Paper The following article is Open access

A master equation for strongly interacting dipoles

and

Published 13 April 2018 © 2018 The Author(s). Published by IOP Publishing Ltd on behalf of Deutsche Physikalische Gesellschaft
, , Citation Adam Stokes and Ahsan Nazir 2018 New J. Phys. 20 043022 DOI 10.1088/1367-2630/aab29d

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1367-2630/20/4/043022

Abstract

We consider a pair of dipoles such as Rydberg atoms for which direct electrostatic dipole–dipole interactions may be significantly larger than the coupling to transverse radiation. We derive a master equation using the Coulomb gauge, which naturally enables us to include the inter-dipole Coulomb energy within the system Hamiltonian rather than the interaction. In contrast, the standard master equation for a two-dipole system, which depends entirely on well-known gauge-invariant S-matrix elements, is usually derived using the multipolar gauge, wherein there is no explicit inter-dipole Coulomb interaction. We show using a generalised arbitrary-gauge light-matter Hamiltonian that this master equation is obtained in other gauges only if the inter-dipole Coulomb interaction is kept within the interaction Hamiltonian rather than the unperturbed part as in our derivation. Thus, our master equation depends on different S-matrix elements, which give separation-dependent corrections to the standard matrix elements describing resonant energy transfer and collective decay. The two master equations coincide in the large separation limit where static couplings are negligible. We provide an application of our master equation by finding separation-dependent corrections to the natural emission spectrum of the two-dipole system.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Dipole–dipole interactions are central to several important effects in atomic and molecular physics. Early studies by Eisenschitz, London and Förster [1, 2] treated dipolar interactions as perturbative effects arising from direct electrostatic coupling. Molecular quantum electrodynamics (QED) extends these treatments by incorperating retardation effects due to finite signal propagation. As was first shown by Casimir and Polder [3], a striking retardation effect occurs at large separations R/λ ≫ 1 where the R−6 dependence of the dispersion energy is increasingly replaced by an R−7 dependence.

In order to study the dynamics of systems of interacting dipoles open quantum systems theory has proven useful [4]. The master equation formalism can be used to obtain dynamical information about state populations and coherences, and to obtain fluorescence spectra [58]. As will be confirmed in this work, the standard second-order Born–Markov-secular master equation describing two dipoles within a common radiation reservoir depends entirely on well-known QED matrix elements. These matrix elements describe dipole–dipole coupling and decay with retardation effects included. This master equation is obtained by treating the direct electrostatic coupling between the dipoles as a perturbation along with the coupling to transverse radiation. However, it is clear that if the former is sufficiently strong this approach may not be justified, in analogy with the case of externally imposed interactions [9]. Here we consider a system of free dipoles strongly coupled by dipole–dipole interactions. Our focus is on discerning the full dependence of the physics on the inter-dipole separation. We also delineate how microscopic gauge-freedom effects the ensuing master equation derivation.

An important class of systems strongly coupled by dipole–dipole interactions are Rydberg atoms, which have been of interest for some time [10]. In recent years dipole–dipole interactions of Ryberg atoms have been the subject of numerous experimental and theoretical works [1123]. Recently the first experimental confirmation of Förster resonant energy transfer was demonstrated using two Rydberg atoms separated by 15 μm [14]. This type of resonant energy transfer is an important mechanism within photosynthesis, whose quantum nature is of continued interest within open quantum systems theory [24]. Dipole–dipole interactions of Rydberg atoms also offer promising means of implementing quantum gates in which adjacent Rydberg states are treated as effective two-level systems and dipolar interactions are tuned with the use of lasers [21].

Such adjacent Rydberg states are typically separated by microwave transitions, which for small enough separations can be matched or even exceeded by the electrostatic dipole–dipole interaction strength divided by ℏ. Thus, a novel regime of strong electrostatic coupling occurs, in which the usual weak-coupling theory is expected to break down. A repartitioning of the Hamiltonian is necessary in order to identify a genuinely weak system-reservoir interaction, which can then constitute the starting point for perturbation theory. More specifically, we include the direct inter-dipole Coulomb energy within the unperturbed part of the Hamiltonian and only treat the coupling to transverse radiation as a weak perturbation. The master equation we derive exhibits a different dependence on the inter-dipole separation, and this has important consequences for the predicted physics. The rates of collective decay and resonant energy transfer are altered, as are the properties of the light emitted by the system.

There are five sections in this paper. We begin in section 2 by reviewing the standard one and two dipole master equations in the Born–Markov and secular approximations. We show how the standard two-dipole master equation can be obtained for various choices of gauge for the microscopic Hamiltonian. Our purpose is to clearly identify limitations in the standard derivation, which is usually always performed using the multipolar Hamiltonian [4]. This concrete form of the Hamiltonian is the form obtained by choosing the multipolar gauge, also known as the Poincaré gauge [25]. In section 3 we derive an alternative master equation describing the two-dipole system, which only reduces to the standard result in the limit of vanishing direct electrostatic coupling between the dipoles. This occurs in the limit of large separation. In section 4 we solve the master equation derived in section 3 and compare the solution with that of the standard master equation. We also obtain corrections to the emission spectrum of the two-dipole system. Finally in section 5 we summarise our findings. We assume natural units ${\hslash }={\epsilon }_{0}=c=1$ throughout.

2. Gauge-invariant master equations

2.1. Single-dipole Hamiltonian and master equation

Here we identify sufficient conditions in order that the same master equation can be obtained from different microscopic Hamiltonians. This will be important when it comes to deriving the two-dipole master equation in the following sections. Let us consider a single dipole within the electromagnetic bath, and assume that there are only two relevant states ($| g\rangle ,| e\rangle $) of the dipole separated by energy ω0 = ωe − ωg. Associated raising and lowering operators are defined by ${\sigma }^{+}=| e\rangle \langle g| $ and ${\sigma }^{-}=| g\rangle \langle e| $. The electromagnetic bath is described by creation and annihilation operators ${a}_{{\bf{k}}\lambda }^{\dagger },\,{a}_{{\bf{k}}\lambda }$ for a single photon with momentum k and polarisation λ. The photon frequency is denoted ${\omega }_{k}=| {\bf{k}}| $.

The energy of the dipole-field system is given by a Hamiltonian of the form H = H0 + V, where

Equation (1)

defines the free (unperturbed) Hamiltonian and V denotes the interaction Hamiltonian. Gauge-freedom within the microscopic description results in the freedom to choose a number of possible interaction Hamiltonians. We define the generalised-gauge transformation [26]

Equation (2)

In this expression the ${\alpha }_{k}$ are real and dimensionless, the ekλ, λ = 1, 2 are mutually orthogonal polarisation unit vectors, which are both orthogonal to k, L3 is the volume of the assumed fictitious quantisation cavity, and $\hat{{\bf{d}}}$ denotes the dipole moment operator. By making the two-level approximation after having transformed the Coulomb gauge Hamiltonian using the unitary operator ${R}_{\{{\alpha }_{k}\}}$ we obtain the Hamiltonian H = H0 + V where [26]

Equation (3)

and

Equation (4)

The term ${V}^{(2)}$ is a self-energy term, which does not act within the two-level dipole Hilbert space and which depends on the field

Equation (5)

The coupling constant gk λ and the (real) coefficients ${u}_{k}^{\pm }$ are defined as

Equation (6)

where ${\bf{d}}$ and ${\omega }_{0}$ denote the two-level transition dipole moment and transition frequency, respectively. The real numbers αk can be chosen arbitrarily. Choosing αk = 0 yields the Coulomb-gauge Hamiltonian while choosing αk = 1 yields the multipolar-gauge Hamiltonian. Letting αk = 1 in equation (2) yields the well-known Power–Zienau–Woolley (PZW) transformation that relates the Coulomb and multipolar gauges. While the relation between the Coulomb and multipolar gauge has been discussed extensively [2731], the PZW transformation is in fact a special case of a broader class of unitary gauge-fixing transformations [3133]. More generally still, the freedom to choose the αk within the canonical transformation (2) implies redundancy within our mathematical description and is henceforth referred to as generalised gauge-freedom. A third special case of equation (3) is afforded by making the choice αk = ω0/(ω0 + ωk), which specifies a symmetric mixture of Coulomb and multipolar couplings. This representation has proved useful in both photo-detection theory [31, 34] and open quantum systems theory [26], because within this representation ${u}_{k}^{+}\equiv 0$. The counter-rotating terms in the linear dipole-field interaction term V − V(2) are thereby eliminated without use of the rotating-wave approximation.

