Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Beyond photon pairs—nonlinear quantum photonics in the high-gain regime: a tutorial

Open Access Open Access

Abstract

Integrated optical devices will play a central role in the future development of nonlinear quantum photonics. Here we consider the generation of nonclassical states of light within them with a focus on Gaussian states beyond the low-gain, single photon pair regime accurately described by perturbation theory. Starting from the solid foundation provided by Maxwell’s equations, we then move to applications by presenting a unified formulation that allows for a comparison of stimulated and spontaneous experiments in ring resonators and nanophotonic waveguides and leads directly to the calculation of the quantum states of light generated in high-gain nonlinear quantum photonic experiments.

© 2022 Optica Publishing Group

1. Introduction

Spontaneous emission, where a single photon is emitted as an atom decays from an excited state to its ground state, was the earliest source of nonclassical light. Within a decade of the invention of the laser, however, researchers began to explore the possibility of using parametric fluorescence to generate other nonclassical states [1,2]. Here the state of a material medium is left unchanged after the passage of a strong pump pulse, but because of optical nonlinearities it is possible for a pair of new photons to be generated. Much of the interest has focused on two processes. In one of them, called spontaneous parametric downconversion (SPDC), a pump photon splits into two daughter photons, usually referred to as “signal” and “idler,” in what can be thought of as a “photon fission” process. This is sufficiently strong only in materials without inversion symmetry, where a $\chi _2$ nonlinear susceptibility is present. In the other process, called spontaneous four-wave mixing (SFWM), what can be thought of as an “elastic scattering” between two pump photons occurs to make two new photons, again usually referred to as “signal” and “idler,” with one of the new photons at a frequency higher than that of the pump photons, and the other at a frequency lower than that. This process is governed by a $\chi _3$ nonlinear susceptibility, and so in principle can occur in any material medium. In both cases the pairs of photons produced are typically entangled, and thus of interest in quantum information processing protocols.

It was soon realized that not only could a strong enough pump pulse generate a pair of photons with significant probability, but that multiple pairs might also be generated. The quantum superposition of these states with different numbers of pairs of photons leads to a state of light called squeezed, in that the uncertainty in the amplitude of one quadrature is suppressed below the usual quantum limit, whereas the uncertainty in the amplitude of the other is correspondingly increased so as not to violate the uncertainty relation [3,4]. These squeezed states are central to current efforts in quantum information processing based on continuous variables [512] and to quantum metrology [1315].

One may think that the low efficiency of SPDC or SFWM is fundamentally related to their quantum nature. However, this is not the case, for it rather depends on the weakness of the nonlinear light–matter interaction. In this respect, there are at least two strategies to tackle this problem, either by increasing the nonlinear interaction length, e.g., by working with bigger and bigger crystals, or by enhancing the intensity of the electromagnetic field interacting with the matter.

In the case of SFWM, which is the weaker of the two above-mentioned nonlinear processes, significant progress was made at the beginning of this century with the use optical fibers [16,17]. These allow one to simultaneously enhance the electromagnetic field intensity by leveraging the transverse light confinement and, at the same time, increasing the nonlinear interaction lengths from a few millimeters to a few kilometers.

More recently, studies of squeezed light and its generation have intersected with impressive improvements in the design and fabrication of integrated photonic structures involving channel waveguides and resonators. In these systems, the spatial and temporal confinement of light can enhance the nonlinear interaction strength up to ten orders of magnitude with respect to what can be achieved in bulk systems, leading to efficient parametric fluorescence with continuous wave (CW) excitation at milliwatt pump powers.

One important consequence is that the generation of nonclassical light “beyond photon pairs” is becoming commonplace [1826]. A second consequence is that a more detailed comparison of theory and experiment is now possible, because implementations of well-characterized integrated structures do not suffer from the uncertainties that can plague experiments with bulk systems, such as details of beam shape and propagation.

From a theoretical point of view, the description of the nonlinear light–matter interaction in micro- and nanostructures is far more complex than that required in bulk systems, starting with the quantization of the electromagnetic field and moving on to the construction of the Hamiltonian describing a system that can be composed of several optical elements. Finally, the strong enhancement of the nonlinear response of the system necessitates tools able to describe fast, nontrivial dynamics, which cannot always by approached using perturbation theories.

In this tutorial we present an overview of the physics of nonlinear quantum optics “beyond photon pairs.” We begin in Sec. 2 with a treatment of the quantization of light in integrated photonic structures. Although correct treatments of this subject can of course be found in the literature, there are pitfalls in adding to the linear Hamiltonian what one might think would be the obvious nonlinear contribution. We discuss the subtleties involved, and detail the mode expansion and nonlinear Hamiltonians for treating channel waveguides and microring resonators, the latter being the cavity structure we use as an example in this tutorial. In Sec. 3 we consider general “squeezing Hamiltonians” that result when the usual classical approximation is made for the pump pulse; these Hamiltonians then only involve sums of products of pairs of operators. We consider low-gain solutions, and then develop a general framework in which the solution of the Heisenberg equations of motion can be leveraged to build the full ket describing the squeezed state of light. This can be used to treat both spontaneous and stimulated parametric fluorescence, and we present results for both photon-number and homodyne statistics of the generated light. In Sec. 4 we focus on waveguides, and use a particular example to illustrate the low-gain regime and the joint spectral amplitude (JSA) of pairs of photons that can be generated, and how it is modified upon entering the high-gain region. In Sec. 5 we turn to microring resonators, and consider in detail both single- (SP-) and dual-pump (DP-)SFWM, as well as SPDC. With all this as background we turn in Sec. 6 to connections between classical and quantum nonlinear optics, and how the spontaneous processes we have considered can be understood as “stimulated” by vacuum power fluctuations. In Sec. 7 we move beyond the world of coherent states and squeezed states, both of which fall in the category of Gaussian states. We present a leading strategy, based on “matrix product states” (MPSs), to describe states more complicated than Gaussian. As moving beyond information processing that involves only Gaussian states and measurements is necessary to achieve universal quantum computing [27,28], the description and characterization of non-Gaussian states is a current area of active research, with many different approaches being explored [2936]. Indeed, it can be hoped that at some point in the future a tutorial that could be considered a successor to this one, and might be called “Beyond Gaussian states,” will become available. Finally, we conclude in Sec.8.

2. Quantization in Integrated PhotonicStructures

In this section, we develop the quantization of nonlinear optics in an approach particularly suitable for treating integrated photonic structures. In linear quantum optics the usual treatments are naturally based on treating the electric and magnetic fields $\boldsymbol{E}(\boldsymbol{r},t)$ and $\boldsymbol{B}(\boldsymbol{r},t)$ as fundamental, but a straightforward extension of this to nonlinear optics is invalid [37]; it leads to incorrect expressions for the Hamiltonian that can be found even in well-respected references [38]. In fact, for nonlinear quantum optics a treatment going back to the early work of Born and Infeld [39], which treats $\boldsymbol{D}(\boldsymbol{r},t)$ and $\boldsymbol{B}(\boldsymbol{r},t)$ as the fundamental fields, seems the simplest way to develop the Hamiltonian treatment directly without first introducing an underpinning Lagrangian framework and building up the Hamiltonian treatment from that, as is done by, e.g., Drummond and Hillery [40]. In Sec. 2.1 we illustrate the problems that can arise if a Hamiltonian treatment of quantum nonlinear optics is based on $\boldsymbol{E}(\boldsymbol{r},t)$ and $\boldsymbol{B}(\boldsymbol{r},t)$ as the fundamental fields, and as a preamble to nonlinear quantum optics we develop linear quantum optics based on $\boldsymbol{D}(\boldsymbol{r},t)$ and $\boldsymbol{B}(\boldsymbol{r},t)$ as the fundamental fields in Sec. 2.2.

Of particular interest are channel waveguide and ring resonator structures, and after introducing the general quantization procedure we treat those special cases in Sec. 2.2.4. Initially we neglect dispersion, treating the relative dielectric constant $\varepsilon _{1}(\boldsymbol{r})$ as independent of frequency, but in Sec. 2.2.5 we indicate how dispersion can be included in frequency regions where absorption can be neglected, such as below the band-gap of a semiconductor, which is the usual regime of interest for integrated photonic structures. In Sec. 2.3 we present a treatment of the coupling between channel waveguides and ring resonators in the linear regime that will be useful in later sections, and in Sec. 2.4 we include nonlinearities by introducing the usual nonlinear susceptibilities. The particular results for the Hamiltonians describing self-phase modulation (SPM) and cross-phase modulation (XPM) in channels and rings are detailed in Secs. 2.4.1 and 2.4.2, respectively; in later sections of this tutorial we introduce the appropriate Hamiltonians for other nonlinear processes. Of course, other treatments are possible. One can start from the classical equations relating the nonlinear polarization and the electric field to each other, and “elevate” the classical fields to their quantum counterparts [41]. Or one could model a crystal as a lattice of isolated molecules and work out the nonlinear response of the crystal from that model [42]. However, we feel that the treatment given here provides both the trustworthiness and the generality appropriate for the problems at hand.

Throughout this tutorial we generally assume that the relative dielectric constant $\varepsilon _{1}(\boldsymbol{r})$, or, if dispersion effects are included, $\varepsilon _{1}(\boldsymbol{r};\omega )$, can be treated as a scalar. Structures of interest, such as those involving lithium niobate, are clearly exceptions to this. The generalization to a relative dielectric tensor is straightforward, but we do not do it explicitly here because it complicates the formulas and, we believe, makes the physics we want to highlight in this tutorial less accessible.

2.1 Trouble with $E$

The quantization of the electromagnetic field in vacuum is a standard exercise in elementary quantum optics, and is naturally formulated in terms of the electric and magnetic fields $\boldsymbol{E}(\boldsymbol{r},t)$ and $\boldsymbol{B}(\boldsymbol{r},t)$ [43]. The generalization to include a uniform, frequency-independent dielectric constant is just as easy, and including even a position dependent relative dielectric constant $\varepsilon _{1}(\boldsymbol{r})$, where the constitutive relations are taken to be

$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})\boldsymbol{E}(\boldsymbol{r},t),\\ & \boldsymbol{B}(\boldsymbol{r},t)=\mu_{0}\boldsymbol{H}(\boldsymbol{r},t), \end{aligned}$$
with $\varepsilon _{1}(\boldsymbol{r})$ real, is straightforward.

For our discussion in this section we just quote the results for a dielectric ring shown in Fig. 1; we give a derivation of the results in Sec. 2.2.4. Cylindrical coordinates $z$, $\rho =\sqrt {x^{2}+y^{2}}$, and the angle $\phi$ are the natural variables; it is convenient to introduce a nominal radius $R$ and use $\zeta =R\phi$ in place of $\phi$. Using the notation $\boldsymbol{r}_{\perp }=(\rho,z)$, a volume element is $\textrm{d}\boldsymbol{r}=\textrm{d}\boldsymbol{r}_{\perp }\textrm{d}\zeta$, where $\textrm{d}\boldsymbol{r}_{\perp }=R^{-1}\rho \textrm{d}\rho \textrm{d}z$, and $\zeta$ varies from $0$ to $\mathcal {L}=2\pi R$. With the relative dielectric constant $\varepsilon _{1}(\boldsymbol{r})=\varepsilon _{1}(\boldsymbol{r}_{\perp })$ in the ring large enough to confine modes, the modes will be labeled by integer quantum numbers $m_{J}$, where the wavenumber associated with propagation around the ring is $\kappa _{J}=2\pi m_{J}/\mathcal {L}$; for the moment we restrict ourselves to one relevant transverse field structure for each $\kappa _{J}$, identify its frequency by $\omega _{J}$, and consider only these bound modes. Then introducing raising and lowering operators $c_{J}^{\dagger }$ and $c_{J}$ for each mode, neglecting zero-point energies the Hamiltonian is

$${H}_{\textrm{ring}}^{\textrm{L}}=\sum_{J}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J},$$
and the electric field operator is written in terms of them as
$$\boldsymbol{E}(\boldsymbol{r})=\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2\mathcal{L}}}c_{J}\mathsf{e}_{J}(\boldsymbol{r}_{{\perp}})e^{i\kappa_{J}\zeta}+\textrm{H.c.},$$
where the $\mathsf{e}_{J}(\boldsymbol{r}_{\perp })$ are determined so that the Heisenberg operator corresponding to (3) satisfies Maxwell’s equations, and then they are appropriately normalized; here “+H.c.” indicates that the Hermitian conjugate of the preceding should be added. In writing (3) we have assumed for simplicity that for the modes of interest the amplitudes $\mathsf{e}_{J}(\boldsymbol{r}_{\perp })$ lie in the $z$ direction; we write down more general expressions later.

 figure: Figure 1.

Figure 1. Schematic representation of an isolated ring resonator. The inset shows the typical field profile of the fundamental transverse-electric mode supported by the bent waveguide. In the simulation the waveguide is 800 nm wide and 400 nm high, is made of silicon nitride fully clad in silica, and has a bending radius of 10 $\mu$m. The simulation is performed at a wavelength of 1550 nm. The contour plot accounts for the $E_{\zeta }$ component, whereas the arrow field represents $\boldsymbol{E}_{\boldsymbol{r}_\perp }$

Download Full Size | PDF

This is all uncontroversial. All of the possible derivations of (2) and (3) rely on identifying the energy density $\mathfrak {h}(\boldsymbol{r})$ at a point $\boldsymbol{r}$ in space, and noting that for a change in the fields at this point the energy density changes according to [44]

$$d\mathfrak{h}=\boldsymbol{H}\cdot \textrm{d}\boldsymbol{B}+\boldsymbol{E}\cdot \textrm{d}\boldsymbol{D},$$
and so for the constitutive relations of (1) we find an expression for the classical energy density,
$$\mathfrak{h}(\boldsymbol{r},t)=\frac{1}{2\mu_{0}}\boldsymbol{B}(\boldsymbol{r},t)\cdot\boldsymbol{B}(\boldsymbol{r},t) +\frac{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}{2}\boldsymbol{E}(\boldsymbol{r},t)\cdot\boldsymbol{E}(\boldsymbol{r},t),$$
with the volume integral of this equal to the total energy. In accordance with the prescription for canonical quantization, the operator version of that integral is set equal to ${H}_{\textrm{ring}}^{\textrm{L}}$ after the subtraction of the zero-point energy.

However, what would seem to be a straightforward extension of this to include nonlinear effects, where then

$${H}_{\textrm{ring}}={H}_{\textrm{ring}}^{\textrm{L}}+{H}_{\textrm{ring}}^{\textrm{NL}},$$
leads to erroneous results. To see the difficulty, note that with the approximation that the $\mathsf{e}_{J}(\boldsymbol{r}_{\perp })$ of interest lie purely in the $z$ direction, then $\boldsymbol{E}(\boldsymbol{r})$ will also be, thus $\boldsymbol{B}(\boldsymbol{r})$ can correspondingly be approximated as lying purely in the radial direction, and we denote those components by $E(\boldsymbol{r})$ and $B(\boldsymbol{r})$, respectively. Assuming the usual series expansion for the nonlinear response up to some power $N$,
$$\begin{aligned}& D(\boldsymbol{r},t)=\epsilon_{0}\sum_{n=1}^{N}\varepsilon_{n}(\boldsymbol{r})E^{n}(\boldsymbol{r},t),\\ & B(\boldsymbol{r},t)=\mu_{0}H(\boldsymbol{r},t), \end{aligned}$$
from (4) we find
$$\mathfrak{h}(\boldsymbol{r},t)=\frac{1}{2\mu_{0}}B^{2}(\boldsymbol{r},t)+\epsilon_{0}\sum_{n=1}^{N}\frac{n}{n+1}\varepsilon_{n}(\boldsymbol{r})E^{n+1}(\boldsymbol{r},t).$$
Considering only a third-order nonlinearity $(n=3)$ as well as the usual linear ($n=1$) response, a natural argument would be to note that the magnetic term and the first electric term in (8) give the usual linear terms (2), when the usual expressions (3) for the electric and magnetic fields are used, and so taking the integral of (8) over all space we can identify
$${H}_{\textrm{ring}}^{\textrm{NL}}=\frac{3}{4}\epsilon_{0}\int\varepsilon_{3}(\boldsymbol{r})E^{4}(\boldsymbol{r})\textrm{d}\boldsymbol{r}.$$
Using the electric field expression (3) in this expression (9), and combining it with ${H}_{\textrm{ring}}^{\textrm{L}}$ to form the total Hamiltonian (6), it would seem we could obtain the nonlinear equations for the dynamics of the $c_{J}$.

However, this leads to nonsense. To see that, consider just one mode $P$ and its nonlinear interaction with itself. Neglecting normal ordering corrections in the nonlinear term in (9), we find that the equation for $c_{P}$ is

$$\frac{\textrm{d}{c_{P}}}{\textrm{d}t}={-}i\omega_{P}c_{P}-\frac{9i}{\hbar}\left(\frac{\hbar\omega_{P}}{2\mathcal{L}}\right)^{2}Kc_{P}^{{\dagger}}c_{P}c_{P},$$
where
$$K=\epsilon_{0}\int\varepsilon_{3}\left(\boldsymbol{r}\right)\left[\mathsf{e}_{P}^{*}\left(\boldsymbol{r}_{{\perp}}\right)\right]^{2} \left[\mathsf{e}_{P}\left(\boldsymbol{r}_{{\perp}}\right)\right]^{2}\textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta.$$
In the classical limit, where $c_{P}(t)$ and $c_{P}^{\dagger }(t)$ ($=c_{P}^{*}(t))$ are just variables, we see that this leads to a prediction of a shift in frequency
$$\omega_{P}\rightarrow\omega_{P}+\frac{9}{\hbar}\left(\frac{\hbar\omega_{P}}{2\mathcal{L}}\right)^{2}Kc_{P}^{{\dagger}}c_{P}.$$
That is, for a positive $\varepsilon _{3}(\boldsymbol{r}$), for which $K$ is positive, the frequency is predicted to increase with increasing intensity. Yet that is incorrect: initially we can always write the frequency $\omega _{P}$ as $c\kappa _{P}/n_{\text {eff}}$, where $n_{\text {eff}}$ is an effective index of refraction of the mode in the ring, and a positive $\varepsilon _{3}(\boldsymbol{r})$ will certainly lead to an increase with intensity of the effective index of the ring, and thus to a frequency that decreases with increasing intensity. In fact, rather than (10), the correct shift in frequency is given by
$$\omega_{P}\rightarrow\omega_{P}-\frac{3}{\hbar}\left(\frac{\hbar\omega_{P}}{2\mathcal{L}}\right)^{2}Kc_{P}^{{\dagger}}c_{P},$$
as we show in Sec. 2.4.2. Thus, (12) fails to correctly describe both the size of the effect and its sign, and it would lead to the wrong prediction, for example, of the group velocity dispersion (GVD) regime in which soliton propagation would exist.

What has gone wrong? The Hamiltonian density (9) is certainly equal to the energy density, as it should be for canonical quantization. However, the use of the form (3) of the electric and magnetic fields in that Hamiltonian, together with the usual commutation relations for the operators $c_{P}$ and $c_{P}^{\dagger }$, does not lead to the correct equations, Maxwell’s equations, in the nonlinear regime. This is apparent because, very generally, the equation for the Heisenberg operator $\boldsymbol{B}(\boldsymbol{r},t)$ would be

$$\frac{\partial\boldsymbol{B}(\boldsymbol{r},t)}{\partial t}=\frac{1}{i\hbar}\left[\boldsymbol{B}(\boldsymbol{r},t),{H}_{\textrm{ring}}^{\textrm{L}}\right]+\frac{1}{i\hbar}\left[\boldsymbol{B}(\boldsymbol{r},t),{H}_{\textrm{ring}}^{\textrm{NL}}\right].$$
Keeping only the first commutator on the right we obtain Faraday’s law, $\dot {\boldsymbol{B}}(\boldsymbol{r},t)=-\nabla \times \boldsymbol{E}(\boldsymbol{r},t)$; the theory was built to give this in the linear regime. However, it is easy to see that the second commutator on the right is nonvanishing, and then leads to a violation of Faraday’s law when nonlinear effects are included.

A hint of the problem can be gleaned from the differential form (4) for the change in the energy density, which suggests that $\boldsymbol{B}$ and $\boldsymbol{D},$ rather than $\boldsymbol{B}$ and $\boldsymbol{E}$, should be thought of as the fundamental fields. This is in contrast to a Lagrangian approach, where the differential form for the change in the Lagrangian density is given [39] by

$$\textrm{d}\mathfrak{L=}-\boldsymbol{D}\cdot \textrm{d}\boldsymbol{E}+\boldsymbol{H}\cdot \textrm{d}\boldsymbol{B}.$$
Indeed, one way to proceed is to return to the Lagrangian formulation, include the nonlinear interactions, and “rebuild” the Hamiltonian formulation from that starting point. This has been done at a microscopic level involving two- or many-level atoms [4547], and from a macroscopic perspective using susceptibilities [48,49]. For an overview, see Drummond and Hillery [40]. Alternately, as was first pointed out by Born and Infeld [39], one can begin with $\boldsymbol{B}$ and $\boldsymbol{D}$ as the fundamental fields and directly introduce the Hamiltonian formulation, without relying on an earlier Lagrangian formulation. When the focus is on the Hamiltonian framework, as it is often in quantum optics, this is a simpler strategy; it is the one we follow here [50]. In the next section, we implement this in the linear regime, and in Sec. 2.4 we extend it to nonlinear optics. For a recent tutorial on these issues, see Raymer [51].

2.2 Linear Quantum Optics

We begin by adopting the constitutive relations (1), assuming a real relative dielectric constant $\varepsilon _{1}(\boldsymbol{r})$ that is dispersionless but arbitrarily position dependent. In Sec. 2.2.5 we generalize to include material dispersion, which can have important consequences for experiments in quantum photonics.

2.2.1 Modes

We begin with the classical Maxwell equations

$$\begin{aligned} & \frac{\partial\boldsymbol{B}(\boldsymbol{r},t)}{\partial t}={-}\boldsymbol{\nabla}\times\boldsymbol{E}(\boldsymbol{r},t),\\ & \frac{\partial\boldsymbol{D}(\boldsymbol{r},t)}{\partial t}=\boldsymbol{\nabla}\times\boldsymbol{H}(\boldsymbol{r},t),\\ & \boldsymbol{\nabla}\cdot\boldsymbol{B}(\boldsymbol{r},t)=0,\\ & \boldsymbol{\nabla}\cdot\boldsymbol{D}(\boldsymbol{r},t)=0 \end{aligned}$$
and look for a solutions in the linear regime where the constitutive relations are (1). Looking for solutions of the form
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\boldsymbol{D}_{\alpha}(\boldsymbol{r})e^{{-}i\omega_{\alpha}t}+\textrm{c.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\boldsymbol{B}_{\alpha}(\boldsymbol{r})e^{{-}i\omega_{\alpha}t}+\textrm{c.c.}, \end{aligned}$$
with $\omega _{\alpha }$ positive, we have a solution of (16) if
$$\begin{aligned} & \boldsymbol{\nabla}\times\left[\frac{\boldsymbol{\nabla}\times\boldsymbol{B}_{\alpha}(\boldsymbol{r})}{\varepsilon_{1}(\boldsymbol{r})}\right]=\frac{\omega_{\alpha}^{2}}{c^{2}}\boldsymbol{B}_{\alpha}(\boldsymbol{r}),\\ & \boldsymbol{\nabla}\cdot\boldsymbol{B}_{\alpha}(\boldsymbol{r})=0,\\ & \boldsymbol{D}_{\alpha}(\boldsymbol{r})=\frac{i}{\omega_{\alpha}\mu_{0}}\boldsymbol{\nabla}\times\boldsymbol{B}_{\alpha}(\boldsymbol{r}). \end{aligned}$$
The first of these is the “master equation” familiar from work on photonic crystals [52]. Whatever the functional form of $\varepsilon _{1}(\boldsymbol{r}),$ we refer to a solution $\left (\boldsymbol{D}_{\alpha }\left (\boldsymbol{r}\right ),\boldsymbol{B}_{\alpha }\left (\boldsymbol{r}\right )\right )$ of (18) as a “mode” of the system. Note that for every such mode, we can identify another mode
$$\left(\boldsymbol{D}_{\alpha'}\left(\boldsymbol{r}\right),\boldsymbol{B}_{\alpha'}\left(\boldsymbol{r}\right)\right)= \left(-\boldsymbol{D}_{\alpha}^{*}\left(\boldsymbol{r}\right),\boldsymbol{B}_{\alpha}^{*}\left(\boldsymbol{r}\right)\right)$$
with $\omega _{\alpha '}=\omega _{\alpha }$. The organization of all modes into pairs, such as $\left (\boldsymbol{D}_{\alpha }\left (\boldsymbol{r}\right ),\boldsymbol{B}_{\alpha }\left (\boldsymbol{r}\right )\right )$ and $\left (\boldsymbol{D}_{\alpha '}\left (\boldsymbol{r}\right ),\boldsymbol{B}_{\alpha '}\left (\boldsymbol{r}\right )\right )$, can be done just as was outlined earlier for acoustic modes [53], but note that different sign conventions have been used in the past for partner modes. Following the notation there we refer to $\alpha '$ as the “partner mode” of $\alpha$.

We begin by looking at modes that are also chosen to satisfy periodic boundary conditions over a normalization volume $V$. Then the operator $\mathcal {M}$,

$$\mathcal{M}(\cdots)\equiv\boldsymbol{\nabla}\times\left[\frac{\boldsymbol{\nabla}\times(\cdots)}{\varepsilon_{1}(\boldsymbol{r})}\right]$$
is Hermitian, in that
$$\int_{V}\boldsymbol{A}^{*}(\boldsymbol{r})\cdot\mathcal{M}(\boldsymbol{C}(\boldsymbol{r}))\textrm{d}\boldsymbol{r}=\int_{V}\left(\mathcal{M}(\boldsymbol{A}(\boldsymbol{r}))\right)^{*}\cdot\boldsymbol{C}(\boldsymbol{r})\textrm{d}\boldsymbol{r},$$
for any such periodic functions $\boldsymbol{A}(\boldsymbol{r})$ and $\boldsymbol{C}(\boldsymbol{r})$. From this it follows immediately that
$$\begin{aligned} & \int_{V}\boldsymbol{B}_{\alpha}^{*}(\boldsymbol{r})\cdot\boldsymbol{B}_{\beta}(\boldsymbol{r})\textrm{d}\boldsymbol{r}=0,\\ & \omega_{\alpha}\neq\omega_{\beta}, \end{aligned}$$
with the orthogonality also holding if $\beta =\alpha '$ and the pairs of modes are properly chosen [53]. We normalize the modes according to
$$\int_{V}\frac{\boldsymbol{B}_{\alpha}^{*}(\boldsymbol{r})\cdot\boldsymbol{B}_{\alpha}(\boldsymbol{r})}{\mu_{0}}\textrm{d}\boldsymbol{r}=1,$$
and it then follows from (18) that
$$\int_{V}\frac{\boldsymbol{D}_{\alpha}^{*}(\boldsymbol{r})\cdot\boldsymbol{D}_{\alpha}(\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}\textrm{d}\boldsymbol{r}=1.$$

2.2.2 Canonical Formulation

It is useful to use a slightly more general canonical formulation than often employed, but we begin by reviewing the traditional approach. There one starts with a set of functions of the canonical coordinates and momenta, $q_{i}$ and $p_{j}$ respectively, and with the Poisson bracket of two such functions $f(q_{i},p_{j})$, $g(q_{i},p_{j})$ defined according to

$$\left\{ f,g\right\} =\sum_{k}\left(\frac{\partial f}{\partial q_{k}}\frac{\partial g}{\partial p_{k}}-\frac{\partial f}{\partial p_{k}}\frac{\partial g}{\partial q_{k}}\right),$$
from which follows the properties
$$\begin{aligned} & \left\{ f,g\right\} ={-}\left\{ g,f\right\},\\ & \left\{ f+g,b\right\} =\left\{ f,b\right\} +\left\{ g,b\right\},\\ & \left\{ fg,b\right\} =f\left\{ g,b\right\} +g\left\{ f,b\right\},\\ & \left\{ f,\left\{ g,b\right\} \right\} +\left\{ g,\left\{ b,f\right\} \right\} +\left\{ b,\left\{ f,g\right\} \right\} =0. \end{aligned}$$
We then find the familiar results
$$\begin{aligned} & \left\{ q_{i},q_{j}\right\} =\left\{ p_{i},p_{j}\right\} =0,\\ & \left\{ q_{i},p_{j}\right\} =\delta_{ij}. \end{aligned}$$
The dynamical equations are then taken to be given by
$$\frac{\textrm{d}f}{\textrm{d}t}=\left\{ f,H\right\}.$$
Now consider a more general approach. Suppose we have a set of quantities $(f,g,b,\ldots )$ and a definition of their Poisson brackets satisfying (26) such that the set is closed under the operations specified by (26); then those quantities together with their Poisson brackets form a representation of a Lie algebra. If there is a Hamiltonian $H$ that is an element of that set, and the dynamics of the quantities are given by (28), we take this to constitute a canonical formulation.

As an example, suppose that instead of deriving (27) from (25) we simply assert (27) for the $q_{i}$ and the $p_{j}$, and take our set of quantities to be the $q_{i}$, the $p_{j}$, and all polynomial functions of them. Then, together with the definition that the Poisson bracket of two numbers vanishes, and the definition that the Poisson bracket of a number with any of the quantities vanishes, the Poisson brackets of all such quantities can be determined from (26) and (27), and the set of quantities is closed under those operations. If the Hamiltonian is taken as a polynomial function of the $q_{i}$ and the $p_{j}$, then the dynamics given by (28) are precisely what we would expect from the more traditional approach. This more general approach is often useful. One example is if one wants to treat the angular momentum components $J_{i}$ of an object as fundamental, as one does for the spin of a particle; then there are no canonical position and momentum in the description, but the Lie algebra involves the Poisson bracket

$$\left\{ J_{x},J_{y}\right\} =J_{z},$$
and the brackets that follow from cyclic permutations. Of course, particle spin is not a classical concept, but one could use this strategy to develop a canonical formulation for the angular momentum of a rigid body without reverting to a description of the particles constituting the body and their momenta.

