Você está na página 1de 137

Light-Matter Interactions and Quantum Optics.

Jonathan Keeling

http://www.st-andrews.ac.uk/~jmjk/teaching/qo/

Contents
Contents Introduction 1 Quantisation of electromagnetism 1.1 Revision: Lagrangian for electromagnetism 1.2 Eliminating redundant variables . . . . . . . 1.3 Canonical quantisation; photon modes . . . 1.4 Dipole approximation and coupling strength 1.5 Further reading . . . . . . . . . . . . . . . . iii vii 1 2 3 5 6 8 9 9 10 11 15 15 16 18 19 20 21 21 22 25 27 29 29 33 35 37 38 42 44

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

2 Jaynes Cummings model 2.1 Derivation of Jaynes-Cummings model . . . . . . . . . . . . 2.2 Semiclassical limit . . . . . . . . . . . . . . . . . . . . . . . 2.3 Quantum behaviour . . . . . . . . . . . . . . . . . . . . . . 3 Decay of a two-level system 3.1 Many mode quantum model irreversible decay . . . . . 3.2 Density matrix equation for relaxation of two-level system 3.3 Eective decay rate . . . . . . . . . . . . . . . . . . . . . . 3.4 Further reading . . . . . . . . . . . . . . . . . . . . . . . . 3.A WignerWeisskopf approach . . . . . . . . . . . . . . . . . 4 Power broadening 4.1 Equations of motion with coherent driving 4.2 Dephasing in addition to relaxation . . . . 4.3 Power broadening of absorption . . . . . . 4.4 Further reading . . . . . . . . . . . . . . .

. . . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

5 Purcell Eect 5.1 The Purcell eect in a 1D model cavity . . . . . . . . . . . 5.2 Weak to strong coupling via density matrices . . . . . . . . 5.3 Further reading . . . . . . . . . . . . . . . . . . . . . . . . . 6 Cavity Quantum Electrodynamics 6.1 Optical transitions of atoms . . . . . . . . . . . . . . . . . . 6.2 Microwave transitions of atoms . . . . . . . . . . . . . . . . 6.3 Superconducting qubits in microwave resonators . . . . . . iii

iv

CONTENTS 6.4 Quantum-dot excitons in semiconductor microstructures . . 6.5 Further reading . . . . . . . . . . . . . . . . . . . . . . . . . 6.A Transfer matrix approach to calculating reectivity of a Bragg mirror . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 50 50 51 51 52 54 56 59 59 61 63 65 67 67 69 70 73 74 75 75 78 79 80 81 83 83 85 87 89

7 Resonance Fluorescence 7.1 Spectrum of emission into a reservoir 7.2 Quantum regression theorem . . . 7.3 Resonance uorescence spectrum . . 7.4 Further reading . . . . . . . . . . . . 8 Quantum stochastic methods 8.1 Quantum jump formalism . . . 8.2 Heisenberg-Langevin equations 8.3 Fluctuation dissipation theorem 8.4 Further reading . . . . . . . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

9 Quantum theory of the laser 9.1 Density matrix equation for a model laser system 9.2 Maxwell Bloch equations . . . . . . . . . . . . . . 9.3 Reduction to photon density matrix equation . . 9.4 Properties of the laser rate equations . . . . . . . 9.5 Further reading . . . . . . . . . . . . . . . . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

10 Laser uctuations: linewidth and correlations 10.1 Laser linewidth . . . . . . . . . . . . . . . . . . . . . . . 10.2 Spontaneous emission and the parameter . . . . . . . 10.3 Fano factor . . . . . . . . . . . . . . . . . . . . . . . . . 10.4 Further reading . . . . . . . . . . . . . . . . . . . . . . . 10.A Derivation of o diagonal eld density matrix equations 11 Stronger coupling, micromasers, 11.1 Micromaser gain . . . . . . . . 11.2 Micromaser and noise . . . . . 11.3 Single atom lasers . . . . . . . 11.4 Further reading . . . . . . . . . single atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . lasers . . . . . . . . . . . . . . . .

. . . . .

. . . . .

. . . .

. . . .

. . . .

. . . .

12 Superradiance 91 12.1 Simple density matrix equation for collective emission . . . 91 12.2 Beyond the simple model . . . . . . . . . . . . . . . . . . . 96 12.3 Further reading . . . . . . . . . . . . . . . . . . . . . . . . . 100 13 The 13.1 13.2 13.3 13.4 13.5 Dicke model Phase transitions, spontaneous superradiance . . No-go theorem: no vacuum instability . . . . . . Radiation in a box; restoring the phase transition Dynamic superradiance . . . . . . . . . . . . . . Further Reading . . . . . . . . . . . . . . . . . . 101 101 103 104 104 106

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

CONTENTS 14 Three levels, and coherent control 14.1 Semiclassical introduction . . . . . 14.2 Coherent evolution alone; why does 14.3 Dark state polaritons . . . . . . . . 14.4 Further reading . . . . . . . . . . . A Problems for lectures 14 B Problems for lectures 59 C Problems for lectures 1014 Bibliography

v 109 109 113 113 116 117 121 123 127

. . . . . . . EIT occur . . . . . . . . . . . . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

Introduction
The title quantum optics covers a large range of possible courses, and so this introduction intends to explain what this course does and does not aim to provide. Regarding the negatives, there are several things this course deliberately avoids: It is not a course on quantum information theory. Some basic notions of coherent states and entanglement will be assumed, but will not be the focus. It is not a course on relativistic gauge eld theories; the majority of solid state physics does not require covariant descriptions, and so it is generally not worth paying the price in complexity of using a manifestly covariant formulation. As far as possible, it is not a course on semiclassical electromagnetism. While at times radiation will be treated classically, this will generally be for comparison to a full quantum treatment, or where such an approximation is valid (for at least part of the radiation). Regarding the positive aims of this course, they are: to discuss how to model the quantum behaviour of coupled light and matter; to introduce some simple models that can be used to describe such systems; to discuss methods for open quantum systems that arise naturally in the context of coupled light and matter; and to discuss some of the more interesting phenomena which may arise for matter coupled to light. Semiclassical behaviour will be discussed in some sections, both because an understanding of semiclassical behaviour (i.e. classical radiation coupled to quantum mechanical matter) is useful to motivate what phenomena might be expected; and also as comparison to the semiclassical case is important to see what new physics arises from quantised radiation. The kind of quantum optical systems discussed will generally consist of one or many few-level atoms coupled to one quantised radiation elds. Realisations of such systems need not involve excitations of real atoms, but can instead be articial atoms, i.e. well dened quantum systems with discrete level spectra which couple to the electromagnetic eld. Such concepts therefore apply to a wide variety of systems, and a variety of characteristic energies of electromagnetic radiation. Systems currently studied experimentally include: real atomic transitions coupled to optical cavities[1]; vii

viii

INTRODUCTION

Josephson junctions in microwave cavities (waveguides terminated by reecting boundaries)[2, 3]; Rydberg atoms (atoms with very high principle quantum numbers, hence small dierences of energy levels) in GHz cavities[4]; and solid state excitations, i.e. excitons or trions localised in quantum dots, coupled to a variety of optical frequency cavities, including simple dielectric contrast cavities, photonic band gap materials, and whispering gallery modes in disks[5]. These dierent systems provide dierent opportunities for control and measurement; in some cases one can probe the atomic state, in some cases the radiation state. To describe experimental behaviour, one is in general interested in calculating a response function, relating the expected outcome to the applied input. However, to understand the predicted behaviour, it is often clearer to consider the evolution of quantum mechanical state; thus, both response functions and wavefunctions will be discussed. As such, the lectures will switch between Heisenberg and Schr odinger pictures frequently according to which is most appropriate. When considering open quantum systems, a variety of dierent approaches; density matrix equations, Heisenberg-Langevin equations and their semiclassical approximations, again corresponding to both Schr odinger and Heisenberg pictures. The main part of this course will start with the simplest case of a single two-level atom, and discuss this in the context of one or many quantised radiation modes. The techniques developed in this will then be applied to the problem of many two-level atoms, leading to collective eects. The techniques of open quantum systems will also be applied to describing lasing, focussing on the more quantum examples of micromasers and single atom lasers. The end of the course will consider atoms beyond the two-level approximation, illustrating what new physics may arise. Separate to this main discussion, the rst two lectures stand alone in discusing where the simple models of coupled light and matter used in the rest of the course come from, in terms of the quantised theory of electromagnetism.

Lecture 1

Quantisation of electromagnetism in the Coulomb gauge


Our aim is to write a theory of quantised radiation interacting with quantised matter elds. Such a theory, e.g. the Jaynes-Cummings model (see next lecture) has an intuitive form:

HJ.C. =
k

k a k ak +
i,k

z i i

+ + gi,k i ak + H.c. .

(1.1)

The operator a k creates a photon in the mode with wavevector k , and so this Hamiltonian describes a process where a two-level system can change its state with the associated emission or absorption of a photon. The term k a k ak then gives the total energy associated with occupation of the mode with energy k . While the rest of the course is dedicated to studying such models of coupled light-matter system, this lecture will show the relation between such models and the classical electromagnetism of Maxwells equations. To reach this destination, we will follow the route of canonical quantisation; our rst aim is therefore to write a Lagrangian in terms of only relevant variables. Relevant variables are those where both the variable and its time derivative appear in the Lagrangian; if the time derivative does not appear, then we cannot dene the canonically conjugate momentum, and so cannot enforce canonical commutation relations. The simplest way of writing the Lagrangian for electromagnetism contains irrelevant variables i.e. the electric scalar potential and gauge of the vector potential A; that irrelevant variables exist is due to the gauge invariance of the theory. Since we are not worried about preserving manifest Lorentz covariance, we are free to solve this problem in the simplest way eliminating the unnecessary variables. 1

LECTURE 1. QUANTISATION OF ELECTROMAGNETISM

1.1

Revision: Lagrangian for electromagnetism

To describe matter interacting with radiation, we wish to write a Lagrangian whose equations of motion will reproduce Maxwells and Lorentzs equations: B = 0 J + 0 0 E B=0 B(r )]. r = q [E(r ) + r m E = /0 E = B (1.2) (1.3) (1.4)

Equations (1.3) determine the structure of the elds, not their dynamics, . Let and are immediately satised by dening B = A and E = A us suggest the form of Lagrangian L that leads to Eq. (1.2) and Eq. (1.4): L=

1 0 m r 2 + 2 2

dV E2 c2 B2 +

q [ r A(r ) (r )] .

(1.5) Here, the elds E and B should be regarded as functionals of and A. Note also that in order to be able to extract the Lorentz force acting on individual charges, the currents and charge densities have been written as: (r) =

q (r r ),

J(r) =

(r r ). q r

(1.6)

The identication of the Lorentz equation is simple: d L d + q A(r )] = m = [m r r + q ( r )A(r ) + q A(r ) dt r dt t L = = q [ r A(r ) (r )] r ( A(r )) (r )] , = q [( r ) A(r ) + r thus one recovers the Lorentz equation, [ A(r )] (r ) m r = q r A(r ) . t (1.7)

Similarly, the equation that results from can be easily extracted; since = 0, this becomes L/ L = 0 E q (r r ) = 0.

(1.8)

Finally, the equations for A are more complicated, requiring the identity ( A)2 = 2 ( A), A which then gives: d L d L = 0 E = = 0 c2 ( A) + dt A dt A (r r ), q r

(1.9)

(1.10)

1.2. ELIMINATING REDUNDANT VARIABLES which recovers the required Maxwell equation 0 d 1 E= B+ dt 0 (r r ). q r

(1.11)

Thus, the Lagrangian in Eq. (1.5), along with the denitions of E and B in terms of A and produce the required equations.

1.2

Eliminating redundant variables

As mentioned in the introduction, we must remove any variable whose time derivative does not appear in the Lagrangian, as one cannot write the required canonical commutation relations for such a variable. It is clear from Eq. (1.5) that the electric scalar potential is such a variable. does not appear, it is also possible to eliminate directly from Since the equation L/; using Eq. (1.8) and the denition of E, this equation gives: 0 2 (r) = 0. 0 A (1.12) Rewriting this in Fourier space, one has: (k) = 1 k2 (k) ( k) . + ik A 0 (1.13)

We can now try to insert this denition into the Lagrangian, to eliminate . To do this, we wish to write E2 and B2 in terms of and A; it is therefore useful to start by writing j (k) = i kj (k) + jk kj kk Ej (k) = ikj (k) + A 0 k 2 k2 k (k). (1.14) A

This means that the electric eld depends on the charge density, and on the transverse part of the vector potential, which will be written: A j k= jk kj kk k2 k (k). A (1.15)

The transverse1 part of the vector potential is by denition orthogonal to the wavevector k, and so the electric eld is the sum of two orthogonal vectors, and so: (k)|2 + 1 |(k)|2 . |E(k)|2 = |A (1.18) 2 2 0k
1

The combination:
jk (k ) =

jk

kj kk k2

(1.16)

is the reciprocal space representation of the transverse delta function; with appropriate regularisation[6, Complement AI ], it can be written in real space as:
jk (r) =

2 1 rj rk (r)jk + 3 2 jk . 3 4r3 r

(1.17)

LECTURE 1. QUANTISATION OF ELECTROMAGNETISM

Similarly, the squared magnetic eld in reciprocal space is given by: |B(k)|2 = (k A(k)) (k A (k)) = k 2 jk kj kk k2
2 2 Aj (k)A k (k) = k |A (k)| .

(1.19)

Thus, the eld part of the Lagrangian becomes: 0 2 dV E2 c2 B2 = 1 1 d3 k 2 |(k)|2 0 k (k)|2 c2 k 2 |A (k)|2 . (1.20) + 0 d3 k |A The notation has been introduced to mean integration over reciprocal half-space; since A(r) is real, A (k) = A(k), thus the two half spaces are equivalent. This rewriting is important to avoid introducing redundant elds in the Lagrangian; the eld is either specied by one real variable at all points in real space, or by two real variables at all points in reciprocal half-space. Similar substitution into the coupling between elds and matter, written in momentum space gives: Lemmatter = 2 d3 k [J(k) A (k) (k) (k)] = 2 d3 k J(k) A (k) (k) (k) ik 2 A (k) 0 k 2 k . (1.21) This can be simplied by adding a total time derivative, L L + dF/dt; such transformations do not aect the equations of motion, since they add only boundary terms to the action. If: F = 2 d3 k(k) then |(k)|2 . 0 k 2 (1.23) Then, using conservation of current, (k) + ik J(k) = 0, one nally has: Lemmatter + dF = 2 d3 k dt J(k) ( k) ik k2 A ( k) Lemmatter + dF |(k)|2 = 2 d3 k J (k) A (k) . dt 0 k 2 (1.24) ik A (k) , k2 (1.22)

Note that this set of manipulations, adding dF/dt has eliminated the longitudinal part of the vector potential from the Lagrangian. The form chosen for F is such that this procedure is equivalent to a gauge transformation; the chosen gauge is the Coulomb gauge. Putting everything together, one has: Lcoulomb =

1 1 1 m r 2 d3 k 2 |(k)|2 2 0 k J (k) A (k) . (1.25)

(k)|2 c2 k 2 |A (k)|2 + 2 + 0 d3 k |A

1.3. CANONICAL QUANTISATION; PHOTON MODES

Thus, the nal form has divided the interaction into a part mediated by transverse elds, described by A , and a static (and non-retarded) Coulomb interaction. Importantly, there are no irrelevant variables left in Eq. (1.25) The Coulomb term can also be rewritten: Vcoul = 1 1 d3 k 2 |(k)|2 = 0 k q q . 80 |r r | (1.26)

Note that since the Coulomb interaction is non-retarded, both the Coulomb and transverse parts of interaction must be included to have retarded interactions between separated charges.

1.3

Canonical quantisation; photon modes

We have in Eq. (1.25) a Lagrangian which can now be treated via canonical quantisation. Since only the transverse part of the eld A appears in Eq. (1.25), we can drop the superscript label in A from here on. To proceed, we should rst identify the canonical momenta and the Hamiltonian, and then impose canonical commutation relations. Thus, Lcoulomb + q A(r ) = mr r Lcoulomb (k). (k) = = 0 A (k) A p = Then, constructing the Hamiltonian by H = H=
i Pi Ri

(1.27) (1.28)

L, one nds:

1 [p q A(r )]2 + Vcoul 2m |(k)|2 + c2 k 2 |A(k)|2 . 2 0 (1.29)

+ 0 d3 k

In order to quantise, it remains only to introduce commutation relations for the canonically conjugate operators. Noting that A(r) has only two independent components, because it is transverse, it is easiest to write its commutation relations in reciprocal space, introducing directions ek,n orthogonal to k with n = 1, 2 ; then: [ri, , pj, ] = i ij Aek,n (k), ek ,n (k ) = i (k k )nn . (1.30) (1.31)

This concludes the quantisation of matter with electromagnetic interactions in the Coulomb gauge. It is however instructive to rewrite the transverse part of the elds in terms of their normal modes. The second line of Eq. (1.29) has a clear similarity to the harmonic oscillator, with a separate oscillator for each polarisation and momentum. Rewriting in normal modes thus means introducing the ladder operators: ak,n = 0 i ckAek,n (k) + ek,n (k) , 2 ck 0 (1.32)

LECTURE 1. QUANTISATION OF ELECTROMAGNETISM

or, inverted one has: A(r) =


k n=1,2 i k r n ak,n eikr + a . e k,n e

20 k V

(1.33)

Inserting this into Eq. (1.29) gives the nal form: H=

1 [p q A(r )]2 + 2m

k a k ak + Vcoul .
k n=1,2

(1.34)

1.4

Dipole approximation and coupling strength

Equation (1.34) applies to any set of point charges interacting with the electromagnetic eld. In many common cases, one is interested in dipoles, with pairs of opposite charges closely spaced, and much larger distances between the dipoles. In this case it is possible to make the dipole approximation, by assuming that the electromagnetic eld does not vary across the scale of one dipole. We consider a system with just two charges: charge +q mass m1 at R + r/2 and q mass m2 at R r/2. If r where is a characteristic wavelength of light, then one may assume that A(R + r/2) A(R r/2) A(R), and so the remaining coupling between radiation and matter is of the form p1 p2 HAp = q A(R). m1 m2 Then, in the case that one can write H0 , a Hamiltonian for the dipole without its coupling to radiation one can use p/m = x = i[H0 , x]/ , thus giving: q HAp = i [H0 , r] A(R) (1.35) If one considers a given basis of eigenstates |n for the free Hamiltonian H0 obeying H0 |n = n |n one may then rewrite the above in the form: k,n gnm ikR HAp = (1.36) |n m| ak,n eikR + a k,n e 2 n,m
k,n where we have introduced the coupling strengths gnm . These describe the transitions between bare levels that are driven by coupling to the photon eld, and are given by the expression k,n gnm q = ek,n n|[H0 , r]|m = 2 20 k V nm ek,n

dnm , 2 0 k V

(1.37)

where dnm = q n|r|m is the dipole matrix element and nm = n m . We have assumed H0 commutes with the parity operator P : P x = x, and so the coupling to radiation appears only in the o-diagonal terms in the two-level basis.

1.4. DIPOLE APPROXIMATION AND COUPLING STRENGTH

Oscillator strength
The expression in Eq. (1.37) shows that the coupling strength between light and matter depends both on properties of the atom and the radiation modes. The relevant property of the radiation mode is the eld strength, describing the electric eld (or vector potential) associated with a single photon. The relevant propert of the atom is a combination of its energy and dipole matrix element. There is another characterisation of an atoms coupling to light, which is given by the atomic oscillator strength. We will here restrict attention to transitions involving coupling to electrons, hence q = e, m = m0 . fnm = 2m0 |dnm |2 2 e2
nm .

(1.38)

This dimensionless combination is important for several reasons, which are discussed partly here, and further elaborated in the problems. The rst is that the oscillator strength describes the relative contribution of a given atomic transition to the atomic susceptibility, as compared to an harmonic oscillator. The second is that there is a sum rule which states n fnm = 1 (at least for transitions involving only a single electron changing state), implying that the dimensionless combination fnm states how large the coupling is as compared to the maximum value it could possibly take. The susceptibility of a Lorentz oscillator is calculated by nding P ( ) = ( ) 0 E ( ) = ex( ), where x is the displacement of the oscillator, and the oscillator obeys: 2 mx = m0 x 2mx eE (t) (1.39) It is easy to then decompose this into frequency components and thus nd LO ( ) = e2 1 2. 2 m0 0 + 2i 0 (1.40)

To see the signicance of the oscillator strength fnm , this susceptibility should be compared to that corresponding to Eq. (1.36). This means calculating: P (t) = dP ( )eit = |
mn m m |m

dmn |m n|

(1.41)

in linear response theory, with H0 = H1 = i


nm

m| and the perturbation d E ( ) (it) e . (1.42)

A(t)

nm dnm |n

m|,

A(t) = i

Here, A(t) is a classical eld corresponding to the probe eld E (t), and is an innitessimal shift to ensure the eld vanishes at t . Exercise A1 shows that the linear response found here evaluates to: ( ) = e2 m0 0 fnm
nm

( +

i )2

1 (

nm /

)2

(1.43)

indicating that each transition contributes like a harmonic oscillator, but with a reduced strength given by fnm .

LECTURE 1. QUANTISATION OF ELECTROMAGNETISM

1.5

Further reading

The discussion of quantisation in the Coulomb gauge in this chapter draws heavily on the book by Cohen-Tannoudji et al. [6].

Lecture 2

The Jaynes-Cummings model


The aim of this lecture is to discuss the interaction between a single twolevel system and various dierent congurations of photon elds. We start by introducing the Jaynes-Cummings model, and then consider various results of this model. These will include the interaction with a semiclassical light eld, and the behaviour of a single quantised mode of light in an initially coherent state.

2.1

Derivation of Jaynes-Cummings model

At the end of the previous lecture, we began to discuss approximations that result in simplied descriptions of coupled lightmatter systems. In this section, we continue this programme of simplications in order to reduce the model of coupled matter and radiation modes to the Jaynes-Cummings model, of a two-level atom coupled to a single photon model. To eect this reduction, three further simplications are required in addition to the dipole approximation discussed previously: neglect of the A2 terms, projection to two-level systems, and the rotating wave approximation. Neglect of the A2 terms in expanding [p q A(r )]2 can be justied in the limit of low density of atoms; considering only a single radiation mode, the contribution of the A2 term can be rewritten using Eq. (1.33) as: 2 N q2 HA2 = a + a . (2.1) V 4m0 Thus, this term adds a self energy to the photon eld, which scales like the density of atoms. The relative importance of this term can be estimated by comparing it to the other term in the Hamiltonian which is quadratic in the photon operators, k a k ak . Their ratio is given by: q2 N N a3 V 4m0 (k )2 V B Ryd k
2

(2.2)

thus, assuming particles are more dilute than their Bohr radius, neglect of HA2 is valid for frequencies of the order of the Rydberg for the given 9

10

LECTURE 2. JAYNES CUMMINGS MODEL

bound system of charges. To further simplify, one can now reduce the number of states of the atom that are considered; currently, there will be a spectrum of eigenstates of H0 , and transitions are induced between these states according to f |r|i . Then, restricting to only the two lowest atomic levels and to a single radiation mode, one has a model of two-level systems (describing matter) coupled to bosonic modes (describing radiation). H= 1 2 g (a + a ) g (a + a ) + 0 a a. (2.3)

The matrix notation represents the two levels of the atom. We have introduced the energy splitting between the lowest two atomic levels, and we have used the notation g for the eective coupling strength gab as dened in the previous lecture: ek,n dba g , (2.4) = 2 2 0 k V The nal approximation to be discussed here, the rotating wave approximation, is appropriate when g 0 . Considering g as a perturbation, one can identify two terms: co = g 2 0 a a 0 , counter = g 2 0 a a 0 , (2.5)

where co , the co-rotating terms, conserve energy particle number; and counter do not. More formally, the eects of counter give second order perturbation terms like g 2 /(0 + ), while co give the much larger g 2 /(0 ). Including this approximation, one then derives the Jaynes-Cummings model: 1 g H = z + (a + a + ) + 0 a a. (2.6) 2 2 where represent Pauli matrices.

2.2

Semiclassical limit

On applying a classical light eld to a two-level system, we can adapt the Jaynes-Cummings model by neglecting the dynamics of the photon eld, and replacing the photon creation and annihilation operators with the amplitude of the time-dependent classical eld. This gives the eective Hamiltonian: 1 gei0 t . (2.7) H= i t 2 ge 0 We have made the rotating wave approximation, and hence only included the positive/negative frequency components of the classical eld in the raising/lowering operators for the two-level system. To solve this problem, and nd the time-dependent state of the two-level atom, it is convenient to make a transformation to a rotating frame. In general, if U= eif (t) 0 0 eif (t) , (2.8)

2.3. QUANTUM BEHAVIOUR then = HH

11

H11 + f(t) H12 e2if (t) H21 e2if (t) H22 f(t)

(2.9)

thus, in this case we use f = t/2, giving: H= 1 2 g g ( ) . (2.10)

Since this transformation aects only the phases of the wavefunction, we can then nd the absorption probability by considering the time averaged probability of being in the upper state. The eigenvalues of Eq. (2.8) are /2 where = ( )2 + g 2 2 . Writing the eigenstates as (cos , sin )T , and ( sin , cos )T corresponding respectively to the roots (assuming tan() > 0). Substituting these forms, one nds the condition: tan(2) = g/( ). This form is sucient to nd the probability of excitation. If the two-level system begins in its ground state, then the time-dependent state can be written as: = sin cos sin eit/2 + cos sin cos eit/2 (2.11)

1 cos(t) (g)2 . 2 (g)2 + ( )2 (2.12) Thus, the probability of excitation oscillates at the Rabi frequency = ( )2 + (g)2 , and the amplitude of oscillation depends on detuning. On resonance ( = ), one can engineer a state with | | = | | = 1/ 2, or a state with | | = 1 by applying a pulse with a duration (g)t = /2 or respectively. Pex = | |2 = |2 sin cos sin(t/2)|2 =

2.3

Quantum behaviour

Let us now repeat this analysis in the case of a quantised mode of radiation, starting with the Jaynes-Cummings model in the rotating wave approximation. 1 ga H= + 0 a a. (2.13) ga 2

Rabi oscillations in the Jaynes-Cummings model


In the rotating wave approximation the total number of excitations Nex = z + a a is a conserved quantity, thus if we start in a number state of the light eld: (a )n |n, = |0, , (2.14) n! then the general state at later times can always be written: | (t) = (t) |n 1, + (t) |n, . (2.15)

12

LECTURE 2. JAYNES CUMMINGS MODEL

Inserting this ansatz in the equation of motion gives 1 1 ( ) (t) g n = n it 1 + (t) g n ( ) 2 2

(t) (t)

(2.16) After removing the common phase variation exp[i(n 1/2)t[, the general solution can be written in terms of the eigenvectors and values of the matrix in Eq. (2.16), and reusing the results for the semiclassical case, we have the state (t) (t) = ei(n1/2)t sin eit/2 cos sin + cos eit/2 sin , cos (2.17)

and the excitation probability is thus: 1 cos(t) g2n (2.18) 2 g 2 n + ( )2 where we now have tan(2) = g n/( ) and = ( )2 + g 2 n. Pex =

Collapse and revival of Rabi oscillations


Restricting to the resonant case = , let us discuss what happens if the initial state does not have a well dened excitation number. Since the oscillation frequency depends on n, the various components of the initial state with dierent numbers of excitations will oscillate at dierent frequencies. We can therefore expect interference, washing out the signal. This is indeed seen, but in addition one sees revivals; the signal reappears at much later times, when the phase dierence between the contributions of succesive number state components is an integer multiple of 2 . Let us consider explicitly how the probability of being in the excited state evolves for an initially coherent state | = e||
2 /2

exp(a ) |0, = e||

2 /2

n |n, . n!