Given the above arbitrary generalised-gauge description, it is clear that arbitrary matrix elements ${M}_{{fi}}(t)=\langle f| M(t)| i\rangle $ between eigenstates $| f\rangle $ and $| i\rangle $ of H0 will not be the same when the evolution of the operator M is determined by different total Hamiltonians H and H', that have been obtained by making different choices of αk in equation (3). In contrast on-energy-shell QED S-matrix elements are necessarily the same for two interaction Hamiltonians V and V', constrained such that the corresponding total Hamiltonians are related by a unitary transformation eiT as [25, 28, 35]

Equation (7)

For gauge-invariance to hold the unperturbed Hamiltonian H0 must be identified as the same operator before and after the transformation by eiT. Note however, that the unperturbed Hamiltonian H0 given in equation (1) does not commute with the unitary transformation ${R}_{\{{\alpha }_{k}\}}$ given in equation (2) meaning that this H0 represents a different physical observable depending on the choice of interaction. Despite this, S-matrix elements based on the partition H = H0 + V are invariant, because H0 in equation (1) does not explicitly depend on the αk and is therefore the same for each different choice of generalised-gauge.

Having determined the conditions under which QED matrix elements are gauge-invariant we now turn our attention to deriving a master equation describing the two-level dipole within the radiation field. The conventional derivation of the second order quantum optical master equation, as found in [36] for example, does not at any point involve self-energy contributions due to the V(2) term within the interaction V. In general however, this term does contribute to dipole level-shifts, as is shown in appendix A.1. In the general case that the temperature of the radiation field is arbitrary, the self-energy contributions from V(2) can be incorporated into the master equation by defining the Hamiltonian

Equation (8)

where the excited and ground state self-energy shifts are defined as

Equation (9)

in which n = e, g and ${\rho }_{F}^{\mathrm{eq}}(\beta )={{\rm{e}}}^{-\beta {\sum }_{{\bf{k}}\lambda }{\omega }_{k}{a}_{{\bf{k}}\lambda }^{\dagger }{a}_{{\bf{k}}\lambda }}/\mathrm{tr}({{\rm{e}}}^{-\beta {\sum }_{{\bf{k}}\lambda }{\omega }_{k}{a}_{{\bf{k}}\lambda }^{\dagger }{a}_{{\bf{k}}\lambda }})$ with $\beta $ the inverse temperature of the radiation field. Since V(2) is gauge-dependent we cannot include the self-energy shifts within the unperturbed Hamiltonian H0 without ruining the gauge-invariance of any S-matrix elements obtained using the unperturbed states. Instead we replace the free system Hamiltonian in the usual Born–Markov master equation with the shifted Hamiltonian ${\tilde{H}}_{d}$ directly to obtain the second order master equation

Equation (10)

This master equation automatically includes the level shifts due to V(2) within the unitary evolution part, but the rest of the master equation is expressed in terms of the original partition H = H0 + V. Using this partition where H0 and V are given in equations (1) and (3) respectively, equation (10) yields the αk-independent result

Equation (11)

Here $N=1/({{\rm{e}}}^{\beta {\omega }_{0}}-1)$, ${\tilde{\omega }}_{0}={\omega }_{0}+{\rm{\Delta }}$ and

Equation (12)

where the continuum limit for wavevectors k has been applied and ${N}_{k}=1/({{\rm{e}}}^{\beta {\omega }_{k}}-1)$. Further details of the calculations leading to the final result for Δ in equation (12) are given in appendix A.1. We note that for Nk = 0 we have

Equation (13)

The αk-independence (generalised gauge-invariance) of the master equation (11) can be understood by noting that the spontaneous emission rate γ and level-shift Δ in equation (13) are gauge-invariant QED matrix elements that can be obtained directly using second order perturbation theory.

In summary, we have shown that the master equation obtained from different, unitarily equivalent microscopic Hamiltonians is the same provided it depends only on S-matrix elements. S-matrix elements are invariant if the bare Hamiltonian H0 is kept the same for each choice of total Hamiltonian. In what follows this will be seen to be significant for the derivation of the master equation describing two strongly coupled dipoles.

2.2. Arbitrary gauge derivation of the standard two-dipole master equation

Let us now turn our attention to obtaining the analogous result to equation (11) for the case of two identical interacting dipoles at positions R1 and R2 within a common radiation reservoir. The transition dipole moments and transition frequencies of the dipoles are independent of the dipole label and are denoted d and ω0 respectively. The wavelength corresponding to ${\omega }_{0}$ is denoted with ${\lambda }_{0}$. An important quantity in the two-dipole system dynamics is the inter-dipole separation $R=| {{\bf{R}}}_{2}-{{\bf{R}}}_{1}| $. In terms of R we can identify in the usual way three distinct parameter regimes: $R\ll {\lambda }_{0}$ is the near zone in which R−3-dependent terms dominate, $R\sim {\lambda }_{0}$ is the intermediate zone in which R−2-dependent terms may become significant, and $R\gg {\lambda }_{0}$ is the far-zone (radiation zone) in which R−1-dependent terms dominate. We give a general derivation of the standard two dipole master equation, in which the gauge freedom within the microscopic Hamiltonian is left open throughout. This reveals limitations within the conventional derivation using the multipolar gauge. To begin we define the two-dipole generalised PZW gauge transformation by [26, 34]

Equation (14)

where dμ denotes the dipole moment operator of the μ'th dipole. We now transform the dipole approximated Coulomb gauge Hamiltonian using ${R}_{\{{\alpha }_{k}\}}$ and afterwards make the two-level approximation for each dipole. This implies that ${{\bf{d}}}_{\mu }={\bf{d}}({\sigma }_{\mu }^{+}+{\sigma }_{\mu }^{-})$, and that the Hamiltonian can be written H = H0 + V with

Equation (15)

and

Equation (16)

Analogously to the single-dipole case the first term in equation (16) defines a linear dipole-field interaction component while the term ${V}_{\{{\alpha }_{k}\}}^{(2)}$ consists of self-energy contributions for each dipole and the radiation field;

Equation (17)

where m is the dipole mass. Due to the two-level approximation the first term in (17) is proportional to the identity, while the second term is a radiation self-energy term. The field $\tilde{A}$ is defined as in the singe-dipole case by equation (5). The final term in equation (16) ${C}_{\{{\alpha }_{k}\}}$, has no analogue in the single-dipole Hamiltonian. This term gives a static Coulomb-like interaction between the dipoles, which is independent of the field;

Equation (18)

where ${\sigma }_{\mu }^{x}={\sigma }_{\mu }^{+}+{\sigma }_{\mu }^{-}$. In the Coulomb gauge (αk = 0) the term ${C}_{\{{\alpha }_{k}\}}$ reduces to the usual dipole–dipole Coulomb interaction. In the multipolar gauge ${C}_{\{{\alpha }_{k}\}}$ vanishes, and the interaction in equation (16) therefore reduces to a sum of interaction terms for each dipole.