That approach is also useful here. Returning to our Eqs. (16), we note that the divergence equations can be taken as initial conditions, for if they are satisfied at an initial time and the curl equations are satisfied at all later times, the divergence equations will also be satisfied at all later times. Thus, we need only construct dynamical equations for the curl equations, and for ultimate use in quantization we seek to do this with a Hamiltonian that is numerically equal to the energy. From (5) we see that the Hamiltonian should then be

$$\textrm{H}^{\textrm{L}}=\int_{V}\left(\frac{\boldsymbol{D}(\boldsymbol{r},t)\cdot\boldsymbol{D}(\boldsymbol{r},t)}{2\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})} +\frac{\boldsymbol{B}(\boldsymbol{r},t)\cdot\boldsymbol{B}(\boldsymbol{r},t)}{2\mu_{0}}\right)\textrm{d}\boldsymbol{r},$$
where $V$ is the normalization volume.

As first pointed out by Born and Infeld [39], to obtain the correct dynamical equations the fundamental Poisson brackets should be taken as

$$\begin{aligned} & \left\{ D^{i}(\boldsymbol{r},t),D^{j}(\boldsymbol{r'},t)\right\} =\left\{ B^{i}(\boldsymbol{r},t),B^{j}(\boldsymbol{r'},t)\right\} =0,\\ & \left\{ D^{i}(\boldsymbol{r},t),B^{j}(\boldsymbol{r'},t)\right\} =\epsilon^{ilj}\frac{\partial}{\partial r^{l}}\delta(\boldsymbol{r}-\boldsymbol{r'}), \end{aligned}$$
for using these with (30) we recover our desired dynamics,
$$\begin{aligned}& \frac{\partial\boldsymbol{D}(\boldsymbol{r},t)}{\partial t}=\left\{ \boldsymbol{D}(\boldsymbol{r},t),\textrm{H}^{\textrm{L}}\right\} =\frac{1}{\mu_{0}}\boldsymbol{\nabla}\times\boldsymbol{B}(\boldsymbol{r},t),\\ & \frac{\partial\boldsymbol{B}(\boldsymbol{r},t)}{\partial t}=\left\{ \boldsymbol{B}(\boldsymbol{r},t),\textrm{H}^{\textrm{L}}\right\} ={-}\boldsymbol{\nabla}\times\left(\frac{\boldsymbol{D}(\boldsymbol{r},t)}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}\right). \end{aligned}$$

2.2.3 Quantization

We now quantize in the usual way, by taking

$$\left\{ \;\right\} \Rightarrow\frac{1}{i\hbar}\left[\;\right],$$
where $\left [\;\right ]$ indicates the commutator. The equal time commutation relations for Heisenberg operators are then given by
$$\begin{aligned} & \left[D^{i}(\boldsymbol{r},t),D^{j}(\boldsymbol{r'},t)\right]=\left[B^{i}(\boldsymbol{r},t),B^{j}(\boldsymbol{r'},t)\right]=0,\\ & \left[D^{i}(\boldsymbol{r},t),B^{j}(\boldsymbol{r'},t)\right]=i\hbar\epsilon^{ilj}\frac{\partial}{\partial r^{l}}\delta(\boldsymbol{r}-\boldsymbol{r'}), \end{aligned}$$
and to satisfy the initial conditions we expand our fields in modes,
$$\begin{aligned}& \boldsymbol{D}(\boldsymbol{r},t)=\sum_{\alpha}\mathcal{C}_{\alpha}^{(1)}(t)\boldsymbol{D}_{\alpha}(\boldsymbol{r}),\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{\alpha}\mathcal{C}_{\alpha}^{(2)}(t)\boldsymbol{B}_{\alpha}(\boldsymbol{r}), \end{aligned}$$
where here the $\mathcal {C}_{\alpha }^{(1)}(t)$ and the $\mathcal {C}_{\alpha }^{(2)}(t)$ are operators. Whatever their time dependence, the field operators $\boldsymbol{D}(\boldsymbol{r},t)$ and $\boldsymbol{B}(\boldsymbol{r},t)$ will be divergenceless at all times, because $\boldsymbol{D}_{\alpha }(\boldsymbol{r})$ and $\boldsymbol{B}_{\alpha }(\boldsymbol{r})$ are divergenceless.

However, because those field operators must be Hermitian, the operators $\mathcal {C}_{\alpha }^{(1)}(t)$ and $\mathcal {C}_{\alpha }^{(2)}(t)$ are not all independent; we easily find the conditions

$$\begin{aligned}& \mathcal{C}_{\alpha'}^{(1)}(t)={-}\left(\mathcal{C}_{\alpha}^{(1)}(t)\right)^{{\dagger}},\\ & \mathcal{C}_{\alpha'}^{(2)}(t)={+}\left(\mathcal{C}_{\alpha}^{(2)}(t)\right)^{{\dagger}}, \end{aligned}$$
where we have used (19). We can ensure these conditions are satisfied by introducing new operators $a_{\alpha }(t)$ with no restrictions, such that
$$\begin{aligned}& \mathcal{C}_{\alpha}^{(1)}(t)=\sqrt{\frac{\hbar\omega_{\alpha}}{2}}\left(a_{\alpha}(t)-a_{\alpha'}^{{\dagger}}(t)\right),\\ & \mathcal{C}_{\alpha}^{(2)}(t)=\sqrt{\frac{\hbar\omega_{\alpha}}{2}}\left(a_{\alpha}(t)+a_{\alpha'}^{{\dagger}}(t)\right), \end{aligned}$$
where the factor $\sqrt {\left (\hbar \omega _{\alpha }\right )/2}$ is introduced for later convenience; here $\alpha '$ is the partner mode of $\alpha$. We then have
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\sum_{\alpha}\sqrt{\frac{\hbar\omega_{\alpha}}{2}}\left(a_{\alpha}(t)\boldsymbol{D}_{\alpha}(\boldsymbol{r})+a_{\alpha}^{{\dagger}}(t)\boldsymbol{D}_{\alpha}^{*}(\boldsymbol{r})\right),\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{\alpha}\sqrt{\frac{\hbar\omega_{\alpha}}{2}}\left(a_{\alpha}(t)\boldsymbol{B}_{\alpha}(\boldsymbol{r})+a_{\alpha}^{{\dagger}}(t)\boldsymbol{B}_{\alpha}^{*}(\boldsymbol{r})\right), \end{aligned}$$
where we have again used (19). The equal time commutation relations (34) are then satisfied by taking
$$\begin{aligned}& \left[a_{\alpha}(t),a_{\beta}(t)\right]=0,\\ & \left[a_{\alpha}(t),a_{\beta}^{{\dagger}}(t)\right]=\delta_{\alpha\beta}, \end{aligned}$$
which are indeed required if the set of modes is complete. Using the field expansions (38) in the expression (30) for the Hamiltonian, we find
$$\begin{aligned} \textrm{H}^{\textrm{L}} & =\sum_{\alpha}\frac{\hbar\omega_{\alpha}}{2}\left(a_{\alpha}^{{\dagger}}(t)a_{\alpha}(t)+a_{\alpha}(t)a_{\alpha}^{{\dagger}}(t)\right)\\ & =\sum_{\alpha}\hbar\omega_{\alpha}\left(a_{\alpha}^{{\dagger}}(t)a_{\alpha}(t)+\frac{1}{2}\right)\\ & \rightarrow\sum_{\alpha}\hbar\omega_{\alpha}a_{\alpha}^{{\dagger}}(t)a_{\alpha}(t), \end{aligned}$$
where in the last expression we have neglected the zero-point energy, as we do henceforth. The equation for any Heisenberg operator $\mathcal {O}(t)$,
$$i\hbar\frac{\textrm{d}\mathcal{O}(t)}{\textrm{d}t}=\left[\mathcal{O}(t),\textrm{H}^{\textrm{L}}\right],$$
[compare with (28)] leads to
$$\frac{\textrm{d}a_{\alpha}(t)}{\textrm{d}t}={-}i\omega_{\alpha}a_{\alpha}(t),$$
as expected.

2.2.4 Special Cases

The usual first case of interest is a uniform medium. We treat that for completeness in Appendix A. Here we consider the form of the field operators for two simple structures from integrated optics, the first a channel waveguide as indicated schematically in Fig. 2, and the second a ring resonator, initially considered isolated and indicated schematically in Fig. 1; of interest in itself, the ring resonator can also be taken as a simple example of an integrated optical cavity.

 figure: Figure 2.

Figure 2. Schematic representation of a channel waveguide. (a) Typical field intensity profile of the fundamental transverse-electric mode, computed by an eigenmode solver. The waveguide and simulation parameters are the same as in Fig. 1, but with no curvature. By comparison with Fig. 1 it is possible to appreciate the different field distribution produced by the waveguide curvature. The contour plot accounts for the $E_{x}$ component, whereas the arrow field represents $\boldsymbol{E}_{y,z}$. (b) Scanning electron microscopy (SEM) image of a typical high index contrast waveguide cross section. Reprinted from [54]. Copyright 2020 Optical Society of America.

Download Full Size | PDF

2.2.4.1 Channel Waveguide

Looking first at the waveguide structure, we take $\boldsymbol{\hat {z}}$ normal to the substrate and $\boldsymbol{\hat {x}}$ the direction along which the waveguide runs; then $\varepsilon _{1}(\boldsymbol{r})=\varepsilon _{1}(y,z)$ here.

We are interested in modes confined to the waveguide, and so we need only introduce periodic boundary conditions in the $\boldsymbol{\hat {x}}$ direction, with periodicity over a length $L$. Then because of the translational symmetry in the $x$ direction we can choose the modes $(\boldsymbol{D}_{\alpha }(\boldsymbol{r}),\boldsymbol{B}_{\alpha }(\boldsymbol{r}))$ satisfying (18) to be of the form

$$\begin{aligned} & \boldsymbol{D}_{Ik}(\boldsymbol{r})=\frac{\boldsymbol{d}_{Ik}(y,z)}{\sqrt{L}}e^{ikx},\\ & \boldsymbol{B}_{Ik}(\boldsymbol{r})=\frac{\boldsymbol{b}_{Ik}(y,z)}{\sqrt{L}}e^{ikx}, \end{aligned}$$
where $k=2\pi m/L,$ with $m$ an integer; the index $I$ labels the different modes, which in their polarization profile will generally be more complicated than either circular or linear polarization; we label the frequency of such a mode by $\omega _{Ik}$, and the partner mode of a mode characterized by $k$ is one characterized by $-k$. In the normalization conditions (23) and (24) we can let $y$ and $z$ vary over all values, with only $x$ limited to a region of length $L$, yielding
$$\int\frac{\boldsymbol{b}_{Ik}^{*}(y,z)\cdot\boldsymbol{b}_{Ik}(y,z)}{\mu_{0}}\textrm{d}y\textrm{d}z=1,$$
and
$$\int\frac{\boldsymbol{d}_{Ik}^{*}(y,z)\cdot\boldsymbol{d}_{Ik}(y,z)}{\epsilon_{0}\varepsilon_{1}(y,z)}\textrm{d}y\textrm{d}z=1,$$
as the normalization conditions of $\boldsymbol{b}_{Ik}(y,z)$ and $\boldsymbol{d}_{Ik}(y,z).$ The general field expansions (38) then take the form
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\sum_{I,k}\sqrt{\frac{\hbar\omega_{Ik}}{2L}}a_{Ik}(t)\boldsymbol{d}_{Ik}(y,z)e^{ikx}+\textrm{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I,k}\sqrt{\frac{\hbar\omega_{Ik}}{2L}}a_{Ik}(t)\boldsymbol{b}_{Ik}(y,z)e^{ikx}+\textrm{H.c.}, \end{aligned}$$
where $\omega _{Ik}$ is the frequency of mode $I,k,$
$$\begin{aligned} & \left[a_{Ik}(t),a_{I'k'}(t)\right]=0,\\ & \left[a_{Ik}(t),a_{I'k'}^{{\dagger}}(t)\right]=\delta_{II'}\delta_{kk'}, \end{aligned}$$
and the Hamiltonian is
$$\textrm{H}^{\textrm{L}}=\sum_{I,k}\hbar\omega_{Ik}a_{Ik}^{{\dagger}}(t)a_{Ik}(t).$$
We consider passing to the limit of an infinite normalization length $L$, which involves a continuously varying $k$. We do this in the usual heuristic way by replacing the sum over $k$ by an integral over $k$, taking into account the density of points in $k$ space. That is, we take
$$\sum_{k}\rightarrow\frac{\int \textrm{d}k}{\left(\frac{2\pi}{L}\right)}=\frac{L}{2\pi}\int \textrm{d}k.$$
We also have to replace the operators $a_{Ik}(t)$ associated with discrete modes by operators associated with modes labeled by a continuous varying $k;$ those latter operators we label $a_{I}(k,t).$ The connection between the two can be identified by recalling that for the discrete modes we have (47), and so, in particular,
$$\begin{aligned}& \sum_{k'}\left[a_{Ik}(t),a_{I'k'}^{{\dagger}}(t)\right]=\delta_{II'},\\ & \sum_{k'}\left[a_{Ik}(t),a_{I'k'}(t)\right]=0. \end{aligned}$$
Recalling (49), we see that if we take
$$a_{Ik}(t)\rightarrow\sqrt{\frac{2\pi}{L}}a_{I}(k,t),$$
we arrive at
$$\begin{aligned}& \int\left[a_{I}(k,t),a_{I}^{{\dagger}}(k',t)\right]\textrm{d}k'=\delta_{II'},\\ & \int\left[a_{I}(k,t),a_{I}(k',t)\right]\textrm{d}k'=0, \end{aligned}$$
and because the operators at different $k$ are to be independent, this yields
$$\begin{aligned} & \left[a_{I}(k,t),a_{I'}(k',t)\right]=0,\\ & \left[a_{I}(k,t),a_{I'}^{{\dagger}}(k',t)\right]=\delta_{II'}\delta(k-k'). \end{aligned}$$
Thus, the $a_{I}(k,t)$ are convenient operators for the expansion of the electromagnetic field. Using (49) and (51) in (46) we find
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\sum_{I}\int\sqrt{\frac{\hbar\omega_{Ik}}{4\pi}}a_{I}(k,t)\boldsymbol{d}_{Ik}(y,z)e^{ikx}\textrm{d}k+\textrm{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I}\int\sqrt{\frac{\hbar\omega_{Ik}}{4\pi}}a_{I}(k,t)\boldsymbol{b}_{Ik}(y,z)e^{ikx}\textrm{d}k+\textrm{H.c.}, \end{aligned}$$
and using them in (48) we find
$$\textrm{H}_{\textrm{chan}}^{\textrm{L}}=\sum_{I}\int\hbar\omega_{Ik}a_{I}^{{\dagger}}(k,t)a_{I}(k,t)\textrm{d}k.$$
Here we have not changed the notation for $\omega _{Ik}$ nor for $\boldsymbol{d}_{Ik}(y,z)$ and $\boldsymbol{b}_{Ik}(y,z)$, although $k$ now various continuously; the normalization conditions remain (44) and (45). Of course, the integration over $k$ here only ranges over the values for which the modes are confined to the waveguide structure; that is, the part of the electromagnetic field we include in (54).

2.2.4.2 Ring Resonators

Next we turn to the ring structure shown in Fig. 1, and consider the form of the modes $\left (\boldsymbol{D}_{\alpha }\left (\boldsymbol{r}\right ),\boldsymbol{B}_{\alpha }\left (\boldsymbol{r}\right )\right )$ that can at least be approximated as confined to the ring. Here we use indices $J$ to denote the modes, and $\omega _{J}$ their frequencies; the modes can be taken to be of the form

$$\begin{aligned} & \boldsymbol{D}_{J}(\boldsymbol{r})=\frac{\textsf{d}_{J}(\boldsymbol{r}_{{\perp}};\zeta)}{\sqrt{\mathcal{L}}}e^{i\kappa_{J}\zeta},\\ & \boldsymbol{B}_{J}(\boldsymbol{r})=\frac{\textsf{b}_{J}(\boldsymbol{r}_{{\perp}};\zeta)}{\sqrt{\mathcal{L}}}e^{i\kappa_{J}\zeta}, \end{aligned}$$
where, as in Sec. 2.1, $\kappa _{J}=2\pi m_{J}/\mathcal {L}$ where $m_{J}$ is an integer; we use $J$ to denote the type of the mode. The dependence of $\textsf {d}_{J}(\boldsymbol{r}_{\perp };\zeta )$ and $\textsf {b}_{J}(\boldsymbol{r}_{\perp };\zeta )$ on $\zeta$ arises because of components of the vectors in the $xy$ plane. As $\zeta$ varies the direction of those vectors will change, although $\textsf {b}_{J}^{*}(\boldsymbol{r}_{\perp };\zeta )\cdot \textsf {b}_{J}(\boldsymbol{r}_{\perp };\zeta )$ and $\textsf {d}_{J}^{*}(\boldsymbol{r}_{\perp };\zeta )\cdot \textsf {d}_{J}(\boldsymbol{r}_{\perp };\zeta )$ will be independent of $\zeta$ and we can put $\textsf {b}_{J}^{*}(\boldsymbol{r}_{\perp };\zeta )\cdot \textsf {b}_{J}(\boldsymbol{r}_{\perp };\zeta )=\textsf {b}_{J}^{*}(\boldsymbol{r}_{\perp };0)\cdot \textsf {b}_{J}(\boldsymbol{r}_{\perp };0)$ and $\textsf {d}_{J}^{*}(\boldsymbol{r}_{\perp };\zeta )\cdot \textsf {d}_{J}(\boldsymbol{r}_{\perp };\zeta )=\textsf {d}_{J}^{*}(\boldsymbol{r}_{\perp };0)\cdot \textsf {d}_{J}(\boldsymbol{r}_{\perp };0)$.

The electromagnetic field operators associated with the modes (56) can then be written, using (38), as

$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2\mathcal{L}}}c_{J}(t)\textsf{d}_{J}(\boldsymbol{r}_{{\perp}},\zeta)e^{i\kappa_{J}\zeta}+\textrm{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2\mathcal{L}}}c_{J}(t)\textsf{b}_{J}(\boldsymbol{r}_{{\perp}},\zeta)e^{i\kappa_{J}\zeta}+\textrm{H.c.}, \end{aligned}$$
where $J$ indicates both $\kappa _{J}$ and the type of mode,
$$\begin{aligned} & \left[c_{J}(t),c_{J'}(t)\right]=0,\\ & \left[c_{J}(t),c_{J'}^{{\dagger}}(t)\right]=\delta_{JJ'}, \end{aligned}$$
and the Hamiltonian is
$${H}_{\textrm{ring}}^{\textrm{L}}=\sum_{J}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J},$$
as in (2).

The general normalization conditions (23) and (24) here become

$$\int\frac{\textsf{b}_{J}^{*}(\boldsymbol{r}_{{\perp}};0)\cdot\textsf{b}_{J}(\boldsymbol{r}_{{\perp}};0)}{\mu_{0}}\textrm{d}\boldsymbol{r_{{\perp}}}=1,$$
$$\int\frac{\textsf{d}_{J}^{*}(\boldsymbol{r}_{{\perp}};0)\cdot\textsf{d}_{J}(\boldsymbol{r}_{{\perp}};0)}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r_{{\perp}}})}\textrm{d}\boldsymbol{r}_{{\perp}}=1.$$
The second of these gives the normalization condition for $\mathsf{e}_{J}(\boldsymbol{r}_{\perp };\zeta$), because comparing (3) with the first of (57) we see that $\textsf {d}_{J}(\boldsymbol{r}_{\perp };\zeta )=\epsilon _{0}\varepsilon _{1}(\boldsymbol{r})\mathsf {e}_{J}(\boldsymbol{r}_{\perp };\zeta ).$

For the curvature of the ring not too great, the fields $(\textsf{d}_{J}(\boldsymbol{r}_{\perp };\zeta ),\textsf{b}_{J}(\boldsymbol{r}_{\perp };\zeta ))$ will be related to those fields $(\boldsymbol{d}_{Ik}(y,z),\boldsymbol{b}_{Ik}(y,z))$ of a channel structure of the same cross section according to

$$\begin{aligned} & \textsf{d}_{J}(\rho,z;0)\approx\boldsymbol{d}_{I_{\alpha}k_{\alpha}}(R-\rho,z),\\ & \textsf{b}_{J}(\rho,z;0)\approx\boldsymbol{b}_{I_{\alpha}k_{\alpha}}(R-\rho,z), \end{aligned}$$
[and compare (44) and (45) with (60) and (61)].

2.2.5 Including Dispersion

Up to this point we have assumed that the relative dielectric constant $\varepsilon _{1}(\boldsymbol{r})$ is independent of the frequency of the optical field. More generally, however, if we have electric and displacement fields

$$\begin{aligned} & \boldsymbol{E}(\boldsymbol{r},t)=\boldsymbol{E}(\boldsymbol{r})e^{{-}i\omega t}+\textrm{c.c.},\\ & \boldsymbol{D}(\boldsymbol{r},t)=\boldsymbol{D}(\boldsymbol{r})e^{{-}i\omega t}+\textrm{c.c.}, \end{aligned}$$
which for the moment we consider classical, they are related by a frequency dependent $\varepsilon _{1}(\boldsymbol{r},\omega ),$
$$\boldsymbol{D}(\boldsymbol{r})=\epsilon_{0}\varepsilon_{1}(\boldsymbol{r};\omega)\boldsymbol{E}(\boldsymbol{r}).$$
The quantity $\varepsilon _{1}(\boldsymbol{r};\omega )$ is in general complex, with its real part describing dispersive effects and its imaginary part describing absorption; they are linked by Kramers–Kronig relations [55]. This more complicated electrodynamics arises, of course, because of the involvement of degrees of freedom of the underlying material medium. Quantizing the dynamics then involves quantizing both the “bare” electromagnetic field and those material degrees of freedom to which it is coupled. Here we follow a strategy introduced earlier [56]. Although the dynamics over a wide frequency range was considered there, for integrated photonic structures one is mainly interested in frequencies below the bandgap of all materials, for which $\varepsilon _{1}(\boldsymbol{r};\omega )$ may be a strong function of frequency but is purely real. That is the regime we consider here; the treatment of quantum electrodynamics in the regime where absorption is presented has been investigated using this strategy elsewhere [57].

We begin by noting that for $\varepsilon _{1}(\boldsymbol{r};\omega )$ purely real we can still consider “modes” for $\boldsymbol{D}(\boldsymbol{r},t)$ and $\boldsymbol{B}(\boldsymbol{r},t)$, as we did when dispersion was neglected. Seeking solutions of Maxwell’s Eqs. (16) of the form (17), but now using the constitutive relations

$$\begin{aligned} & \boldsymbol{D}_{\alpha}(\boldsymbol{r})=\epsilon_{0}\varepsilon_{1}(\boldsymbol{r},\omega_{\alpha})\boldsymbol{E}_{\alpha}(\boldsymbol{r}),\\ & \boldsymbol{B}_{\alpha}(\boldsymbol{r})=\mu_{0}\boldsymbol{H}_{\alpha}(\boldsymbol{r}), \end{aligned}$$
we find we require
$$\begin{aligned} & \boldsymbol{\nabla}\times\left[\frac{\boldsymbol{\nabla}\times\boldsymbol{B}_{\alpha}(\boldsymbol{r})}{\varepsilon_{1}(\boldsymbol{r};\omega_{\alpha})}\right]=\frac{\omega_{\alpha}^{2}}{c^{2}}\boldsymbol{B}_{\alpha}(\boldsymbol{r}),\\ & \boldsymbol{\nabla}\cdot\boldsymbol{B}_{\alpha}(\boldsymbol{r})=0,\\ & \boldsymbol{D}_{\alpha}(\boldsymbol{r})=\frac{i}{\omega_{\alpha}\mu_{0}}\boldsymbol{\nabla}\times\boldsymbol{B}_{\alpha}(\boldsymbol{r}). \end{aligned}$$
The first of these is more complicated than the first of (18), with $\omega _{\alpha }$ here appearing on both sides of the eigenvalue equation. In general, for example for a photonic crystal, determining the allowed $\omega _{\alpha }$ must be done self-consistently, with some iteration. For the simpler structures we consider here, however, that will not be a problem. Note also that the mode fields $\boldsymbol{B}_{\alpha }(\boldsymbol{r})$ and $\boldsymbol{B}_{\beta }(\boldsymbol{r})$ associated with different frequencies $\omega _{\alpha }\neq \omega _{\beta }$ are, in general, not orthogonal; the condition (22) does not hold here. Mathematically, this arises because such a $\boldsymbol{B}_{\alpha }(\boldsymbol{r})$ and a $\boldsymbol{B}_{\beta }(\boldsymbol{r})$ are eigenfunctions of different Hermitian operators, because in (20) $\varepsilon _{1}(\boldsymbol{r})$ must be replaced by $\varepsilon _{1}(\boldsymbol{r};\omega_\alpha )$ when it acts on $\boldsymbol{B}_{\alpha }(\boldsymbol{r})$, and by $\varepsilon _{1}(\boldsymbol{r};\omega _{\beta })$ when it acts on $\boldsymbol{B}_{\beta }(\boldsymbol{r})$. Physically, it arises because the true modes of the system involve both the electromagnetic field and the degrees of freedom of the material medium; when these full “polariton modes” are considered orthogonality reappears [56]. Nonetheless, one finds that one can proceed with just $(\boldsymbol{D}_{\alpha }(\boldsymbol{r}),\boldsymbol{B}_{\alpha }(\boldsymbol{r}))$, satisfying (66), in Eqs. (38) if the normalization conditions (23) and (24) are modified; one formally introduces the material degrees of freedom to show this, and then one need not explicitly deal with them again. Once $\boldsymbol{D}_{\alpha }(\boldsymbol{r})$ and $\boldsymbol{B}_{\alpha }(\boldsymbol{r})$ are found from (66), for future use the appropriate normalization condition is most usefully written in terms of $\boldsymbol{D}_{\alpha }(\boldsymbol{r})$, and for the case of periodic boundary conditions is given by [56,58]
$$\int_{V}\frac{\boldsymbol{D}_{\alpha}^{*}(\boldsymbol{r})\cdot\boldsymbol{D}_{\alpha}(\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r};\omega_{\alpha})} \frac{v_{p}(\boldsymbol{r};\omega_{\alpha})}{v_{g}(\boldsymbol{r};\omega_{\alpha})}\textrm{d}\boldsymbol{r}=1$$
instead of (24). Here
$$v_{p}(\boldsymbol{r};\omega)=\frac{c}{n(\boldsymbol{r};\omega)}$$
is the local phase velocity at frequency $\omega$, where $n(\boldsymbol{r};\omega )=\sqrt {\varepsilon _{1}(\boldsymbol{r};\omega )}$ is the local index of refraction at that frequency, and
$$v_{g}(\boldsymbol{r};\omega)=\frac{v_{p}(\boldsymbol{r};\omega)}{1+\frac{\omega}{n(\boldsymbol{r};\omega)}\frac{\partial n(\boldsymbol{r};\omega)}{\partial\omega}}$$
is the local group velocity at frequency $\omega.$

In Appendix A we consider the inclusion of dispersion in a description of light in a uniform medium. Here we begin with the inclusion of dispersion in the description of channel waveguide structures, first considering periodic boundary conditions. The mode frequencies $\omega _{Ik}$ must be found by solving (66) with modes of the form (43); following the discussion above, it is easy to show from (67) that the normalization condition (45) is then replaced by

$$\int\frac{\boldsymbol{d}_{Ik}^{*}(y,z)\cdot\boldsymbol{d}_{Ik}(y,z)}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\frac{v_{p}(y,z;\omega_{Ik})}{v_{g}(y,z;\omega_{Ik})}\textrm{d}y\textrm{d}z=1.$$
With this new normalization and the corresponding scaling for $\boldsymbol{b}_{Ik}(y,z;)$, the expressions (46), (47), and (48) for the field expansions, the commutation relations, and the Hamiltonian still all hold. Forming the Poynting vector $\boldsymbol{S}(\boldsymbol{r},t)=\boldsymbol{E}(\boldsymbol{r},t)\times \boldsymbol{H}(\boldsymbol{r},t)$, if we look at a one-photon state
$$\left|\Psi\right\rangle =a_{Ik}^{{\dagger}}(0)\left\vert\text{vac}\right\rangle$$
we find
$$\int\left\langle \Psi|\boldsymbol{S}(\boldsymbol{r},t)|\Psi\right\rangle \cdot\boldsymbol{\hat{x}}\textrm{d}y\textrm{d}z=\frac{\hbar\omega_{Ik}}{L}v_{Ik},$$
where of course $(\hbar \omega _{Ik}/L)$ is the photon energy per unit length in the waveguide, and
$$v_{Ik}\equiv\frac{\partial\omega_{Ik}}{\partial k}=\frac{1}{2}\int\left(\boldsymbol{e}_{Ik}(y,z)\times\boldsymbol{h}_{Ik}^{*}(y,z)+\boldsymbol{e}_{Ik}^{*}(y,z)\times\boldsymbol{h}_{Ik}(y,z)\right)\cdot\boldsymbol{\hat{x}}\textrm{d}y\textrm{d}z$$
is the group velocity of the mode, including both modal and material dispersion; here
$$\begin{aligned} & \boldsymbol{e}_{Ik}(y,z)=\frac{\boldsymbol{d}_{Ik}(y,z)}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})},\\ & \boldsymbol{h}_{Ik}(y,z)=\frac{\boldsymbol{b}_{Ik}(y,z)}{\mu_{0}}, \end{aligned}$$
and (73) can be derived from the normalization condition (70), which we do in Appendix B. The appearance of the local group velocity in the normalization expressions (67) and (70) is perhaps not surprising, as the group velocity (73) of the mode, which determines the rate at which energy is transported, depends on it [55]. If we pass to an infinite length waveguide structure, then (53), (54), and (55) all hold, with the $\boldsymbol{d}_{Ik}(y,z)$ and $\boldsymbol{b}_{Ik}(y,z)$ found as in the periodic boundary condition analysis, normalized according to (70).

Finally, for the isolated ring the inclusion of dispersion leads to using the mode forms (56) in (66) to find the mode frequencies $\omega _{J}$ and the fields $\textsf{d}_{J}(\boldsymbol{r}_{\perp };\zeta )$ and $\textsf{b}_{J}(\boldsymbol{r}_{\perp };\zeta )$. Instead of the normalization condition (61) we use

$$\int\frac{\textsf{d}_{J}^{*}(\boldsymbol{r}_{{\perp}};0)\cdot\textsf{d}_{J}(\boldsymbol{r}_{{\perp}};0)}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r_{{\perp}}};\omega_{J})} \frac{v_{p}(\boldsymbol{r}_{{\perp}};\omega_{J})}{v_{g}(\boldsymbol{r}_{{\perp}};\omega_{J})}\textrm{d}\boldsymbol{r}_{{\perp}}=1.$$
The field expansion (57), the commutation relations (58), and the form of the Hamiltonian (59) are all unchanged.