(2.19)

If resonant, the general result in Eq. (2.17) simplies, as = /4. Following the previous analysis, the state at any subsequent time is given by: | = e(||
n
2 +it)/2

(eit )n cos n!

g nt 2

|n, + sin

g nt 2

|n 1,

. (2.20)

Thus, the probability of being in the excited atomic state, traced over all possible photon states, is given by ||2n 2 ||2 2 g nt Pex = | n, | | = e sin n! 2 n n = 1 1 ||2 e 2 2
n

||2n cos g nt . n!

(2.21)

2.3. QUANTUM BEHAVIOUR

13

We consider the case 1. In this case, one may see that the coecients ||2n /n! are sharply peaked around n ||2 by writing: ||2n 1 = exp[f (n)], n! 2n f (n) = n ln(n) n n ln(||2 ), (2.22)

By dierentiating f twice, one nds f (n) = ln n/||2 , f = 1/n, and hence one may expand f (n) near its minimum at n = ||2 to give: f (n = ||2 + x) ||2 + x2 2||2 (2.23)

Using this in Eq. (2.21) gives the approximate result: Pex 1 2 2 1 2 ||2
m

exp

m2 + igt ||2 + m 2||2

(2.24)

where it is assumed that ||2 is large enough that the limits of the sum may be taken to . Three dierent timescales aect the behaviour of this sum. Concentrating on the peak, near m = 0, there is a fast oscillation frequency g, describing Rabi oscillations. Then, considering whether the terms in the sum add in phase or out of phase, there are two time scales: collapse of oscillations occurs when there is a phase dierence of 2 across the range of terms |m| < m contributing to this sum, i.e. when gTcollapse ( ||2 + m ||) = . Taking m = || from the Gaussian factor, this condition becomes gTcollapse 2 . On the other hand, if the phase dierence between each successive term in the sum is 2 , then they will rephase, and a revival occurs, with the condition gTrevival ( ||2 + 1 ||) = 2 , giving gTrevival = 4 ||. Thus, the characteristic timescales are given by: Toscillation 1 , g || Tdecay 1 , g Trevival || . g (2.25)

and the associated behaviour is illustrated in Fig. 2.1. Let us now consider how this behaviour may be approximated analytically. The simplest approach might be to replace the sum over number states by integration, however this would inevitably lose the revivals, which are associated with the discreteness of the sum allowing rephasing. Thus, it is necessary to take account of the discreteness of the sum, and hence the dierence between a sum and an integral. For this, we make use of Poisson summation which is based on the result: (x m) =
m r

e2irm

f (m) =
r

dxe2irx f (x). (2.26)

This allows us to replace the summation by integration, at the cost of having a sum of nal results, however in the current context this is very helpful, as the nal sum turns out to sum over dierent revivals. Applying 1 this formula, we may write Pex 2 [1 r {(r, ||, t)}] where (r, ||, t) = dx 2 ||2 exp 2irx x2 x2 x + igt || + 2 2|| 2|| 8||3 (2.27)

14

LECTURE 2. JAYNES CUMMINGS MODEL

1 1 0.8 0.6 0.4 0.2 0 0

Excitation probability

0.5 0 0

20

40

60

80 100 120 140 160 180 200

Time g
Figure 2.1: Collapse and revival of Rabi oscillations, plotted for
= 200.

Completing the square yields: (r, ||, t) = = eigt|| 2 ||2 eigt|| 2 ||2 = dx exp exp 1 2||2 1+ igt 4|| x2 + i 2r + gt 2|| x

||2 (2r + gt/2||)2 2 1 + igt/4|| 1 2||2 1+ igt 4|| x i||2 2r + gt/2|| 1 + igt/4||
2

dx exp eigt||

||2 (2r + gt/2||)2 exp . 2 1 + igt/4|| 1 + igt/4||

(2.28)

From this nal expression one may directly read o: the oscillation frequency g ||; the characteristic time for the rst collapse 2 2/g (by taking r = 0 and assuming gt/|| is small); the delay between successive revivals 4 ||/g ; and nally one may also see that the revival at gt = 4r|| has an increased width and decreased height compared to the initial collapse, given by a factor 1 + 2 r2 as is visible in gure 2.1.

Further reading
The contents of most of this chapter is discussed in most quantum optics books. For example, it is discussed in Scully and Zubairy [7], or Yamamoto and Imamo glu [8]. Further properties of collapse and revival of JaynesCummings oscillations are discussed by Gea-Banacloche [9].

Lecture 3

Two level systems coupled to many modes: Decay and density matrix equations
The aim of this lecture is to introduce the behaviour of a two-level system coupled to many photon modes, such that irreversible decay now occurs. We will use the density matrix description of the two-level system, and show how coupling to a reservoir of bath modes induces irreversible decay. The formalism we develop here will be useful also when we later consider the laser, and its more quantum mechanical variants of single atom lasers and micromasers.

3.1

Many mode quantum model irreversible decay

In the previous lecture we discussed a two level system coupled to a single mode, considering both classical and quantum light. One important difference between these situations is that for classical light, if the light eld vanishes, then there can be no transitions; in contrast, for the quantum description, spontaneous emission also existed, in that the state |0, hybridises with the state |1, , due to the +1 in the n + 1 matrix elements. We will now consider the eects of this +1 factor when we couple a single two level system to a continuum of radiation modes. Our Hamiltonian will remain in the rotating wave approximation, but no longer restricted to a single mode of light, and so we have: H = z + 2 where k = ck and ek,n dba gk,n = . 2 20 k V 15 (3.2) k a k,n ak,n +
n,k k

gk,n ak,n + H.c. 2

(3.1)

16

LECTURE 3. DECAY OF A TWO-LEVEL SYSTEM

3.2

Density matrix equation for relaxation of two-level system

We now discuss how the irreversible decay arising from the coupling to many photon modes may be described in the density matrix approach. We proceed by eliminating the the behaviour of the continuum of radiation modes in order to write a closed equation for the density matrix of the system. If we dene the system density matrix as the result of tracing over all reservoir degrees of freedom, we can write an equation of motion for this quantity as: d S = iTrR {[HSR , SR ]} . (3.3) dt in which S means state of the system, R the reservoir (in this case, the continuum of photon modes), and TrR traces over the state of the reservoir. After removing the free time evolution of the reservoir and system elds, the remaining system-reservoir Hamiltonian can be written in the interaction picture as: HSR =
k

gk i(k a ke 2

)t

+ + ak ei(

k )t

(3.4)

In general Eq. (3.3) will be complicated to solve, as interactions between the system and reservoir lead to entanglement, and so the full density matrix develops correlations. This means that the time evolution of the system would depend on such correlations, and thus on the history of system reservoir interactions. However this will simplify is we assume a smooth density of states for the bath, in which case it has a short correlation (memory) time, and memory eects can be neglected. This will allow us to make a Markov approximation, meaning that the evolution of the system depends only on its current state i.e. neglecting memory eects of the interaction. For a general interaction, there are a number of approximations which are required in order to reduce the above equation to a form that is straightforwardly soluble. We assume the interaction is weak, so SR = S R + SR , where SR O(gk ). Since S = TrR (SR ), it is clear that TrR (SR ) = 0, however the result of TrR (HSR SR ) may be non-zero. Such correlations will be assumed to be small, but must be non-zero in order for there to be any inuence of the bath on the system. In order to take account of these small correlation terms, generated by the coupling, we use the formal solution of the density matrix equation t SR (t) = i dt [HSR (t ), SR (t )] to write: d S (t) = TrR dt
t

dt HSR (t), HSR (t ), SR (t )


0

(3.5)

We may now make the assumption of small correlations by assuming that SR (t ) = S (t )R (t ), and that only the correlations generated at linear order need be considered.

3.2. DENSITY MATRIX EQUATION FOR RELAXATION OF TWO-LEVEL SYSTEM

17

We further assume the reservoir to be large compared to the system, and so unaected by the evolution, thus R (t ) = R (0). Along with the previous step, this is the Born approximation. Finally, if the spectrum of the bath is dense (i.e. the spacing of energy levels is small), then the trace over bath modes of the factors eik t will lead to delta functions in times. This in eect means that we may replace S (t ) = S (t). This is the Markov approximation. One thus has the nal equation: d S (t) = TrR dt
t

dt HSR (t), HSR (t ), S (t) R (0)


0

(3.6)

To evaluate the trace over the reservoir states, we need the trace of the various possible combinations of reservoir operators coming from HSR . These will involve pairs of operators ak , a k . Using cyclic permutations, these trace terms can be written as thermal average X = TrR (R X ). For bulk photon modes, the relevant expectations are: = kk (nk + 1). (3.7) In these expressions we have assumed nk = 0, i.e. we have allowed for a thermal occupation of the photon modes. The results of the previous lecture assumed an initial vacuum state, and thus nk = 0. With these expectations, we can then write the equation of motion for the density matrix explicitly, writing k = exp[i(k )(t t )], to give: ak ak = 0, = kk nk , d S (t) = dt gk 2
2 0 t = a k ak

a k ak

ak a k

dt

+ nk k + + (nk + 1)k s

) + (nk + 1) + + nk (k + k +s + nk k + + (nk + 1)k

. (3.8)

In this expression, the rst term in parentheses comes from the expression TrR [HSR (t)HSR (t )SR ], the second comes from the conjugate pair )], and the nal term TrR [HSR (t)SR HSR (t )] and t t [hence (k + k comes from TrR [SR HSR (t )HSR (t)]. At this point, we can now simplify the above expression by restricting to the case where there is a smooth density of states. We dene (t t ) =
k

gk 2

eik (tt )

( ) = 2
k

gk 2

( k ). (3.9)

Then, the restriction that is a smooth function means that ( ) does not signicantly vary over the range , i.e. d/d 1. In this case (which corresponds to the Markov approximation, meaning a at eective density of states) the integral over in Eq. (3.23) becomes a delta function, so that: gk 2 gk 2 k = (t t ), nk k = n (t t ). (3.10) 2 2
k k

18

LECTURE 3. DECAY OF A TWO-LEVEL SYSTEM

Here we have assumed that not only is the density of states at, but that also nk is suciently at, that we might write an averaged occupation n . Such an approximation is only valid at high enough temperatures we will discuss this in some detail in a few lectures time. d S (t) = n + S 2 + S + S + dt 2 ( n + 1) + S 2 S + + S + 2

(3.11)

The factor of 1/2 here comes from the regularised half integral of the delta function. The two lines in this expression correspond to stimulated absorption (which exists only if n > 0), and leads to excitation of the two-level system, and to emission (both stimulated and spontaneous). If n = 0, only emission survives, and the results of the previous lecture are recovered for the relaxation to equilibrium. One can see the behaviour from this equation most clearly by writing the equations for the diagonal components: t = n (2 ) ( n + 1) (2 ) = t 2 2 (3.12)

Then, using = 1 , one may nd the steady state probability of excitation is given by: 0 = [ n(1 ) ( n + 1) ] = n . 2 n+1 (3.13)

This formula is as expected, such that if one uses the Bose-Einstein occupation function for the photons, n = [exp( ) 1]1 , the excitation probability of the two-level system is the thermal equilibrium expression.

3.3

Eective decay rate

We have not yet calculated an actual expression for the eective decay rate appearing in the previous expression. Let us now use the form of gk in Eq. (3.2) to evaluate ( ). Writing the wavevector in polar coordinates, with respect to the dipole matrix element dab pointing along the z axis (see Fig. 3.1), we then have: ( ) = 2 V (2 )3 dd sin k 2 dk ( ck )
2 |d |2 ab

20 ckV

kn ) 2 . ( ze
n

(3.14) e k2 = 0 and one may see that z From Fig. 3.1 it is clear that z ek1 = sin . For concreteness, the k direction and polarisations can be parametrised by the orthogonal set: sin cos cos cos sin = sin sin , e k1 = cos sin , e k2 = cos . k cos sin 0 (3.15)

3.4. FURTHER READING

19

e2 e1

Figure 3.1: Directions of polarisation vectors in polar coordinates

Putting all the terms together, one nds: ck cdk 1 4|dab |2 3 ( ck ) = 3 c3 40 3 3 c3 (3.16) This formula therefore gives the decay rate of an atom in free space, associated with the strength of its dipole moment and its characteristic energy. It is useful to rewrite this decay rate in terms of the oscillator strength fab of the transition involved. Recalling the denition of fab , and continuing to use the shorthand = ab we have ( ) =
2 |d |2 ab (2 )2 20

d sin3

2 e2 4 1 1 e2 2 f = fab ab 3 3 3 40 c 2m0 3 40 c m0 c2

(3.17)

Then, using the additional results: = e2 , 40 c 1 Ryd = 2 m0 c2 2 (3.18)

where Ryd is the Rydberg energy and the ne structure constant, one can write the decay rate in the easily interpretable form: 1 = 3 fab . 3 Ryd (3.19)

The smallness of is thus crucial in the fact that atomic lines are well dened (i.e. have linewidths smaller than their separation).

3.4

Further reading

Once again, the contents of this chapter can be found in most standard quantum optics books, e.g.[7, 10].

20

LECTURE 3. DECAY OF A TWO-LEVEL SYSTEM

3.A

WignerWeisskopf approach

An alternative method to show the decay rate in simple situations is to consider solving the Schrodinger equation directly, to nd the probability that the two-level system remains excited. To study spontaneous emission, we start from the photon state |0, , in which there are no photons. The subsequent state can be written as: | = (t)|0 +
k

k,n (t)|1k,n ,

(3.20)

where |1k,n contains a single photon in state k, n. Then, writing the Schr odinger equations as: it = + 2 gk,n k,n , 2 gk,n it k,n = + k k,n + , 2 2 (3.21)

k,n

k ei( /2k )t , one may solve these equations by rst writing = ei t/2 , k = and then formally solving the equation for k , with the initial condition k (0) = 0, to give: k,n = i gk,n 2
t

dt ei(k

)t

(t ).

(3.22)

Substituting this into the equation for gives:


t

t =
t

dt
k,n

gk,n 2

ei(k

)(t t)

(t ) (3.23)

dt

d ( )ei( 2

)(t t)

(t )

where we have used the same denition of as previously: ( ) = 2


k,n

gk,n 2

( k,n )

(3.24)

We may then make the same Markov approximation as before to yield


t

t =

dt ( ) (t t) (t ) =

( ) (t) 2

(3.25)

Thus, the probability of remaining in the excited state decays exponentially due to spontaneous decay, with a decay rate ( ) for the probability. The approach used in this section is known as the WignerWeisskopf approach.

Lecture 4

Density matrix equations with decay and coherent driving


This lecture applies the framework of density matrix equations introduced in the previous lecture in order to incorporate other relaxation and decoherence processes, such as those arising due to noise sources, or inhomogeneous broadening in an ensemble of atoms. We will then illustrate the use of this approach by discussing the behaviour of a two-level system illuminated by a coherent state of light, but able to radiate into a continuum of modes. In this lecture we will study the steady state of this system, and in subsequent lectures we will look at the spectrum of emitted light, which will require knowing more about the full time-evolution of the system.

4.1

Equations of motion with coherent driving

Before introducing extra contributions to dephasing, we begin by applying the method developed in the previous lecture to consider the combination of coherent driving and decay. We will assume from hereon that the reservoir photon modes are empty.1 The evolution of the density matrix is then controlled by: t = i [H, ] + S 2 S + + S + 2 (4.1)

where H describes the coherent, Hamiltonian dynamics, which we will take to be given by: g it + H = z + e + H.c.. (4.2) 2 2
This limit is frequently relevant in cavity quantum electrodynamics experiments, since the conned photon modes start at energies of the order of 1eV 104 K, so the thermal population of such photon modes is negligible.
1

21

22

LECTURE 4. POWER BROADENING

One may then write out the equations of motion for each component explicitly, giving t = i t g it g e eit , (4.3) 2 2 2 g it g e eit . (4.4) = i + 2 2 2 2 2 2 +

These two equations are sucient, since the unit trace property of a density matrix implies = 1 , and the o diagonal elements are complex conjugate of one another by the Hermitian structure of the density matrix. To solve these equations, it is convenient to rst go to a rotating frame, so that one writes = P eit , and secondly to introduce the inversion, Z = ( )/2. In terms of these variables, one has: g 1 (P P ) Z + 2 2 g t P = i( )P + i 2Z P. 2 2 t Z = i (4.5) (4.6)

To further simplify, we may separate the real and imaginary parts of P = X + iY , and write = for the detuning of the optical pump, so that one has the three coupled equations: X 2

t X = Y

(4.7) (4.8) (4.9)

t Y = +X + gZ

Y 2 1 . t Z = gY Z + 2

These are Bloch equations for the Bloch vector parametrisation of the density matrix (in the rotating frame): = +Z X + iY
1 2

X iY 1 2 Z

(4.10)

The three terms, , g, correspond to: rotation about the Z axis (i.e. phase evolution); rotation about the X axis (i.e. excitation); and causes the length of the Bloch vector to shrink. Note that the rotation rate around the X axis is g, so a duration t = /g would lead to the transition from ground to inverted state. Before solving Eq. (4.74.9), we will rst consider other kinds of dephasing that may aect the density matrix evolution, leading to a generalised version of these equations, allowing for both relaxation and pure dephasing.

4.2

Dephasing in addition to relaxation

In addition to relaxation, where excitations are transfered from the system to the modes of the environment, pure dephasing is also possible, in

4.2. DEPHASING IN ADDITION TO RELAXATION

23

which the populations are unaected but their coherence is reduced. In considering this kind of dephasing, it is useful to also broaden our view from considering a single two level system, to describing measurements on an ensemble of many two level systems. For the moment we will assume that even if there are many two level systems, they all act independently. As such, the expectation of any measurement performed on this ensemble can be given by: X =
i

T r(i X ) = T r( X ),

where

=
i

i .

(4.11)

We thus have two types of decoherence one may consider: Broadening within a two-level system, coming from time dependent shifts of energies (and possibly also coupling strengths). A variety of sources of such terms exist; examples include collisional broadening of gaseous atoms, where the rare events in which atoms approach each other lead to shifts of atomic energies. In solid state contexts, motions of charges nearby the articial atom can also lead to time dependent noise. Inhomogeneous broadening, where the parameters of each two-level system vary, and so o-diagonal matrix elements, involving time dependent coherences, are washed out in . This is particularly an issue in solid state experiments, where the two-level system energy may involve the dimensions of a fabricated or self-assembled articial atom, and these dimensions may vary between dierent systems. To describe both such eects together, we consider adding noise, + (t). (4.12)

This noise will then lead to a decay of the o diagonal correlations. To see this, consider Eq. (4.6), with the above replacement, and then perform a gauge transform so as to remove the explicit time dependence. One nds:
t

P (t) P (t) exp i


0

(t )dt

(4.13)

We will take (t) = 0, and two distinct limits for the correlations of . Fast noise limit In order to extract a tractable model of time-dependent noise, the simplest limit to consider is that (t) has a white noise spectrum, i.e. it is Gaussian correlated with: (t) (t ) = 20 (t t ). (4.14) This assumption is equivalent to assuming that the noise is fast, i.e. that any correlation timescale it does possess is suciently short compared to the dynamics of the system that it may be neglected.

24

LECTURE 4. POWER BROADENING

With this assumption of instantaneous Gaussian correlations, one may then expand the exponential in Eq. (4.13) and write:
t

P (t) P (t) 1 + i
0

dt (t )
t 0

1 2!

dt dt (t ) (t ) + . . .
0 t 2

20 =P (t) 1 2!

(20 )2 dt + 3 4!

dt
0

+ . . . = P (t)e0 t (4.15)

with the combinatoric factors coming from the number of possible pairings in the Wick decomposition of n . This noise thus leads to an enhanced decoherence rate for the o diagonal terms, so that /2 should be replaced by 1/T2 = (/2) + 0 in the density matrix equations. One may equivalently dene the relaxation time 1/T1 = . In general T2 < T1 , as dephasing is faster than relaxation, however the factors of two in the above mean that the strict inequality is T2 < 2T1 , with equality holding when there is only relaxation. The Bloch equations are thus given by: t X = Y 1 X T2 1 Y T2 1 Z+ 2 (4.16) (4.17) . (4.18)

t Y = +X + gZ t Z = gY 1 T1

A more rigorous approach to deriving the eect of dephasing is discussed in question A3, in analogy to the way we derived decay in the density matrix formalism

Static limit (inhomogeneous broadening)


To describe inhomogeneous broadening, should represent the systemdependent variation of energies, and should have no time dependence. The same behaviour is also relevant if time-dependent noise terms vary slowly on the timescale of an experiment. (For example, slow noise may arise from a nearby impurity have multiple stable charging states, so that the charge environment can take multiple dierent values; if the charging and discharging of this impurity is slow compared to the experiment, this energy shift will eectively give inhomogeneous broadening when averaged over multiple experimental shots.) For this case is static, but randomly distributed, and Eq. (4.13) becomes: P (t) P (t) dp( )eit . (4.19) If we consider a Lorentzian probability distribution for with width 1 , 2 + 2 ), then one nds P (t) P (t) exp( |t|). Alteri.e. p( ) = 21 /(1 1 natively, a Gaussian probability distribution leads to a Gaussian decay of correlations.

4.3. POWER BROADENING OF ABSORPTION

25

If one has both noise terms as in the previous section, and inhomoge, neous broadening, one may then distinguish T2 as dened above, and T2 where 1/T2 = (/2) + 0 + 1 .
Distinguishing T2 and T2

Since both 0 and 1 lead to decay of the coherence functions, a simple . However, the experiment measuring coherence will see the lifetime T2 dynamics of each individual two level system remains coherent for the longer timescale T2 . Such extra coherence can easily be seen in any suciently non-linear measurement. An example of how the dierence of T2 and T2 can be measured is given by photon echo, illustrated in Fig. 4.1.
Z Z Z g t= Z

X Y

X Y Y

X Y

g t=/2

Figure 4.1: Cartoon of photon-echo experiment, which removes


dephasing due to inhomogeneous broadening, leaving only true dephasing rate to reduce intensity of revival.

From the Bloch equations, Eq. (4.164.18), it is clear that a resonant pulse with duration gt = /2 brings the Bloch vector to the equator. From there, inhomogeneous broadening means that (in a rotating frame at the mean frequency) the Bloch vectors for each individual two-level system start to spread out. However, by applying a second pulse with gt = , a rotation about the X axis means that whichever Bloch vector spread at the fastest rate is now furthest from the Y axis, and the subsequent evolution sees the vectors re-converge. At the nal time, the coherence of the resulting pulse will have been reduced by 0 and , but 1 will have had no eect.

4.3

Power broadening of absorption

Having introduced the idea of extra dephasing terms, we may now include these in a re-examination of the question we started an earlier lecture with what is the eect of shining classical light on a two-level atom? However, we may now account for decay and dephasing. Whereas previously the absence of decay meant that the result was Rabi oscillations, with decay a steady state is eventually found. Looking for steady state solutions of Eq. (4.164.18) one nds from the equations for X and Z that: Y = 1 gT1 Z+ 1 2 , X = T2 Y = T2 gT1 Z+ 1 2 . (4.20)

26

LECTURE 4. POWER BROADENING

Substituting these into the equation for Y , and recalling Pex = Z + 1/2, gives: T2 1 1 g 0 = Pex + + g (4.21) gT1 T2 gT1 2 After re-arranging, one then nds: Pex = T1 T2 (g)2 (T1 /T2 )(g)2 1 1 = 2 2 (T2 )2 + 1 + T1 T2 (g)2 2 2 + [1 + T1 T2 (g)2 ] /T2 (4.22)

The addition of damping to the equations of motion has thus had a number of important consequences: It is responsible for a steady state existing at all (but note that this steady state is in a frame rotating at the pump frequency , so the physical polarisation is time dependent.) It gives a nite width to the resonance, even at weak pump powers in the Hamiltonian case, as g 0, the width of the resonance peak vanished. It modies the overall amplitude at resonance, by a factor depending on the power, such that at g 0, there is no response. An important feature of this excitation probability is that the linewidth depends on the intensity of radiation. This is a consequence of the nonlinearity implicit in a two-level system. Such dependence is not so surprising given that in the Hamiltonian dynamics, the resonance width already depended on the eld strength. One may note that if one considered an harmonic atomic spectrum, rather than a two-level atom, no such broadening would be seen (see question A4).
0.5

0.4

0.3 Pex 0.2 0.1 0 -30

-20

-10

10

20

30

Figure 4.2: Power broadening of absorption probability of excitation vs atom-photon detuning for increasing eld strength. This problem, of a single atom coherently pumped, and incoherently decaying also shows interesting behaviour if we look at its incoherent emission spectrum. This is however something we will postpone until lecture 7.