It is important to note that as in the single-dipole case ${R}_{\{{\alpha }_{k}\}}$ does not commute with H0 given in equation (15) implying that H0 represents a different physical observable for each choice of αk. More generally, since ${R}_{\{{\alpha }_{k}\}}$ is a non-local transformation, which mixes material and transverse field degrees of freedom, the canonical material and field operators are different for each choice of αk. This implies that the master equation for the dipoles will generally be different for each choice of αk. We can, however, obtain a gauge-invariant result by ensuring that the master equation depends only on gauge-invariant S-matrix elements. These matrix elements are gauge-invariant despite the implicit difference in the material and field degrees of freedom within each generalised gauge. Usually the two-dipole master equation is derived using the specific choice αk = 1 (multipolar gauge) for which the direct Coulomb-like coupling ${C}_{\{{\alpha }_{k}\}}$ vanishes identically. To obtain the same master equation for any other choice of ${\alpha }_{k}\ne 1$, we must include ${C}_{\{{\alpha }_{k}\}}$ within the interaction Hamiltonian V. The reason is that H0 must be identified as the same operator for each choice of αk in order that the gauge-invariance of the associated S-matrix holds.

We now proceed with a direct demonstration that the standard two-dipole master equation can indeed be obtained for any other choice of αk, provided ${C}_{\{{\alpha }_{k}\}}$ is kept within the interaction Hamiltonian V. To do this we substitute the interaction Hamiltonian in equation (16) into the second order Born–Markov master equation in the interaction picture with respect to H0, which is given by

Equation (19)

where VI(t) denotes the interaction Hamiltonian in the interaction picture and ρI(t) denotes the interaction picture state of the two dipoles. We retain contributions up to order e2 and perform a further secular approximation, which neglects terms oscillating with twice the transition frequency ω0. Transforming back to the Schrödinger picture and including the single-dipole self-energy contributions as in equation (10), we arrive after lengthy but straightforward manipulations at the final αk-independent result

Equation (20)

This equation is identical in form to the standard two-dipole master equation, which can be found in [4] for example. The coefficients within the master equation are as follows. The decay rates γμν are given by

Equation (21)

where

Equation (22)

In equation (21) and throughout we denote spatial components with Latin indices and adopt the convention that repeated Latin indices are summed. The quantity γ12 denotes an R-dependent collective decay rate. The third equality in equation (21) wherein γ12 has been expressed as a matrix element involving V makes the reason for the αk-independence of this rate clear. We now turn our attention to the master equation shifts ${\rm{\Delta }}={\tilde{\omega }}_{0}-{\omega }_{0}$ and Δ12. The single-dipole shift Δ includes all self-energy contributions, which have been dealt with in the same way as for the single-dipole master equation (see equation (10)). The shift Δ is therefore as in equation (12). Details of the calculation of the joint shift Δ12 are given in appendix A.2 with the final result being

Equation (23)

where $| f\rangle =| g,e,0\rangle $, $| i\rangle =| e,g,0\rangle $ and $| n\rangle $ are eigenstates of H0 and

Equation (24)

As indicated by the second equality in equation (23) Δ12 is nothing but the well-known gauge-invariant QED matrix-element describing resonant energy-transfer.

We have therefore obtained the standard result, equation (20), without ever making a concrete choice for the αk. In order that the standard result is obtained the direct Coulomb-like interaction ${C}_{\{{\alpha }_{k}\}}$ must be kept within the interaction Hamiltonian V. This ensures that for all αk the unperturbed Hamiltonian H0 is that used in conventional derivations wherein αk = 1. The S-matrix elements involving ${C}_{\{{\alpha }_{k}\}}$ that appear as coefficients in the master equation are then αk-independent and are the same as those obtained in the conventional derivation. Our derivation makes it clear that when ${C}_{\{{\alpha }_{k}\}}$ is sufficiently strong compared with the linear dipole-field coupling term, its inclusion within the interaction Hamiltonian rather than H0 may be ill-justified. The standard master equation may therefore be inaccurate in such regimes. This fact is obscured within conventional derivations that use the multipolar gauge αk = 1, because in this gauge ${C}_{\{{\alpha }_{k}\}}$ vanishes identically. However, one typically still assumes weak-coupling to the radiation field in the multipolar gauge, and this leads to the standard master equation (20). If instead we adopt the Coulomb gauge αk = 0 we obtain the static dipole–dipole interaction ${C}_{\{0\}}=C{\sigma }_{1}^{x}{\sigma }_{2}^{x}$, where in the mode continuum limit

Equation (25)

This quantity coincides with the near-field limit of the resonant energy transfer element Δ12 given in equation (23). In the near-field regime $R/\lambda \ll 1$, ${C}_{\{0\}}$ may be too strong to be kept within the purportedly weak perturbation V and the standard master equation, which only results when one treats C{0} as a weak perturbation, should then break down. This will be discussed in more detail in the following section.

3. Corrections to the standard master equation

3.1. Derivation of an alternative master equation

In the near-field regime $R/{\lambda }_{0}\ll 1$ the rate of spontaneous emission into the transverse field is much smaller than the direct dipolar coupling; $C/\gamma \gg 1$. Moreover, for a system of closely spaced Rydberg atoms, the electrostatic Coulomb interaction may be such that $C\sim {\omega }_{0}$. For example, given a Rydberg state with principal quantum number n = 50 we can estimate the maximum associated dipole moment as $(3/2){n}^{2}{a}_{0}e\sim {10}^{-26}\,\mathrm{Cm}$ where a0 is the Bohr radius and e the electronic charge. For a $1\mu {\rm{m}}$ separation, which is approximately equal to $10{n}^{2}{a}_{0}$, the electrostatic dipole interaction $C/{\omega }_{0}\sim 1$ for ω0 corresponding to a microwave frequency. In such situations it is not clear that the Coulomb interaction can be included within the perturbation V with the coupling to the transverse field. In the multipolar gauge where no direct Coulomb interaction is explicit the same physical interaction is mediated by the low frequency transverse modes, which must be handled carefully. A procedure which separates out these modes should ultimately result in a separation of the Coulomb interaction, which is of course already explicit within the Coulomb gauge. We remark that when considering realistic Rydberg atomic systems within the strong dipole–dipole coupling regime the validity of the two-level model should also be considered. However, moving beyond the two-level approximation is beyond the scope of this paper. Our aim is to consider general atomic, molecular and condensed matter systems strongly-coupled by dipole–dipole interactions for which two-level models are typically used [4, 37, 38]. Retaining the two-level model for each dipole allows us to succinctly compare with existing literature and thereby determine the relative difference produced by our non-perturbative treatment of dipole–dipole interactions.

In the Coulomb gauge the interaction Hamiltonian V coupling to the transverse radiation field is

Equation (26)

with ${\sigma }_{\mu }^{y}=-{\rm{i}}({\sigma }_{\mu }^{+}-{\sigma }_{\mu }^{-})$ and

Equation (27)

The contribution of the transverse field to Δ12 in equation (21) is found using equation (26) to be

Equation (28)

When the contribution $C=\langle 0,e,g| {C}_{\{0\}}| 0,g,e\rangle $ resulting from the direct Coulomb interaction is added to ${\tilde{{\rm{\Delta }}}}_{12}$ the fully retarded result Δ12 is obtained. The two matrix elements Δ12 and ${\tilde{{\rm{\Delta }}}}_{12}$ therefore only differ in their near-field components, which vary as R−3 and which we denote by ${{\rm{\Delta }}}_{12}^{\mathrm{nf}}$ and ${\tilde{{\rm{\Delta }}}}_{12}^{\mathrm{nf}}$, respectively. According to equations (23) and (28) the components ${{\rm{\Delta }}}_{12}^{\mathrm{nf}}$ and ${\tilde{{\rm{\Delta }}}}_{12}^{\mathrm{nf}}$ dominate at low frequencies ω0. Since Δ12 is evaluated at resonance ωk = ω0, it follows that within the multipolar-gauge the low ωk modes within the system-reservoir coupling give rise to a strong dipole–dipole interaction in the form of ${{\rm{\Delta }}}_{12}^{\mathrm{nf}}\approx C$. In such regimes the multipolar interaction Hamiltonian cannot be classed as a weak perturbation. On the other hand, the matrix element ${\tilde{{\rm{\Delta }}}}_{12}$ is obtained using the Coulomb gauge interaction in equation (26), and is such that ${\tilde{{\rm{\Delta }}}}_{12}^{\mathrm{nf}}={{\rm{\Delta }}}_{12}^{\mathrm{nf}}-C\approx 0$. Within the Coulomb gauge the interaction equivalent to the low frequency part of the multipolar gauge system-reservoir coupling is a direct dipole–dipole Coulomb interaction C{0}. This appears explicitly in the Hamiltonian, but has not been included within equation (26), which therefore represents a genuinely weak perturbation.