2.3 Channel Fields and Ring–Channel Coupling

We now turn to the coupling between the channel and ring, in a structure such as in Fig. 3. We identify the isolated ring resonances $\omega _{J}$, restricting ourselves here to resonances associated with one transverse mode; recall that the isolated ring fields are given by (57), with in general the normalization condition (75). Our electromagnetic fields associated with the channel are expanded as (54), with commutation relations (53) and a Hamiltonian (55); the mode normalization condition is generally given by (70). In those expressions we have used the dummy index $I$ to indicate the mode type. For the channel we now just consider one type, and use the dummy index $J$ to label a range of $k$ (with a corresponding range of $\omega _{Jk}$), introducing different wavenumber ranges, each centered at a ring mode resonance.

 figure: Figure 3.

Figure 3. Schematic representation of a point-coupled ring resonator. The coupling is modeled by a simple $\begin {pmatrix} \sigma & i\kappa \\ i\kappa & \sigma \\ \end {pmatrix}$ matrix, where $(\sigma, \kappa ) \in \mathbb {R}$ are the self-coupling and cross-coupling coefficients, respectively. The coupler is assumed lossless, hence yielding $\sigma ^2 + \kappa ^2 = 1$.

Download Full Size | PDF

If we denote the range of $k$ allotted to $J$ by $\mathscr {R}(J)$, (see Fig. 4) we can then write

$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\sum_{J}\int_{k\in\mathscr{R}(J)}\sqrt{\frac{\hbar\omega_{Jk}}{4\pi}}a_{J}(k,t)\boldsymbol{d}_{Jk}(y,z)e^{ikx}\textrm{d}k+\textrm{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{J}\int_{k\in\mathscr{R}(J)}\sqrt{\frac{\hbar\omega_{Jk}}{4\pi}}a_{J}(k,t)\boldsymbol{b}_{Jk}(y,z)e^{ikx}\textrm{d}k+\textrm{H.c.}, \end{aligned}$$
where
$$\begin{aligned} & \left[a_{J}(k,t),a_{J'}(k',t)\right]=0,\\ & \left[a_{J}(k,t),a_{J'}^{{\dagger}}(k',t)\right]=\delta_{JJ'}\delta(k-k'), \end{aligned}$$
and the Hamiltonian is
$$\textrm{H}_{\textrm{chan}}^{\textrm{L}}=\sum_{J}H_{J}^{\text{L}},$$
where
$$H_{J}^{\text{L}}=\int_{k\in\mathscr{R}(J)}\hbar\omega_{Jk}a_{J}^{{\dagger}}(k,t)a_{J}(k,t)\textrm{d}k.$$
For some applications it is convenient to introduce “channel field operators,”
$$\psi_{J}(x,t)\equiv\int_{k\in\mathscr{R}(J)}\frac{\textrm{d}k}{\sqrt{2\pi}}a_{J}(k,t)e^{i(k-k_{J})x},$$
where we use $k_{J}$ to denote a reference $k$ in the center of the range $\mathscr {R}(J)$, and before considering the coupling of the channel to the ring, we discuss these operators. From the Hamiltonian (79) their dynamics are given by
$$\begin{aligned} \frac{\partial\psi_{J}(x,t)}{\partial t} & ={-}i\int_{k\in\mathscr{R}(J)}\frac{\textrm{d}k}{\sqrt{2\pi}}\omega_{Jk}a_{J}(k,t)e^{i(k-k_{J})x}\\ & ={-}i\int_{k\in\mathscr{R}(J)}\frac{\textrm{d}k}{\sqrt{2\pi}}\left(\omega_{J}+v_{J}(k-k_{J})+\frac{1}{2}v'_{J}(k-k_{J})^{2}+\cdots\right)a_{J}(k,t)e^{i(k-k_{J})x}, \end{aligned}$$
where in the second line we have put $\omega _{J}\equiv \omega _{Jk_{J}}$ and assumed that in the range $k\in \mathscr {R}(J)$ the full modal dispersion is small enough that an expansion of the dispersion relation is reasonable; we have put $v_{J}\equiv v_{Jk_{J}}$ and $v'_{J}\equiv v'_{Jk_{J}}$, where
$$v'_{Jk}=\frac{\partial^{2}\omega_{Jk}}{\partial k^{2}}.$$
Then we can write (81) as
$$\frac{\partial\psi_{J}(x,t)}{\partial t}={-}i\omega_{J}\psi_{J}(x,t)-v_{J}\frac{\partial\psi_{J}(x,t)}{\partial x}+\frac{1}{2}iv'_{J}\frac{\partial^{2}\psi_{J}(x,t)}{\partial x^{2}}+\cdots.$$
For many applications the GVD within each channel range $\mathscr {R}(J)$ can be neglected, and we can take simply
$$\frac{\partial\psi_{J}(x,t)}{\partial t}={-}i\omega_{J}\psi_{J}(x,t)-v_{J}\frac{\partial\psi_{J}(x,t)}{\partial x}.$$

 figure: Figure 4.

Figure 4. Typical transmission of a point-coupled ring resonator. Here $\mathscr {R}(J)$ represents the wave vector range associated with each resonance $J$, centered at $k_J$.

Download Full Size | PDF

Under certain conditions the Hamiltonian (79) and the derivation of the dynamics of $\psi _{J}(x,t)$ that follow from it can be simplified. We clearly have

$$\left[\psi_{J}(x,t),\psi_{J'}(x',t)\right]=0,$$
and
$$\left[\psi_{J}(x,t),\psi_{J'}^{{\dagger}}(x',t)\right]=0,$$
for $J\neq J'$, since different frequency ranges are involved. In addition,
$$\left[\psi_{J}(x,t),\psi_{J}^{{\dagger}}(x',t)\right]=\int_{k\in\mathscr{R}(J)}\frac{\textrm{d}k}{2\pi}e^{i(k-k_{J})(x-x')},$$
where we have used the commutation relations (77). Now suppose, for example, $k\in \mathscr {R}(J)$ for $-\Delta \kappa /2<k-k_{J}<\Delta \kappa /2;$ then
$$\left[\psi_{J}(x,t),\psi_{J}^{{\dagger}}(x',t)\right]=\frac{1}{2\pi}\frac{\sin\left(\Delta\kappa\frac{x-x'}{2}\right)}{\frac{x-x'}{2}}.$$
We will usually be interested in incident fields and generated fields near the frequencies of ring resonances; if field excitations associated with the channel field $\psi _{J}(x,t)$ are centered at $k=k_{J}$ and extend over a range $\Delta k\ll \Delta \kappa$, then for calculations involving those fields and other independent fields we can take
$$\left[\psi_{J}(x,t),\psi_{J}^{{\dagger}}(x',t)\right]\approx\delta(x-x'),$$
and if we neglect GVD for each channel field, an expansion of the Hamiltonian $H_{J}^{\text {L}}$ (79) along the lines used in (81) leads to
$$\begin{aligned} H_{J}^{\text{L}} & =\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x,t)\psi_{J}(x,t)\textrm{d}x\\ & \quad -\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x,t)\frac{\partial\psi_{J}(x,t)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x,t)}{\partial x}\psi_{J}(x,t)\right)\textrm{d}x, \end{aligned}$$
with the neglect of GVD terms. With the full Hamiltonian (78) and the commutation relations
$$\begin{aligned} & \left[\psi_{J}(x,t),\psi_{J'}^{{\dagger}}(x',t)\right]=\delta_{JJ'}\delta(x-x'),\\ & \left[\psi_{J}(x,t),\psi_{J'}(x',t)\right]=0, \end{aligned}$$
we indeed find Eqs. (84).

To lowest order in $\Delta k/\Delta \kappa$ we can write the electromagnetic field operators (76) in terms of the channel field operators by neglecting the variation of $\omega _{Jk}$, and $\boldsymbol{b}_{Jk}(y,z)$ and $\boldsymbol{d}_{Jk}(y,z)$, over the $\Delta k$ characterizing the excitation bandwidth. Putting $\boldsymbol{b}_{J}(y,z)\equiv \boldsymbol{b}_{Jk_{J}}(y,z)$ and $\boldsymbol{d}_{J}(y,z)\equiv \boldsymbol{d}_{Jk_{J}}(y,z)$, we have

$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)\approx\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2}}\boldsymbol{d}_{J}(y,z)\psi_{J}(x,t)e^{ik_{J}x}+\textrm{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)\approx\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2}}\boldsymbol{b}_{J}(y,z)\psi_{J}(x,t)e^{ik_{J}x}+\textrm{H.c.} \end{aligned}$$
At the same level of approximation we have
$$\begin{aligned} & \boldsymbol{E}(\boldsymbol{r},t)\approx\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2}}\boldsymbol{e}_{J}(y,z)\psi_{J}(x,t)e^{ik_{J}x}+\textrm{H.c.},\\ & \boldsymbol{H}(\boldsymbol{r},t)\approx\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2}}\boldsymbol{h}_{J}(y,z)\psi_{J}(x,t)e^{ik_{J}x}+\textrm{H.c.}, \end{aligned}$$
where
$$\begin{aligned} & \boldsymbol{e}_{J}(y,z)=\frac{\boldsymbol{d}_{J}(y,z)}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{J})},\\ & \boldsymbol{h}_{J}(y,z)=\frac{\boldsymbol{b}_{J}(y,z)}{\mu_{0}}, \end{aligned}$$
[cf. (74)]. The total power flow in the waveguide is
$$P(x,t)=\int \textrm{d}y\textrm{d}z\boldsymbol{S}(\boldsymbol{r},t)\cdot\boldsymbol{\hat{x}},$$
with $\boldsymbol{S}(\boldsymbol{r},t)=\boldsymbol{E}(\boldsymbol{r},t)\times \boldsymbol{H}(\boldsymbol{r},t)$. Even a classical field oscillating at a single frequency will have terms in the Poynting vector that vary as twice that frequency; as usual we are interested in the slowly varying part of the total power flow, and so we omit terms in $\boldsymbol{S}(\boldsymbol{r},t)$ that will arise from the terms $\psi _{J}(x,t)\psi _{J'}(x,t)$ and $\psi _{J}^{\dagger }(x,t)\psi _{J'}^{\dagger }(x,t)$. We use the commutation relations (91) to write $\psi _{J}(x,t)\psi _{J'}^{\dagger }(x,t)$ as $\psi _{J}^{\dagger }(x,t)\psi _{J'}(x,t)$ plus a formally divergent [see the discussion before (89)] term proportional to $\delta _{JJ'}$ that will, however, give a vanishing contribution when channel modes going in the $+x$ and $-x$ directions are considered. For the terms remaining proportional involving $\psi _{J}^{\dagger }(x,t)\psi _{J'}(x,t)$, we neglect the contributions for $J\neq J'$ under the assumption that any distinction frequencies $\omega _{J}$ and $\omega _{J'}$ are far enough apart that interference terms between them will be rapidly varying. The final result is then [53]
$$P(x,t)\rightarrow\sum_{J}\hbar\omega_{J}v_{J}\psi_{J}^{{\dagger}}(x,t)\psi_{J}(x,t),$$
where $v_{J}$ is that of (84), and includes both material and modal dispersion; corrections to (96) would arise if we considered variations in $\psi _{J}(x,t)$ more rapid than could be well described by (84).

We now consider the channel and ring close enough that coupling can occur. The coupling arises because the evanescent fields associated with the ring and the channel overlap, and light can move from the channel to the ring and vice versa. The standard phenomenological model for this coupling [59] begins by treating the field in the ring in the same manner as one treats the field in the channel; in our notation this involves introducing a field $\psi _r(\zeta,t)$, where $\zeta =0$ at the nominal coupling point, to describe the propagation in the ring. If we let $x=0$ denote the nominal coupling point for the channel field $\psi _J(x,t)$ associated with the ring resonance of interest (see Fig. 4), we then take the Fourier components of $\psi _J(x,t)$ and $\psi _r(\zeta,t)$,

$$\psi_{J}(x,\omega)=\int\frac{\textrm{d}t}{\sqrt{2\pi}}\psi_{J}(x,t)e^{i\omega t}, \psi_{r}(\zeta,\omega)=\int\frac{\textrm{d}t}{\sqrt{2\pi}}\psi_{r}(\zeta,t)e^{i\omega t},$$
and introduce self- and cross-coupling coefficients, $\sigma$ and $\kappa$, respectively, that relate the fields near the nominal coupling point,
$$\begin{aligned} \begin{pmatrix} \psi_{J}(0^{+},\omega) \\ \psi_{r}(0^{+},\omega) \end{pmatrix} = \left( \begin{array}{cc} \ \sigma & i\kappa \\ i\kappa & \sigma \\ \end{array} \right) \begin{pmatrix} \psi_{J}(0^{-},\omega) \\ \psi_{r}(0^{-},\omega) \end{pmatrix}, \end{aligned}$$
(see Fig. 5).

 figure: Figure 5.

Figure 5. Enlarged view of the point coupler of Fig. 3 between a channel and a ring, with particular emphasis on the conventional coordinate systems used in the two waveguide.

Download Full Size | PDF

Here we assume that the waveguide properties of the ring and channel are the same, although the more general case can easily be considered. Conventionally both $\sigma$ and $\kappa$ are taken to be real and positive; then energy conservation yields

$$\sigma^{2}+\kappa^{2}=1.$$
Generalizations to this are also straightforward, including the introduction of different $\sigma$ and $\kappa$ for frequencies close to different ring resonance frequencies. Now because light propagates freely from $\zeta =0^{+}$ to $\zeta =0^{-}$ within the ring, for fields of the form (97) we have
$$\psi_{r}(0^{-},\omega)=\psi_{r}(0^{+},\omega) e^{ik\mathcal{L}},$$
where if the frequency $\omega$ of interest is close to the frequency $\omega _{J}$ of the $J$th resonance of the ring we can write
$$ \begin{aligned}k &=\frac{2\pi n_{J}}{\mathcal{L}}+(\omega-\omega_{J}) \left (\frac{dk}{d\omega} \right)_{\omega_{J}}+\cdots\\ &= \frac{2\pi n_{J}}{\mathcal{L}}+\frac{\omega-\omega_{J}}{v_{J}}+\cdots. \end{aligned}$$
Stopping the expansion at this point we have $e^{ik\mathcal {L}}=e^{i\eta }$, where
$$\eta=\left(\frac{\omega-\omega_J}{v_J} \right) \mathcal{L}.$$
This can also be derived directly from the version of (84) that would hold for $\psi _{r}(\zeta,t)$ in the ring. With this result used in (101), when that equation is inserted into (98) we can solve for $\psi _{J}(0^{+},\omega )$ in terms of $\psi _{J}(0^{-},\omega )$,
$$\psi_{J}(0^{+},\omega)=\left( \frac{\sigma-e^{i\eta}}{1-\sigma e^{i\eta}} \right) \psi_{J}(0^{-},\omega),$$
where we have used (99). Recalling that we have taken $\sigma$ to be real, from (103) we find that $|\psi _{J}(0^{+},\omega )|^2=|\psi _{J}(0^{-},\omega )|^2$. That is, the coupled channel–ring structure functions as an “all pass filter,” as would be expected.

We are typically interested in frequencies $\omega$ such that $|\omega -\omega _{J}|$ is much less than the frequency difference between neighboring resonances; then we have $|\eta | \ll 1$, and we can approximate $e^{i\eta } \approx 1+i\eta$. We are also typically interested in small ring–channel coupling, such that $\kappa \ll \sigma$; then because $|\eta |\ll 1$, we can take $\eta \sigma \approx \eta$, and using these approximations in (103) we have

$$\psi_{J}(0^{+},\omega) \approx \frac{\omega-\omega_{J}-i \left( \frac{(1-\sigma) v_{J}}{ \mathcal{L}} \right)} {\omega-\omega_{J}+i \left( \frac{(1-\sigma) v_{J}}{ \mathcal{L}} \right)} \psi_{J}(0^{-},\omega).$$
Note that the use of these two approximations respects the all-pass nature of the coupled channel–ring structure.

Although a whole range of linear and nonlinear optical phenomena in channel–ring structures can be considered at this level [59], for quantum nonlinear optics we want to construct a quantum treatment of the channel–ring coupling that can be built into a Hamiltonian framework. Beginning with the linear regime, we can do this by formally introducing a point coupling model [60], with a coupling constant $\gamma _{J}$. The full linear Hamiltonian is taken to be

$$\begin{aligned} \textrm{H}^{\textrm{L}} & =\sum_{J}\left(\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x)\psi_{J}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x)\frac{\partial\psi_{J}(x)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x)}{\partial x}\psi_{J}(x)\right)\textrm{d}x\right)\\ & \quad +\sum_{J}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J}+\sum_{J}\left(\hbar\gamma_{J}^*c_{J}^{{\dagger}}\psi_{J}\left(0\right)+\textrm{H.c.}\right). \end{aligned}$$
The first term on the right-hand side is the Hamiltonian (78) for the channel fields, the second term is the Hamiltonian (59) for the modes in the ring, and the third term represents a coupling that can either remove a photon from the channel and create one in the ring or vice versa. Here, as in the above, we consider only modes with $v_{J}>0$. If we take $\zeta =0$ to identify the point on the ring closest to the channel, then because the coupling involves the fields in the ring and channel near that point we can expect the $\gamma _{J}$ to be only weakly dependent on $J$; were $\zeta =\pi$ taken to identify that point we would expect $\gamma _{J}\propto (-1)^{J}$; in SCISSOR structures (“side-coupled integrated spaced sequences of optical resonators”), for example, both types of behavior arise [61].

The Heisenberg equations of motion following from the Hamiltonian (105) then are

$$\begin{aligned} & \left(\frac{\partial}{\partial t}+v_{J}\frac{\partial}{\partial x}+i\omega_{J}\right)\psi_{J}(x,t)={-}i\gamma_{J}c_{J}(t)\delta(x),\\ & \left(\frac{d}{dt}+i\omega_{J}\right)c_{J}(t)={-}i\gamma_{J}^{*}\psi_{J}(0,t). \end{aligned}$$
The formal solution of the first of these is
$$\psi_{J}(x,t)=\psi_{J}(x-v_Jt,0)e^{{-}i\omega_{J}t}-\frac{i\gamma_{J}}{v_{J}}c_{J}(t-\tfrac{x}{v_{J}})\left[\theta(x)-\theta(x-v_{J}t)\right] e^{{-}i\omega_{J}x/v_{J}}.$$
We then substitute this expression, evaluated at $x=0$, back into the second of (106). Note that the expression (107) respects the discontinuity of $\psi _{J}(x,t)$ across $x=0$ that was apparent in the standard phenomenological model discussed above. As was implicitly done there, we introduce one-sided limits
$$\psi_{J}(0^{{\pm}},t)=\lim_{x\rightarrow0^{{\pm}}}\psi_{J}(x,t),$$
and define $\psi _{J}(x,t)$ at $x=0$ via
$$\psi_{J}(0,t)\rightarrow\frac{1}{2}\left(\psi_{J}(0^{-},t)+\psi_{J}(0^{+},t)\right);$$
see [60]. Using the limits (108) in (107) we find
$$\psi_{J}(0^{+},t)=\psi_{J}(0^{-},t)-\frac{i\gamma_{J}}{v_{J}}c_{J}(t),$$
and then with $\psi _{J}(0,t)$ constructed from (109) and used in the second of (106) we find
$$\left(\frac{\textrm{d}}{\textrm{d}t}+\Gamma_{J}+i\omega_{J}\right)c_{J}(t)={-}i\gamma_{J}^{*}\psi_{J}(0^{-},t),$$
where
$$\Gamma_{J}=\frac{\left|\gamma_{J}\right|^{2}}{2v_{J}},$$
appears as an effective damping term in (111) characterizing the rate at which energy in the ring can be lost to the channel. The full solution of Eqs. (106) then follows from solving for $c_{J}(t)$ in terms of $\psi _{J}(0^{-},t)$ from (111), and then the determination of $\psi _{J}(0^{+},t)$ from $\psi _{J}(0^{-},t)$ from (110).

If we then introduce Fourier components of $\psi _{J}(0^{\pm },t)$ as in the first of (97), we find [60]

$$\psi_{J}(0^{+},\omega)=\left[\frac{\omega_{J}-\omega+i\Gamma_{J}}{\omega_{J}-\omega-i\Gamma_{J}}\right]\psi_{J}(0^{-},\omega),$$
This can be compared with (104), and the coupling constant $\gamma _{J}$ of the Hamiltonian model can be related to the self-coupling constant $\sigma$ of the standard phenomenological model,
$$\frac{\left|\gamma_{J}\right|^{2}}{2v_{J}}=\frac{(1-\sigma)v_J}{\mathcal{L}}.$$
Ubiquitous scattering losses lead to the break-down of the all-pass nature of this structure. We can model these by introducing a “phantom channel” [60] (see Fig. 6) that models the loss of light into the environment, and more generally the possible scattering of light into the system from the environment. Note that we indicate the positive $x$ direction in the phantom channel in the opposite direction of that in the real channel. Including the phantom channel and the coupling of light into and out of it by taking
$$\textrm{H}^{\textrm{L}}\rightarrow \textrm{H}^{\textrm{L}}+{\textrm{H}^{\textrm{phantom}}},$$
where
$$\begin{aligned} {\textrm{H}^{\textrm{phantom}}} & =\sum_{J}\left(\hbar\omega_{J}\int\psi_{J\text{ph}}^{{\dagger}}(x)\psi_{J\text{ph}}(x)\textrm{d}x \right.\\&\quad\left.-\frac{1}{2}i\hbar v_{J\text{ph}}\int\left(\psi_{J\text{ph}}^{{\dagger}}(x)\frac{\partial\psi_{J\text{ph}}(x)}{\partial x}-\frac{\partial\psi_{J\text{ph}}^{{\dagger}}(x)}{\partial x}\psi_{J\text{ph}}(x)\right)\textrm{d}x\right)\\ & \quad +\sum_{J}(\hbar \gamma_{J\text{ph}}c_{J}^{{\dagger}}\psi_{J\text{ph}}(0)+\textrm{H.c.}). \end{aligned}$$
Here the $\psi _{J\text {ph}}(x,t)$ are the fields in the phantom channel, taken to satisfy
$$\begin{aligned}& \left[\psi_{J\text{ph}}(x,t),\psi_{J'\text{ph}}^{{\dagger}}(x',t)\right]=\delta_{JJ'}\delta(x-x'),\\ & \left[\psi_{J\text{ph}}(x,t),\psi_{J'\text{ph}}(x',t)\right]=0, \end{aligned}$$
with an obvious generalization of the real channel notation for the phantom channel parameters. In place of (106) we have the more general equations
$$\begin{aligned} & \left(\frac{\partial}{\partial t}+v_{J}\frac{\partial}{\partial x}+i\omega_{J}\right)\psi_{J}(x,t)={-}i\gamma_{J}c_{J}(t)\delta(x),\\ & \left(\frac{\partial}{\partial t}+v_{J\text{ph}}\frac{\partial}{\partial x}+i\omega_{J}\right)\psi_{J\text{ph}}(x,t)={-}i\gamma_{J\text{ph}}c_{J}(t)\delta(x),\\ & \left(\frac{\textrm{d}}{\textrm{d}t}+i\omega_{J}\right)c_{J}(t)={-}i\gamma_{J}^{*}\psi_{J}(0,t)-i\gamma_{J\text{ph}}^{*}\psi_{J\text{ph}}(0,t). \end{aligned}$$
Following the same procedure as used above but taking both the real and phantom channel into account, and introducing
$$\begin{aligned}& \psi_{J\text{ph}}(0^{{\pm}},\omega)=\int\frac{\textrm{d}t}{\sqrt{2\pi}}\psi_{J\text{ph}}(0^{{\pm}},t)e^{i\omega t},\\ & c_{J}(\omega)=\int\frac{\textrm{d}t}{\sqrt{2\pi}}c_{J}(t)e^{i\omega t}, \end{aligned}$$
we find
$$c_{J}(\omega)={-} \frac{\gamma_{J}^{*}\psi_{J}(0^{-},\omega)+\gamma_{J\text{ph}}^{*}\psi_{J\text{ph}}(0^{-},\omega)}{\omega_{J}-\omega-i\overline{\Gamma}_{J}},$$
where
$$\overline{\Gamma}_{J}=\Gamma_{J}+\Gamma_{J\text{ph}},$$
and
$$\Gamma_{J\text{ph}}=\frac{\left|\gamma_{J\text{ph}}\right|^{2}}{2v_{J\text{ph}}},$$
describes the scattering losses. For the usual situation where $\left \langle \psi _{J\text {ph}}(0^{-},\omega )\right \rangle =0$ we have
$$|\left\langle c_{J}(\omega)\right\rangle|^{2}=\frac{|\gamma_{J}^{*}\left\langle\psi_{J}(0^{-},\omega)\right\rangle|^{2}}{(\omega_{J}-\omega)^{2}+\overline{\Gamma}_{J}^{2}}.$$
From this we can identify the full-width at half-maximum as $\Delta \omega _{J}=2\overline {\Gamma }_{J}$, and so the “loaded” quality factor $Q$ for the resonance is given by
$$Q_{J}=\frac{\omega_{J}}{\Delta\omega_{J}}=\frac{\omega_{J}}{2\overline{\Gamma}_{J}}.$$
Defining an intrinsic $Q_{J\text {int}}$ due to scattering losses and an extrinsic $Q_{J\text {ext}}$ due to the coupling of the ring to the channel, we have
$$\frac{1}{Q_{J}}=\frac{1}{Q_{J\text{int}}}+\frac{1}{Q_{J\text{ext}}},$$
where
$$Q_{J\text{ext}}=\frac{\omega_{J}}{2{\Gamma}_{J}}$$
and
$$Q_{J\text{int}}=\frac{\omega_{J}}{2{\Gamma}_{J\text{ph}}}.$$

 figure: Figure 6.

Figure 6. Schematic representation of a point-coupled ring resonator with and additional fictitious coupler and waveguide, referred to as “phantom channel.”

Download Full Size | PDF

The solution of our general equations (118), again following the same procedure as used above, gives

$$\psi_{J}(0^{+},\omega)=\left[\frac{\omega_{J}-\omega+i(\Gamma_{J}-\Gamma_{J\text{ph}})}{\omega_{J}-\omega-i\overline{\Gamma}_{J}}\right],\,\psi_{J}(0^{-},\omega) +\left[\frac{2i\Gamma_{J\text{ph}}}{\omega_{J}-\omega-i\overline{\Gamma}_{J}}\right]\psi_{J\text{ph}}(0^{-},\omega)$$
where the term proportional to $\psi _{J\text {ph}}(0^{-},\omega )$ in (128) describes the input noise from the loss channel; and again for the usual situation where $\left \langle \psi _{J\text {ph}}(0^{-},\omega )\right \rangle =0$ we have
$$\left\langle \psi_{J}(0^{+},\omega)\right\rangle =\left[\frac{\omega_{J}-\omega+i(\Gamma_{J}-\Gamma_{J\text{ph}})}{\omega_{J}-\omega-i\overline{\Gamma}_{J}}\right]\left\langle \psi_{J}(0^{-},\omega)\right\rangle,$$
and so-called “critical coupling” is achieved if $\Gamma _{J}=\Gamma _{J\text {ph}}$, for then $\left \langle \psi _{J}(0^{+},\omega _{J})\right \rangle =0.$

To avoid dealing with the fields $\psi _J(x,t)$ that have a discontinuity at $x=0$, it is convenient to introduce fields $\psi _{J<}(x,t)$ and $\psi _{J>}(x,t)$ associated with each $\psi _J(x,t)$. The first of these is equal to $\psi _J(x,t)$ for $x<0$ and, for $x\ge 0$, follows from the evolution equation (84) that would hold were there no coupling to the ring. Similarly, $\psi _{J>}(x,t)$ is given by $\psi _J(x,t)$ for $x>0$ and, for $x\le 0$, follows from that evolution equation. Thus, the fields $\psi _{J<}(x,t)$ and $\psi _{J>}(x,t)$ are continuous even at $x=0$, and we have $\psi _J(0^-,t)=\psi _{J<}(0,t)$ and $\psi _J(0^+,t)=\psi _{J>}(0,t)$.

Finally, because the free spectral range (FSR) in the neighborhood of the $J$th resonance is given by

$$\delta\omega_{J}=v_{J}\delta\kappa_{J}=v_J\frac{2\pi}{\mathcal{L}},$$
the finesse in the neighborhood of the $J$th resonance is
$$\mathcal{F}=\frac{\delta\omega_{J}}{\Delta\omega_{J}}=\frac{\pi v_{J}}{\overline{\Gamma}_{J} \mathcal{L}}.$$
From this expression one can see that the finesse is a measurement of the effect of the resonant spatial and temporal confinement of light inside the ring. Indeed, it is inversely proportional to the resonator length and is directly proportional to the photon dwelling time in the resonator. This quantity can be very useful, because it is experimentally accessible from the structure transmission spectrum, which gives information about the FSR and the quality factor, and yet is also simply related to the on-resonance field enhancement, $|\text {FE}(\omega _J)|$. This can be defined as the modulus of the ratio between the value of the displacement field measured in the center of the waveguide in any point of the ring and the value measured in the center of the channel waveguide before the point coupling. One can show that, at critical coupling (i.e., $\Gamma _J=\Gamma _{J\text {ph}}$),
$$\left\vert\text{FE}_\text{crit}\left(\omega_J\right)\right\vert^{2}\simeq\frac{\mathcal{F}}{\pi},$$
which means that at critical coupling the intensity inside the ring is increased by a factor approximately equal to a third of the finesse, with respect to that in the input channel. A similar expression can be derived in the case of over coupling (i.e., $\Gamma _J\gg \Gamma _{J\text {ph}}$)
$$\left\vert\text{FE}_\text{over}\left(\omega_J\right)\right\vert^{2}\simeq\frac{2}{\pi}\mathcal{F}.$$
Although the point coupling model introduced above has a wide range of applicability for ring–channel systems [59], we note that it is also possible to construct a more general treatment of the response of these systems by the use of the asymptotic-in and -out fields familiar from scattering theory [62], which also fits within the Hamiltonian framework we describe here.