4.4. FURTHER READING

27

4.4

Further reading

The discussion of the resonance uorescence problem in particular can be found in chapter 4 of Meystre and Sargent III [10].

Lecture 5

Two-level atom in a cavity: Purcell eect and Cavity QED


In the previous few lectures we have discussed a single two-level atom coupled either to a continuum of radiation modes, or considering a cavity, which picks out a particular mode. In this lecture we will consider more carefully the situations in which a cavity will pick out a single mode, by describing how the system changes as one goes from no cavity, via a bad cavity (which modies the eld strength but does not pick out a single mode) to a good cavity, which can pick out a single mode. To study this problem throughout this crossover, we will consider a two level system coupled both to a cavity pseudo-mode (which itself decays), and also coupled to non-cavity modes, providing incoherent decay. We will show how in this model, a bad cavity can lead to either enhanced decay, and a good cavity can lead to periodic exchange of energy between the cavity and the two-level system, i.e. the Rabi oscillations we previously discussed for a perfect cavity. Before discussing the crossover between bad and good cavities, we rst investigate a toy model of a 1D cavity, to put into context the meaning of the pseudo-mode description of the cavity mode.

5.1

The Purcell eect in a 1D model cavity

The aim of this section is to describe how a cavity modies the rate of decay of a two-level system, coupled to one-dimensional radiation modes. Following the earlier discussion of system-reservoir coupling, the decay is characterised by the combination of reservoirs density of states and the coupling of each reservoir mode to the atom. This gave an expression for the decay rate of the form: ( ) = 2
k

( ck ) 29

gk 2

(5.1)

30

LECTURE 5. PURCELL EFFECT

Recalling that the coupling strength gk depends on the mode structure via gk Ek , where Ek is the eld strength associated with a single photon, we can nd the eect of the cavity on ( ) by determining the spatial proles of the modes of the full system, and inserting these into the above sum.
L/2 a/2

Figure 5.1: Toy model of cavity, length a, in quantisation volume


of length L.

We will consider a cavity of size a, embedded in a quantisation volume of length L, where we will take L later on. (See gure 5.1). Our aim is to nd the spatial prole of the electric eld modes, and thus to calculate ( )
k

( ck )|Ek (x = 0)|2 .

(5.2)

The electric eld obeys the wave equation: d2 Ek (x) = k 2 r (x)Ek (x), dx2 (5.3)

where r (x) is the (spatially varying) dielectric constant, and k is the wavevector in vacuum. Both in the cavity, and outside, the modes will be appropriate combinations of plane waves. Restricting to symmetric solutions (as antisymmetric solutions will vanish at x = 0), and matching the boundary at x = L/2, we may write: Ek (x) = f (x) Ak cos(kx) g (x) Bk sin k |x| |x| < |x| >
a 2 a 2

L 2

(5.4)

As L , the normalisation condition will approach the simple result Bk = 2/L, since the normalisation integral will be dominated by the parts outside the cavity. We now need to nd the eect of the imperfect barriers on this mode in order to relate Ak , Bk , and thus nd the quantisation condition specifying k , and the mode amplitude, Ek (x = 0) = Ak . As a simple model, let us consider a varying dielectric index, = 0 [1 + (|x| a/2)], giving the equation: d2 a Ek (x) = k 2 1 + |x| 2 dx 2 Ek . (5.5)

Let us focus on x > 0; then, given the form of Ek (x) = (a/2 x)f (x) + (x a/2)g (x) with f = k 2 f and g = k 2 g , we may write: d2 a Ek = k 2 Ek + 2 x dx2 2 g (x) f (x) + x a [g (x) f (x)] 2 (5.6)

5.1. THE PURCELL EFFECT IN A 1D MODEL CAVITY Thus, to solve Eq. (5.5) we require that:

31

a a a a k 2 a a =g , g f = f +g (5.7) 2 2 2 2 4 2 2 Substituting the forms of f (x), g (x) into this equation, one may eliminate Ak , Bk to give the eigenvalue condition: f cos kL 2 + k cos 2 ka 2 sin ka kL 2 2 = 0. (5.8)

For nite L, this provides a restriction on the permissible values of k . As L , the permissible values of k become more dense, such that, in this limit, one may instead consider this equation as depending on two separate values kL and ka; if L a, it is possible to signicantly change kL without modifying ka. Thus, we will instead rewrite this equation as a condition that species the value of kL given a xed ka: cos kL 2 1+ k sin 2 ka 2 cos ka 2 = sin kL 2 k cos2 2 ka 2 .

(5.9) Using this condition, one may then eliminate kL, and write Ak in terms of Bk = 2/L and functions of ka, i.e.: Ak = 2 L 1 1 + k sin(ka/2) cos(ka/2) + (k/2)2 cos2 (ka/2) (5.10)

If we dene = k/2, then we may use this formula to give the eective decay rate: k= c = = 0 2 1 + sin(ka) + 1 2 (1 + cos(ka)) 0 1+
2 2

1+

2 4

tan 0 =

cos(ka 0 )

2 . (5.11)

The form of this function in general is shown in Fig. 5.2. For small k , the eect of the cavity is weak (i.e. k 1), and so there is little modication compared to the result without the cavity. For larger k , there are sharp peaks, which can be described by an approximately Lorentzian form (see the inset of Fig. 5.2). This Lorentzian form can be found by expanding Eq. (5.11) near its peaks, which are at: k0 a = 0 + (2n + 1). We can then expand ( ) near 0 = ck0 . If its k dependence, then we nd: 0 1 1+
1 2 2 2 2 2

(5.12) 1, and if we may neglect

1+ .

2 2 2 4

1 2

a( 0 ) c

2 a2 ( 4c2

0 )2

(5.13)

32

LECTURE 5. PURCELL EFFECT

25 100 20 15 10 5 0 0 1 2 3 4 /0 5 6 7 8 10 1 ()/0 0.1 0.01 7.2 7.6 8

Figure 5.2: Eective decay rate vs frequency for the toy model of
a 1D imperfect cavity. Inset compares the exact result, Eq. (5.11) to the Lorentzian approximation of Eq. (5.14)

Let us introduce the full width half maximum, , such that this expression becomes: P (/2)2 (5.14) = 0 ( 0 )2 + (/2)2 which leads to the denitions: 2c = 2 , 2 a P = 2 = 4c . a (5.15)

Here 0 is the result for a 1D system in the absence of a cavity. The factor P describes the maximum value of /0 at resonance. At resonance, the presence of the cavity enhances the decay rate, and P is the Purcell enhancement factor. It is helpful for comparison to later results to rewrite this as: 0 c 2 0 P =4 = Q , Q (5.16) 0 a a where Q is the quality factor of the cavity mode. The enhancement on resonance therefore depends both on the quality factor, and on the mode volume. Away from resonance, the decay rate decreases instead of increasing, since there are no modes with signicant weight inside the cavity; the minimum value of /0 occurs when k0 a = 0 + 2n , and is given by /0 = 1/2 = 1/P , hence for this 1D case, P describes both the enhancement on resonance, and the reduction o resonance. As increases further, the Lorentzian density of states becomes sharply peaked, and so the Markovian approximation for decay of the atom no longer holds, instead one can use: d = dt =
t

dt (t )
t

d 0 P (/2)2 ei( )(t t) 2 ( 0 )2 + (/2)2


)(t t)(tt )/2

0 P 4

dt (t )ei(

(5.17)

5.2. WEAK TO STRONG COUPLING VIA DENSITY MATRICES 33 This form describes the behaviour one would see if the two-level system were coupled to another degree of freedom, with frequency and decay rate /2. Hence, we are led in this limit to introduce the cavity pseudomode as an extra dynamical degree of freedom. The density of states implies that the cavity mode cannot be treated as a Markovian system, it has a non-negligible memory time 1/, and so we must consider its dynamics.

5.2

Weak to strong coupling via density matrices

This section considers the case in which a single cavity mode must be treated beyond a Markovian approximation, by considering the JaynesCummings model, along with relaxation of both the two-level system and the cavity mode. The model is thus given by: d = i [H, ] a a 2aa + a a dt 2 + 2 + + + 2 g H= + a + a + z + 0 a a. 2

(5.18) (5.19)

One should note that in this equation, a is the creation operator for a pseudo-mode of the system, i.e. it does not describe a true eigenmode; the true eigenmodes are instead superpositions of modes inside and outside the cavity, just as in the previous section. Because the pseudo-mode overlaps with a range of true eigenmodes, the probability of remaining in the pseudomode will decay (see question B1 for the 1D case). This coupling between cavity modes and modes outside the cavity can be described in a Markovian approximation, leading to the decay rate in the above density matrix equation. In addition to the decay of the pseudo-mode, we also include a rate describing decay of the two level system into modes other than the cavity mode. In the one dimensional description considered previously, such a channel must correspond to non radiative decay. However, in three dimensions, if the cavity is not spherical, then decay into non-cavity directions is possible and is described by . In the three dimensional case, one may consider = (1 /4 ), depending on solid angle the cavity encloses. For the experimental systems discussed below, this is generally the case, due to the need to have access to insert atoms into the cavity. More generally describes the possibility of relaxation into any mode other than the cavity mode; in solid state contexts, other possible reservoirs often exist (e.g phonons, other quasiparticle excitations etc). Let us solve the equation of motion, Eq. (5.18), starting in an initially excited state | , 0 . From this initial state, the only possible other subsequent states are | , 1 and | , 0 . As the last of these states cannot evolve into anything else it may be ignored, and we can write a closed set of equations for three elements of the density matrix: PA = 0,0 , PB = 1,1 , CAB = 0,1 (5.20)

34

LECTURE 5. PURCELL EFFECT

By taking appropriate matrix elements, one nds: d g PA = i (CAB CAB ) PA dt 2 d g PB = i (CAB CAB ) PB dt 2 g + d CAB = i( 0 )CAB + i (PA PB ) CAB dt 2 2 (5.21) (5.22) (5.23)

One may further simplify these equations by noting that PA PB |CAB |2 = 0 is conserved, and thus writing PA = ||2 , PB = | |2 , CAB = . Substituting this leads to the simpler equations: d dt = /2 ig/2 ig/2 i /2 (5.24)

where we have written = 0 . In general this leads to two frequencies for decaying oscillations of the excitation probability: i 2 i + i g2 + = 0. 2 4 (5.25)

Bad cavity limit Purcell eect in 3D


In the limit of a bad cavity, i.e. g, , the two frequencies corresponding to the above equation correspond separately to decay of the excited twolevel system and decay of photons. Because these are on very dierent timescales, one may treat the coupling perturbatively writing = i /2 + , where at O( ) the determinant equation becomes: i i and since + 2 2 + g2 = 0, 4 (5.26)

this bad cavity limit gives: = i 2 + g 2 2 + 42 + g 2 . 2 + 42 (5.27)

This describes a cavity-enhanced decay rate, e = +cav , where the cavity decay rate describes the Purcell eect, as discussed above. On resonance we have: g 21 (5.28) cav = 4 2 and recalling the earlier 3D results: =
3 |d |2 0 ab , 30 c3

g 2

0 |dab |2 2 0 V

(5.29)

(where we have assumed resonance, so 0 = ), one may write: cav 3c3 1 3 =P =4 2 = 4 2 Q 2V 0 3 V . (5.30)

5.3. FURTHER READING

35

This closely resembles the one dimensional result before; the Purcell enhancement depends both on the quality factor, and also on the mode volume. An estimate of the minimum decay rate o resonance in this case yields cav,min = cav,max
20 2

3 1 16 2 Q

3 V

. However, if the back-

ground decay rate also exists, this o-resonant reduction is not as relevant as the on-resonant enhancement.

Strong coupling Rabi oscillations


If the cavity is suciently good, then rather than decay, the excitation probability oscillates. In the resonant case of = 0, one can write the general solution for the oscillation frequency as: ( + ) = i 4 g2 4 4
2

(5.31)

In order that Rabi oscillations exist, it is thus necessary that g > ( )/2. In order that they should be visible before decay has signicantly reduced the amplitude, it is necessary that one has g > , . In this limit, one has strong coupling, and decay is indeed very strongly modied.

5.3

Further reading

For a discussion of the Purcell eect, and the crossover between weak and strong coupling, see e.g. Meystre and Sargent III [10].

Lecture 6

Features of various of Cavity QED systems


Having introduced all of the concepts required to understand a single atom coupled to radiation in a cavity, this lecture now discusses a number of systems which can realise this situation. The aim here is to understand what parameter ranges can physically be achieved in dierent systems, to understand the advantages and disadvantages of the dierent systems, and to understand the physical origin of the various coupling strengths and decay rates. A summary of the characteristic values is given in table 6.1. The discussion below explains the origin and signicance of these numbers. In each case there are several stages to consider: the nature of the two-level system, and the energies, decay rates, and coupling strengths of the transition involved. In addition, one must also consider what designs of cavities are compatible with the transition and nature of the two-level system.

System /2 /2 g/2 /2 3 /V Q Other tmax


a b

Atom[1, 11] Atom[4] Optical Microwave 1 MHz 1 kHz a 6 MHz 30 Hz 10 MHz 50 kHz 390 THz 50 GHz 5 10 101 8 10 108 tight 100s

SC qubit[3] Microwave 30 MHz 3 MHz b 100 MHz 10 GHz /a = 1 104 1/T2 3MHz

Exciton[5] Optical 30 GHz 0.1 GHz 100 GHz 400 THz 101 104 1/T2 3GHz

Recent work reaches 10Hz[12] Not well known, dominated by cavity induced decay

Table 6.1: Characteristic energy scales for dierent cavity QED


realisations, as discussed in the text.

37

38

LECTURE 6. CAVITY QUANTUM ELECTRODYNAMICS

6.1

Optical transitions of atoms

We start by considering the conceptually simplest case, in which the two level system is a transition between two low lying levels of a real atom. Although real atoms have many levels, the level spacing is anharmonic (E Ryd/n2 ), hence it is possible for the cavity to be near to resonance only for a single transition between the ground and excited states. This is only possible because the decay rate / f 3 ( /Ryd), so that the natural linewidth is much smaller than the transition. Such transitions involve light at optical frequencies, and so require cavities designed to produce high reectivity at these frequencies.

Atomic energy levels; ne structure, hyperne structure and linewidths


Considering optically allowed transitions m = 0, 1, l = 1, the simplest spectra will correspond to alkali atoms, involving s p transitions. However, to properly understand these transitions, one must consider both the ne structure, and the hyperne structure. Fine structure arises due to the spin-orbit coupling: ESO = 1 2m2 c2 1 dV r dr SL (6.1)

The order of magnitude of this expression is: ESO


2 1 Ryd2 e2 2 Ryd. 2 mc2 2ma2 4 a mc 0 B B

(6.2)

The power of 2 indicates that this is much smaller than the principal electronic level splitting, but still bigger than the linewidth, hence it can be resolved. The spin orbit term increases with atomic number due to the larger potential gradient that exists at small radii within the screening cloud. Since the term produces an L S coupling, its eigenstates are given by the quantum numbers of |L|, |S|, |J| and one other commuting observable, e.g. Jz . Thus, for alkali atoms, the s state is not split, but the p state is split into J = 1/2, 3/2 states. Hyperne splitting arises due to coupling between the electronic angular momentum J and the nuclear angular momentum I. The coupling I J means that rather than using Jz , Iz to label eigenstates, one should consider eigenstates of |F|, Fz with F = J + I. A level scheme for 87 Rb is shown in Fig. 6.1, based on values in Steck [13]. 87 Rb is a boson, as it has an odd number of electrons and an odd number of nucleons. Its ground state nuclear spin is |I| = 3/2, hence the |J| = 3/2 state splits into four levels |F| = 0 . . . 3. The notation F = 3 indicates an excited electronic state with hyperne state F = 3, whereas F = 2 indicates a ground electronic state with hyperne state F = 2. The transition between the ne structure level S1/2 P3/2 is known as the D2 line. The dominant transition within this line is the F = 3 F =

6.1. OPTICAL TRANSITIONS OF ATOMS

39

P3/2
7THz

266MHz

F=3 F=2 F=1 F=0

0.5GHz

P1/2

F=2 F=1

1.6eV=390THz=780nm

S1/2
7GHz
8

F=2 F=1

Figure 6.1: Level scheme for 7Rb (bosonic), showing ne and hyperne structure. For further details, see[13]

. Using the transition 2. This has a dipole matrix element dD2 = 1.3eA energy ED2 = 1.6eV, one nds fD2 = 0.7, so this transition is close to saturating the oscillator sum rule. Using this transition strength, the above estimate of decay rate gives: 1 1.6eV 1 350THz 0.7 3 3 137 13.6eV 3MHz (6.3)

This estimate is smaller than the actual width, = 38MHz because it does not properly include summation over possible spin states. However, this linewidth is rather small, and so it is possible to resolve even hyperne structure, given the scales in Fig. 6.1. To have a true two-level system, one has two choices: either to avoid any stray magnetic elds, and rely on the fact that the cavity mode polarisation will couple to a particular superposition of the 2F + 1 = 7 levels labelled by mF = 3 . . . 3. Alternatively, this remaining degeneracy could be split by a magnetic eld.

Optical cavities
From the above considerations, in order to have a cavity couple to a an individual hyperne level transition, the cavity linewidth must also be smaller than the hyperne splitting, i.e 10MHz. Given the transition frequency of E = 390THz, this implies a cavity quality factor of Q 108 is required. One may also consider the quality factor required for Purcell enhancement. The minimum size of the cavity is constrained by the need to be able to controllably insert atoms, and a typical cavity length[1, 11] is L 200m,

40

LECTURE 6. CAVITY QUANTUM ELECTRODYNAMICS

and beam waist w 20m. This gives a ratio of wavelength to mode volume of 3 /V 105 . Recalling the condition that controls the Purcell eect, this means again that a very high Q factor, Q 108 is needed to reach strong coupling. In the following we consider two possible approaches; metallic and dielectric mirrors, and see which is capable of producing such high quality factors. In relating the quality factor the reectivity, it is important to note that they depend dierently on the mode index involved. In particular, if one introduces the nesse F , then: n Q= = , F= R 1R (6.4)

where is the mode spacing (which depends on the length of the cavity), and R is the intensity reection coecient of the mirrors. The above approximation holds for good mirrors. For very good mirrors, i.e. T = 1R 1, one has Q = nF n/(1 R), and the cases discussed below will be in this range. For a cavity with length L 200m, the mode index will be n 250, and so F 4 105 would be sucient. Metallic mirrors As a brief reminder, the reectivity of a metallic mirror can be found by considering a plane wave incident on its surface, and nding the reection and transmission coecients associated with the eective (frequency dependent) dielectric constant arising from its conductivity. For a plane wave with frequency , one has: B = 0 ( + ir 0 )E (6.5)

Thus the refractive index n = r i/0 and for a good conductor at low enough frequencies n = (1 i) /20 . Then, matching electric and magnetic elds at the interface: Ei + Er = Et Bi + Br = Bt 1+r = t 1 r = nt (6.6)

so r = (n 1)/(n + 1). Thus, for a good conductor one nds: R=12 20 (6.7)

For a typical metal, with 107 1 m1 , and optical frequencies = 400THz, this gives R = 1 0.05, hence F = 60. In order that the quality factor reach the required value, one would have to use a very high index (i.e. a very large cavity), with n 107 . For such a cavity, the mode volume would be huge, and so strong coupling unattainable. Clearly metal cavities are not relevant.

6.1. OPTICAL TRANSITIONS OF ATOMS Dielectric mirrors: Bragg reectors

41

Much higher quality mirrors for optical frequencies can be made using dielectric Bragg mirrors. These consist of alternating layers of materials with dierent dielectric constant, with spatial period /4, so that even if the dielectric contrast between adjacent layers is not sucient to cause strong reection, the interference of multiple reections will lead to strong overall reection. This is illustrated in Fig. 6.2.
1 0.8 0.6 0.4 0.2 0 0 0.5 1 /0 1.5 2 26 layers na=1.5, nb=2.5

Cavity

Figure 6.2: Left: Cartoon of Bragg mirror cavity. The two


colours indicate material of two diering refractive indices. Right: reectivity spectrum of a Bragg mirror

The reectivity of a Bragg mirror can be found using the transfer matrix approach, discussed in the appendix to this lecture. At resonance (i.e. at a frequency such that wa na = wb nb = (c/ )(/2), where wa,b are layer thicknesses and na,b the refractive indices), the reection coecient is given by: 1 (nb /na )N r= . (6.8) 1 + (nb /na )N where N is the number of pairs of layers. For the case illustrated in Fig. 6.2, with a dielectric contrast nb /na = 5/3, N = 13 layers is sucient to reach a nesses of F = 4105 . Because the overall reection depends on destructive interference between the partial reections at each interface, the reectance is high only over a range of frequencies, known as the stop band. It is therefore important that the frequency of all normal modes of the coupled cavity QED system lie within this stop band if one wants to ignore the frequency dependence of the mirrors. Other considerations The background atomic decay rate is determined by the intrinsic properties of the atomic transition, reduced by the geometry of the cavity, and so this sets the scale that must be overcome for strong coupling. Typical values of the corresponding g, , are given in Table 6.1. In addition to the timescales included within Eq. (5.18), in some experimental systems, there is another timescale, that of the time for an atom to leave the cavity, however this is typically larger than all other timescales.

R=|r|2

42

LECTURE 6. CAVITY QUANTUM ELECTRODYNAMICS

6.2

Microwave transitions of atoms

Remaining with atoms as the matter part of the Jaynes-Cummings Hamiltonian, some of the problematic features above can be removed if one considered instead microwave frequency transitions of the atoms. This increases the wavelength, allowing cavities comparable to the wavelength, and can also signicantly improve the cavity quality. For microwave frequencies, superconducting cavities can be used (since the frequency is less than the BCS gap, BCS THz, so the superconducting mirrors are diamagnetic at these frequencies. The remaining limitation on cavity quality instead comes from scattering o the mirror surface, and any gaps required in the cavity for atom injection. For niobium cavities operating at 50GHz ( 6mm), values of 3 /V 0.1 and Q 1010 are possible[12]. The relevant atomic transitions at these frequencies are transitions between highly excited atomic states (Rydberg states). For states with n, l 1, the outermost electron remains far from the nucleus and so sees only the screened charge of +1, so the energy levels of such orbits are almost Hydrogenic. For large n but l 0, the orbit is perturbed by the enhanced potential near the nucleus, which can be incorporated by a quantum defect, l to write: Ry En,l , (6.9) (n l )2 with l 0 for large l. It is thus clear that transitions between n, n 1 will have small energies, En 1/n3 , and so for n 50 one can have microwave frequency transitions. Atoms can be prepared in such states by using tuned laser pulses.

(a)

h
(b)

4 3

4 3

2 2
Figure 6.3: Cartoon of atomic level scheme, and transitions from
a highly excited state with (a) l n, or (b) l = n 1. Note that the actual values used are of the order n, l 50, much larger than shown here.

6.2. MICROWAVE TRANSITIONS OF ATOMS

43

As well as allowing better cavities at these frequencies, Rydberg atoms also have signicantly reduced spontaneous decay rates without sacricing coupling to the cavity mode. The total decay rate from atomic state nl is given by: tot
nl 3 2 nl,n l |dnl,n l | .

(6.10)

Selection rules imply that l = l 1, which means that two dierent limits of decay rates exist, depending on whether l = n 1 or l n, as illustrated in Fig. 6.3. l n In this case, decay to almost all other values n < n is possible (see arrows (a) in Fig. 6.3). We may consider the characteristic decay rate to nearby levels n n, or to the ground state n = 1. Decay to n n has 1/n3 , and the dipole matrix element depends on the characteristic size of the orbit for large n, i.e. d n2 (this holds because both initial and nal states have similar extensions). These combine to give n n n5 . Alternatively, for transitions to the ground state, is n independent, being more or less xed at the the Rydberg energy. In this case the dipole matrix element is however much smaller, as it now involves the overlap between a large Rydberg state and a much smaller ground state orbital. Thus, the dipole matrix element depends on the overlap d 1/ n3 , giving n 1 n3 . As such, the latter process dominates for large n, and so the overall decay rate is a factor n3 slower than for low excitation states. l n 1 In this case, the only transition allowed by selection rules is to n = n 1. Thus, this has the decay rate in the rst part of the above case, n5 and is yet even slower than for l n. These states with l = n 1 are known as circular Rydberg states, as they correspond closely to classical circular atomic orbits, as expected from the correspondence principle.

While the decay rates are signicantly reduced by the reduced energy of nearby transitions, and reduced overlap of remote transitions, the coupling to conned radiation for nearby transitions is not reduced in the same way. This is because the continuum density of states does not enter the calculation of g . Instead, one has g d for resonant transitions, giving g n, which increases with n. However, compared to optical transitions, the value of g is signicantly reduced because g 1/ V , and the cavity has V 3 . Thus, the value of g in table 6.1 is reduced compared to optical transitions, but is reduced by a smaller factor than the reduction of the atomic decay rate.

44

LECTURE 6. CAVITY QUANTUM ELECTRODYNAMICS

6.3

Superconducting qubits in microwave resonators

Continuing with microwave frequencies, one can also consider superconducting qubits coupled to microwave cavities. These have the advantage of considerably increasing the coupling strength, by (in this case) increasing the number of electrons involved in the articial atom. There are several possible superconducting qubit schemes that have been considered, the following will focus on the transmon qubit[14], where the qubit eigenstates corresponds to quantised modes plasmon oscillations, modied by the transmission line. In choosing a qubit scheme, the intention is to minimise noise (dephasing) by reducing sensitivity to the environment, while at the same time allowing strong coupling to the cavity. This concern about sensitivity to the environment is an unavoidable consequence of considering solid state systems for the articial atom, and is much more signicant for solid state systems than it is for isolated atoms. In addition, as with the previous systems, it is also necessary to have strong enough anharmonicity to allow restriction to two levels.

