<<

Fundamentals of

Marco Ornigotti Lukas Maczewsky Alexander Szameit Summer Semester 2016 Friedrich-Schiller Universität - Jena

June 19, 2016 Contents

1 HAMILTONIAN FORMULATION OF THE ELECTROMAGNETIC FIELD 1 1.1 Coulomb Gauge ...... 2 1.2 Field Density ...... 3

2 QUANTIZATION OF THE FREE ELECTROMAGNETIC FIELD 7

3 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD 10 3.1 Number States ...... 10 3.1.1 Properties of Fock States ...... 10 3.2 Coherent states ...... 12 3.2.1 Expectation Value of the electric Field over Coherent States . 13 3.2.2 Action of the Number on Coherent States ...... 14 3.2.3 Probability ...... 15 3.2.4 Summary ...... 15 3.3 Phase Space and Quadrature Operators for the Electromagnetic Field 16 3.3.1 Number states in phase space ...... 17 3.3.2 Coherent states in phase space ...... 17 3.4 ...... 19 3.5 Squeezed States ...... 21

4 Interaction of Matter with Quantized Light -I 24 4.1 Electromagnetic Potentials in Presence of Matter ...... 24 4.2 Introduction of Atoms ...... 25 4.2.1 Charge Density ...... 26 4.2.2 Current Density ...... 26 4.3 Atom-Field Interaction ...... 27 4.3.1 Electric Part ...... 28 4.3.2 Magnetic Part ...... 29 4.3.3 Orders of Magnitude ...... 29 4.4 The Interaction Hamiltonian ...... 30 4.5 ...... 32 4.5.1 Second Quantization of the Atomic Hamiltonian ...... 32 4.5.2 Dipole Operator ...... 34 4.5.3 Jaynes-Cummings Model ...... 35 4.5.4 States Of The Joint System ...... 37 4.6 Schrödinger, Heisenberg and Interaction Picture ...... 37 4.6.1 Schrödinger Picture ...... 37 4.6.2 Heisenberg Picture ...... 38 4.6.3 Interaction Picture ...... 38 5 Interaction of Matter with Quantized Light - II 39 5.1 Photon Emission and Absorption Rates ...... 39 5.1.1 Absorption ...... 39 5.1.2 Emission ...... 40 5.1.3 Rate Equation ...... 41 5.2 Quantum Rabi Oscillations ...... 42 5.2.1 The Semiclassical Rabi Model in a Nutshell ...... 42 5.2.2 The quantum Rabi Model ...... 44 5.2.3 Interaction of an Atom with a ...... 46 5.3 Wigner-Weißkopf-Theory [5] ...... 51

6 of Beam Splitters 58 6.1 Classical theory ...... 58 6.2 Quantum Theory ...... 61 6.3 Classical VS Quantum Beam Splitter ...... 65 6.3.1 Classical BS ...... 65 6.3.2 Quantum BS ...... 65 6.4 Effects of Vacuum in BS: Coherent State Input ...... 67 6.5 Effects of Vacuum in BS: Single Photon Input ...... 68 6.6 Hong-Ou-Mandel Effect [9] ...... 68

7 Quantum Theory of Photodetection 70 7.1 Direct Detection ...... 70 7.2 Homodyne Detection ...... 73

8 Degree of Coherence and Quantum Optics 76 8.1 First Order Degree of Coherence ...... 76 8.2 Quantum Theory of g(1) (τ) ...... 79 8.3 Second Order Degree of Coherence ...... 80 8.4 Quantum Theory of g(2) (τ) ...... 82 8.4.1 Number States ...... 82 8.4.2 Coherent States ...... 83 8.4.3 Chaotic Light ...... 83 8.4.4 Squeezed Vacuum ...... 83 8.5 Comparison and Nonclassical Light ...... 83

9 Elements of Quantum Nonlinear Optics 85 9.1 Nonlinear Wave Equation ...... 85 9.2 Quantization of the Electromagnetic Field in Dielectric Media . . . . 87 9.3 Second Harmonic Generation ...... 91 9.4 Parametric Down Conversion ...... 96 10 Appendix A: The Harmonic Oscillator 98

11 Appendix B: The Quantization of a Paraxial Electromagnetic Field 99 11.1 From the Vector Potential to the Paraxial Operator ...... 99 11.2 The Paraxial Dispersion Relation ...... 101 11.3 Back to Continuous Mode Operators ...... 103 11.4 The Maxwell-Paraxial Modes ...... 104 11.5 The Paraxial States of the Electromagnetic Field ...... 105 11.6 An Example of Paraxial Mode Operators ...... 105

12 Bibliography 106 HAMILTONIAN FORMULATION OF THE ELECTROMAGNETIC FIELD

1 HAMILTONIAN FORMULATION OF THE ELEC- TROMAGNETIC FIELD

First of all let us write down the Maxwell equations in vacuum:

∇ · E (r, t) = 0, (1.1a)

∂B (r, t) ∇ × E (r, t) = − , (1.1b) ∂t ∇ · B (r, t) = 0, (1.1c) 1 ∂E (r, t) ∇ × B (r, t) = . (1.1d) c2 ∂t Eq.(1.1c) is also satisfied if B (r, t) is derived from a potential A (r, t), i.e.,

B (r, t) = ∇ × A (r, t) . (1.2)

Substituting Eq.(1.2) into Eq.(1.1c) leads to

∇ · (∇ × A) = 0. (1.3)

And substituting Eq.(1.2) into Eq.(1.1b) we get

∂ ∇ × E + (∇ × A) = 0. (1.4) ∂t

In Eq.(1.4) we can factor out the curl operator to obtain

 ∂  ∇ × E + A = 0. (1.5) ∂t

Furthermore, if we define ∂A E + = −∇φ, (1.6) ∂t we have that  ∂  ∇ × E + A = −∇ × ∇φ = 0. (1.7) ∂t So now we have the electric and magnetic field, given as a function of the vector potential A and the scalar potential φ. The evolution equations for the potentials we can get by using eq.(1.1a) and eq.(1.1d). Equation (1.1a) leads to

 ∂  ∇ · − A − ∇φ = 0 ∂t ∂ ⇒ (∇ · A) + ∇2φ = 0 (1.8) ∂t

1 HAMILTONIAN FORMULATION OF THE ELECTROMAGNETIC FIELD and from eq.(1.1d), using the vector identity ∇ × ∇ × A = ∇ (∇ · A) − ∇2A, we get:

1 ∂  ∂A  ∇ (∇ · A) − ∇2A = − − ∇φ c2 ∂t ∂t 1 ∂2A 1 ∂∇φ ⇒ ∇ (∇ · A) − ∇2A = − − (1.9) c2 ∂t2 c2 ∂t

Note: Eq.(1.8) and eq.(1.9) Maxwell’s equations for the field potentials. However A and φ have no physical meaning. In order to show what is meant by this statement let us have a look at the following transformations

A = A0 − ∇Λ, (1.10a)

∂Λ φ = φ0 + , (1.10b) ∂t where Λ is an arbitrary (but well behaved) scalar field, which is called gauge field. We now calculate the fields E0 and B0 generated by A0 and φ0. For the magnetic field B0 we obtain following result:

B0 = ∇ × A0  = ∇ × A + ∇ ×(∇Λ) = ∇ × A = B, (1.11) while for the electric field E0 we have: ∂A0 E0 = − − ∇φ0 ∂t   ∂A ∂∇Λ ∂∇Λ = − − − ∇φ+ ∂t  ∂t  ∂t ∂A = − − ∇φ = E. (1.12) ∂t

So we conclude: • Different field potentials generate the same E and B (gauge transformations)

• While E and B are manifestly gauge invariant, the field potentials are not. It is therefore important when dealing with field potentials to fix the gauge by fixing the choice of Λ.

1.1 Coulomb Gauge

Assume to consider the condition

∇ · A = 0 (1.13)

2 HAMILTONIAN FORMULATION OF THE ELECTROMAGNETIC FIELD to be valid. Therefore using this gauge we obtain for the scalar potential from Eq.(1.8) that ∇2φ = 0, (1.14) from which it follows that φ = 0. This result is in accordance with the fact that the scalar potential is associated to the charge distribution (and, ultimately, to the electrostatic field) and therefore since in vacuum there are no charges, the scalar potential is zero. Substituting this result into Eq.(1.9) then gives

2 ¨  2 1 ∂ A 1 ∂¨∇φ ∇(∇ · A) − ∇ A = − −¨¨  c2 ∂t2 ¨ c2 ∂t 1 ∂2A → ∇2A − = 0. (1.15) c2 ∂t2

From both, Eq.(1.14) and Eq.(1.15) we can see that in the Coulomb (radiation) gauge in vacuum, E and B are completely determined by the vector potential solely, i.e., ∂A E = − (1.16a) ∂t and B = ∇ × A. (1.16b) What about Λ? Once the gauge condition is fixed, it must be respected by all (A, φ). Therefore:

 ∇ · A = ∇ · A0 − ∇ · (∇Λ) = 0 ⇒ ∇2Λ = 0 (1.17) and ∂Λ ∂Λ φ = φ0 + = 0 ⇒ = 0 (1.18)  ∂t ∂t This means, that Λ is a harmonic function, i.e., it is a solution of the wave equation

1 ∂2 ∇2Λ − Λ = 0. (1.19) c2 ∂t2

1.2 Field Energy Density

Let us consider the electromagnetic field Hamiltonian: Z   1 2 1 2 H = dV ε0|E (r, t) | + |B (r, t) | . (1.20) 2 µ0

Our goal is now to rewrite Eq. (1.20) in terms of the vector potential. To do that, however, we make sure to follow the following prescriptions, which will turn out to be useful for quantization. 3 R P a) Finite Volume V = L =⇒ dV → k This is done because by introducing a finite volume we can imagine to write the

3 HAMILTONIAN FORMULATION OF THE ELECTROMAGNETIC FIELD

field as a superposition of cavity modes, that constitute a finite set of discrete vectors; b) Allowed Modes 2π k = (nxˆx + ny ˆy + nzˆz) (1.21) L These are the k-vectors of the modes sustained by the fictitious cavity intruduced in a); c) Polarization The electromagnetic field can be characterized by 2 orthogonal states of polar- ization in the transverse plane (e.g., {|Hi , |V i} , {|Li , |Ri} , {|TEi , |TMi}...).

For each mode we then have two possible polarization states, namely eˆkλ, with

λ = 1, 2 and ˆekλ · ˆek0λ0 = δλλ0 . The vector potential then can be written in the following way: 2 X X A (r, t) = ˆekλAkλ (r, t) (1.22) k λ=1 Moreover, A must obey Coulomb gauge condition, i.e., ∇ · A = 0. This leads to

k · ˆekλ = 0, (1.23) which states that the vector potential (and therefore E and B) does not possess components along the propagation direction k. In other words, A is a purely trans- verse field. Another condition that must be fulfilled, is that A must obey the wave equation 1 ∂2A ∇2A − = 0. (1.24) c2 ∂t2 Inserting Eq.(1.22) leads to

2  2  X X 2 1 ∂ Akλ (r, t) ˆekλ ∇ Akλ (r, t) − = 0. (1.25) c2 ∂t2 k λ=1

Let us now assume that

ik·r ∗ −ik·r Akλ (r, t) = Akλ (t) e + Akλ (t) e , (1.26) thus we have:

2  ik·r ∗ −ik·r ∇ Akλ (r, t) = −k · k Akλ (t) e + Akλ (t) e . (1.27)

Defining ωk = ck we have from Eq.(1.25)

2 ∂ Akλ (t) 2 + ω Akλ (t) = 0 (1.28) ∂t2 k

4 HAMILTONIAN FORMULATION OF THE ELECTROMAGNETIC FIELD and the same expression for the complex conjugate. The solution to Eq.(1.28) is then −iωkt Akλ (t) = Akλe , (1.29) and for the complex conjugate

∗ ∗ iωkt Akλ (t) = Akλe . (1.30)

Using this result and Eqs. (1.16), the electric and magnetic fields can be then written as follows:

2 X X  −iωkt+ik·r ∗ iωkt−ik·r E (r, t) = iωkˆekλ Akλe − Akλe , (1.31) k λ=1 and 2 X X  −iωkt+ik·r ∗ iωkt−ik·r B (r, t) = i (k × ˆekλ) Akλe − Akλe . (1.32) k λ=1 Now we can calculate the electromagnetic field Hamiltonian H which is given by Eq. (1.20) and wich we will write in following form: Z   1 2 1 2 H = dV ε0|E (r, t) | + |B (r, t) | = HE + HB, (1.33) 2 µ0 where HE and HB are the Hamiltonians of the electric and magnetic field, respec- tively. In the following we will calculate explicitly the electric field Hamiltonian, using the vector Potential A. We obtain:

ε Z H = 0 dV |E (r, t) |2 E 2 ε Z = 0 dV E∗ (r, t) · E (r, t) 2 2 2 ε Z h 0 0 i 0 X X X X −iωk0 t+ik ·r ∗ iωk0 t−ik ·r = dV (−iωk)(iωk0 ) Ak0λ0 e − A 0 0 e 2 k λ k0 λ0=1 k λ=1  ∗ iωkt−ik·r −iωkt+ik·r × Akλe − Akλe ˆek0λ0 · ˆekλ (1.34)

NOTE: The only quantities that remain inside of the integral are the exponentials. By recalling that Z ±i(k−k0)·r dV e = V δk,k0 , (1.35a) and Z ±i(k+k0)·r dV e = V δk,−k0 , (1.35b)

5 HAMILTONIAN FORMULATION OF THE ELECTROMAGNETIC FIELD we obtain, substituting the above result into Eq. (1.20),

V ε ωkωk0 X X 0 ∗ i(ωk0 −ωk)t HE = − A 0 0 Akλe δk,k0 ˆek0λ0 · ˆekλ + 2 k λ kk0 λλ0 | {z } δλλ0

∗ ∗ i(ωk0 +ωk)t −i(ωk0 +ωk)t + Ak0λ0 Akλe δk,−k0ˆek0λ0 · ˆekλ + Ak0λ0 Akλe δk,−k0ˆek0λ0 · ˆekλ

∗ −i(ωk0 −ωk)t + Ak0λ0 Akλe δk,k0 ˆek0λ0 · ˆekλ | {z } δλλ0 V ε ω2 X X 2 2 X X 0 k  ∗ i2ωkt = − ε V ω |Akλ| − A A−k,λ0ˆe−k,λ0 · ˆekλe + 0 k 2 kλ k λ k λ −i2ωkt + AkλA−k,λ0ˆe−k,λ0 · ˆekλe . (1.36)

The second term in Eq. (1.36) compensates a similar term in the magnetic field Hamiltonian, which is given by:

1 Z h 0 0 i X X −iωk0 t−ik ·r ∗ iωk0 t−ik ·r HB = − dV Ak0λ0 e + Ak0λ0 e 2µ0 kk0 λλ0  ∗ iωkt−ik·r −iωkt+ik·r 0 × Akλe − Akλe (k × ˆek0λ0 )(k × ˆekλ) (1.37)

Analogously to the electric field Hamiltonian, the integration gives factors δk,k0 and

δk,−k0 . With this we obtain:

2 δk,k0 → (k × ˆekλ0 )(k × ˆekλ) = k δλλ0 (1.38a)

2 δk,−k0 → (k × ˆekλ)(k × ˆe−k,λ0 ) = −k ˆekλ · ˆe−k,λ0 (1.38b)

Combining HE and HB we obtain the full electromagnetic field Hamiltonian

2 X X 2 ∗ ∗ H = ε0V ωk (AkλAkλ + AkλAkλ) . (1.39) k λ=1

For the calculation we have used that

2  2  2 k 2 k ε0ωkˆe−kλ0 · ˆekλ − ˆekλ · ˆe−kλ0 = ε0ωk − ˆekλ · ˆe−kλ0 = µ0 µ0  2  2 2 2 1 2 2 2 2 2 k ε0µ0c k − k c2 c k − k = ε0c k − = = = 0, (1.40) µ0 µ0 µ0 where ωk = ck. NOTE: Although for the classic EM field AA∗ = A∗A we still write this as AA∗ + A∗A instead of 2|A|2 in order to obtain a direct connection with the quantized form, which we will see later, where A and A∗ will be promoted to operators.

6 QUANTIZATION OF THE FREE ELECTROMAGNETIC FIELD

2 QUANTIZATION OF THE FREE ELECTRO- MAGNETIC FIELD

We consider the EM field Hamiltonian as a collection of cavity modes, as given by Eq. (1.39), i.e., 2 X X 2 ∗ ∗ H = ε0V ωk (AkλAkλ + AkλAkλ) . k λ=1 We now introduce the creation and annihilation operators and express the Hamil- tonian in terms of these. If we compare the above expression of the field Hamiltoinan of a harmonic oscillator (see Eq. (10.3)), we see that we can obtain the quantized version of the field Hamiltonian if we promote the mode coefficients A and A∗ to operators in such a way that A ' a† and A∗ ' a, namely r ~ † Akλ → aˆkλ, (2.1a) 2ε0V ωk r ∗ ~ Akλ → aˆkλ. (2.1b) 2ε0V ωk Substituting this into Eq. (1.39) brings us the following form for the quantized field Hamiltonian:   ˆ X ~ω † † H = aˆ aˆkλ +a ˆkλaˆ . (2.2) 2 kλ kλ k,λ The following commutation rules are understood

h † i aˆkλ, aˆk0λ0 = δkk0 δλλ0 , (2.3a)

h † † i [ˆakλ, aˆk0λ0 ] = 0 = aˆkλ, aˆk0λ0 . (2.3b)

If we make use of the above commutation relation, Eq. (2.2) can be written in the following, more inspiring, form:

X   Hˆ = ~ω aˆ† aˆ +a ˆ† aˆ + 1 2 k,λ kλ kλ kλ kλ k,λ 2   X X † 1 = ωk,λ aˆ aˆkλ + . (2.4) ~ kλ 2 k λ=1

This result can be interpreted in a very intuitive way: thanks to the above equation, the total energy of the electromagnetic field can be interpreted as given by the sum of the contribution of several uncoupled harmonic oscillators, each one characterized by its own frequency ωk,λ. Since the field is composed by several uncoupled oscillators, it is reasonable to assume that for each oscillator (i.e., each field mode), there should be a state |nkλi associated. This allows us to write the global state of the

7 QUANTIZATION OF THE FREE ELECTROMAGNETIC FIELD electromagnetic field as a direct product of all such states, i.e.,

|nk1,1, nk1,2, nk2,1, nk2,2,...i = |nk1,1i |nk1,2i |nk2,1i ... ≡ |nkλi (2.5)

These states are known in literature as Fock states, and constitute the energy eigen- states of the electromagnetic field. To understand that, let us analyze how the Hamilton operator (2.4) acts on such states, namely

2 X X  1 Hˆ |n i = ω aˆ† aˆ + |n i kλ ~ k kλ kλ 2 kλ k λ=1 2 X X  √ 1  = ω aˆ† n |n − 1i + |n i ~ k kλ kλ kλ 2 kλ k λ=1 2 X X  1 = ωk nkλ + |nkλi , (2.6) ~ 2 k λ=1 where the following relations have been used: √ aˆ |ni = n |n − 1i (2.7a) √ aˆ† |ni = n |n − 1i (2.7b)

NOTE: Combining the above relations we can introduce the photon number oper- ˆ ator Nkλ as follows ˆ † Nkλ =a ˆkλaˆkλ. (2.8)

The matrix elements of the Hamilton operator of the free field are then obtained by multiplying eq. (2.6) from the left side with hnkλ| we obtain:

2   ˆ X X 1 hnkλ| H |nkλi = ω nkλ + (2.9) ~ 2 k λ=1

This is the total energy of the field expressed as the sum of quanta distributed among all the possible field modes. We used hnkλ| nk0λ0 i = δkk0 δλλ0 . IMPORTANT NOTE:

• nkλ is the number of quanta (field excitations - ) in the mode (k, λ);

• Classical Electromagnetism: NO FIELD = NO ENERGY;

• Quantum Optics: For the vacuum state nkλ = 0 we obtain:

2   2 ˆ X X 1 X X ~ωk H |0i = ωk nkλ + |0i = 6= 0. (2.10) ~ 2 2 k λ=1 k λ=1

8 QUANTIZATION OF THE FREE ELECTROMAGNETIC FIELD

P So the energy of the vacuum state is k,λ ~ωk/2. It must be nonzero because of the Heisenberg principle (compare to particle in a box)

To conclude the quantization procedure, let us now write the vector potential, the electric and the magnetic field operators in terms of aˆ and aˆ†. The expressions of the vector potential, electric and magnetic fields in terms of the vector potential coefficients A and A∗ are given as follows:

2 X X  −iωkt+ik·r ∗ iωkt−ik·r A (r, t) = ˆekλ Akλe + Akλe , (2.11a) k λ=1

2 X X  −iωkt+ik·r ∗ iωkt−ik·r E (r, t) = ˆekλ · iωk Akλe − Akλe , (2.11b) k λ=1

2 X X  −iωkt+ik·r ∗ iωkt−ik·r B (r, t) = i (k × ˆekλ) Akλe − Akλe . (2.11c) k λ=1 q q ~ † ∗ ~ Akλ → aˆ A → aˆkλ By recalling that 2ε0V ωk kλ and kλ 2ε0V ωk , we can therefore write the field operators in the following form:

A (r, t) = A+ (r, t) + A− (r, t) , (2.12a)

E (r, t) = E+ (r, t) + E− (r, t) , (2.12b)

B (r, t) = B+ (r, t) + B− (r, t) . (2.12c)

Where + indicates the positive frequency part and − the negative one. Furthermore:

2 r + X X ~ −iωkt+ik·r A (r, t) = ˆekλ aˆkλe , (2.13a) 2ε0V ωk k λ=1 2 r − X X ~ † iωkt−ik·r A (r, t) = ˆekλ aˆkλe , (2.13b) 2ε0V ωk k λ=1

2 r ω + X X ~ k −iωkt+ik·r E (r, t) = i ˆekλ aˆkλe , (2.14a) 2ε0V k λ=1

2 r ω − X X ~ k † iωkt−ik·r E (r, t) = −i ˆekλ aˆkλe , (2.14b) 2ε0V k λ=1

2 r + X X ~ −iωkt+ik·r B (r, t) = i (k × ˆekλ)a ˆkλe , (2.15a) 2ε0V ωk k λ=1 2 r − X X ~ † iωkt−ik·r B (r, t) = −i (k × ˆekλ)a ˆkλe . (2.15b) 2ε0V ωk k λ=1

9 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD

3 QUANTUM STATES OF THE ELECTROMAG- NETIC FIELD

We now want to study the possible states of the electromagnetic field and their properties. Without loss of generality, we consider only single mode fields, i.e., the electromagnetic field is represented by a single harmonic oscillator corresponding to a single degree of freedom of the field (polarization, frequency, spatial mode, etc.). To this aim, let us fix the polarization of the field to be xˆ and let us assume that the field is characterised by a single k-vector k and that it propagates along the z-direction. . Using these assumptions, the electric field operator defined in Eq. (2.11b) assumes then the following form:

ˆ  i(kz−ωt) † −i(kz−ωt) E(z, t) = iE0 aeˆ − aˆ e , (3.1)

p where E0 = ~ω/2ε0V . Notice, moreover, that for monochromatic plane wave fields |B| = |E|/c and the magnetic field can be therefore be neglected.