2.4 Including Nonlinearities

We now turn to including the effects of nonlinearities. Considering a very general dielectric structure, and beginning in the classical regime, the usual description of the nonlinear response is in terms of nonlinear response coefficients $\chi _{2}^{ijk}(\boldsymbol{r})$ and $\chi _{3}^{ijkl}(\boldsymbol{r}),$ etc., with the components of the nonlinear polarization $\boldsymbol{P}_{\text {NL}}(\boldsymbol{r},t)$ given by

$$P_{\text{NL}}^{i}(\boldsymbol{r},t)=\epsilon_{0}\chi_{2}^{ijk}(\boldsymbol{r})E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t) +\epsilon_{0}\chi_{3}^{ijkl}(\boldsymbol{r})E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t)E^{l}(\boldsymbol{r},t)+\cdots.$$
The displacement field $\boldsymbol{D}(\boldsymbol{r},t)$ is then given by $\boldsymbol{D}(\boldsymbol{r},t)=\epsilon _{0}\boldsymbol{E}(\boldsymbol{r},t)+\boldsymbol{P}(\boldsymbol{r},t),$ where $\boldsymbol{P}(\boldsymbol{r},t)$ is the full polarization, so
$$D^{i}(\boldsymbol{r},t)=\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})E^{i}(\boldsymbol{r},t)+\epsilon_{0}\chi_{2}^{ijk}(\boldsymbol{r})E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t) +\epsilon_{0}\chi_{3}^{ijkl}(\boldsymbol{r})E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t)E^{l}(\boldsymbol{r},t)+\cdots.$$
Here we have neglected material dispersion, and will formally do so in the nonlinear components of the Hamiltonian in the rest of this tutorial. It many cases the effects of material dispersion on the nonlinear components can be reasonably supposed to be negligible. Of course, when particular nonlinear terms are calculated the appropriate $\chi _{n}$ for the frequencies of the fields involved can be employed; we implicitly assume that in the following. In addition, in the following equations that involve $\varepsilon _{1}(\boldsymbol{r})$, we use the $\varepsilon _{1}(\boldsymbol{r};\omega )$ for the frequencies involved. However, the full inclusion of dispersive effects in the $\chi _{n}$ themselves, for $n>1$, is an outstanding problem even at frequencies where absorption is not present.

Equation (135) is a usual starting point for investigating the propagation of light in nonlinear media. However, because our fundamental fields are taken to be $\boldsymbol{D}(\boldsymbol{r},t)$ and $\boldsymbol{B}(\boldsymbol{r},t)$, we want to rewrite this as an expression for $\boldsymbol{E}(\boldsymbol{r},t)$ in terms of $\boldsymbol{D}(\boldsymbol{r},t).$ To do that, write (135) as

$$E^{i}(\boldsymbol{r},t)=\frac{D^{i}(\boldsymbol{r},t)}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}-\frac{\chi_{2}^{ijk}(\boldsymbol{r})}{\varepsilon_{1}(\boldsymbol{r})} E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t)-\frac{\chi_{3}^{ijkl}(\boldsymbol{r})}{\varepsilon_{1}(\boldsymbol{r})}E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t)E^{l}(\boldsymbol{r},t)+\cdots.$$
In the absence of any nonlinearity $E^{i}(\boldsymbol{r},t)=D^{i}(\boldsymbol{r},t)/[\epsilon _{0}\varepsilon _{1}(\boldsymbol{r})]$, of course, and we can iterate to find
$$E^{i}(\boldsymbol{r},t)=\frac{D^{j}(\boldsymbol{r},t)}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}-\frac{\Gamma_{2}^{ijk}(\boldsymbol{r})}{\epsilon_{0}}D^{j} (\boldsymbol{r},t)D^{k}(\boldsymbol{r},t)-\frac{\Gamma_{3}^{ijkl}(\boldsymbol{r})}{\epsilon_{0}}D^{j}(\boldsymbol{r},t)D^{k}(\boldsymbol{r},t)D^{l}(\boldsymbol{r},t)+\cdots,$$
where
$$\begin{aligned} & \Gamma_{2}^{ijk}(\boldsymbol{r})=\frac{\chi_{2}^{ijk}(\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}^{3}(\boldsymbol{r})},\\ & \Gamma_{3}^{ijkl}(\boldsymbol{r})=\frac{\chi_{3}^{ijkl}(\boldsymbol{r})}{\epsilon_{0}^{2}\varepsilon_{1}^{4}(\boldsymbol{r})}-2\frac{\chi_{2}^{ijm}(\boldsymbol{r}) \chi_{2}^{mkl}(\boldsymbol{r})}{\epsilon_{0}^{2}\varepsilon_{1}^{5}(\boldsymbol{r})}, \end{aligned}$$
and we have used the fact that the $\chi _{n}(\boldsymbol{r})$ are invariant under permutation of their Cartesian components. Note that the $\Gamma _{n}$, which are dimensionless, are then also invariant under permutation of their Cartesian components; they are typically positive. Using (137) in the expression (4) we find
$$\begin{aligned}H & =\int\left(\frac{\boldsymbol{D}(\boldsymbol{r},t)\cdot\boldsymbol{D}(\boldsymbol{r},t)}{2\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}+\frac{\boldsymbol{B}(\boldsymbol{r},t)\cdot\boldsymbol{B}(\boldsymbol{r},t)}{2\mu_{0}}\right)\textrm{d}\boldsymbol{r}\\ & \quad -\frac{1}{3\epsilon_{0}}\int\Gamma_{2}^{ijk}(\boldsymbol{r})D^{i}(\boldsymbol{r},t)D^{j}(\boldsymbol{r},t)D^{k}(\boldsymbol{r},t)\textrm{d}\boldsymbol{r}\\ & \quad -\frac{1}{4\epsilon_{0}}\int\Gamma_{3}^{ijkl}(\boldsymbol{r})D^{i}(\boldsymbol{r},t)D^{j}(\boldsymbol{r},t)D^{k}(\boldsymbol{r},t)D^{l}(\boldsymbol{r},t)\textrm{d}\boldsymbol{r}+\cdots, \end{aligned}$$
for the total energy. Adopting this as the Hamiltonian, and using either the Poisson bracket expressions (31) (in the classical regime), or the corresponding commutation relations (34) (in the quantum regime), we find that the Hamiltonian leads to the dynamics
$$\begin{aligned}& \frac{\partial\boldsymbol{D}(\boldsymbol{r},t)}{\partial t}=\frac{1}{\mu_{0}}\boldsymbol{\nabla}\times\boldsymbol{B}(\boldsymbol{r},t),\\ & \frac{\partial\boldsymbol{B}(\boldsymbol{r},t)}{\partial t}={-}\boldsymbol{\nabla}\times\boldsymbol{E}(\boldsymbol{r},t), \end{aligned}$$
where the components of $\boldsymbol{E}(\boldsymbol{r},t)$ are given by (137). For the very general mode expansion (38) we then find the Schrödinger Hamiltonian
$$H=\sum_{\alpha}\hbar\omega_{\alpha}c_{\alpha}^{{\dagger}}c_{\alpha}+{\textrm{H}^{\textrm{NL}}},$$
where
$$\begin{aligned} {\textrm{H}^{\textrm{NL}}} & ={-}\frac{1}{3\epsilon_{0}}\int\Gamma_{2}^{ijk}(\boldsymbol{r}):\left(D^{i}(\boldsymbol{r})D^{j}(\boldsymbol{r})D^{k}(\boldsymbol{r})\right):\textrm{d}\boldsymbol{r}\\ & \quad -\frac{1}{4\epsilon_{0}}\int\Gamma_{3}^{ijkl}(\boldsymbol{r}):\left(D^{i}(\boldsymbol{r})D^{j}(\boldsymbol{r})D^{k}(\boldsymbol{r})D^{l}(\boldsymbol{r})\right):\textrm{d}\boldsymbol{r}+\cdots. \end{aligned}$$
In (141) we have neglected the zero point energy in the linear regime, and in (142) the notation “$:(\quad ):$” indicates that the operator expressions inside should be normal ordered. The normal ordering of a term will lead to new terms involving fewer raising and lowering operators than the original term, and those new terms can be considered as corrections to terms in the full Hamiltonian that, before normal ordering, involved those fewer raising and lowering operators. We consider all such corrections to be included in the values of the appropriate material coefficients that we adopt.

Using an expansion of $\boldsymbol{D}(\boldsymbol{r})$ in terms of the linear modes of the system the form of ${\textrm{H}^{\textrm{NL}}}$ in terms of raising and lowering operators can be explicitly constructed. We do this later in this tutorial for SPDC and SFWM. In the rest of this section we treat the special cases of SPM and XPM. For simplicity we assume that $\Gamma _{3}^{ijkl}$ receives significant contributions only from $\chi _{3}^{ijkl}$, and not from $\chi _{2}^{ijk}$ [see (138)]; that usually holds, but if it does not the generalization of the formulas given in the following is straightforward.

2.4.1 SPM and XPM in Channels

Considering an isolated channel, we still assume that we can take our fields to be of the form (92), although the frequencies $\omega _{J}$ here are not linked to any ring resonances, but merely identify the center frequencies of different components of light that are well-separated from each other in frequency. Assuming that we have just one such component; taking $J=P$, its self-interaction is described by $\chi _{3}^{ijkl}(\boldsymbol{r})$, and using (92) in (142) the term describing the resulting SPM is

$$\begin{aligned} {H}_{\textrm{chan}}^{\textrm{SPM}} & ={-}\frac{1}{4\epsilon_{0}}\left(\frac{\hbar\omega_{P}}{2}\right)^{2}\frac{4!}{2!2!}\left(\int \textrm{d}y\textrm{d}z\Gamma_{3}^{ijkl}(y,z)\left(d_{P}^{i}(y,z)d_{P}^{j}(y,z)\right)^{*}d_{P}^{k}(y,z)d_{P}^{l}(y,z)\right)\\ & \quad\times\int \textrm{d}x\psi_{P}^{{\dagger}}(x)\psi_{P}^{{\dagger}}(x)\psi_{P}(x)\psi_{P}(x), \end{aligned}$$
where the combinatorial factor $4!/(2!2!)$ counts the number of ways, after normal ordering, that the indicated combination of fields $\psi _{P}(x)$ and $\psi _{P}^{\dagger }(x)$ arise. We use (138) to write $\Gamma _{3}^{ijkl}(y,z)$ in terms of $\chi _{3}^{ijkl}(y,z)$, and use (94) to write the fields $\boldsymbol{d}_{P}(y,z)$ in terms of $\boldsymbol{e}_{P}(y,z)$. Then (143) can be written as
$${H}_{\textrm{chan}}^{\textrm{SPM}} ={-}{\gamma_{\textrm{chan}}^{\textrm{SPM}}}\frac{\hbar^{2}\omega_{P}}{2}v_{P}^{2}\int \textrm{d}x\psi_{P}^{{\dagger}}(x)\psi_{P}^{{\dagger}}(x)\psi_{P}(x)\psi_{P}(x),$$
where
$${\gamma_{\textrm{chan}}^{\textrm{SPM}}}=\frac{3\omega_{P}\epsilon_{0}}{4v_{P}^{2}}\int \textrm{d}y\textrm{d}z\chi_{3}^{ijkl}(y,z)\left(e_{P}^{i}(y,z)e_{P}^{j}(y,z)\right)^{*}e_{P}^{k}(y,z)e_{P}^{l}(y,z).$$
The approximations involved in deriving (145), and other equations like it in the following, lead to nonlinear coefficients evaluated only at a center frequency, despite the fact that the light may be broadband. No simple prescription can be given for determining the validity of this assumption, for it will depend on the nature of the waveguide modes, how close one might be to a cutoff frequency, etc. However, a first check of the validity of the assumption can be made by evaluating (145) at a range of frequencies corresponding to the pulse bandwidth, and checking to see whether the variation in its value is indeed small.

Taking as our Hamiltonian the sum of the linear Hamiltonian (90) and the nonlinear contribution (144) we find the Heisenberg equations of motion yield

$$\frac{\partial\psi_{P}(x,t)}{\partial t}={-}i\omega_{P}\psi_{P}(x,t)-v_{P}\frac{\partial\psi_{P}(x,t)}{\partial x}+i{\gamma_{\textrm{chan}}^{\textrm{SPM}}}\hbar\omega_{P}v_{P}^{2}\psi_{P}^{{\dagger}}(x,t)\psi_{P}(x,t)\psi_{P}(x,t).$$
In the classical limit, letting
$$\begin{aligned} & \psi_{P}(x,t)\rightarrow\phi_{P}(x,t),\\ & \psi_{P}^{{\dagger}}(x,t)\rightarrow\phi_{P}^{*}(x,t), \end{aligned}$$
where $\phi _{P}(x,t)$ is a classical field, we find a solution of the form $\phi _{P}(x,t)=\phi _{P}(x)\exp (-i\omega _{P}t)$, where
$$\begin{aligned} \frac{\textrm{d}\phi_{P}(x)}{\textrm{d}x} & =i{\gamma_{\textrm{chan}}^{\textrm{SPM}}}\hbar\omega_{P}v_{P}\left|\phi_{P}(x)\right|^{2}\phi_{P}(x)\\ & =i{\gamma_{\textrm{chan}}^{\textrm{SPM}}} P_{P}(x)\phi_{P}(x), \end{aligned}$$
where $P_{P}(z)$ is the power in the channel [cf. (96)]. Thus, we can identify ${\gamma_{\textrm{chan}}^{\textrm{SPM}}}$ as the SPM coefficient for the channel [63].

To write ${\gamma_{\textrm{chan}}^{\textrm{SPM}}}$ in a more transparent form, divide the expression (145) for it by the square of

$$\epsilon_{0}c\int\frac{n(y,z;\omega_{P})}{v_{g}(y,z;\omega_{P})}\boldsymbol{e}_{P}^{*}(y,z)\cdot\boldsymbol{e}_{P}(y,z)\textrm{d}y\textrm{d}z=1,$$
where the identify follows from the normalization condition (70) and the relation (94) between $\boldsymbol{d}_{P}(y,z)$ and $\boldsymbol{e}_{P}(y,z)$, as well as the expression (68) for the local phase velocity; we have also introduced the local index of refraction $n(y,z;\omega )=\sqrt {\varepsilon _{1}(y,z;\omega )}$. Then taking $\overline {n}_{P}$ to be a characteristic value of that index in the region where $\boldsymbol{e}_{P}(y,z)$ is nonvanishing, and $\overline {\chi }_{3}$ to be a characteristic value of the nonlinear susceptibility tensor components in that region, we can combine (145) and (149) to write
$${\gamma_{\textrm{chan}}^{\textrm{SPM}}}=\frac{3\omega_{P}\overline{\chi}_{3}}{4\epsilon_{0}\overline{n}_{P}^{2}c^{2}}\frac{1}{{\textrm{A}_{\textrm{chan}}^{PPPP}}},$$
where the effective area ${\textrm{A}_{\textrm{chan}}^{PPPP}}$ is given by
$$\frac{1}{{\textrm{A}_{\textrm{chan}}^{PPPP}}}=\frac{\int \textrm{d}y\textrm{d}z\left(\chi_{3}^{ijkl}(y,z)/\overline{\chi}_{3}\right)\left(e_{P}^{i}(y,z)e_{P}^{j}(y,z)\right)^{*}e_{P}^{k}(y,z)e_{P}^{l} (y,z)}{\left(\int\frac{n(y,z;\omega_{P})/\overline{n}_{P}}{v_{g}(y,z;\omega_{P})/v_{P}}\boldsymbol{e}_{P}^{*}(y,z)\cdot\boldsymbol{e}_{P}(y,z)\textrm{d}y\textrm{d}z\right)^{2}}.$$
Note that if we define an effective nonlinear index $n_{2}$ in terms of $\overline {\chi }_{3}$ and $\overline {n}_{P}$ in the usual way [63]
$$n_{2}=\frac{3\overline{\chi}_{3}}{4\epsilon_{0}c\overline{n}_{P}^{2}},$$
as is appropriate for a uniform medium, we have
$${\gamma_{\textrm{chan}}^{\textrm{SPM}}}=\frac{\omega_{P}}{c}n_{2}\frac{1}{{\textrm{A}_{\textrm{chan}}^{PPPP}}},$$
the usual expression for ${\gamma_{\textrm{chan}}^{\textrm{SPM}}}$ in terms of an effective area [63]. Further, if we consider an electric field with only one component, put $\overline {\chi }_{3}$ equal to a typical value of the relevant component of $\chi _{3}^{ijkl}(y,z)$ in the region where $\boldsymbol{e}_{P}(y,z)$ is nonvanishing, and set $n(y,z;\omega _{P})/\overline {n}_{P}\approx v_{g}(y,z;\omega _{P})/v_{P}\approx 1$ as well in that region, we have
$$\frac{1}{{\textrm{A}_{\textrm{chan}}^{PPPP}}}\approx\frac{\int\left|e_{P}(y,z)\right|^{4}\textrm{d}y\textrm{d}z}{\left(\int\left|e_{P}(y,z)\right|^{2}\textrm{d}y\textrm{d}z\right)^{2}},$$
the usual expression for the effective area in that limit [63].

We now consider XPM, a consequence of the nonlinear interaction between a field $\psi _{S}(x,t)$ with frequency centered at $\omega _{S}$ and a field $\psi _{P}(x,t)$ with frequency centered at $\omega _{P}$. The contribution to the nonlinear Hamiltonian responsible for XPM is

$$\begin{aligned} {H}_{\textrm{chan}}^{\textrm{XPM}} & ={-}\frac{1}{4\epsilon_{0}}\left(\frac{\hbar\omega_{P}}{2}\right)\left(\frac{\hbar\omega_{S}}{2}\right)4!\left(\int \textrm{d}y\textrm{d}z\Gamma_{3}^{ijkl}(y,z)\left(d_{P}^{i}(y,z)d_{S}^{j}(y,z)\right)^{*}d_{P}^{k}(y,z)d_{S}^{l}(y,z)\right)\\ & \quad\times\int \textrm{d}x\psi_{P}^{{\dagger}}(x)\psi_{S}^{{\dagger}}(x)\psi_{P}(x)\psi_{S}(x), \end{aligned}$$
where here the combinatorial factor arising from normal ordering is $4!$, because there are four distinct field operators. In many applications one of the fields is strong, say that centered at $\omega _{P}$, and we are interested in describing its effect on the phase of the weaker field, say that centered at $\omega _{S}$. We introduce a XPM coefficient ${\gamma_{\text{chan}}^{\text{XPM}}}$ that is appropriate in that scenario,
$${\gamma_{\text{chan}}^{\text{XPM}}}=\frac{3\epsilon_{0}\omega_{S}}{4v_{P}v_{S}}\int \textrm{d}y\textrm{d}z\chi_{3}^{ijkl}(y,z)\left(e_{P}^{i}(y,z)e_{S}^{j}(y,z)\right)^{*}e_{P}^{k}(y,z)e_{S}^{l}(y,z),$$
[cf. the expression (145) for ${\gamma_{\textrm{chan}}^{\textrm{SPM}}}$] in terms of which we can write
$${H}_{\textrm{chan}}^{\textrm{XPM}}={-}2{\gamma_{\text{chan}}^{\text{XPM}}}\hbar^{2}\omega_{P}v_{P}v_{S}\int \textrm{d}x\psi_{P}^{{\dagger}}(x)\psi_{S}^{{\dagger}}(x)\psi_{P}(x)\psi_{S}(x).$$
Following the same strategy used above (146), (147), and (148) in identifying the physical significance of ${\gamma_{\textrm{chan}}^{\textrm{SPM}}}$, taking the Hamiltonian to be the sum of the linear terms (90) and the nonlinear contribution (157), in the classical limit we find
$$\begin{aligned}\frac{\textrm{d}\phi_{S}(x)}{\textrm{d}x} & =2i{\gamma_{\text{chan}}^{\text{XPM}}}\hbar\omega_{P}v_{P}\left|\phi_{P}(x)\right|^{2}\phi_{S}(x)\\ & =2i{\gamma_{\text{chan}}^{\text{XPM}}} P_{P}(x)\phi_{S}(x), \end{aligned}$$
with the usual factor of $2$ for XPM explicit [cf. (148)]. Dividing (156) by unity using the left-hand side of (149), once at $\omega _{P}$ and once at $\omega _{S}$, we find
$${\gamma_{\text{chan}}^{\text{XPM}}}=\frac{3\omega_{S}\overline{\chi}_{3}}{4\epsilon_{0}\overline{n}_{P}\overline{n}_{S}c^{2}}\frac{1}{{\textrm{A}_{\textrm{chan}}^{PSPS}}},$$
where
$$\frac{1}{{\textrm{A}_{\textrm{chan}}^{PSPS}}}=\frac{\int \textrm{d}y\textrm{d}z\left(\chi_{3}^{ijkl}(y,z)/\overline{\chi}_{3}\right)\left(e_{P}^{i}(y,z)e_{S}^{j}(y,z)\right)^{*}e_{P}^{k}(y,z)e_{S}^{l} (y,z)}{\left(\int\frac{n(y,z;\omega_{P})/\overline{n}_{P}}{v_{g}(y,z;\omega_{P})/v_{P}}\boldsymbol{e}_{P}^{*}(y,z)\cdot\boldsymbol{e}_{P}(y,z)\textrm{d}y\textrm{d}z\right)\left(\int\frac{n(y,z;\omega_{S})/\overline{n}_{S}}{v_{g}(y,z;\omega_{S})/v_{S}}\boldsymbol{e}_{S}^{*}(y,z)\cdot\boldsymbol{e}_{S}(y,z)\textrm{d}y\textrm{d}z\right)},$$
with $\overline {n}_{P}$ and $\overline {n}_{S}$ here typical values of $n(y,z;\omega _{P})$ and $n(y,z;\omega _{S})$, respectively, in regions where the associated electric fields are nonvanishing. Comparing these equations with the corresponding expressions (150) and (151) for SPM, we see that for $\omega _{P}$ and $\omega _{S}$ close we have ${\gamma_{\textrm{chan}}^{\textrm{SPM}}} \approx {\gamma_{\text{chan}}^{\text{XPM}}}$, as expected.

2.4.2 SPM and XPM in Rings

In a ring coupled to a channel, the usual approximation is to assume that the important nonlinear interaction is in the ring, because in usual applications the field is concentrated there. Thus, to construct the appropriate Hamiltonian for SPM here we use the expression (57) for $\boldsymbol{D}(\boldsymbol{r})$ in the ring, keeping the contribution from just one mode $P$ with frequency $\omega _{P}$, in the general nonlinear Hamiltonian (142). Then keeping the contribution from $\chi _{3}$, the form of the nonlinear Hamiltonian closely follows that for a channel, and we have

$${\begin{aligned} H_{\text{ring}}^{\text{SPM}} & ={-}\frac{1}{4\epsilon_{0}}\left(\frac{\hbar\omega_{P}}{2\mathcal{L}}\right)^{2}\frac{4!}{2!2!}\left(\int \textrm{d}\boldsymbol{r}_{{\perp}}d\zeta\Gamma_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})\left(\textsf{d}_{P}^{i}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{d}_{P}^{j} (\boldsymbol{r}_{{\perp}};\zeta)\right)^{*}\textsf{d}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{d}_{P}^{l}(\boldsymbol{r}_{{\perp}};\zeta)\right)\\ & \quad\times c_{P}^{{\dagger}}c_{P}^{{\dagger}}c_{P}c_{P}. \end{aligned}}$$
We can write this as
$$H_{\text{ring}}^{\text{SPM}}={-}\frac{\hbar^{2}\omega_{P}v_{P}^{2}{\gamma_{\textrm{ring}}^{\textrm{SPM}}}}{2\mathcal{L}}c_{P}^{{\dagger}}c_{P}^{{\dagger}}c_{P}c_{P},$$
where for convenience later we have introduced $v_{P}$ as the group velocity at $\omega _{P}$ for light propagating in a channel with a structure matching that of the ring, and
$${\gamma_{\textrm{ring}}^{\textrm{SPM}}}=\frac{3\omega_{P}\epsilon_{0}}{4v_{P}^{2}}\frac{1}{\mathcal{L}}\int\chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})\left(\textsf{e}_{P}^{i} (\boldsymbol{r}_{{\perp}};\zeta)\textsf{e}_{P}^{j}(\boldsymbol{r}_{{\perp}};\zeta)\right)^{*}\textsf{e}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta) \textsf{e}_{P}^{l}(\boldsymbol{r}_{{\perp}};\zeta)\textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta,$$
where we have followed the procedure used for the channel and written $\textsf {d}_{P}(\boldsymbol{r}_{\perp };\zeta )=\epsilon _{0}n^{2}(\boldsymbol{r}_{\perp };\omega _{P})\textsf {e}_{P}(\boldsymbol{r}_{\perp };\zeta ),$ and we have put $n(\boldsymbol{r}_{\perp };\omega )=\sqrt {\varepsilon _{1}(\boldsymbol{r}_{\perp };\omega )}$. We then use the normalization condition (75) to write
$$\epsilon_{0}c\int\frac{n(\boldsymbol{r}_{{\perp}};\omega_{P})}{v_{g}(\boldsymbol{r}_{{\perp}};\omega_{P})}\textsf{e}_{P}^{*}(\boldsymbol{r}_{{\perp}};0) \cdot\textsf{e}_{P}(\boldsymbol{r}_{{\perp}};0)\textrm{d}\boldsymbol{r}_{{\perp}}=1,$$
and again introducing characteristic values $\overline {n}_{P}$ and $\overline {\chi }_{3}$ for, respectively, $n(\boldsymbol{r}_{\perp };\omega _{P})$ and the nonvanishing components of $\chi _{3}^{ijkl}(\boldsymbol{r}_{\perp })$ in the region where these quantities are important in the integrals appearing in (163) and (164), we can write ${\gamma_{\textrm{ring}}^{\textrm{SPM}}}$ as
$${\gamma_{\textrm{ring}}^{\textrm{SPM}}}=\frac{3\omega_{P}\overline{\chi}_{3}}{4\epsilon_{0}\overline{n}_{P}^{2}c^{2}}\frac{1}{{\textrm{A}_{\textrm{ring}}^{PPPP}}},$$
where
$$\frac{1}{{\textrm{A}_{\textrm{ring}}^{PPPP}}}=\frac{\mathcal{L}^{{-}1}\int\left(\chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})/\overline{\chi}_{3}\right)\left(\textsf{e}_{P}^{i}(\boldsymbol{r}_{{\perp}};\zeta) \textsf{e}_{P}^{j}(\boldsymbol{r}_{{\perp}};\zeta)\right)^{*}\textsf{e}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{e}_{P}^{l}(\boldsymbol{r}_{{\perp}};\zeta) \textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta}{\left(\int\frac{n(\boldsymbol{r}_{{\perp}};\omega_{P})/\overline{n}_{P}}{v_{g}(\boldsymbol{r}_{{\perp}};\omega_{P})/v_{P}}\textsf{e}_{P}^{*} (\boldsymbol{r}_{{\perp}};0)\cdot\textsf{e}_{P}(\boldsymbol{r}_{{\perp}};0)\textrm{d}\boldsymbol{r}_{{\perp}}\right)^{2}}.$$
Comparing the results for a ring with those (150) and (151) for a channel, we see that if the ring structure is just the channel structure formed in a ring, and if the ring is large enough that the mode fields are approximately the same (62) then we can expect ${\gamma_{\textrm{ring}}^{\textrm{SPM}}} \approx {\gamma_{\textrm{chan}}^{\textrm{SPM}}}$ if the electric field amplitude is largely in the $\boldsymbol{\hat {z}}$ direction. However, if other components are also important, differences between ${\gamma_{\textrm{ring}}^{\textrm{SPM}}}$ and ${\gamma_{\textrm{chan}}^{\textrm{SPM}}}$ can arise, depending on the nonvanishing components of $\chi _{3}^{ijkl}$ and their relative sizes.

The treatment of XPM also follows that of the channel. Considering ring fields at $\omega _{P}$ and $\omega _{S}$, the contribution to the nonlinear Hamiltonian associated with XPM is

$${\begin{aligned}H_{\text{ring}}^{\text{XPM}} & ={-}\frac{1}{4\epsilon_{0}}\left(\frac{\hbar\omega_{P}}{2\mathcal{L}}\right)\left(\frac{\hbar\omega_{S}}{2\mathcal{L}}\right)4!\left(\int \textrm{d}\boldsymbol{r}_{{\perp}}d\zeta\Gamma_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})\left(\textsf{d}_{P}^{i}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{d}_{S}^{j}(\boldsymbol{r}_{{\perp}};\zeta) \right)^{*}\textsf{d}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{d}_{S}^{l}(\boldsymbol{r}_{{\perp}};\zeta)\right)\\ & \quad\times c_{P}^{{\dagger}}c_{S}^{{\dagger}}c_{P}c_{S}. \end{aligned}}$$
As in the channel calculation, we introduce a nonlinear coefficient ${\gamma_{\textrm{ring}}^{\textrm{XPM}}}$ appropriate to describe XPM where there is a strong field at $\omega _{P}$ modifying the phase of one at $\omega _{S}$, writing
$$H_{\text{ring}}^{\text{XPM}}={-}\frac{2{\gamma_{\textrm{ring}}^{\textrm{XPM}}}\hbar^{2}\omega_{P}v_{P}v_{S}}{\mathcal{L}}c_{P}^{{\dagger}}c_{S}^{{\dagger}}c_{P}c_{S},$$
where
$${\gamma_{\textrm{ring}}^{\textrm{XPM}}}=\frac{3\omega_{S}\epsilon_{0}}{4v_{P}v_{S}}\frac{1}{\mathcal{L}}\int\chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})\left(\textsf{e}_{P}^{i}(\boldsymbol{r}_{{\perp}};\zeta) \textsf{e}_{S}^{j}(\boldsymbol{r}_{{\perp}};\zeta)\right)^{*}\textsf{e}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{e}_{S}^{l}(\boldsymbol{r}_{{\perp}};\zeta)\textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta,$$
where as for ${\gamma_{\textrm{ring}}^{\textrm{XPM}}}$ we have written the expression in terms of the $\textsf {e}_{P,S}(\boldsymbol{r_{\perp }};\zeta )$. Again using the normalization condition (164), here for both $\omega _{P}$ and $\omega _{S}$, we can write
$${\gamma_{\textrm{ring}}^{\textrm{XPM}}}=\frac{3\omega_{S}\overline{\chi}_{3}}{4\epsilon_{0}\overline{n}_{P}\overline{n}_{S}c^{2}}\frac{1}{{\textrm{A}_{\textrm{ring}}^{\textrm{PSPS}}}},$$
where
$$\frac{1}{{\textrm{A}_{\textrm{ring}}^{PSPS}}}=\frac{\mathcal{L}^{{-}1}\int\left(\chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})/\overline{\chi}_{3}\right)\!\left(\textsf{e}_{P}^{i}(\boldsymbol{r}_{{\perp}};\zeta) \textsf{e}_{S}^{j}(\boldsymbol{r}_{{\perp}};\zeta)\right)^{*}\textsf{e}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{e}_{S}^{l}(\boldsymbol{r}_{{\perp}};\zeta) \textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta}{\left(\int\frac{n(\boldsymbol{r}_{{\perp}};\omega_{P})/\overline{n}_{P}}{v_{g}(\boldsymbol{r}_{{\perp}};\omega_{P})/v_{P}} \textsf{e}_{P}^{*}(\boldsymbol{r}_{{\perp}};0)\cdot\textsf{e}_{P}(\boldsymbol{r}_{{\perp}};0)\textrm{d}\boldsymbol{r}_{{\perp}}\right)\!\left(\int\frac{n(\boldsymbol{r}_{{\perp}}; \omega_{S})/\overline{n}_{S}}{v_{g}(\boldsymbol{r}_{{\perp}};\omega_{S})/v_{S}}\textsf{e}_{S}^{*}(\boldsymbol{r}_{{\perp}};0)\cdot\textsf{e}_{S} (\boldsymbol{r}_{{\perp}};0)\textrm{d}\boldsymbol{r}_{{\perp}}\right)}.$$
As in the channel calculation, here we have ${\gamma_{\textrm{ring}}^{\textrm{SPM}}} \approx {\gamma_{\textrm{ring}}^{\textrm{XPM}}}$ as long as $\omega _{J}$ and $\omega _{K}$ are close.