Qubit eigenstates

Waveguide

Cin Lr Cr V

Cg CQ Cg EJ EJ

Superconductor

Figure 6.4: Schematic diagram of superconducting qubit capacitively coupled to a stripline microwave resonator (left), and the equivalent circuit (right), coloured for comparison of diagrams after Koch et al. [14].

Figure 6.4 illustrates the transmon qubit design, and its coupling to a microwave stripline resonator. The equivalent circuit shows that the articial atom consists of a pair of Josephson junctions, shunted by a large capacitance. EC ). Hatom = ( n nbias )2 EJ cos( (6.11) 2 describe the number of Cooper pairs transfered between The operators n , the two islands, and the phase dierence respectively. The charging energy EC depends on an eective capacitance of the islands via EC = (2e)2 /C , and the bias nbias depends on both DC bias between the transmission lines, and the AC eld due to microwaves on the transmission line. The eective Josephson energy EJ is controllable by the ux through the SQUID loop. i.e., for a phase dierence between the blue and red superconducting

6.3. SUPERCONDUCTING QUBITS IN MICROWAVE RESONATORS 45 islands, the energy is: EJ,0 cos 2 20 + cos + 2 20 = 2EJ,0 cos cos() 0 EJ,e cos (6.12)

2 /2, For low energies one may expand to write H = E n nbias )2 + EJ C ( and hence nd a harmonic spectrum En = (n + 1/2) EJ EC . Including the 4 perturbation, the anharmonicity depends on EJ /EC . (In general, the solutions of Eq. (6.11) are Mathieu functions, as can be seen by regarding as a position, and the charging energy as a kinetic energy in a rotating frame.) For the transmon, EJ EC , and so the anharmonicity is relatively small. However, as long as the linewidths are also small, this can be sucient. The reason for choosing small EJ /EC is that it makes the eigenstates very insensitive to any DC bias, and thus insensitive to any charge noise. Despite this insensitivity to DC noise, there can still be strong coupling to photons in the resonator.

Coupling to cavity
This qubit is capacitively coupled to the resonator mode, i.e. the charge difference across the resonator conductors voltage biases the qubit. Including the coupling to the resonator, the Hamiltonian may be written as: H = Hatom + C a a + (a + a ) n, (6.13)

where depends on the eld strength (characteristic voltage for a single photon in the resonator), and the geometry (fraction of this voltage that is across the capacitor). Considering the ground and excited states of the qubit, and introducing the qubit raising operator: 1 + = 2
4

EC EC n i4 EJ EJ

(6.14)

one has 0|n |1 = 1|n |0 = 4 EJ /4EC , 0|n |0 = 1|n |1 = 0. This coupling means that for EC , the perturbation due to static noise is much smaller than the energy level splitting, so is a weak perturbation. However, for cavity mode, one has: He = EJ EC z + C a a + 2
4

EC x (a + a ) EJ

(6.15)

In this case, the relevant scale of the coupling depends on the qubit-cavity detuning, and the eective linewidths, rather than on the energy splitting of the qubit. Table 6.1 shows that strong enough coupling strengths are indeed attainable, while still having g .

46

LECTURE 6. CAVITY QUANTUM ELECTRODYNAMICS

Other considerations
Unlike the microwave system with Rydberg atoms, where atoms eventually leave the cavity and can be measured, the observation of these superconducting systems is typically via the emitted radiation. For this reason is chosen to be larger than physical constraints would require, so that sucient photons escape to allow measurement. The origin of the articial atom decay rate in these experiments is not particularly clear, since coupling to bulk radiation modes should be negligible. However, since the system is in a solid state environment, other degrees of freedom exist that can lead to relaxation. These other degrees of freedom certainly lead to dephasing, so that in these systems, 1/T2 is a signicant rate, arising from charge and ux noise on the superconducting circuit.

6.4

Quantum-dot excitons in semiconductor microstructures

Finally, we consider articial atoms at optical frequencies. One example of this concerns excitons in quantum dots, coupled to semiconductor microstructures. Excitons are bound electron hole states, that can be found by evaluating the eigenstates of: H= p2 p2 e2 e + h 2me 2mh 40 |re rh | (6.16)

The bound states will be Hydrogen-like wavefunctions of the electron hole separation, but with a reduced binding energy, and increased Bohr radius due to the dielectric screening rel 10 and reduced electronic mass. One may write: m0 1 Eex = Ryd , aex = aB (6.17) 2 m0
1 1 where 1 = m e + mh is the reduced mass. In GaAs, the bandstructure of which is indicated in Fig. 6.5 one has me = 0.07m0 , mh = 0.5m0 , = 13 leading to Eex = 5meV.

A() Excitons Electronhole continuum

Figure 6.5: Left: Band structure of an direct gap semiconductor,


e.g. GaAs. Right: Absorption spectrum, indicating bound excitons below the band edge of the electron hole continuum. (Continuum is for a 3D system, neglecting the Coulomb enhancement.)

6.4. QUANTUM-DOT EXCITONS IN SEMICONDUCTOR MICROSTRUCTURES

47

Exciton bound states then appear as a series of Rydberg levels below the band edge associated with creating unbound electron hole pairs. This is illustrated in Fig. 6.5. The optically active transition here is the process of creating or destroying an exciton. The energy for this process is therefore the = Egap Eex , and so despite the small exciton binding energy, the relevant transitions are at optical frequencies. In order to realise a cavity QED system, it is necessary that there should be discrete excitonic states; in bulk there are a continuum of excitonic states with dierent centre of mass momenta. Discrete states can be found by conning the exciton in a quantum dot, as discussed below.

Exciton matrix elements


Before considering conned excitons, we rst discuss the oscillator strength of the transition between states of the semiconductor with and without an exciton. For the delocalised exciton states, it is easier to directly calculate pex = 0|p|ex than to calculate the dipole matrix element d. In terms of this, we have: g e p 2p2 ex = f = ex . (6.18) 2 m m0 20 k V (Note that the operator p measures the physical, rather than crystal, momentum). For an exciton with zero centre-of-mass momentum, in a quantisation volume V , one has: |ex = 1s (r) r r d3 Rd3 r c cv R | c R+ 2 2 V (6.19)

where c c(v ) create electrons in the conduction (valence) band, 1s (r ) is the normalised 1s exciton wavefunction in terms of electron-hole separation, 1/ V is the zero-momentum centre of mass wavefunction, and | represents a lled Fermi sea. In order to evaluate the matrix element, it will be useful to work in momentum space, and write the conduction and valence band creation operators in terms of Bloch states, and nally to revert to a rst quantised notation. In momentum space, the exciton wavefunction becomes: |ex = 1 V =
p

c c,p cv,q |
pq

1s (r) d3 Rd3 r ei(pq)R+i(p+q)r/2 V 1s (r) d3 r eipr = V 1s (p)c c,p cv,p | . (6.20)


p

c c,p cv,p |

The conduction and valence band functions appearing here can be written as: eipr uc,p (r) a c,p = d3 r c (r ). (6.21) V Here uc,p (r) is a periodic Bloch functionfunction, such that uc,p (r + R) = uc,p (r), and normalised per unit cell.

48

LECTURE 6. CAVITY QUANTUM ELECTRODYNAMICS

With the above ingredients, the value of pex can be evaluated by writing out the Slater determinants corresponding to initial and nal states, in which only one electron changed state. One then has: pex =
k

1s (p)

d3 r u (r ) p uck (r). V vk

(6.22)

In general, the exciton wavefunction varies slowly in space, and so the typical values of k obey k kBZ where kBZ is the Brillouin zone boundary. In this case, the periodic Bloch functionfunctions depend only weakly on k , and this dependence can be dropped. Recalling the denition of the Fourier transform used above, one may then write: pex = V (r = 0)pcv pcv =
uc

d3 r d u (r) i Vuc v dr

uc (r).

(6.23)

where uc (r) is the periodic Bloch functionfunction at the Gamma point and the integral is over the unit cell. The quantity pcv is a material property, dependent on the band structure. As one might anticipate, the crucial property of the excitonic wavefunction that determines its dipole moment is the probability that electron and hole are at the same point. Inserting the above into the expression for f one nds: f 2p2 cv | (0)|2 V. mEgap (6.24)

We nd the oscillator strength is extensive, and clearly can be much larger than 1. This is also easy to understand; we are calculating the electronic response of an entire semiconductor, not of a single atom, and so the number of electrons appearing is extensive.

Quantum dots
In order to have a two-level system to consider for cavity QED, we must restrict the number of exciton states that are coupled to the cavity mode. This can be done by conning the exciton in a quantum dot. This is a region of semiconductor with a smaller band gap, such that it is an energetic trap for both electrons and holes, as illustrated in Fig. 6.6. There then exist a number of discrete trapped exciton states, lying below the bulk exciton states. For the exciton conned in a quantum dot, the above calculation of the matrix elements is modied. In the simplest case, where the dot is larger than the exciton Bohr radius, one may consider replacing 1/ V by (R). If the range of (R) is much larger than that of (r), one may approximate: d3 Rd3 r1s (r)(R)ei(pq)R+i(p+q)r/2 and so, writing d3 R(R) f Vdot one nds: (6.26) p,q 1s (p) V d3 R(R) (6.25)

2p 2 cv | (0)|2 Vdot . mEgap

6.4. QUANTUM-DOT EXCITONS IN SEMICONDUCTOR MICROSTRUCTURES


000000000000000000000000000000000000 111111111111111111111111111111111111 111111111111111111111111111111111111 000000000000000000000000000000000000 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111

49

111111111111111111111111111111111111 000000000000000000000000000000000000 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111 000000000000000000000000000000000000 111111111111111111111111111111111111

Figure 6.6: Cartoon of band structure in space for a quantum


dot.

This is no longer extensive, but can still be larger than 1, since the quantum dot contains many electrons that can be involved in the transition.

Cavities for excitons, and properties of excitons as a cavity-QED system


As the quantum dots are xed inside the semiconductor, microstructures can be grown with characteristic sizes comparable to the wavelength of light. To conne light, several options have been used: Bragg mirrors can again be produced. If these are grown by epitaxy of semiconductors, the available dielectric contrast is typically smaller than the SiO2 /TiO2 discussed above. For lattice matched semiconductors, typical values (e.g. GaAs/Al0.3 Ga0.7 As are na /nb 3.3/3.17). A much higher dielectric contrast is available between the semiconductor and air. Thus, photonic crystals cavities, where a defect is introduced into a regular pattern of holes, can produce high quality cavities. Q 6000 has been seen [15]. Dielectric contrast between the semiconductor and air also be used to conne whispering gallery modes in circular resonators. Whispering gallery modes (named after the whispering gallery in St Pauls Cathedral, London) involve rays with a shallow angle of incidence with the edge of the cavity, and so are conned by total internal reection. Q 12, 000 has been achieved this way.[16] A comparison of these approaches is discussed in [5]. Compared to real atoms, the notable improvement in table 6.1 is the signicantly enhanced g ; this arises from the increased Bohr radius, and hence larger dipole matrix element, as well as the much reduced mode volume. Cavity decay rates for systems involving excitons are however typically much larger than cavities used for atoms, due to the lower quality of mirrors in integrated semiconductor heterostructures. The exciton decay rate is much larger than for real atoms due to the solid state environment. Just as for superconducting articial atoms, this again leads both to reduced T1 and also T2 dephasing.

50

LECTURE 6. CAVITY QUANTUM ELECTRODYNAMICS

6.5

Further reading

The discussion of particular cavity QED systems in this chapter is based on various reviews: for Rydberg atoms, see Raimond et al. [4] or Gallagher [17]; for superconducting qubits, see Blais et al. [3]; for excitons in semiconductor microcavities, see Khitrova et al. [5].

6.A

Transfer matrix approach to calculating reectivity of a Bragg mirror

The transfer matrix relates the electric and magnetic eld at position x0 + x to the elds at position x, by building up individual transfer matrices for separate layers of material. Starting from Maxwells equations for plane waves in a one dimensional geometry, and considering solutions: cB = eit n(A+ eikz A eikz ) y (6.27) where = (c/n)k , one may relate A to E, cB at z = 0, and thus write: E cB =
z

E = eit (A+ eikz + A eikz ) x,

cos(kz ) i sin(kz )/n in sin(kz ) cos(kz )

E cB

(6.28)
0

Writing the above matrix as T = T (n, kz ), the transfer matrix for the DBR can be written as T = [T (na , )T (nb , )]N where = ka za = kb zb = (/0 )(/2). i.e., both layers have equal optical depth, chosen to be = /2 at frequency 0 . To nd the transmission and reection coecients, one then solves: t t t =T 1+r 1r , (6.29)

by considering the values of E, cB for an incoming wave from one side. Hence, explicitly: 1 (1 1)T 1 r= (6.30) 1 (1 1)T 1 On resonance, an explicit expression can be found, as: T na , T nb , 2 2 = = 0 i/na ina 0 0 i/nb inb 0 (6.31) (6.32)

nb /na 0 0 na /nb

Thus the reection coecient is given by: r= (nb /na )N (na /nb )N (nb /na )N + (na /nb )N (6.33)

Lecture 7

Resonance Fluorescence
At the end of lecture 4, we found the steady state excitation probability for a two-level system pumped by a coherent light eld. This steady state emerged due to the competition between coherent pumping, and incoherent decay into the continuum of radiation modes. The aim of this lecture is to discuss the spectrum of this uorescence, when the atom is driven near resonance. This is intended both as an illustration of applying the density matrix equation of motion approach to more complex problems, and also to reveal further aspects of the behaviour of a two-level system.

7.1

Spectrum of emission into a reservoir

Our aim is to calculate the spectrum of the emission into the continuum of photon modes. Formally, the spectrum of radiation is given (via the Wiener-Khintchine theorem, applicable to a stationary process, which ours is) by:

I ( ) =

eit E + (t)E (0) = 2


0

eit E + (t)E (0)

(7.1)

where we have divided: E (t) = E + (t) + E (t), E (t) =


k

Ek ak,n (t)en,k ,

(7.2)

with Ek = k /20 V . The rst challenge is to write the time dependent reservoir operators in terms of system operators. If we work in the Heisenberg picture, then this is straightforward, the interaction Hamiltonian was: HSR =
k

gk i(k a ke 2

)t

+ + a k e i(

k )t

(7.3)

and so the solution to the Heisenberg equation for ak (t) can be written as: ak (t) = gk 2
t

dt ei(k

)(t t)

(t ) + ak ().

(7.4)

51

52

LECTURE 7. RESONANCE FLUORESCENCE

We will neglect second term ak (), since we assume the state at t = to be a vacuum, i.e. the only photons are those coming from the atom. In this case we may write: E (t) =
t

dt
k

Ek gk i(k e 2

)(t t)

(t )en,k ,

(7.5)

Because (at least near resonance) Ek gk the sum in the above expression is Markovian as in the previous lecture, and so we may write E (t) (t), hence up to an overall constant, we have:

I ( ) 2
0

eit + (t) (0)

(7.6)

This reduces the problem to that of nding a two-time correlation function of the open quantum system. This is however something we have not yet calculated the density matrix completely describes single time correlation functions, but in general more knowledge is required to nd the correlation of operators at two dierent times. However, for Markovian baths it turns out that it is possible to relate this two-time correlation to the density matrix evolution via the quantum regression theorem.

7.2

Quantum regression theorem

The quantum regression theorem can be stated in terms of the time evolution of single-time correlation functions, governed by: B (t + ) =
i

i ( ) Bi (t) ,

(7.7)

where Bi (t) is some complete set of operators, and the functions i ( ) solve the averaged equations of motion for Bi (t) . Then, the theorem states that one can write: A(t)B (t + )C (t) =
i

i ( ) A(t)Bi (t)C (t) ,

(7.8)

Note that ordering is important, as A, Bi , C are non-commuting operators. In order to prove the theorem, it is rst necessary to state more explore further what is implied by the Markovian approximation.

Signicance of the Markovian approximation


Formally stated, a Markovian system is one where the future evolution is governed by the current state, and not by any history of the system. In the current context, current state means the state of the system, not including the baths. Thus, the assumption is that one can write an equation of motion for the system density matrix that depends only on the current value of the system density matrix. Therefore, the baths must have no memory of previous states of the system.

7.2. QUANTUM REGRESSION THEOREM

53

Generically, without the Markovian approximation, the equation of motion of the system density matrix could be written as dS (t) = i[HS , S (t)] + dt
t

dt (t t )S (t ).

(7.9)

The term (t t ) describes how the state of the bath inuences the system at time t, and depends on the state of the system at the earlier time t . The Markovian approximation is that (t t ) = (t t ). If one describes the bath by a sum over many independent modes, these delta correlated response functions imply a dense spectrum of bath modes; this is what was used in Eq. (3.10). There is another consequence of the Markovian limit, which was also used earlier in Eq. (3.6). The full evolution really depends on SR ; validity of the Markovian approximation then requires that the state of the system is suciently described by S = TrR SR . This means that correlations between the system and the reservoir must be unimportant, and so it is sucient to write SR (t) = S (t) R (t). It is this statement, as we see next, that implies the quantum regression theorem. Under what physical conditions is the Markovian approximation a good description of a real system? Firstly, to have truly delta correlated noise requires a at spectrum of the baths, as we have already assumed. However, a at spectrum is only required for the bath modes that couple almost resonantly to the system (i.e. |j | ). If one considers a reservoir which is not in its vacuum state i.e. with some thermal occupation then a Markovian description of the system requires that not only is the density of states at, but that we may also approximate: nk
k

gk 2

ei(k )t = n (t).

i.e. we require that nk varies slowly for over the range of k for which |j | ; e.g. for a thermal distribution, we require T . This is clearly always an approximation; the consequences of this approximation will be investigated further in the next lecture.

Proof of regression theorem


The Markovian approximation implies that the evolution of single-time expectations can be written as follows, in terms of the unitary evolution U ( ) of system and reservoir: B (t + ) = TrS TrR U ( )B (t)U ( )S (t) R (t) = TrS B (t)TrR U ( )S (t) R (t)U ( ) = TrS B (t)TrR S (t + ) R (t + ) . (7.10)

Here we made use of the cyclic property of the trace, and the relation between evolution of density matrices and that of Heisenberg operators. In

54

LECTURE 7. RESONANCE FLUORESCENCE

an analogous way, the two-time expectation evolves according to: A(t)B (t + )C (t) = TrS TrR A(t)U ( )B (t)U ( )C (t)S (t) R (t) = TrS B (t)TrR U ( ) [C (t)S (t)A(t)] R (t)U ( ) = TrS B (t)TrR [CS A] (t + ) R (t + ) . (7.11) Thus, the only dierence between the right hand sides of Eq. (7.10) and Eq. (7.11) is to replace the initial system density matrix S with the product CS A. However, in the Markovian approximation, we know that Eq. (7.10) can be solved using Eq. (7.7), for any initial density matrix. Thus, one can use this solution to give A(t)B (t + )C (t) =
i

i ( ) Bi (t)

s (t)C (t)S (t)A(t)

=
i

i ( )TrS Bi (t)C (t)S (t)A(t) i ( )TrS A(t)Bi (t)C (t)S (t)


i

(7.12)

which is the desired result in Eq. (7.12). Thus, to nd a two-time correlation in practise, the scheme is this: nd the general time-dependent solution of one of the operators, and how it depends on initial expectation values nd the explicit form of Eq. (7.7), and then insert the second operator(s) in these initial expectations.

7.3

Resonance uorescence spectrum

From the quantum regression theory, we now have that:

I ( ) 2
0

eit Tr + (t)

(t=0)= 0

(7.13)

i.e. we should solve the density matrix equation of motion to nd (t) given the initial condition 0 , where 0 is the equilibrium density matrix. Since the equation of motion for the density matrix is a rst order dierential equation, t = M [], where M is a linear super-operator, we can easily solve by Laplace transforming to give: M [(s)] s(s) = 0 , (7.14)

where the right hand side encodes the initial conditions. From the Laplace transform (s), the Fourier transform corresponding to the intensity spectrum is then given by: I ( ) 2 Tr + (s = i + 0) . (7.15)

7.3. RESONANCE FLUORESCENCE SPECTRUM

55

Let us now specify the various parts of this for the case of resonance uorescence, i.e. = = 0. WE will work in the rotating frame as in the previous lecture. The steady state condition then becomes: 0 = 1 2[1 + T1 T2 (g)2 ] T1 T2 (g)2 iT2 (g) iT2 (g) 2 + T1 T2 (g)2 (7.16)

(making use of X = 0 and Y, Z as calculated in the previous lecture.) The important part of the spectrum however comes from the superoperator in Eq. (7.14). If we write density matrices as a vector: = = (7.17) , then in this four dimensional space, the super-operator can be written as a 4 4 matrix, following Eq. (4.5) 1/T1 ig/2 ig/2 0 ig/2 1/T2 0 ig/2 . M = (7.18) ig/2 0 1/T2 ig/2 1/T1 ig/2 ig/2 0 One can thus nd the full density matrix as:

0 0 + 1 (0 0 1 0)[ M ( i + 0 ) 1 ] I ( ) 2 2 T1 T2 (g) . (7.19) +iT2 (g)

When inverting the 4 4 matrix, there will in general be four poles. However, because the density matrix evolution is trace-preserving, one of these poles will necessarily be at zero, corresponding to the conserved quantity. The corresponding eigenvalue will be the steady state (the trace-preserving property ensures that the steady state is non-trivial). Since we know that in the steady state there is a non-zero polarisation, the weight of this zero pole will not in general vanish. This means there is a delta-function peak in the spectrum. Such a peak describes the elastic scattering of pump photons into the radiation modes. The remainder of the poles will describe inelastic scattering. To nd these poles, we should solve: 0 = Det [M + i ] = i i 1 T2 i 1 T2 i 1 T1 + (g)2 . (7.20) This has the zero mode as discussed above, and three other poles: = i , T2 = i 1 1 + 2T1 2T2 (g)2 1 1 2T1 2T2
2

(7.21)

56

LECTURE 7. RESONANCE FLUORESCENCE

The rst of these will describe a broadened resonance near elastic scattering (remembering that this is in the rotating frame). The second term describes (at large enough ), a resonance shifted by . When the weights of these poles are calculated, as shown in Fig. 7.1, the central peak has twice the weight of the other two peaks. This characteristic shape is known as the Mollow triplet, and has a simple interpretation.

= T1=T2=2 g=10 Intensity -15

-10

-5

10

15

Figure 7.1: Spectrum of resonance uorescence, from Eq. (7.19).

Interpretation: of the Mollow triplet


The origin of the Mollow triplet is the eective energy level scheme of the two-level system in the presence of a coherent eld. If one considers in general a two-level system coupled to photons, there are Rabi oscillations. In the frequency domain, these oscillations correspond to a splitting of the eigenstates resulting from hybridising the states with n excitations; i.e. the states |n 1, , |n, mix and lead to eigenstates split by g n (if on resonance). If the system can now decay, it will emit an excitation into the continuum, and undergo a transition to a state with one fewer excitation in the system. However, starting from one of the hybridised eigenstates with n excitations, it is possible to undergo a transition to either of the eigenstates with n 1 excitations non-zero matrix elements connect them all. Thus, in general there are four transition frequencies. In the limit of a strong classical eld, all four of these matrix elements have the same strength, and moreover, as indicated in Fig. 7.2, the Rabi splitting becomes almost n independent when n 1, so two of the frequencies will match. This then leads to the characteristic 1:2:1 ratio seen in Figure 7.1.

7.4

Further reading

The discussion of the resonance uorescence problem in particular can be found in chapter 16 of Meystre and Sargent III [10]. The discussion of the

7.4. FURTHER READING

57

g/2 g/2

g/2 g/2

Figure 7.2: Characteristic behaviour at high excitation number,


with n ||2 , demonstrating a triplet of transition frequencies, with the intensity ratio 1:2:1.

quantum regression theorem in this lecture is however more closely related to the density matrix based approach given in Scully and Zubairy [7]. The Onsager-Lax (quantum regression) theorem was introduced by Lax [18].

Lecture 8

Quantum stochastic methods and limitations of the Markovian approximation.


The previous two lectures, have introduced the density matrix equations of motion to describe a system coupled to a bath, and have shown how the Markovian approximation implicit in these equations of motion leads one to the quantum regression theorem. This lecture reviews the foundations of these methods for open systems. Two further formalisms are presented, which can be seen as the stochastic equations of motion for which the probability evolution is that of the density matrix. In the latter of these the Heisenberg-Langevin equations the consequences of the approximations required for a Markovian equation of motion are more explicit; these are discussed in terms of the uctuation dissipation theorem. We will consider just the problem of a system coupled to a single decay bath, which gave us the equation of motion: d (t) = n aa 2a a + aa dt 2 + ( n + 1) a a 2aa + a a . (8.1)

8.1

Quantum jump formalism

Consider evolution under Eq. (8.1) when n = 0, i.e. when only decay can occur. Then, after a short time t, one has: S (t + t) = s (t) = t a as + s a a + tas a + O(t2 ) 2 t t 1 a a s 1 a a + tas ta + O(t2 ). 2 2 (8.2)

These two contributions to the density matrix can be interpreted as the conditional density matrices that arise under no photon loss and one photon loss; they are added together because of the trace over states of 59

60

LECTURE 8. QUANTUM STOCHASTIC METHODS

the bath. The probability of these two nal states can be found from the trace of the conditional density matrices, hence: Pone photon loss = Tr tas a = t a a . (8.3)

This is as one intuitively expects; the probability of losing a single photon in [t, t + t] is given by the rate of loss of each photon multiplied by the total number of photons. Thus, the evolution can be written in terms of the probabilities of transition to the (suitable normalised) a| P = t |a a| |a a| | (8.4) 1 ta a/2 | P = 1 t |a a| |1 ta a| One may further then note that the second case given above corresponds almost to the evolution that the system would undergo with the nonHermitian Hamiltonian He = H i(/2)a a. This statement becomes exact if one disregards the normalisation, and allows the wavefunction to evolve in a non-unitary manner. This then gives a recipee for the evolution of the wavefunction: For each time step, choose a random number, and use it to decide whether a jump occurs or not. A jump occurs with probability t |a a| / | . If a jump occurs, replace | a| / |a a| .