3.1 Number States

According to the results of the previous section, the Hamiltonian of a single mode fieldis given by  1 Hˆ = ω nˆ + , (3.2) ~ 2 where nˆ =a ˆ†aˆ is the number operator, which counts the number of excitations in a mode, and therefore its eigenvalues n give the number of excitations (i.e., photons) in the mode. It makes then sense to introduce the Fock (number) states as follows: Def: Fock (number) states are eigenstates of the number operator, such that

nˆ |ni = n |ni . (3.3)

3.1.1 Properties of Fock States

Fock states form a complete set of orthonormal vectors, i.e.,

• Orthogonality:

hn| mi = δnm, (3.4)

• Completeness: ∞ X |ni hn| = 1. (3.5) n=0

The operators aˆ† and aˆ create and annihilate, respectively, one photon in the

10 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD

field i.e., √ aˆ† |ni = n + 1 |n + 1i , (3.6a) √ aˆ |ni = n |n − 1i . (3.6b)

In order to create a state from the vacuum we need to apply this relations recursively:

aˆ†n |ni = √ |0i , (3.7) n! while the vacuum state is defined as

aˆ |0i = 0. (3.8)

The expectation value of the Hamilton operator of the field over the Fock states is given by

 1  H = hn| Hˆ |ni = ω hn| nˆ |ni + hn| ni ~ 2  1  = ω n hn| ni + hn| ni ~ 2  1 = ω n + hn| ni , (3.9) ~ 2 and therefore:  1 H = ω n + . (3.10) ~ 2 The expectation value of the electric field operator (3.1) is instead given by

ˆ  i(kz−ωt) † −i(kz−ωt) hn| E(z, t) |ni = iE0 hn| aˆ |ni e − hn| aˆ |ni e √ h√ i(kz−ωt) −i(kz−ωt)i = iE0 ne hn| n − 1i − n + 1 hn| n + 1i e = 0. (3.11)

The electric field is not well defined for number states!! D E2 However Eˆ is nonzero since it is connected with the field energy

 1 hn| Eˆ 2 |ni = 2E 2 n + . (3.12) 0 2

We can therefore calculate the field variance:

D E2 D E D E2  1 ∆E(ˆ z, t) = Eˆ 2 (z, t) − E(ˆ z, t) = 2E 2 n + . (3.13) 0 2

The above equation has an interesting consequence: letting n = 0, i.e. no photons

11 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD in the field, the variance D E2 ∆E(ˆ z, t) 6= 0.

This means that the vacuum state |0i shows field fluctuations (quantum noise). Moreover the number operator and the electric field operator do not commute, i.e., h i ˆ  i(kz−ωt) † −i(kz−ωt) n,ˆ E(z, 0) = iE0 −aeˆ +a ˆ e . (3.14)

Summarizing: • Number states are quantum states of well definite photon number. As a con- sequence, the electric field is not well defined for these states;

• vacuum field fluctuations.

3.2 Coherent states

As we have seen number states do not have a well defined E and therefore are not fitted to describe a classical field. So how can we find then a classical expression for a of the e.m. field?

Considerations: 1. Large numbers of photons in the field |Ni ,N  1?

FALSE: we just saw, that no matter how big N is, a number state has an ill-defined E and therefore the cannot come from |Ni

2. Since Eˆ ∼ aˆ − aˆ† a possible trial solution could be |ψi ∼ |ni + |n + 1i?

Good: we generalize this concept and therefore obtain the definition:

Def. Coherent States: Coherent states are eigenstates of the annihilation opera- tor and therefore obtain following eigenvalue equation:

aˆ |αi = α |αi , α ∈ C (3.15) (hα| aˆ† = hα| α∗).

Let us write the coherent states in number state basis:

∞ X |αi = cn |ni , (3.16) n=0 from this two equations we can calculate the expansion coefficients:

∞ ∞ ∞ X X √ X aˆ |αi = cnaˆ |ni = ncn |n − 1i = α cn |ni . (3.17) n=0 n=1 n=0

12 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD

Now, in order to let both sums on the r.h.s. of eq.(3.17) start from the same number, P∞ P∞ i.e. n = 1, we need to change the expression n=0 cn |ni to n=1 cn−1 |n − 1i. This leads to √ ncn = αcn−1. (3.18)

Applying successively the annihilation operator on equation (3.17) it follows that

αn cn = c0 √ , (3.19) n! where c0 is defined by normalization, i.e., by imposing hα| αi = 1. This leads to

∞ ∞ ∗ n m 2 X X (α ) α hα| αi = |c0| √ hn| mi n=0 m=0 n!m! | {z } δnm ∞ 2n 2 X |α| 2 |α|2 = |c0| = |c0| e . (3.20) n! n=0

Therefore:

" 2 # ∞ " 2 # ∞ n n |α| X αn |α| X α aˆ† |αi = exp − √ |ni = exp − √ |0i . (3.21) 2 2 n=0 n! n=0 n!

3.2.1 Expectation Value of the electric Field over Coherent States

Let us now calculate the expectation value of the electric field in coherent state basis, using

ˆ i(kz−ωt) † −i(kz−ωt) hα|E(z, t)|αi = iE0hα|aeˆ − aˆ e |αi (3.22)  i(kz−ωt) ∗ −i(kz−ωt) = iE0 αe − α e (3.23)  i(kz−ωt+φ) −i(kz−ωt+φ) = iE0|α| e − e (3.24)

= −2E0|α| sin(kz − ωt + φ), (3.25) where in passing from the second to the third line we have used the fact that α = |α| exp (iφ). As can be seen, moreover, the expectation value of the electric field operator over the coherent state |αi gives as result the classical electric field. Now let us calculate the variance of the electric field operator, i.e.,

 2 hα|∆Eˆ2(z, t)|αi = hα|Eˆ2(z, t)|αi − hα|Eˆ(z, t)|αi

2 ~ω = E0 = . (3.26) 2ε0V

A coherent state is a nearly classical state beause it reproduces correctly the classical field although it still contains field fluctuations (that are the same of the vacuum field).Note, moreover, coherent states are also defined for the harmonic oscillator as

13 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD states of minimal uncertainty.

3.2.2 Action of the Number Operator on Coherent States

Let us now investigate how the number operator acts on coherent states. Therefore we will calculate the expectation value of the number operator in coherent state basis, as follows

2 2 hα| nˆ |αi = hα| aˆ†aˆ |αi = |α| hα| αi = |α| ≡ n.¯ (3.27)

This means that coherent states do not possess a well defined number of photons, as we will see now. 2 In fact, compared to the number states where the variance h∆ˆni = 0, the variance of the photon number doesn’t vanish for coherent states. In order to convince ourselves, that this is the case, let us calculate

h∆ˆni2 = nˆ2 − hnˆi2 = hα| aˆ†aˆaˆ†aˆ |αi − hα| aˆ†aˆ |αi . (3.28)

Here we are going to use the commutation rule

  a,ˆ aˆ† =a ˆaˆ† − aˆ†aˆ = 1, (3.29) which leads to

hα| aˆ†aˆaˆ†aˆ |αi = hα| aˆ† 1 +a ˆ†aˆ aˆ |αi = hα| aˆ†aˆ |αi + hα| aˆ†aˆ†aˆaˆ |αi 2 4 = |α| + |α| . (3.30)

Combining our results and plugging them back into eq.(3.28), we obtain the variance of the photon number operator for the case of coherent states as follows:

2 2 4 4 2 h∆ˆni = |α| + |α| − |α| = |α| , (3.31) hence 2 2 hnˆi = h∆ˆni = |α| (3.32)

14 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD

3.2.3 Photon Probability

The probability of finding n photons in a coherent state |αi can be calculated as follows:

2 " 2 # ∞ m 2 |α| X α P = |hn| αi| = exp − √ hn| mi n 2 m=0 m! n 2 n  2 |α|  2 n¯ = exp − |α| = exp −n¯ . (3.33) n! n!

This satisfies the Poissonian statistic (mean value = variance):

n n¯ 2 Pn = exp[−n¯ ] (3.34) n! and moreover: ∆n 1 = √ (3.35) n¯ n¯

3.2.4 Summary

Number States: Number states are described by |ni. They are eigenstates of the number operator nˆ =a ˆ†aˆ and fulfil the equation

nˆ |ni = n |ni .

The mean value of the electric field in number state basis is:

hn| Eˆ |ni = 0

but the variance is:

D E2  1 ∆Eˆ ∼ n + (quantum noise) 2

Coherent States: Coherent states are the most classical among the quantum states. Tey are described by |αi. They are eigenstates of the creation operator aˆ and fulfil the equation aˆ |αi = α |αi

We can transform them to number state basis by:

" 2 # ∞ |α| X αn |αi = exp − √ |ni 2 n=0 n!

15 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD

The mean value of the electric field in coherent state basis is:

DˆE E = −2E0 |α| sin (kz − ωt + φ)

and the variance: r D ˆE ~ω ∆E = E0 ≡ 2ε0V

3.3 Phase Space and Quadrature Operators for the Electro- magnetic Field

Classical mechanics: H (q, p) → each point in phase space represents a configuration of the system (see fig.3.1)

Figure 3.1: Representation of the system configuration in phase space

We can introduce this also in quantum optics. In order to do this, we first remember the harmonic oscillator and adapt the same rules to our operator:

1 † ˆ † ˆx ∼ aˆ +a ˆ → X1 = aˆ +a ˆ (3.36a) 2 1 † ˆ † ˆp ∼ aˆ − aˆ → X2 = aˆ − aˆ (3.36b) 2i Now we can build a 2D phase space associated to the single mode quantum field ˆ ˆ using X1 and X2, with hˆ ˆ i i X1, X2 = . (3.37) 2

16 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD

3.3.1 Number states in phase space

Expectation value: ˆ ˆ hn| X1 |ni = 0 = hn| X2 |ni (3.38) but variance: 1 hn| ∆Xˆ 2 |ni = (2n + 1) = hn| ∆Xˆ 2 |ni (3.39) 1 4 2 Proof:

D E D E ZD E2 ˆ 2 ˆ 2 ˆZ ∆X1 = X1 − X1 ZZ 1 = hn| aˆ +a ˆ† aˆ +a ˆ† |ni 4 1 ¨¨ = hn|aˆaˆ +a ˆaˆ† +a ˆ†aˆ + ¨aˆ†aˆ† |ni 4 1 h √ √ i = hn| aˆ n + 1 |n + 1i + hn| aˆ† n |n − 1i 4 1 1 = [(n + 1) hn| ni + n hn| ni] = (2n + 1) (3.40) 4 4

hˆ ˆ i Since the variances are non-zero and X1, X2 6= 0 we can apply an indetermination relation 1 ∆Xˆ 2∆Xˆ 2 ≥ (3.41) 1 2 16 ˆ → ∆X1 6= 0, even if n = 0 (v hacuum fluctuations). Moreover:

D E D E 1 1 1 ∆Xˆ 2 + ∆Xˆ 2 = (2n + 1) = n + = hn| ˆn+ |ni , (3.42) 1 2 2 2 2 the vacuum fluctuations are automatically taken into account.

Phase Space Representation(fig.3.2) p • circle of radius n + 1/2 (radius is well defined)

• phase completely undefined (circle)

• vacuum state = circle with radius p1/2 Figure 3.2: Representation of number states in phase space

3.3.2 Coherent states in phase space

For coherent states we have a peculiar thing:

ˆ 1 ∗ hα| X1 |αi = (α + α ) = Re {α} (3.43a) 2

17 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD

ˆ 1 ∗ hα| X2 |αi = (α − α ) = Im {α} (3.43b) 2i This means that the complex α-plane and the phase space for a coherent state are isomorph with each other. Therefore it is useful to use the complex polar represen- tation for α, i.e.: α = |α| eiϑ. (3.44)

The phase space representation of a coherent state is illustrated in Fig. 3.3

Remember:

hα| ˆn |αi = |α|2

there is no well defined number of pho- tons for a coherent state. Furthermore there is also a phase uncertainty (be- cause of Heisenberg uncertainty princi- ple)

Figure 3.3: Representation of co- herent states in phase space

Moreover:

D E  2 D E ˆ 2 ˆ 2 ˆ 1 ˆ 2 ∆X = hα| X |αi − hα| X1 |αi = = ∆X (3.45) 1 1 4 2 the variance of the quadrature operator is the same as the quantum vacuum. Espe- cially: ˆ ˆ 1 ∆X1∆X2 = (3.46) 4 and therefore coherent states are states of minimum uncertainty. Note: Looking at the figure, one can imagine (also motivated by the fact, that coherent states have the same ∆ˆxk as the vacuum state) that coherent states can be viewed as suitably displaced vacuum states.

18 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD

3.4 Displacement Operator

Figure 3.4: Schematic illustration of the creation of a coherent state by displacing the vacuum state

How is this possible? Let’s assume that there is an operator Dˆ (α) such that:

Dˆ (α) |0i = |αi . (3.47)

To obtain an expression for Dˆ (α) let us note that With: We

" 2 # ∞ ∞ n |α| X αn X aˆ† |αi = exp − √ |ni |ni = √ |0i (3.49) 2 n! n=0 n! n=0 " 2 # ∞ n |α| X αaˆ† = exp − |0i 2 n! n=0 " 2 # |α|   = exp − exp αaˆ† |0i(3.48) 2 are tempted to say that eq. (3.48) describes the displacement operator. In order to understand why this is the case let us consider the Baker-Campbell-Hausdorff- Formula: A+ˆ Bˆ 1 [Aˆ ,Bˆ] Bˆ Aˆ e = e 2 e e (3.50)

19 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD let us choose: Aˆ = αaˆ† and Bˆ = α∗a,ˆ then we obtain: h i Aˆ, Bˆ = αaˆ†, α∗aˆ = |α|2 aˆ†, aˆ = |α|2

∗ However, we are still missing a factor eα aˆ. But note that:

∞ m ∗ X (−αaˆ) e−α aˆ |ni = |0i m! m=0 ∞ m X (−α) m = (ˆa) |0i . (3.51) m! | {z } m=0 =0

If we therefore define † ∗ Dˆ (α) = eαaˆ −α aˆ (3.52)

∗ as the displacement operator, the factor eα aˆ has no effect but contributes to make the displacement operator Dˆ (α) more symmetric. Note: The displacement operator is unitary:

• Dˆ † (α) = Dˆ (−α)

• Dˆ (α) Dˆ † (α) = Dˆ † (α)Dˆ (α) = 1

∗ Note:You can prove, that without eα aˆ the displacement operator is not unitary! More on Dˆ (α): It is useful and constructive to calculate the following quantity

hα| aˆ |αi = h0| Dˆ † (α)a ˆDˆ (α) |0i

What is then Dˆ † (α)a ˆDˆ (α)? It is hard to calculate this by hand, but we will use the: Decomposition Theorem for Exponential Operators

ˆ ˆ h i 1 h h ii 1 h h h iii eABˆe−A = Bˆ + Aˆ, Bˆ + Aˆ, Aˆ, Bˆ + Aˆ, Aˆ, Aˆ, Bˆ + ... (3.53) 2! 3! For our purpose we chose: Aˆ = −αaˆ† + α∗aˆ and Bˆ =a ˆ. So we obtain:

Dˆ † (α)a ˆDˆ (α) =a ˆ + −αaˆ† + α∗a,ˆ aˆ     =a ˆ + −α aˆ†, aˆ +α∗ [ˆa, aˆ]  | {z } |{z} =−1 =0 =a ˆ + α (3.54)

20 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD

Summarizing:

Dˆ † (α)a ˆDˆ (α) =a ˆ + α (3.55) Dˆ † (α)a ˆ†Dˆ (α) =a ˆ† + α∗ (3.56)

In this way we can evaluate each expectation value as expectation value done on the vacuum state. And so we obtain, that:

hα| f a,ˆ aˆ† |αi = h0| Dˆ † (α) f a,ˆ aˆ† Dˆ (α) |0i ∗ = h0| f aˆ + α, aˆ† + α |0i (3.57)

Coherent States: Classical field (α) plus quantum fluctuations. Squeezed States: States of the electromagnetic field, for which it is possible for a quadrature to beat the Heisenberg limit. Of course the price to pay is that the uncertainty on the other quadrature gets bigger.

3.5 Squeezed States

So far we have considered number states (well defined n but ill-defined Eˆ and Bˆ) and coherent states (almost classical, minimal uncertainty) and their phase space representations (see Fig. 3.3). They fulfil the uncertainty relation:

ˆ ˆ 1 ∆X1∆X2 ≥ . (3.58) 4

Is this the limit for manipulating states or is there another possibility to play around beating the fundamental Heisenberg limit?

Figure 3.5: Comparison of a squeezed state with the vacuum state in phase space

ˆ ˆ As before ∆X1∆X2 = const but in the certain case the uncertainty on x2 is much less than in the cases before. How can we define squeezed states and how are they generated?

21 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD

ˆ 2 ˆ 2 1 We seek for a state |ξi with ∆X1∆X2 ≥ 16 , but:

1 ∆Xˆ 2 = γ (3.59) 1 4 1 ∆Xˆ 2 = , (3.60) 2 4γ where γ ≤ 1 so that it is a reduction of the variance (uncertainty) for x1 but at the same time it is an amplification for the variance of x2 (so that the product is still constant). We employ:

1. Since we learned, that every (coherent) state can be displaced from the vac- uum, let us consider the vacuum itself and squeeze it (Squeezed Vacuum, actually this is, what we generate in the lab).

2. Let us assume to be able to define an operator S(ˆ ξ) similar to the displacement operator Dˆ (α), namely

h 2i 1 ξ∗aˆ2−ξ(aˆ†) S(ˆ ξ) = e 2 . (3.61)

Eq.(3.61) is the squeezing operator, where ξ ∈ C and ξ = reiϑ.

Figure 3.6: Schematic illustration of squeezed vacuum

ˆ 1 ˆ 1 ˆ ˆ Note: For the squeezed state in fig.3.6 we have ∆X1 < 2 , ∆X2 > 2 but ∆X1∆X2 = 1 4 (For more information on dynamics in phase space see Ref. [4]) 2 2 Note: S(ˆ ξ) contains aˆ and aˆ† , i.e. the squeezing operator is a nonlinear op- erator. If we want to generate squeezed states, therefore, we must do it through a

22 QUANTUM STATES OF THE ELECTROMAGNETIC FIELD nonlinear process (typically SPDC). Furthermore we require S(ˆ ξ) to be unitary:

Sˆ† (ξ) = S(ˆ −ξ) , (3.62) Sˆ† (ξ) S(ˆ ξ) = S(ˆ ξ) Sˆ† (ξ) = 1. (3.63)

Therefore we can generate a squeezed vacuum state by applying the squeezing op- erator on the vacuum state: |ξi = S(ˆ ξ) |0i . (3.64) ˆ The variance of X1 is obtained by

D E D E D E2 ˆ 2 ˆ 2 ˆ ∆X1 = X1 − X1 . (3.65)

ˆ For this we calculate the expectation value of X1:

1 1 † ˆ† † ˆ hξ| ˆx1 |ξi = hξ| aˆ +a ˆ |ξi = h0| S (ξ) aˆ +a ˆ S(ξ) |0i . (3.66) 2 2

As we did for the displacement operator for the coherent states, we look here for:

Sˆ† (ξ)a ˆS(ˆ ξ) =a ˆ cosh r − aˆ†eiϑ sinh r, (3.67) Sˆ† (ξ)a ˆ†S(ˆ ξ) =a ˆ†cosh (r) − aeˆ iϑ sinh r. (3.68)

Therefore we obtain:

ˆ iϑ † hξ| X1 |ξi = cosh r h0| aˆ |0i − e sinh r h0| aˆ |0i + + cosh r h0| aˆ† |0i − eiϑ sinh r h0| aˆ |0i = 0. (3.69)

For ϑ = 0 we obtain at the end: 1 1 hξ| ∆Xˆ 2 |ξi = e−2r → hξ| ∆Xˆ 2 |ξi = e2r, (3.70) 1 4 2 4 r is then called the "squeezing parameter". Final Notes: aˆ |0i = 0, but then also: S(ˆ ξ)a ˆ |0i = 0.

Since S(ˆ ξ) is unitary and therefore Sˆ† (ξ) S(ˆ ξ) = 1 and therefore:

S(ˆ ξ)a ˆ |0i = S(ˆ ξ)a ˆSˆ† (ξ) S(ˆ ξ) |0i = S(ˆ ξ)a ˆSˆ† (ξ) |ξi = 0. (3.71) | {z } =|ξi

This allows us to define the squeezed states as eigenstates of the operator S(ˆ ξ)a ˆSˆ† (ξ)

23 Interaction of Matter with Quantized Light -I with the eigenvalue zero. To Summarize:

Number States: nˆ |ni = n |ni

Coherent States: aˆ |αi = α |αi

Squeezed States: S(ˆ ξ)a ˆSˆ† (ξ) |ξi = 0

4 Interaction of Matter with Quantized Light -I

After studying the free electromagnetic field and introducing its quantum states, the next step is to consider the interaction of the E.M. field with matter. Up to now we considered the E.M. field in free space, therefore the first step is to rewrite the electromagnetic potentials in presence of matter. Afterwards we will introduce atoms to our calculations and then analyse their interaction with the quantized field. As a result of our considerations we will be able to introduce the interaction Hamiltonian of the E.M. field with matter at the end of this section.

4.1 Electromagnetic Potentials in Presence of Matter

Since we are not considering the case of vacuum anymore we need to take into account charge and current densities. Therefore the E.M. potentials in presence of matter are now given as follows

2  2 1 ∂∇φ 1 ∂ A ∇(∇ · A) − ∇ A + + = µ0J (4.1a)  c2 ∂t c2 ∂t2 σ − ∇2φ = (4.1b) ε0 ∇ · A = 0 (Coulomb Gauge) (4.1c)

In general eqs. (4.1a)-(4.1c) are difficult to solve, since we cannot say anymore that φ = 0. In presence of matter the scalar potential φ (r) is given as

Z 0 1 σ (r ) 0 φ (r) = 0 dr . (4.2) ε0 |r − r |

Moreover we know that since ∇A = 0 the vector potential A needs to be transversal, and so the correspondent E field must also be transversal. In order to separate the transversal and longitudinal part of a vector field we use the Helmholtz theorem. The Helmholtz theorem states that it is possible to split any vector field in three dimensions into the sum of an irrotational (curl-free) vector field and a solenoidal (divergence-free) vector field. Applying the H.T. on eq. (4.1a) we obtain for the current J that

J = JL + JT , (4.3)

24 Interaction of Matter with Quantized Light -I with

∇JT = 0 (solenoidal) , (4.4a)

∇ × JL = 0 (irrotational) . (4.4b)

The Coulomb gauge fixes the vector potential to be transverse. So the µ0JL term does not contribute to A. Therefore we obtain after separating the longitudinal and transverse parts: 2 2 1 ∂ A − ∇ A + = µ0JT , (4.5a) c2 ∂t2 1 ∂∇φ = µ0JL, (4.5b) c2 ∂t σ − ∇2φ = . (4.5c) ε0 From the last two equations we obtain the continuity equation

∂σ ∇ · JL = − . (4.6) ∂t

Splitting J in its transversal and longitudinal parts allows us to split the fields in the same way, namely

E = ET + EL, (4.7a)

B = BT + BL, (4.7b) with ∂σ ET = − , BT = ∇ × A, (4.8) ∂t EL = −∇φ, BL = 0. (4.9)

While the longitudinal field is basically the electrostatic field generated by the charge distribution σ, the transverse field is the one that propagates. BL = 0 because of ∇ · B = 0.