We close this section by deriving the promised correct expression for the SPM term arising in Sec. 2.1. In the special case considered there, we can identify $\varepsilon _{3}(\boldsymbol{r})$ with $\chi _{3}^{zzzz}(\boldsymbol{r}),$ and from (163) we find

$${\gamma_{\textrm{ring}}^{\textrm{SPM}}}=\frac{3\omega_{J}}{4v_{J}^{2}\mathcal{L}}K,$$
where $K$ is given by (11). Using this in (162) we have
$$H_{\text{ring}}^{\text{SPM}}={-}\frac{3}{2}\left(\frac{\hbar\omega_{J}}{2\mathcal{L}}\right)^{2}Kc_{J}^{{\dagger}}c_{J}^{{\dagger}}c_{J}c_{J},$$
which leads to the expression (13).

3. Obtaining the Ket from the Equations of Motion

In this section we develop mathematical tools to describe the quantum states that are typically generated in quantum nonlinear optical processes. The Hamiltonians describing these processes are polynomials with degree strictly greater than two in the electromagnetic fields, such as those developed in the previous section. However, it is often the case that some of the fields can be described classically, that is, replaced by their classical expectation values, leading to Hamiltonians that, while still nonlinear in the fields, are only quadratic in quantum operators. The validity and usefulness of this approximation is made apparent in subsequent sections, and thus for the moment we focus on the mathematical tools necessary to obtain the quantum state generated by these interactions. We stress that although there is a considerable reduction in complexity working with Hamiltonians at most quadratic in the operators, there are still significant complications stemming from the fact that these Hamiltonians are still time-dependent and have nonvanishing commutators at different times.

Before heading on directly to the most general problem one can tackle with the tools from Gaussian quantum optics and the symplectic formalism [5,6467] we consider simple examples that help build intuition toward the most general results in Sec. 3.1. We assume the reader is familiar with basic quantum mechanics, including the Schrödinger and Heisenberg pictures [68,69]. In Sec. 3.2 we introduce the Dyson, Magnus, and Trotter–Suzuki expansions as methods for the solution of linear differential equations, including the Schrödinger and the Heisenberg equation. In Sec. 3.3 we use the first-order Magnus expansion to solve the low-gain multimode squeezing problem and to introduce the concept of Schmidt modes. In Sec. 3.4 we use the fact that the solutions to the Heisenberg equations must preserve equal-time commutation relations to derive important properties of the Heisenberg propagator. In Sec. 3.5 we introduce the characteristic function and the Heisenberg propagator to derive the form of the quantum state generated by a general quadratic Hamiltonian in the spontaneous regime when the initial state is vacuum. We specialize some of the results derived to the case of nondegenerate squeezing in Sec. 3.6. Then we generalize the results related to the spontaneous problem, when the input state is vacuum, to arbitrary input states in Sec. 3.7. In the last two subsections we study the effect of loss in Sec. 3.8 and finally show how to calculate the statistics of homodyne and photon-number resolved measurements in Sec. 3.9.

3.1 Squeezing Hamiltonians

To gain some insight into the Gaussian formalism we study a very simple and well-known problem in quantum optics. We solve, both in the Schrödinger and Heisenberg pictures, the dynamics generated by the Hamiltonian

$$H = i \frac{\hbar g}{2}(e^{i \phi} a^{{\dagger} 2} - e^{{-}i \phi} a^2 ),$$
where $a,a^\dagger$ are the annihilation and creation operators of a harmonic oscillator that satisfy the usual commutation relations $[a,a^\dagger ] = 1$. The evolution of any quantum mechanical system is dictated by its Hamiltonian $H(t)$ via the Schrödinger equation,
$$i \hbar \frac{\textrm{d}}{\textrm{d}t} \mathcal{U}(t,t_{\text{i}}) = H(t) \mathcal{U}(t,t_{\text{i}}),$$
where $\mathcal {U}(t,t_{\text {i}})$ is the time evolution operator that satisfies the boundary condition $\mathcal {U}(t_{\text {i}},t_{\text {i}}) = \mathbb {I}$ and $\mathbb {I}$ is the Hilbert space identity.

For the very simple time-independent Hamiltonian of (174), we can immediately write

$$\mathcal{U}(t_{\text{f}},t_{\text{i}}) = \exp\left( - i \frac{H}{\hbar} (t_{\text{f}}-t_{\text{i}}) \right) = \exp\left(\tfrac{r}{2}\left[e^{i \phi} a^{{\dagger} 2} - e^{{-}i \phi} a^2 \right] \right), r = g (t_{\text{f}}-t_{\text{i}}).$$
The unitary operator above corresponds to the so-called squeezing operator that when acted on vacuum yields a single-mode squeezed vacuum state, which can be written as [43]
$$|{\Psi(t_{\text{f}})} = \exp\left(\tfrac{r}{2}\left[ e^{i \phi} a^{{\dagger} 2} - e^{{-}i \phi} a^2 \right] \right)|0 \rangle = \frac{1}{\sqrt{\cosh r}}\sum_{n=0}^\infty e^{i n\phi} \tanh^n r \frac{\sqrt{2n!}}{2^n n!} |2n \rangle,$$
where in the last equation we introduced the Fock states
$$|{n} = \frac{a^{{\dagger} n} }{ \sqrt{n!}} |{0},$$
and used the well-known disentangling formula (cf. Appendix 5 of Barnett and Radmore [70])
$$\begin{aligned}\exp\left(\tfrac{r}{2}\left[ e^{i \phi} a^{{\dagger} 2} - e^{{-}i \phi} a^2 \right] \right) & = \exp\left( \tfrac{1}{2} e^{i \phi } \tanh r \ a^{{\dagger} 2}\right) \\ & \quad\times \exp\left( - \left[ a^\dagger a+\tfrac{1}{2} \right] \ln \cosh r \right)\\ & \quad\times \exp\left( -\tfrac{1}{2} e^{{-}i \phi } \tanh r \ a^{ 2}\right). \end{aligned}$$
One can easily verify the following expectation values:
$$\begin{aligned} \langle{a}_{t_{\text{f}}}\rangle & = \langle{\Psi(t_{\text{f}})\rangle |a|\Psi(t_{\text{f}})} = 0, \\ n_{t_{\text{f}}} = \langle{a^\dagger a}\rangle_{t_{\text{f}}} & = \langle{\Psi(t_{\text{f}})|a^\dagger a|\Psi(t_{\text{f}})}\rangle = \sinh^2 r,\\ m_{t_{\text{f}}} =\langle{a^2}\rangle_{t_{\text{f}}} & = \langle{\Psi(t_{\text{f}})|a^2|\Psi(t_{\text{f}})}\rangle = \tfrac{1}{2}e^{i \phi} \sinh 2r,\\ \langle{a^\dagger a a^\dagger a}\rangle_{t_\text{f}} & = 3 n_{t_\text{f}}^2+ 2n_{t_\text{f}}. \end{aligned}$$
Introducing the Hermitian quadrature operators
$$q = \sqrt{\frac{\hbar}{2}} \left(a + a^\dagger \right), \quad p ={-}i \sqrt{\frac{\hbar}{2}} \left(a - a^\dagger \right)$$
one can also verify the moments
$$\begin{aligned}\langle{q}\rangle_{t_{\text{f}}} & = \langle{p}\rangle_{t_{\text{f}}} = 0, \\ \langle{q^2}\rangle_{t_{\text{f}}} & = \tfrac{\hbar}{2}\left(2n_{t_{\text{f}}}+1+ m_{t_{\text{f}}}+m_{t_{\text{f}}}^* \right) = \frac{\hbar}{2} \left(\sinh (2 r) \cos (\phi )+2 \sinh ^2(r)+1 \right),\\ \langle{p^2}\rangle_{t_{\text{f}}} & = \tfrac{\hbar}{2}\left(2n_{t_{\text{f}}}+1- m_{t_{\text{f}}}-m_{t_{\text{f}}}^* \right)=\frac{\hbar}{2} \left(-\sinh (2 r) \cos (\phi )+2 \sinh ^2(r)+1 \right),\\ \frac{\langle{qp+pq}\rangle_{t_{\text{f}}}}{2} & = i \tfrac{\hbar}{2} (m^* - m) =\frac{\hbar}{2} \sin(\phi) \sinh(2r). \end{aligned}$$
These moments can be arranged into a covariance matrix
$${\begin{aligned}V_{t_{\text{f}}} =\begin{pmatrix} \langle{q}\rangle_{t_{\text{f}}}^2 & \frac{\langle{qp+pq}\rangle_{t_{\text{f}}}}{2} \\ \frac{\langle{qp+pq}\rangle_{t_{\text{f}}}}{2} & \langle{p^2}\rangle_{t_{\text{f}}} \end{pmatrix} = \frac{\hbar}{2} R(\phi/2) \begin{pmatrix} e^{2r} & 0 \\ 0 & e^{{-}2r} \end{pmatrix} R(\phi/2)^T, \quad R(\phi) = \begin{pmatrix} \cos \phi & - \sin \phi \\ \sin \phi & \cos \phi \end{pmatrix}. \end{aligned}}$$
We see that the quadrature along the angle $\phi /2$ has increased fluctuations by amount $e^{2r}$ relative the vacuum level, whereas the quadrature in the direction perpendicular to $\phi /2$ has decreased (squeezed) fluctuations by amount $e^{-2r}$. This is illustrated in Fig. 7.

 figure: Figure 7.

Figure 7. Wigner functions of the vacuum state and a squeezed vacuum state.

Download Full Size | PDF

One can also solve the dynamics of this problem in the Heisenberg picture. The operators are now dynamical quantities defined by $O(t) = \mathcal {U}^\dagger (t,t_{\text {i}})O(t_{\text {i}}) \mathcal {U}(t,t_{\text {i}})$ and their dynamics is determined by the Heisenberg equation of motion,

$$\frac{\textrm{d}}{\textrm{d}t} O(t) = \frac{i}{\hbar} [H(t),O(t)],$$
with the boundary condition $O(t_{\text {i}}) \equiv O$ where $O$ is simply the Schrödinger picture operator. For the Hamiltonian of (174) and for $O \in \{a,a^\dagger \}$ we find
$$\frac{\textrm{d}}{\textrm{d}t} a = \frac{i}{\hbar} [H,O] = g e^{i \phi} a^\dagger, \frac{\textrm{d}}{\textrm{d}t} a^\dagger{=} \frac{i}{\hbar} [H,O] = g e^{{-}i \phi} a,$$
or, more compactly,
$$\begin{aligned}\frac{\textrm{d}}{\textrm{d}t} \begin{pmatrix} a \\ a^\dagger \end{pmatrix} & ={-}i \underbrace{\begin{pmatrix} 0 & i g e^{i \phi}\\ i g e^{{-}i \phi} & 0 \end{pmatrix} }_{{\equiv} \boldsymbol{A}}\begin{pmatrix} a \\ a^\dagger \end{pmatrix}. \end{aligned}$$
Note that the matrix $\boldsymbol{A}$ defined in the last equation is not Hermitian. It is instead an element of the Lie algebra $\mathfrak {su}(1,1)$ that generates a Heisenberg propagator which is an element of the Lie group $SU(1,1)$ (cf. Appendix 11.1.4 of Klimov and Chumakov [71]).

We can immediately find the solution of the last differential equation by exponentiation,

$${\begin{aligned}\begin{pmatrix} a(t_{\text{f}}) \\ a^\dagger(t_{\text{f}}) \end{pmatrix} = \exp({-}i \boldsymbol{A}(t_{\text{f}}-t_{\text{i}})) \begin{pmatrix} a(t_{\text{i}}) \\ a^\dagger(t_{\text{i}}) \end{pmatrix} = \left( \begin{array}{cc} \cosh g (t_{\text{f}} - t_{\text{i}}) & e^{i \phi } \sinh g (t_{\text{f}} - t_{\text{i}}) \\ e^{{-}i \phi } \sinh g(t_{\text{f}} - t_{\text{i}}) & \cosh g (t_{\text{f}} - t_{\text{i}}) \\ \end{array} \right) \begin{pmatrix} a(t_{\text{i}}) \\ a^\dagger(t_{\text{i}}) \end{pmatrix}, \end{aligned}}$$
from which we can recover the moments derived in the Schrödinger picture by the usual rule $\langle {O}_t\rangle = \langle {\Psi (t_{\text {i}})|O(t)|\Psi (t_{\text {i}})}\rangle = \langle {\Psi (t)|O(t_{\text {i}})|\Psi (t)}\rangle$.

At this point it might seem like overkill to have derived the same expectation value using two different methods. However, as we show in a moment, it is useful to have two ways of looking at the problem because there are situations where solving the Schrödinger equation for the ket $| {\Psi (t)}$ is hopeless whereas solving the Heisenberg equations is practical. Moreover, once the Heisenberg equations are solved one can then write the sought after time-evolved Schrödinger picture ket.

3.2 Multimode Squeezing: Dyson, Magnus, and Trotter–Suzuki

Let us consider a slightly more complicated Hamiltonian given by

$$H(t) = \hbar \left\{ \sum_{i,j =1}^\ell \Delta_{ij}(t) a_i^\dagger a_j + i \frac12 \sum_{k,l=1}^\ell \left[ \zeta_{kl}(t) a_k^{{\dagger} } a_{l}^{{\dagger}} -\text{H.c.} \right] \right\},$$
which is the most general bosonic quadratic Hamiltonian on $\ell$ modes. In the last equation we have introduced $\ell$ bosonic modes satisfying the usual canonical commutation relations
$$[a_i,a_j] = [a_i^\dagger,a_j^\dagger] = 0, \quad [a_i,a_j^\dagger] = \delta_{ij},$$
we demand that $\Delta _{ij} = \Delta _{ji}^*$ for the Hamiltonian to be Hermitian, and assume without loss of generality that $\zeta _{kl} = \zeta _{lk}$, because any anti-symmetric contribution to $\zeta _{kl}$ will vanish upon contraction with the permutation symmetric term $a_k^\dagger a_l^\dagger$.

In this section we do not attach any particular meaning to the mode labels $i$ and $j$ in the last set of equations. In subsequent sections these labels will correspond to, for example, discretized wave vectors near a reference or central wave vector used to obtain a simplified dispersion relation. In this case, the term $\Delta _{ij}$ will correspond to a detuning between the different wave vectors (and possibly XPM of a classical pump on the quantum modes) whereas $\zeta _{kl}$ will correspond to wave-mixing induced by a nonlinear process.

As before, we would like to find the ket obtained by evolving the vacuum under this Hamiltonian, the so-called “spontaneous problem.” Note that the Hamiltonian describing the dynamics is now time-dependent and has three different terms that do not commute with each other, making a more complex dynamics than that of the single-mode problem analyzed in Sec. 3.1. In what follows we use boldface to refer to matrices $\boldsymbol{M}$ while referring to their entries as $M_{ij}$. In matrix notation, the constraints described in the previous paragraph are simply $\boldsymbol{\Delta } = \boldsymbol{\Delta }^\dagger$ and $\boldsymbol{\zeta } = \boldsymbol{\zeta }^T$ where $\dagger$ and $T$ indicate the conjugate transpose and the transpose, respectively.

We can formally solve the dynamics by writing the time-evolution operator associated with this Hamiltonian as

$$\mathcal{U}(t,t_{\text{i}}) = \mathcal{T} \exp\left(-\frac{i}{\hbar} \int_{t_{\text{i}}}^t \textrm{d}t' H(t') \right),$$
where $\mathcal {T}$ is the so-called time-ordering operator. When applied to a product of Hamiltonians at different times $t_1,\ldots, t_n$ it simply orders them chronologically
$$\mathcal{T}\left[H(t_1) H(t_2) \ldots H(t_n) \right] = H(t_{i_1}) H(t_{i_2}) \ldots H(t_{i_n}),$$
where $i_1,i_2\ldots,i_n$ is a permutation of the set $[1,2,\ldots,n]$ such that $t_{i_1} \geq t_{i_2}\geq \cdots \geq t_{i_n}$ [68,69].

In practice, the expression (190) can be interpreted as a power series of the Dyson [72] or Magnus [73,74] type or can be approximated using Trotterization [75,76]. In the Dyson series the unitary evolution operator is written as an infinite power series

$$\mathcal{U}(t_{\text{f}},t_{\text{i}}) = \mathbb{I}+({-}i)\int_{t_{\text{i}}}^{t_{\text{f}}} \textrm{d}t' \frac{ H(t')}{\hbar}+({-}i)^2\int_{t_{\text{i}}}^{t_{\text{f}}} \textrm{d}t' \int_{t_{\text{i}}}^{t'} \textrm{d}t^{\prime\prime} \frac{H(t') H(t^{\prime\prime})}{\hbar^2}+\cdots ,$$
whereas in the Magnus series one writes the solution as the exponential of a series of nested commutators at different times
$$\begin{aligned}\mathcal{U}(t_{\text{f}},t_{\text{i}}) & =\exp\big( \Omega_1(t_{\text{f}},t_{\text{i}})+ \Omega_2(t_{\text{f}},t_{\text{i}})+ \Omega_3(t_{\text{f}},t_{\text{i}})+\cdots \big), \end{aligned}$$
$$\begin{aligned}\Omega_1(t_{\text{f}},t_{\text{i}}) & ={-}\frac{i}{\hbar} \int_{t_{\text{i}}}^{t_{\text{f}}} \textrm{d}t H(t), \end{aligned}$$
$$\begin{aligned} \Omega_2(t_{\text{f}},t_{\text{i}}) & =\frac{({-}i)^2}{2 \hbar^2}\int_{t_{\text{i}}}^{t_{\text{f}}} \textrm{d}t \int_{t_{\text{i}}} ^{t} \textrm{d}t' \left[ H(t), H(t') \right], \end{aligned}$$
$$\begin{aligned} \Omega_3(t_{\text{f}},t_{\text{i}}) & =\frac{({-}i)^3}{6 \hbar^3}\int_{t_{\text{i}}}^{t_{\text{f}}} \textrm{d}t \int_{t_{\text{i}}} ^{t} \textrm{d}t' \int_{t_{\text{i}}} ^{t'} \textrm{d}t^{\prime\prime} \Big( \left[ H(t),\left[ H(t'), H(t^{\prime\prime}) \right] \right]\\ & \quad +\left[\left[ H(t), H(t') \right], H(t^{\prime\prime}) \right] \Big). \end{aligned}$$

For the quadratic Hamiltonians we investigate here, the unitary evolution operator resulting from the Magnus expansion at any order respects the photon statistics that the exact operator is known to have, whereas that resulting from the Dyson expansion does not. Furthermore, in the Magnus expansion all terms beyond $\Omega _1$ are clearly time-ordering corrections, related to the fact that $\left [ H(t), H(t') \right ] \neq 0$ whereas in the Dyson series each successive term contains a part associated with time-ordering corrections as well as a part associated with the solution that would be obtained by ignoring time-ordering corrections and naïvely writing $\mathcal {U}(t,t_{\text {i}}) = \exp \left (-\frac {i}{\hbar } \int _{t_{\text {i}}}^t \textrm{d}t' H(t') \right )$.

Finally, a third approach is to use the Trotter–Suzuki expansion to approximate the time evolution operator as a product of time evolution operators $\mathcal {U}(t_i,t_i+\Delta t)$ where each Hamiltonian is assumed to be roughly constant over small intervals of duration $\Delta t$, thus writing

$$\mathcal{U}(t_{\text{f}},t_{\text{i}}) = \exp\left({-}i \frac{H(t_{N})}{\hbar} \Delta t \right) \exp\left({-}i \frac{H(t_{N-1})}{\hbar} \Delta t \right) \ldots \exp\left({-}i \frac{H(t_{\text{i}})}{\hbar} \Delta t \right),$$
where $t_j = t_{\text {i}} + j \Delta t |_{j=0}^N$ and $\Delta t = (t_{\text {f}}-t_{\text {i}})/N$. Regardless of which strategy is chosen, one will have to deal with infinite-dimensional creation and destruction operators acting on a Hilbert space. Even if these operators are truncated at a finite Fock cutoff $c$, the dimensionality of the operators appearing in any of the equations above will be $c^{2\ell }$, which will easily fit into computer memory only for very modest $c$ and $\ell$.

3.3 Low-Gain Solutions

The three strategies in the previous section have different strengths and weaknesses. If the Hamiltonian satisfies $[H(t),H(t')]=0$ for all the times $t,t'$ between $t_{\text {i}}$ and $t_{\text {f}}$ then the Magnus expansion at the first level immediately gives the exact solution. If the Hamiltonian $H(t)$ is “weak” we can treat the Magnus or Dyson series perturbatively and keep only the first few terms of either of them.

The simplest approximation is to keep only the first term of either of them. In this perturbative or low-gain limit one can obtain a simple solution to the spontaneous problem

$$\begin{aligned} |{\overline{\Psi}(t_\text{f})}\rangle & \approx \overline{\mathcal{U}}(t_\text{f}, t_\text{i})\, |{\text{vac}}\rangle, \end{aligned}$$
$$\begin{aligned}\overline{\mathcal{U}}(t_\text{f}, t_\text{i}) & = \exp\left(-\frac{i}{\hbar} \int_{t_{\text{i}}}^{t_\text{f}} \textrm{d}t' H(t') \right) = \exp\left[ \tfrac{1}{2}\sum_{k,l} \overline{\mathit{J}}_{kl} a_l^\dagger a_k^\dagger{-} \text{H.c.} \right], \end{aligned}$$
where, for simplicity, we have assumed that $\Delta _{ij}(t) = 0$ [77] and introduced $\overline {\boldsymbol{J}} = \int _{t_{\text {i}}}^{t_{\text {f}}} \textrm{d}t \ \boldsymbol{\zeta }(t)$, the so-called joint amplitude of the squeezed state above. As detailed in the following, it is often useful to rewrite this state by making use of the Takagi–Autonne decomposition of the symmetric matrix $\overline {\boldsymbol{J}} = \overline {\boldsymbol{J}}^T = \overline {\boldsymbol{F}} [\oplus _{\lambda =1}^\ell \overline {r}_\lambda ] \overline {\boldsymbol{F}}^T$. Here $\overline {\boldsymbol{F}}$ is unitary $\overline {\boldsymbol{F}} \overline {\boldsymbol{F}}^\dagger = \mathbb {I}_\ell$, the low-gain assumption is reflected in the fact that $\overline {\mathit{r}}_\lambda \ll 1$, we have used the direct-sum notation $\oplus _{\lambda =1}^\ell \overline {\mathit{r}}_\lambda$ to indicate a diagonal matrix square with entries $\{\overline {\mathit{r}}_\lambda \}$, and have used overbars to indicate that quantities are associated with a low-gain solution. Note that the symmetry of $\boldsymbol{J}$ simply follows from the symmetry of $\boldsymbol{\zeta }$. The Takagi–Autonne decomposition (cf. Corollary 4.4.4 of Horn and Johnson [78]) is a singular value decomposition (SVD) where it is made explicit that the matrix being decomposed is symmetric. Numerical routines to perform this decomposition can be found in the Python packages Strawberry Fields [79] and The Walrus [80].

Using the Takagi–Autonne decomposition lets us introduce the Schmidt (or broadband, or supermode) operators

$$\overline{A}_\lambda^\dagger{=} \sum_{k=1}^\ell a_k^\dagger \overline{F}_{k\lambda} \Longleftrightarrow a_k^\dagger{=} \sum_{i=1}^\ell \overline{F}^*_{k\lambda} \overline{A}_{\lambda}^\dagger,$$
which, due to the unitarity of $\overline {\boldsymbol{F}}$, satisfy bosonic canonical commutation relations
$$[\overline{A}_{\lambda}, \overline{A}_{\lambda'}] = 0, \quad [\overline{A}_{\lambda}, \overline{A}_{\lambda'}^\dagger] = \delta_{\lambda,\lambda'}$$
that allows us to finally write [81,82]
$$\begin{aligned}|{\overline{\Psi}(t_\text{f})}\rangle & = \exp\left[\tfrac12 \sum_{k,l=1}^\ell \overline{J}_{kl} a^\dagger_{k} a^\dagger_l - \text{H.c.}\right] |{\text{vac}}\rangle = \exp \left[ \tfrac12 \sum_{\lambda=1}^\ell \overline{\mathit{r}}_\lambda \overline{A}_\lambda^{{\dagger} 2} -\text{H.c.} \right] |{\text{vac}}\rangle \\ & = \bigotimes_{\lambda=1}^\ell \exp \left[ \tfrac12 \overline{\mathit{r}}_\lambda \overline{A}_\lambda^{{\dagger} 2} -\text{H.c.} \right] |{\text{vac}}\rangle. \end{aligned}$$
The diagonal form of the exponential argument allows us to identify the state generated in the spontaneous problem as a manifold of squeezed states over the modes defined by $\overline {\boldsymbol{F}}$, i.e., the Schmidt modes of the system for which we can easily write their photon number expansion as
$$|{\overline{\Psi}(t_\text{f})}\rangle = \bigotimes_{\lambda=1}^\ell \left[ \frac{1}{\sqrt{\cosh \overline{\mathit{r}}_\lambda}}\sum_{n=0}^\infty \tanh^n \overline{\mathit{r}}_\lambda \frac{\sqrt{(2n)!}}{2^n n!} |2n \rangle_\lambda \right],$$
where $| {n}\rangle_\lambda \equiv \frac {\overline {A}_\lambda ^{\dagger n}}{\sqrt {n!}} | {0}\rangle$ is a Fock state in the Schmidt mode labeled by $\lambda$.

We can also use the form of the low-gain time-evolution operator to transform the operators. Using the definitions of the Schmidt modes one can easily find

$$\begin{aligned}a_i(t_{\text{f}}) & = \overline{\mathcal{U}}(t_\text{f}, t_\text{i})^\dagger a_i( t_\text{i}) \overline{\mathcal{U}}(t_\text{f}, t_\text{i}) = \sum_{j=1}^\ell \overline{V}_{ij} a_{j}(t_{\text{i}}) + \sum_{j=1}^\ell \overline{W}_{ij} a^\dagger _{j} (t_{\text{i}}), \end{aligned}$$
$$\begin{aligned}a_i^\dagger(t_{\text{f}}) & = \overline{\mathcal{U}}(t_\text{f}, t_\text{i})^\dagger a_i^\dagger( t_\text{i}) \overline{\mathcal{U}}(t_\text{f}, t_\text{i}) = \sum_{j=1}^\ell \overline{V}^*_{ij} a^\dagger_{j}(t_{\text{i}}) + \sum_{j=1}^\ell \overline{W}^*_{ij} a _{j}(t_{\text{i}}), \end{aligned}$$
where now
$$\overline{\boldsymbol{V}} = \overline{\boldsymbol{F}} \left[ \oplus_{\lambda=1}^\ell \cosh \overline{\mathit{r}}_{\lambda} \right] \overline{\boldsymbol{F}}^\dagger, \quad \overline{\boldsymbol{W}} = \overline{\boldsymbol{F}} \left[ \oplus_{\lambda=1}^\ell \sinh \overline{\mathit{r}}_{\lambda} \right] \overline{\boldsymbol{F}}^T.$$
To lowest nonvanishing order in the squeezing parameters $\overline {\mathit{r}}_\lambda \ll 1$ one has $\overline {\boldsymbol{V}} \approx \mathbb {I}_\ell$ and $\overline {\boldsymbol{W}} \approx \overline {\boldsymbol{J}}$.

3.4 Properties of the Solution to the Heisenberg Equations of Motion: Bogoliubov Transformations

Let us now analyze this same problem in the Heisenberg picture for arbitrary gain. We now find

$$\begin{aligned} \frac{\textrm{d}}{\textrm{d}t} \begin{pmatrix} \boldsymbol{a} \\ \boldsymbol{a}^\dagger \end{pmatrix} & ={-}i \underbrace{\begin{pmatrix} - \boldsymbol{\Delta}(t) & i \boldsymbol{\zeta}(t) \\ i \boldsymbol{\zeta}^*(t) & \boldsymbol{\Delta}(t) \end{pmatrix} }_{{\equiv} \boldsymbol{A}(t)}\begin{pmatrix} \boldsymbol{a} \\ \boldsymbol{a}^\dagger \end{pmatrix}, \end{aligned}$$
where we have introduced
$$\begin{aligned}\boldsymbol{a} = \begin{pmatrix} a_1 \\ \vdots \\ a_\ell \end{pmatrix}, \quad \boldsymbol{a}^\dagger{=} \begin{pmatrix} a^\dagger_1 \\ \vdots \\ a^\dagger_\ell \end{pmatrix}. \end{aligned}$$
Just like the Schrödinger equation for the time evolution operator of (175), for quadratic bosonic Hamiltonians, the Heisenberg equations of motion are linear and dictated by the matrix $\boldsymbol{A}$. We can formally solve these equations using a time-ordered exponential as previously,
$${\begin{aligned} \begin{pmatrix} \boldsymbol{a}(t_{\text{f}}) \\ \boldsymbol{a}^\dagger(t_{\text{f}}) \end{pmatrix} = \mathcal{U}^\dagger(t_{\text{f}},t_{\text{i}}) \begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix} \mathcal{U}(t_{\text{f}},t_{\text{i}})= \mathcal{T} \exp\left( - i \int_{t_{\text{i}}}^{_{\text{f}}} \textrm{d}t'\boldsymbol{A}(t') \right) \begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix} =\boldsymbol{K}(t_{\text{f}},t_{\text{i}})\begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix}. \end{aligned}}$$
Yet while we still need to consider a time-ordered exponential here, this problem is actually much simpler than the one associated with (190). In particular, here $\boldsymbol{A}(t)$ is simply a $2 \ell \times 2 \ell$ matrix. It easily fits in the memory of a computer, and using Trotter–Suzuki we can approximate
$$\boldsymbol{K}(t_{\text{f}},t_{\text{i}}) \approx \exp\left({-}i \Delta t \boldsymbol{A}(t_{N}) \right) \exp\left({-}i \Delta t \boldsymbol{A}(t_{N-1}) \right) \ldots \exp\left({-}i \Delta t \boldsymbol{A}(t_2) \right) \exp\left({-}i \Delta t \boldsymbol{A}(t_{\text{i}}) \right),$$
where $t_j = t_{\text {i}} + j \Delta t |_{j=0}^N$ and $\Delta t = (t_{\text {f}}-t_{\text {i}})/N$ and we have discretized the time evolution into $N+1$ slices. The solution of the problem now boils down to the multiplication of matrices with size proportional to the number of modes, and there is no need to even introduce a cutoff in Fock space. Note that an alternative approach to approximate the Heisenberg propagator is to employ the Magnus expansion to a finite level as done by Lipfert et al. [83]. Moreover, this approach can also be applied to evolution equations in space instead of time, as done in Ref. [84].