Otherwise, evolve with the non-Hermitian Hamiltonian He = H i(/2)a a. This gives a particular trajectory. By repeating this procedure many times, one has an ensemble of trajectories which describe the statistics of photon counting events, and the average evolution of the system. This quantum jump formalism can be understood as describing the loss process in terms of the environment repeatedly measuring the number of photons in the cavity, and hence decohering the states with dierent numbers of photons.

Example: Photon loss from a single-photon state


The above procedure can be illustrated for a particularly simple case: if the initial state contains a single photon then from the density matrix equation: t n + k ||n = (2n + k ) n + k ||n 2 + (n + 1)(n + k + 1) n + k + 1||n + 1

(8.5)

it is clear that the initial condition n : n + k ||n = 0 is preserved except for k = 0, and so the only equations to solve are t 1||1 = t 0||0 = 1||1 . (8.6)

8.2. HEISENBERG-LANGEVIN EQUATIONS so the full solution is: = et |1 1| + (1 et )|0 0|.

61

(8.7)

In the quantum jump formalism it is easy to reproduce this result; at each time step we have that: |1 |0 |1 P = t P = 1 t (8.8)

A more complicated example is given in Question B3, for decay of a coherent state. In that case, since a coherent state is unchanged under the loss of a photon, one has the counterintuitive result that it is failure to observe a photon that leads to the density matrix evolving. This is less surprising if one considers that the state has an uncertain number of photons to start with, and failure to observe any photons escape indicates that the wavefunction has been projected by observation onto a state with fewer photons.

8.2

Heisenberg-Langevin equations

Just as one can solve a closed quantum-mechanical problem in either the Schr odinger or Heisenberg pictures, quantum stochastic approaches can be constructed in either picture. The previous sections generally described a stochastic evolution of the wavefunction, this section instead describes stochastic evolution of the system operators. Starting once again from the same coupled bath and system, and working in the rotating frame as normal, one has the Hamiltonian of Eq. (3.4), HSR =
k i( k )t + a bk ei(k )t . k ab ke

(3.4)

From this Hamiltonian, the Heisenberg operator equations of motion are: ia = [a, HSR ] =
k

k bk ei(k )t

(8.9) (8.10)

k = [bk , HSR ] = k aei(k )t . ib

To eliminate the bath degrees of freedom, one can integrate Eq. (8.10), and substitute it in Eq. (8.9). Thus,
t

bk (t) = ik
0

dt ei(k )t a(t ) + bk (0)

(8.11)

which, substituted into Eq. (8.9), gives:


t

a =
0

dt
k

2 i( k )(t t) a(t ) i k e k

k bk (0)ei(k )t . (8.12)

62

LECTURE 8. QUANTUM STOCHASTIC METHODS

Then, dening the second term in brackets as Fa (t), and making the same Markovian approximation:
2 k F (k ) k

d F ( ) 2

(8.13)

as in Eq. (3.10), so that the rst term in brackets becomes (t t ), this equation becomes: 1 a = a + Fa (t). (8.14) 2 Here Fa (t) is a stochastic operator; it has quantum mechanical commutation relations related to its denition in terms of bk ; but bk is a bath operator, and the state of the bath is random drawn from a thermal ensemble, so bk has dierent expectations for each realisation of the bath.

Preservation of commutation relations


The operator nature of Fa (t) is apparent if one considers the commutation relations:
[Fa (t), Fa (t )] = k,k i( k )t+i( k k k [bk , b k ]e 2 i( k )(tt ) k e = (t t ). k )t

(8.15)

This non-commuting nature is essential to ensure the preservation of commutation relations in the time-evolution of the operators a(t). The solution of Eq. (8.14) can be written a(t) = a(0)et/2 +
0 t

dt Fa (t )e(tt )/2 .

(8.16)

Using this, one can write: a(t), a (t) = a(0), a (0) et


t t

+
0

dt
0

dt

Fa (t ), Fa (t ) et+(t +t

)/2

, (8.17)

in which we have assumed no correlation between the initial state of the system and that of the bath. Then, using [a, a ] = 1 in the initial state, and the commutator in Eq. (8.15) this becomes: a(t), a (t) = et +
0 t

dt et+t (8.18)

= et + 1 et

This shows that the existence of the non-commuting stochastic terms was essential to preserve the operator commutation relations.

8.3. FLUCTUATION DISSIPATION THEOREM

63

Finite bath occupation; other correlations


To reproduce all the results of the density-matrix formalism, it is necessary also to have expressions for the correlation functions of the stochastic operators Fa (t) as well as their commutator. Given the form for the commutator, all correlation functions can be found from this and the anticommutator,
Fa (t), Fa (t )

=
k,k

k k

bk , b k

ei(k )t+i(k

)t

=
k

2 k (2nk + 1)ei(k )(tt )

d ei( )(tt ) coth 2 2 coth (t t ). 2

(8.19) (8.20)

The last line here makes the additional assumption of T , i.e. that the temperature is large compared to the bandwidth, with the bandwidth taken approximately equal to the imaginary width. This assumption is clearly necessary if the entire formalism is to be Markovian (i.e. to depend only on current state, rather than history), and is identical to the approximation made in Eq. (3.10). The validity and limitations of this last approximation are the subject of the next section.

8.3

Fluctuation dissipation theorem and the Markovian approximation

The uctuation-dissipation theorem relates the correlation function of an operator with its response function. For the photon operators considered here, the correlation function can be dened by:

C () =

d C ( )ei ,

C ( ) =

1 2

a(t), a (t + )

(8.21)

The response function can be found by considering the response to a perturbation: H = f eit a + H.c. (8.22)

The response function () is given by a(t) =

()f eit .

(8.23)

The discussion below uses the Heisenberg-Langevin formalism to evaluate the relation between these operators with and without the second Markovian approximation above. The response function is the same in either case, since it is averaged and so does not depend on the stochastic operator Fa (t). In the following

64

LECTURE 8. QUANTUM STOCHASTIC METHODS

it will be necessary to distinguish between the bare operators a(t) and the gauge transformed versions a (t) = a(t)eit . Thus, t + a (t) = i f ei()t (8.24) 2

where we have used the Heisenberg equation following from the gauge transformed perturbation Hamiltonian. This yields a(t) = eit a (t) = eit

if ei()t = i( ) + /2

f eit (8.25) ( ) i/2

hence () = 1/[( ) i/2]. Next, consider the uctuation correlation function at long times, so that the initial conditions are not involved. In this case, Eq. (8.16) should become:
t

a(t) = eit

dt Fa (t )e(tt )/2

(8.26)

and so: C ( ) = ei 2
t t+

dt

dt

Fa (t ), Fa (t )

e(2t+ t t

)/2

(8.27)

We now have two choices; we may either use the exact relation in Eq. (8.19), or the Markov approximation in Eq. (8.20). Without second Markovian approximation Considering rst Eq. (8.19), we have: C ( ) = ei 2
t t+

dt

dt 2

d coth 2

ei( )(t

t )(2t+ t t )/2

ei = 2 1 = 2

d coth 2 2

ei( )(t+ t) [i( ) + /2][i( ) + /2] (8.28)

d coth 2 2

ei ( )2 + (/2)2 2

hence it is immediately clear that: C () = coth /2 = coth ( )2 + (/2)2 [()] (8.29)

which is the required uctuation dissipation theorem. With second Markovian approximation In contrast, if we use Eq. (8.20) we have that: C ( ) = = ei 2 ei 2
t t+

dt

dt coth 2 2

(t t )e(2t+ t t

)/2

min[t,t+ ]

coth

dt e(2t+ 2t )/2 (8.30)

ei = coth 2

e| |/2 ,

8.4. FURTHER READING from which it follows straightforwardly that C () = coth 2 [()].

65

(8.31)

Thus, use of the Markovian approximation means that the uctuation dissipation theory fails, unless T ( ) > . Such a result can also be seen to be a limit for the use of the quantum regression theorem. In that case, the high temperature (classical) answer is known as the Onsager theorem. The quantum extension of the Onsager theorem is an approximation, that derives from the Markov approximation in the dynamics

8.4

Further reading

The topic of Heisenberg-Langevin equations, and stochastic operator equations is discussed briey in Scully and Zubairy [7], and more fully in Gardiner and Zoller [19] The implicit dependence of the quantum regression theorem on the use of the Markovian approximation is highlighted in the papers of Ford and OConnel [20, 21], and the response by Lax [22].

Lecture 9

Quantum theory of the laser


In this and the next two lectures, we move to discuss a somewhat more classical system, the laser. The laser is classical in the sense that far above the lasing threshold, it produces large coherent states with classical correlations, and it is based on stimulated emission of radiation, which can be described classically. However, we will address the quantum theory of this process, which allows one also to calculate the uctuations that occur near the lasing transition. In the following lectures, we will also consider more exotic stimulated emission processes, such as the micromaser, where a quantum theory will always be necessary.

9.1

Density matrix equation for a model laser system

The model that we will consider for a laser consists of a single cavity mode coupled to many atoms, which are pumped by some external process. Our eventual aim will be to describe the quantum dynamics of the light in the cavity, working in a limit where the atomic dynamics can be adiabatically eliminated. The dynamics due to the leakage of photons out of the cavity is simply given by the now familiar decay term: d dt =
decay

a a 2aa + a a . 2

(9.1)

In the following, it will often be useful to write equations in the Fock basis, and the decay contribution in this basis has the form d nn dt =
decay

(n + n )nn 2 2

(n + 1)(n + 1)n+1,n +1 .

(9.2)

For the gain medium, we will consider a three level conguration, consisting of two active lasing levels e, g , and a reservoir level 0. This is illustrated in Fig. 9.1). Although ve levels are drawn in this gure, our description eectively involves only three levels, as we assume recycling to the reservoir state is fast following any transitions out of the lasing levels, thus only the ground and excited states of the lasing transition, g, e and the reservoir state 0 need explicitly appear in our equations. 67

68

LECTURE 9. QUANTUM THEORY OF THE LASER

e g 0

lasing transition

Figure 9.1: Labelling of three level scheme

The coherent dynamics is therefore controlled by the Hamiltonian H= g 2


eg ge ai + a i . i

(9.3)

eg The operator i introduced here is the operator that takes atom i from state g to state e. Since we have three active levels, such operators should be written as 3 3 matrices, rather than the 2-level system denitions of , z . As well as the Hamiltonian, there are incoherent processes describing pumping, decay and dephasing. The pumping that we will consider corresponds to coupling the 0 e transition to an inverted reservoir. i.e., there is incoherent pumping from the reservoir to the excited lasing state. While this model may seem a crude approximation, it gives the same result as can be found by considering coherent pumping to a higher excited state, followed by incoherent decay to the excited lasing level. Problem C1 considers this more detailed description of the pumping scheme. In describing pumping and decay, we assume these processes act on each atom in the gain medium separately. There are three incoherent processes that we include: pumping from the reservoir to the excited lasing level at rate , decay from the lasing levels to the reservoir at rate , and pure dephasing of the lasing levels with a rate . Putting these ingredients together we have:

t = i[H, ]
i

a a 2aa + a a 2

00 e0 0e 00 2i i + i 2 i

gg 0g g0 gg 0e e0 ee ee 2i i + i + i 2i i + i 2 i gg 2 gg gg gg 2 ee ee ee ee (i i ) 2 (i i ) (i i ) + (i i ) 4 (9.4) + In writing this, we have used the fact that each term acts on a single atom to collapse adjacent atomic operators, e.g. 0e e0 = 00 , hence the simpler form of the pumping and decay terms. In the following, we will drop the subscript i on the atomic degrees of freedom, assuming all atoms

9.2. MAXWELL BLOCH EQUATIONS

69

are equivalent, and so the sum can be replaced by the number of atoms NA .

9.2

Maxwell Bloch equations

Before presenting the quantum theory of the photons, let us rst a simpler description of the laser, the semiclassical approximation which leads to the Maxwell-Bloch equations. This should be valid when large numbers of both photons and atoms are present. We will thus use the density matrix equations to write equations for expectations of three operators: = a , P = NA ge , N = NA ee gg , (9.5)

corresponding to the electric eld, polarisation and inversion respectively. The semiclassical approximation will be made by replacing expectations of products of operators by products of expectations. From the density matrix equations of motion, one immediately nds: g g t = i P , t P = +i N t P . (9.6) 2 2 2 For the equation for N we must write: g t ee = i (P P ) + 00 ee 2NA g t gg = +i (P P ) gg 2NA t 00 = 00 + ee + gg Let us now introduce the approximation of fast atomic dynamics, which we will also use in the quantum case. We can then nd the atomic steady state of these last three equations, assuming that remains constant over the timescales required for the atomic degrees of freedom to relax; in particular we assume . Combinging the last equation with the constraint 00 + ee gg + = 1, one then has 00 = /( +) and then taking the dierence of the rst two equations yields Na dN = ig (P P ) + N dt + (9.7)

The combination r = Na /( + ) describes the pumping rate to the excited laser level, accounting for the depletion of the reservoir. We next consider Eq. (9.6) and Eq. (9.7), and look at their steady states, and normal mode spectrum, to understand their behaviour.

Steady states of Maxwell Bloch equations


For a steady state, Eq. (9.6) immediately implies that: ig P=i = N . g 2t (9.8)

This implies either that = 0 or that N = 2t /g 2 . The latter solution (corresponding to lasing) implies that inversion locks at its threshold value; this is known as gain clamping. From Eq. (9.7), we have either that:

70

LECTURE 9. QUANTUM THEORY OF THE LASER

Below threshold = 0. In this case, t N = 0 implies N = r/ . Hence, threshold occurs at r = 2(t /g 2 ). Above threshold N = Nsat = 2t /g 2 . In this case, we may use the gain equation as a condition for , using Eq. (9.8) to eliminate P , and nd: 0 = 2||2 + r Nsat ||2 = 2 r Nsat (9.9)

Thus, above threshold ||2 increases linearly with pump rate in this mean eld theory. This is illustrated in Fig. 9.2.
||2 N

0 -/2 Im[] - Nsat r rth r

||2

Nsat

Figure 9.2: Left: Field strength, and inversion arising from mean
eld theory. Right: Decay rates of uctuations of Maxwell-Bloch equations.

Fluctuations of Maxwell-Bloch equations


The instability of the non-lasing state at threshold can be understood by considering uctuations, and linearising the Maxwell-Bloch equations around = P = 0, N = r/ . One then nds: /2 ig/2 0 0 P t P = igr/2 t (9.10) N 0 0 N Thus, inversion uctuations are clearly damped. The photon eld and polarisation have collective oscillations at a frequency given by: = i t + i 4 2 t 4 2
2

g2r . 4

(9.11)

This frequency is purely imaginary, so the modes are either growing or decaying. If r > rth , one of these roots becomes positive, and such uctuations will grow. These modes are shown in Fig. 9.2.

9.3

Reduction to photon density matrix equation

We will now reduce Eq. (9.4), which contains both eld and atomic degrees of freedom, to an equation for just the eld degrees of freedom. This means

9.3. REDUCTION TO PHOTON DENSITY MATRIX EQUATION we want to introduce: = Tratoms () = ee + gg + 00

71

(9.12)

The extra notation ij indicates an operator in the photon space, but a c-number in the atomic degrees of freedom; i.e. we dene: ij = Tratoms ( ji ) (9.13)

where i, j are photon number states. With this notation, taking a trace of Eq. (9.4) over atom states gives: t = i gNA age + a eg ge a eg a 2 a a 2a a + a a . (9.14) 2

Thus, to proceed, we must also nd equations for eg etc. Taking a trace over atoms with a factor of ge then gives: t eg = eg ig (agg ee a) a a 2aeg a + eg a a 2 2 (1 + 1)eg (1 + 2 + 1)eg . (9.15) 2 4

To make further progress, we will assume that the atomic dynamics can be adiabatically eliminated, i.e. . This rstly means we can drop the term in the above equation. It secondly means that, for the purpose of substituting into Eq. (9.14), we may nd the steady state of the atomic degrees of freedom. Dening t = + , this gives the condition: 0 = t eg = giving the result: eg = ig (agg ee a) . 2t (9.17) ig (agg ee a) t eg , 2 (9.16)

We can now repeat the above logic to nd nd the relation of the diagonal elements ee , gg , 00 . Dropping the terms, and looking for steady states, one nd: ig eg a ge a gg 2 g2 = a agg + gg a a 2a ee a gg 4t ig 0= age eg a ee + 00 2 g2 aa ee + ee aa 2agg a ee + 00 = 4t 0 = 00 + (ee + gg ) . 0=

(9.18)

(9.19) (9.20)

72

LECTURE 9. QUANTUM THEORY OF THE LASER

Equations (9.18) and (9.19) have made use of the result in Eq. (9.17). Furthermore, Eq. (9.20), along with the denition in Eq. (9.12) implies that: 00 = , ee + gg = . (9.21) + + We then have a closed set of three equations for , ee , gg . However, because these equations involve photon operators, it is rather involved to invert them to get a single equation for . To do this inversion, it is easiest to work in the photon number basis. Restricting to the diagonal elements, g we have equations for pn , pe n and pn , which become: g2 e g 2npg n 2npn1 pn 4t g2 g e 0= 2(n + 1)pe pn . n 2(n + 1)pn+1 pn + 4t + 0= (9.22) (9.23)

Shifting n n 1 in Eq. (9.23), it is clear one can combine the two equations, by taking their dierence to give: 0= g2 e g e n pg pn1 . n pn1 pn pn1 t + 1 pn1 . 2 + g n/t + (9.24)

e This equation can be solved for pg n pn1 to give: e pg n pn1 =

(9.25)

Then, noticing that the atom-photon coupling terms in Eq. (9.18) and Eq. (9.19) add up to give the required term in Eq. (9.14), one has the eld equation in the form: dpn g 2 NA g e e = 2n(pg n pn1 ) 2(n + 1)(pn+1 pn ) dt 4t npn + (n + 1)pn+1 (9.26) and substituting Eq. (9.25) gives the form Na g 2 n/t g 2 (n + 1)/t dpn = p pn n 1 dt 2( + ) + g 2 n/t + g 2 (n + 1)/t [npn (n + 1)pn+1 ] (9.27) This can be written more simply if we introduce: r= to give: dpn r n (n + 1) = pn1 pn [npn (n + 1)pn+1 ] (9.29) dt 2 1 + n 1 + (n + 1) Clearly, if , then r NA , meaning that the eective pumping rate is just the bare pumping rate. However, if becomes too large, then the eective pumping rate is limited by the recycling rate , as the reservoir becomes depleted. Na , + = g2 t (9.28)

9.4. PROPERTIES OF THE LASER RATE EQUATIONS

73

9.4

Properties of the laser rate equations

Having established the quantum description of the photon state in the lasing cavity, given by Eq. (9.29), we now examine the consequences of this equation. We begin by looking at the steady states, and seeing how to recover the results we found from the semiclassical analysis.

Laser threshold condition


Lasing occurs when the vacuum state becomes unstable, i.e. when t n > 0. Thus, let us nd: t n = nt pn . (9.30)
n

Let us however assume that n is small i.e. that the average population is small compared to 1/ , so that the denominators in the gain can be expanded. Then, t n =
n

r [n (1 n) pn1 (n + 1) (1 (n + 1)) pn ] 2 [npn (n + 1)pn+1 ] . (9.31)

Now, shifting the various sums, one can rewrite this as t n =


n

pn

r (n + 1)2 n(n + 1) 2 (n + 1)3 n(n + 1)2 2 n2 n(n 1)

, (9.32)

which can be written as: t n = r 2 r r n + 2 (n + 1)2 . 2 2 (9.33)

This can be understood as three terms; the rst describes competition between loss and stimulated emission; the second describes spontaneous emission; the third describes the nonlinear susceptibility of the two-level systems, reducing the gain they provide when n is large. The rst and third of these contributions could be recovered from a semiclassical theory of lasing, but the second term describes spontaneous emission which requires a fully quantum theory. However, if our aim is the lasing condition, this can easily be identied from the rst term: stimulated emission outstrips decay if r/2 > , and lasing then occurs. This matches the threshold found from the Maxwell-Bloch equations.

Steady state probabilities


The coecients in Eq. (9.29) depend on n in such a way that one may write: dpn = Bn pn + Bn1 pn1 Cn pn + Cn+1 pn+1 , (9.34) dt

74

LECTURE 9. QUANTUM THEORY OF THE LASER

with excitation rate Bn = r (n + 1)/[1 + (n + 1)] and decay rate Cn = n. To nd a steady state that is valid for all n, we need to nd a relation between terms on the right hand side such that this will always vanish. The only consistent equality is to choose: Bn pn = Cn+1 pn+1 Bn1 pn1 = Cn pn . (9.35)

The steady state distribution is then given by:


n

pn =
m=1

Bm1 p0 = Cm

n m=1

r 2

1 + m

p0 ,

(9.36)

with p0 set by normalisation. This product can in general be expressed in terms of a hypergeometric function. Let us consider two limits where a simpler expressions result. Below threshold If we are at small powers, then small values of n will be most probable. In this case, one may neglect the term involving m/ 2 in the denominator. Thus, at low pumping rates, below threshold, one has pn (r/2)n , which is a thermal distribution, with exp(Ee ) = (r/2), from which one recovers n = 1 eEe 1 = r , 2 r (9.37)

in agreement with the small pumping rate limit of Eq. (9.33). Clearly, such an approximation is only valid below threshold, otherwise pn cannot be normalised. In this low pump limit, there is an eective pump rate, limited by the ratio of g/ , indicating that emission competes with relaxation. Far above threshold The opposite limit occurs when n is large enough that the term 1 in the denominator can be neglected. In this case one has pn (r/2)n /n!. This is a Poissonian distribution, as one would expect for a coherent state, with average population r/2, set only by the balance of pump rate and decay rate, since strong elds mean that stimulated emission beats any other relaxation process, thus for every excited atom created, half a photon is emitted.

9.5

Further reading

The laser rate equations presented here follow closely the presentation in Scully and Zubairy [7]. For a dierent presentation of laser theory, the review article by Haken [23]. The discussion of laser linewidth in this Chapter is based on Scully and Zubairy [7].

Lecture 10

Laser uctuations: linewidth and correlations


In this lecture we continue to discuss the quantum theory of the laser, now looking at uctuation properties, the linewidth, and the photon correlations.

10.1

Laser linewidth

To nd the laser linewidth, we want to calculate the spectrum of the emission from laser cavity. As in lecture 7, this can be found by using the quantum regression theorem to calculate a two time correlation. In the current case we need to nd

I ( ) 2
0

dteit a (t)a(0)

(10.1)

The quantum regression theorem means this is equivalent to nding the time evolution of a (t) , when the initial state at t = 0 is taken to result from acting with a on the equilibrium density matrix. If we have the time-dependent solution of the equation of motion for the density matrix, then we can write a (t) = Tr a (t) = n + 1n,n+1 (t). (10.2)
n

We will dene the shorthand n,n+k (t) = k n (t), and so our task is to nd 1 ( t ). Before writing its equation of motion, let us anticipate how it should n behave. Regardless of the initial conditions, at long times this expectation should decay to zero, as it is not invariant under phase rotations, while the steady state ensemble is invariant under phase rotation. However, since the initial conditions are those set by aeqbm , the expectation of a (t) will be non-zero; we are interested in how this function decays in time.

Equation of motion
Repeating the same procedure as in the last lecture, of eliminating atomic degrees of freedom, one may recover the equation of motion for the o75

76

LECTURE 10. LASER FLUCTUATIONS: LINEWIDTH AND CORRELATIONS

diagonal density matrix. dk n k k k k k = Ak n n + Bn1 n1 + Cn+1 n+1 . dt With , r as dened in the previous lecture, these coecients are:
k Bn =

(10.3)

(n + 1)(n + 1 + k ) r , 2 1 + (n + 1 + k/2) + 2 k 2 /16 n(n + k ), . r (n + 1 + k/2) + 2 k 2 /8 k + n+ 2 2 2 1 + (n + 1 + k/2) + k /16 2

k Cn =

Ak n =

Details of the derivation of these coecients are presented in appendix 10.A. k , C k have been written in such a way that for k = 0 Note that the terms Bn n they reduce to the previous Bn , Cn discussed above.

Approximate solution
k k 0 0 0 For k = 0, the term Ak n is simply related to the Bn , Cn by An = Bn + Cn . However, for k = 0 no such simple relation exists in general. Since it was this simple relation that lead to the detailed balance condition for the probability distribution, no equivalent condition exists with k = 0. The previous steady state solution required this detailed balance, and so no such steady state solution exists for k = 0. However, the previous solution does motivate a possible extension, which gives a solution decaying in time: n k Bm 1 k 0 , k Cm

k n (t)

=e

kt Dn

(10.4)

m=1

k is a parameter to be determined self consistently. Substituting where Dn this ansatz for k n into Eq. (10.3) gives: k Cn k n k Bn 1
k k

k k k k (Dn1 Dn )t Dn n = Ak n n + Bn1 e

k (Dn+1 Dn )t + Cn +1 e

k Bn k n k Cn +1

(10.5)

k The terms in brackets are the expressions for k n1 in terms of n from k k Eq. (10.4). If Dn Dn1 , then the exponential terms may be dropped, giving: k k k Dn = Ak (10.6) n Cn B n . k Thus, if we solve Eq. (10.6), and nd it varies slowly with n, then Dkn Dn denes the decay rate of the o-diagonal density matrix. We therefore want to nd D D1n to solve for the linewidth. If we assume we are far above

10.1. LASER LINEWIDTH threshold, so n to give: D

77

1 and B 1 , 1, then we can expand the square roots in Cn n

r ( n + 1 + 1/2) + 2 /8 ( n + 1 + 1/2) 2 n + 1 + (1 + 1/2) + 2 /16 1 1 1 + n + n 1+ +O 2 2n 8n2 r g2 1 + 8n 2 2

1 n2 (10.7)

Thus, using this time-dependence in the two-time correlation, we have: g1 ( ) eD ; (10.8)

We have been working in an interaction picture, neglecting free evolution of the photon elds at frequency . Restoring this, the frequency dependence of the intensity is then given by: I ( ) = D ( )2 + D 2 (10.9)

Hence the FWHM of this Lorentzian lineshape is given by: 2D = 1 4n r g2 + 2 t (10.10)

Interpretation of linewidth
This result for the lineshape has a simple interpretation in terms of phase diusion. If we consider the radiation eld of a laser to be described by a coherent state, with eectively xed magnitude || = n , but uctuating phase. Decay of correlations occurs because of variation of the phase. Each time a photon is incoherently emitted emitted into, or lost from the cavity the phase changes. This is illustrated in Fig. 10.1.