4.2 Introduction of Atoms

Let us now specify our analysis to the case of an atom (with Ne−, see Fig. 4.1) as a field source and try to solve equations (4.5).

25 Interaction of Matter with Quantized Light -I

Figure 4.1: Illustration of the choice of a coordinate system in presence of an atom, where rα represent the position vector of the α-th electron with respect to the nucleus

4.2.1 Charge Density

All the electrons and protons are treated like point charges, this is why we will write the charge density σ (r) as

X σ (r) = −e δ (r − rα) + eZδ (r) . (4.10) α

This charge density generates a longitudinal field EL, which is essentially the Coulomb potential of the atom).

4.2.2 Current Density

Each electron orbits the nucleus, this is why we write the total current density J (r) as a sum of the currents produced by each electron

X J (r) = −e r˙ αδ (r − rα) , (4.11) α where the dot stands for the time derivative.

26 Interaction of Matter with Quantized Light -I

4.3 Atom-Field Interaction

It is convenient to treat the interaction of the atom with the electromagnetic field in terms of the polarization P (r) and magnetization M (r), since they are associated with the macroscopic atomic charges and currents. Using the standard results of electromagnetism (see, e.g., [3]), the polarization and magnetization are linked to the charge density and current (respectively) via the following relations:

σ (r) = −∇ · P (r) , (4.12)

J (r) = P˙ (r) + ∇ × M (r) . (4.13)

The polarization P can be thought to be the sum of all the contributions coming from the dipoles composing the medium. In our case, such elementar dipoles are represented by each single electron and the atom’s nucleus. The polarization of such an elementar dipole is then given by

P = qαrαδ (r − rα) (4.14)

In order to simplify our calculations we will use a trick in the choice of expressing the polarization, namely we imagine to cover the distance between the nucleus (situated at r = 0) and the electron at rα with a series of non-point dipoles oriented along

0rα, in such a way that −qα of one dipole is compensated by +qα of the preceding one (see Fig. 4.2).

Figure 4.2: Split polarization vector between nucleus and electron

Therefore, since for a single dipole we have that

 1  p + 2 Pn (r) = (qαrα/n) δ r − rα , (4.15) n summing all the contributions as in Fig. 4.2 and taking the limit n → ∞ gives

27 Interaction of Matter with Quantized Light -I

(4.16) 1 X Z P (r) = −e rα dξ δ (r − ξrα) . (4.16) a 0 Reasoning in an analogous way for M we obtain:

1 X Z M (r) = −e rα × ˙rα dξ ξδ (r − ξrα) . (4.17) a 0

Recalling that potential energy of a permanent electric and a permanent magnetic dipole are given by −d · E and −µ · B, respectively, we will define the electric VE and magnetic VM contribution to the atom’s potential energy as

1 Z Z 3 X VE = − d r P (r) · ET (r) = e dξ rα · ET (ξrα) , (4.18) α 0 and

1 Z Z 3 X VM = − d r M (r) · B (r) = −e dξ (rα × ˙rα) · B (ξrα) . (4.19) α 0

Usually the typical scale of atoms (a few Ångström) is much smaller than the wave- length of light (400 to 700 nm), therefore let us now expand the fields around ξrα = 0 in order to evaluate the various interaction contributions:

4.3.1 Electric Part 1 Z   X 1 2 2 VE = e dξ 1 + ξ (rα · ∇) + ξ (rα · ∇) + ... rα · ET (0) . (4.20) 2! a 0 Here we can perform the ξ-Integration and obtain:     X  1 1 2  VE = e 1 + (rα · ∇) + (rα · ∇) + ... rα · ET. (0) . (4.21) |{z} 2! 3! a dipole | {z } | {z }   quadrupole hexapole 

Equation (4.21) is the multipole expansion of the electric field potential. In partic- ular we notice that P Dipole: −eD = − α erα ⇒ eD · ET (0)

1 P Quadrupole: Q = − 2 α erαrα ⇒ −∇ · Q · ET (0)

28 Interaction of Matter with Quantized Light -I

4.3.2 Magnetic Part

Now we will do the same expansion for the magnetic field.     e X  1 2 3 2  VM = + (rα · ∇) + (rα · ∇) + ... Iα · B (0) (4.22) m 2! 3! 4! a |{z} | {z } | {z }  dipole quadrupole hexapole 

drα th where Iα = mrα × dt is the angular of the α electron. Eq. (4.22) is then a multipole expansion of the magnetic field potential. The explicit expression of the magnetic dipole, e.g., is then given by

1 X e − eDM = − Iα ⇒ eDM · B (0) (4.23) 2 m α

4.3.3 Orders of Magnitude

In order to get an idea of which contributions of the multipole expansion we have to take in account in our calculations, let us consider the following argument. To −10 do that, let us take, approximately, rα ≈ a0 = 10 m (Bohr radius), Iα ≈ ~ = −34 ω 15 10 (quantum of angular momentum) and ∇ET (0) ≈ c ET (0) , ω ≈ 3 · 10 Hz (visible light). If we substitute these numbers in the definitions of electric dipole and quadrupole, as well as magnetic dipole we obtain:

2 4πε0~ • de = eD · ET (0) ≈ me ET (0)

3e~ • qe = −∇ · Q · ET (0) ≈ 16mc ET (0)

e~ e~ E • dm = eDM · B (0) ≈ 2m B (0) ≈ 2mc ET (0) (plane wave: B ∼ c )

If we now compare the different coefficients with each other:

1. q 3e E (0) 2mc 3 e = ~ T · = ∼ 1, dm 16mc e~ET (0) 8 i.e., the electric quadrupole and magnetic dipole terms are approximately of the same order of magnitude

2. 2 qe 3e~ET (0) me e 1 = · 2 ≈ ∼ ≡ α, de 16mc 4πε0~ ET (0) 4πε0~c 137 i.e., the electric dipole is roughly 2 orders of magnitude bigger than the quadrupole (and αn orders of magnitude bigger than the n-th pole). It is therefore also α-times bigger than the magnetic dipole!

Conclusion: It makes sense to consider interactions of an atom with an electro- magnetic field in the dipole approximation, since this is the term that gives the

29 Interaction of Matter with Quantized Light -I biggest contribution. Note: Some cases forbid the transition to the electric dipole (selection rules). In all those cases we need to consider contributions from higher order of the electric multipole expansion, or, if the atom is magnetically active, then take in account the magnetic multipole expansion.

4.4 The Interaction Hamiltonian

Now we are going to discuss, how to construct the dipole-Hamiltonian out of our theory. We are interested in quantized fields, so we need to take the approach of quantizing E through quantizing A. We therefore need to write D · E starting from A.

Hamiltonian of the Atom-Light System

Since we still want to have a Hamiltonian representation, we need to take into account the canonical momentum of the system. Furthermore we still want to be able to use the canonical quantization scheme to quantize the field. In our first attempt let us assume that the Hamiltonian we are looking for has this form:

2 Z Z   (1) X pα 1 3 1 3 2 1 2 H = + d r σ(r)φ(r) + d r ε0|E (r) | + |B (r) | 2m 2 2 µ0 α | {z } | {z } Atomic Coulomb potential | {z } field (4.24) Considering this Hamiltonian first thing we notice is that there is no interaction part. We could add an interaction term by hand but which kind of? Moreover, it can be demonstrated, that pα is not a canonical momentum anymore when the field is there. h ˆ i Hence it is better to use the minimal coupling principle where pα → pˆα + eA(r) . This definition takes automatically into account the atom-field interaction through the vector potential A (r). As a consequence of this consideration our new Hamil- tonian should have following form:

1 X h i2 1 Z 1 Z  1  Hˆ0 = pˆ + eAˆ (r) + d3r σ(r)φ(r) + d3r ε Eˆ 2 (r) + Bˆ 2 (r) . 2m α 2 2 0 T µ α 0 (4.25) Where the first term is the Hamiltonian of a in the presence of an electromagnetic field. The second term is the Coulomb potential of the atom (com- pletely classical, no operators). Finally the third term is the Hamiltonian of the free electromagnetic field.

30 Interaction of Matter with Quantized Light -I

Let us develop the first term of eq. (4.25) to extract the interaction Hamiltonian:

h i2 2  2  1 X ˆ X pˆα e X ˆ e X ˆ 2 pˆα + eA(r) = + A(rα)pˆα+ A (rα). (4.26) 2m 2m m 2m α α α α | {z } Kinetic Energy

Since the first term on the right hand side of this equation describes the kinetic energy of the isolated atom, we obtain the interaction Hamiltonian as:

 2  ˆ0 e X ˆ e X ˆ 2 H = A(rα)pˆα + A (rα). (4.27) I m 2m α α

Note:

• The interaction Hamiltonian is written as a function of A and is therefore not gauge invariant

• We want gauge invariance and will write it therefore as a function of E · D

For this we need to apply the so called Power-Zienau-Wooley-transformation

 i Z  Uˆ = exp d3r pˆ (r) Aˆ (r) . (4.28) ~

The application of Uˆ to the system Hamiltonian is such that

Hˆ = Uˆ †Hˆ0Uˆ. (4.29)

In doing that, however, the state |ψ0i of the system must also be transformed ac- cording to |ψi = Uˆ † |ψ0i . (4.30)

After some calculation we obtain the full Hamiltonian describing the atom-field interaction:

X p2 1 Z 1 Z  1  X Hˆ = α + d3r σ(r)φ(r) + d3r ε Eˆ 2 (r) + Bˆ 2 (r) + e r · Eˆ (0), 2m 2 2 0 T µ α T α 0 α | {z } | {z } | {z } Hˆ F Hˆ A Hˆ ED (4.31) where now the interaction Hamiltonian is in the right dipole form:

ˆ X ˆ HED = e rα · ET (0) = eD · ET (0) . α

31 Interaction of Matter with Quantized Light -I

4.5 Second Quantization

4.5.1 Second Quantization of the Atomic Hamiltonian

We have seen (see eq. (4.31)) that the atomic Hamiltonian is given by:

X p2 1 Z Hˆ = α + d3r σ(r)φ(r). A 2m 2 α

This form, however, is not useful to us, since we would prefer a second quantized approach, i.e., we would like to express the atomic Hamiltonian in terms of "cre- ation" and "annihilation" operators, so that we can unify the formalism with the one used for describing the electromagnetic field. To do so, we consider the atomic energy eigenstates |ii, with eigenvalues ~ωi, so that:

ˆ HA |ii = ~ωi |ii . (4.32)

P Here we will use the completeness of |ii, i.e., 1 = 1 |ii hi| in order to write ! ! ˆ X ˆ X HA = |ii hi| HA |ji hj| i j X ˆ = |ii hi| HA |ji hj| ij X ˆ = |ii HA hi| ji hj| ij X = ~ωiδij |ii hj| ij = ~ωi |ii hi| . (4.33)

The second quantized form of the atomic Hamiltonian is then given by:

ˆ X HA = ~ωi |ii hi| (4.34) i

ˆ In this form, HA is written as a sum of projections onto the energy eigenstates |ii, where each of them has the eigenvalue ~ωi. In order to catch the essential physics behind the atom-field interaction in a simple and intuitive way, we will consider the case of a two-level system, i.e., we assume the atomic system only to possess a ground state, |1i, and an excited state, |2i. The transition frequency is defined as

ω0 = ω2 − ω1. A pictorial representation of such a system is given in Fig. 4.3.

32 Interaction of Matter with Quantized Light -I

Figure 4.3: Schematic sketch of a two level system

The Hamiltonian of a two level system is then given by:

ˆ HA = ~ω1 |1i h1| + ~ω2 |2i h2| . (4.35)

For a two level system it is useful to introduce the following operators:

σˆ− = |1i h2| (4.36a)

† σˆ+ = (ˆσ−) = |2i h1| . (4.36b)

To understand their meaning, let us consider how they act on the atomic states, namely:

σˆ− |1i = |1i h2| 1i = 0, (4.37a)

σˆ− |2i = |1i h2| 2i = |1i , (4.37b)

σˆ+ |1i = |2i h1| 1i = |2i , (4.37c)

σˆ+ |2i = |2i h1| 2i = 0. (4.37d)

Then, σˆ+ brings the atom from |1i to |2i, while σˆ− brings the atom from |2i to |1i. Moreover

σˆ−σˆ+ = |1i h2| 2i h1| = |1i h1| , (4.38a)

σˆ+σˆ− = |2i h1| 1i h2| = |2i h2| , (4.38b) i.e., this two different combinations of the operators project directly into the eigen- states of the system. It is also useful to introduce the population inversion operator,

σˆ3, as:

σˆ3 = |2i h2| − |1i h1| . (4.39)

Then, the set {σˆ+, σˆ−, σˆ3} obeys the usual Pauli algebra:

[ˆσ+, σˆ−] =σ ˆ3, (4.40a)

33 Interaction of Matter with Quantized Light -I

[ˆσ3, σˆ±] = ±2ˆσ±. (4.40b)

The atomic Hamiltonian can be then rewritten in terms of σˆ± as follows:

ˆ HA = ~ω1 |1i h1| + ~ω2 |2i h2| = ~ω1σˆ−σˆ+ + ~ω2σˆ+σˆ−. (4.41)

Convention: Since the energy is usually defined up to a constant, we have the freedom to decide where to put the zero of the energy. We have two choices:

1. E = 0 for the ground state (Fig. 4.3). Therefore we obtain:

ˆ HA = ~ω0σˆ+σˆ−,

with ω0 = ω2.

ω0 2. E = 0 in the middle of the two levels (see fig. 4.4). Here we obtain ω1 = − 2 ω0 and ω2 = 2 .

Figure 4.4: Sketch of two level system with E = 0 between the two states

If not specified explicitly, in the rest of this notes we will use the second convention. The Hamiltonian is therefore given by:

ω ω ω Hˆ = −~ 0 |1i h1| + ~ 0 |2i h2| = ~ 0 σˆ . A 2 2 2 3

4.5.2 Dipole Operator

It is now easy to second quantize the dipole operator. If we assume that the con- sidered two level system has a transition allowed to the electric dipole then the representation of this operator in second quantization must be such that it real- ized the transition |1i → |2i (and vice versa for the complex conjugate). If we call D12 = D21 ∈ R the dipole moment of the atom (assuming it real), we obtain the dipole operator: ˆ D = D12 (ˆσ+ +σ ˆ−) (4.42)

34 Interaction of Matter with Quantized Light -I

4.5.3 Jaynes-Cummings Model

We have now all the elements to build a consistent model of the interaction of a two level atom with a quantized field. If we recall that the (single-mode) electric field operator is given by ˆ ikz † −ikz E (z, 0) = iE0 aeˆ − aˆ e xˆ, (4.43) ˆ (t = 0 → Schrödinger picture, see below) we can then rewrite HED, in the following way: ˆ ˆ ˆ ikz † −ikz HED = eD · E = ieE0 (D12 · x) aeˆ − aˆ e (ˆσ+ +σ ˆ−) . | {z } interaction parameter If we introduce ~g = eE0 (D12 · x) , (4.44) where g describes the strength of the interaction, we obtain, assuming that the atom is locates at position z = 0,

ˆ † †  HED = ~g aˆσˆ+ +a ˆσˆ− +a ˆ σˆ+ +a ˆ σˆ− . (4.45)

What is the meaning of the 4 terms appearing in (4.45)? To understand that, let us consider the diagrammatic representation of the atom-field interaction given by Fig. 4.5, where

Figure 4.5: Interaction of electric field with two level system

the photon is illustrated as a solid line (the arrow shows in which direction the photon is propagating), the dashed line symbolizes the time evolution of the two level system and the atom is represented by a black dot. The initial state of the atom is on the right side of the diagram (at t = −∞), while the final state is on the left side (at t = ∞). We assume for simplicity that the interaction occurs at t = 0. Now we will make sketches, analogous to Fig. 4.5, for all four terms of Eq. (4.45), in order to figure out which terms are contributing to the interaction:

35 Interaction of Matter with Quantized Light -I

Fig. 4.6 :aˆσˆ− The atom is initially in |2i. It absorbs a photon and de- cays to |1i. This does not conserve energy and is therefore excluded! Figure 4.6: Atom in excited state, photon gets absorbed

Fig. 4.7 :aˆσˆ+ The atom is initially in |1i. It absorbs a photon and gets excited to |2i. This process is allowed. Figure 4.7: Atom in ground state, photon gets absorbed

† Fig. 4.8 :aˆ σˆ− The atom is initially in |1i. It emits a photon and gets excited to |2i. This does not conserve energy Figure 4.8: Atom in ground state, and is therefore ex- photon is emitted cluded!

† Fig. 4.9 :aˆ σˆ+ The atom is initially in |2i. It emits a photon and de- cays to |1i. Figure 4.9: Atom in excited state, This process is allowed. photon is emitted

† Then, only aˆσˆ+ and aˆ σˆ+ give contributions to our calculations and the interac- tion Hamiltonian can be written as follows:

ˆ † HED = ~g σˆ+aˆ +σ ˆ−aˆ . (4.46)

36 Interaction of Matter with Quantized Light -I

The full Hamiltonian in the Jaynes-Cummings model is then given by:

1 ˆ † † H = ω0σˆ3 + ωaˆ aˆ + g σˆ+aˆ +σ ˆ−aˆ , (4.47) 2~ ~ ~ together with the commutation relations

  a,ˆ aˆ† = 1, (4.48a)

[ˆσ+, σˆ−] =σ ˆ3, (4.48b)

[ˆσ3, σˆ±] = ±2ˆσ±, (4.48c) and the other commutators are zero.

4.5.4 States Of The Joint System

The form of the Hamiltonian (4.47) suggests that the considered system can be thought as the direct product of two disjoint Hilbert spaces, a 2D

HA belonging to the atom and spanned by {|1i , |2i}, and an infinite- dimensional

Hilbert space describing the e.m. field HE and spanned, e.g. by the Fock states {|ni}. Since these two Hilbert spaces are disjointed, one can define the state of the joint system atom + field simply as the direct product between these two spaces, i.e.,

HJ = HA ⊗ HE. The joint Hilbert space HJ is then spanned by {|n, ii = |ni |ii}.

4.6 Schrödinger, Heisenberg and Interaction Picture

So far we considered all the operators to be time-independent. The question is how to introduce the time dependence in our model. We have three possibilities:

4.6.1 Schrödinger Picture

The Schrödinger picture is typically used in quantum mechanics. Here operators are time-independent ˆ dOS = 0, (4.49) dt whereas states evolve in time according to the Schrödinger equation

∂ΨS (t) ˆ i = HΨS (t) . (4.50) ~ ∂t

The formal solution is given by

ˆ − iHt ΨS (t) = e ~ ΨS (0) . (4.51) and dHˆ/dt = 0.

37 Interaction of Matter with Quantized Light -I

4.6.2 Heisenberg Picture

Operators vary in time, while the states are time independent (typical in quantum field theory and quantum optics), namely

dΨ H = 0, dt and the operators evolve according to the Heisenberg equation:

ˆ dOH h ˆ ˆi i = OH , H . (4.52) ~ dt

The total Hamiltonian, however, is time independent in the Heisenberg picture! The connection to the Schrödinger picture is given as:

iHˆ t ΨH = e ~ ΨS (t) = ΨS (0) , (4.53) and ˆ ˆ ˆ iHt ˆ − iHt OH (t) = e ~ OSe ~ . (4.54)

Proof:

ˆ ˆ ˆ iHt ˆ − iHt ˆ hΨ(t)| OS |Ψ(t)i = hΨ(0)| e ~ OSe ~ |Ψ (0)i = hΨ (0)| OH (t) |Ψ (0)i | {z } =Oˆ H (t)

4.6.3 Interaction Picture

Intermediate picture where the time evolution is shared between the state and the operator. In the interaction picture the is fefined as:

ˆ ˆ ˆ iH0t iH0t − iHt ΨI (t) = e ~ ΨS (t) = e ~ e ~ ΨS (0) , (4.55)

ˆ where H0 is the free Hamiltonian of the total system. For the case of a two- level ˆ ˆ ˆ system interacting wiht the e.m. field, e.g., H0 = HA + HF . The equation of motion is then given by: ( ) ˆ ˆ ˆ dΨI (t) iH0t iH0t iH0t dΨS (t) i~ = i~ e ~ ΨS (t) + e ~ dt ~ dt ˆ ˆ iH0t nˆ o ˆ ˆ = −H0ΨI (t) + e ~ HΨS (t) = −H0ΨI (t) + HΨS (t)  ˆ ˆ ˆ  = −H0 + H0 + HED ΨS (t) .

And therefore: dΨI (t) ˆ i = HEDΨS (t) (4.56) ~ dt

38 Interaction of Matter with Quantized Light - II

In the interaction picture the time evolution associated to operators is realized by making the Schrödinger operator evolving with the free Hamiltonian:

ˆ ˆ ˆ iH0t ˆ − iH0t OI = e ~ OSe ~ , (4.57) and accordingly: ˆ dOI (t) h ˆ ˆ i i = OI (t) , H0 . (4.58) ~ dt The interaction picture is useful because all the dynamics are held by the interaction Hamiltonian. The free Hamiltonian only accounts for a global phase factor. We can therefore employ the Heisenberg representation to solve the atom-field dynamics. In this case, the time-evolution of the atomic and field operators is given by  −iω0t σˆ− (t) =σ ˆ−e , σˆ± → (4.59) iω0t σˆ+ (t) =σ ˆ+e , and  aˆ (t) =ae ˆ −iωt, a,ˆ aˆ† → (4.60) aˆ† (t) =a ˆ†eiωt.

The time dependent interaction Hamiltonian thus becomes:

ˆ i(ω0−ω)t † −i(ω0−ω)t HED (t) = ~g σˆ+aeˆ +σ ˆ−aˆ e . (4.61)

† Note: We assume ω ≈ ω0 (nearly resonant field). Having eliminated σˆ−aˆ and σˆ+aˆ corresponding to eliminate the terms that oscillate with ω+ω0 ≈ 2ω0 (rotating wave approximation).

5 Interaction of Matter with Quantized Light - II

5.1 Photon Emission and Absorption Rates

Let us now evaluate the matrix elements for absorption and emission processes driven ˆ by HED. Here, |ii and |fi describe the initial and final state of the joint system, respectively.

5.1.1 Absorption

First of all, let us have a look at the absorption process. Let us assume, that at the beginning of this process, the atom is in its ground state and the e.m. field has n photons. The absorption one photon of the e.m. field leads to the excitation of the atom. Therefore let us write the initial and final joint state for the absorption

39 Interaction of Matter with Quantized Light - II process |ii = |n, 1i = |ni |1i , (5.1a)

|fi = |n − 1, 2i = |n − 1i |2i . (5.1b)

Then

ˆ  i(ω0−ω)t −i(ω0−ω)t † hf| HED (t) |ii = ~g e hf| σˆ+aˆ |ii + e hf| σˆ−aˆ |ii

i(ω0−ω)t = ~g{e hn − 1| aˆ |ni h2| σˆ+ |1i

−i(ω0−ω)t † + e hn − 1| aˆ |ni h2| σˆ− |1i}. (5.2) | {z } | {z } 0 0

Using

h2| σˆ+ |1i = h2| 2i h1| 1i = 1, (5.3a) √ √ hn − 1| aˆ |ni = n hn − 1| n − 1i = n, (5.3b) we finally obtain √ ˆ i(ω0−ω)t hf| HED (t) |iiABS = ~g ne . (5.4) ˆ Comment: If there are no photons in the field (n = 0), then hf| HED (t) |iiABS = 0 (the atom cannot absorb photons). This is in accordance with the result from the semi-classical theory of light-matter interaction.