From the fact that the transformations connecting initial- and final-time Heisenberg operators are linear and the fact that these solutions must respect the canonical commutation relation at each time, one can obtain a number of useful properties. To simplify notation, from now on we drop initial and final times from matrices, writing, for example, $\boldsymbol{K}$ in place of $\boldsymbol{K}(t_{\text {f}},t_{\text {i}})$. Let us start by writing the Heisenberg propagator in block form

$$\begin{aligned}\boldsymbol{K} = \begin{pmatrix} \boldsymbol{V} & \boldsymbol{W} \\ \boldsymbol{W}^* & \boldsymbol{V}^* \end{pmatrix}, \end{aligned}$$
along with
$$\begin{aligned} a_i(t_{\text{f}}) & = \mathcal{U}^\dagger a_i \mathcal{U} = \sum_{j=1}^\ell V_{ij} a_{j}(t_{\text{i}}) + \sum_{j=1}^\ell W_{ij} a^\dagger _{j} (t_{\text{i}}), \end{aligned}$$
$$\begin{aligned} a_i^\dagger(t_{\text{f}}) & = \mathcal{U}^\dagger a_i^\dagger \mathcal{U} = \sum_{j=1}^\ell V^*_{ij} a^\dagger_{j}(t_{\text{i}}) + \sum_{j=1}^\ell W^*_{ij} a _{j}(t_{\text{i}}). \end{aligned}$$
As $\mathcal {U}$ is a unitary operator in Hilbert space, the bosonic operators at the final time must satisfy the same equations as their initial time versions, i.e., (189) must hold if we replace $a_i \to a_i(t)$. From this alone we infer that
$$ [a_i(t_{\text{f}}), a_j(t_{\text{f}})] = 0 \Leftrightarrow \boldsymbol{W} \boldsymbol{V}^T = \boldsymbol{V} \boldsymbol{W}^T, $$
$$ [a_i(t_{\text{f}}), a^\dagger_j(t_{\text{f}})] = \delta_{ij} \Leftrightarrow \boldsymbol{V}^* \boldsymbol{V}^T - \boldsymbol{W}^* \boldsymbol{W}^T = \mathbb{I}_M. $$
The constraints above imply that one can write the following joint SVDs [6,85,86]
$$\boldsymbol{V} = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \cosh r_{\lambda} \right] \boldsymbol{G}, \quad \boldsymbol{W} = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \sinh r_{\lambda} \right] \boldsymbol{G}^*,$$
where $\boldsymbol{F}$ and $\boldsymbol{G}$ are unitary matrices. As we show later, the matrix $\boldsymbol{F}$ defines the (output) Schmidt modes of the problem, the matrix $\boldsymbol{G}$ defines the input Schmidt modes of the problem, and the quantities $r_\lambda$ are the squeezing parameters. Linear transformation equations as in Eq. (212) that preserve the canonical commutation relation are known as Bogoliubov transformations.

It is useful to note that the entries of the matrix

$$\boldsymbol{M} = \boldsymbol{M}^T = \boldsymbol{W} \boldsymbol{V}^T$$
coincide with the following vacuum expectation value
$$M_{ij} = \langle{a_i(t_{\text{f}})a_j(t_{\text{f}})}\rangle_{|{\text{vac}}\rangle} = \rangle{\text{vac}}|\mathcal{U}^\dagger a_i a_j \mathcal{U} |{\text{vac}}\rangle = \left( \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \tfrac{1}{2}\sinh 2r_\lambda \right] \boldsymbol{F}^T \right)_{ij}.$$
Similarly we find that the entries of the Hermitian matrix
$$\boldsymbol{N} = \boldsymbol{N}^\dagger{=} \boldsymbol{W}^* \boldsymbol{W}^T$$
coincide with
$$N_{ij} = \langle{a_i^\dagger(t_{\text{f}})a_j(t_{\text{f}})}\rangle_{|{\text{vac}}\rangle} = \rangle{\text{vac}}|\mathcal{U}^\dagger a_i^\dagger a_j \mathcal{U} |{\text{vac}}\rangle = \left( \boldsymbol{F}^* \left[ \oplus_{\lambda=1}^\ell \sinh^2r_\lambda \right] \boldsymbol{F}^T \right)_{ij}.$$
Finally, based on the decompositions of (216) we can easily write
$$\begin{aligned} \boldsymbol{K} = \begin{pmatrix} \boldsymbol{F} & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{F}^* \end{pmatrix} \begin{pmatrix} \oplus_{\lambda=1}^\ell \cosh r_{\lambda} & \oplus_{\lambda=1}^\ell \sinh r_{\lambda} \\ \oplus_{\lambda=1}^\ell \sinh r_{\lambda} & \oplus_{\lambda=1}^\ell \cosh r_{\lambda} \end{pmatrix} \begin{pmatrix} \boldsymbol{G} & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{G}^* \end{pmatrix}, \end{aligned}$$
and the inverse
$$\begin{aligned} \boldsymbol{K}^{{-}1} = \begin{pmatrix} \boldsymbol{G}^\dagger & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{G}^T \end{pmatrix} \begin{pmatrix} \oplus_{\lambda=1}^\ell \cosh r_{\lambda} & -\oplus_{\lambda=1}^\ell \sinh r_{\lambda} \\ -\oplus_{\lambda=1}^\ell \sinh r_{\lambda} & \oplus_{\lambda=1}^\ell \cosh r_{\lambda} \end{pmatrix} \begin{pmatrix} \boldsymbol{F}^\dagger & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{F}^T \end{pmatrix}= \begin{pmatrix} \tilde{\boldsymbol{V}} & \tilde{\boldsymbol{W}} \\ \tilde{\boldsymbol{W}}^* & \tilde{\boldsymbol{V}}^* \end{pmatrix}, \end{aligned}$$
where we have introduced
$$\tilde{\boldsymbol{V}} = \boldsymbol{G}^\dagger \left[ \oplus_{\lambda=1}^\ell \cosh r_\lambda \right] \boldsymbol{F}^\dagger{=} \boldsymbol{V}^\dagger, \quad \tilde{\boldsymbol{W}} ={-}\boldsymbol{G}^\dagger \left[ \oplus_{\lambda=1}^\ell \sinh r_\lambda \right] \boldsymbol{F}^T ={-} \boldsymbol{W}^T,$$
for the blocks of the matrix $\boldsymbol{K}^{-1}$. These matrices are quite useful when later we need to write the backward-evolved Heisenberg operators
$$\begin{aligned} \mathcal{U}(t_{\text{f}},t_{\text{i}}) \begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix} \mathcal{U}^\dagger(t_{\text{f}},t_{\text{i}}) = \boldsymbol{K}^{{-}1} \begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix} = \begin{pmatrix} \tilde{\boldsymbol{V}} & \tilde{\boldsymbol{W}} \\ \tilde{\boldsymbol{W}}^* & \tilde{\boldsymbol{V}}^* \end{pmatrix} \begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix}. \end{aligned}$$
Finally, we note that these transformations, that preserve the commutation relations, are called symplectic transformations. They form a group and have been studied extensively [5,6467]; we provide some details of their properties in Appendix D.

3.5 Obtaining the Ket from the Bogoliubov Transformations

In the previous section we showed how to solve the Heisenberg equations of motion, and, based on the fact that they correspond to unitary operations in Hilbert space, we derived some of their properties. In this section, we show how one can obtain the state that evolved under $\mathcal {U}$. Before doing this we introduce a useful object that uniquely describes a quantum state, its characteristic function [70]

$$\chi_{\rho}(\boldsymbol{\alpha}) = \text{tr}\left( \mathcal{D}(\boldsymbol{\alpha}) \rho \right),$$
where $\rho$ is the density matrix of quantum system. The characteristic function is also the starting point for the calculation of any quasi-probability distribution, such as the Wigner function, of a quantum state via Fourier transform over the real and imaginary parts of $\boldsymbol{\alpha }$ [70]. For a pure state $\rho = | {\Psi }\rangle \rangle {\Psi }|$ and we write its characteristic function as $\chi _{| {\Psi }\rangle}(\boldsymbol{\alpha })$. The displacement operator is defined as [87]
$$\begin{aligned} \mathcal{D}(\boldsymbol{\alpha}) & = \exp\left[ \boldsymbol{\alpha}^T \boldsymbol{a}^\dagger{-} \boldsymbol{\alpha}^\dagger \boldsymbol{a} \right] = \exp\left[ \boldsymbol{\alpha}^T \boldsymbol{a}^\dagger \right] \exp\left[ - \boldsymbol{\alpha}^\dagger \boldsymbol{a} \right] \exp[ -\tfrac12 \|\boldsymbol{\alpha} \|^2], \end{aligned}$$
$$\begin{aligned} \boldsymbol{\alpha} & = (\alpha_1,\ldots,\alpha_M)^T \in \mathbb{C}^\ell, \end{aligned}$$
and in the last line we used a well-known disentangling formula (cf. Appendix 5 of Barnett and Radmore [70]).

For the vacuum state the characteristic function is easily calculated, and found to be a Gaussian in the complex amplitudes $\boldsymbol{\alpha }$

$$\chi_{|{\text{vac}}\rangle}(\boldsymbol{\alpha}) = \exp[ -\tfrac12 \|\boldsymbol{\alpha} \|^2],$$
where we used the disentangling formula for the displacement operator and the cyclic property of the trace $\text {tr}(A B C) = \text {tr}(BCA)$ together with $\rangle {\text {vac}}| \exp \left [ \boldsymbol{\alpha } \boldsymbol{a}^\dagger \right ] = \rangle {\text {vac}}|$ and $\exp \left [ - \boldsymbol{\alpha }^* \boldsymbol{a} \right ] | {\text {vac}}\rangle = | {\text {vac}}\rangle$. The vacuum state is the simplest state that has a Gaussian characteristic function. More generally, any state that has a Gaussian characteristic function is called a Gaussian state, and we show now that the state $\mathcal {U} | {\text {vac}}\rangle$ is a member of this set. To this end consider
$$\begin{aligned} \chi_{\mathcal{U}|{\text{vac}}\rangle}(\boldsymbol{\alpha}) & = \text{tr}\left( \mathcal{D}(\boldsymbol{\alpha}) \mathcal{U}|{\text{vac}}\rangle \rangle{\text{vac}}| \mathcal{U}^\dagger \right) \\ & = \text{tr}\left(\mathcal{U}^\dagger \mathcal{D}(\boldsymbol{\alpha}) \mathcal{U}|{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \text{tr}\left(\mathcal{U}^\dagger \exp\left[ \boldsymbol{\alpha}^T \boldsymbol{a}^\dagger{-} \boldsymbol{\alpha}^\dagger \boldsymbol{a} \right]\mathcal{U}|{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \text{tr}\left( \exp\left[ \boldsymbol{\alpha}^T \mathcal{U}^\dagger \boldsymbol{a}^\dagger \mathcal{U} - \boldsymbol{\alpha}^\dagger \mathcal{U}^\dagger \boldsymbol{a} \mathcal{U}\right]|{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \text{tr}\left( \exp\left[ \boldsymbol{\alpha}^T \left\{ \boldsymbol{V}^*\boldsymbol{a}^\dagger{+} \boldsymbol{W}^* \boldsymbol{a} \right\} - \boldsymbol{\alpha}^\dagger \left\{ \boldsymbol{V}\boldsymbol{a} + \boldsymbol{W} \boldsymbol{a}^\dagger \right\}\right]|{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \text{tr}\left( \exp\left[ \left\{ \boldsymbol{\alpha}^T \boldsymbol{V}^* - \boldsymbol{\alpha}^\dagger \boldsymbol{W} \right\} \boldsymbol{a}^\dagger{-} \left\{ \boldsymbol{\alpha}^\dagger\boldsymbol{V} - \boldsymbol{\alpha}^T \boldsymbol{W}^* \right\} \boldsymbol{a} \right]|{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \text{tr}\left( \mathcal{D}( \boldsymbol{V}^\dagger \boldsymbol{\alpha} - \boldsymbol{W}^T \boldsymbol{\alpha}^* ) |{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \exp[ -\tfrac12 \| \boldsymbol{\xi} \|^2], \quad \boldsymbol{\xi} = \boldsymbol{V}^\dagger \boldsymbol{\alpha} - \boldsymbol{W}^T \boldsymbol{\alpha}^*. \end{aligned}$$
We can now write
$$\| \boldsymbol{\xi}\|^2 = \boldsymbol{\xi}^\dagger \boldsymbol{\xi} = \boldsymbol{\alpha}^\dagger \boldsymbol{V} \boldsymbol{V}^\dagger \boldsymbol{\alpha} - \boldsymbol{\alpha}^\dagger \boldsymbol{V} \boldsymbol{W}^T \boldsymbol{\alpha}^* - \boldsymbol{\alpha}^T \boldsymbol{W}^* \boldsymbol{V}^\dagger \boldsymbol{\alpha} + \boldsymbol{\alpha}^T \boldsymbol{W}^* \boldsymbol{W}^T \boldsymbol{\alpha}^*,$$
and using the SVDs (216) we find
$$ \boldsymbol{V} \boldsymbol{V}^\dagger = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \cosh^2 r_{\lambda} \right] \boldsymbol{F}^\dagger{=} \mathbb{I}_{M}+ \boldsymbol{N}^*, $$
$$ \boldsymbol{V} \boldsymbol{W}^T = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \tfrac12 \sinh 2r_{\lambda} \right] \boldsymbol{F}^T = \left( \boldsymbol{W}^* \boldsymbol{V}^\dagger \right)^\dagger{=} \boldsymbol{M}, $$
$$ \boldsymbol{W}^* \boldsymbol{W}^T = \boldsymbol{F}^* \left[ \oplus_{\lambda=1}^\ell \sinh^2 r_{\lambda} \right] \boldsymbol{F}^T = \boldsymbol{N}, $$
from which we conclude that $\mathcal {U}| {\text {vac}}\rangle$ only depends on the left singular vectors $\boldsymbol{F}$ and the singular values $r_{\lambda }$ or, equivalently, on the moments $\boldsymbol{M}$ and $\boldsymbol{N}$ of the state $\mathcal {U}| {\text {vac}}\rangle$ introduced in (218) and (220), respectively. Explicitly, we have
$$\chi_{\mathcal{U}|{\text{vac}}\rangle}(\boldsymbol{\alpha}) = \exp\left[- \tfrac12 \left\{ \|\boldsymbol{\alpha}\|^2 + \boldsymbol{\alpha}^\dagger \boldsymbol{N}^* \boldsymbol{\alpha} - \boldsymbol{\alpha}^\dagger \boldsymbol{M} \boldsymbol{\alpha}^* - \boldsymbol{\alpha}^T \boldsymbol{M}^* \boldsymbol{\alpha} + \boldsymbol{\alpha}^T \boldsymbol{N} \boldsymbol{\alpha}^* \right\}\right].$$
Based on this observation we propose that [88]
$$ |{\Psi(t_{\text{f}})}\rangle \equiv \mathcal{U}|{\text{vac}}\rangle = \exp\left[ \frac12 \sum_{k,l=1}^\ell J_{kl} a^\dagger_{k} a^\dagger_l - \text{H.c.}\right]|{\text{vac}}\rangle, $$
$$ \boldsymbol{J} = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell r_{\lambda} \right] \boldsymbol{F}^T = \boldsymbol{J}^T, $$
where $\boldsymbol{J}$ is the (arbitrary gain) joint amplitude of the squeezed state. To show that the equation above holds we need to show that the $\boldsymbol{M}$ and $\boldsymbol{N}$ moments when calculated with the right-hand side of the equation above coincide with the ones obtained for the state $\mathcal {U}| {\text {vac}}\rangle$. As these uniquely determine the characteristic function, this is straightforward to do by reusing the results from Sec. 3.3 but now using the joint amplitude defined in the last equation. In fact, noting that the solution above has exactly the same form as the low-gain solution found at the end of Sec. 3.3, we can write exactly the same equation as in that section but now removing the overbars to indicate an arbitrary gain solution. In addition, just like in the low-gain regime, we can introduce Schmidt operators and squeezing parameters. The fundamental difference is that here the joint amplitude is determined by the Takagi–Autonne singular values and vectors of the matrix moment $\boldsymbol{M}$ as defined in (218) which is obtained by solving the Heisenberg equations of motion. The general procedure to solve for the ket is schematically represented in Fig. 8 [89].

 figure: Figure 8.

Figure 8. Block diagram describing how to calculate the state generated by a quadratic bosonic Hamiltonian.

Download Full Size | PDF

3.6 Nondegenerate Squeezing

As we show later, it is often useful to split our $\ell$-modes into two separate groups, called the signal and idler modes, for which we write ladder operators $b_i$ and $c_j$, respectively, satisfying canonical commutation relations

$$[b_i, b_j] = [c_i,c_j] = [b_i,c_j] = [b_i,c_j^\dagger] = 0, \quad [b_i,b_j^\dagger] = [c_i,c_j^\dagger] = \delta_{ij}.$$
The signal and idlers have $\ell _b$ and $\ell _c$ modes ($\ell _b+\ell _c = \ell$), respectively, and we write their nondegenerate squeezing Hamiltonian as
$$H(t) = \hbar \left\{\sum_{i,j=1}^{\ell} \Delta^{b}_{ij}(t) b_i^\dagger b_j + \sum_{i,j=1}^{\ell} \Delta^{c}_{ij}(t) c_i^\dagger c_j + i \left[ \sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_c} \zeta^{b,c}_{ij}(t) b_i^\dagger c_j^\dagger{-} \text{H.c.} \right] \right\}.$$
The separation between the $b$ and $c$ modes is physically motivated by observing that there might be situations where two different polarizations, or spectral regions that are widely separated and have disjoint support, interact. It is then convenient to assign different operators to the two sets of modes participating in the Hamiltonian. Whenever this is the case, we call the process described by the Hamiltonian above nondegenerate squeezing; this is to contrast it with the degenerate squeezing discussed in the previous section.

The Hamiltonian in (238) looks superficially different to the one in (188) but it is easy to show that the former is just a special case of the latter by identifying

$${\begin{aligned}\boldsymbol{a} = \begin{pmatrix} \boldsymbol{b} \\ \boldsymbol{c} \end{pmatrix}, \;\; \boldsymbol{b} = \begin{pmatrix} b_1 \\ \vdots \\ b_{\ell_b} \end{pmatrix}, \quad \boldsymbol{c} = \begin{pmatrix} c_1 \\ \vdots \\ c_{\ell_c} \end{pmatrix}, \;\; \boldsymbol{\Delta}(t) = \begin{pmatrix} \boldsymbol{\Delta}_b(t) & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{\Delta}_c(t) \end{pmatrix}, \;\; \boldsymbol{\zeta}(t) = \begin{pmatrix} \boldsymbol{0} & \boldsymbol{\zeta}^{b,c}(t)\\ [\boldsymbol{\zeta}^{b,c}(t)]^T & \boldsymbol{0} \end{pmatrix}. \end{aligned}}$$
Having written this problem in the general framework of the previous sections, we could simply use the techniques developed previously to solve it. However, it is often more useful to study this problem separately from the general problem discussed before. Thus, we start by looking at perturbative solutions, where we assume for simplicity that $\boldsymbol{\Delta }^{b,c}(t) = 0$. Note that if these terms are time independent, then one can without loss of generality transform to a rotating frame where the only term appearing in the Hamiltonian is that corresponding to $\boldsymbol{\zeta }^{b,c}$. Under these assumptions we can use the Magnus series at the first order to write the perturbative solution to the spontaneous problem as [86,90,91]
$$ |{\overline{\Psi}(t_\text{f})}\rangle \approx \overline{\mathcal{U}}(t_\text{f}, t_\text{i}) |{\text{vac}}\rangle $$
$$ \overline{\mathcal{U}}(t_\text{f}, t_\text{i}) = \exp\left(-\frac{i}{\hbar} \int_{t_{\text{i}}}^{t_\text{f}} \textrm{d}t' H(t') \right) = \exp\left[ \sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_c} \overline{J}^{b,c}_{ij} b_i^\dagger c_j^\dagger{-} \text{H.c.}\right], $$
where $\overline {\boldsymbol{J}}^{b,c} = \int _{t_{\text {i}}}^{t_{\text {f}}} dt\ \boldsymbol{\zeta }^{b,c}(t)$ is the joint amplitude of the nondegenerate squeezed state. As before, it is often useful to rewrite this state by making use of a decomposition of $\boldsymbol{J}$, here the SVD
$$\overline{\boldsymbol{J}}^{b,c} = \overline{\boldsymbol{F}}^{b} \ \overline{\boldsymbol{D}} \ [\overline{\boldsymbol{F}}^{c} ]^T,$$
where $\overline {\boldsymbol{F}}^{b/c}$ are $\ell _{b/c} \times \ell _{b/c}$ unitary matrices and $\boldsymbol{D}$ is an $\ell _b \times \ell _{c}$ diagonal matrix with entries $\{\overline {r}_1,\ldots, \overline {r}_{\ell _{\rm min} } \}$ with $\ell _{\rm min} = \min (\ell _b, \ell _c)$. We can use this SVD to introduce Schmidt modes for the signal and idler operators
$$ \overline{B}_\lambda^\dagger{=} \sum_{k=1}^{\ell_b} b_k^\dagger \overline{F}^b_{k,\lambda} \Longleftrightarrow b_k^\dagger{=} \sum_{i=1}^{\ell_b} [\overline{F}^b]^*_{k,\lambda} \overline{B}_{\lambda}^\dagger, $$
$$ \overline{C}_\lambda^\dagger{=} \sum_{k=1}^{\ell_c} c_k^\dagger \overline{F}^c_{k,\lambda} \Longleftrightarrow c_k^\dagger{=} \sum_{i=1}^{\ell_c} [\overline{F}^c]^*_{k,\lambda} \overline{C}_{\lambda}^\dagger. $$
The Schmidt operators satisfy bona fide bosonic commutation relations and we can finally write the output ket as
$$\begin{aligned} |{\overline{\Psi}(t_{\text{f}})}\rangle& = \exp\left[ \sum_{\lambda=1}^{\ell_{\rm min}} \overline{r}_{\lambda} \overline{B}_{\lambda}^\dagger \overline{C}_{\lambda}^\dagger{-} \text{H.c.}\right] |{\text{vac}}\rangle = \bigotimes_{\lambda=1}^{\ell_{\rm min}} \exp\left[ \overline{r}_{\lambda} \overline{B}_{\lambda}^\dagger \overline{C}_{\lambda}^\dagger{-} \text{H.c.}\right] |{\text{vac}}\rangle\\& = \bigotimes_{\lambda=1}^{\ell_{\rm min}} \left[ \frac{1}{\cosh \overline{r}_\lambda} \sum_{n=0}^\infty \tanh^n \overline{r}_\lambda |{n,n}\rangle_{\lambda} \right],\end{aligned}$$
where in the last line we have used a well-known disentangling identity (cf. Appendix 5 of Barnett and Radmore [70])
$$\begin{aligned} \exp(r e^{i \phi} b^\dagger c^\dagger{-}\text{H.c.}) = & \exp\left( e^{i \phi} \tanh r \ c^{{\dagger} } b^\dagger\right) \\ & \times \exp\left( - \left[ c^\dagger c+ b^\dagger b +1\right] \ln \cosh r \right)\\ & \times \exp\left( - e^{{-}i \phi} \tanh r \ c b\right), \end{aligned}$$
and introduced
$$|{n,n}\rangle_{\lambda} = \frac{\overline{B}_\lambda^{{\dagger} n} \overline{C}_\lambda^{{\dagger} n}}{n!} |{\text{vac}}\rangle.$$
We can also use the low-gain perturbative solution in (240) to transform the operators, for which we find
$$ b_i(t_{\text{f}}) = \overline{\mathcal{U}}^\dagger b_i(t_{\text{i}}) \overline{\mathcal{U}} = \sum_{j=1}^\ell \overline{V}^{b,b}_{ij} b_{j}(t_{\text{i}}) + \sum_{j=1}^\ell \overline{W}^{b,c}_{ij} c^\dagger _{j} (t_{\text{i}}), $$
$$ c_i^\dagger(t_{\text{f}}) = \overline{\mathcal{U}}^\dagger c_i^\dagger(t_{\text{i}}) \overline{\mathcal{U}} = \sum_{j=1}^\ell \left[\overline{V}^{(c,c)}_{ij}\right]^* c^\dagger_{j}(t_{\text{i}}) + \sum_{j=1}^\ell \left[\overline{W}^{c,b}_{ij}\right]^* b _{j}(t_{\text{i}}), $$
where now
$$ \overline{\boldsymbol{V}}^{b,b} = \overline{\boldsymbol{F}}^{b} \left[ \oplus_{\lambda=1}^\ell \cosh \overline{r}_{\lambda} \right] \left[\overline{\boldsymbol{F}}^{b} \right]^\dagger, \quad \overline{\boldsymbol{V}}^{(c,c)} = \overline{\boldsymbol{F}}^{c} \left[ \oplus_{\lambda=1}^\ell \cosh \overline{r}_{\lambda} \right] \left[\overline{\boldsymbol{F}}^{c} \right]^\dagger, $$
$$ \overline{\boldsymbol{W}}^{b,c} = \overline{\boldsymbol{F}}^{b} \left[ \oplus_{\lambda=1}^\ell \sinh \overline{r}_{\lambda} \right] \left[\overline{\boldsymbol{F}}^{c} \right]^T, \quad \overline{\boldsymbol{W}}^{c,b} = \overline{\boldsymbol{F}}^{c} \left[ \oplus_{\lambda=1}^\ell \sinh \overline{r}_{\lambda} \right] \left[\overline{\boldsymbol{F}}^{b} \right]^T = \left[ \overline{\boldsymbol{W}}^{b,c} \right]^T. $$
To lowest nonvanishing order in the squeezing parameters $\overline {r}_\lambda \ll 1$, one has
$$\overline{\boldsymbol{V}}^{b,b} \approx \mathbb{I}_{\ell_b}, \overline{\boldsymbol{V}}^{c,c} \approx \mathbb{I}_{\ell_c} \quad \text{and} \quad \overline{\boldsymbol{W}}^{b,c} = \left[\overline{\boldsymbol{W}}^{c,b} \right]^T \approx \overline{\boldsymbol{J}}^{b,c}.$$
As the inverse of the unitary operator $\overline {\mathcal {U}}$ can be obtained by reversing the sign of $\overline {\boldsymbol{J}}^{b,c}$ inside the exponential, it is straightforward to show the backward-Heisenberg evolved operators $\overline {\mathcal {U}} b_i(t_{\text {i}}) \overline {\mathcal {U}} ^\dagger$ can be obtained from (248) by negating the sign of the terms $\overline {\boldsymbol{W}}^{b,c}$ and $\overline {\boldsymbol{W}}^{c,b}$.

We can now move to the high-gain regime, where we find that the solution to the spontaneous problem has exactly the same form as in the low-gain regime, but now the joint amplitude of the nondegenerate squeezed state will be inferred from the SVD of a second-order moment of the operators. As previously, we write the equations of motion for $\boldsymbol{b}$ and $\boldsymbol{c}^\dagger$ as

$$\begin{aligned}\frac{\textrm{d}}{\textrm{d}t} \begin{pmatrix} \boldsymbol{b} \\ \boldsymbol{c}^\dagger \end{pmatrix} ={-}i \underbrace{\begin{pmatrix} - \boldsymbol{\Delta}_b(t) & i \boldsymbol{\zeta}^{b,c}(t) \\ i [\boldsymbol{\zeta}^{b,c}(t)]^\dagger & \boldsymbol{\Delta}_c(t) \end{pmatrix} }_{{\equiv} \boldsymbol{A}^{b,c}(t)} \begin{pmatrix} \boldsymbol{b} \\ \boldsymbol{c}^\dagger \end{pmatrix}. \end{aligned}$$
The solution to this equation can be found with the same techniques as before, yielding
$$\begin{aligned}\begin{pmatrix} \boldsymbol{b}(t_{\text{f}}) \\ \boldsymbol{c}^\dagger(t_{\text{f}}) \end{pmatrix} = & \mathcal{T}\exp\left[ - i \int_{t_{\text{i}}}^{t_\text{f}} \textrm{d}t \boldsymbol{A}^{b,c}(t) \right] \begin{pmatrix} \boldsymbol{b}(t_{\text{i}}) \\ \boldsymbol{c}^\dagger(t_{\text{i}}) \end{pmatrix} = \begin{pmatrix} \boldsymbol{V}^{b,b} & \boldsymbol{W}^{b,c} \\ (\boldsymbol{W}^{c,b})^* & (\boldsymbol{V}^{c,c})^* \end{pmatrix} \begin{pmatrix} \boldsymbol{b}(t_{\text{i}}) \\ \boldsymbol{c}^\dagger(t_{\text{i}}) \end{pmatrix} \\ = & \boldsymbol{K}^{b,c}(t_{\text{f}},t_{\text{i}}) \begin{pmatrix} \boldsymbol{b}(t_{\text{i}}) \\ \boldsymbol{c}^\dagger(t_{\text{i}}) \end{pmatrix}. \end{aligned}$$
These transfer functions can be accessed using classical intensity measurements as shown in Fig. 9. By correctly modeling the different terms in the equations of motion one can obtain excellent agreement between theory and simulation.

 figure: Figure 9.