I()

R()
Figure 10.1: Cartoon of coherent state undergoing a random walk
due to spontaneous emission events, leading to phase diusion for the coherent state.

To turn this cartoon into a physical estimate, we start from the rate of such changes per unit time, which is d = + rg 2 /2t , by adding rates

78

LECTURE 10. LASER FLUCTUATIONS: LINEWIDTH AND CORRELATIONS

of photon decay and spontaneous emission (hence the appearance of the eective pumping rate, as appearing in the threshold condition, rather than the total pump rate, as in the probability distribution far above threshold). If we assume a random walk, with d events per unit time, the change to the coherent state is governed by the random walk probability distribution: P () = td ()2 exp td (10.11)

However, the variation is entirely phase, so = n . Using this, the decay of correlations after time can be found by averaging the state: g1 (t) = = a (t)a(0) = ei[+(t)] ei | a |2 d n ()2 td exp i() td = exp td 4n , (10.12)

which reproduces the previous result, D = d/4 n .

10.2

Spontaneous emission and the parameter

We now return to considering single time properties of the laser system, but consider further the role of uctuations. At the start of the previous lecture we discussed the Maxwell-Bloch equations, which are valid when uctuations are small (i.e. when photon numbers are large). When numbers of photons are small, spontaneous emission becomes important, which was not considered previously. However, its leading order eects can be found by writing: t n = Bn Cn (10.13) with Bn , Cn as derived in the previous lecture, and re-derived above. If we make the mean-eld assumption Bn B n , and we write n for n , then we have: r g 2 (n + 1) t n = n (10.14) 2 t + g 2 (n + 1) We will investigate here the eects of this +1 we should be aware however that whenever the eects of this term become signicant, so do uctuations (as we will investigate more fully below), in which case no mean-eld theory can be applicable i.e. it is necessary to study the full probability distribution, not just n . However, if we can understand when the +1 term will matter, then we can see the conditions under which semiclassics is a reasonable approximation. The important question regarding the eects of the +1 is to understand whether single photons will aect the dynamics. In order for single photons to have signicant eects, they must be able to saturate the transition. Whether or not this is possible is determined by the ratio = g 2 /t that appeared frequently in the density matrix equations. Rather than working

10.3. FANO FACTOR

79

with , it is more common to introduce the parameter which is dened as: g2 = . (10.15) 2 t + g 1+ If g 2 t then single photons have a negligible eect, and this correspond to 1. If g 2 t , single photons are important, and 1. In terms of , we may rewrite Eq. (10.14) as: t n = r n + 1 n 2 1 + n (10.16)

Based on the semiclassical threshold discussed above (valid if 1, and 2 so g /t ) we will dene rth = 2/ . In terms of this threshold, we can then solve the steady state number of photons to be given by r(n +1) = rth n(1 + n), with the solution: 2 r r 1 r (10.17) 1 1 + 4 n= 2 rth rth rth This equation (normalised by plotting n) is shown in Fig. 10.2. For small it is clear that one recovers the sharp step like feature at r = rth . However as increases, the step becomes more smeared, and eventually, at = 1, one has n = r/rth for all r, i.e. no step survives at all. This indicates that in the limit of strong coupling, there is no clear laser threshold, but rather a thresholdless lasing behaviour.
15 F 10 5 0 101 10 n=n/nsat 10-1 10-2 10
-3

=10-3 =10-2 =0.1 =0.9

2 1 0 0 0.1 1 (r/ 2) = P/Pth 1 2 3 10

10-4

Figure 10.2: Numbers of photons vs pump rate (normalised by


rth , and Fano factor, for dierent values of

10.3

Fano factor

As well as the sharpness of n vs pump power, further insight into the role of can be found by considering the full probability distribution, and the

80

LECTURE 10. LASER FLUCTUATIONS: LINEWIDTH AND CORRELATIONS

relative size of uctuations. We choose to measure the Fano factor, dened by: n2 n 2 F = . (10.18) n This parameter measures deviations from Poissonian statistics, as F = 1 for a Poisson distribution. The uctuations could also equivalently be characterised by g2 (0) g2 (0) = n2 n a a aa = . n2 a a 2 (10.19)

Let us now determine F from the full probability distribution as derived in the last lecture, and assuming is not too large. Far above threshold, the distribution is Poissonian, and so F 1. Below the threshold, we may use the Bose-Einstein distribution pn (r/rth )n to give: n = and so: F =1+ r/rth 1 r/rth (10.21) r/rth , 1 r/rth n2 = r/rth (1 + r/rth ) (1 r/rth )2 (10.20)

which again becomes 1 if far from threshold. Exactly at threshold, one can make an approximation for small that:
n

pn
m

(1 + m)

exp
m=0

exp n2 /2 .

(10.22)

This is valid if 1, in which case, pn has decayed to a small value well before n 1. With this one-sided Gaussian distribution, one then nds: n = 2 , n2 = 1 , 1 F = 2 2 . (10.23)

Thus, if 1, there is a sharp spike of Fano factor at the transition the uncertainty in photon number becomes very large here in the semiclassical limit. On the other hand, everywhere else, the Fano factor is small. The full behaviour of the Fano factor vs is shown in Fig. 10.2. It is clear that for 1, there is no clear spike, and instead one has a broad region near the ill-dened threshold where uctuations are large.

10.4

Further reading

Discussion of density matrix equations and Maxwell-Bloch equations for our model laser can be found in many textbooks, e.g. Meystre and Sargent III [10], Yamamoto and Imamo glu [8]. The eect of the parameter, and the nature of the threshold is discussed in Rice and Carmichael [24].

10.A. DERIVATION OF OFF DIAGONAL FIELD DENSITY MATRIX EQUATIONS 81

10.A

Derivation of o diagonal eld density matrix equations

This appendix presents the intermediate steps required in getting from the combined atomeld equations of motion given in Sec. 9.3 to the form shown gg in Eq. (10.3). Introducing probabilities pg,k n = n| |n + k and similar e,k o,k expressions for pn , pn we may proceed as previously. Equation (9.20) k once again immediately yields po,k n = pn /( + ) and so Eq. (9.18) and Eq. (9.19) can be written as: e,k g,k (2n + k )pg,k n 2 n(n + k )pn1 pn 4 g,k 0 = (2n + k )pe,k pk pe,k n1 2 n(n + k )pn n1 + 4 + n 0= (10.24) (10.25)

Note that in Eq. (10.25), we have shifted n n 1 as compared to Eq. (9.19), so that both the above equations involve the same probability terms. Dening: k n = 1 + (2n + k ), 4 one may rewrite these equations as:
k k n n k k n n k n =

n(n + k )

(10.26)

pg,k n e,k pn1

pk n1 +

0 1

(10.27)

e,k which can be inverted to give pg,k n , pn1 . These can then be subsituted into the density matrix equation for the eld, coming from Eq. (9.14). Substituting Eq. (9.17) into Eq. (9.14), and taking matrix elements, one has:

t p k n

gain

N (2n + k + 2)pe,k n 2 4

(n + 1)(n + k + 1)pg,k n+1

e,k +(2n + k )pg,k n 2 n(n + k )pn1 . (10.28)

Substituting Eq. (10.27) into this then yields: t p k n = r k pn 4


k k (2n + k + 2)n +1 2 (n + 1)(n + k + 1)n+1 k k 2 2 (n +1 ) (n+1 )

gain

+pk n1

k 2 n(n + k )k (2n + k )n n k )2 ( k )2 (n n

. (10.29)

These can then be reduced to the nal form by using: (2n + k )2 4n(n + k ) = k 2 n(n + k )k = k 2 and noting the cancellation occuring in (2n + k )n n 2 n(n + k )

Lecture 11

Stronger coupling, micromasers, single atom lasers


In the previous lectures we considered lasing from an ensemble of three-level atoms. The density matrix equations of motion for the cavity eld allowed a quantum description of the photon eld. In general however, classical descriptions gave adequate answers, unless the parameter became large. In this lecture, we consider lasing in other systems, where quantum eects are more signicant. The majority of the lecture will concentrate on the micromaser, and the latter parts will consider single atom lasers.

11.1

Micromaser gain

r t=0 t=
Figure 11.1: Micromaser scheme, atoms in their excited state are
injected at a rate r, and fall through the cavity in time .

The micromaser consists of a sequence of excited state atoms passed into the cavity, which then fall through, in a given time, leading to gain for the photon eld. We will dene as the time during which each atom interacts with the cavity, and r as the rate at which atoms are injected. We will again assume fast atomic dynamics, so we will restrict 1. In this limit, the gain will have a periodic dependence on , as longer interaction times will cause the coupled atom-cavity system to perform Rabi oscillations. More 83

LECTURE 11. STRONGER COUPLING, MICROMASERS, SINGLE 84 ATOM LASERS crucially, there will be a non-trivial (non-monotonic) dependence of the gain on the number of photons already emitted, producing complicated steady state distributions. To nd the gain supplied by the two-level systems, we want to nd the evolution of the eld density matrix. If interaction between the twolevel system and the cavity for time causes the change eld ( ), and new excited two-level systems are created at rate r, then one can write the evolution of the eld as d eld dt = reld ( ),

(11.1)

Thus, we require the change of the eld state due to evolution for time : eld ( ) = Tratom [(t0 + )] eld (t0 ) = eld, (t0 + ) + eld, (t0 + ) eld, (t0 ). (11.2)

In the second line we have used the fact that the two-level system is by denition excited at t = t0 . Now, using results of Sec. 2.3, we can follow the evolution of a state |, n with n photons to time : |, n cos n+1 2 + i cos(2) sin n+1 2 n+1 2 |, n (11.3)

+ i sin(2) sin

|, n 1 ,

with n = ( )2 + g 2 n, tan(2n ) = g n/( ) as in chapter 2. For simplicity, let us assume resonance, = , so 2 = /2. Then, using this wavefunction evolution, one nds: nn ( ) = nn cos g g n + 1 cos n + 1 1 2 2 g g + n1,n 1 sin n sin n 2 2

. (11.4)

It is apparent this equation describes gain, since it describes transfer of probability from the state with n 1 photons to the state with n photons. Note that as anticipated, the xed time would lead to possible cancellation if integer numbers of Rabi oscillations occur. However, since the oscillation period is dierent for each number of photons, the general result is very complicated, and can lead to a very non-smooth probability distribution.

Random interaction times and recovery of laser limit


It is interesting to note that if the interactio time is not taken as constant, but is instead drawn from a Poisson distribution, one recovers the laser density matrix equations we considered in the previous lectures. Physically, this arises because r plays the role of the eective rate of excitation to the upper lasing levels, and the poissonian removal of interacting atoms is equivalent to the relaxation of atoms from the lasing levels back to the resrvoir.

11.2. MICROMASER AND NOISE

85

To see this mathematically, one may write that for a Possionian distribution of interaction times, with mean time 1/ one has: d eld dt

=r
0

d eld ( )e .

(11.5)

Using the form of given in Eq. (11.4), we can then evaluate these averages. For simplicity, we will present here only the calculations for the diagonal elements, pn = nn dpn dt = r sin2
gain

g n+1 2

pn + sin2

g n 2

pn1 . (11.6)

Then, using the identity sin2 2 = = d e sin2 4 d e ei + ei 2 = 2 4 1 1 2 2 /2 2 + = , (11.7) + i i 1 + 2 / 2

one nds that the rate of gain due to a such a distribution of atomic coherence times is given by: dpn dt =
gain

r (g 2 n/ 2 ) (g 2 (n + 1)/ 2 ) p pn . n 1 2 1 + (g 2 n/ 2 ) 1 + (g 2 (n + 1)/ 2 )

(11.8)

which matches the expression we found previously, with = g 2 / 2 here. (Only atomic decay, and no dephasing is present, hence t = .)

11.2

Micromaser and noise

Let us now consider the lasing properties of the micromaser. Given that Rabi frequencies associated with dierent photon numbers appear in Eq. (11.4), the density matrix equation will have a non-trivial evolution with both g and r. Although the behaviour is non-trivial, the steady state can still be written explicitly as: n r n sin2 g m/2 pn . (11.9) m
m=1

Several features may be noted about this distribution. The rst is that it will depend irregularly on g , as dierent photon number components perform incomensurate Rabi oscillations. This is illustrated in the lower panel of Fig. 11.2 which shows the evolution of n with g . Irregular jumps in the average occupation occur when destructive interference (g m (2N + 1)/2) causes high photon number components of the distribution to become unreachable. A more important feature of the distribution given in Eq. (11.9) is that the the dependence on n, can lead to number squeezing (i.e. sub-poissonian noise). In the previous lecture, we showed how the Fano factor F evolves

LECTURE 11. STRONGER COUPLING, MICROMASERS, SINGLE 86 ATOM LASERS


5

F 0 20 15 10 5 0 0 2 4 g 6 8 10 n

Figure 11.2: Numbers of photons and Fano factor vs coupling g for a micromaser. Blue squares show the most probable number of photons. (r/ is xed at 20)

through the laser transition, however we always had F > 1 in all cases, meaning that the noise is always at least poissonian, and is sometimes worse. On the other hand, the micromaser can have sub-poissonian noise, i.e. ucutations reduced below the classical case. This sub-poissonian noise occurs because the number dependence of the gain means that the distribution is attracted to particular number states. This is illustrated schematically in Fig. 11.3, which plots G(n) = r sin2 (g n + 1/2) n as a function of n. The value of this function determines whether there is net gain (G(n) > 0) or net loss from a given photon number state, i.e. whether such photon number states are driven toward higher or lower photon numbers. The photon number distribution is dominated by semiclassical congurations where G(n) = 0 and dG/dn < 0, so that small uctuations are self-correcting.

Net gain

n
Figure 11.3: Gain vs n for micromaser equation Going beyond semiclassics, there will be number uctuations near these semiclassical congurations, and the variance of photon number will depend on the size of these uctuations. Because the width (i.e. range of n) of the attraction regions shown in Fig. 11.3 varies in a non-smooth way, the width

11.3. SINGLE ATOM LASERS

87

of the distribution does not simply relate to the mean number of photons. Thus, the variance can be far less than it would be for a Poisson distribution with the same mean number of photons. The dependence of photon number and noise on the atom pumping rate r is also non-trivial, but here a semiclassical picture can be of more use in explaining what is seen. Consider the semiclassical states with G(n) = 0, illustrated in Fig. 11.3. As r increases, new semiclassical attractors appear, i.e. new peaks of sin2 g n + 1/2 fall within the range n > r/. The locations of these attractors are shown in the bottom panel of Fig. 11.4 by the dashed lines. While it is clear that the values of both n and the most probable photon numbers do correspond to these semiclassical states, further physics is required to understand when jumps between these semiclassical states occur. Once again, one sees that sub-poissonian noise is generic; super-poissonian noise only occurs at the boundaries between stable values of n , and is a natural consequence of bimodal photon number distributions.
6 5 4 3 2 1 0 60 50 40 30 20 10 0 0 50 100 r/ 150 200 n

Figure 11.4: Numbers of photons and Fano factor vs pump rate


r/ for a micromaser. Blue squares show the most probable number of photons. Dotted lines correspond to the mean-eld results, r sin2 (g n + 1) = n. (g is xed at 2.5)

11.3

Single atom lasers

In this nal section on the laser, we consider what happens when the gain medium consists of only a single quantum emitter. In such systems, semiclassical descriptions will naturally be of little use. One may connect this expected behaviour to the results we have already found relating the laser parameter to the validity of semiclassics. To do this, we may start by noting that the density matrix equation in Sec. 9.1 could reasonably describe a single atom. However, it does rely on the separation of timescales, and if very strong coupling is required, then the approximations used in deriving it will become invalid. To see why this

LECTURE 11. STRONGER COUPLING, MICROMASERS, SINGLE 88 ATOM LASERS might occur, let us rewrite the threshold condition as: r= NA 2 = = rth + (11.10)

If NA = 1, and cannot be signicantly tuned (as it describes intrinsic decay rates of atoms), then to reach threshold with a single atom, it is generally necessary to increase , so that one will no longer have 1. In this case, uctuations will matter more, the behaviour will be thresholdless, and if g becomes very strong then no separation of timescales is possible. Thus, while single atom lasing need not automatically imply a breakdown of the above description, reaching threshold with a small number of atoms generally requires strong coupling, and this in term means large , hence thresholdless lasing. In addition, the above requirement to reach threshold with a single atom will generally imply that one no longer has a separation of timescales i.e. the atomic decay is typically slower, and so the coupled dynamics of atom and eld must insteady be studied. The following discusses a few examples of such systems that have been studied.

Optical cavity, real atom

Trap

e g

Trap
Figure 11.5: Schematic of single atom optical lasing scheme, and
level scheme

Lasing from a single atom, using an optical transition and a regular cavity has been achieved by [25]. The system and level scheme are shown schematically in Fig. 11.5. The parameters for this experiment were (g, , ) = 2 (16, 4.2, 2.6)MHz. This implies 0.97, and so the lasing will be thresholdless. In addition, because of the strong coupling, the adiabatic elimination of the non-lasing levels is not a good approximation in this system. This allows the possibility of sub-poissonian behaviour, which was observed experimentally, and can be described theoretically by modelling the full four level system illustrated in Fig. 11.5.

Josephson qubit micromaser


The other example we will mention is a superconducting qubit as the articial atom, coupled to a mechanical resonator, thus this is phonon lasing rather than photon lasing[2628]. The circuit is shown schematically in Fig. 11.6

11.4. FURTHER READING


Resonator SC Island

89

Vg + Vds

EJ

Vg

Figure 11.6: Circuit diagram of single articial atom laser

The two-level atom here consists of two charge states of the island, having either zero or two extra electrons on the island (two electrons because of Cooper pairing, and the consequent odd/even energy gap). Thus one has + = 20 etc. This gives the Hamiltonian: H = [EC + (a + a )] z + 2EJ x + a a (11.11)

where EC depends on the capacitance of the island, and describes the shift due to the location of the mechanical resonator. An equivalent model would also describe a capacitively coupled stripline resonator. As well as this Hamiltonian behaviour, the bias Vds leads to a single-electron current through the island, with individual electrons hopping in a two step process. Eliminating the singly charged state (which decays fast), one has the eective density matrix equation: t = i[H, ] a a 2aa + a a 2 + 2 + + + 2

(11.12)

Lasing can occur if EC + 0, in which case, there is a cycle of incoherent transition from 2 0, followed by coherent transition from 0 2 with emission of a phonon to conserve energy. However, because the photon number aects the energy of the state EC , this leads to micromaser like behaviour.

11.4

Further reading

A discussion of the micromaser is given in Scully and Zubairy [7]. A review on cavity QED particularly focusing on micromaser behaviour can be found in Walther et al. [29].

Lecture 12

Collective eects of two-level atoms: open systems, superradiance


So far in these lectures we have either considered behaviour of a single two-level atom, or if we have considered ensembles then we have assumed the atoms act independently. However, as rst pointed out by Dicke[30], this approach cannot be correct if the atoms are close together, as they see the same electromagnetic modes with suciently similar phases, and so emission is a collective process. This collective behaviour can signicantly change how decay occurs, even when the electromagnetic modes are treated as a Markovian bath, with short memory times. In this lecture we will consider the case of many atoms coupled to a continuum of radiation modes (i.e. without a cavity); the next lecture will consider collective eects of many atoms with a cavity.

12.1

Simple density matrix equation for collective emission

We may begin by considering coupling between a collection of atoms and radiation modes: HSR =
k i

gk i(k i ak e 2

)tikri

+ + i ak ei(

k )t+ikri

(12.1)

Repeating the derivation of the system density matrix equation, we recover: d (t) = dt
t

dt
ij k

gk 2

eik(ri rj )

+ + + i j k j i (k + k ) + i j k . (12.2)

Two extremes of this equation exist; if |ri rj | 0 , then the atoms will act independently. In the other extreme, we may neglect all phase factors, 91

92 in which case we can introduce: J=

LECTURE 12. SUPERRADIANCE

i
i

(12.3)

and write:

d (t) = J + J J J + + J + J . (12.4) dt 2 This second limit is what we will consider in this section. We must however note that when separations are small, one must also consider the eects of Coulomb interactions between the two-level systems. In the Coulomb gauge, this means the explicit Coulomb term will have an eect; in the dipole gauge, the problem is instead that long-wavelength modes (for which phase factors can not be neglected) describe the eects of Coulomb interaction, leading to important energy shifts. We will discuss this eect in section 12.2. Putting aside these issues for the moment, let us study the consequences of Eq. (12.4) Dicke Enhancement of emission rate A simple understanding of how collective emission diers from independent emission can be found by considering eigenstates of |J |, J z . These are spin states and so obey: J 2 |J, M = J (J + 1)|J, M J |J, M = M |J, M J |J, M = J (J + 1) M (M 1)|J, M 1
z

(12.5) (12.6) (12.7)

We may then use these states as a basis for the density matrix equation, and writing equations for the diagonal elements PM J = M J,M J we nd: d PM,J = [J (J + 1) M (M 1)]PM,J dt + [J (J + 1) M (M + 1)]PM +1,J

(12.8)

The rate of emission is thus I = [J (J + 1) M (M 1)] = (J + M )(J M + 1), whereas if incoherent, the total emission rate would be N . If we consider cases where J = N/2, if J takes the maximum value possible consistent with the number of two-level systems, then the radiation rates at characteristic values of M are: M M M =N I = N 2 2 =0 I = N 4 = N 2 + 1 I = N

(12.9)

Thus, near M = 0, there is macroscopic enhancement of the emission rate compared to the incoherent case. Even when M = 1 N/2, the emission is much greater than it would be for a single excitation if considering independent atoms.

12.1. SIMPLE DENSITY MATRIX EQUATION FOR COLLECTIVE EMISSION 93

Solving the simple model


The above estimates indicate the general idea of superradiance; as M decreases, the collective eects increase in importance, leading to an increasing rate of transitions through the super-radiant cascade. Our aim in the following will be to describe this cascade. Before discussing a tractable approximation scheme, we may note that the Laplace transformed equations can be straightforwardly solved to give: [s + 2J ]PJ,J (s) = 1 [s + (J + M )(J M + 1)]PM,J (s) = (J + M + 1)(J M )PM +1,J (s) The general result will thus be a sum of exponential decays with dierent time constants, and prefactors linear in t (degenerate decay rates exist, with M 1 M having the same rate, hence the matrix problem for eigenvalues is defective). While this approach may be appropriate numerically, it does not help to produce a qualitative understanding of the decay, so we will instead concentrate on approximately solving the density matrix equations. Semiclassical evolution after initial times As a rst approach, we may write a semiclassical approximation for how M evolves. Since J does not change during the time evolution, we will drop the label J , and assume J = N/2 throughout. d M = dt M
M

d PM = dt

[M (M 1)](J + M )(J M + 1)PM


M

= (J + M )(J M 1)

(12.10)

The semiclassical approximation is to assume that (M M )2 M 2 so that expectations of products of operators can be represented as products of expectations, and one has: d M dt (J 2 M 2 ) (12.11)

We have also assumed here that J, M 1, such an approximation will be seen later to arise naturally whenever the semiclassical approximation is valid. These equations can be solved by substituting M = J tanh(), which lead to d/dt = J , hence the general solution is: M = J tanh[J (t tD )], I = dM = J 2 sech2 [J (t tD )] dt (12.12)

Semiclassical results in Heisenberg picture The same results can be recovered in the Heisenberg picture by considering equations of motion for the operators J z , J . This approach will be useful later on, when considering an extended system. Using the commutation relations [J + , J ] =

94

LECTURE 12. SUPERRADIANCE

J J

-J 0 tD tD

Figure 12.1: Time evolution of M and the associated rate of


radiation calculated semiclassically

2J z , [J z , J ] = J , we may write: = Tr [J J + J + J ]J = J z J + . 2 (12.13) z + Similarly, one nds (d/dt) J = J J . The semiclassical approximation corresponds again to factorising products of operators, thus yielding: d z J = | J |2 , dt d J = Jz J dt (12.14) d d J = Tr J dt dt

Clearly J z = M , and so the previous equations can be recovered by writing J z = J tanh() along with J = J sech()ei . The phase is constant, and is not determined by the equations of motion. While this substitution is most suitable to solve the equations, it is also worth noting (for future reference) that another substitution is more natural in order to understand the behaviour. By noting that the components J , J z represent parts of a collective spin, one is led to write: J z = J cos(), J = J sin()ei , = J sin(). (12.15)

Hence, superradiance corresponds to the collective spin behaving as a damped pendulum, initially in its inverted state. Early time evolution In the above description, it is clear that at early times, when fully inverted, the classical equations will not start. However, at early times the semiclassical approximation fails. In order that semiclassics is valid, the distribution of M should not spread too much, and so the M dependence of the time evolution rate should be small compared to the mean evolution, i.e. d (J + M )(J M + 1) = |2M 1| dM J 2 M 2. (12.16)

It is clear that this is true as long as |M | J , thus semiclassical evolution describes the late time evolution, but not the early time. At early times,

-d <M> / dt

<M>

12.1. SIMPLE DENSITY MATRIX EQUATION FOR COLLECTIVE EMISSION 95 M spreads, but at later times no signicant spread occurs. This allows us to write: PM (t) = dsPJ s (t0 ) M M (t) |
M (t0 ) =J s

(12.17)

i.e., if we nd the density matrix at early times, then for each possible value of M arising from it, we may evolve forward semiclassically. The initial state, can be matched to the solution in Eq. (12.12), expanded for small s to give: M (t0 ) = J s J 1 2e2J (t0 tD ) (12.18)

thus the delay time appearing in the semiclassical equation is given by: tD (s) = t0 + 1 ln 2J 2J s . (12.19)

This formula allows one to translate the result of the early time evolution at t0 into the subsequent evolution of M . Small values of s, describing small deviations from an initially inverted state, correspond to longer delay times; because the inverted state is unstable the initial motion away from this point is exponential growth, hence the logarithmic dependence of delay time on the value of s. To complete the problem, we must nd how the distribution of M spreads for early times. At early times, excitation numbers remain of order M J , i.e. we may write PJ =M s and expand Eq. (12.8) for small s to give: d Ps = (2J s)(s + 1)Ps + (2J s + 1)sPs1 dt 2J [(s + 1)Ps sPs1 ].