5.1.2 Emission

For the emission process we will consider an atom in its excited state and an e.m. field with n photons, i.e., the initial and final joint states of the emission process are given as |ii = |n, 2i = |ni |2i , (5.5a)

|fi = |n + 1, 1i = |n + 1i |1i . (5.5b)

Then

ˆ  i(ω0−ω)t −i(ω0−ω)t † hf| HED (t) |ii = ~g e hf| σˆ+aˆ |ii + e hf| σˆ−aˆ |ii

i(ω0−ω)t = ~g{e hn − 1| aˆ |ni h2| σˆ+ |1i | {z } | {z } 0 0 −i(ω0−ω)t † + e hn − 1| aˆ |ni h2| σˆ− |1i}. (5.6)

Using

h1| σˆ− |2i = h1| 1i h2| 2i = 1, (5.7a) and √ √ hn + 1| aˆ |ni = n + 1 hn + 1| n + 1i = n + 1, (5.7b)

40 Interaction of Matter with Quantized Light - II we obtain √ ˆ −i(ω0−ω)t hf| HED (t) |iiEMISS = ~g n + 1e . (5.8) √ Comment: Since the matrix element depends on n + 1 it is different from zero also in the case in which there are no photons in the field! This means that an atom in the excited state can interact with the vacuum state of the field, resulting in decaying by emitting a photon (spontaneous emission)

5.1.3 Rate Equation

In order to derive the rate equations we will have a look at a two level system in a single-mode e.m. field. The time evolution of the population changes due to continuous absorption and emission of photons. This change is given by:

dN1 = −WABSN1 + WEMISSN2, (5.9a) dt

dN1 = −WABSN1 + WEMISSN2, (5.9b) dt where N1 is the population of level 1, N2 describes the population of level 2, WABS stands for the absorption probability and WEMIS gives the emission probability. In thermal equilibrium we have:

dN1 dN2 = 0 = (5.10) dt dt and therefore:

WABSN1 = WEMISSN2 (5.11)

And so: ˆ N1 WEMISS hf| HED (t) |ii n + 1 = = EMISS = . (5.12) N W ˆ n 2 ABS hf| HED (t) |iiABS At thermal equilibrium Boltzmann statistics imposes that   N1 ~ω0 = exp . (5.13) N2 kT

Comparing this two equations leads to   n + 1 ~ω0 = exp , (5.14) n kT which fulfils the Bose-Einstein-Statistics 1 n = .5 (5.15)  ~ω0  exp kT − 1

41 Interaction of Matter with Quantized Light - II

5.2 Quantum Rabi Oscillations

5.2.1 The Semiclassical Rabi Model in a Nutshell

In order to have a comparison case, let us first briefly revise the results and the main dynamics of the semiclassical Rabi model. Assuming a monochromatic classical electric field E = E0 cos ωt, the semiclassical dipole Hamiltonian can be written ˆ ˆ ˆ ˆ as Hsc(t) = V0 cos ωt, where V0 = −d · E0. The time dependent state vector can be written as a superposition (with time dependent coefficients) of the ground and excited states as follows:

−iω1t −iω2t |ψ(t)i = C1(t)e |1i + C2(t)e |2i. (5.16)

Using the Schrödinger equation we can obtain a set of two differential equations describing the evolution of the time dependent coefficients (that represent the am- plitude probability of being in the correspondent level). The result is the following:

iV ˙ i(ω−ω0)t C1 = − e C2, (5.17a) ~ iV ˙ −i(ω−ω0)t C2 = e C1, (5.17b) ~ ˆ where V = h2| − d · E0|1i ∈ R and ω0 = ω2 − ω1 is the transition frequency. In order to obtain the previous equations, moreover, we assumed that the electromag- netic field is nearly resonant with the atomic transition. With this approximation the terms oscillating at a frequency (ω + ω0) have been neglected (Rotating Wave Approximation [2]). Solutions of this system of differential equations can be sought for example by taking the derivative of the second equation and substituting the first one into it, thus obtaining the equation of motion for the of the excited state 2 ¨ ˙ V C2 + i(ω − ω0)C2 + C2 = 0. (5.18) 4~2 Solution of this second order equation with the initial condition that the ground state is the only populated state (C1(0) = 1 and C2(0) = 0) brings to the following- expressions for the probability amplitudes C1(t) and C2(t):      i∆t/2 ΩRt i∆ ΩRt C1(t) = e cos − sin , (5.19a) 2 ΩR 2   iV i∆t/2 ΩRt C2(t) = e sin , (5.19b) ~ΩR 2 where ∆ = ω−ω0 is the detuning between the electric field frequency and the atomic transition, and the quantity ΩR is known as Rabi frequency and its expression is the

42 Interaction of Matter with Quantized Light - II

1.0

0.8

0.6 L D=0 t H 2 D=0.5 P 0.4 D=1 0.2

0.0 0 5 10 15 20 25

WRt

Figure 5.1: Plot of P2(t) versus t for various values of the detuning ∆. Figure adapted from Ref. [2]. following: r 2 2 V ΩR = ∆ + . (5.20) ~2 The Rabi frequency represents the frequency at which the oscillation of population between the ground and the excited state of the atom takes place, triggered by the interaction with the electromagnetic field. Theprobability that the atom is in the excited state is in fat given by

2   2 V 2 ΩRt P2(t) = |C2(t)| = 2 2 sin , (5.21) ~ ΩR 2 which for the case of perfect detuning (∆ = 0) assumes the easier form   2 V t P2(t) = sin . (5.22) 2~

The time evolution of the population probability of the excited state is reported in Fig. 5.1 for various values of ∆. In the case of perfect resonance, a complete transfer of population between the ground and the excited state takes place at the time t = π~/V . An electromagnetic field characterized by the frequency ω that realized this condition is said to be a π-pulse. It is also useful to introduce the population inversion probability       2 V t 2 V t V t W (t) = P2(t) − P1(t) = sin − cos = − cos . (5.23) 2~ 2~ ~

43 Interaction of Matter with Quantized Light - II

At the time t = π~/2V , one can note that the population inversion is W (π~/2V ) = 0, and the population is coherently shared between the ground and the excited state, i.e., the state of the system is given by

1 |ψ(t)i = √ (|1i + i|2i) . (5.24) 2

5.2.2 The quantum Rabi Model

We then want to investigate if the process of Rabi oscillations is influenced by the quantization of the field or not. Before starting solving the problem of calculating the dynamics of the Jaynes-Cummings model, let us make some preliminary obser- vations. First of all we assume that no electrons (or atomic species in general) are lost by the system, and that therefore the population of the atomic system must remain constant over time. If we define the electron number operator as

ˆ PE =σ ˆ−σˆ+σˆ+σˆ− = |1ih1| + |2ih2| = 1, (5.25) then it is immediate to see that

h ˆ ˆi PE, H = 0. (5.26)

Since we are considering a closed system, not only the energy is conserved, but also the total number of excitations is conserved. This means that whenever a photon gets annihilated the atom (in order to conserve energy) must absorb it and make the transition to the excited astate, and every time a photon is created, the atom must decay to the ground state. We can formalize this by defining the following excitation number operator ˆ † NE =a ˆ aˆ +σ ˆ−σˆ+, (5.27) so that h ˆ ˆi Ne, H = 0. (5.28)

This allows us to rewrite the total Jaynes-Cummings Hamiltonian as the sum of ˆ ˆ two parts, one depending on PE and NE and the other one containing the dipolar interaction. Without loss of generality, we can consider first the simple case of perfect resonance, thus putting ∆ = ω − ω0 = 0. As we did for the semiclassical case, the state of the system can be written as a superposition of the initial state |ii ≡ |n, 2i and the final state |fi ≡ |n + 1, 1i (i.e. we are considering the case in which the atom is excited and decays to the ground state by emitting a photon) as follows:

|ψ(t)i = Ci(t)|ii + Cf (t)|fi, (5.29)

44 Interaction of Matter with Quantized Light - II

with the initial conditions imposing that Ci(0) = 1 and Cf (0) = 0. If we now use the above expression for the state and the interaction Hamiltonian ito solve the the Schrödinger equation for the considered system, we obtain the following set of coupled differential equations for the coefficients Ci(t) and Cf (t): √ ˙ Ci = −ig n + 1Cf , (5.30a) √ ˙ Cf = −ig n + 1Ci. (5.30b)

Taking the derivative of the first equation and eliminating Cf (t) by sung the second one we obtain the following equation of motion for the amplitude probability Ci(t):

¨ 2 Ci + g (n + 1)Ci = 0, (5.31) whose solution compatible with the initial condition is

Ω(n)t Ci(t) = cos , (5.32a) 2 Ω(n)t Cf (t) = −i sin , (5.32b) 2 √ where Ω(n) = 2g n + 1 is the quantum Rabi frequency, whose expression should be compared with the semiclassical result (at prefect resonance) ΩR = V/~. In order to draw some conclusion, we can define also in this case the probabilities that the system is (at a certain time t) in the initial or final state as   2 2 Ω(n)t Pi(t) = |Ci(t)| = cos , (5.33a) 2   2 2 Ω(n)t Pf (t) = |Cf (t)| = sin , (5.33b) 2 and the population inversion probability as

W (t) = hψ(t)|σˆ3|ψ(t)i = Pi(t) − Pf (t) = cos Ω(n)t, (5.34) that, apart from amines sign due to the different choice of initial condition, coincides with the result of Eq. (5.23), where the semiclassical Rabi frenquecy ΩR has to be substituted by its quantum counterpart Ω(n). In the semiclassical case, the Rabi frequency depends on the amplitude of the ˆ electric field through V = h2| − d · E0|1i, and if the amplitude of the field is zero

(meaning vettE0 = 0), there is no field, the Rabi frequency is consequently zero and there is no population oscillation. In the fully quantized model, however, the Rabi frequency depends on the number of photons in the field through the quantity √ n + 1. As can be noted, Ω(0) 6= 0. This has a straightforward physical meaning. If the electromagnetic field is in its vacuum state, the atom has then a nonzero

45 Interaction of Matter with Quantized Light - II probability of spontaneously emitting a photon and subsequently re-absorbing it and then re-emitting it etc, originating the so-called vacuum Rabi fluctuations. This process, of course, is a manifestation of the quantum nature of the electromagnetic field and has no classical counterpart. However, apart from this difference, the dynamics of an atom interacting with an electromagnetic field with a definite number of photons (i.e., in a Fock state) is very much similar to the semiclassical case, as it gives rise to a regular and periodic dynamics, namely the Rabi oscillations. This could be a bit confusing, since we said in the previous chapters that number states are among the most nonclassical states of the electromagnetic field (they have no definite electric field!). However, the induced dynamics are practically the semiclassical one. One would have expected this result by using a coherent state, that is the most classical state of the field. However, in this case, the dynamics is much richer, and very different from the semiclassical case.

5.2.3 Interaction of an Atom with a Coherent State

To analyze the interaction of a two level system with an electromagnetic field de- scribed by a coherent state, we must first of all generalize the solution given above to the case of an atom interacting with a field in a superposition of Fock states. Once we have this result, in fact, writing down the solution for the coherent state case will be very easy. To do that we assume the atomic state at the time t = 0 to be in a superposition of ground and excited state

|ψ(0)iatom = C1|1i + C2|2i, (5.35) and the electromagnetic field to be in a superposition of number states

∞ X |ψ(0)ifield = Cn|ni. (5.36) n=0

As we said before, the initial state of the joint system atom+field can be written as the direct product of the corresponding Hilbert spaces, resulting in

( ∞ ) X |ψ(0)i = |ψ(0)iatom ⊗ |ψ(0)ifield = Cn|ni {C1|1i + C2|2i} . (5.37) n=0

At the generic time t, the state of the electromagnetic field can only be changed of one photon, either created (emission) or annihilated (absorption) and therefore

∞ X |ψ(t)ifield = {Cn|ni + Cn±1|n ± 1i} . (5.38) n=0

46 Interaction of Matter with Quantized Light - II

For the atom, instead, the coefficients C1 and C2 are time dependent and therefore

|ψ(t)iatom = C1(t)|1i + C2(t)|2i. (5.39)

The joint state of the system atom+field at the generic time t is therefore given by

|ψ(t)i = |ψ(t)iatom ⊗ |ψ(t)ifield. (5.40)

This state must be a solution of the Schrödinger equation. Therefore:

∂ X h i i |ψ(t)i = i (C |ni + C |n + 1i + C |n − 1i) C˙ (t)|1i + C˙ (t)|2i , ~∂t ~ n n+1 n−1 1 2 n (5.41) and

ˆ X † HED|ψ(t)i = ~g σˆ+aˆ +σ ˆ−aˆ (Cn|ni + Cn+1|n + 1i + Cn−1|n − 1i) n × [C1(t)|1i + C2(t)|2i] X n√ √ = ~g nCnC1(t)|n − 1, 2i + n + 1CnC2(t)|n + 1, 1i n √ √ o + n + 1Cn+1C1(t)|n, 2i + nCn−1C2(t)|n, 1i , (5.42) where in the last line we neglected the terms containing |n ± 2i since they represent states of the electromagnetic field that cannot be reached (since at maximum the number of photons in the field can change by one unity by assumption). If we want to obtain the equation of motion for C1(t) and C2(t) we have to project the Schrödinger equation ∂ ˆ i |ψ(t)i = HED|ψ(t)i (5.43) ~∂t onto the four different states |n, 1i (corresponding to the atom in the ground state and the field with n photons), |n, 2i (atom in the excited state and n photons in the field), |n + 1, 1i (the atom decayed to the ground state by emitting a photon) and |n − 1, 2i (the atom absorbed a photon to go to the excited state). We then obtain the following four equations of motion:

¨ 2 CnC1(t) + g nCnC1(t) = 0, (5.44a) ¨ 2 Cn+1C1(t) + g (n + 1)Cn+1C1(t) = 0, (5.44b) ¨ 2 CnC2(t) + g (n + 1)CnC2(t) = 0, (5.44c) ¨ 2 Cn−1C1(t) + g nCn−1C1(t) = 0. (5.44d)

These equations, together with the initial condition (5.37) (and after having properly shifted the indices so that it is possible to write the solution as a single summation

47 Interaction of Matter with Quantized Light - II in n) bring the following result:

∞ X n h  √   √ i |ψ(t)i = C2Cn cos gt n + 1 − iC1Cn+1 sin gt n + 1 |2i n=0  √  √  o + −iC2Cn−1 sin gt n + C1Cn cos gt n |1i |ni. (5.45)

In general, this state is an entangled state, meaning that it is not possible anymore to write this state as a simple product between a state of the field and a state of the atomic system, as it is done in Eq. (5.37). Therefore, in the general case of interaction of the field with an atomic system, it does not make sense anymore to talk about the field and the atom, but rather we must speak of the atom-field system, as the general wave function is non-separable anymore (entangled).

If we assume that the atom is initially in the excited state (i.e., we set C1(0) = 0 and C2(0) = 1), the previous solution can be still written as a product between an evolved field state and the atomic states as follows:

|ψ(t)i = |ψ1(t)i|1i + |ψ2(t)i|2i, (5.46) where

∞ X  √  |ψ1(t)i = −i Cn sin gt n + 1 |n + 1i, (5.47a) n=0 ∞ X  √  |ψ2(t)i = Cn cos gt n + 1 |ni. (5.47b) n=0

In this case we can calculate analytically the atomic inversion probability as

W (t) = hψ(t)|σˆ3|ψ(t)i

= hψ2(t)|ψ2(t)i − hψ1(t)|ψ1(t)i ∞ √ X 2   = |Cn| cos 2gt n + 1 . (5.48) n=0

We now have a very general expression for the atomic inversion as a function of the expansion coefficients Cn. If we now want to study the dynamics of the system under the action of a coherent state, we just have t substitute in the previous expression the expression for the expansion coefficients of a coherent state in terms of number states, namely n −|α|2/2 α Cn = e √ . (5.49) n! If we do that, eq. (5.48) becomes (we call n¯ = |α|2 as the mean number of photons

48 Interaction of Matter with Quantized Light - II

1.0 HaL

0.5 L t H 0.0 W

-0.5

0 5 10 15 20 25 T

1.0 HbL

0.5 L t H 0.0 W

-0.5

0 10 20 30 40 50 T

Figure 5.2: (a) Atomic inversion with the field initially in a coherent state with n¯ = 5. (b) Same as (a) but showing the evolution for a longer time, beyond the first revival time. Both graphs are plotted against the normalized time T = gt. The figures are adapted from Ref. [2]. in the coherent state)

∞ X n¯n  √  W (t) = e−n¯ cos 2gt n + 1 . (5.50) n! n=0

In Fig. 5.2 a plot of the atomic inversion is given. As it appears clear by simply

49 Interaction of Matter with Quantized Light - II looking at the graph, the dynamics of the system in this case is profoundly different from the semiclassical case. Let us try to understand what are the characteristics of this dynamics. First of all, from Fig. 5.2(a) we note that the Rabi oscillation appears to damp out after a certain time. If we however wait long enough [Fig. 5.2(b)], the oscillations start to revive, although not in phase with the initial oscillations. This dynamics repeats almost periodically in time, each time making more difficult to distinguish between a collapse and a revival. We now try to interpret this strange behavior. Let us start with the following observation: Eq. (5.50) is basically a series of cosines, each one characterized by a frequency Ω(n) that depends on the number of photons in the field. If the average number of photons is n¯, then it makes sense to say that the dominant Rabi frequency will be Ω(¯n), namely

p √ Ω(¯n) = 2g n¯+1 ' 2g n, (5.51) where the last approximation is valid in the limit n¯  1, as it is usually the case for a coherent state, where n¯ ' 106. Since the photon number is not perfectly determined, however, there will be a range of frequencies around Ω(¯n) that will contribute sensibly to the dynamics. Calling ∆n the uncertainty in the photon number, the relevant frequency range should be between Ω(¯n − ∆n) and Ω(¯n + ∆n). If we want to say something about the dynamics of the system, we therefore have to investigate this frequency range

Collapse The collapse time tC is defined as the instant of time in which the relative dephasing between the various frequency components that are present in the initial expression of W (t) is such that they all interfere destructively, leaving no atomic inversion, therefore collapsing the dynamics. Using the uncertainty relation existing between time and frequency, we can estimate the collapse time tC as follows:

tC [Ω(¯n + ∆n) − Ω(¯n − ∆n)] ' 1. (5.52) √ Since for a coherent state ∆n = n¯, we have that

q √ √  1 1/2 ω(¯n ± ∆n) ' 2g n¯ ± n¯ ' 2g n¯ 1 ± √ n¯ √  1  √ ' 2g n¯ 1 ± √ = 2g n¯ ± g, (5.53) 2 n¯ and therefore the collapse time is given by

1 tC ' . (5.54) 2g

50 Interaction of Matter with Quantized Light - II

Interestingly, the collapse time is independent of the average number of photons in the coherent state. This derivation is not fully rigorous, nevertheless it catches the salient features. For a more rigorous derivation, the reader is addressed to Ref. [2].

Revivals We now turn to our attention to determine the revival time tR, corre- sponding to the time in which all the relevant frequencies are oscillating in phase (or most of them are in the case of a fractional revival), thus reproducing the ini- tial oscillatory pattern. By means of simple considerations based on interference of adjacent frequencies, the revival time can be estimated by means of the following relation:

[Ω(¯n + 1) − Ω(¯n)] tR = 2πk, (5.55) where k ∈ N. By using the approximated expressions of Ω(¯n + 1) and Ω(¯n) given above, we can estimate the revival time to be given by

2π  √ tR ' k n.¯ (5.56) g

In this case, the revival time depends on the average number of photons in the field as a monotonically increasing function. The more photons are in the field, therefore, the more is tR close to infinity. As for the case of the collapse time, a more rigorous derivation of tR is given in ref. [2].

5.3 Wigner-Weißkopf-Theory [5]

In this section we will study the interaction of a 2 level atom with a multimode field, where the model multimode field is seen as a continuum of states. In order to study the Wigner-Weißkopf model, we need to introduce the continuous limit to our discrete model of quantization. This can be done by considering two things, which are first taking the limit V → ∞ for the cavity volume and then, accordingly we will need to generalize the definition of a,ˆ aˆ† to the continuous case.

Figure 5.3: Geometry of optical cavity (taken from [1])

51 Interaction of Matter with Quantized Light - II

Since we consider the case V → ∞, the number of allowed k-vectors inside the cavity rises, so that we can consider them as continuous. Therefore we will substitute the sum over the k-vectors by an integration over the density of states in k-space D (k): 2 2 X X X Z −→ d3k D (k) . (5.57) k λ=1 λ=1 In order to obtain D (k), we consider a quadratic cavity as shown in Fig. 5.3, therefore: 2π k = (nx, ny, nz) , (5.58) L with ni ∈ N, but only one of the ni can be zero, since the field in the cavity vanishes if two or three n’s are zero. For a given combination (nx, ny, nz), there is only one (2π)3 state in the volume V . Therefore

V D (k) = . (5.59) (2π)3

How do we generalize operators?

With this we get: √ aˆkλ → ∆ωaˆλ (ω) (5.60a) √ † † aˆkλ → ∆ωaˆλ (ω) (5.60b)

h † i 0 aˆkλ, aˆkµ = δ (ω − ω ) δλµ, (5.60c)

2πc where ∆ω = c∆k = L is a normalization factor introduced to preserve the bosone nature of the commutator (5.60c).The electric field operator (for a linearly polarized field propagating along z-axis) is given by:

1   2 Z ω h z z i ˆ ~ −iω(t− c ) † iω(t− c ) ET (z, t) = i dω aˆ (ω) e +a ˆ (ω) e . (5.61) 4πε0cA

The magnetic field operator Bˆ is defined accordingly.

The Model

Our model considers the interaction of a 2-level atom with a single-mode field. For simplicity we only take single photon processes into account. In the beginning of our considerations we assume the atom to be in the excited state and assume no photons in the field, i.e., |Φ2i |0i. After the interaction we want our atom to be in the ground state (decay by emission of a photon) |Φ1i |Φ(k)i, and therefore to have one photon in the field. The emitted photon can have any momentum k, meaning that it can be emitted in any direction! This is due to the fact that the atom is inserted in an isotropic vacuum, and there is no preferred direction of emission. Therefore, to

52 Interaction of Matter with Quantized Light - II correctly account for all the possibilities, we need to integrate over all the possible values of k.