Figure 9. Experimental and theoretical prediction for the blocks of the transfer function $\boldsymbol{K}^{b,c}$ for nondegenerate squeezing in a periodically poled potassium titanyl phosphate SPDC source. The labels $U_{s,s}, U_{i,s}, U_{s,i}, U_{i,i}$ in the figure correspond to $V_{b,b}, W_{c,b}, W_{b,c}, W_{c,c}$ in the notation used in this paper. Figure 9 reprinted with permission from Triginer et al., Phys. Rev. X 10, 031063 (2019) [92]. Copyright 2019 by the American Physical Society.

Download Full Size | PDF

Using this input–output relation we can now calculate the (nonzero) second-order moments

$$\begin{aligned}N^b_{ij} = \rangle{\text{vac}}| b_i^\dagger(t_{\text{f}}) b_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger b_i^\dagger b_j \mathcal{U} |{\text{vac}}\rangle, \end{aligned}$$
$$\begin{aligned}N^c_{ij} = \rangle{\text{vac}}| c_i^\dagger(t_{\text{f}}) c_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger c_i^\dagger c_j \mathcal{U} |{\text{vac}}\rangle, \end{aligned}$$
$$\begin{aligned}M^{b,c}_{ij} = \rangle{\text{vac}}| b_i(t_{\text{f}}) c_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger b_i c_j \mathcal{U} |{\text{vac}}\rangle, \end{aligned}$$
and verify that the following moments are identically zero
$$\begin{aligned}N^{b,c}_{ij} = \rangle{\text{vac}}| b_i^\dagger(t_{\text{f}}) c_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger b_i^\dagger c_j \mathcal{U} |{\text{vac}}\rangle = 0, \end{aligned}$$
$$\begin{aligned}M^{b}_{ij} = \rangle{\text{vac}}| b_i(t_{\text{f}}) b_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger b_i b_j \mathcal{U} |{\text{vac}}\rangle =0, \end{aligned}$$
$$\begin{aligned}M^{c}_{ij} = \rangle{\text{vac}}| c_i(t_{\text{f}}) c_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger c_i c_j \mathcal{U} |{\text{vac}}\rangle =0. \end{aligned}$$
Note that if we identify again the $\boldsymbol{b}$ and $\boldsymbol{c}$ modes as a subset of the larger set of $\boldsymbol{a}$ modes we can write the moments of the latter as
$$\begin{aligned} \boldsymbol{N} = \begin{pmatrix} \boldsymbol{N}^b & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{N}^c \end{pmatrix}, \quad \boldsymbol{M} = \begin{pmatrix} \boldsymbol{0} & \boldsymbol{M}^{b,c} \\ [\boldsymbol{M}^{b,c}]^T & \boldsymbol{0} \end{pmatrix}. \end{aligned}$$
Finally, after having the moments we simply need to find the SVD [92]
$$\boldsymbol{M}^{b,c} = {\boldsymbol{F}}^{b} \ {\boldsymbol{D}} \ [{\boldsymbol{F}}^{c} ]^T,$$
where $\boldsymbol{F}^{b/c}$ are $\ell _{b/c} \times \ell _{b/c}$ unitary matrices and $\boldsymbol{D}$ is an $\ell _b \times \ell _{c}$ diagonal matrix with entries $\{\tfrac 12 \sinh 2{r}_1,\ldots, \tfrac 12 \sinh 2{r}_{\ell _{\rm min} } \}$ where $\ell _{\rm min} = \min (\ell _b, \ell _c)$.

Note that the output of the spontaneous problem is also a Gaussian state, characterized by its second-order moments and that we can write the output ket as

$$|{\Psi(t_{\text{f}})}\rangle = \mathcal{U} |{\text{vac}}\rangle = \exp\left[ \sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_c} {J}^{b,c}_{ij} b_i^\dagger c_j^\dagger{-} \text{H.c.}\right] |{\text{vac}}\rangle,$$
where now $\boldsymbol{J}^{b,c} = {\boldsymbol{F}}^{b} \ {\boldsymbol{R}} \ [{\boldsymbol{F}}^{c} ]^T$ where $\boldsymbol{R}$ is an $\ell _b \times \ell _{c}$ diagonal matrix with entries $\{{r}_1,\ldots, {r}_{\ell _{\rm min} } \}$ that is directly related to the matrix $\boldsymbol{D}$ in the SVD of $\boldsymbol{M}^{b,c}$ of (262). This form is functionally the same as the ket in the low-gain regime, and thus arbitrary-gain Schmidt modes can be introduced in exactly the same way as before.

3.7 Solution to the Stimulated Problem

Having solved the spontaneous problem associated with a general quadratic Hamiltonian, we now solve the associated stimulated problem, i.e., we study the evolution of other states different from vacuum. One can write an arbitrary input state as

$$|{\Psi(t_{\text{i}})}\rangle = f(a_i^\dagger) |{\text{vac}}\rangle,$$
where $f$ is either a polynomial or a power series of the creation operators. The parametrization above is the so-called Bargmann–Segal or stellar representation of a bosonic quantum state [29,93,94]. For a single photon and a coherent state we can write, respectively,
$$|{1_{\boldsymbol{\alpha}}}\rangle = \sum_{i=1}^\ell \alpha_i a_i^\dagger |{\text{vac}}\rangle, \quad |{\boldsymbol{\alpha}}\rangle = \mathcal{D}(\boldsymbol{\alpha})|{\text{vac}}\rangle = e^{-\tfrac12 \| \boldsymbol{\alpha}\|^2} \exp\left[ \sum_{i=1}^\ell \alpha_i a_i^\dagger \right] |{\text{vac}}\rangle.$$
We can then write the output ket as
$$|{\Psi(t_{\text{f}})}\rangle = \mathcal{U}|{\Psi(t_{\text{i}})}\rangle = \mathcal{U} f(a_i^\dagger)\mathcal{U}^\dagger \mathcal{U}|{\text{vac}}\rangle,$$
the part $\mathcal {U}| {\text {vac}}\rangle$ corresponds to the spontaneous problem solved in the previous section. For the second part, $\mathcal {U} f(a_i^\dagger )\mathcal {U}^\dagger$ we can write
$$\mathcal{U} f(a_i^\dagger)\mathcal{U}^\dagger{=} f(\mathcal{U} a_i^\dagger \mathcal{U}^\dagger) = f\left(\sum_{k=1}^\ell \tilde{V}^*_{ik} a_k^\dagger{+} \sum_{k=1}^\ell \tilde{W}^*_{ik} a_k \right),$$
where in the last line we used the fact that $\mathcal {U} O^m \mathcal {U}^\dagger = \left [ \mathcal {U} O \mathcal {U}^\dagger \right ]^m$ and the backward-Heisenberg transformation of (222). Going back to the initial states considered before in (266) we find that an input single photon becomes a superposition of a photon added and a photon subtracted squeezed state
$$\mathcal{U}|{1_{\boldsymbol{\alpha}}}\rangle = \left(\sum_{k=1}^\ell \tilde{V}^*_{ik} a_k^\dagger{+} \sum_{k=1}^\ell \tilde{W}^*_{ik} a_k \right) \mathcal{U} |{\text{vac}}\rangle,$$
whereas for an input coherent state, the reader can confirm that
$$\mathcal{U}|{\boldsymbol{\alpha}}\rangle = \mathcal{D}\left( \boldsymbol{V}\boldsymbol{\alpha} + \boldsymbol{W} \boldsymbol{\alpha}^* \right) \mathcal{U} |{\text{vac}}\rangle$$
and, thus, the output is a displaced-squeezed state, where the input coherent displacement is transformed in precisely the same way as the bosonic operators [cf. (212)]. Note also that the information about the JSA of the squeezed part of the state $\mathcal {U} | {\text {vac}}\rangle$ is imprinted in the displacement since it depends on $\boldsymbol{V}$ and $\boldsymbol{W}$ which uniquely determine $\boldsymbol{J}$. For the case of nondegenerate squeezing this connection can be made more explicit in the low-gain regime discussed at the beginning of Sec. 3.6. Consider the following coherent state input in the idler modes:
$$|{\boldsymbol{\nu}}\rangle = \mathcal{D}_c(\boldsymbol{\nu}) |{\text{vac}}\rangle = \exp\left[ \boldsymbol{\nu}^T \boldsymbol{c}^\dagger{-} \boldsymbol{\nu}^\dagger \boldsymbol{c} \right] |{\text{vac}}\rangle.$$
Using the results from that section, and to lowest nonvanishing order in the squeezing parameters, we find
$$\begin{aligned}\overline{\mathcal{U}} |{\boldsymbol{\nu}}\rangle & = \overline{\mathcal{U}} \exp\left[ \boldsymbol{\nu}^T \boldsymbol{c}^\dagger{-} \boldsymbol{\nu}^\dagger \boldsymbol{c} \right] \overline{\mathcal{U}}^\dagger \overline{\mathcal{U}} |{\text{vac}}\rangle \\ & =\exp\left[ \boldsymbol{\nu}^T \overline{\mathcal{U}} \boldsymbol{c}^\dagger \overline{\mathcal{U}}^\dagger{-} \boldsymbol{\nu}^\dagger \overline{\mathcal{U}}\boldsymbol{c}\overline{\mathcal{U}}^\dagger \right] \overline{\mathcal{U}} |{\text{vac}}\rangle. \end{aligned}$$
Recalling that the backward-Heisenberg operator $\overline {\mathcal {U}} \boldsymbol{c}^\dagger \overline {\mathcal {U}}^\dagger$ can be obtained from the forward-Heisenberg operators in the low-gain regime (248) and (252) by letting $r_\lambda \to -r_\lambda$ (equivalently letting $\boldsymbol{J}^{b,c} \to - \boldsymbol{J}^{b,c}$), we obtain
$$\begin{aligned} \overline{\mathcal{U}} |{\boldsymbol{\nu}}\rangle & =\exp\left[ \boldsymbol{\nu}^T \left(\boldsymbol{c}^\dagger{-} [\overline{\boldsymbol{J}}^{c,b}]^* \boldsymbol{b} \right) - \boldsymbol{\nu}^\dagger \left(\boldsymbol{c} - \overline{\boldsymbol{J}}^{c,b} \boldsymbol{b}^\dagger \right) \right] \overline{\mathcal{U}} |{\text{vac}}\rangle \nonumber\\ & =\exp\left[ \boldsymbol{\nu}^T \boldsymbol{c}^\dagger{-} \boldsymbol{\nu}^\dagger \boldsymbol{c} \right] \exp\left[ \boldsymbol{\nu}^\dagger \overline{\boldsymbol{J}}^{c,b} \boldsymbol{b}^\dagger{-} \boldsymbol{\nu}^T [\overline{\boldsymbol{J}}^{c,b}]^* \boldsymbol{b} \right] \overline{\mathcal{U}} |{\text{vac}}\rangle \\ & = \mathcal{D}_{\boldsymbol{c}}(\boldsymbol{\nu}) \mathcal{D}_{\boldsymbol{b}}( [\overline{\boldsymbol{J}}^{c,b}]^T \boldsymbol{\nu}^*) \overline{\mathcal{U}} |{\text{vac}}\rangle\nonumber\\ & = \mathcal{D}_{\boldsymbol{c}}(\boldsymbol{\nu}) \mathcal{D}_{\boldsymbol{b}}( \overline{\boldsymbol{J}}^{b,c} \boldsymbol{\nu}^*) \exp\left[ \sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_c} \overline{J}^{b,c}_{ij} b_i^\dagger c_j^\dagger{-} \text{H.c.}\right] |{\text{vac}}\rangle. \end{aligned}$$
Note that now the joint amplitude of the nondegenerate squeezed state not only determines the squeezed part of the state but it also appears as a response function telling us the displacement imparted to the signal modes ($\boldsymbol{b}$) when the idler modes ($\boldsymbol{c}$) are seeded by a coherent displacement amplitude $\boldsymbol{\nu }$.

3.8 Passive Operations and Loss

An important set of processes in the evolution of bosonic modes are given by passive operations, i.e., linear operations where the energy is not increased. In these process, the modes of an environment prepared in the vacuum couple to the modes of interest via a beam-splitter-like interaction, which in the Heisenberg picture we write as [9597]

$$\boldsymbol{a}(t_\text{f}) = \mathcal{U}_{\text{BS}}^\dagger \boldsymbol{a}(t_{\text{i}}) \mathcal{U}_{\text{BS}} = \boldsymbol{L} \boldsymbol{a}(t_{\text{i}}) + \sqrt{\mathbb{I}_\ell - \boldsymbol{L} \boldsymbol{L}^\dagger } \boldsymbol{e},$$
where $\boldsymbol{L}$ is a loss matrix, that in order to represent a physical process satisfies that all its singular values are bounded by 1 and $\boldsymbol{e}$ are bosonic destruction operators of the environment. Note, in particular, that if the losses are uniform, then $\boldsymbol{L} = \sqrt {\eta } \mathbb {I}_\ell$ with $0 \leq \eta \leq 1$ being the net energy transmission. The case of nonlossy operation is obtained when the matrix $\boldsymbol{L}$ is unitary, in which case $\boldsymbol{L} = \boldsymbol{U}$ and, thus,
$$\boldsymbol{a}(t_\text{f}) = \mathcal{U}_{\text{BS}}^\dagger \boldsymbol{a}(t_{\text{i}}) \mathcal{U}_{\text{BS}} = \boldsymbol{U} \boldsymbol{a}(t_{\text{i}}).$$
As the transformations above are linear, we can show directly that if the input state at time $t_{\text {i}}$ is Gaussian, then the output state is also Gaussian. This implies that the output state is uniquely characterized by the $\boldsymbol{M}$ and $\boldsymbol{N}$ moments. We can see how they transform by writing, for example,
$$\begin{aligned}M_{ij} (t_{\text{f}}) & = \text{tr}\left( a_i a_j \mathcal{U}_{\text{BS}} \left[ \rho_{S} \otimes |{\text{vac}_e}\rangle \rangle{\text{vac}_e}| \right] \mathcal{U}_{\text{BS}}^\dagger \right) \\ & =\text{tr}\left( \mathcal{U}_{\text{BS}}^\dagger a_i a_j \mathcal{U}_{\text{BS}} \left[ \rho_{S} \otimes |{\text{vac}_e}\rangle \rangle{\text{vac}_e}| \right] \right)\\ & =\text{tr}\left( \left[\boldsymbol{L} \boldsymbol{a}(t_{\text{i}}) + \sqrt{\mathbb{I}_\ell - \boldsymbol{L} \boldsymbol{L}^\dagger } \ \boldsymbol{e}\right]_i \left[\boldsymbol{L} \boldsymbol{a}(t_{\text{i}}) + \sqrt{\mathbb{I}_\ell - \boldsymbol{L} \boldsymbol{L}^\dagger } \ \boldsymbol{e}\right]_j \left[ \rho_{S} \otimes |{\text{vac}_e}\rangle \rangle{\text{vac}_e}| \right] \right)\\ & =\text{tr}\left( \left[\boldsymbol{L} \boldsymbol{a}(t_{\text{i}}) \right]_i \left[\boldsymbol{L} \boldsymbol{a}(t_{\text{i}}) \right]_j \left[ \rho_{S} \otimes |{\text{vac}_e}\rangle \rangle{\text{vac}_e}| \right] \right)\\ & =\text{tr}\left( \sum_{k=1}^\ell L_{ik} {a}_k(t_{\text{i}}) \sum_{l=1}^\ell L_{jl} {a}_l(t_{\text{i}}) \rho_{S} \right) = \left[ \boldsymbol{L} \boldsymbol{M}(t_{\text{i}}) \boldsymbol{L}^T \right]_{ij}, \end{aligned}$$
where we used the cyclic property of the trace and the fact that $\boldsymbol{e} | {\text {vac}_e}\rangle = 0$, and wrote $\rho _S$ for the density operator of the input modes. Similarly, by using the cyclic property of the trace, one finds that
$$\boldsymbol{N}(t_{\text{f}}) = \boldsymbol{L}^* \boldsymbol{N}(t_{\text{i}}) \boldsymbol{L}^T.$$
Note that in the special case where the loss is uniform we have
$$\boldsymbol{N}(t_{\text{f}}) =\eta \boldsymbol{N}(t_{\text{i}}), \quad \boldsymbol{M}(t_{\text{f}}) = \eta \boldsymbol{M}(t_{\text{i}}),$$
which gives the identity operation when $\eta =1$ (no loss) and the complete loss channel when $\eta =0$.

3.9 Photon-Number and Homodyne Statistics

In this section we study properties of the photon number and quadrature statistics of the Gaussian states described in the previous sections. We first look at pure lossless states of the form given in (235) and ask what can be said about the statistical moments of the total photon number observable

$$N_{\boldsymbol{a}} = \sum_{i=1}^\ell a_i^\dagger a_i.$$
Although the mean photon number is simply the sum of the means of the different “bare” modes labeled by $i$, the variance of the total photon number is not the sum of the variances of the “bare” modes because these modes will be, in general, correlated. We can rewrite the total photon number in terms of the Schmidt operators finding
$$N_{\boldsymbol{a}} = \sum_{\lambda=1}^\ell A_\lambda^\dagger A_\lambda,$$
and, thus, since the Schmidt modes are not correlated, we can use (180) to obtain
$${\displaystyle\langle{N_{\boldsymbol{a}}}\rangle = \sum_{\lambda=1}^{\ell} n_\lambda, \quad \text{Var}(N_{\boldsymbol{a}}) = \sum_{\lambda=1}^{\ell} \text{Var}( A^\dagger_\lambda A_\lambda) = \sum_{\lambda=1}^{\ell} 2 n_\lambda (1+n_\lambda), \quad n_\lambda = \langle{A^\dagger_\lambda A_\lambda}\rangle = \sinh^2 r_\lambda,}$$
where we introduced the variance as $\text {Var}(O) = \langle {O^2}\rangle - \langle {O}\rangle^2$. Similarly, for nondegenerate squeezed states we can easily use their Schmidt decomposition to write
$$\begin{aligned}\langle{N_{\boldsymbol{b}}}\rangle & = \langle{N_{\boldsymbol{c}}}\rangle = \sum_{\lambda=1}^{\min\{\ell_b, \ell_c\}} n_\lambda, \end{aligned}$$
$$\begin{aligned}\text{Var}(N_{\boldsymbol{b}}) & = \text{Var}(N_{\boldsymbol{c}})\\& = \sum_{\lambda=1}^{\min\{\ell_b, \ell_c\}} \text{Var}( B^\dagger_\lambda B_\lambda) = \sum_{\lambda=1}^{\min\{\ell_b, \ell_c\}} n_\lambda (1+n_\lambda), \quad n_\lambda = \langle{A^\dagger_\lambda A_\lambda}\rangle = \sinh^2 r_\lambda. \end{aligned}$$
More generally, assume that we have a state $\rho$ and that we wish to evaluate some normal-ordered moment of the form
$$\left\langle \prod_{i=1}^\ell a_i^{{\dagger} m_i } \prod_{i=1}^\ell a_i^{n_i } \right\rangle.$$
These moments can be obtained by taking derivatives of the normal-ordered characteristic function $\chi ^{N}(\boldsymbol{\alpha })$ which is related to the (symmetric-)ordered characteristic function introduced in (229) as
$$\chi^{N}_{\rho}(\boldsymbol{\alpha}) = \chi(\boldsymbol{\alpha}) e^{\tfrac12 \|\boldsymbol{\alpha}\|^2} = \left\langle \exp\left[ \boldsymbol{\alpha}^T \boldsymbol{a}^\dagger \right] \exp\left[ - \boldsymbol{\alpha}^\dagger \boldsymbol{a} \right] \right\rangle_\rho,$$
from which we easily confirm that
$$\left\langle \prod_{i=1}^\ell a_i^{{\dagger} m_i } \prod_{i=1}^\ell a_i^{n_i } \right\rangle = \left. ({-}1)^{\sum_{i=1}^\ell n_i}\frac{\partial^{m_1}}{\partial \alpha_1^{m_1}} \ldots \frac{\partial^{m_\ell}}{\partial \alpha_\ell^{m_\ell}} \frac{\partial^{n_1}}{\partial \alpha_1^{*n_1}} \ldots \frac{\partial^{n_\ell}}{\partial \alpha_\ell^{* n_\ell}} \chi^{N}_\rho(\boldsymbol{\alpha}) \right|_{\boldsymbol{\alpha}=\boldsymbol{\alpha}^* = 0}.$$
For a Gaussian state (having a Gaussian characteristic function) one easily finds
$$\begin{aligned} \langle{n_i}\rangle & = N_{i,i}, \end{aligned}$$
$$\begin{aligned}\text{Var}(n_i) & = \langle{n_i^2}\rangle - \langle{n_i}\rangle^2 = |M_{i,i}|^2 + N_{i,i}(N_{i,i}+1), \end{aligned}$$
$$\begin{aligned}\text{Cov}(n_i,n_j) & = |M_{ij}|^2 + |N_{ij}|^2, \end{aligned}$$
$$\begin{aligned}\text{Var}(n_i-n_j) & = \text{Var}(n_i) + \text{Var}(n_j) - 2 \ \text{Cov}(n_i,n_j) \\ & = |M_{i,i}|^2 + N_{i,i}(N_{i,i}+1) + |M_{j,j}|^2 + N_{j,j}(N_{j,j}+1) - 2 \left( |M_{ij}|^2 + |N_{ij}|^2 \right),\\ \end{aligned}$$
$$\begin{aligned}\text{Var}(N) & = \sum_{i=1}^\ell \text{Var}(n_i) + 2 \sum_{i<j}^\ell \text{Cov}(n_i,n_j), \end{aligned}$$
with $n_i = a_i^\dagger a_i$ and $n_i^2 = a^\dagger _i a_i a^\dagger _i a_i = a_i^{\dagger 2}a_i^2 + a^\dagger _i a_i$.

Let us apply these results to study so-called coherence functions [98]. For a beam with destruction operators $b_i$ the second-order coherence function is defined as

$$g^{(2)}_{\boldsymbol{b}} = \frac{\langle{N_{\boldsymbol{b}}^2 - N_{\boldsymbol{b}}}\rangle}{\langle{N_{\boldsymbol{b}}}\rangle^2} = \frac{\sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_b} \langle{b_i^\dagger b_j^\dagger b_i b_j}\rangle }{\left( \sum_{i=1}^{\ell_b} \langle{b_i^\dagger b_i}\rangle\right)^2},$$
where in analogy to (278) we define the total photon number of the signal modes $\boldsymbol{b}$ as $N_{\boldsymbol{b}} = \sum _{i=1}^{\ell _b} b_i^\dagger b_i$. Note that a similar quantity can be introduced for the idler modes by letting $b \to c$ in the last equation. One can also introduce the cross-beam second-order coherence function as
$$g^{(1,1)}_{\boldsymbol{b},\boldsymbol{c}} = \frac{\langle{N_{\boldsymbol{b}} N_{\boldsymbol{c}}}\rangle}{\langle{N_{\boldsymbol{b}}}\rangle \langle{N_{\boldsymbol{c}}}\rangle} = \frac{\sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_c} \langle{b_i^\dagger c_j^\dagger b_i c_j}\rangle }{\left( \sum_{i=1}^{\ell_b} \langle{b_i^\dagger b_i}\rangle\right) \left( \sum_{i=1}^{\ell_c} \langle{c_i^\dagger c_i}\rangle\right)}.$$

The first interesting property of these quantities is that they are not affected by a loss. Assume, for example, that a beam is prepared in certain quantum state $\rho _{\boldsymbol{b}}$ and then undergoes uniform loss by energy transmission $\eta _{\boldsymbol{b}}$ as described in Sec. 3.8. It is straightforward to see that because the numerator and denominator of (291) are normal ordered in the creation and annihilation operators of the beam $\boldsymbol{b}$, then each of them will pick up a factor of $\eta _{\boldsymbol{b}}^2$ that will cancel out. Similarly, for the case of the cross-beam second-order correlation function, if beams $\boldsymbol{b}$ and $\boldsymbol{c}$ undergo uniform loss by energy transmissions $\eta _{\boldsymbol{b}}$ and $\eta _{\boldsymbol{c}}$, respectively, then both the numerator and denominator in the right-hand side of (292) will pick up a factor $\eta _{\boldsymbol{b}} \eta _{\boldsymbol{c}}$ that will cancel out. We emphasize that these results are true irrespective of the state prepared before loss. We now use these results to investigate the correlation functions of degenerate and nondegenerate squeezed light. For nondegenerate squeezed states it is easy to show [99]

$$g^{(2)}_{\boldsymbol{b}} = g^{(2)}_{\boldsymbol{c}} = 1 + \frac{1}{K}, \text{ where } K = \frac{ \left[ \sum_{\lambda=1}^{\min(\ell_c,\ell_b)} \langle{A_\lambda^\dagger A_\lambda}\rangle \right]^2 }{\sum_{\lambda=1}^{\min(\ell_c,\ell_b)} \langle{A_\lambda^\dagger A_\lambda}\rangle^2}.$$
The quantity $K$ is the so-called Schmidt number of the of nondegenerate squeezed state. Note that this quantity is equal to one only if one of the Schmidt modes is occupied $\langle {A_1^\dagger A_1}\rangle >0$ and all the rest ($\lambda >1$) have zero photons $\langle {A_\lambda ^\dagger A_\lambda }\rangle =0$. If more than one Schmidt is occupied, then $K>1$ and $g^{(2)}_{\boldsymbol{b}/\boldsymbol{c}} <2$. Similarly, we find that the cross correlation function is given by
$$g^{(1,1)}_{\boldsymbol{b},\boldsymbol{c}} = 1 + \frac{1}{K} + \frac{1}{\langle{N}\rangle},$$
where we wrote the total photon number of either the signal or idler beam as $\langle {N}\rangle = \sum _{\lambda =1}^{\min (\ell _c,\ell _b)} \langle {A_\lambda ^\dagger A_\lambda }\rangle$. Note that this is the total photon number before any propagation loss occurs, thus, by subtracting the second-order correlation function of either of the beams from the cross correlation coherence function one has access to the absolute number of photons generated by the source. These can be compared with the total average photon numbers measured in the detectors allowing to obtain the inefficiencies, or energy transmission factors, $\eta _{\boldsymbol{b}},\eta _{\boldsymbol{c}}$; if we write the measured mean photon numbers after transmission $\eta _{\boldsymbol{b}/\boldsymbol{c}}$ as
$$\langle{N_{\boldsymbol{b}/\boldsymbol{c}}}\rangle = \eta_{\boldsymbol{b}/\boldsymbol{c}} \langle{N}\rangle,$$
then we can obtain an absolute determination of the transmission by forming
$$\eta_{\boldsymbol{b}/\boldsymbol{c}} = \langle{N_{\boldsymbol{b}/\boldsymbol{c}}}\rangle \left( g^{(2)}_{\boldsymbol{b}/\boldsymbol{c}} - g^{(1,1)}_{\boldsymbol{b},\boldsymbol{c}} \right),$$
which is a photon-number resolved extension of the elegant Klyshko method [42].

The question of calculating individual photon-number probabilities of a multimode Gaussian state has led to the introduction of the Gaussian boson sampling problem [100102]. In general the evaluation of these probabilities is computationally hard. Moreover, even the computational task of generating samples that distribute according to this probability distribution is believed to be hard (modulo reasonable complexity-theoretic assumptions [103]). The study of this problem is beyond the scope of this tutorial, however we point the reader to a number of recent theoretical [9,97,104] and experimental advances [105,106].

One can also study homodyne measurements where now one is interested in statistics of the quadrature observable:

$$X_{\phi} = e^{i \phi} \sum_{i=1}^\ell \alpha_i^* a_i + \text{H.c.}$$
It is direct to see that this observable has $\langle {X_\phi }\rangle = 0$ and variance
$$V_{\phi} = \|\boldsymbol{\alpha}\|^2 + 2\boldsymbol{\alpha}^T \boldsymbol{N} \boldsymbol{\alpha}^* + \left( e^{2i\phi} \boldsymbol{\alpha}^\dagger \boldsymbol{M} \boldsymbol{\alpha}^* +\text{c.c.} \right).$$
Note that only the phase-sensitive moment $\boldsymbol{M}$ couples to the local oscillator (LO) phase $\phi$. The phase insensitive moment $\boldsymbol{N}$ only adds noise to the variance. Moreover, if we select $\alpha _k = F_{k,\lambda }$ for a fixed Schmidt mode $\lambda$ then the maximum and minimum variance as $\phi$ is varied are given by
$$V_{\phi}^{\rm max/min} = e^{{\pm} 2 r_\lambda},$$
where $r_\lambda$ is the squeezing parameter of the Schmidt mode $\lambda$ and we have used the fact that the Schmidt mode functions have unit norm. This is exactly like the single-mode case described at the end of Sec. 3.1.

Considering now nondegenerate squeezing, assuming the LOs for the signal and idler modes are given by the amplitudes $\boldsymbol{\mu }$ and $\boldsymbol{\nu }$ respectively and defining the quadrature observable

$$X_{\phi} = e^{i \phi} \left[ \sum_{i=1}^{\ell_b} \mu_i^* b_i + \sum_{j=1}^{\ell_c} \nu_j^* c_j\right] + \text{H.c.},$$
we find that the variance is given by
$$V_{\phi} = \|\boldsymbol{\nu}\|^2 + \|\boldsymbol{\mu}\|^2 + 2\boldsymbol{\mu}^T \boldsymbol{N}^{b} \boldsymbol{\mu}^* + 2 \boldsymbol{\nu}^T \boldsymbol{N}^{c} \boldsymbol{\nu}^* +\left[ 2 e^{2i \phi} \boldsymbol{\mu}^\dagger \boldsymbol{M}^{b,c} \boldsymbol{\nu}^* + \text{c.c.} \right].$$
To obtain this result we use (261) mapping a nondegenerate state into a degenerate squeezer and write the equivalent degenerate squeezer LO as amplitude $\boldsymbol{\alpha } = \left ( \begin {smallmatrix} \boldsymbol{\mu } \\ \boldsymbol{\nu } \end {smallmatrix} \right )$.