The collective decay rate ${\gamma }_{12}$ as given in equation (21) does not involve the direct Coulomb interaction C{0} in any way, and can be obtained from the transverse field interaction in equation (26) or from the multipolar interaction. Crucially, in the near-field regime $R/{\lambda }_{0}\ll 1$ the terms γ, γ12 and ${\tilde{{\rm{\Delta }}}}_{12}$, which result from the interaction in equation (26), are several orders of magnitude smaller than the direct electrostatic coupling C. Motivated by the discussion above, we include the Coulomb interaction within the unperturbed Hamiltonian, but continue to treat the interaction with the transverse field as a weak perturbation. This gives rise to a master equation depending on different S-matrix elements.

The unperturbed Hamiltonian H0 = Hd + HF is defined by

Equation (29)

where $C\in {\mathbb{R}}$ is given by equation (25). The corresponding interaction Hamiltonian is then given in equation (26). We begin by diagonalising Hd as

Equation (30)

where

Equation (31)

and

Equation (32)

with $\eta =\sqrt{{\omega }_{0}^{2}+{C}^{2}}$. Next we move into the interaction picture with respect to H0 and substitute the interaction picture interaction Hamiltonian into equation (19). Moving back into the Schrödinger picture we eventually obtain

Equation (33)

Here, ω1 = η − C and ω2 = η + C, while ${A}_{\mu (-\zeta )}={A}_{\mu \zeta }^{\dagger }$ and ${A}_{\mu {\zeta }_{n}}\equiv {A}_{\mu n}\,(n=1,2)$ with

Equation (34)

where

Equation (35)

The coefficients Γμν(ω) are defined by

Equation (36)

where AI(x, t) denotes the field A(x) in equation (27) once transformed into the interaction picture, and $\langle \cdot {\rangle }_{\beta }$ denotes the average with respect to the radiation thermal state at temperature β−1. The Γμν(ω) are symmetric ${{\rm{\Gamma }}}_{\mu \nu }(\omega )={{\rm{\Gamma }}}_{\nu \mu }(\omega )$ and can be written

Equation (37)

where

Equation (38)

The frequency integrals in equation (38) are to be understood as principal values. The decay rates γμν(ω) in equation (38) coincide with those found in the standard master equation (20) when evaluated at ω0, though are here evaluated at the frequencies ω1,2. The quantities Sμν are related to the shifts Δ and Δ12, defined in equations (12) and (23) respectively, by

Equation (39)

In deriving equation (33) we have not yet performed a secular approximation, in contrast to the derivation of equation (20). However, naively applying a secular approximation that neglects off-diagonal terms for which $\zeta \ne \zeta ^{\prime} $ in the summand in equation (33) would not be appropriate, because this would eliminate terms that are resonant in the limit $C\to 0$. Instead we perform a partial secular approximation which eliminates off-diagonal terms for which ζ and ζ' have opposite sign. These terms remain far off-resonance for all values of C. The resulting master equation is given by

Equation (40)

We are now in a position to compare our master equation (40) with the usual result in (20). In the limit $C\to 0$ we have $\eta \to {\omega }_{0}$ so that ${\omega }_{\mathrm{1,2}}\to {\omega }_{0}$. The rates and shifts in equation (38) are then evaluated at ω0 within equation (40). Also, the Hamiltonian Hd tends to the bare Hamiltonian ${\omega }_{0}({\sigma }_{1}^{+}{\sigma }_{1}^{-}+{\sigma }_{2}^{+}{\sigma }_{2}^{-})$, and furthermore we have that

Equation (41)

Thus, taking the limit $C\to 0$ in equation (40) one recovers equation (20) with Δ12 replaced by ${\tilde{{\rm{\Delta }}}}_{12}$ given in equation (28). However, since ${\tilde{{\rm{\Delta }}}}_{12}\to {{\rm{\Delta }}}_{12}$ when $C\to 0$, equations (40) and (20) coincide in this limit. For finite C, equation (40) offers separation-dependent corrections to the usual master equation and is the main result of this section.

3.2. Discussion: gauge-invariance of the new master equation

It is important to note that while our master equation (40) is generally different to the usual gauge-invariant result (equation (20)) there is no cause for concern regarding the issue of gauge-invariance. As we have shown the standard master equation can be obtained when αk = 0 provided one uses a partitioning of the Hamiltonian in the form $H={H}_{0}+{V}^{\mathrm{usual}}$ where H0 is given by equation (15) and

Equation (42)

Our master equation (40) has also been obtained by choosing αk = 0, but our derivation makes use of the different partitioning $H={\tilde{H}}_{0}+V$ where ${\tilde{H}}_{0}$ is defined as in equation (29) and $V={V}^{\mathrm{usual}}-{C}_{\{0\}}$ is defined as in equation (26). The two different partitionings of the same Hamiltonian yield two different second order master equations.

As we have shown the standard Born–Markov-secular master equation (20) can be obtained for any other choice of αk provided that the unperturbed Hamiltonian is always defined as in equation (15). Similarly a full secular approximation of our master equation (40) can also be obtained for any other choice of αk provided the unperturbed Hamiltonian is always defined as in equation (29). We note further that the secular approximation is well justified within the near-field regime of interest $R\ll {\lambda }_{0}$. Let us consider for example the multipolar gauge obtained by choosing αk = 1. In order to achieve the appropriate partitioning of the multipolar Hamiltonian for derivation of our master equation one must add C{0} to the usual multipolar unperturbed Hamiltonian H0 given in equation (15), and simultaneously subtract C{0} from the usual multipolar interaction Hamiltonian. Using this repartitioning of the multipolar Hamiltonian the Born–Markov-secular master equation is found to coincide with our master equation (40) once a full secular approximation is performed within the latter. This derivation is however, more cumbersome than the Coulomb gauge derivation. Since the Coulomb energy is naturally explicit within the Coulomb gauge, the latter is the most natural gauge to choose for the purpose of including the relatively strong static interaction within the unperturbed Hamiltonian.

Any difference between the master equation (40) and the corresponding partially-secular result found using ${\alpha }_{k}\ne 0$ is contained entirely within non-secular contributions. These contributions are negligible within the regime of interest R ≪ λ0 and have only been retained within equation (40) to facilitate comparison with the standard result equation (20). Moreover, in the far-field regime $R\gg {\lambda }_{0}$ the master equations (20) and (40) coincide, so the master equation (40) is also gauge-invariant within this regime.

4. Solutions and emission spectrum

4.1. Solutions

For large inter-dipole separations R ≫ λ the master equations (20) and (40) coincide and they therefore yield identical physical predictions. However, in the near-zone R ≪ λ0 the master equations generally exhibit significant differences. To compare the two sets of predictions we assume a vacuum field N = 0 and consider the experimental situation in which the system is prepared in the symmetric state $| {\epsilon }_{3}\rangle $. This state is a simultaneous eigenstate of the dipole Hamiltonian ${\omega }_{0}({\sigma }_{1}^{+}{\sigma }_{1}^{-}+{\sigma }_{2}^{+}{\sigma }_{2}^{-})$ appearing in the standard master equation (20), and of the Hamiltonian Hd appearing in our master equation (40). Experimentally, one expects to find that the system initially prepared in the state $| {\epsilon }_{3}\rangle $ decays into the stationary state. Theoretically, different stationary states are predicted by the two master equations (20) and (40), and the rates of decay into these respective stationary states are also different. Figures 1 and 2 compare the symmetric and stationary state populations found using master equations (20) and (40) when the system starts in the symmetric eigenstate $| {\epsilon }_{3}\rangle $. For small separations the ground and symmetric state populations obtained from our master equation (40) crossover earlier, which indicates more rapid symmetric state decay than is predicted by equation (20) (see figure 1). This gives rise to the different starting values at R = ra of the curves depicted in figure 2. For larger separations the solutions converge and become indistinguishable for all times.