Since both {|Φ1i , |Φ2i} and |Φ(k)i are orthogonal basis in their respective Hilbert spaces, |Ψ (0)i = |Φ2i |0i, and we can write the state at time t as superposition of the atomic and field states as follows:   Z   −iEt 3 −iε (k) t |Ψ(t)i = a (t) exp |Φ2i + d k b (k, t) exp |Φ(k)i (5.62) ~ ~ where ε (k) is the energy of the photon with momentum k. In this case, for the sake of simplicity, we assume to use the ground state as a reference for the energy scale, so that E = E2 − E1 = E2. Eq. (5.62) must solve the Schrödinger equation:

∂  ˆ ˆ  i |Ψ(t)i = H0 + V |Ψ(t)i , (5.63) ~∂t where Vˆ is the atom-field interaction, e.g. the dipole Hamiltonian. Calculating the left-hand-side (LHS) and the right-hand-side (RHS) separately, we have: LHS

∂ da (t) −iEt −iEt i~ |Ψ(t)i = i~ exp |Φ2i + Ea (t) exp |Φ1i ∂t dt ~ ~ Z   3 db (k, t) −iE (k) t + i~ d k exp |Φ(k)i dt ~ Z −iE (k) t + d3k E (k) b (k, t) exp |Φ(k)i ~

RHS

    Z   ˆ ˆ −iEt 3 −iE (k) t H0 + V |Ψ(t)i = Ea (t) exp |Φ2i + d k E (k) b (k, t) exp |Φ(k)i ~ ~   Z   −iEt ˆ 3 −iE (k) t ˆ + a (t) exp V |Φ2i + d k b (k, t) exp V |Φ(k)i ~ ~

Remember Vˆ is a dipole-line separator given by

ˆ hΦk| V |Φki = 0, (5.64) and the photon states are orthogonal, i.e.,

hΦ(q)| Φ(k)i = δ (q − k) . (5.65)

If we project the Schrödinger eq. (5.63) onto hΦ2| and hΦ(k)|:

da (t) Z i = d3k b (k, t) M (k) e−iω(k)t, (5.66a) ~ dt

53 Interaction of Matter with Quantized Light - II

db (k, t) i = a (t) M ∗ (k) eiω(k)t, (5.66b) ~ dt ˆ where ~ω (k) = ε (k) − E is the transmission energy, M (k) = hΦ2| V |Φ(k)i gives the matrix element and with the initial conditions a (0) = 1, b (k, 0) = 0. Eq. (5.66b) can be formally solved by multiplying by dt and integrating, namely

t t ∗ Z M (k) Z 0 db (k, t0) = dt0 a (t0) eiω(k)t , i~ 0 0 thus obtaining t ∗ M (k) Z 0 b (k, t) = dt0 a (t0) eiω(k)t . (5.67) i~ 0 Substituting the expression of b (k, t) by Eq. (5.67) we obtain for Eq. (5.66a)

t 2 da (t) 1 Z |M (k)| Z 0 = d3k eiω(k)t dt0 a (t0) eiω(k)t , dt i~ i~ 0 i.e., t Z Z da (t) 1 2 0 = − d3k |M (k)| eiω(k)t dt0 a (t0) eiω(k)t . (5.68) dt ~2 0 The last term of Eq. (5.68) makes the problem non-Markovian, i.e. the system has memory and therefore depends on its past history. We need to make it Markovian, i.e. a (t2 > t1) does not have to depend from all a (t) with t ∈ [0, t1]. Here we use a little trick: we will multiply the equation by e−λt and integrate in time. We then have from Eq. (5.68) LHS

∞ ∞ Z da (t) Z  d  dt e−λt = dt a (t) e−λt + λa (t) e−λt dt dt 0 0 ∞ Z −λt ∞ −λt = a (t) e 0 + λ dt a (t) e 0 ∞ Z = λ dt a (t) e−λt − 1,

0

54 Interaction of Matter with Quantized Light - II

RHS

∞ t 1 Z Z Z 0 − d3k |M (k)|2 dt e−λte−iω(k)t dt0 a (t0) eiω(k)t ~2 0 0 | {z } t ∞ 1 Z Z 0 Z = − d3k |M (k)|2 dt0 a (t0) eiω(k)t dt e−[λ−iω(k)]t ~2 0 t0 t Z Z   1 3 2 0 0 iω(k)t0 −1 −[λ−iω(k)]t ∞ = − d k |M (k)| dt a (t ) e e 0 ~2 λ + iω (k) t 0 t 2 1 Z |M (k)| Z 0 = − d3k dt0 a (t0) eλt . ~2 λ + iω (k) 0

Here we used the fact that the underlined integral can be solved in two ways (see Fig. R ∞ 5.4), i.e., whether considering t as fixed and integrating along a vertical strip, 0 dt, R ∞ R t 0 R ∞ which corresponds to 0 dt 0 dt , or integration along a horizontal strip (i.e. t0 dt) R t 0 where we first integrate over 0 dt , which corresponds to swithcing the integrations R t 0 R ∞ in Eq. (5.68), i.e., 0 dt t0 dt.

Figure 5.4: Interpretation of the integral dt0dt

In this case, we use the second possibility, to integrate the RHS of Eq. (5.68)and then obtaining

∞ " #−1 Z 1 Z |M (k)|2 dt e−λta (t) = λ + d3k . (5.69) ~2 λ + iω (k) 0

2 Note: If we go back to Eq. (5.68) and assume |M (k)| to be independent of k we

55 Interaction of Matter with Quantized Light - II obtain:

t Z Z da (t) 1 2 0 = − d3k |M (k)| e−iω(k)t dt0 a (t0) eiω(k)t (5.70) dt ~2 0 t 2 |M| Z Z 0 = − d3k e−iω(k)t dt0 a (t0) eiω(k)t (5.71) ~2 0 t 2 2 |M| Z Z 0 |M| = − dt0 a (t0) d3k e−iω(k)(t−t ) = − a (t) , (5.72) ~2 | {z } ~2 0 ∼δ(t−t0) and therefore: a (t) = a (0) e−γt, (5.73)

2 γ = |M| with ~2 . This solution is consistent with the assumption of having an isotropic vacuum. In this case, in fact, there is no reason why the emitted photon should prefer one k vector to the other one, as they are all equally probable. Therefore, the matrix element M (k) cannot (by symmetry considerations) depend on k. Now let us assume that this would still be a good solution even for the case that M (k) depends on k and solve the problem for λ → 0. Keeping a (t) = e−zt as Ansatz, we then have: ∞ Z −λt −zt 1 −(z+λ)t ∞ 1 dt e e = − e = . (5.74) z + λ 0 z + λ 0 and therefore Eq. 5.69 becomes

" #−1 1 i Z |M (k)|2 = λ − d3k , (5.75) z + λ ~2 ω (k) − iλ where we have used the fact that 1 1 1 1 i 1 = = − . ~2 λ + iω ~2 i (ω − iλ) ~2 ω − iλ

As λ is an artificial parameter inserted for facilitating the solution of Eq. (5.68), the real solution must not depend on it. To do so, we then take the limit of (5.75) for λ → 0, thus obtaining ( ) i Z |M (k)|2 z = lim − d3k λ→0+ ~2 ω (k) − iλ ( ) i Z |M (k)|2 [ω (k) + iλ] = lim − d3k λ→0+ ~2 ω2 (k) + λ2 ( ) i Z ω (k) |M (k)|2 1 Z λ |M (k)|2 = lim − d3k + d3k , (5.76) λ→0+ ~2 ω2 (k) + λ2 ~2 ω2 (k) + λ2

56 Interaction of Matter with Quantized Light - II

By using λ lim = πδ (ω (k)) , (5.77a) λ→0+ ω2 (k) + λ2 and 1 δ (ax) = δ (x) . (5.77b) |a| Eq. (5.76) becomes

Z 2 Z i 3 |M (k)| π 3 2 z = − d k + d k |M (k)| δ (~ω (k)) . (5.78) ~2 ω (k) ~

In order to better understand this result, we define

2 ˆ Z hΦ1| V |Φ(k)i E0 = E + d3k , (5.79a) E − ε (k)

Z 2 2π 3 ˆ γ = d k hΦ1| V |Φ(k)i δ (ε (k) − E) , (5.79b) ~ so that z = iE0/~ + πγ/(2~). Then, the solution for a (t) reads as follows:

0 − iE t − γ t a (t) = e ~ e 2 , (5.80)

2 and the occupation probability can be simply calculated as |a (t)| , i.e.,

2 −γt Pl (t) = |a (t)| = e . (5.81)

The excited state of the atom decays exponentially in time due to the interaction with the electromagnetic field in the vacuum state.

Let us then now calculate, what is the probability that at time t → ∞ the sys- tem is in the state |Φ(k)i (atom decayed and emitted a photon). To do that, we calculate b (k, t) using the above result a (t) = e−zt, thus obtaining

∞ ∗ M (k) Z 0 b (k, ∞) = dt0 e−[z−iω(k)]t i~ 0  ∗   Z −1 M (k) π 3 0 2 = d k |M (k)| δ (~ω (k)) − i∆ , (5.82) i~ ~ where ε (k) − E 1 Z |M (k)|2 ∆ = − d3k0 . (5.83) ~ ~ E − ε (k0)

57 Quantum Mechanics of Beam Splitters

Rearranging the terms gives the following final result:

M ∗ (k) i~γ b (k, ∞) = 2 + , (5.84) ε (k0) − E − 1 R d3k0 |M(k)| 2 ~ E−ε(k0) and therefore the probability that a photon with momentum k has been emitted is given by 2 2 |M (k)| Pk = |b (k, ∞)| = 2 . (5.85) 0 2 ~γ  [ε (k ) − E] + 2 Usually, for simplicity, in our models we assume that the transition of of electrons between energetic level has a delta like resonance function, i.e., the emitted photon is monochromatic. But as seen in Eq. (5.85) this is not true, because the transition probability has a Lorentzian shape with non zero bandwidth. Therefore the emitted photons show a spectral broadening, which we call natural linewidth. The natural linewidth gives the lower limit to the spectral bandwidth of our photon.

6 Quantum Mechanics of Beam Splitters

6.1 Classical theory

Figure 6.1: Schematic sketch of a classical beam splitter with incoming and outgoing fields

The beam splitter in Fig. 6.1 is assumed to be ideal, i.e., it has two different surfaces and its thickness can be assumed to be small enough to be neglected. The reflection and transmission coefficients are (r, t) and (r0, t0) for the first (solid line in Fig. 6.1) and second (dashed line in Fig. 6.1) surface, respectively. The incoming and

58 Quantum Mechanics of Beam Splitters outgoing fields can be then written as

1 −iωt  ik1·r ik2·r EIN (r, t) = e a1e + a2e , (6.1a) 2 1 −iωt  ik3·r ik4·r EOUT (r, t) = e a3e + a4e , (6.1b) 2

Where ai are the amplitudes of the ingoing (a1, a2) and outgoing (a3, a4) fields, respectively. We can write this in a matrix form, obtaining ! ! ! ! a t r0 a a 4 = 1 = B 1 , 0 (6.2) a3 r t a2 a2

2 2 2 2 where due to energy conservation |a1| + |a2| = |a3| + |a4| . If we define: ! a1 v = , (6.3a) a2 and u = Bv, (6.3b) then the total energy reads

2 Ein = |Ein| = (v, v) , (6.4) and †  Eout = (u, u) = (Bv,Bv) = v,B Bv = Ein = (v, v) , (6.5)

†  † where we used (AB, C) = A, B C . We get that Eout = Ein if B B = 1. What does this mean? If B†B = 1 → B† = B−1, then ! ! t r0 t∗ r∗ B = → B† = , t t0 r0∗ t0∗ and ! 1 t0 −r0 B−1 = , tt0 − rr0 −r t | {z } where tt0 − rr0 is just a phase factor. Therefore  t0 = t∗, B†B = 1 → r0 = −r∗,

59 Quantum Mechanics of Beam Splitters and so the beam splitter can be represented by the following matrix ! t −r∗ B = . (6.6) r∗ t∗

Moreover it follows that: t0  t ∗ = − . (6.7) r0 r This result allows us to obtain the following relation for the phases of the transmis- sion and reflexion coefficients:

(Φt0 − Φr0 ) + (Φt − Φr) = ∓π + 2nπ.

If we choose π Φ − Φ = Φ 0 − Φ 0 = , t r t r 2 this leads to ! t ir B = . (6.8) ir t Note:

1. Energy conservation is fulfilled:

|t|2 + |r|2 = 1

2 2 2. because of the first point we can write t = cos ϑ, r = sin ϑ, so that |t| + |r| = cos2 ϑ + sin2 ϑ = 1 and so ! cos ϑ i sin ϑ B = . i sin ϑ cos ϑ

60 Quantum Mechanics of Beam Splitters

6.2 Quantum Theory

Figure 6.2: Schematic sketch of a QM beam splitter with incoming and outgoing field

Let us consider a beam splitter with very high transmission (ε  1). We want an operator representation for a beam splitter from a quantum mechanical point of view, therefore the operator must be unitary and linear in aˆ†, aˆ. Hence let us look for an exponential operator. We obtain the simple result: ! ! aˆ4 aˆ1 = B . (6.9) aˆ3 aˆ2

Proof:

We want a unitary operator Uˆ such that1:

ˆ ˆ Uaˆ1U =a ˆ4, (6.10a)

ˆ ˆ Uaˆ2U =a ˆ3. (6.10b)

1 Note that the transformation between the input mode operators aˆ1 and aˆ2 and the correspon- dent output mode operators aˆ3 and aˆ4 implemented by the unitary operator Uˆ is a particular case of a more general class of transformations, namely the so-called Bogoliubov transformations. Bogoli- ubov transformations are unitary transformations from a unitary representation of a commutation (or anti-commutation) algebra into another, equivalent and unitary representation. They are often employed in theoretical physics and condensed matter theory to diagonalize Hamiltonians, leading to the steady-state solutions of the correspondent Schrödinger equations. Moreover, Bogoliubov transformation are an essential ingredient to understand the Unruh effect and the Hawking radi- ation. The interested reader can consult Refs. [6, 7, 8] for more detailed informations about the use of Bogolyubov transformations in quantum field theory and condensed matter theory.

61 Quantum Mechanics of Beam Splitters

Since Uˆ is unitary, we can write it in the form of an exponential operator, namely

Uˆ = e−iHˆ , with Hˆ † = H,ˆ so that ˆ |outi = Uˆ |ini = e−iH |ini . (6.11)

If we assume low reflectivity (ε  1), then Hˆ = Hˆ (ε) can be expanded in a Taylor series with respect to ε, i.e.,

ˆ ˆ ˆ 0 2 H (ε) = H0 + EH + O E .

Plugging Hˆ (ε) in (6.11) and Taylor expanding of the exponential we obtain:

ˆ ˆ ˆ 0 |outi = e−iH |ini = e−i[H0+EH ] |ini   ˆ 0 ˆ 0 ≈ 1 − iEH |ini = |k1i − iEH |k1i

∝ |k1i + E |k2i . (6.12)

Note that ˆ 0 H |k1i ∝ |k2i which leads to ˆ 0 hk2| H |k1i = c − number (6.13)

How can we find an expression for Hˆ 0? Given Eq. (6.13), we can say that

ˆ 0 † ˆ H =a ˆ2aˆ1 + Q, with ˆ hk2| Q |k1i = 0.    † From Hˆ 0 = Hˆ 0 , follows that

† ˆ † ˆ† aˆ2aˆ1 + Q =a ˆ1aˆ2 + Q , and therefore it is not difficult to prove that

ˆ † Q =a ˆ1aˆ2.

Therefore we obtain: ˆ 0 † † H =a ˆ1aˆ2 +a ˆ2aˆ1, (6.14)

62 Quantum Mechanics of Beam Splitters which in general reads ˆ 0 ∗ † † H = g aˆ1aˆ2 + gaˆ2aˆ1. (6.15)

The Beam Splitter operator can be then written as follows:

∗ † † iE(g aˆ aˆ2+gaˆ aˆ1) Uˆ = e 1 2 . (6.16)

Now we will prove, that the matrix representation of the beam splitter operator Uˆ introduced above is given by B, therefore we introduce: ! aˆ1 x = , (6.17) aˆ2 and build the following scalar product

(x,Mx) = x†Mx † † † † ˆ = m11aˆ1aˆ1 + m12aˆ1aˆ2 + m21aˆ2aˆ1 + m22aˆ2aˆ2 = B, (6.18) with M = M †. Let us now assume to introduce an operator Uˆ (λ) such that

ˆ Uˆ (λ) = eiλB. (6.19)

ˆ In this case Uˆ (1) = eiB and calculate

UˆxUˆ † = eiλBˆ xe−iλBˆ ≡ fˆ(λ) .

By using the Heisenberg equation we can see how fˆ(λ) evolves with respect to λ, i.e., dfˆ h i i = i B,ˆ fˆ(λ) , (6.20) dλ with fˆ(0) =x ˆ. Calculating the RHS gives h i h i B,ˆ fˆ(λ) = B,ˆ eiλBˆ xeˆ −iλBˆ

= Beˆ iλBˆ xeˆ −iλBˆ − eiλBˆ xeˆ −iλBˆ Bˆ ˆ h i ˆ = eiλB B,ˆ xˆ e−iλB (6.21)

ˆ (note that Bˆ and e±iλB commute). Computing the matrix element of the commu- ˆ † tator, with B = mijaˆi aˆj gives h i h i h i ˆ † † B, xˆ = mijaˆi aˆj, aˆk = mij aˆi aˆj, aˆk

n † h † i o = mij aˆi [ˆaj, aˆk] + aˆi , aˆk aˆj

= −mijδikaˆj = −mkjaˆj = (Mxˆ)k . (6.22)

63 Quantum Mechanics of Beam Splitters

By then generalizing the previous result we get: h i B,ˆ fˆ(λ) = −eiλBˆ Mxeˆ −iλBˆ

= −MeiλBˆ xeˆ −iλBˆ = −Mfˆ(λ) (6.23)

(note that M is a c-number matrix), i.e., it is not an operator anymore. Eq. (6.20) now becomes dfˆ = −iMf,ˆ dλ whose solution is given by fˆ(λ) = e−iλM x. (6.24)

Observe that Uˆ †xUˆ can be then represented in two equivalent ways, i.e.,

∗ † † † ∗ † † i(g aˆ aˆ2+gaˆ aˆ1) −i(gaˆ aˆ1+g aˆ aˆ2) Uˆ xUˆ = e 1 2 xe 2 1 , (6.25a) and iλBˆ −iλBˆ im aˆ†aˆ −im aˆ aˆ† Uˆ †xUˆ = e xe = e ij i j Me ij i j . (6.25b)

Comparing these two equivalent representations, i.e., the RHS of (6.25a) with the RHS of (6.25b) we get ! 0 g∗ M = . (6.26) g 0

Exponentiating M we have the following result

∞ k g∗ ! X (−i) cos |g| −i sin |g| e−iM = M k = |g| . (6.27) k! g k=0 −i |g| sin |g| cos |g|

g g∗ If the phases |g| and |g| are properly chosen, then

g∗ ! ! cos |g| −i |g| sin |g| t ir g → , −i |g| sin |g| cos |g| ir t therefore t = cos |g| , (6.28a) and r = −ieiφg sin |g| . (6.28b)

64 Quantum Mechanics of Beam Splitters

6.3 Classical VS Quantum Beam Splitter

6.3.1 Classical BS

We consider a classical beam splitter, where we have a field in only one arm. There- fore the beam splitter equation reads ! ! ! a4 t ir a1 = . (6.29) a3 ir t 0

The incoming and outgoing energy then are given by

2 Ein = |v1| , (6.30a) and

2 2 2 2 Eout = |a3| + |a4| = |irv1| + |ta1| 2 2 2 2 = |r| + |t| |v1| = |a1| . (6.30b) | {z } =1

As one can expect the energy conservation is fulfilled in this case.

6.3.2 Quantum BS

As in the case of a classical beam splitter we want to study the case of a quantum beam splitter where we have a field only in one input. The beam splitter equation for the quantum case then reads ! ! ! aˆ4 t ir aˆ1 = . (6.31) aˆ3 ir t 0

Since the beam splitter is a passive device, the nature of the incoming radiation must not be changed. Therefore, if the bosonic commutation relations (see Eq. (6.32)) hold for the incoming field, they must also hold for the outgoing fields.

h †i aˆi, aˆj = δij, (6.32a)

h † †i [ˆai, aˆj] = 0 = aˆi , aˆj . (6.32b)

h †i In order to see if above statement is valid we need to compute aˆ3, aˆ3 (where we h †i h †i expect result=1), aˆ4, aˆ4 (=1) and aˆ3, aˆ4 (=0)

h †i h ∗ †i 2 h †i 2 aˆ3, aˆ3 = iraˆ1, −ir aˆ1 = |r| aˆ1, aˆ1 = |r| 6= 1 (6.33a)

65 Quantum Mechanics of Beam Splitters

h †i h ∗ †i 2 h †i 2 aˆ4, aˆ4 = taˆ1, t aˆ1 = |t| aˆ1, aˆ1 = |t| 6= 1 (6.33b)

h †i h ∗ †i ∗ h †i ∗ aˆ3, aˆ4 = iraˆ1, −ir aˆ1 = irt aˆ1, aˆ1 = irt 6= 0 (6.33c)

Why are this results different from our expectations? In order to understand why the results are not the expected ones let us consider one little (subtle) detail (see Fig. 6.3):

Figure 6.3: Comparison classical vs quantum BS

As we see from Fig. 6.3 we need to account for the presence of the quantum vacuum in the second arm of the BS. We indicate this by assigning an operator aˆ0 on the other input port of the BS. In this case Eq. (6.31) becomes ! ! ! aˆ4 t ir aˆ1 = , (6.34) aˆ3 ir t aˆ0

h †i if we now calculate for example aˆ3, aˆ3 , we get

h †i h ∗ † ∗ †i aˆ3, aˆ3 = iraˆ1 + taˆ0, −ir aˆ1 + t aˆ0

2 h †i ∗ h †i ∗ h †i 2 h †i = |r| aˆ1, aˆ1 + irt aˆ1, aˆ0 − itr aˆ0, aˆ1 + |t| aˆ0, aˆ0 = |r|2 + |t|2 = 1, which is the expected result. The other commutator follows analogously.

66 Quantum Mechanics of Beam Splitters

6.4 Effects of Vacuum in BS: Coherent State Input

We now want to ask ourselves if taking into account the presence of the vacuum state in the unused port of the BS has some sort of consequence or not. To this aim, let us first consider the case of a coherent state as input, and assume, for simplicity, to √ consider a 50% BS, i.e., t = r = 1/ 2. The input state is then given by (remember displacement operator, see 3.52):

ˆ αaˆ†−α∗aˆ |ini = |αi1 |0i2 = D (α) |0i1 |0i2 = e |0i1 |0i2 . (6.35)

And, according to Eq. (6.34) ! ! ! aˆ 1 1 i aˆ 4 = √ 1 . aˆ3 2 i 1 aˆ2

Since we want to calculate how the input state gets transformed by the BS, it is more useful to express the input operators aˆ1 and aˆ2 as a function of the output operators (ˆa3, aˆ4). This is obtained by inverting the above equation, thus obtaining ! ! ! aˆ1 1 1 −i aˆ4 = √ , (6.36) aˆ2 2 −i 1 aˆ3 we then have that the input state |ini tronasforms to the following output state:

α † † α∗ √ aˆ +iaˆ − √ (ˆa4−iaˆ3) ˆ 0 2 ( 4 3) 2 |outi = D (α) |0i3 |0i4 = e |0i3 |0i4 , (6.37)

Moreover, since aˆ4 and aˆ3 commute

" † † # " † † # α † † α∗ ∗ ∗ √ (aˆ +iaˆ )− √ (ˆa4−iaˆ3) iαaˆ3 iα aˆ3 αaˆ4 α aˆ4 e 2 4 3 2 = exp √ − √ exp √ − √ (6.38) 2 2 2 2     ˆ iα ˆ α = D3 √ D4 √ . (6.39) 2 2

As a result we have that the input coherent state is split (50%) into two coherent states. The average number of photons per output port is therefore 50% of the initial one, since coherent states contain the quantum noise, this corresponds to the classical case

67 Quantum Mechanics of Beam Splitters

6.5 Effects of Vacuum in BS: Single Photon Input

We now consider the case of a single photon in a Fock state impinging on a 50% BS. r = t = √1 Again, 2 , and we have

1   |1i |0i =a ˆ† |0i |0i −−→BS √ aˆ† + iaˆ† |0i |0i 1 2 1 1 2 2 4 3 3 4 1 = √ (i |1i |0i + |0i |1i ) . 2 3 4 3 4

We have therefore: 1 |1i |0i → √ (i |1i |0i + |0i |1i ) . (6.40) 1 2 2 3 4 3 4

This means that the output state of the BS is, because of the impact of the vacuum state in the second port of the input, an entangled state, where we can not tell anymore, where the photon is.