(12.20)

Solving the rst two cases, with the initial condition P0 = 1, Ps>0 = 0, one nds: P0 = e2Jt , P1 = e2Jt (1 e2Jt ), ... (12.21)

from which one may guess the general solution: Ps = e2Jt (1 e2Jt )s This can be shown to satisfy the equation of motion, i.e.: d Ps = 2J Ps + s(e2Jt 1 + 1)Ps1 ) dt = 2J [(1 + s)Ps sPs1 ] . (12.22)

(12.23)

This is a Bose-Einstein distribution, Ps z s with z = 1exp(2Jt) which has mean excitation s = z/(1 z ) = e2Jt 1, which grows exponentially at early times.

96

LECTURE 12. SUPERRADIANCE

If we choose t0 such that 2Jt0 1, then (for large enough J ) it is possible simultaneously to full the condition for semiclassical evolution, and also to expand the above distribution for large t as giving: Ps (t0 ) = exp 2Jt0 se2Jt0 (12.24)

Putting together all of the above ingredients, one has the result: PM (t) = ds exp 2Jt0 se2Jt0 M + J tanh J t t0 1 ln 2J 2J s (12.25)

Since t0 was somewhat arbitrary, this result should not depend on t0 , and indeed it can be shown that dPM (t)/dt0 = 0 1 . Thus, one can set t0 = 0 and nd a full semiclassical approximation of the time evolution of PM . Early time evolution for Heisenberg picture The above analysis of PM for short times has an equivalent manifestation in the Heisenberg picture. For early times, one cannot factorise J + J as | J |2 , but instead, one needs to worry about correlations of the operators. This can be treated approximately by assuming that the initial value of J should be drawn from a Gaussian distribution of mean 0, and variance JJ |J + J |JJ = 2J . These quantum uctuations are then amplied by the semiclassical behaviour, just as the semiclassical evolution of PJ amplies small initial dierences in M via the dierent classical trajectories they lead to.

12.2

Beyond the simple model

The above discussion assumed a contradictory set of conditions; it assumed that the atoms are close enough that phase dierences between their coupling to light could be neglected, and that also assumed that the Coulomb interactions between the atoms could be neglected. As the following will show, one or other of these conditions can hold, but not both, and their violation will modify the superradiant behaviour.

Coulomb interactions and dephasing


If the atoms remain close together, the above treatment of coupling to propagating radiation modes is correct, but Coulomb interactions between atoms will break the indistinguishability that led to the collective enhancements. Indistinguishability is broken as each atom sees a dierent environ1 To see this, it is helpful to note that the dierential of the delta function may be rewritten using: d 1 s d f t0 ln = 2Js f [. . .] dt0 2J 2J ds

and then use integration by parts.

12.2. BEYOND THE SIMPLE MODEL

97

ment of Coulomb interactions from its neighbours. The Coulomb interaction can be written as: Hcoulomb =
i>j + + ij i j + j i .

(12.26)

Treating atoms in the dipole approximation, the dipole-dipole coupling is given by: rij ) (d |dab |2 13 . (12.27) ij = 3 2 40 rij rij Before discussing the eect of this term, we may rst notice that it depends on rather similar parameters to , and in terms of the characteristic strength, one may write: ij |dab |2 30 c3 3 = 3 3 2 32 3 40 rij |dab | 0 rij
3

(12.28)

Thus, the same condition required in order to neglect phase dierences also implies that ij . Superradiance occurred because of constructive interference between the indistinguishable pathways of atomic de-excitation. The Coulomb term makes atoms distinguishable. Physically this means that while the coupling to light in Eq. (12.4) causes transitions between symmetric atomic states, Coulomb interaction make symmetric superpositions evolve into less symmetric superpositions, thus reducing J = |J. Coulomb interactions do not however change the excitation level M , it merely transfers excitation between atoms. As such, one may estimate the relative size of eect of the Coulomb term vs M by considering the number of possible states with a given M . When M = N/2, only one state is possible, the symmetric state with J = M = N/2, thus Coulomb interactions may shift the total energy, but not modify the state. For smaller values of M , one may use the number of representations of N spins with modulus J , given by: N (J ) = (2J + 1)N ! . (N/2 + J + 1)!(N/2 J )! (12.29)

The number of possible states of given M is the number of representations N/2 having J > |M |, i.e. J =M N (J ). This behaviour leads to the following scenario. For M J , Coulomb terms do not signicantly eect the decay, but as M approaches 0 (in fact, when M J ), Coulomb terms rapidly dephase the atoms, transferring the system to states with J N , and thus suppressing superradiance. If one were to work in the electric Dipole gauge, the equations may appear dierent, as no static Coulomb term exists in the Hamiltonian. However, in this case, the low k part of the photon mediated interaction describes the same physics. This means that even though a Markovian approximation can be made for the sum in Eq. (12.2) for nearly resonant modes, the low k modes must be treated carefully, and contribute an important imaginary part (energy shift).

98

LECTURE 12. SUPERRADIANCE

Extended systems
To avoid the eects of Coulomb interactions, one may instead consider extended systems, where rij > , and so Coulomb eects become weaker. However, in this case, we can no longer neglect the phase dierences seen by dierent atoms, and so the evolution will be more complicated that Eq. (12.4). The simplest possible extended system is a one dimensional system2 . In this case, we may replace the phase factors by exp[ik (zi zj )], and forget about summation over directions of k . Note that in this replacement, we have chosen to specically consider forward propagating waves in the z direction. An equivalent description of backward propagating waves will exist, however for simplicity we neglect this complication. In this approximation, our equation of motion becomes: d (t) = dt
t

dt
ij k

gk 2

eik(zi zj )

+ + + ) + i j k . (12.30) i (k + k i j k j

and we may note that:


t = dt k

1 i(k ) + 0+

= iP

1 k

+ (k )

(12.31)

In the previous discussion of decay, we have only considered the delta function, and have neglected the principal value part. This can sometimes be valid, as the imaginary part gives an energy shift i.e. it contributes a term that looks like [He , ], describing Hamiltonian dynamics. If this Lamb shift is site independent (or if we only have a single atom) then such a shift can be absorbed into renormalisation of the bare Hamiltonian. However, it can also lead to eective interactions between the atoms. We have already mentioned one example of this; Coulomb interactions in the Dipole gauge originate from the low k contribution of these imaginary parts. In the current case, we work in the Coulomb gauge, so these low k terms are not important, however the imaginary parts of the resonant terms (i.e. for k ) will play an important role in the extended system. These terms lead to phase shifts that are responsible for causality emerging in the nal result. To see this causality, we consider: ij =
k

gk 2

eik(zi zj ) . i(k ) + 0+

(12.32)

2 To be able to truly treat a 1D system, one has two opposing constraints; the system must be narrow enough that there are not signicant numbers of transverse modes. At the same time, diraction must be small enough that propagation remains axial. These constraints are best satised by a tube of waist w L.

12.2. BEYOND THE SIMPLE MODEL

99

Concentrating on nearly resonant terms, we may approximate the density of states by /2 , and extend the sum over k to , yielding: ij 2

cdk

eik(zi zj ) = (zi zj )eik0 (zi zj ) i(ck ) + 0+

(12.33)

where k0 = /c. In the case that we may neglect phases, this becomes /2, due to the step function. If we phase shift the system operators ik0 zi i i e then we nd: d = dt z >z
i

+ + + + i j j i i j + j i
j

(12.34)

This describes causality, in that operators must always proceed from small z to larger z . Note that the previous factor of 1/2 is replaced by the restricted range of the sum. Because we have written a Markovian density matrix equation, we have neglected propagation time of electric elds, and so the above equation only applies for intermediate sized systems. For larger systems Heisenberg equations can be used to extend the range of validity. To describe the behaviour more concretely, let us consider the Heisenberg equations within this Markovian approximation. Following the procedure of Sec. 12.1 we may write: d z Tr = dt k z >z
i j

+ + z z ] , i ]j j [k , i [k

=
zj <zk

+ + k j + j k z k j .
j <zk

(12.35) (12.36)

d = 2 dt k z

These equations cannot directly be factorised if they apply to single two-level systems, but a coarse grained version, in terms of: N (z ) =
i z (z zi ) i ,

P (z ) =
i

(z zi ) i ,

(12.37)

can be considered instead. These give the semiclassical approximation dN = dt dP = 2 dt If we dene E = tions:
z z

dz P (z )P (z ) + P (z )P (z )

(12.38) (12.39)

dz N (z )P (z )

dz P (z ) then this reduces to the simpler set of equadP = 2N E , dt dE = P. dz

dN = [PE + EP ] , dt

(12.40)

100

LECTURE 12. SUPERRADIANCE

Here, E can be interpreted as proportional to the electric eld strength, resulting from the integrated polarisation of the medium to the left of the current position. (Were we to have instead written Heisenberg equations from the begining, and were we to have avoided making the Markovian approximation, then a similar set of equations would have been found, but with the retarded time = t z/c appearing in place of the t.) Equations (12.40) can then be understood by the same substitution as in Sec. 12.1, mapping them to a generaised damped pendulum problem. Writing N = N0 cos(), P = N0 sin()ei one has the equations: d i e = 2E , dt giving the result: d2 = 2N0 sin(), dzdt I |E|2 d dt
2

dE = N0 sin()ei dz

(12.41)

(12.42)

Compared to the small system equation, the principal dierence here is that this equation is now second order, i.e. it contains both damping and inertia. For suciently long samples, one thus has ringing oscillations, as each atom accumulates a dierent phase depending on where it is. This causes overshoot, and hence oscillations.

12.3

Further reading

A comprehensive review of superradiance may be found in Gross and Haroche [31].

Lecture 13

More collective eects: Cavity system and the Dicke model


In lecture 12, collective atomic decay was discussed for coupling to a continuum of modes. This lecture considers the case of collective atomic behaviour in a cavity. Two particular cases are considered: the rst section is on the possibility of spontaneous coherence in thermal equilibrium, and the second section discusses the general time evolution starting from a fully excited atomic state. In the rst section we ask whether the model in thermal equilibrium can undergo a phase transition in which the two-level systems polarise, and generate an expectation of the photon eld.

13.1

Phase transitions, spontaneous superradiance

In this rst section, we will consider the Dicke model without making the rotating wave approximation. Written in terms of spin operators si for the two-level systems, this is: g H= sz a + a s+ (13.1) i + a a + i + si . 2
i i

We wish to consider the thermodynamics of this model, by considering a mean-eld theory, in which we assume a coherent mean-eld for the photon state, and nd the associated free energy. If the free energy is minimised by a non-zero photon eld, (i.e. if such a state has a lower free energy than the vacuum state), then there is a phase transition to a spontaneous superradiant phase. For a coherent photon state exp(||2 /2 + a )|0 , the associated partition function can be written as: Z [] = e||
2

z exp 2g sx i + si i si

(13.2)

Here, we have written real and imaginary parts of as = + i , and x have used s+ i + si = 2si . One may then evaluate the sum over states of 101

102

LECTURE 13. THE DICKE MODEL

the two-level system by transforming to the diagonal basis in the presence of eld . The eigenvalues of the 2 2 matrices are then E/2, with 2 + 4g 2 2 , and so the free energy is given by: E= F () = ||2 N ln cosh 2
2

+ 4g 2 2

(13.3)

To determine where a non-zero value of can minimise F , we should nd the derivative of F with respect to , to locate possible stationary points. Clearly, dF/d = 2 , and so for any stationary point we must have = 0. For the real part of we have instead dF N = 2 d 2 4g 2
2

+ 4g 2 2

tanh

+ 4g 2 2 .

(13.4)

Using the eigenvalue energies E dened above, this may be rewritten in the simpler form: dF N g2 E . (13.5) = 2 tanh d E 2 A non-zero solution exists if the term in brackets can be made to vanish. In order that this vanishes at some temperature T > 0, we require that the function tanh(E/2)/E crosses the value /N g 2 . As illustrated in Fig. 13.1, at small temperatures, tanh(E/2)/E 1/E , while at large temperatures it becomes 1/2T . Thus, for a crossing to occur, one requires 1/E > /N g 2 , and since E > | |, a transition may exist if < N g 2 / .

Small T: tanh(E/2)/E = 1/E

tanh(E/2)/E

0.6 0.4 0.2 0 0

reas i

ng E

0.8

/N g2 Large T: tanh(E/2)/E = 1/2T

Dec
1

Temperature

Figure 13.1: Graphical representation of Eq. (13.5). It is clear


that for a solution to exist, it is necessary for 1/| | > 1/E > /N g 2 .

Now, recalling the denition of g in terms of microscopic parameters and matrix elements in Eq. (1.37), we have the condition: < N V 2|dab |2 0
2

(13.6)

13.2. NO-GO THEOREM: NO VACUUM INSTABILITY

103

Thus, apparently a transition can occur if the density of two-level systems is large enough. This phase transition was originally discussed by Hepp and Lieb [32, 33], and in the form presented here by Wang and Hioe [34].

13.2

No-go theorem: no vacuum instability

If the density of two-level systems is large enough, the Dicke model is invalid, as we have neglected the matter-radiation coupling originating from the A2 terms. Let us restore these terms, and see at what density of twolevel systems they become important. In the same single-mode approximation in which we are treating the quantised radiation, the correction due these terms can be written following Eq. (2.1) as: HA2 = N (a + a )2 ; = q2 . 4m0 V (13.7)

(NB, the mass m appearing here for the contribution from a single atom is 1/m = 1/me + 1/mi , i.e. is the reduce electron-ion mass.) Including this term in the Hamiltonian, the function F () acquires an extra dependence on , giving F () = ||2 + N 4 2 N ln cosh 2
2

+ 4g 2 2

(13.8)

In terms of the condition for a phase transition to occur, this means one should replace + 4N ; i.e. the condition is now ( + 4N ) < N g 2 Physically, this means that the A2 terms in the Hamiltonian describe a dielectric response of the atoms, opposing large transverse elds; as a result, the energy cost of spontaneous polarisation has increased, and sponatneous eects can arise only if the two-level system susceptibility is larger than it was above. Combining the dentions of and g , we may write the new condition as: N q2 N 2|dab |2 2 + < . (13.9) V m0 V 0 As a function of density of two-level systems, this condition is clearly not satised at N/V = 0, but will be satised for some nite density of two-level systems if and only if: q2 2|dab |2 2 < . (13.10) m0 0 Thus, spontaneous polarisation requires that the energy and dipole matrix elements of an atom can satisfy 2|dab |2 > q 2 /m. Reaclling the denition of the oscillator strength, the above requirement can be written as fab = 2m|dab |2 /q 2 > 1. However, the oscillator sum rule imples that b fab = 1. Since fab = 0 when a is the ground state, it is then clear that it is impossible for any term in this sum to exceed 1. Hence, the sum rule imposes the opposite inequality to the one we require for a phase transition: the sum rule prevents such a phase transition from occurring. This observation was rst pointed out by Rz az ewski, W odkiewicz, and Zakowicz [35].

104

LECTURE 13. THE DICKE MODEL

13.3

Radiation in a box; restoring the phase transition

The above result shows that the phase transition by which the vacuum state becomes unstable, and spontaneously polarises is an artefact of that model, and not physical. This however is not the end of the story. There is a simple extension to the Dicke model, which is appropriate in a variety of recent works on conned quantum optical systems, which can restore the phase transition. That extension is to consider a closed system with a density of excitations. This means considering the at-rst counterintuitive idea of a chemical potential for photons. The idea of a chemical potential for photons is not generally considered because photons cannot easily be conned, and so a xed density of photons is hard to achieve. Without connement, photons can be exchanged with the bulk, which acts as a reservoir at zero chemical potential. However, in engineered cavity quantum electrodynamics systems, photon connement is exactly what is being created, and so for such systems, including polaritons in semiconductor microcavities, Josephson junctions in microwave waveguides, and atoms in superconducting mirror cavities, a chemical potential for photons is a useful concept. Adding a chemical potential, we have = H M, HH M=
i

sz i +

1 2

+ a a.

(13.11)

The net result is to replace = and = in the previous Hamiltonian (although the factors of and in the coupling strengths remain unmodied). In this case, it is clear that one can satisfy the condition ( + 4N ) < N g 2 by ensuring the chemical potential is close enough to the two-level system energy. Since this may be done at any density of two-level systems, we may assume a low density, and so neglect i.e. neglect the A2 terms in the Hamiltonian. Since a transition may occur, it is interesting to nd the critical temperature of this transition. From the adapted form of Eq. (13.5), this condition is: ( ) ( ) = N g 2 tanh . (13.12) 2T This can be combined with the equilibrium expectation of the number of excitations M in the absence of an expectation of , which gives. M 1 = 1 tanh N 2 2T . (13.13)

13.4

Dynamic superradiance

In this section, we consider the general time evolution of the Dicke model, starting from a fully inverted atomic state. This describes the analog of the superradiance discussed in Sec. 12 in the case where there is a good cavity.

13.4. DYNAMIC SUPERRADIANCE

105

For illustration, we discuss here the simplest case of = , and so our model is: g H = J z + a J + aJ + + a a. (13.14) 2 To study the semiclassical dynamics, we rst write the Heisenberg equations of motion: x = i[H, Jx ] = J y + g i a a J z , J 2 g y x = i[H, Jy ] = +J J a + a J z , 2 z = i[H, Jz ] = g i a a J x + g a + a J y , J 2 2 g x y ia = [H, a] = (J iJ ) + a. 2 (13.15) (13.16) (13.17) (13.18)

The semiclassical approximation is then to replace these equations for noncommuting operators for equations for their commuting expectations, and to factorise the expectations, so (a + a )J z a + a J z . Doing this, and writing a = = + i the above equations can be written as: g g = h J, h = g , i = (Jx iJy ) + . (13.19) J 2 In order to solve these equations, it is rst convenient to transform to a rotating frame. This means substituting eit , and J J eit , i.e. cos(t) sin(t) 0 cos(t) 0 J. J sin(t) (13.20) 0 0 1 This then gives = h J, J g h = g , 0 i = g (Jx iJy ) . 2 (13.21)

, we can make the ansatz: At this point, starting in the excited state J = J0 z 0 0 cos() . = J0 J (13.22) J = J0 sin() , cos() sin() Then, substituting this into the equations of motion, we have: 0 cos() cos() = gJ0 cos() , J0 sin() sin()

(13.23)

= g . Then, the equation for = gives: which implies = 0 and = gJ0 sin(). 2 (13.24)

106

LECTURE 13. THE DICKE MODEL

Thus, the angle dening the Bloch vector obeys the equation of motion:
2 = g J0 sin(). = g 2

(13.25)

This is the equation of motion for an (initially) inverted pendulum, and so the Bloch vector makes traces out great circles passing through the entirely inverted, and entirely empty states. One should compare this second order equation for dynamical superradiance in a cavity to the rst order equation in the decaying case discussed in Sec. 12.1 of the previous lecture. When non-resonant, i.e. = , the comparable equations are more complicated, however the general result is similar: after transformation to a rotating frame, the Bloch vector traces out circles (see Fig. 13.2), and can always be mapped to the problem of an inverted pendulum.

Jz

Jy Jx

Figure 13.2: Bloch sphere, and paths that Bloch vector takes in
the rotating frame, in the general (non-resonant) case.

13.5

Further Reading

The original phase transition of the Dicke model is discussed in Refs.[32 34]. The no-go theorem is introduced in Ref.[35] and elaborated further by, e.g. Ref.[36]. The phase transition of a model with a chemical potential is discussed by Eastham and Littlewood [37, 38]. As hinted to in the discussion on dynamic superradiance, the Dicke model is integrable, and this integrability extends even to the case of a distribution of two-level system energies: H=
i z i si

+ a a +
i

g a si + as+ . i 2

(13.26)

The integrability of this model can be proven by dening a vector function L(z ), having the following properties: The modulus of the vector L(z ) is conserved.

13.5. FURTHER READING

107

The Taylor expansion in z contains as many powers of z as there are two-level systems. Together, these mean that L(z ) contains sucient conserved quantities that the system is integrable. This method is discussed in specic case of the Dicke model by Yuzbashyan, Kuznetsov, and Altshuler [39]. The history of this integrability, and its relation to a wider class of models is discussed in the review article by Dukelsky, Pittel, and Sierra [40]. The time-dependent solution, starting from the excited state is given by Barankov and Levitov [41].

Lecture 14

Three levels, electromagnetically induced transparency, and coherent control


Most of the lectures so far has been restricted to two-level atoms. While there many are obvious extensions to the case of more than two levels, this section focuses only on one phenomena which requires three levels: That is the coherent interference of pathways for atomic transitions, and the possibility of suppressed transitions. By using an applied electric eld, one can suppress transitions between levels, and thus prevent absorption, leading to Electromagnetically Induced Transparency. This lecture starts by discussing the phenomena in terms of the density matrix evolution, allowing incoherent decay of atomic energy levels, and considering radiation semiclassically. In this way, we will see that conditions under which the group velocity of light is signicantly reduced while absorption of light remains small is possible in this system a combination of features that does not occur for a two level system. Then, by considering the atomic wavefunction evolution, it is possible to understand the underlying interference phenomena which lead to the cancellation that prevents absorption. Finally, we return to a quantum optics description of the states that are free from absorption dark state polaritons discussing how the nature of these coherent quantum states can be controlled.

14.1

Semiclassical introduction

We consider the level scheme shown in Fig. 14.1. Dipole matrix elements exist between the ground state b and the excited state a, and also between the metastable state c and excited state a. The transition between b and c is however dipole forbidden. The states a and c are coupled by a strong pumping radiation eld, described only by the eective Rabi splitting it induces, p . The transition between a and b is then probed by a weaker 109

110

LECTURE 14. THREE LEVELS, AND COHERENT CONTROL

eld. We allow the possibility of incoherent decay from a to b, and also a (much weaker) decay from c to b. The coherent part of the evolution can

a E.dab ab b
driven, transitions between them.

p c cb

Figure 14.1: Scheme of atomic energy levels and allowed, and

be described by the Hamiltonian: a Edab eit p eip t it . b 0 H = Ed ab e i t p p e 0 c

(14.1)

Susceptibility; slow light.


Let us consider the absorption and emission coecients, by nding the susceptibility of the three level system; i.e. nding how the polarisation, P = (Natoms /V )dab ab , depends on the eld E applied resonantly with the a b transition. We assume the eld E is weak, so we consider only linear response, and can assume the populations of the a and c levels remain small. , We are interested in the evolution of the density matrix, = i[H, ] + L represented a superoperator, describing the incoherent evolution, where L described in lecture 3. It is convenient to make a gauge transform so as to remove the explicit time dependence in the above Hamiltonian, this gives: a Edab p . 0 H = Ed (14.2) ab b + p 0 c + p Let us now assume that the pump is resonant, so that c + p = a , and let us rewrite b + = a , so characterises the detuning of the probe below the transition. Since Tr = 1, small populations of a, c imply bb 1 and ac 0, and Hermiticity means the only relevant components are ab and cb . Thus: ab = i [(Haa ab + Hab 1 + Hac cb ) (0 Hab + ab Hbb 0 Hcb )] ab ab = (i + ab ) ab + ip cb + iEdab , cb = i [(Hca ab + 0 1 + Hcc cb ) (0 Hab + cb Hbb 0 Hcb )] cb cb = (i + cb ) cb + i p ab . (14.4) (14.3)

14.1. SEMICLASSICAL INTRODUCTION These can then be written as t ab cb = (i + ab ) ip ip (i + cb ) ab cb + iEdab 0

111

(14.5) = M X + A, and we want the steady state solution (afThis has the form X ter decay of any initial transients), which is clearly given by X = M 1 A. Thus we can extract the polarisation by nding ab : P = Natoms Natoms (i + cb )iEdab dab ab = dab V V (i + ab )(i + cb ) + |p |2 (14.6)

Then one can easily identify the susceptibility, by P = 0 E . i.e. = i(i + cb ) Natoms |dab |2 . V 0 (i + ab )(i + cb ) + |p |2 (14.7)

Let us now discuss the properties of this susceptibility. It is clear that () = () , thus if = + i , it is clear that is an odd function of (and vanishes at = 0), and an even function. However, ( = 0) cb , which we assumed earlier to be small, as this transition is dipole forbidden. To be precise, ( = 0) cb /(ab cb + |p |2 ); thus the absorption part is small if cb p ; the coherent driving eld has induced transparency. The full real and imaginary parts of susceptibility are plotted in Fig. 14.2, both with and without a driving eld p .
1

Susceptibility

0.5 0 -0.5 -1 2 1 0

p=1 p=0 Real part

Susceptibility

Imaginary part
-1 -3 -2

Detuning from resonance

-1

Figure 14.2: Real and imaginary parts of susceptibility, plotted


for 1 = 0.5, 3 = 0.01, and two values of p as indicated.