Figure 1.

Figure 1. The populations of the stationary state ($| {\epsilon }_{1}\rangle $ or $| g,g\rangle $) and symmetric state $| {\epsilon }_{3}\rangle $ are plotted as functions of t for fixed separation R = 10ra, where ra = n2a0 is a characteristic Rydberg atomic radius, with n = 50 and a0 the Bohr radius. We have assumed no thermal occupation of the field, N = 0 and that the transition dipole moment d is orthogonal to the separation vector R. The transition frequency is chosen in the microwave regime ω0 = 1010. We use pg and ps to denote the stationary and symmetric state populations respectively, and we use p and p0 to denote populations obtained from master equations (40) and (20), respectively. In the case of equation (40) the stationary state is $| {\epsilon }_{1}\rangle $ whereas in the case of equation (20) the stationary state is simply $| g,g\rangle $. The initial condition chosen is ps = 1. The ground and symmetric state populations obtained from the master equation (40) crossover significantly earlier than those obtained from equation (20).

Standard image High-resolution image
Figure 2.

Figure 2. The populations of the stationary state ($| {\epsilon }_{1}\rangle $ or $| g,g\rangle $) and symmetric state $| {\epsilon }_{3}\rangle $ are plotted as functions of R for fixed time t = 1/8γ. The remaining parameters are as in figure 1. For the separations considered the solutions to equation (20) do not vary significantly, while the solutions to equation (40) converge to those of equation (20) only for larger values of R. The subplot shows these solutions over a much larger scale of separations up to $O({10}^{6}{r}_{a})$, for which the solutions to equations (20) and (40) are indistinguishable.

Standard image High-resolution image

The different behaviour in figures 1 and 2 can be understood by looking at a few relevant quantities. The matrix element of the combined dipole moment between ground and symmetric eigenstates is found to be

Equation (43)

which is different to the usual transition dipole moment $\langle {\epsilon }_{3}| {{\bf{d}}}_{1}+{{\bf{d}}}_{2}| {gg}\rangle =\sqrt{2}{\bf{d}}$. Since $a\to 1/\sqrt{2}$ as $C\to 0$, the dipole moment d31 reduces to $\sqrt{2}{\bf{d}}$ when $R\to \infty $. As R decreases, however, d13 becomes increasingly large compared with $\sqrt{2}{\bf{d}}$. This is consistent with the more rapid decay observed in figure 1. A more complete explanation of this behaviour can be obtained by calculating the rate of decay of the symmetric state into the vacuum, which we denote γs. Using Fermi's golden rule, and the eigenstates given in equation (32), we obtain

Equation (44)

Only when $C\to 0$, such that ${\omega }_{2}=\eta +C\to {\omega }_{0}$ and $c\to 1/\sqrt{2}$, does this decay rate reduce to that obtained when using the bare eigenstates $| i,j\rangle ,(i,j=e,g)$, which is

Equation (45)

where γ12(ω0) is given in equation (21). As shown in figure 3, for sufficiently small R the decay rate γs(ω2) is significantly larger than γs,0.

Figure 3.

Figure 3. Comparison of the symmetric state decay rates γs (from our master equation) and γs,0 (from the standard master equation) as functions of separation R. All parameters are chosen as in figure 1.

Standard image High-resolution image

In contrast to the decay behaviour of the symmetric state, the predictions of the master equations (20) and (40) are the same if the system is assumed to be prepared in the anti-symmetric state $| {\epsilon }_{2}\rangle $, which like $| {\epsilon }_{3}\rangle $ is a simultaneous eigenstate of ${\omega }_{0}({\sigma }_{1}^{+}{\sigma }_{1}^{-}+{\sigma }_{2}^{+}{\sigma }_{2}^{-})$ and Hd. Both master equations predict that the population of the state $| {\epsilon }_{2}\rangle $ remains stationary, i.e., that it is a completely dark state. This can be understood by noting that the collective dipole moment associated with the anti-symmetric to stationary state transition vanishes when either stationary state, $| g,g\rangle $ or $| {\epsilon }_{1}\rangle $, is used. Finally, the predicted behaviour by our master equation (40) of the standard stationary state $| g,g\rangle $ is illustrated in figure 4. For an initial state $| {\epsilon }_{3}\rangle $ the population pgg(t) of the state $| g,g\rangle $ at a given time t, is identical to that predicted by equation (20) only for sufficiently large R whereby $| {\epsilon }_{1}\rangle \approx | g,g\rangle $.

Figure 4.

Figure 4. The population of the state $| g,g\rangle $ found using equation (40) is plotted as a function of separation R for various times. All remaining parameters are as in figure 1. The dashed lines give the corresponding populations found using equation (20), which are insensitive to variations in R over the range considered. For large R the two sets of solutions agree. In particular, the steady state population ${p}_{g,g}({\rm{\infty }})=| \langle g,g| {\epsilon }_{1} \rangle {| }^{2}$ is equal to unity only for sufficiently large R.

Standard image High-resolution image

4.2. Emission spectrum

In this section we apply our master equation (40) to calculate the emission spectrum of the two-dipole system initially prepared in the symmetric state $| {\epsilon }_{3}\rangle $. This provides a means by which to test experimentally whether our predictions are closer to measured values than the standard approach. The spectrum of radiation is defined according to the quantum theory of photodetection by [39, 40]

Equation (46)

where for simplicity we assume that the field is in the vacuum state. Since the master equations (20) and (40) yield different predictions for this experimentally measurable quantity, an experiment could be used to test which master equation is the most accurate.

The detector is located at position x with $x\gg R$, so that only the radiative component ${{\bf{E}}}_{s,\mathrm{rad}}$ of the electric source field, which varies as $| {\bf{x}}-{{\bf{R}}}_{\mu }{| }^{-1}$, need be used. This is the only part of the field responsible for irreversibly carrying energy away from the sources. The positive and negative frequency components of the radiation source field are given within both rotating-wave and Markov approximations by

Equation (47)

where ${{\bf{r}}}_{\mu }={\bf{x}}-{{\bf{R}}}_{\mu }$, and ${t}_{\mu }=t-{r}_{\mu }$ is the retarded time associated with the μ'th source. For $x\gg R$ we have to a very good approximation that ${{\bf{r}}}_{1}={{\bf{r}}}_{2}={\bf{r}}$, where r is the relative vector from x to the midpoint of R1 and R2. Substituting equation (47) into equation (46) within this approximation yields

Equation (48)

where

Equation (49)

To begin with, let us use the standard master equation (20) to find the required two-time correlation function. Assuming that the system is initially prepared in the symmetric state $| {\epsilon }_{3}\rangle $, using the standard master equation (20) together with the method of calculation given in appendix A.3 we obtain the two-time correlation function

Equation (50)

where γs,0 and Δ12 are given in equations (45) and (21), respectively. By direct integration of equation (50) one obtains the corresponding Lorentzian spectrum

Equation (51)

Let us now turn our attention to the spectrum obtained from our new master equation (40). We have seen that the solutions of equations (40) and (20) differ only in the near field regime $R\ll \lambda $. For sufficiently small R we have that ${\omega }_{0}\sim C$, and the frequency difference ${\omega }_{2}-{\omega }_{1}=2C\sim 2{\omega }_{0}$ is large. In this situation we can perform a full secular approximation within equation (40) to obtain the master equation

Equation (52)

with

Equation (53)

and

Equation (54)

Solving this secular master equation allows us to obtain a simple expression for the emission spectrum.