6.6 Hong-Ou-Mandel Effect [9]

Consider a 50% BS with two photons as input, one single photon from the field in first arm and one single photon from the vacuum in second port. What will be our output? Let us have a look at the possibilities how two photons would behave the classical case:

Figure 6.4: Classical prediction of the behaviour of two photons in a BS

In order to check if our expectations are correct, let us do some QM calculations.

68 Quantum Mechanics of Beam Splitters

First of all let us rewrite the input operators as a function of the output ones, namely ! ! ! aˆ 1 1 −i aˆ 1 = √ 4 aˆ2 2 −i 1 aˆ3, and similarly, for the hermite conjugate ! ! ! aˆ† 1 1 i aˆ† 1 = √ 4 . † † (6.41) aˆ2 2 i 1 aˆ3

Since the incoming field is given by:

† † |ini = |1i1 |1i2 =a ˆ1aˆ2 |0i , (6.42) the outgoing field is then given by

 1    1   |outi = √ aˆ† + iaˆ† √ iaˆ† +a ˆ† |0i 2 4 3 2 4 3 1   2  2 = i aˆ† +a ˆ†aˆ† − aˆ†aˆ† + i aˆ† |0i 2 4 4 3 3 4 3  2 2   †  † 1 aˆ4 aˆ3 = √ i √ + i √  |0i 2  2 2 

1 = √ (|2i |0i + |0i |2i ) . (6.43) 2 4 3 4 3

Figure 6.5: Schematic sketch of a HOM interferometer

Fig. 6.5 shows schematically how a HOM interferometer works. Here a nonlinear process (PDC-cell) is used to generate single photon states. The BS works as seen

69 Quantum Theory of Photodetection above. If we slightly move the BS, we can introduce a delay between the two counts. The detected signal is illustrated in Fig. 6.6.

Figure 6.6: Coincidence rate detected by a HOM interferometer depending on the delay between the photons

The coincidence rate can be written as follows

 2 2 Rcoincidence ≈ 1 − exp −∆ω (τs − τi) , (6.44) where −∆ω2 is the bandwidth of the photons (non-monochromatic signal/idler) and

τs−τi is proportional to the distance difference that the two photons travelled. If τs−

τi  τcorr we obtain a maximum, i.e. the photons are not impinging simultaneously on the BS, which shows classical behaviour (simultaneous clicks possible). But for τs − τi = 0 we obtain that Rcoincidence = 0, i.e. the photons are impinging simultaneously on the BS, therefore we observe bunching (no simultaneous click possible) and hence, the coincidence is zero, as can be seen.

7 Quantum Theory of Photodetection

7.1 Direct Detection

Number of photons that arrive at the detector during the integration time t is given by t+τ Z Mˆ (t, τ) = dt0aˆ† (t0)a ˆ (t0) (7.1) t just like the time averaged classical intensity (average over integration time). Let us assume, that the light is nearly monochromatic and that T < ∞. This is the ideal

70 Quantum Theory of Photodetection case. In reality we have no detector with a 100% conversion efficiency (photons may fail to trigger ionization events, dead time masks subsequent ionization events). Dead time is the time that the photo-tube needs to recover, therefore we have a certain quantum efficiency η < 1. Therefore we can model a real (imperfect/ inefficient) photo-detector as shown in fig. 7.1.

Figure 7.1: Schematic sketch of the model of a real detector

Fig. 7.1 shows the actual model of a real detector. The BS models the inefficiency and vˆ (t) accounts for the quantum noise. The transmittance and reflectivity of the BS are given by: √ T = η (7.2a) p R = i 1 − η (7.2b)

• The output on arm 3 is lost (detector inefficiency)

• the output on arm 4 enters a perfect detector

• vˆ (t) is a continuous vacuum

We have: √ p dˆ(t) = ηaˆ† (t) + i 1 − ηvˆ (t) (7.3)

71 Quantum Theory of Photodetection

The photo-count operator Mˆ is now defined as

t+τ Z ˆ 0 ˆ† 0 ˆ 0 MD (t, τ) = dt d (t ) d (t ) . (7.4) t

Mean Photocount:

* t+τ + D E Z ˆ 0 ˆ† 0 ˆ 0 hmi ≡ MD (t, τ) = dt d (t ) d (t ) t * t+τ + Z h√ i h√ i = dt0 ηaˆ† (t0) − ip1 − ηvˆ† (t0) ηaˆ (t0) + ip1 − ηvˆ (t0)

t * t+τ + D E Z  ˆ p 0 † 0 0 = η M (t, τ) + i η (1 − η) dt aˆ (t)v ˆ (t ) t * t+τ + * t+τ + Z  Z  p 0 † 0 0 0 † 0  0 − i η (1 − η) dt vˆ (t )a ˆ (t ) + (1 − η) dt vˆ (t )v ˆ (t ) t t and therefore: D E hmi ≡ η Mˆ (t, τ) (7.5)

Second Moment:

t+τ t+τ Z Z D E hm (m − 1)i = dt0 dt00 dˆ† (t0) dˆ† (t00) dˆ(t00) dˆ(t0) = ...

t t D h iE = η2 Mˆ (t, τ) Mˆ (t, τ) − 1 and therefore D h iE hm (m − 1)i = η2 Mˆ (t, τ) Mˆ (t, τ) − 1 (7.6)

We can calculate now the variance photocount:

2 D 2E D E h∆mi = η2 ∆Mˆ (t, τ) + η (1 − η) Mˆ (t, τ) (7.7)

The first term of the variance photocount expresses the variance of the integrated photon number of the light beam. The second term expresses the partition noise, which is based on a random selection of a fraction η of the incident photons by the imperfect photo-detector. Note: with direct detection we are only able to retrieve properties pertaining to the incident intensity. No phase information can be retrieved.

72 Quantum Theory of Photodetection

7.2 Homodyne Detection

Homodyne detection is a method for measuring the electric field operator (actu- ally its quadratic operator) and therefore allows phase measurement (particularly important for squeezed light because its nonclassical properties are encoded in the phase).

Figure 7.2: Sketch of a balanced homodyne detector

50% |R| = |T | = √1 φ − φ = π We have a BS, i.e. 2 and R T 2 . The signal that we want to measure is in arm 1 of the BS. If we place photodetectors in both output arms, then we speak of balanced homodyne detection. We measure the difference of the two output signals, during the integration time. This corresponds to:

t+τ Z h i ˆ 0 † 0 0 † 0 0 M− (t, τ) = dt aˆ3 (t )a ˆ3 (t ) − aˆ4 (t )a ˆ4 (t ) (7.8) t in terms of the output signal this means that ! ! ! aˆ4 1 1 −i aˆ = √ (7.9) aˆ3 2 −i 1 aˆL where:

• 1 nh i o aˆ† (t0)a ˆ (t0) = iaˆ† (t0) +a ˆ† (t0) [−iaˆ (t0) +a ˆ (t0)] 3 3 2 L L  1 n 0 0 0 0 † 0 0 † 0 0 o = aˆ†(t)a ˆ (t ) + iaˆ† (t )a ˆ (t ) − iaˆ (t )a ˆ (t ) + aˆ (t )a ˆ (t ) 2  L L L L

73 Quantum Theory of Photodetection

• 1 nh i o aˆ† (t0)a ˆ (t0) = aˆ† (t0) + iaˆ† (t0) [ˆa (t0) − iaˆ (t0)] 4 4 2 L L  1 n 0 0 0 0 † 0 0 † 0 0 o = aˆ†(t)a ˆ (t ) − iaˆ† (t )a ˆ (t ) + iaˆ (t )a ˆ (t ) + aˆ (t )a ˆ (t ) 2  L L L L

Therefore: t+τ Z h i ˆ 0 † 0 0 † 0 0 M− (t, τ) = i dt aˆ (t )a ˆL (t ) − aˆL (t )a ˆ (t ) (7.10) t Although the detector measures photocount, the presence of the local oscillator makes it possible to actually measure the field! Note: If we use an inefficient detector:

t+τ Z h i ˆ 0 ˆ† 0 ˆ 0 ˆ† 0 ˆ 0 MH (t, τ) = dt d3 (t ) d3 (t ) − d4 (t ) d4 (t ) . (7.11) t

Therefore the mean homodyne difference photocount is given as

t+τ D E Z D E ˆ 0 † 0 0 † 0 0 hm−i = MH (t, τ) = iη dt aˆ (t )a ˆL (t ) − aˆL (t )a ˆ (t ) (7.12) t and the variance is given as

t+τ D E Z D E 2 2 ˆ 2 0 † 0 0 † 0 0 h∆m−i = η ∆ MH (t, τ) + η (1 − η) dt aˆL (t )a ˆL (t ) +a ˆ (t )a ˆ (t ) . t (7.13) Let us take a specific example: Local oscillator: Single-mode coherent state:

p −iωLt+iϑL αL (t) = FLe (7.14) where ωL is chosen in such a way, that ωSignal = ωL (homodyne). Since aˆ (t) |αi = α (t) |αi we obtain:

t+τ Z p 0 † −iωLt+iϑL iωLt−iϑL hm−i = iη FL dt aˆ (t) e − aˆ (t) e (7.15) t

Let us define:

t+τ 1 Z ˆ 0  † −iωLt+iϑL iωLt−iϑL  EH (χ, t, τ) = √ dt aˆ (t) e − aˆ (t) e (7.16) 2 τ t

74 Quantum Theory of Photodetection

π the homodyne electric field operator, where χ = ϑL + 2 , so that:

p D ˆ E hm−i = 2η FLT EH (χ, t, τ) (7.17a)

2 n D ˆ 2E o h∆m−i = ηFLT 4η ∆EH (χ, t, τ) + 1 − η (7.17b) ˆ And since substantially E ∼ xˆ1 we can conclude that with the homodyne detection we are able to measure the phase space of the field and therefore its phase properties.

Example 1 √ Input: single mode coherent state α (t) = F e−iωLt+iϑL and in this case: √ D ˆ E S ≡ EH (χ, t, τ) = FT cos (χ − ϑ) (7.18) and D ˆ 2E 1 N ≡ ∆EH (χ, t, τ) = (7.19) 4 With this we obtain the signal to noise ratio: D E ˆ2 2 EH S 2 SNR = = D E = 4FT cos (χ − ϑ) (7.20) N ˆ2 ∆EH

For an inefficient detector: SNRH = ηSNR

Example 2

Input: Squeezed state ˆ D ˆ E With EH ∼ xˆ1 follows that EH = 0 for a squeezed state and therefore hm− = 0i. For continuous squeezed state we have that:

0 2 0 aˆ† (ω)a ˆ (ω ) = sinh [s (ω)] δ (ω − ω ) (7.21) with the photon flux 1 Z dω sinh2 [s (ω)] = F (7.22) 2π and 0 1 iϑ(ω) 0 haˆ (ω)a ˆ (ω )i = − e sinh [2s (ω)] δ (ω + ω − ωp) (7.23) 2 Note: Here we used the following representation of a continuous coherent state:

ϕ (ω) = s (ω) eiϑ(ω) therefore:

75 Degree of Coherence and Quantum Optics

• we need to Fourier transform the operators because the properties of squeezed light are best caught in the frequency domain

1 • Assumption: T long enough that s (ω) and ϑ (ω) vary insignificantly over T

After some calculations follows D E 1   1   1  ∆Eˆ (χ, t, τ)2 = e2s(ωL) sin2 χ − ϑ (ω ) + e−2s(ωL) cos2 χ − ϑ (ω ) H 4 2 L 2 L (7.24) and we can show that D ˆ 2E 1 0 ≤ ∆EH (χ, t, τ) ≤ (7.25) 4 and therefore 2 0 ≤ (∆m−) < ηF2T, (7.26) where F2  F . This means that this noise derives only by the beating of the strong coherent field (of the local oscillator) with the noise in the signal. Note: Shot noise of the local oscillator automatically cancels out if we use balanced homodyne detection.

8 Degree of Coherence and Quantum Optics

8.1 First Order Degree of Coherence

Figure 8.1: Representation of a Mach-Zehnder interferometer (taken from [1])

zk • tk = t − c already contains eventual delay coming from different path lengths

• BS → (r, t)

Intensity on output E4 (t) averaged over an oscillation cycle:

1 2 2  2 2 ∗ I4 = ε0c |r| |t| |E1 (t1)| + |E2 (t2)| + 2< [E (t1) E2 (t2)] (8.1) 2 1

76 Degree of Coherence and Quantum Optics

where E1,2 are the fields entering the second BS (E1 → path 1, E2 → path 2). We moreover need to average the result over the detector integration time T (normally assumed to be much larger than the source’s coherence time). DEF:

T 1 Z hE (t)i = dt E (t) (8.2) T 0

The intensity now becomes:

1 2 2  2 2 ∗ hI4 (t)i = ε0c |r| |t| |E1 (t1)| + |E2 (t2)| + 2< [hE (t1) E2 (t2)i] (8.3) 2 1 the third term is responsible for interference and it is related to the first- order degree of coherence. First-Order Degree Of Coherence

hE∗ (t) E (t + τ)i g(1) (τ) = (8.4) hE∗ (t) E (t)i

(1) (1) z1−z2 with g (−τ) = g (τ) and τ = c . It is basically related to the fringe visibility. By collecting hI (t)i = hE∗ (t) E (t)i we have:

1 2   (1)  hI4 (t)i = ε0c |rt| hI (t)i 1 + < g (τ) (8.5) 2

Example: Collision Broadened Source

We consider a source that emits radiation characterized by random phase jumps due to the collisions between the atoms of the source (see fig. 8.2).

Figure 8.2: Phase jumps caused by collisions (taken from [1])

In this case even if each atom emits only the same frequency, the phase of the emitted field is different (assumed to be random). The total emitted electric field is

77 Degree of Coherence and Quantum Optics given by −iω0t X −iϕk(t) E (t) = E0e e (8.6) k Then if we calculate g(1) (τ) with this definition by recalling that the correlation i[ϕ (t+τ)−ϕ (t)] function that we gat for the phase e k k is proportional to the probability that the atom has a period of free flight longer than τ given by:

1  τ  p (τ) = exp (8.7) τ0 τ0 with τ0 ≡ τc being the source’s coherent time. We get at the end   (1) |τ| gCB (τ) = exp −iω0τ − (8.8) τc

Figure 8.3: Degrees of first order coherence. Lorentzian shape: collision broadened source. Gaussian shape: Doppler broadened source

Example: Doppler Broadened Source

In this case all atoms emit coherently but the emission frequency is slightly different due to Doppler broadening, therefore

X E (t) = E0 exp [−iωkt + iϕk (t)] (8.9) k

78 Degree of Coherence and Quantum Optics

noting, that ϕk for k 6= j averages to zero (randomly distributed), after some calcu- lations we arrive at the final result:

"  2# (1) π τ gDB (τ) = exp −iω0τ − (8.10) 2 τc

Message:

• g(1) (0) = 1  1 fully coherent   • g(1) (τ) = 0 fully incoherent  ∈ ]0, 1[ partially coherent

8.2 Quantum Theory of g(1) (τ)

Note: We consider single mode fields The field operator is given by

+ − i  −i(ωt−kz) i(ωt−kz) Eˆ (z, t) = Eˆ (z, t) + Eˆ (z, t) = aeˆ − aˆ†e (8.11) 2

(1) we can generalize g (τ) for two fields at different space-time points (z1, t1) and

(z2, t2) write it as:

∗ (1) hE (z1, t1) E (z2, t2)i g (z1, t1; z2, t2) = . (8.12) q 2 2 |E (z1, t1)| |E (z2, t2)|

If we want to write its quantum version, we just need to replace the fields with their operator versions.

Intensity operator: ˆ † † ˆ ˆ I ∼ aˆ aˆ → for the electric field this means aˆ ∼ E−,aˆ ∼ E+ and therefore Iˆ ∼ Eˆ−Eˆ+

Operator Ordering we always use normal ordering (aˆ† on the left and aˆ on the right). The reason we do this is because with this the vacuum energy becomes zero

we consider light beams that are stationary (fluctuations of the field do net vary in time and so g(1) (τ) does noet depend on τ)

Therefore: D E ˆ− ˆ+ ET (t) ET (t + τ) (1) g (τ) = D E (8.13) ˆ− ˆ+ ET (t) ET (t) and we can calculate the intensity as for the MZ case.

79 Degree of Coherence and Quantum Optics

Single Mode Field

Substitute Eˆ± and calculate

g(1) (τ) = ei(χ1−χ2) (8.14) with χ = ωt − kz Message: We are not really able to extract new information about the nature of a quantum state by simply looking at its first order degree of coherence. Therefore any single mode field is first order coherent for all pairs of space-time points.

8.3 Second Order Degree of Coherence

Instead of measuring correlations in the field we now measure intensity correlations. For simplicity we assume simple polarization and that the intensity readings are taken at a fixed pint in space and at a fixed pint in space and at a fixed delay τ.

I¯(t) I¯(t + τ) hE∗ (t) E∗ (t + τ) E (t) E (t + τ)i g(2) (τ) = = (8.15) I¯2 hE∗ (t) E (t)i2

Properties:

• g(2) (−τ) = g(2) (τ)

• g(2) (0) ≥ 1

Proof: Let us consider two intensity measurements one at t1 and the other one at t2, so that the Cauchy-Schwarz inequality is fulfilled by

¯ ¯ ¯ 2 ¯ 2 2I (t1) I (t2) ≤ I (t1) + I (t2) . (8.16)

Generalization: ( )2 1 X 1 X 2 I (tk) ≤ I (tk) (8.17) N N k k this means that for the statistical average we obtain:

2 2 2 I¯ = I¯(t) ≤ I¯(t) (8.18) and therefore with I¯(t) I¯(t) hI2i g(2) (0) = ≥ = 1 (8.19) I2 I2 we obtain g(2) (0) ≥ 1. q.e.d. In particular it can be demonstrated that there is no upper band

1 ≤ g(2) (0) ≤ ∞ (8.20)

80 Degree of Coherence and Quantum Optics with similar arrangements we can also prove that

0 ≤ g(2) (τ) ≤ ∞, τ 6= 0 (8.21a) and g(2) (τ) ≤ g(2) (0) (8.21b)

Figure 8.4: Degree of second order coherence for various light sources. Dashed line: coherent source. Lorentzian shape: collision broadened. Gaussian shape: doppler broadened (taken from [1])

How to measure g(2) (τ)? → Hanbury-Brown-Twiss Interferometer

Figure 8.5: Schematic illustration of the Hanbury-Brown-Twiss experimental setting (taken from [2])

81 Degree of Coherence and Quantum Optics

8.4 Quantum Theory of g(2) (τ)

Analogously to the case of g(1) (τ) we can easily quantize the expression of g(2) (τ) given above: D E ˆ− ˆ− ˆ+ ˆ+ ET (t) ET (t + τ) ET (t + τ) ET (t) g(2) (τ) = . D E2 (8.22) ˆ− ˆ+ ET (t) ET (t)

For a single mode field, and for the case of zero delay, i.e., τ = 0, this reduces to the following expression † † (2) aˆ aˆ aˆaˆ g (0) = , (8.23) SM haˆ†aˆi2 which we call single-mode second order coherence function. It is now clear that g(2) (τ) will have different outcomes depending on the quantum state considered because aˆ†aˆ†aˆaˆ gives different results for different quantum states. Let us rewrite 2 g(2) (0) in a more easy way as a function of hnˆi and (∆n) as follows:

• aˆ†aˆ = hnˆi ≡ hni  • aˆ†aˆ†aˆaˆ = aˆ† aˆaˆ† − 1 aˆ = aˆ†aˆaˆ†aˆ − aˆ†aˆ = hn (n + 1)i

Therefore:

hn2i − hni hn2i − hni2 + hni2 − hni (∆n)2 − hni g(2) (0) = = = 1 + hni2 hni2 hni2 and so we obtain (∆n)2 − hni g(2) (0) = 1 + . (8.24) hni2 g(2) (0) is therefore able to distinguish different quantum states!

8.4.1 Number States

• hni = n

2 • (∆n) = 0 it follows (2) 1 g (τ) = 1 − , (8.25) N n where n ≥ 1. Note:

(2) 1 (2) • Here gN (0) = 1− n , which is in contrast to the classical case, where g (0) ≥ 1!

• if |ni ≡ |0i then g(2) (0) is undefined.

82 Degree of Coherence and Quantum Optics

8.4.2 Coherent States

2 • hni = |α|

2 2 • (∆n) = |α| and so: (2) gC (0) = 1 (8.26) this gives a classical result (coherent states are the most classical states among the quantum states...).

8.4.3 Chaotic Light

We did not treat chaotic light during this course. Anyway it can be characterized by 2 2 a photon mean number hni and a variance (∆n) = hni + hn2i and a distribution property hat is essentially the Bose-Einstein statistics

hnin pn = (8.27) (1 + hni)1+n

So we obtain X |ψi = pn |ni hn| . (8.28)

With this: g(2) (0) = 2 (8.29) it is also in accordance with the classical result (cfr. both collision/Doppler broad- ened).

8.4.4 Squeezed Vacuum

• hni = sinh2 (s)

2 • (∆n) = 2 hni (hni + 1) so that 1 g(2) (0) = 3 + (8.30) hni Note: When hni = 0 then g(2) (0) 6= 1, so it does not give the coherent state limit. Care must be taken in how the vacuum limit is approached.

8.5 Comparison and Nonclassical Light g(2) (τ) Classical: • 1 ≤ g(2) (0) ≤ ∞, τ = 0 • 0 ≤ g(2) (τ) ≤ ∞, τ 6= 0

83 Degree of Coherence and Quantum Optics g(2) (τ) Single Mode: 1 1 − ≤ g(2) (τ) ≤ ∞ hni for hni ≥ 1 and ∀τ ∈ R

If we compare these ranges, we can see that for the case of quantum (single mode) light we can access also the ... below ..., which is not possible in the classical case, where g(2) (0) ≥ 1, i.e. 1 1 − ≤ g(2) (0) ≤ 1 (8.31) hni for hni ≥ 1. Note: If light lies in this range, then it is called nonlcassical light.

Figure 8.6: Comparison second order coherence of different light forms (taken from [1])

Why are the classical and quantum g(2) (τ) different? Consider measuring intensity/photon number twice in succession:

I2 hn (n − 1)i this leaves the value of I unchanged for It appears explicit the fact that there the second measurement and therefore has been a first measurement followed I2 is to be averaged by a second one, i.e. the measurement interferes with the measured system. While the first measurement counts n photons, the second one (affected by the first one) counts only (n − 1) pho- tons.