Since both real and imaginary parts of are zero, or very small at = 0, let us consider linear expansion; clearly only the real part exists to linear order. Setting cb = 0, one trivially nds: = Natoms |dab |2 + O(2 ). V 0 |p |2 (14.8)

112

LECTURE 14. THREE LEVELS, AND COHERENT CONTROL

The linear part of susceptibility relates to the dispersion of the probe eld; n() = 1 + (), which can be used with = ab , and n( ) = ck . One thus has phase velocity vp = c/n() and group velocity vg = c/[n (ab )dn/d]. Assuming remains small, one can expand: n1 Natoms |dab |2 , V 0 2|p |2 dn Natoms |dab |2 1 . d V 0 2|p |2 c 1+
N ab |dab |2 |p |2 20 V

(14.9)

Thus, to leading order: vp c, vg , (14.10)

where we have rearranged factors to identify the atom-radiation coupling strength g 2 = ab |dab |2 /20 V [cf. the denition in Eq. (1.37), in the case k = = ab ]. Thus, if N g 2 /|p |2 is large, one can have small absorption, yet large enough dispersion to signicantly reduce the group velocity, and thus have non-absorbed slow light.

Decay length of slow light


The previous analysis shows that near zero detuning, one can have a small group velocity, but small absorption. Let us now be more specic about how small the absorption is. What matters in practice is the attenuation of the signal as it propagates, following I (l, ) = I (0)ekl () . Since the attenuation is frequency dependent, this may also lead to distortion of the propagating signal. The frequency dependence of the transmission intensity is shown in Fig. 14.3.
0.4

0.3

Transmission

0.2

0.1

0 -5

-4

-3

-2

-1

Detuning from resonance

Figure 14.3: Transmission after L = (50/ ), for the same parameters as in Fig. 14.2.

Considering the case cb = 0, for small , the second order expansion yields: N g 2 2 ab 2i , (14.11) ck |p |4

14.2. COHERENT EVOLUTION ALONE; WHY DOES EIT OCCUR 113 where we have written ab = ck . One can thus identify a frequency dependent decay length: 1 1 |p |4 L=c . (14.12) 2N g 2 2 ab

14.2

Coherent evolution alone; why does EIT occur

Let us now ignore the decay rates, ab and bc , and look at the evolution of atomic states alone. Thus, since we are not interested in the dielectric response, we may denote Edab = s as the signal eld. Considering the fully resonant case, = a b (i.e. = 0) as well as p = a c , one can write: | = ca eia t |a + cb eib t |b + cc eic t |c 0 Ed p ca ca 0 0 cb . it cb = Ed p 0 0 cc cc (14.13) (14.14)

One can easily nd the eigenvalues and vectors of this problem; in terms of 0 = |Ed|2 + |p |2 one has: 0 1 p ; 0 Ed 0 1 Ed . 20 p

=0:

= :

(14.15)

Thus, if one starts in the state (p |b Ed |c ) /0 , then the evolution never reaches the excited state |a . Since the only substantial decay constant was from the state |a , this then means one has no absorption: The evolution is trapped in this dark state, and shows no decay. One may also describe this process as destructive interference between the transitions |b |a and |b |a |c |a etc. Since such a dark state exists, from which no further excitation is possible, the full evolution (with decay rates) can be understood as follows: Regardless of the initial state, decay out of states |a allows one to reach various superpositions B |b + C |c . Eventually, one will reach dark state, after which no further evolution can occur.

14.3

Coherent control of quantised radiation: Dark state polaritons

The previous sections show that the existence of the classical driving eld, resonant with the c a transition modify the response seen by a probe eld at the b a transition: One can have vanishing absorption, while still having strong dispersion, allowing propagation of slow light. In this section, we consider the probe eld quantum mechanically, and ask how the classical driving changes the evolution of quantised radiation. As illustrated

114

LECTURE 14. THREE LEVELS, AND COHERENT CONTROL


a Quantum Field

Classical Driving c

Figure 14.4: Illustration of which transitions are classically


driven, and which described quantum mechanically

in Fig. 14.4, we consider the driving as still classical, but the probe eld is quantum mechanical. In order to consider propagation through the EIT medium, we consider a continuum of radiation modes, as well as explicitly including a sum over atomic states. Note that in the earlier treatment, the sum over atoms occurred implicitly via the factor Natoms in Eq. (14.7) Assuming resonance xy for the driving eld, and writing i for the three by three matrix describing transitions from level y to level x of atom i, one can write: H=
k>0

cka k ak +
i

ba iab t gi e k

ikxi a ke

ca p i + H.c. .

(14.16) Then, making a gauge transform to a rotating frame one can remove the explicit eiab t time dependence, in favour of replacing ck ck ab . Then, one may make a Fourier transform of the photon eld, dening: a(x) =
k

ak ei(kab /c)x .

(14.17)

By writing k ab /c rather than k , one recovers the correct energy from the derivative w.r.t. x. Thus, one can write H= dxa (x)(icx )a(x) +
i ba ca gi a (xi )eiab xi /c p i + H.c. .

(14.18) Note the factor of eiab xi /c appearing in the photon-atom coupling term. Let us introduce collective variables: X (x) =
i ca i (x xi ) ba i (x xi )eiab xi /c i bc i (x xi )eiab xi /c . i

P (x) = S (x) =

(14.19)

The denitions here are not the most symmetric, but are physically motivated: P represents a polarisation wave, while S represents a spin wave this language is appropriate to cases where the low lying levels b, c arise

14.3. DARK STATE POLARITONS

115

from opposite spin states. The denitions of P (x) and X (x) are easily understood, as they allow one to rewrite Eq. (14.18) as: H= dxa (x)(icx )a(x) gP (x)a (x) + p X (x) + H.c. . (14.20)

To understand the denition of S (x), we must consider commutation relations. Writing commutators using [ ij , kl ] = jk il li kj , the cross commutators are given by: [P (x), X (x )] = S (x) (x x ), [S (x), X (x )] = P (x) (x x ), [P (x), S (x )] = X (x) (x x ). (14.21)

Other than these, and their conjugate forms, all other cross commutators vanish. Thus, it is clear that the phase factors in the denition of S (x) in Eq. (14.19) is required to give these simple forms. We now have a tractable form of the quantum Hamiltonian, which allows us to consider how an initial state will evolve. Let us solve the simpler problem, of nding the quantum states that are trapped in the absorptionfree subspace. It is clear that the atomic ground state, with no photons is an eigenstate, which we will denote |0 . Let us consider excited states of the form: | = |0 = dx(x) a (x) + S (x) |0 , (14.22)

and ask for the condition under which no excitations to the excited atomic state, a, occur. Note the distinction to the discussion in Sec. 14.2; in that case we wanted superpositions of b and c states so that in the presence of a given probe electric eld there were no excitations; in the current case we want a superposition of radiation and spin-wave excitations. Since the trial state in Eq. (14.22) contains no excitations, we require only that the time derivative of number of excitations vanishes. Thus, the condition we require can clearly be written as: 0| P [H, ] |0 = 0. Let us consider the commutator, using the results in Eq. (14.21), along with [a(x), a (x )] = (x x ): H, a (x) + S (x) =

icx a (x) gP (x) + ga (x)X (x) p P (x) . (14.23) The term proportional to a (x)X (x) is second order, involving both atomic and radiation excitations, for low intensities it can be neglected. The number of excitations then vanishes under the condition: g + p = 0. If we wish to nd a normalised wavefunction, we require: (x x ) = [(x), (x )] = ||2 (x x ) + | |2
i

(x x ).

(14.24)

Thus, one has 1 = ||2 + | |2 in addition to the previous condition, thus: = p , |p |2 + g 2 N = g . |p |2 + g 2 N (14.25)

From Eq. (14.23), we nd this state evolves according to the Heisenberg equation of motion, it (x) = [H, ] = icp x a (x). |p |2 + g 2 N (14.26)

To eliminate a (x), we wish to write this in terms of the eld (x). The eld is the linear combination of a and S which is dark i.e. never undergoes a transition to the excited atomic state; another orthogonal linear combination of a and S exists, let us call that . Then, one can write: 1 p g N a . (14.27) = NS g N |p |2 + g 2 N p Here we have included a factor of 1/ N required to achieve bosonic com mutation relations into the denition of a bosonic spin operator S / N . Inverting this to nd a in terms of and , and assuming an initially dark state so that 0, Eq. (14.26) becomes: it (x) = i.e. icp x |p |2 + g 2 N t + p |p |2 + g 2 N (x), (14.28)

c z = 0. (14.29) 1 + g 2 N/|p |2 Thus, we have found a dark state polariton [42], a state which is never excited to the radiative atomic state a, and is half light, half spin-wave excitation. It propagates at a reduced velocity, which is the group velocity found in Eq. (14.10). Importantly, one can control the nature of the state by the ratio of p to g N . In a strong driving eld p g N , the excitation is almost pure photon; for a vanishing eld p g N , the excitation becomes pure spin wave. Thus, starting with a large p , one can inject an arbitrary quantum state of photons. By adiabatically reducing p , the quantum state is exactly transfered to the spin waves. These propagate slowly, until p is increased again, and the state transfered back to photons. Hence, coherent control allows the trapping of quantum states of light.

14.4

Further reading

The general topics of EIT, dark states, and coherent control in three level systems are discussed in a review by Fleischhauer et al. [43]. The question of coherently transferring the quantum state of light to dark state polaritons is introduced in Ref.[42], and discussed further in Ref.[44]. An interesting application of this idea is described in Ref. [45], where a quantum state of light is transported by transferring atoms between two condensates, and then releasing the state of light from the second condensate.

Problem Set A

Problems for lectures 14


Problem A1: Oscillator strength and Thomas-Reiche-Kuhn sum rule (a) Consider the response of the n-level atom to the applied electric eld, as described by the perturbation given in Sec. 1.4, H = H0 + H1 with H0 = m m |m m| and the perturbation: H1 = i
nm

A(t)

nm dnm |n

m|,

A(t) = i

E ( ) (i)t e . (1.41)

where is an innitesimal term added to ensure that A(t) vanishes at t , and nm = n m . If the initial state is |m , show that to rst order in time dependent perturbation theory one has: ( i )t d E ( ) e nm nm | = ei m t/ |m + d |n . (1) 2 + i nm /
n=m

Hence calculate the polarisation, and nd the susceptibility: ( ) = (b) e2 m0 0 fnm


n=m

( +

i )2

1 (

nm /

)2

(1.43)

Prove that for suitably dened H0 the equality: [r , [H0 , r ]] =


2

(2)

holds, stating the most general form of H0 for which this is true. (c) Using the above commutator, show that |dnm |2 (
n n

m)

2q2

2m

(3)

and thus that for transitions involving a single electron, (i.e. if q = e, m = m0 and the sum over contains only one term), then m fnm = 1. 117

Problem A2: Collapse and revival Consider the linear model: H = a a + c c + g (a c + c a). (4)

If one starts from a coherent state of photons, exp(a )|0, 0 , what is the expectation of occupation of the atomic mode c after time t. Explain why there is no collapse of the Rabi oscillations. Problem A3: Dephasing from fast-correlated noise As another way to consider dephasing due to noise with fast correlations, consider interaction between the two-level system and a bath of radiation modes given by the interaction picture coupling: HSR =
n in t z n Bn ein t + Bn e

(5)

where Bn are the quantised modes of a eld whose eld strength shifts the energies of the two-level system. (a) By making the Born-Markov approximation, show that this leads to a density matrix equation of the form: t |noise = 0 ( z z ) , 2 (6)

and determine how 0 depends on the temperature of the reservoir if the reservoir follows a thermal Bose-Einstein distribution. (b) Show that this describes pure dephasing, i.e. it has no eect on the diagonal elements of the density matrix, and leads to a decay of the o-diagonal terms t = 0 .

Problem A4: Absence of power broadening for an harmonic oscillator Consider an harmonic oscillator mode of energy , coherently driven by a eld of strength at frequency , and coupled to a continuum of radiation modes. This system is described by the density matrix equation: t = i[H, ] b b 2bb + b b 2 H = (b eit + H.c.) + b b (7) (8)

(a)

Write the density matrix equation in the number basis and show that the ansatz: n m ieit ieit 1 (9) nm i + /2 n!m! i + /2 satises this equation of motion where = .

(b)

Comment on the nature of the state described by Eq.(9)

(c)

From the normalised density matrix, nd the average excitation probability, and show that there is no power broadening.

Problem A5 (Extended problem): Entropy in the collapse and revival of Jaynes Cummings oscillations This question shows that the excited state probability does not fully characterise what is happening in the collapse and revival of oscillations in the Jaynes-Cummings model. To characterise these oscillations further we study the entanglement between the two level system and the photon eld. This requires rst nding the reduced two-level system density matrix tls tls,ab =
n

n, a| |n, b

(10)

where a, b {, } are states of the two level system, and n are photon number states. The degree of entanglement is characterised by the extent to which is a mixed density matrix, one measure of this is R2 = 1/4 Det(). Any density matrix for a two-level system can be written as: tls = so that R2 = X 2 + Y 2 + Z 2 . (a) Find an expression for the matrix X, Y, Z as a sum over photon number states. Show that Z vanishes except near revivals. Considering the time between revivals, where Z can be neglected, show that: 1 gt g 2 t2 R sin exp (12) 2 4 128||4 Comment on the relation between the time dependence of R and the location of the revivals. Why are they related in this way? +Z X iY
1 2

X + iY 1 2 Z

(11)

(b) (c)

(d)

Problem Set B

Problems for lectures 59


Problem B1: Pseudo-mode decay rate. Consider a standing wave conned within a cavity of length a, 0 (x) = 2/a cos[(2n + 1)x/a] (1)

As discussed in Sec. 5.1), if cavity is not perfect, this state is not an eigenstate. (a) By decomposing the wavefunction in Eq. (1) onto the true eigenmodes given in Sec. 5.1), i.e. Ek (|x| < a/2) = f (x) 2 L 1
2 1 + sin(ka) + 1 2 (1 + cos(ka))

cos(kx) (2)

show that: 0 | (t) =


k

a a a sinc (k + kn ) + sinc (k kn ) L 2 2

eickt 2 1 + sin(ka) + 1 2 (1 + cos(ka))

(3)

where kn = (2n + 1)/a. (b) Hence, for a good cavity (where 2 exponentially at rate /2 = 2c/2 a. 1), show that the overlap decays

Problem B2: Fluorescence of a harmonic oscillator Following the same method as in lectures, calculate the uorescence spectrum of a driven harmonic oscillator. Why is this answer so much simpler than the two-level system? Problem B3: Quantum jumps for a decaying coherent state Consider a single leaky photon mode obeying the density matrix equation of motion: (4) = a a + a a 2aa , 2 with an initial a coherent state = e|| ea |0 0|e 121
2 a

(a)

Show that the time dependent density matrix equation is solved by allowing to be time dependent, and nd the equation of motion must satisfy. Now consider this question in the quantum jump formalism. Introducing |n as the normalised quantum state that occurs after having lost exactly n photons, the quantum jump formalism describes the density matrix as: (t) =
n

Pn (t)|n (t) n (t)|

(5)

(b)

Show that |n (t) = |0 (t) ; (i.e. the state is unchanged after photon loss), and hence show that show that (t) = |0 (t) 0 (t)|. Find |0 (t) from the non-Hermitian Hamiltonian part of the quantum jump formalism, and show that the nal result matches your result from the rst part of this question. Give a physical interpretation to the above results.

(c)

(d)

Problem B4: Maxwell-Bloch equations (a) Starting from the Maxwell-Bloch equations given in Eq. (9.5) and Eq. (9.7), show that by an appropriate rescaling = 0 x, P = P0 y, N = N0 ( r z) these can be put in the form: x = (y x), 2 y = t ([ r z ]x y ), z = xy + x y z 2 (6)

Give expressions for 0 , P0 , N0 , r and comment on their signicance. (b) Optional extra Assuming that x, y have the same phase, comment on the known behaviour of these equations when t and r 1. In the limit , t , for values of r not too large compared to threshold, show that these can be reduced to: x = r 1 r |x|2 x 2 (7)

(c)

Compare this to the result found from the Master equation: t n = r 2 r r n + 2 (n + 1)2 . 2 2 (9.33)

where = g 2 /t and comment on where the dierences are important.

Problem Set C

Problems for lectures 1014


Problem C1: Four level laser This question concerns a four level lasing scheme, depicted in Fig. C.1, and how it reduces to the three level scheme discussed in lectures. In addition to the levels 0, e, g considered in the three level case we now also 01 10 have a level 1, with a coherent coupling H H + 1 2 ( + ), and the density matrix equation has an additional term: p 11 t t 2 e1 1e + 11 . (1) 2

1 e g p

lasing transition

0
Figure C.1: Labelling of four level scheme

(a) (b)

Show that in the limit that p is rapid compared to other rates, one may 2 + 2 )]00 where ij = Tr ji write 11 = [2 /(p atoms ( ) as before. Hence show that in the four level scheme the eective pump rate becomes: r= p 2 Na , 2 (p + )2 + p (2)

Problem C2: Superradiance transition from linearisation Consider the Dicke model Hamiltonian: g H = Jz + a J + aJ + + a a. 2

(13.14)

123

(a)

By using the Holstein-Primako representation of a spin, J + = b N b b, J z = b b N/2, and expanding for b b N , write the leading order Hamiltonian in the form: H = (a b )M a b (3)

giving an explicit expression for the matrix M. (b) By nding an appropriate unitary transformation, or otherwise, put this in the form H = E1 c c + E2 d d, where c, d are linear combinations of a, b. Comment on the signature of the Dicke phase transition appearing in the behaviour of E1,2 . (Note that Eq. (13.14) has no counter-rotating terms, and so the critical coupling strength obeys g 2 N = 4 .)

(c)

Problem C3: Superradiance and dephasing Consider two level systems with either individual or collective decay and dephasing: LA [] = 2
z z z z z z [i i 2i i + i i ] i

and LB [] = (a) ( 2 2 (
i z i )( i + i ) 2( i

+ + + i i 2i i + i i i

(4)

z i ) 2( i + i )( i

z i )( i

z i ) + ( i i )( i i

z i )( i + i )

z i )

i )( i

i +)( i

(5)

Calculate the linear absorption spectrum for these two problems, by considering = i[H0 , ] + LA,B [] where H0 =
i + it it . i e + i e

(6)

Work in the weak pump limit where the diagonal parts of the density matrix are unchanged (i.e. the population remains in the ground state), and only coherences between ground state and a single excited state need be considered. Problem C4: EIT with inverted pump and probe Starting from the EIT Hamiltonian in the rotating frame: 0 Ed 0 , H = Ed 0 0

(7)

consider what happens when the level of pump and probe transitions are inverted (see Fig. C.2): = i[H, ] ab ab ba 2 ba ab + ab ba bc bc cb 2 cb bc + bc cb (previously, the second term decay term would have had c b.) (a) (b) (c) Find aa , bb , cc in the absence of the probe E Find ab to leading order in E , using the above results. Find the susceptibility in the case where ab = bc /2 = . (8)

a E.dab ab b cb c
Figure C.2: Inverted level scheme; probe couples metastable
states.

Problem C5 (Extended problem): Second-order correlation functions (a) Using the quantum regression theorem, give an expression for the second order correlation function: g2 ( ) = a (t)a (t + )a(t + )a(t) a (t)a(t)
2

(9)

in terms of the equations of motion and initial conditions for diagonal elements of the density matrix. (b) Say as much as you can about the form this correlation function takes in the limits far below and far above threshold.

Bibliography
[1] R. Miller, T. E. Northup, K. M. Birnbaum, A. Boca, A. D. Boozer, and H. J. Kimble, Journal of Physics B: Atomic, Molecular and Optical Physics 38, S551 (2005), DOI: 10.1088/0953-4075/38/9/007. A. Wallra, D. I. Schuster, A. Blais, L. Frunzio, R.-S. Huang, J. Majer, S. Kumear, S. M. Girvin, and R. J. Schoelkopf, Nature 431, 162 (2004), DOI: 10.1038/nature02851. A. Blais, R.-S. Huang, A. Wallra, S. M. Girvin, and R. J. Schoelkopf, Phys. Rev. A 69, 062320 (2004), DOI: 10.1103/PhysRevA.69.062320. J. M. Raimond, M. Brune, and S. Haroche, Rev. Mod. Phys. 73, 565 (pages 18) (2001), DOI: 10.1103/RevModPhys.73.565. G. Khitrova, H. M. Gibbs, M. Kira, S. W. Koch, and A. Scherer, Nature Physics 3 (2007). C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Photons and Atoms (Wiley, New York, 1989). M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University Press, Cambridge, 1997). Y. Yamamoto and A. Imamo glu, Mesoscopic Quantum Optics (Wiley, New York, 1999). J. Gea-Banacloche, Phys. Rev. DOI: 10.1103/PhysRevLett.65.3385. Lett. 65, 3385 (1990),

[2]

[3]

[4]

[5]

[6]

[7]

[8]

[9]

[10] P. Meystre and M. Sargent III, Elements of Quantum OPtics (Springer, Berlin, 2007), 4th ed. [11] A. Ottl, S. Ritter, M. K ohl, and T. Esslinger, Rev. Sci. Instrum. 77, 063118 (2006). [12] S. Kuhr, S. Gleyzes, C. Guerlin, J. Bernu, U. B. Ho, S. Deleglise, S. Osnaghi, M. Brune, J.-M. Raimond, S. Haroche, et al., Appl. Phys. Lett. 90, 164101 (2007). [13] D. Steck, Tech. Rep., University of Oregon (2008), URL http:// steck.us/alkalidata. 127

[14] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Phys. Rev. A 76, 042319 (2007), DOI: 10.1103/PhysRevA.76.042319. [15] E. Peter, P. Senellart, D. Martrou, A. Lema tre, J. Hours, J. G erard, and J. Bloch, Phys. Rev. Lett. 95, 067401 (2005), ISSN 0031-9007, DOI: 10.1103/PhysRevLett.95.067401. [16] T. Yoshie, A. Scherer, J. Hendrickson, G. Khitrova, H. M. Gibbs, G. Rupper, C. Ell, O. B. Shchekin, and D. G. Deppe, Nature 432, 200 (2004), ISSN 1476-4687, DOI: 10.1038/nature03119. [17] T. F. Gallagher, Rydberg Atoms, Cambridge Monographs on Atomic, Molecular and Chemical Physics (Cambridge University Press, Cambridge, 1994). [18] M. Lax, Phys. Rev. 129, Rev.129.2342. 2342 (1962), DOI: 10.1103/Phys-

[19] C. W. Gardiner and P. Zoller, Quantum Noise (Springer-Verlag, Berlin, 2004), 3rd ed. [20] G. W. Ford and R. F. OConnel, Phys. Rev. Lett. 77, 798 (1996), DOI: 10.1103/PhysRevLett.77.798. [21] G. W. Ford and R. F. OConnel, Optics Commun 179, 451 (2000), DOI: 10.1016/S0030-4018(99)00608-2. [22] M. Lax, Optics Commun. 179, 463 (2000), DOI: 10.1016/S00304018(00)00622-2. [23] H. Haken, Rev. Mod. Phys. 47, 67 (1975). [24] P. R. Rice and H. J. Carmichael, Phys. Rev. A 50, 4318 (1994), DOI: 10.1103/PhysRevA.50.4318. [25] J. McKeever, A. Boca, A. D. Boozer, J. R. Buck, and H. J. Kimble, Nature 425, 268 (2003). [26] O. Astaev, K. Inomata, A. O. Niskanen, T. Yamamoto, Y. A. Pashkin, Y. Nakamura, and J. S. Tsai, Nature 440, 588 (2007). [27] D. A. Rodrigues, J. Imbers, and A. D. Armour, Phys. Rev. Lett. 98, 067204 (2007), DOI: 10.1103/PhysRevLett.98.067204. [28] D. A. R. J. Imbers, T. J. Harvey, and A. D. Armour, New J. Phys 9, 84 (2007). [29] H. Walther, B. T. H. Varcoe, B.-G. Englert, and T. Becker, Rep. Prog. Phys. 69, 1325 (2006). [30] R. H. Dicke, Phys. Rev. 93, 99 (1954), DOI: 10.1103/PhysRev.93.99. [31] M. Gross and S. Haroche, Phys. Rep. 93, 301 (1982).

[32] K. Hepp and E. Lieb, Phys. Rev. A 8, 2517 (1973), DOI: 10.1103/PhysRevA.8.2517. [33] K. Hepp and E. Lieb, Ann. Phys 76, 360 (1973), DOI: 10.1016/00034916(73)90039-0. [34] Y. K. Wang and F. T. Hioe, Phys. Rev. A 7, 831 (1973), DOI: 10.1103/PhysRevA.7.831. [35] K. Rz az ewski, K. W odkiewicz, and W. Zakowicz, Phys. Rev. Lett 35, 432 (1975), DOI: 10.1103/PhysRevLett.35.432. [36] K. Rz az ewski and K. W odkiewicz, Phys. Rev. A 43, 593 (1991). [37] P. R. Eastham and P. B. Littlewood, Solid State Commun. 116, 357 (2000), DOI: 10.1016/S0038-1098(00)00350-1. [38] P. R. Eastham and P. B. Littlewood, Phys. Rev. B 64, 235101 (2001), DOI: 10.1103/PhysRevB.64.235101. [39] E. A. Yuzbashyan, V. B. Kuznetsov, and B. L. Altshuler, Phys. Rev. B 72, 144524 (2005), DOI: 10.1103/PhysRevB.72.144524. [40] J. Dukelsky, S. Pittel, and G. Sierra, Rev. Mod. Phys. 76, 643 (2004), DOI: 10.1103/RevModPhys.76.643. [41] R. A. Barankov and L. S. Levitov, Phys. Rev. Lett. 93, 130403 (2004), DOI: 10.1103/PhysRevLett.93.130403. [42] M. Fleischhauer and M. D. Lukin, Phys. Rev. Lett. 84, 5094 (2000), DOI: 10.1103/PhysRevLett.84.5094. [43] M. Fleischhauer, A. Imamoglu, and J. P. Marangos, Rev. Mod. Phys. 77, 633 (2005), DOI: 10.1103/RevModPhys.77.633. [44] M. D. Lukin, Rev. Mod. Phys. 75, 457 (2003), DOI: 10.1103/RevModPhys.75.457. [45] N. S. Ginsberg, S. R. Garner, and L. V. Hau, Nature 445, 623 (2007), DOI: 10.1038/nature05493.

Você também pode gostar