The correlation function in equation (46) defines the radiation intensity when it is evaluated at t = t'. Naively calculating the quantity ${\sum }_{\mu ,\nu =1}^{2}\langle {\sigma }_{\mu }^{+}(t){\sigma }_{\nu }^{-}(t){\rangle }_{1}$ taken in the stationary state $| {\epsilon }_{1}\rangle $ yields a non-zero stationary intensity, because $| {\epsilon }_{1}\rangle $ is a superposition involving both $| g,g\rangle $ and the doubly excited bare state $| e,e\rangle $. A non-zero radiation intensity even in the stationary state is clearly non-physical. However, a more careful analysis recognises that when the radiation source fields are to be used in conjunction with equation (40) the optical approximations used in their derivation should be applied in the interaction picture defined in terms of the dressed Hamiltonian Hd given in equation (30). One then obtains the source field

Equation (55)

where ${\sigma }_{\mu ,{nm}}={\sigma }_{\mu ,{nm}}^{+}+{\sigma }_{\mu ,{nm}}^{-}$ in which ${\sigma }_{\mu ,{nm}}^{\pm }$ denotes the nm'th matrix element of ${\sigma }_{\mu }^{\pm }$ in the basis $| {\epsilon }_{n}\rangle $. The transition frequencies associated with this basis are denoted ${\epsilon }_{{nm}}={\epsilon }_{n}-{\epsilon }_{m}$, and the raising and lowering operators are denoted ${\theta }_{{nm}}=| {\epsilon }_{n}\rangle \langle {\epsilon }_{m}| ,\,n\ne m$. The derivation of equation (55) is given in appendix A.4. According to equation (55) the annihilation (creation) radiation source field is now associated with lowering (raising) operators in the dressed basis $| {\epsilon }_{i}\rangle $ rather than in the bare basis $| n,m\rangle ,(n,m=e,g)$. Substitution of equation (55) into equation (46) yields

Equation (56)

where we have again assumed that ${{\bf{r}}}_{1}={{\bf{r}}}_{2}={\bf{r}}$. Unlike the correlation function in equation (92), when t = t' the correlation functions $\langle {\theta }_{{pq}}(t){\theta }_{{nm}}(t^{\prime} )\rangle ,\,p\gt q,\,m\gt n$ vanish in the stationary (ground) state θ11. The radiation intensity is therefore seen to vanish in the stationary limit as required physically.

Taken in the symmetric state θ33 the only non-zero two-time correlation function that contributes to equation (56) is found to be

Equation (57)

where

Equation (58)

is the shifted symmetric to ground transition frequency. Integration of equation (57) according to equation (56) then yields the Lorentzian spectrum

Equation (59)

Full details of the calculation of the spectrum in equation (59) are given in appendix A.4. In the limit of large separation $C\to 0$, which implies that ${\omega }_{2}\to {\omega }_{0}$, ${\tilde{\omega }}_{2}\to {\tilde{\omega }}_{0}+{{\rm{\Delta }}}_{12}$, ${\gamma }_{s}({\omega }_{2})\to {\gamma }_{s,0}$, and $a\to 1/\sqrt{2}$. As a result $s(\omega )\to {s}_{0}(\omega )$ for large R and the predicted spectra coincide. On the other hand, for sufficiently small R the spectrum s(ω) again offers separation-dependent corrections to the standard result s0(ω).

The two spectra ${s}_{0}(\omega )$ and s(ω) are compared in figures 5 and 6. As their relative widths are proportional to the rates, they are given in equations (45) and (44), respectively. These quantities have been plotted already in figure 3. The relative heights of the spectral peaks are ${s}_{0}({\tilde{\omega }}_{0}+{{\rm{\Delta }}}_{12})/\mu $ and $s({\tilde{\omega }}_{2})/\mu $ respectively, which are plotted in figure 7. This figure shows that the peak heights in the spectra begin to diverge as R decreases. At a separation of 15ra, where ra = n2a0, n = 50 is a characteristic Rydberg atomic radius, the peak value of s(ω) is around two times larger than the peak value of s0(ω) for the parameters chosen here. The positions of the peaks are ${\tilde{\omega }}_{0}+{{\rm{\Delta }}}_{12}$ and ${\tilde{\omega }}_{2}$, respectively, and these are plotted in figure 8. The ultra-violet cut-off chosen for the calculation of the single-dipole shift components corresponds to the inverse dipole radius wavelength, namely 2π c /ra. This value is chosen for consistency with the electric dipole approximation that we have used throughout. For small R the spectrum s(ω) is blue-shifted relative to s0(ω). Figure 8 shows that the ratio of peak positions approaches a constant value around two for very small R. These differences could in principle be detected in an experiment. At a separation of 20ra, which is roughly 2.5 μm, for instance, the difference in shifted frequencies ${\tilde{\omega }}_{2}-{\tilde{\omega }}_{0}-{{\rm{\Delta }}}_{12}$ is around 1 Ghz for the parameters chosen in figure 1. This is similar in magnitude to the Lamb-shift in atomic hydrogen.

Figure 5.

Figure 5. The spectra s(ω) and s0(ω) are plotted with R = 10ra and with all remaining parameters as in figure 1. For this separation the peak heights and centres are quite different as shown in figures 7 and 8, respectively. Here, for illustrative purposes, the spectra have both been centred at zero and normalised by their respective peak heights.

Standard image High-resolution image
Figure 6.

Figure 6. The spectra s(ω) and s0(ω) are plotted with R = 50ra and with all remaining parameters as in figure 1. For this separation the positions of the peak centres remain quite different on the frequency scale set by the width ${\gamma }_{s}({\omega }_{2})\approx {\gamma }_{s,0}$. Here, for illustrative purposes, the spectra have both been centred at zero. However, for this value of R the peak heights are effectively the same. Therefore, the spectra have been normailsed by the same peak value ${s}_{0}({\tilde{\omega }}_{0}+{{\rm{\Delta }}}_{12})$.

Standard image High-resolution image
Figure 7.

Figure 7. The relative heights of the peaks in the spectra s(ω) and s0(ω) as a function of the separation R. We have chosen a normalisation factor $n={s}_{0}({\tilde{\omega }}_{0}){| }_{R=50{ra}}$. We have chosen all remaining parameters as in figure 1.

Standard image High-resolution image
Figure 8.

Figure 8. The positions of the peaks in the spectra s(ω) and s0(ω) as functions of the separation R. We have chosen all remaining parameters as in figure 1. The upper subplot shows the ratio of the two peak positions over the same range of values of R, while the lower subplot shows the difference in peak positions over the same range of values of R.

Standard image High-resolution image

5. Conclusions

In this paper we have derived a partially secular master equation valid for arbitrarily separated dipoles within a common radiation field at arbitrary temperature. The equation is intended for the modelling of dipolar systems in which static dipole–dipole interactions are strong compared with the coupling to transverse radiation. This situation can arise in systems of Rydberg atoms and other molecular systems [1023, 4144].

We have shown that the standard gauge-invariant two-dipole master equation can only be derived in gauges other than the multipolar gauge if the direct inter-dipole Coulomb energy is included within the interaction Hamiltonian rather than the unperturbed part. Our arbitrary gauge approach makes a particular limitation of this method clear. Specifically, the usual approach can only be justified when the direct Coulomb interaction is weak along with the coupling to transverse radiation. In situations in which this is not the case our master equation, which is based on a repartitioning of the Hamiltonian into unperturbed and interaction parts, yields significant corrections to previous results. In addition to corrections to the decay of the excited states of the system, we have found corrections to the natural emission spectrum of the initially excited system. In principle, spectroscopy could be used to determine which predictions are closer to the measured values. A possible extension of our result would be to include an external driving Hamiltonian that represents coherent irradiation. The techniques employed here could then be used to calculate the fluorescence spectrum of the driven system.