84 Elements of Quantum Nonlinear Optics

9 Elements of Quantum Nonlinear Optics

In the first part of this chapter we will introduce the nonlinear susceptibility, which is responsible for nonlinear processes. Furthermore the quantum optical expressions for the electric and magnetic field operators in the presence of a medium will be introduced. We will start with the classic formulation of the Maxwell equations in presence of a medium and derive the wave equation. At this point we will in- troduce the nonlinear polarization of the medium which is caused by the nonlinear susceptibility. In the second part we will treat nonlinear effects like second harmonic generation (SHG) and parametric up/down conversion.

9.1 Nonlinear Wave Equation

In order to derive the wave equation w need to start with the Maxwell equations in a medium: ∂B (r, t) ∇ × E (r, t) = − , (9.1a) ∂t

∇ · D (r, t) = ρfree, (9.1b)

∇ · B (r, t) = 0, (9.1c) 1 ∂D (r, t) ∇ × H (r, t) = + j. (9.1d) c2 ∂t

We assume to consider a material for which the free charges ρfree = 0 and the current j = 0. Moreover, let us assume to consider a non-magnetic material, i.e.

B = µ0H. In this case we also have that (see e.g. [3])

D (r, t) = ε0E (r, t) + P (r, t) , (9.2) where P (r, t) is the polarization vector, which accounts for the effect of th material on the propagation of the electromagnetic field inside the medium. In the following we will drop the dependence of the fields, but we will keep in mind that all fields depend on r and t if not mentioned explicitly. We obtain the wave equation by applying the curl operator two times resulting in:

2 2 ∂ D ∇ × ∇ × E = ∇ (∇ · E) − ∇ E = −µ0 . (9.3) ∂t2

In general the first term on the r.h.s. of this equation, i.e. ∇ (∇ · E), is nonzero, because in a material ∇ · D = 0 but ∇ · E 6= 0. However, for practical cases in nonlinear optics we can treat this term as being very small, i.e. ∇ · E ≈ 0. Furthermore with Eq. (9.2) we obtain:

∂2D 1 ∂2E ∂2P µ0 = + µ0 . (9.4) ∂t2 c2 ∂t2 ∂t2

85 Elements of Quantum Nonlinear Optics

Substituting this result into Eq. (9.3) and keeping in mind that ∇ · E ≈ 0 leads to

 2  2 2 1 ∂ ∂ P ∇ − E = µ0 . (9.5) c2 ∂t2 ∂t2

For the case of a nonlinear medium the electric polarization P (z, t) consists of two contributions

P = PL + PNL, (9.6) where PL = ε0χ (ω) E is the linear part of the polarization, which is responsible for P (k) k the refractive index. PNL = k χ (ω) E is the nonlinear part of the polarization, which is related to anharmonic terms in the Lorentz model (see [11]), e.g.

(2) (2) χ (ω) → PNL = χ E · E.

As it can be seen, the second order nonlinear polarization is proportional to the product of the square of the field, with the proportionality constant being the second order nonlinear susceptibility χ(2)(ω). For third order processes, the nonlinear polarization can be analogously written as follows (3) (3) χ (ω) → PNL = χ E · E · E, where χ(3)(ω) is the third order nolinear susceptibility. Substituting these expressions in Eq. (9.5) leads to

2 2 ! 2 2 n (ω) ∂ ∂ PNL ∇ − E = µ0 , (9.7) c2 ∂t2 ∂t2 where we have defined the refractive index

2 n (ω) = 1 + χ (ω) . (9.8)

Normally the electric susceptibility χ (ω) ∈ C, and therefore

n (ω) = n1 (ω) + in2 (ω) , (9.9) where n1 (ω) describes dispersion and n2 (ω) describes absorption. For simplicity we assume n2 (ω) = 0. This is a good approximation as long as we are consider- ing a frequency range where the considered material appear transparten or quasi- transparent. To consider a simple enough system, let us assume an electromagnetic field which 2 ∂2 propagates along the z-axis, so that ∇ → ∂z2 . Moreover, let us take the positive

86 Elements of Quantum Nonlinear Optics frequency part of the Fourier expansion of the field E (z, t) in plane waves, i.e.

∞ Z E (z, t) = dω E˜ (z, ω) e{−i[ωt−k(ω)z]}, (9.10) 0 with k (ω) = k0n (ω). Analogously for the nonlinear polarization we obtain

∞ Z ˜ −iωt PNL (z, t) = dω PNL (z, ω) e . (9.11) 0

Here we took the positive frequency part instead of the full Fourier transform because this would provide an easier way to write the quantum counterparts of these fields, looking back at the electric field operator, which only contains the positive frequency part, i.e. Eˆ+ (z, t). ∂2 Applying ∂z2 on E (z, t) gives

2 2 ˜ ¨¨ ˜ ∂ E ∂ E (z,¨ ω) ∂E (z, ω) 2 ˜ = ¨¨ + 2ik (ω) − k (ω) E (z, ω) . (9.12) ∂z2 ¨ ∂z2 ∂z where we used the slow varying envelope approximation (SVEA), namely we as- sumed ∂2E ∂E  k . (9.13) ∂z2 ∂z Substituting (9.11) and (9.12) in (9.5) leads to the expression

˜ 2 2 ∂E 2 ˜ n (ω) ω ˜ 2 ˜ −ik(ω)z 2ik (ω) − k (ω) E − E = −µ0ω PNLe (9.14) ∂z c2 where 2 2 2 2 ω n (ω) k (ω) = k2n (ω) = (9.15) c2 was used. This can be written in the form ˜ ∂E iω ˜ −ik(ω)z = PNLe , (9.16) ∂z 2ε0n (ω) where we used that µ ω2 µ ω2 µ ω ω − 0 = i 0 · c = i 0 = i . 2ik (ω) 2ωn 2µ0ε0n 2ε0n

9.2 Quantization of the Electromagnetic Field in Dielectric Media

Since Eq. (9.16) is valid in a nonlinear medium, we need to find the correct quan- tization rule for the electromagnetic field in a dielectric medium in order to solve

87 Elements of Quantum Nonlinear Optics it. In the beginning of this course we have already treated the case of quantizing the e.m. field in free space. Furthermore, we already found an expression for the quantized e.m. field for the case of the presence of an atom. Here we have assumed that the atomic medium is sufficiently diluted so that the quantized field can retain its free-space form. For the case of the propagation of the electromagnetic field in a dielectric medium, it is not possible anymore to make the approximation of dilute gas (as in the case of a single atom interacting with the field), as the atoms in a dielectric medium are organized in a compacted crystalline structure. To find an easy way to approach the problem of field quantization in a dielectric, we therefore need to make some assumptions, in order to be able to use the same quantization method we used for the case of free space. First of all, assume that the optical nonlinearities of the medium do not significantly affect the quantization pro- cedure, and thus only the linear effects are included in the determination of the field operators. Furthermore we assume no absorption, i.e. n2 (ω) = 0. This restricts the analysis to the range of frequencies of which the medium appears transparent. The definitions of the field operator are then only valid in this frequency interval. The last assumption that we make, is that the nonlinear effects are not significant ef- fected by the beam configurations. This means we can employ 1D-continuous mode field operators. Therefore we will consider propagation in z-direction and the polar- ization of the e.m. field is such that Eˆ k x and Bˆ k y. The quantization procedure is essentially similar to the free-space quantization, i.e. first we calculate the field Hamiltonian H and express it as a sum of harmonic oscillators. We obtain from this calculations the following electric field operator:

∞ 1 Z Eˆ+ (z, t) = √ dω Eˆ+ (ω) exp {−i [ωt − k (ω) z]} , (9.17) 2π 0 where Eˆ+ (ω) is the Fourier transformed electric field operator and k (ω) is the k- vector in the medium, which accounts for medium dispersion through n (ω). This result can be written in terms of aˆ (ω) as follows:

∞ Z Eˆ+ (z, t) = i dω f (ω)a ˆ (ω) exp {−i [ωt − k (ω) z]} , (9.18) 0 where the normalization factor f (ω) is to be determined. The displacement operator Dˆ + (z, t) in a dispersive material is obtained by means of its classical definition in a dispersive (linear) medium. The classical displacement operator is given by

2 D (z, t) = ε0εRE (z, t) = ε0n E (z, t) , (9.19)

88 Elements of Quantum Nonlinear Optics from which we obtain the nonlinear form as follows

∞ Z ˆ + 2 D (z, t) = iε0 dω f (ω) n (ω) aˆ (ω) exp {−i [ωt − k (ω) z]} . (9.20) 0

From Maxwell’s equations we obtain the magnetic field operator as follows

∞ Z n (ω) Bˆ+ (z, t) = i dω f (ω) aˆ (ω) exp {−i [ωt − k (ω) z]} . (9.21) c 0

To determine f (ω) we make use of the quantum counterpart of the continuity equa- tion for the electromagnetic field. In the classical case the energy-continuity equation reads ∂ ∂ hW (z, t)i = hS (z, t)i (9.22) ∂t ∂z where hW (z, t)i is the time averaged energy density and hS (z, t)i the time aver- aged Poynting vector (see [3]). From which we obtain the similar quantum optical expression of the continuity equation given by

∂Wˆ (z, t) ∂Sˆ = − , (9.23) ∂t ∂z where the operators Wˆ and Sˆ are assumed to be normal-ordered.

Energy Density in a Medium

The energy density is given by

1 W (z, t) = E · D + B · H = E (z, t) D (z, t) + B (z, t) B (z, t) . (9.24) µ0

Therefore we obtain the time derivative ∂W ∂D ∂E 1 ∂B 1 ∂B = E + D + B + B. (9.25) ∂t ∂t ∂t µ0 ∂t µ0 ∂t

Hence, we can write the time derivative of the as follows

∂Wˆ (z, t) ∂Dˆ + (z, t) ∂Dˆ − (z, t) = Eˆ− (z, t) + Eˆ+ (z, t) ∂t ∂t ∂t " # 1 ∂Bˆ+ (z, t) ∂Bˆ− (z, t) + Bˆ− (z, t) + Bˆ+ (z, t) . (9.26) µ0 ∂t ∂t

In the general case, it is not possible to retrieve an expression for Wˆ (z, t) from (9.26), because the r.h.s. is not a total time derivative and cannot be written like one. This is due to the presence of the medium. However, if we first insert the

89 Elements of Quantum Nonlinear Optics expressions of the field operators defined above and then look for an expression for Wˆ (z, t), then Eq. (9.26) can be integrated and leads to the following result:

∞ ∞ Z Z k (ω) − k (ω0) Wˆ (z, t) = ε c dω dω0 f (ω) f (ω0) [n (ω) + n (ω0)] 0 ω − ω0 0 0 0 0 × aˆ† (ω)a ˆ (ω0) ei{(ω−ω )t−[k(ω)−k(ω )]z}, (9.27) and we assume f (ω) ∈ R. Now we can obtain the explicit expression of the normal- ization factor f (ω) by imposing that the z-integrated energy density can be written as a collection of harmonic oscillators, i.e.,

∞ ∞ Z Z A dz Wˆ (z, t) = dω ~ωaˆ† (ω)a ˆ (ω) , (9.28) −∞ 0 where A is the space volume. Therefore we obtain s ~ω f (ω) = . (9.29) 4πε0cAn (ω)

Poynting Vector

Analogously to the classical formulation, we define the Poynting vector operator as follows n o ˆ 2 ˆ− ˆ+ ˆ− ˆ+ S (z, t) = ε0c E (z, t) B (z, t) + B (z, t) E (z, t) . (9.30)

Substituting the expression of the field operators gives

∞ ∞ s Z Z 0 ~ 0 ωω 0 Sˆ (z, t) = dω dω [n (ω) + n (ω )] × (9.31) 4πA n (ω) n (ω0) 0 0 0 0 ×aˆ† (ω)a ˆ (ω0) ei{(ω−ω )t−[k(ω)−k(ω )]z}. (9.32)

If we calculate the total energy that flows through a plane of constant z we have

∞ Z Z A dt Sˆ (z, t) = dω ~ωaˆ† (ω)a ˆ (ω) , (9.33) −∞ since the continuity equation must be valid. Moreover, the total energy flow over the entire length of the z-axis at a given time t is given instead by

∞ Z Z ˆ † A dz S (z, t) = dω ~ωva (ω)a ˆ (ω)a ˆ (ω) , (9.34) −∞

90 Elements of Quantum Nonlinear Optics

where va (ω) denotes the group velocity of the medium. It weights each frequency component of the energy flow. Note: For a quasi-monochromatic beam, i.e. a beam with narrow spectrum distri- bution centred at ω0, the Poynting vector is given as

ω ˆ ˆ− ˆ+ ~ 0 † S (z, t) = 2ε0cn (ω ) E (z, t) E (z, t) = aˆ (tR)a ˆ (tR) , (9.35) 0 A where tR = t − z/va (ω0) is the retarded time and va(ω) is the group velocity of the field inside the dielectric medium.

9.3 Second Harmonic Generation

In this part of our introduction we want to have a closer look at the second harmonic generation in a χ(2) medium. Therefore consider an incident, narrow-band signal at frequency ωp, as it is shown in Fig. 9.1. Furthermore, SHG means that two photons at the frequency ωp are simultaneously annihilated in order to create one photon at

2ωp.

(2) Figure 9.1: Schematic sketch of the SHG in a χ medium, where ωp is the frequency of the pump.

In what follows, we employ the so-called “non-depleted pump approximation”, which implies that the input (pump) beam, initially represented by a very bright coherent state, remains almost unchanged after the interaction with the nonlinear crystal. This means, for exmaple, that is the pump beam was originally containing 109 photons, the SHG interaction maybe converts 50 pump photons into SHG pho- tons, thus leaving 109 − 50 ' 109 photons in the pump beam after the process. The positive frequency part of the nonlinear polarization is then given by

∞ ∞ Z Z (+) ε Pˆ (z, t) = 0 dω0 dω00 χ(2) (ω0 + ω00) Eˆ(+) (ω0) Eˆ(+) (ω00) (9.36) NL 2π 0 0 × exp {−i [(ω0 + ω00) t − [k (ω0) + k (ω00)] z]} , (9.37)

91 Elements of Quantum Nonlinear Optics where ω0 and ω00 are the frequencies of the incident light. This nonlinear polarization acts as a source for the second-harmonic signal, according to

˜ˆ ∂E (z, ω) iω ˜ˆ −ik(ω)z = PNL (z, ω) e . (9.38) ∂z 2ε0n (ω)

Substituting (9.18) in the expression for the nonlinear polarization leads to

∞ ∞ s Z Z ω0ω00 Pˆ(+) (z, t) = − ~ dω0 dω00 χ(2) (ω0 + ω00)a ˆ (ω0)a ˆ (ω00) NL 4πcA n (ω0) n (ω00) 0 0 × exp {−i [(ω0 + ω00) t − [k (ω0) + k (ω00)] z]} , (9.39) where A is the space volume. The component of the nonlinear polarization oscillating at the frequency ω is then obtained by Fourier transform of the expression above, according to

∞ Z 1 ˜ −iωt PNL (z, t) = √ dω PNL (z, ω) e . (9.40) 2π 0

Comparing the expressions above leads to

∞ ∞ s Z Z 0 00 ˜ˆ ~ 0 00 ω ω (2) 0 00 0 00 PNL (z, ω) = −√ dω dω χ (ω + ω )a ˆ (ω )a ˆ (ω ) 2πcA n (ω0) n (ω00) 0 0 × exp {i [k (ω0) + k (ω00)] z} δ (ω − ω0 − ω00) . (9.41)

ˆ With this we can solve Eq.(9.16) with respect to E˜(+) (z, ω)

L Z ˜ˆ(+) iω ˜ˆ(+) −ik(ω)z E (L, ω) = dz PNL (z, ω) e . (9.42) 2ε0n (ω) 0

Using the Fourier relation we can then reconstruct the time-dependent electric field operator, i.e.

∞ L Z Z ˆ(+) i ω ˆ(+) −i[ωt−k(ω)z] E (L, t) = √ dω dz PNL (z, ω) e . (9.43) 2π2ε0c n (ω) 0 0

We can now calculate the intensity of the second harmonic signal as follows:

D E D E ˆ 1 ˆ(−) ˆ(+) I (L, t) = ε0cn (ω) E (L, t) E (L, t) . (9.44) 2

92 Elements of Quantum Nonlinear Optics

In order to get the full expression, let us first calculate Eˆ(+) (L, t)

∞ L Z Z ˆ(+) i ω ˆ(+) −i[ωt+k(ω)z] E (L, t) = √ dω dz PNL (z, ω) e 2π2ε0c n (ω) 0 0 L ∞ ∞ ∞ Z Z Z Z i~ 0 00 ω −i[ωt+k(ω)z] = − 2 dz dω dω dω e 4πε0c A n (ω) 0 0 0 0 s 0 00 ω ω 0 00 × χ(2) (ω0 + ω00) ei[k(ω )+k(ω )]zaˆ (ω0)a ˆ (ω00) n (ω0) n (ω00) × δ (ω − ω0 − ω00) L ∞ ∞ Z Z Z 0 00 i~ 0 00 ω + ω −i[(ω0+ω00)t+k(ω0+ω00)z] = − 2 dz dω dω 0 00 e 4πε0c A n (ω + ω ) 0 0 0 s 0 00 ω ω 0 00 × χ(2) (ω0 + ω00) ei[k(ω )+k(ω )]zaˆ (ω0)a ˆ (ω00) n (ω0) n (ω00)

If we assume that the pump beam can be considered having a narrow-band spectrum centered around the frequency ωp (quasi-monochromatic approximation), then we 0 00 can set ω ' ωp and ω ' ωp in the quantities appearing inside the integral above, except for the operators and time-dependent exponential. By doing so we obtain the following result:

L  i  Z 2ω ˆ(+) ~ p −i[k(2ωp)+2k(ωp)]z E (L, t) = − 2 dz e 8πε0c A n (2ωp) 0 ∞ ∞ Z Z ωp (2) 0 0 −iω0t 00 00 −iω00t × χ (2ωp) dω aˆ (ω ) e dω aˆ (ω ) e . (9.45) n (ωp) 0 0

By noting that the integrals in ω0 and ω00 are esesntially the Fourier transform of the time dependent operators aˆ(t), we can then rewrite the previous equation as follows

L  i ω2  1 Z ˆ(+) ~ p (2) −i[k(2ωp)+2k(ωp)]z E (L, t) = − 2 χ (2ωp)a ˆ (t)a ˆ (t) dz e . 2πε0c A n (2ωp) n (ωp) 0 (9.46) We obtain a equivalent result for Eˆ(+) (L, t), which substituted in the equation for

93 Elements of Quantum Nonlinear Optics the intensity of the second harmonic reads as follows: D E D E ˆ ˆ− ˆ+ ISHG (L, t) = 2ε0cn (2ωp) E (L, t) E (L, t)  2   2  i~ωp i~ωp = 2ε0cn (2ωp) − 2 2 4πε0c A 4πε0c A  2 (2) 2 1 × χ (2ωp) n (2ωp) n (ωp) L 2 Z −i[k(2ωp)+2k(ωp)]z × dze aˆ† (t)a ˆ† (t)a ˆ (t)a ˆ (t) . (9.47)

0

Defining ∆k = k (2ωp) − 2k (ωp) we have

L 2 L L Z Z Z −i∆kz i∆kz −i∆kζ dz e = dz e dζ e

0 0 0  i   −i  L L = e−i∆kz e−i∆kζ  ∆k ∆k 0 0 1 = e−i∆kL − 1 ei∆kL − 1 ∆k2  1 2 = 1 − e−i∆kL − ei∆kL + 1 ∆k 2 = [1 − cos (∆kL)] ∆k2 2  ∆kL = 1 − 1 + 2 sin2 ∆k2 2 sin2 (∆kL/2) = (∆kL/2)2

Summarizing everything leads to the expression for the time averaged second har- monic intensity " #  2ω4  (2) 2 2 Dˆ E ~ p |χ (2ωp)| sin (∆kL/2) ISHG (L, t) = 2 3 2 2 2 8π ε0c A n (2ωp) n (ωp) (∆kL/2) (2) † † × χ (2ωp) aˆ (t)a ˆ (t)a ˆ (t)a ˆ (t) . (9.48)

† † If we consider a continuous wave at the frequency ωp we can write aˆ (t)a ˆ (t)a ˆ (t)a ˆ (t) (2) as a second order correlation function gp,p (0), i.e.

† † (2) † 2 aˆ (t)a ˆ (t)a ˆ (t)a ˆ (t) = gp,p (0) aˆ (t)a ˆ (t) . (9.49)

The second harmonic signal is proportional to g(2) (0) and to the square of the mean photon flux of the incident pump beam. This is the case, because, as already men- tioned SHG is the process of simultaneous two photon absorption at the frequency

94 Elements of Quantum Nonlinear Optics

(2) ωp. Furthermore, g (τ) is a measure for two photon absorption. Therefore, if g(2) (τ) of the pump is zero, we will not obtain second harmonic generation, e.g. consider a single photon state |1i as pump. There is no second photon to absorb (2) (2) and therefore ISHG = 0 and also g (τ) = 0 for |1i, since g (0) = 1 − 1/n for Fock states. Moreover, for a thermal source we obtain

(2) (2) gT HERMAL (0) = 2 6= gCOH (0) , since the availability of photon pairs in thermal light is greater than in coherent (thermal) (coh) light. This leads to the fact, that ISHG /ISHG = 2. In anti-bunched light (e.g. Fock states), we have that:

(2) (2) gA.B. (0) < gCOH (0) , therefore, anti-bunched light will generate less SHG than coherent light.

Phase Matching

For phase matching we need to consider the L-dependent term of the intensity, i.e.

sin2 (∆kL/2) . (∆kL/2)2

If ∆k = 0 this term becomes one, and the second harmonic intensity ISHG reaches its maximum. This means that for ∆k = 0 we obtain

k (2ωp) = 2k (ωp) , (9.50) which is the phase matching condition. This corresponds to the conservation of momentum in passing from ωp to 2ωp.

95 Elements of Quantum Nonlinear Optics

9.4 Parametric Down Conversion

Figure 9.2: Schematic sketch of the quantum optical parametric down conversion in (2) a χ medium, where ωp is the frequency of the pump, ωi is the idler frequency and

ωs the signal frequency.