Acknowledgments

This work was supported by the Engineering and Physical Sciences Research Council grant number EP/N008154/1. We thank Jake Iles-Smith and Victor Jouffrey for useful discussions.

Appendix

A.1. Self-energy contributions and the Gauge-invariance of the single dipole-shift

Here we determine the contribution of self-energy terms to dipole level-shifts and demonstrate that the single-dipole transition shift is gauge-invariant. The self-energy term V(2) is given in equation (4). The shifts arising from this term are divergent in the mode continuum limit ${\omega }_{k}\to \infty $, but this divergence is not unexpected within the non-relativistic dipole approximated treatment. It is typically handled through the introduction of an ultra-violet cut-off. In the treatment of the Lamb-shift in atomic hydrogen the Coulomb gauge self-energy V(2) with αk = 0 is independent of the atomic electron levels and is therefore ignored within the calculation of the measurable shift [35]. In the multipolar gauge V(2) represents a polarisation self-energy term and when its contribution is combined with the remaining contribution to the shift coming from the linear part of the multipolar interaction Hamiltonian one obtains the same result as the Coulomb gauge treatment. In all cases mass renormalisation must also be performed to obtain the correct shift.

In the Coulomb gauge V(2) does not contribute to the master equation transition shift of the two-level dipole, which is the difference between excited and ground state shifts. This is independent of whether the two-level approximation has been made. However, even within the Coulomb gauge it is important to note that one must generally account for all self-energy contributions when explicitly verifying that quantities are gauge-invariant. In particular, to verify that the ground and excited level-shifts are separately gauge-invariant, the contributions ${\langle n| }^{(2)}V| n\rangle ,\,n\,=\,e,g$ must be taken into account.

Using the Hamiltonian in equation (3), the standard Born-Markov master equation has form given in equation (11), in which the decay rate γ is independent of αk. In this equation, the transition shift can be expressed as the difference between excited and ground state shifts as ${\tilde{\omega }}_{0}-{\omega }_{0}={\rm{\Delta }}={\delta }_{e}^{(1)}-{\delta }_{g}^{(1)}$ where

Equation (60)

are αk-dependent. This αk-dependence is due to the lack of any contribution from the self-energy term V(2) in equation (60).

The αk-dependence within the master equation is eliminated when one accounts for the self-energy contributions and the effect of the two-level approximation, recalling that the latter was made after the transformation ${R}_{\{{\alpha }_{k}\}}$ was performed. More specifically it is possible to demonstrate that the single dipole master equation (10) is αk-independent, and that it coincides with equation (11). First we note that we can continue to express the second line in equation (10) in terms of the original partition H = H0 + V. Thus, provided H0 is kept the same for each choice of the αk the dissipative part of the master equation is αk-independent.

It remains to show that when one adds the shift contributions ${\delta }_{e,g}^{(1)}$ coming from the second line in equation (10) to the corresponding self-energy contribution in equation (9) one obtains gauge-invariant total shifts. To this end let us first consider the Coulomb gauge αk = 0. The total excited and ground state shifts are

Equation (61)

The components ${\delta }_{e,\mathrm{CG}}^{(1)}$ are obtained by setting αk = 0 in equation (60), while the remaining component

Equation (62)

is the Coulomb gauge self-energy shift due to the ${{\bf{A}}}_{{\rm{T}}}^{2}$ part of the Coulomb gauge interaction Hamiltonian. Since this term is independent of the dipole, the shift ${\delta }_{\mathrm{CG}}^{(2)}$ is the same for the ground and excited levels. The single-dipole transition shift Δ given in equation (12) in the main text can be expressed in terms of Coulomb gauge shifts as

Equation (63)

More generally, for arbitrary αk the total ground and excited state level shifts are denoted δe, g. In what follows we will show that

Equation (64a)

and

Equation (64b)

from which it follows using equation (63) that δe − δg = Δ for all choices of αk.

In order to show that equations (64a) and (64b) hold we must carefully account for the two-level approximation, which was performed after the gauge transformation ${R}_{\{{\alpha }_{k}\}}$. Let us consider a general shift of the m'th level of the dipole with the form

Equation (65)

where v is arbitrary. If we restrict ourselves to two levels e and g, and if m = e in the above, then the sum includes only one other level n = g, so we get for the shift

Equation (66)

where ${\omega }_{0}:= {\omega }_{{eg}}=-{\omega }_{{ge}}$ and ${\bf{d}}:= {{\bf{d}}}_{{eg}}={{\bf{d}}}_{{ge}}^{* }$. If instead m = g then the shift is

Equation (67)

The shift is clearly different in the m = e and m = g cases when considering a two-level system. However, for an infinite-dimensional dipole the shift is independent of m being given by

Equation (68)

where we have made use of the identity

Equation (69)

The difference between the finite and infinite-dimensional cases arises because the proof of equation (69) rests directly on the CCR algebra $[{r}_{i},{p}_{j}]={\rm{i}}{\delta }_{{ij}}$, which can only be supported in infinite-dimensions. When the algebra is truncated to su(2), the same shift comes out level-dependent. Since the gauge transformation ${R}_{\{{\alpha }_{k}\}}$ is made on the infinite-dimensional dipole it is necessary to employ equation (68) in order to exhibit gauge-invariance of the shifts. Thus, in order to get the correct level-shifts within the two-level approximation, when dealing with the excited level shift m = e we use equations (66) and (68), which imply

Equation (70)

but when dealing with the ground level shift m = g we use equations (67) and (68), which imply

Equation (71)

We now proceed to verify that equations (64a) and (64b) hold. The complete shifts δe, g are obtained by taking the shifts in equation (60) and adding their respective self-energy contributions. Subtracting ${\delta }_{\mathrm{CG}}^{(2)}$ in equation (62) from δe and subsequently using equation (70), which is appropriate for the excited state shift, we obtain

Equation (72)

Using equation (6) we express the bracket within the integrand in this expression in terms of αk. The part independent of Nk is

Equation (73)

In this expression we identify the coefficient of ${\alpha }_{k}^{2}$ as

Equation (74)

and the coefficient of 2αk as

Equation (75)

Thus, equation (73) is αk-independent. The remaining part is

Equation (76)

The Nk-dependent parts of ${\delta }_{e}-{\delta }_{\mathrm{CG}}^{(2)}$ can be dealt with in a similar manner. The coefficient of ${\alpha }_{k}^{2}$ in the Nk-dependent part of the bracket within the integrand of the expression for ${\delta }_{e}-{\delta }_{\mathrm{CG}}^{(2)}$ is

Equation (77)

Similarly, the coefficient of αk is

Equation (78)

The remaining Nk-dependent part is

Equation (79)

Combining equations (73) and (79) we obtain the αk-independent result

Equation (80)

which completes the proof of equation (64a).

The shift appearing on the left-hand side of equation (64b) is found using equation (71) to be

Equation (81)

Similar calculations to those above for the excited state yield the final result

Equation (82)

This completes the proof that the transition shift δe − δg is αk-independent and that it equals Δ given in equation (12).

We remark that the need to account for the self-energy contributions along with the effect of the two-level truncation is a peculiarity of the single-dipole shift term Δ. The same need does not arise in the case of the remaining coefficients γ, γ12 and Δ12 in the standard two-dipole master equation (20). These coefficients are immediately seen to coincide with gauge-invariant matrix elements.

A.2. Calculation of the standard joint shift

The joint shift Δ12 resulting from the arbitrary gauge master equation derivation is given by

Equation (83)

Using equation (6) all αk-dependence can be shown to vanish in the same way as with the single-dipole shifts dealt with in appendix A.1. The final result is

Equation (84)

Evaluating the angular integral and polarisation summation yields

Equation (85)

The integral is regularised by introducing a convergence factor ${{\rm{e}}}^{-\epsilon {\omega }_{k}}$ under the integral, and finally taking the limit $\epsilon \to {0}^{+}$. We substitute τij given in equation (22) into equation (85) and evaluate the resulting integrals term by term. The integral arising from the first part of τij is

Equation (86)

We now make the substitution z = ωk R, and make a suitable choice of contour C such that by the residue theorem we obtain

Equation (87)

Thus, the part of the shift Δ12 arising from the first part (R−1 component) of τij is

Equation (88)

which we recognise as the R−1 component of Δ12 in equation (21). The remaining parts of equation (85) can be evaluated in a similar way, which yields the final result given in equation (21).

A.3. Method of calculation of the spectrum

We denote the dynamical map governing evolution of the reduced density matrix by F(t, t'), which is such that F(t, t')ρ(t') = ρ(t). A general two-time correlation function for arbitrary system observables O and O' can be written [36]

Equation (89)

We define the super-operator Λ by $\dot{\rho }(t)={\rm{\Lambda }}\rho (t)$ using the master equation (equation (20) or equation (40)). Since Λ is time-independent, from the initial condition $F(0,0)\equiv I$ we obtain the general solution $F(t,t^{\prime} )={{\rm{e}}}^{{\rm{\Lambda }}(t-t^{\prime} )}$. For convenience we write F(t,0) = F(t), so that $F(t,t^{\prime} )=F(t-t^{\prime} )$.

In order to calculate the two-time correlation functions we first find a concrete representation of the maps Λ and F(t). For this purpose we introduce a basis of operators denoted $\{{x}_{i}:i=1,\,\ldots ,\,16\}$, which is closed under Hermitian conjugation. The trace defines an inner-product $\langle O,O^{\prime} \rangle =\mathrm{tr}({O}^{\dagger }O^{\prime} )$ with respect to which the basis xi is assumed to be orthonormal. We identify two resolutions of unity as ${\sum }_{i}\mathrm{tr}({x}_{i}^{\dagger }\cdot ){x}_{i}=I={\sum }_{i}\mathrm{tr}({x}_{i}\cdot ){x}_{i}^{\dagger }$, which imply that any operator O can be expressed as $O={\sum }_{i}\mathrm{tr}({x}_{i}^{\dagger }O){x}_{i}={\sum }_{i}\mathrm{tr}({x}_{i}O){x}_{i}^{\dagger }$. Expressing both sides of the equation $\dot{F}(t)={\rm{\Lambda }}F(t)$ in the basis xi yields the relation

Equation (90)

where

Equation (91)

Equation (90) can be written in the matrix form $\dot{{\bf{F}}}={\boldsymbol{\Lambda }}{\bf{F}}(t)$ whose solution is expressible in the matrix exponential form ${\bf{F}}(t)={{\rm{e}}}^{{\boldsymbol{\Lambda }}t}$. A general two-time correlation function of system operators can then be expressed using equation (89) as

Equation (92)

where ${O}_{i}=\mathrm{tr}({{Ox}}_{i})$, ${\rho }_{i}=\mathrm{tr}({x}_{i}^{\dagger }\rho )$ and ${O}_{{ij}}^{{\prime} }=\mathrm{tr}({x}_{i}^{\dagger }O^{\prime} {x}_{j})$. Choosing the basis $\{{x}_{i}\}$ to be the operators obtained by taking the outer products of the bare states $| n,m\rangle ,\,(n,m=e,g)$, the above machinery can be used to obtain the correlation function (50).

A.4. Derivation of spectrum associated with the new master equation

The mode expansion for the transverse field canonical momentum ${{\boldsymbol{\Pi }}}_{{\rm{T}}}$ is

Equation (93)

This operator represents a different physical observable for each choice of ${\alpha }_{k}$, because it does not commute with the generalised gauge transformation ${R}_{\{{\alpha }_{k}\}}$. Similarly the photonic operators ${a}_{{\bf{k}}\lambda }$ are implicitly different for each choice of αk. In the multipolar gauge the field canonical momentum coincides with the total electric field away from the sources; ${{\boldsymbol{\Pi }}}_{{\rm{T}}}({\bf{x}})=-{\bf{E}}({\bf{x}}),\,{\bf{x}}\ne {{\bf{R}}}_{\mu }$. The positive frequency (annihilation) and negative frequency (creation) components of the electric field are therefore defined for ${\bf{x}}\ne {{\bf{R}}}_{\mu }$ by

Equation (94)

where akλ is the photon annihilation operator within the multipolar gauge. For a system of two dipoles the integrated Heisenberg equation for the multipolar photon annihilation operator yields the source component

Equation (95)

Since the dipole moment operators dμ commute with the transformation ${R}_{\{{\alpha }_{k}\}}$ they represent the same physical observable for each choice of αk. This implies that equation (95) can be expressed in terms of Coulomb gauge raising and lowering operators ${\sigma }_{\mu }^{\pm }$ in the two-level approximation, despite the implicit difference between these operators and their counterparts defined within the multipolar gauge. We subsequently express the Coulomb gauge operators ${\sigma }_{\mu }^{\pm }$ in the dressed basis $| {\epsilon }_{n}\rangle $ to obtain

Equation (96)

where ${\sigma }_{\mu ,{nm}}={\sigma }_{\mu ,{nm}}^{+}+{\sigma }_{\mu ,{nm}}^{-}$, ${\epsilon }_{{nm}}={\epsilon }_{n}-{\epsilon }_{m}$, and ${\theta }_{{nm}}=| {\epsilon }_{n}\rangle \langle {\epsilon }_{m}| $. We now perform a rotating-wave approximation, which eliminates terms that are rapidly oscillating within the interaction picture defined by the dressed Hamiltonian Hd given in equation (30). Substitution of the resulting expression into equation (94) yields in the mode continuum limit

Equation (97)

where ${\tilde{\theta }}_{{nm}}(t^{\prime} )$ denotes the operator ${\theta }_{{nm}}(t^{\prime} )$ transformed into the interaction picture with respect to Hd, and ${{\bf{r}}}_{\mu }={\bf{x}}-{{\bf{R}}}_{\mu }$. Performing the angular integration and polarisation summation, and retaining only the radiative component yields

Equation (98)

Finally, using the Markov approximation

Equation (99)

valid for a suitably behaved function f, we obtain the final result equation (55) given in the main text.

To calculate the spectrum according to our master equation (40) we choose the basis of operators {xi} used within the general method laid out in appendix A.3 as that obtained by taking the outer-products of the basis states $| {\epsilon }_{n}\rangle $ given in equation (32). Using equation (92) we define the array of correlation functions

Equation (100)

where p is restricted to values such that xp is diagonal, and where the matrix ${{\bf{X}}}_{m}$ has elements ${({{\bf{X}}}_{m})}_{{jk}}=\mathrm{tr}({x}_{j}^{\dagger }{x}_{m}{x}_{k})$. Taken in the symmetric state θ33 the correlations appearing in equation (56) are all elements of the array ${C}_{{nm}}(t,t^{\prime} )$, which is given by equation (100) with xp = θ33. We choose a labelling whereby the xi are given by

Equation (101)

In this case the only non-zero off-diagonal element of ${C}_{{nm}}(t,t^{\prime} )$ is ${C}_{\mathrm{1,11}}(t,t^{\prime} )$ where x1 = θ11 and x11 = θ33. Furthermore the diagonal elements ${C}_{{nn}}(t,t^{\prime} )$ are zero unless n is odd. It follows that the only non-vanishing correlations in equation (56) are C33(t, t') and C77(t, t'). Moreover, since ${\sum }_{\mu ,\nu =1}^{2}{\sigma }_{\mu ,32}{\sigma }_{\nu ,23}=0$ only the term involving C33(t, t') contributes. This term describes correlations associated with the symmetric to ground state transition and is given by equation (57) in the main text. Integration of this correlation function according to equation (46) then yields the spectrum in equation (59).

Please wait… references are loading.