The nonlinear polarization is given by

∞ ∞ ε Z Z Pˆ(+) (z, t) = 0 dω0 dω00 χ(2) (ω00 − ω0) Eˆ(−) (z, ω0) Eˆ(+) (ω00) NL 2π 0 0 × exp {i [(ω0 − ω00) t − [k (ω0) − k (ω00)] z]} , (9.51) where ω0 and ω00 are the frequencies of the signal and the pump, respectively. If we take the Fourier transform of the nonlinear polarization and evaluate it at the idler frequency ωi = ωp − ω we have:

∞ ∞ Z Z ˜ˆ(+) ε0 0 00 (2) 00 0 ˆ(−) 0 ˆ(+) 00 P (z, ωp − ω) = √ dω dω χ (ω − ω ) E (z, ω ) E (ω ) NL 2π 0 0 −i[k(ω0)−k(ω00)]z 0 00 × e δ (ωp − ω + ω − ω ) . (9.52)

For our calculations we will assume that the pump is a bright coherent state, namely

00 00 aˆ (ω ) |αpi = αp (ω ) |αpi . (9.53)

Furthermore, we assume that we have quasi monochromatic light, i.e., we have a narrow band around ωp which we express as

00 p iϑp 00 αp (ω ) = 2πFpe δ (ω − ωp) , (9.54) where Fp is the mean photon flux and ϑp is the pump phase. The phase matching condition is given by

∆k = k (ωp) − k (ω) − k (ωp − ω) , (9.55)

96 Elements of Quantum Nonlinear Optics where we have for ∆k 6= 0 the non-degenerate and for ∆k = 0 the degenerate para- metric down conversion, respectively. Since signal and idler have different frequencies, we define two different field opera- tors (they are, in fact, two different field modes): s ˆ(+) ~ω E (z, ω) = i aˆz (ω) , (9.56a) 2ε0cAn (ω) as electric field operator of the signal and s ˆ(+) ~ (ωp − ω) ˆ E (z, ωp − ω) = i bz (ω) , (9.56b) 2ε0cAn (ωp − ω) is the electric field operator of the idler. They both obey (9.16), i.e.

ˆ(+) s ∂E (z, ωp − ω) ωp − ω ~ωpFp (2) ˆ(−) = 3 χ (ωp − ω) E (z, ω) ∂z n (ωp − ω) 8ε0c An (ωp)

× exp {i [ϑp + (k (ωp) − k (ω) − k (ωp − ω)) z]} , (9.57a) and

ˆ(−) s ∂E (z, ωp) ω ~ωpFp (2)∗ ˆ(+) = − 3 χ (ω) E (z, ωp − ω) ∂z n (ω) 8ε0c An (ωp)

× exp {−i [(k (ωp) − k (ω) − k (ωp − ω)) z]} . (9.57b)

This set of coupled wave equations can be written with respect to aˆ and ˆb. After introducing s iϑ(ω) ~ωpω (ωp − ω) Fp (2) iϑpL s (ω) e = − 3 χ (ωp − ω) e (9.58) 8ε0c An (ωp) n (ω) n (ωp − ω) we obtain ˆ ∂bz (ωp − ω) s (ω) = − eiϑ(ω)aˆ† (ω) , (9.59a) ∂z L z and ∂aˆz (ω) s (ω) −iϑ(ω)ˆ = − e bz (ωp − ω) . (9.59b) ∂z L whose solution is given by

† † ˆ −iϑ(ω) aˆL (ω) =a ˆ0 (ω) cosh [s (ω)] − b0 (ωp − ω) e sinh [s (ω)] , (9.60a) and

ˆ ˆ † iϑ(ω) bL (ωp − ω) = b0 (ωp − ω) cosh [s (ω)] − aˆ0 (ω) e sinh [s (ω)] , (9.60b)

97 Appendix A: The Harmonic Oscillator for the case of non-degenerate parametric down conversion. ˆ For the degenerate case we have the same theory as before, just with a ≡ ˆb. There- fore, the solution of the coupled equations becomes

† † iϑ(ω) aˆL (ω) =a ˆ0 (ω) cosh [s (ω)] − aˆ0 (ωp − ω) e sinh [s (ω)] , (9.61) which corresponds to a single mode squeezed state. This means, that PDC and degenerate PDC are natural sources of squeezed light (and they possess all the non-classical properties of squeezed light). Therefore we can write PDC |ψi −−−→|ξ (ω)i , (9.62) with ξ (ω) = s (ω) eiϑ(ω). (9.63)

10 Appendix A: The Harmonic Oscillator

The Hamiltonian of a harmonic oscillator is given by

p2 mω2 Hˆ = + x2. (10.1) 2m 2

To solve the problem, we introduce the so called creation (aˆ†) and annihilation (aˆ) operators, defined as follows: r ~  xˆ = aˆ +a ˆ† , (10.2a) 2mω r mω~  pˆ = i aˆ† − aˆ . (10.2b) 2 Substituting this into Eq. (10.1), we obtain the so-called second quantized form of the harmonic oscillator Hamiltonian, i.e.,

~ω  Hˆ = aˆ†aˆ +a ˆaˆ† (10.3) 2 with the commutation rules

  a,ˆ aˆ† =a ˆaˆ† − aˆ†aˆ = 1 (10.4a)

  [ˆa, aˆ] = 0 = aˆ†, aˆ† . (10.4b)

Using the commutation rule (10.4), the Hamiltonian of the harmonic oscillator can be rewritten as follows:  1 Hˆ = ω aˆ†aˆ + . (10.5) ~ 2

98 Appendix B: The Quantization of a Paraxial Electromagnetic Field

If |ni is the state of the harmonic oscillator with n quanta of energy on it, then the Hamiltonian operator written above acts on such a state in the following way:

Hˆ |ni = n~ω |ni . (10.6)

11 Appendix B: The Quantization of a Paraxial Elec- tromagnetic Field

The quantization procedure described in the first chapters of these notes has been carried for the electromagnetic field in free space, with the aid of a fictitious cavity, whose main aim was to introduce a suitable (and regular) set of modes upon which the general electromagnetic field could be expanded. In carrying out that quanti- zation procedure, however, no attention at all was put on the spatial properties of the field itself, as the main result of the quantization of the electromagnetic field was the introduction of Fock states, i.e. energy eigenstates associated to the field. The analysis carried out throughout all these notes, therefore, only focused on the energy-eigenstates description of the quantized field, and the spatial structure of the field itself has not been taken into account, assuming that the field could always be represented by a plane wave in space. In this last chapter, however, we intend to overcome this limitation, by discussing how can we quantize a paraxial electromagnetic field, and how can we introduce field operators, which are able to create, for example, a single photon in a given spatial mode (for example, in a Hermite-Gauss or Laguerre-Gauss mode). To do that, however, we need to reconsider the problem of field quantization, putting particular attention to the fact that the quantized field must be a solution of the paraxial equation, and therefore the corespondent field state must be defined in a suitably chosen Hilbert space, which guarantees the feasibility of the quantization procedure and at the same time ensures the field to be a solution of the paraxial equations. The problem of paraxial quantization was first solved by Deutsch and Garrison in 1991 [12] and can be found (in a slightly more didactical version) in Chapter 6 of Ref. [13]. However, the approach presented in Ref. [14] is much more clear and easier to adapt to the scope of these notes. Therefore, the procedure of quantization of a paraxial light beam will be explained in this chapter taking Ref. [14] as a main guidance. The results presented here, however, are in complete agreement with the work by Deutsch and Garrison presented in Ref. [12].

11.1 From the Vector Potential to the Paraxial Operator

As a starting point for our analysis, let us consider the expression of the positive fre- quency part of the quantized vector potential in the Coulomb gauge, for a continuous

99 Appendix B: The Quantization of a Paraxial Electromagnetic Field mode field, i.e.,

2 s X Z ˆ (+) 3 ~ ˆ(λ) i(k·r−c|k|t) A (r, t) = d k 3 e (k)ˆaλ(k)e . (11.1) 16π cε0|k| λ=1

To correctly carry out the quantization procedure for a paraxial beam, we use a trick similar to the one used for the general case in chapter 1, i.e., we introduce a longi- tudinal quantization length L, so that the longitudinal component kz of the field’s wave vector can be discretized as follows: ζn = (2π/L)n, where n ≥ 0 selects only forward propagating waves. By doing this, the kz-integration in the above equation reduces to a summation over the allowed values of n, and only the integration over the transverse momentum kT = kxxˆ + kyyˆ remains. Therefore

Z 2π X Z d3k → d2k , L T n where the prefecture 2π/L has been inserted to correctly account for normalization. Together with this substitution, we must also redefine de annihilation operator ap- pearing in Eq. (11.1) as follows r L aˆ (k) → aˆ (k, n), λ 2π λ in such a way that the bosonic commutation relation

 † 0  0 aˆλ(k, n), aˆµ(k , m) = δλµδnmδ(k − k ) (11.2) continues to be valid. After these transformations, Eq. (11.1) can be then rewritten as follows: s X Z ˆ (+) 2 ~ ˆ(λ) i[(kt+ζnzˆ)·r−c|k|t] A (r, t) = d kT 2 e (k)ˆaλ(kT, n)e . (11.3) 8π cε0|k|L λ,n

Our aim is to be able to write the expression of the vector potential operator in such a way that it could be interpreted as a solution of the paraxial wave equation. To do that, we shoudlrecall that a paraxial field can be always written as the product of an envelope function ψ(x, y, z), which varies slowly with z0 and a plane wave term exp [i(k0z − ω0t)], which accounts for the propagation along the z direction and the harmonic time dependence of the field itself [13]. A closer inspection to Eq. (11.3), however, reveals that we do not have such exponential factor. However, if we introduce the following, trivial identity

L Z L/2π ei(k0z−ω0t) 1 = dk0 , i(k0z−ω0t) 2π 0 e

100 Appendix B: The Quantization of a Paraxial Electromagnetic Field into Eq. (11.3), we can rewrite it in the following form:

L Z L/2π ˆ (+) −i(k0z−ω0t) ˆ A (r, t) = dk0e ψ(r, t), (11.4) 2π 0 where s X Z ˆ 2 ~ ˆ(λ) i[(k−k0zˆ)−(c|k|−ω0)t] ψ(r, t) = d kT 2 e (k)ˆaλ(kT, n)e , (11.5) 8π cε0|k|L λ,n and k = kT + ζnzˆ. Written in this form, Eq. (11.4) is an exact solution of the paraxial wave equa- tion, if the field operator ψˆ (r, t) is itself an exact solution of the paraxial equation, i.e.,   ∂ 2 ˆ 2ik0 + ∇ ψ(r, t) = 0. (11.6) ∂z T Substituting Eq. (11.5) into the above equation brings to the following dispersion relation: 2 ζn |kT | = 1 − 2 , (11.7) k0 2k0 which is the usual paraxial dispersion relation [15]. It is worth noticing that the above relation constitutes a constraint on the allowed valued of ζn that a paraxial wave can take, as a function of the transverse wave vector kT . This, in our case, constitutes a constraint that defines a two dimensional sub-space in the k-space, where Eq.(11.4) is at the same time an exact solution of the paraxial equation (because of the way we defined ψˆ (r, t)) and also an exact solution of the wave equation, a condition which is necessary for being able to write the quantized form of the vector potential as given by Eq. (11.1). This 2D sub-space is the paraxial Hilbert space introduced in Ref. [12]. In what follows, we will try to rewrite Eq. (11.4) in a form which automatically accounts for this constraint, thus making not necessary to introduce the paraxial Hilbert space explicitly, thus making the quantization procedure easier to understand.

11.2 The Paraxial Dispersion Relation

To implement a form of the paraxial operator ψˆ (r, t) which automatically accounts for the paraxial constraint, we first need to elaborate a bit more eq. (11.7). To do that, let us rename the transverse wave vector as q = kt and let us define q = |kT |. Moreover, let us introduce the following quantity √ q θ 2 = , (11.8) k0

101 Appendix B: The Quantization of a Paraxial Electromagnetic Field which is a generalization of the well known divergence angle of a Gaussian beam [16]. By then introducing the angle φ formed by the propagation direction zˆ and the field k-vector k, it is not difficult to prove that q q tan φ = = . k · zˆ ζn

Using this result and Eq.(11.8) we can rewrite the dispersion relation (11.7) in the following, dimensional form:

√ p θ 2 = − cot φ + 2 + cot2 φ. (11.9) √ The above equation is exact and links the beam divergence θ 2 with the angle formed by the beam with the propagation axis. For the paraxial case, φ  1, and a Taylor expansion of the above relation brings to the following result

√ φ3 θ 2 ' φ − + O(φ5). (11.10) 6

With this approximation, the dispersion relation (11.7) becomes

2 ζn ' k0(1 − θ ). (11.11)

We can use this result to write the k-vector and the polarization vectors appear- ing in Eq. (11.5) as θ-dependent quantities, in such a way that it becomes very easy to isolate their paraxial contribution. In fact, in analogy with paraxial gaus- sian beams [17], the purely paraxial terms are the one independent on θ, and the terms which depend from θ2 represent the lowest order corrections to the paraxial approximation. Using Eq. (11.10) we can then write the following, approximated, quantities:

2 k → k ' qqˆ + k0(1 − θ )zˆ, (11.12a) eˆ(2)(k) → eˆ(2)(q, θ) ' zˆ × qˆ, (11.12b) eˆ(1)(k) → eˆ(1)(q, θ) ' eˆ(1)(q, θ) × k. (11.12c)

Moreover, Eq. (11.11) needs to be compared with the longitudinal quantization condition ζn = (2π/L)n that we assumed at the beginning of our analysis. Since the allowed longitudinal k-vectors are only the ones which have the form ζn = (2π/L)n, the paraxial dispersion relation given by Eq. (11.11) forces the integer n to acquire a precise value, i.e.   k0L 2 n ≡ n(θ) = 1 − θ , (11.13) 2π I.P where the subscript I.P. stands for “Integer Part”. This result is exactly what we were looking for. Among all the possible value

102 Appendix B: The Quantization of a Paraxial Electromagnetic Field

of ζn deriving from the introduction of a longitudinal quantization length L, the constraint imposed by ψˆ (r, t) being an exact solution of the paraxial equation selects only one single value of ζn, namely the one corresponding to n = n(θ). If we now ˆ insert artificially a condition of the kind δn,n(θ) in the expression of ψ(r, t) given by Eq. (11.5), we therefore have a field operator which automatically accounts for the paraxial constraint (11.10). The final expression of the paraxial field operator ψˆ (r, t) is then given by s X Z ˆ 2 ~ ˆ(λ) i[(k−k0zˆ)−(c|k|−ω0)t] ψ(r, t) d kT 2 e (k)ˆaλ(kT, n)δn,n(θ)e . (11.14) 8π cε0|k|L λ,n

11.3 Back to Continuous Mode Operators

Now that we have obtained our desired result of writing the expression of a field operator, which is automatically defined (through the paraxial constraint δn,n(θ)) in a suitable paraxial Hilbert space, we can restore the continuous character of the longitudinal wave vector kz. We have in fact disctretized the longitudinal component of the wave vector only for convenience, in order to deal with well defined quantities (i.e., in order to avoid singularities and problems with infinities).

If we then make the following substitutions, where now kz has been replaced with cω:

X L Z → dω, 2πc n r2πc aˆλ(q, n) → aˆλ(q, ω), (11.15) L and we take the continuos limit of δn,n(θ) as

2πc 2 δn,n(θ) → δ(ω − ck0(1 − θ )), (11.16) L we can rewrite the expression of the vector potential operator (11.1) as follows:

Z r ˆ (+) X −iω(t−z/c) ~ A (r, t) = dω e 3 16π ε0cω λ Z 2 (λ) i[q·x−q2cz/(2ω)] × d q E (q, ω, z, t)ˆaλ(q, ω)e , (11.17) where x = xxˆ + yyˆ and

 2 1/2 √ 1 + θ 4 E (λ)(q, ω, z, t) = eˆ(λ)(q, θ) √ e−iωt( 1+θ −1). (11.18) 1 + θ4

One can easily prove that Eq. (11.17) represents a bona fide vector potential, which for t = 0 is an exact solution of the paraxial equation and for t ≥ 0 is an exact

103 Appendix B: The Quantization of a Paraxial Electromagnetic Field solution of the wave equation. The vector potential operator (11.17) is therefore defined in the paraxial Hilbert space defined by the paraxial constraint (11.10).

11.4 The Maxwell-Paraxial Modes

Equation (11.17) is already a good form of paraxial vector potential operator. How- ever, the annihilation operator that appears there is written in the frequency space, as it annihilates a photon with a specific transverse momentum q. To be able to manage the creation and annihilation of photons in spatial modes of the field, how- ever, we will need to rewrite Eq. (11.17) in such a way that it contains operators defined in the representation space spanned by x To do that, let us introduce the following Fourier transform for the operators aˆλ(q, ω) Z 1 2 iq·x aˆλ(x, ω) = d q aˆλ(q, ω)e . (11.19) 2π

It is left to the reader to prove that aˆλ(x, ω) obeys the bosonic commutation rule

 †  aˆλ(x, ω), aˆµ(y, Ω) = δλµδ(ω − Ω)δ(x − y). (11.20)

Substituting Eq. (11.19) into Eq. (11.17) and calling k0 = ω/c gives the following, final, result: Z i−ω(t−z/c) ˆ (+) X e ˆ(+) A (r, t) = dω p A (x, z, ω, t), (11.21) λ 4πε0cω/~ where Z ˆ(+) 2 0 (λ) 0 0 A (x, z, ω, t) = d x F (x , z, x, ω, t)ˆaλ(x , ω), (11.22) with F(λ)(x0, z, x, ω, t) are the so-called Maxwell-paraxial modes, whose explicit ex- pression is given by

1 Z 0 2 F(λ)(x0, z, x, ω, t) = d2q E (λ)(q, ω, z, t)ei[q·(x−x )−q cz/(2ω)] (11.23) (2π)2

Let us comment this result. First of all, notice that the Maxwell-paraxial modes F(λ)(x0, z, x, ω, t) represent the field at a plane z at time t generated by a point source located at x0 in the transverse plane z = 0, oscillating at the paraxial frequency ω. From this perspective, F(λ)(x0, z, x, ω, t) can be interpreted as the quantum coun- terpart of the Huygens-Fresnel diffraction integral. Moreover, if the paraxial ap- proximation ids valid, i.e, if the integration in Eq. (11.23) is limited to the domain

Cω = {q : q  ω/c}, the Maxwell-paraxial modes reduce to

2 Z d q 0 2 (λ) 0 ˆ(λ) i[q·(x−x )−q cz/(2ω)] P (x , z, x, ω, t) = 2 e (q)e , (11.24) Cω (2π) which is nothing but the paraxial propagator, i.e. the Fresnel diffraction integral.

104 Appendix B: The Quantization of a Paraxial Electromagnetic Field

Moreover, the Maxwell-paraxial modes are quasi-orthogonal, in the sense that Z 2 (λ) 0 (µ) 00 d x F (x , z, x, ω, t)F (x , z, x, ω, t) = δλµ

2  2 1/2 Z d q 1 + θ 0 00 × √ eiq·(x −x ). (11.25) (2π)2 1 + θ4

Notice that these modes become fully orthogonal in the paraxial regime.

11.5 The Paraxial States of the Electromagnetic Field

Let us now concentrate on Eq. (11.22) and its meaning. According to Eq. (11.22), the Fourier component at frequency ω of the paraxial vector potential operator Aˆ (+)(r, t) is given as the superposition of all the annihilation operators at the po- sition x0, weighted by the corresponding Maxwell-paraxial mode F(λ)(x0, z, x, ω, t). This means, from a very general point of view, that if we want to create a photon in a certain paraxial mode given by F(λ)(x0, z, x, ω, t) (or, equivalently, its paraxial ponterpart P(λ)(x0, z, x, ω, t)), we need to construct a field operator, which is given by the superposition of al the field operators that create a photon in the point of space x0, each of them weighted by the value of the Maxwell-paraxial mode at that spatial point. If we then introduce the following single photon paraxial state

† |x, ω, λi =a ˆλ(x, ω)|0i, (11.26) then h0|Aˆ (+)(r, t)|x, ω, λi ' F(λ)(x0, z, x, ω, t)e−iω(t−z/c), (11.27) and therefore the Maxwell-paraxial modes can be interpreted as the single photon wave function. The single photon state defined in Eq. (11.26) can also be used to find the spatial representation of any state of the electromagnetic field. Given in fact a general state |ψi of the electromagnetic state, the projection onto |x, ω, λi gives in fact

hx, ω, λ|ψi = ψλ(x, ω), (11.28) which is an exact solution of the paraxial equation.

11.6 An Example of Paraxial Mode Operators

As it is well known [16], the fundamental solution of the paraxial equation is given in the form of a Gaussian beam. The form of the higher order solutions, however, depend on the symmetry of the problem. For the case of cartesian symmetry, they are given by the Hermite-Gauss modes, while for cylindrical symmetry, they are given by Laguerre-Gauss modes [16]. In both cases, the modes are defined by two

105 REFERENCES integer indices n and m. Moreover, these modes represent an orthonormal and complete basis in cartesian and cylindrical coordinates, respectively. Without any loss of generality, then we can define ψnm(x, ω) to be one of such paraxial eigenmodes. Any more general paraxial field can be then represented as a suitable superposition of these modes. Let us then define the state of the field corresponding in having a single photon in one of these modes as follows: Z 2 |n, m, ω, λi = d x ψnm(x, ω)|x, ω, λi, (11.29) in such a way that

hy, Ω, µ|n, m, ω, λi = ψnm(y, ω)δλµδ(ω − Ω). (11.30)

Analogously to what we did in Eq. (11.26), we can define the mode operators as follows: † |n, m, ω, λi =a ˆnmλ(ω)|0i. (11.31)

It is then not difficult to prove that the following relations hold: Z 2 ∗ aˆnmλ(ω) = d x ψnm(x, ω)ˆaλ(x, ω), (11.32a) Z † 2 † aˆnmλ(ω) = d x ψnm(x, ω)ˆaλ(x, ω), (11.32b) for the mode operators and

X aˆλ(x, ω) = ψnm(x, ω)ˆanmλ(ω), (11.33a) n,m † X ∗ † aˆλ(x, ω) = ψnm(x, ω)ˆanmλ(ω), (11.33b) n,m (11.33c) for the correspondent spatial operators.

12 Bibliography

References

[1] R. Loudon, “The Quantum Theory of Light” , Third Edition (Oxford, 1997).

[2] C. Gerry and P. Knight, “Introductory Quantum Optics” (Cambridge, 2004).

[3] J. D. Jackson, “Classical Electrodynamics”, Third Edition (Wiley, 1998).

106 REFERENCES

[4] B. Mehmani and A. Aiello, Eur. J. Phys. 33, 1367 (2012).

[5] S. Gasiorowicz, “Quantum Physics", First Edition (Wiley, 1974).

[6] A. Fetter and J. Walecka, “Quantum Theory of Many-Particle Systems” (Dover, 2003).

[7] C. Kittel, “Quantum theory of solids” (Wiley, 1987).

[8] V. Mukhanov and S. Winitzki, “Introduction to Quantum Effects in Gravity” (Cambdrige, 2007).

[9] C. K. Hong, Z. Y. Ou, and L. Mandel, Phys. Rev. Lett. 59, 2044 (1987).

[10] H. Walther, B. T. H. Varcoe, B. G. Englert and T. Becker, Rep. Prog. Phys. 69, 1325 (2006).

[11] R. W. Boyd, “Nonlinear Optics”, Third Edition (Academic Press, 2008).

[12] I. H. Deutsch and J. C. Garrison, Phys. Rev. A 43, 2498 (1991).

[13] J. C. Garrison and R. Y. Chiao, “Quantum Optics" (Oxford, 2008).

[14] A. Aiello and J. P. Woerdman, Phys. Rev. A 72, 060101(R) (2005).

[15] L. Mandel and E. Wolf, “Optical Coherence and Quantum Optics” (Cambridge, 1995).

[16] O. Svelto, “Principles of Lasers”, Fifth Edition (Springer, 2010).

[17] M. Lax, W. H. Louisell, and W. B. McKnight, Phys. Rev. A 11, 1365 (1975).

107