4. Heisenberg Equations for Waveguides

In this section, following [107] and [108], we present a general formalism for considering degenerate as well as nondegenerate squeezing in waveguides where either a second- or third-order nonlinearity is dominant. The six nonlinear optical processes that this includes are SPDC, SP-SFWM, and DP-SFWM for the creation of degenerate squeezed vacuum (DSV) states such as those in Sec. 3.5 as well as SPDC, SP-SFWM, and DP-SFWM for the creation of nondegenerate squeezed vacuum (NDSV) states such as those in Sec. 3.6. Beginning with Hamiltonians of the form presented in Sec. 2, suited to integrated photonic structures, we find that the Heisenberg equations of motion, in the limit of strong classical pumps and the undepleted pump approximation, lead to the expected classical coupled mode equations for the pump fields, as well as coupled operator equations for the generated fields. We then discretize coupled operator equations to arrive at compact equations of the form of (207), which can subsequently be solved according to the methods of Sec. 3.4. Perhaps somewhat surprisingly, despite having different physical origins, all of the coupled operator equations can be cast in the same general form.

All of the processes have the same linear Hamiltonian, which we write as (90) plus the GVD term, i.e.,

$$\begin{aligned} \textrm{H}^{\textrm{L}} & =\sum_{J}\int\textrm{d}x\,\hbar\left\{ \omega_{J}\psi_{J}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)+\frac{i}{2}v_{J}\left[\frac{\partial\psi_{J}^{{\dagger}}\left(x\right)}{\partial x}\psi_{J}\left(x\right)-\psi_{J}^{{\dagger}}\left(x\right)\frac{\partial\psi_{J}\left(x\right)}{\partial x}\right]\right.\\ & \left.\quad+\frac{v_{J}^{\prime}}{2}\frac{\partial\psi_{J}^{{\dagger}}\left(x\right)}{\partial x}\frac{\partial\psi_{J}\left(x\right)}{\partial x}+\cdots\right\}. \end{aligned}$$
Nonlinear Hamiltonians can be constructed from (142) as in Sec. 2.2. As a concrete example, we present the relevant Hamiltonian and resulting mode and coupled operator equations for the process of SP-SFWM to produce a NDSV state here, but note that the remaining five are presented in Appendix  F. Assuming three relevant modes of interest $J=\left \{P,S,I\right \}$ for pump, signal, and idler, respectively, satisfying $k_{S}+k_{I}=2k_{P}$, we find that the Hamiltonian consists of a SP-SFWM term, an SPM term for the pump, an XPM term for the pump’s effect on the signal, and an XPM term for the pump’s effect on the idler. Putting it all together we have
$$\begin{aligned} H_{\text{NDSV}}^{\text{SP-SFWM}} & ={-}\gamma_{\text{chan}}^{SIPP}\hbar^{2}\overline{\omega}_{SIPP}\overline{v}_{SIPP}^{2}\int\textrm{d}{x}\,\psi_{S}^{{\dagger}}\left(x\right)\psi_{I}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)\psi_{P}\left(x\right)+\textrm{H.c.}\\ & \quad-\frac{\gamma_{\text{chan}}^{PPPP}}{2}\hbar^{2}\omega_{P}v_{P}^{2}\int\textrm{d}{x}\,\psi_{P}^{{\dagger}}\left(x\right)\psi_{P}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)\psi_{P}\left(x\right)\\ & \quad-2\gamma_{\text{chan}}^{PSPS}\hbar^{2}\sqrt{\omega_{P}\omega_{S}}v_{P}v_{S}\int\textrm{d}{x}\,\psi_{P}^{{\dagger}}\left(x\right)\psi_{S}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)\psi_{S}\left(x\right)\\ & \quad-2\gamma_{\text{chan}}^{PIPI}\hbar^{2}\sqrt{\omega_{P}\omega_{I}}v_{P}v_{I}\int\textrm{d}{x}\,\psi_{P}^{{\dagger}}\left(x\right)\psi_{I}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)\psi_{I}\left(x\right), \end{aligned}$$
where it is convenient to introduce a general form for the coupling coefficients,
$$\gamma_{\text{chan}}^{J_{1}J_{2}J_{3}J_{4}} = \frac{3\overline{\omega}_{J_{1}J_{2}J_{3}J_{4}}\overline{\chi}_{3}}{4\epsilon_{0}\left(\overline{n}_{J_{1}}\overline{n}_{J_{2}}\overline{n}_{J_{3}}\overline{n}_{J_{4}}\right)^{1/2}c^{2}}\frac{1}{A_{\text{chan}}^{J_{1}J_{2}J_{3}J_{4}}},$$
with effective areas $A_{\text {chan}}^{J_{1}J_{2}J_{3}J_{4}}$ given by
$$\frac{1}{A_{\text{chan}}^{J_{1}J_{2}J_{3}J_{4}}}=\frac{\int\textrm{d}{y}\textrm{d}{z}\frac{\chi^{ijkl}_{3}\left(y,z\right)}{\overline{\chi}_{3}}\left[e_{J_{1}}^{i}\left(y,z\right)e_{J_{2}}^{j}\left(y,z\right)\right]^{*}e_{J_{3}}^{k}\left(y,z\right)e_{J_{4}}^{l}\left(y,z\right)}{\mathcal{N}_{J_{1}}\mathcal{N}_{J_{2}}\mathcal{N}_{J_{3}}\mathcal{N}_{J_{4}}},$$
where
$$\overline{\omega}_{J_{1}J_{2}J_{3}J_{4}}=\left(\omega_{J_{1}}\omega_{J_{2}}\omega_{J_{3}}\omega_{J_{4}}\right)^{1/4}, \quad\overline{v}_{J_{1}J_{2}J_{3}J_{4}}=\left(v_{J_{1}}v_{J_{2}}v_{J_{3}}v_{J_{4}}\right)^{1/4},$$
and
$$\mathcal{N}_{J}=\sqrt{\int\text{d}y\text{d}z\frac{n\left(y,z;\omega_{J}\right)/\overline{n}_{J}}{v_{g}\left(y,z;\omega_{J}\right)/v_{J}}\boldsymbol{e}_{J}^{*}\left(y,z\right)\cdot\boldsymbol{e}_{J}\left(y,z\right)}.$$
One can readily identify $\gamma _{\text {chan}}^{SIPP}$ as characterizing the strength of SP-SFWM, $\gamma _{\text {chan}}^{PPPP}$ as characterizing the strength of SPM of the pump, and $\gamma _{\text {chan}}^{PSPS}$ and $\gamma _{\text {chan}}^{PIPI}$ as characterizing the strength of the XPMs of the signal and the idler, respectively, by the pump. Comparing with our earlier expressions (150) and (159), we see that special cases of the coefficients $\gamma _{\text {chan}}^{J_1J_2J_3J_4}$ are related to the standard coefficient describing SPM of the pump $P$,
$$\gamma_{\text{chan}}^{\text{SPM}}=\gamma_{\text{chan}}^{PPPP},$$
and to the standard coefficient describing XPM of the signal $S$ by the pump $P$,
$$\gamma_{\text{chan}}^{\text{XPM}}=\sqrt{\frac{\omega_S}{\omega_P}}\gamma_{\text{chan}}^{PSPS}.$$

To simplify the equations resulting from the sum of the Hamiltonians $H^{\text {L}}$ and $H_{\text {NDSV}}^{\text {SP-SFWM}}$ above, we put

$$\overline{\psi}_{J}\left(x,t\right)=e^{i\omega_{J}t}\psi_{J}\left(x,t\right),$$
such that $\overline {\psi }$ is slowly varying in both time and space. With $\omega _{S}+\omega _{I}=2\omega _{P}$, we find that the Heisenberg equations of motion yield
$$\left(\frac{\partial}{\partial t}+v_{P}\frac{\partial}{\partial x}-i\frac{v_{P}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle =i\gamma_{\text{chan}}^{PPPP}\hbar\omega_{P}v_{P}^{2}\left|\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle,$$
$$\begin{aligned} \left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right) & =i\gamma_{\text{chan}}^{SIPP}\hbar\overline{\omega}_{SIPP}\overline{v}_{SIPP}^{2}\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle ^{2}\overline{\psi}_{I}^{{\dagger}}\left(x,t\right)\\ & \quad+2i\gamma_{\text{chan}}^{PSPS}\hbar\sqrt{\omega_{P}\omega_{S}}v_{P}v_{S}\left|\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{S}\left(x,t\right),\\ \left(\frac{\partial}{\partial t}+v_{I}\frac{\partial}{\partial x}-i\frac{v_{I}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{I}\left(x,t\right) & =i\gamma_{\text{chan}}^{SIPP}\hbar\overline{\omega}_{SIPP}\overline{v}_{SIPP}^{2}\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle ^{2}\overline{\psi}_{S}^{{\dagger}}\left(x,t\right)\\ & \quad+2i\gamma_{\text{chan}}^{PIPI}\hbar\sqrt{\omega_{P}\omega_{I}}v_{P}v_{I}\left|\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{I}\left(x,t\right), \end{aligned}$$
where, under the assumption of strong classical pumps, we have made the approximation
$$\psi_{P}\left(x,t\right)\rightarrow\left\langle\psi_{P}\left(x,t\right)\right\rangle.$$

With this example in front of us, we note that its coupled operator equations (312) are of the form

$$\begin{aligned} \left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right) & =i\mathcal{\tilde{S}}\left(x,t\right)\overline{\psi}_{I}^{{\dagger}}\left(x,t\right)+2i\mathcal{\tilde{M}_{S}}\left(x,t\right)\overline{\psi}_{S}\left(x,t\right),\\ \left(\frac{\partial}{\partial t}+v_{I}\frac{\partial}{\partial x}-i\frac{v_{I}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{I}\left(x,t\right) & =i\mathcal{\tilde{S}}\left(x,t\right)\overline{\psi}_{S}^{{\dagger}}\left(x,t\right)+2i\mathcal{\tilde{M}_{I}}\left(x,t\right)\overline{\psi}_{I}\left(x,t\right). \end{aligned}$$

In fact, as we show in Appendix  G, whether due to waveguide SPDC, SP-SFWM, or DP-SFWM, coupled operator equations for generating NDSV states can be placed in this general form. Similarly, as shown by Helt and Quesada [107], coupled operator equations for SPDC, SP-SFWM, or DP-SFWM generating DSV states in waveguides can be written in the form

$$\left(\frac{\partial}{\partial t}+v\frac{\partial}{\partial x}-i\frac{v^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}\left(x,t\right)=i\tilde{\mathcal{S}}\left(x,t\right)\overline{\psi}^{{\dagger}}\left(x,t\right)+2i\tilde{\mathcal{M}}\left(x,t\right)\overline{\psi}\left(x,t\right),$$

Equations for mean fields, e.g., (311), can be solved using standard approaches, such as split-step Fourier or finite difference methods [109]. The operator equations, on the other hand, can be solved using methods first sketched by Vidrighin [110] and further elaborated by Helt and Quesada [107] and Quesada et al. [108], which we detail in the following.

4.1 Solving the Coupled Operator Equations

Defining

$$\begin{aligned}\omega_{J}\left(\kappa\right) & =v_{J}\kappa+\frac{v_{J}^\prime}{2}\kappa^{2},\\ \tilde{\mathcal{S}}\left(x,t\right) & =\int\textrm{d}{\kappa}\frac{e^{i\kappa x}}{\sqrt{2\pi}}\mathcal{S}\left(\kappa,t\right),\\ \tilde{\mathcal{M}_{J}}\left(x,t\right) & =\int\textrm{d}{\kappa}\frac{e^{i\kappa x}}{\sqrt{2\pi}}\mathcal{M}_{J}\left(\kappa,t\right), \end{aligned}$$
where $\mathcal {M}_{J}\left (\kappa,t\right )=\mathcal {M}_{J}^{*}\left (-\kappa,t\right )$ and using
$$\overline{\psi}_{J}\left(x,t\right)=\int\textrm{d}{\kappa}\frac{e^{i\kappa x}}{\sqrt{2\pi}}b_{J}\left(\kappa,t\right),$$
where we have put $b_{J}\left (\kappa,t\right )=a_{J\left (k_{J}+\kappa \right )}\left (t\right )$, we can rewrite (315) and (314) as
$${\begin{aligned} \left[\frac{\partial}{\partial t}+i\omega_{S}\left(\kappa\right)\right]b_{S}\left(\kappa,t\right) & =i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{S}\left(\kappa+\kappa^{\prime},t\right)b_{I}^{{\dagger}}\left(\kappa^{\prime},t\right)+2i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{M}_{S}\left(\kappa-\kappa^{\prime},t\right)b_{S}\left(\kappa^{\prime},t\right),\\ \left[\frac{\partial}{\partial t}+i\omega_{I}\left(\kappa\right)\right]b_{I}\left(\kappa,t\right) & =i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{S}\left(\kappa+\kappa^{\prime},t\right)b_{S}^{{\dagger}}\left(\kappa^{\prime},t\right)+2i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{M}_{I}\left(\kappa-\kappa^{\prime},t\right)b_{I}\left(\kappa^{\prime},t\right), \end{aligned}}$$
and
$${\displaystyle\left[\frac{\partial}{\partial t}+i\omega\left(\kappa\right)\right]b\left(\kappa,t\right)=i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{S}\left(\kappa+\kappa^{\prime},t\right)b^{{\dagger}}\left(\kappa^{\prime},t\right)+2i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{M}\left(\kappa-\kappa^{\prime},t\right)b\left(\kappa^{\prime},t\right),}$$
respectively, forms highly amenable to numeric evaluation. We note that higher-order dispersion can be considered by including more terms in the linear Hamiltonian (302), thus resulting in more terms in $\omega _{J}\left (\kappa \right )$ above.

To solve these equations, we discretize $\kappa _{j}=j\Delta \kappa$ and write

$$\begin{aligned}\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{S}\left(\kappa+\kappa^{\prime},t\right)b_{J}^{{\dagger}}\left(\kappa^{\prime},t\right) & \approx\sum_{j^{\prime}}\sqrt{\frac{\Delta\kappa}{2\pi}}\mathcal{S}\left(\kappa_{j}+\kappa_{j^{\prime}},t\right)a_{J;j^{\prime}}^{{\dagger}}\left(t\right),\\ \int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{M}_{J}\left(\kappa-\kappa^{\prime},t\right)b_{J}\left(\kappa^{\prime},t\right) & \approx\sum_{j^{\prime}}\sqrt{\frac{\Delta\kappa}{2\pi}}\mathcal{M}_{J}\left(\kappa_{j}-\kappa_{j^{\prime}},t\right)a_{J;j^{\prime}}\left(t\right), \end{aligned}$$
where $a_{J;j}\left (t\right )=\sqrt {\Delta \kappa }b_{J}\left (\kappa _{j},t\right )$, such that $\left [a_{J;j}\left (t\right ),a^\dagger _{J^\prime ;j^\prime }\left (t\right )\right ]=\delta _{JJ^\prime }\delta _{jj^\prime }$. Introducing
$$\textbf{A}^{J_{1}J_{2}}\left(t\right)=\left(\begin{array}{cc} \boldsymbol{\Delta}_{J_{1}}\left(t\right) & \boldsymbol{\zeta}\left(t\right)\\ -\boldsymbol{\zeta}^{*}\left(t\right) & -\boldsymbol{\Delta}_{J_{2}}^{*}\left(t\right) \end{array}\right),$$
where
$$\begin{aligned}\Delta_{J;jj^{\prime}}\left(t\right) & ={-}\frac{\omega_{J}\left(\kappa_{j}\right)}{\Delta\kappa}\delta_{jj^{\prime}}+2\sqrt{\frac{\Delta\kappa}{2\pi}}\mathcal{M}_{J}\left(\kappa_{j}-\kappa_{j^{\prime}},t\right),\\ \zeta_{jj^{\prime}}\left(t\right) & =\sqrt{\frac{\Delta\kappa}{2\pi}}\mathcal{S}\left(\kappa_{j}+\kappa_{j^{\prime}},t\right), \end{aligned}$$
as well as the vector $\textbf {a}_{J}(t)$ with entries $a_{J;j}(t)$, then allows us to rewrite (318) and (319) compactly as
$$\frac{\partial}{\partial t}\left(\begin{array}{c} \textbf{a}_{S}\left(t\right)\\ \textbf{a}_{I}^{{\dagger}}\left(t\right) \end{array}\right)=i\textbf{A}^{SI}\left(t\right)\left(\begin{array}{c} \textbf{a}_{S}\left(t\right)\\ \textbf{a}_{I}^{{\dagger}}\left(t\right) \end{array}\right),$$
and
$$\frac{\partial}{\partial t}\left(\begin{array}{c} \textbf{a}\left(t\right)\\ \textbf{a}^{{\dagger}}\left(t\right) \end{array}\right)=i\textbf{A}\left(t\right)\left(\begin{array}{c} \textbf{a}\left(t\right)\\ \textbf{a}^{{\dagger}}\left(t\right) \end{array}\right),$$
respectively. Note that the matrix $\boldsymbol{\zeta }=\boldsymbol{\zeta }^{T}$ is symmetric and $\boldsymbol{\Delta }_{J}=\boldsymbol{\Delta }_{J}^{\dagger }$ is Hermitian. For a small enough propagation forward in time $\Delta t$, these have the solution [recall (209)]
$$\left(\begin{array}{c} \textbf{a}_{J_{1}}\left(t+\Delta t\right)\\ \textbf{a}_{J_{2}}^{{\dagger}}\left(t+\Delta t\right) \end{array}\right)=\textbf{K}^{J_{1}J_{2}}(t)\left(\begin{array}{c} \textbf{a}_{J_{1}}\left(t\right)\\ \textbf{a}_{J_{2}}^{{\dagger}}\left(t\right) \end{array}\right),$$
where the (single-time) infinitesimal propagator is defined as
$$\textbf{K}^{J_{1}J_{2}}(t)=\exp\left[i\Delta t\textbf{A}^{J_{1}J_{2}}\left(t\right)\right]=\left(\begin{array}{cc} \textbf{V}_{J_{1}}(t) & \textbf{W}_{J_{1}J_{2}}(t)\\ \textbf{W}_{J_{2}J_{1}}^{*}(t) & \textbf{V}_{J_{2}}^{*}(t) \end{array}\right).$$

Propagation over a finite interval can then be calculated by concatenating infinitesimal propagators $\textbf {K}^{J_{1}J_{2}}(t)$ to form the (two-argument) Heisenberg picture propagator [108]

$$\begin{aligned} \textbf{K}^{J_{1}J_{2}}(t,t_0) = \prod_{j=1}^\ell \textbf{K}^{J_{1}J_{2}}(t_j) = \left(\begin{array}{cc} \textbf{V}_{J_{1}}(t,t_0) & \textbf{W}_{J_{1}J_{2}}(t,t_0)\\ \textbf{W}_{J_{2}J_{1}}^{*}(t,t_0) & \textbf{V}_{J_{2}}^{*}(t,t_0) \end{array}\right), \end{aligned}$$
where $t_j = t_0+j \Delta t$ and $\Delta t = (t-t_0)/\ell$, as
$$\begin{aligned}\left(\begin{array}{c} \textbf{a}_{J_{1}}\left(t\right)\\ \textbf{a}_{J_{2}}^{{\dagger}}\left(t\right) \end{array}\right) = \textbf{K}^{J_{1}J_{2}}(t,t_0) \left(\begin{array}{c} \textbf{a}_{J_{1}}\left(t_{0}\right)\\ \textbf{a}_{J_{2}}^{{\dagger}}\left(t_{0}\right) \end{array}\right), \end{aligned}$$
and the product in (327) is taken in chronological order. Reverting back to continuous notation, we note that we may also write the solutions of (318) and (319) as
$$\begin{aligned}b_{S}(\kappa,t) & = \int \textrm{d}\kappa'\, \mathcal{V}_{S}(\kappa,\kappa';t,t_0) b_{S}(\kappa',t_0)+ \int \textrm{d}\kappa'\, \mathcal{W}_{SI}(\kappa,\kappa';t,t_0) b_{I}^\dagger(\kappa',t_0)\\ b_{I}^\dagger(\kappa,t) & = \int \textrm{d}\kappa'\, \mathcal{W}_{IS}^{*}(\kappa,\kappa';t,t_0) b_{S}(\kappa',t_0) + \int \textrm{d}\kappa'\, \mathcal{V}_{I}^{*}(\kappa,\kappa';t,t_0) b_{I}^\dagger(\kappa',t_0) \end{aligned}$$
and
$$b(\kappa,t) = \int \textrm{d}\kappa'\, \mathcal{V}(\kappa,\kappa';t,t_0) b(\kappa',t_0)+ \int \textrm{d}\kappa'\, \mathcal{W}(\kappa,\kappa';t,t_0) b^\dagger(\kappa',t_0),$$
where
$$\begin{aligned} \mathcal{V}_{J_{1}J_{2}}(\kappa_j,\kappa_{j'};t,t_0) & = V_{J_{1}J_{2};j,j'}(t,t_0)/\Delta \kappa,\\ \mathcal{W}_{J_{1}J_{2}}(\kappa_j,\kappa_{j'};t,t_0) & = W_{J_{1}J_{2};j,j'}(t,t_0)/\Delta \kappa. \end{aligned}$$
We note that similar input–output relations were also derived in [86,111113], albeit with different labels.

4.2 Including Loss

With these expressions known, we may now form the phase-sensitive and phase-insensitive moments (cf. Secs. 3.5 and 3.6)

$$\begin{aligned} N_{J;i,j}\left(t\right) & =\left\langle a_{J;i}^{{\dagger}}\left(t\right)a_{J;j}\left(t\right)\right\rangle,\\ M_{J_{1}J_{2};i,j}\left(t\right) & =\left\langle a_{J_{1};i}\left(t\right)a_{J_{2};j}\left(t\right)\right\rangle. \end{aligned}$$

We note that, because the generated state is Gaussian, it is completely characterized by these two moments, and thus any experimentally accessible quantity, such as coherence functions [38,99], can be written in terms of them. Following Sec. 3.8, including loss is then simply a matter of properly updating the moments. In particular, for uniform waveguide loss

$$\textbf{N}_{J}(t+\Delta t) = \eta_{J} \textbf{N}_{J}(t), \quad \textbf{M}_{J_{1}J_{2}}(t+\Delta t) = \sqrt{\eta_{J_{1}}\eta_{J_{2}}} \textbf{M}_{J_{1}J_{2}}(t),$$
where
$$\eta_{J} = \exp(-\rho_{J} \Delta t),$$
and $\rho _{J}$ can be easily obtained from the standard attenuation constant $\alpha _{J}$ [109].

4.3 Example: “Separable” JSAs

The formalism presented above is quite general in that it allows for the calculation of squeezed vacuum states in waveguides where either a second- or third-order nonlinearity is dominant, correctly accounting for dispersion to any desired order, SPM and XPM, if present, and loss, in either the so-called low- or high-gain regime. However, it is also something of a black box, in that it relies on numerics. Thus, to help build intuition as one moves from the low- to the high-gain regime, here we provide an illustrative calculation of a lossless waveguide SP-SFWM process engineered to produce a separable JSA in the low-gain regime ignoring GVD, SPM, and XPM under three scenarios: (i) ignoring time-ordering corrections to arrive at an idealized analytic result; (ii) considering leading-order time-ordering corrections via the Magnus expansion; and (iii) using the full formalism presented previously. We note that we neglect GVD, SPM, and XPM in (iii) not because the numerical formalism cannot include them, but because doing so simplifies comparisons against the analytic results.

For completeness, using (79) and (303) we write the linear and nonlinear Hamiltonians as

$$\begin{aligned}\textrm{H}^{\textrm{L}} & = \sum_{J=P,S,I}\int\textrm{d}{k}\,\hbar\omega_{Jk}a_{J}^{{\dagger}}\left(k\right)a_{J}\left(k\right) \end{aligned}$$
$$\begin{aligned}{\textrm{H}^{\textrm{NL}}} & ={-}\frac{\gamma_{\text{chan}}^{SIPP}\hbar^{2}\overline{\omega}_{SIPP}\overline{v}_{SIPP}^{2}}{4\pi^{2}}\\ & \quad\times\int\textrm{d}{k}_{1}\textrm{d}{k}_{2}\textrm{d}{k}_{3}\textrm{d}{k}_{4}\textrm{d}{x}\,e^{{-}i\left(k_{1}+k_{2}-k_{3}-k_{4}\right)x}a_{S}^{{\dagger}}\left(k_{1}\right)a_{I}^{{\dagger}}\left(k_{2}\right) a_{P}\left(k_{3}\right)a_{P}\left(k_{4}\right)+\textrm{H.c.}, \end{aligned}$$
respectively. Switching to a description in terms of frequency [114], and moving to the interaction picture, we write the corresponding interaction Hamiltonian as
$$\begin{aligned}H_{I}\left(t\right) & ={-}\frac{\gamma_{\text{chan}}^{SIPP}\hbar^{2}\overline{\omega}_{SIPP}}{4\pi^{2}}\int\textrm{d}{\omega}_{1}\textrm{d}{\omega}_{2}\textrm{d}{\omega}_{3}\textrm{d}{\omega}_{4}\int\textrm{d}{x}\,e^{i\left(\omega_{1}+\omega_{2}-\omega_{3}-\omega_{4}\right)t}\\ & \quad \times e^{{-}i\left[k_{S}\left(\omega_{1}\right)+k_{I}\left(\omega_{2}\right)-k_{P}\left(\omega_{3}\right)-k_{P}\left(\omega_{4}\right)\right]x}a_{S}^{{\dagger}}\left(\omega_{1}\right)a_{I}^{{\dagger}}\left(\omega_{2}\right)a_{P}\left(\omega_{3}\right)a_{P}\left(\omega_{4}\right)+\textrm{H.c.} \end{aligned}$$
Expanding dispersion relations
$$k_{J}\left(\omega\right)=k_{J}+\frac{\omega-\omega_{J}}{v_{J}},$$
and approximating the pump operators as classical waveforms
$$a_{P}\left(\omega\right)\rightarrow\left\langle a_{P}\left(\omega\right)\right\rangle =\sqrt{\frac{N_{P}\tau}{\left(\pi/2\right)^{1/2}}}e^{-\tau^{2}\left(\omega-\omega_{P}\right)^{2}},$$
normalized such that $\int \textrm{d}{\omega }\left \langle a^{\dagger }_{P}\left (\omega \right )a_{P}\left (\omega \right )\right \rangle = N_{P}$, we first focus on the integration over $\omega _{3}$ and $\omega _{4}$. In particular, putting
$$\omega_{{\pm}}=\omega_{3}\pm\omega_{4},$$
we find
$$\begin{aligned}& \int\textrm{d}{\omega_{3}}\textrm{d}{\omega_{4}}\,e^{{-}i\left(\omega_{3}+\omega_{4}\right)t}e^{i\frac{\omega_{3}+\omega_{4}-2\omega_{P}}{v_{P}}x}e^{-\tau^{2}\left(\omega_{3}-\omega_{P}\right)^{2}}e^{-\tau^{2}\left(\omega_{4}-\omega_{P}\right)^{2}}\\ = & \int\textrm{d}{\omega_{+}}\,e^{{-}i\omega_{+}t}e^{i\frac{\omega_{+}-2\omega_{P}}{v_{P}}x}e^{-\frac{\tau^{2}}{2}\left(\omega_{+}-2\omega_{P}\right)^{2}}\int\frac{\textrm{d}{\omega_{-}}}{2}e^{-\frac{\tau^{2}}{2}\omega_{-}^{2}}\\ = & \int\textrm{d}{\omega_{+}}\,e^{{-}i\omega_{+}t}e^{i\frac{\omega_{+}-2\omega_{P}}{v_{P}}x}e^{-\frac{\tau^{2}}{2}\left(\omega_{+}-2\omega_{P}\right)^{2}}\frac{\left(\pi/2\right)^{1/2}}{\tau}, \end{aligned}$$
and, thus, can rewrite
$$\begin{aligned} H_{I}\left(t\right) & ={-}\frac{\gamma_{\text{chan}}^{SIPP}\hbar^{2}\overline{\omega}_{SIPP}L N_{P}}{4\pi^{2}}\int\text{d}\omega_{1}\textrm{d}{\omega_{2}}\textrm{d}{\omega_{+}}e^{i\left(\omega_{1}+\omega_{2}-\omega_{+}\right)t}F\left(\omega_{1},\omega_{2},\omega_{+}\right)\\ & \quad\times e^{-\frac{\tau^{2}}{2}\left(\omega_{+}-2\omega_{P}\right)^{2}}a_{S}^{{\dagger}}\left(\omega_{1}\right)a_{I}^{{\dagger}}\left(\omega_{2}\right)+\textrm{H.c.}, \end{aligned}$$
where the phase-matching function (PMF)
$$\begin{aligned}F\left(\omega_{1},\omega_{2},\omega_{+}\right) & =\int_{{-}L/2}^{L/2}\frac{\textrm{d}{x}}{L}e^{{-}i\left(\frac{\omega_{1}-\omega_{S}}{v_{S}}+\frac{\omega_{2}-\omega_{I}}{v_{I}}-\frac{\omega_{+}-2\omega_{P}}{v_{P}}\right)x}\\ & =\text{sinc}\left[\left(\frac{\omega_{1}-\omega_{S}}{v_{S}}+\frac{\omega_{2}-\omega_{I}}{v_{I}}-\frac{\omega_{+}-2\omega_{P}}{v_{P}}\right)\frac{L}{2}\right], \end{aligned}$$
and we have taken the nonlinear region to extend from $-L/2$ to $L/2$. The associated time evolution operator for this Hamiltonian is
$$\mathcal{U}=\mathcal{T}\exp{\left(-\frac{i}{\hbar}\int_{-\infty}^{\infty} \textrm{d}{t^{\prime}}H_{I}\left(t^{\prime}\right)\right)},$$
with $\mathcal {T}$ the time-ordering operator, which can be applied to a vacuum state to yield the state of generated photons. A similar Hamiltonian can be formed for SPDC generating NDSV states in waveguides [115].

4.3.1 Low-Gain Regime

In the low-gain regime, a common, and safe, approximation is to neglect time-ordering effects and, instead, using $\int _{-\infty }^{\infty }\textrm{d}{t^{\prime }}\,e^{i\Delta t^{\prime }}=2\pi \delta \left (\Delta \right )$, write

$$\begin{aligned} \mathcal{U} & =\exp{\left(-\frac{i}{\hbar}\int_{-\infty}^{\infty} \textrm{d}{t^{\prime}}H_{I}\left(t^{\prime}\right)\right)}\\ & =\exp{\left(i\int\textrm{d}{\omega_{1}}\textrm{d}{\omega_{2}}\,J_{1;SI}\left(\omega_{1},\omega_{2}\right)a_{S}^{{\dagger}}\left(\omega_{1}\right)a_{I}^{{\dagger}}\left(\omega_{2}\right)+\textrm{H.c.}\right)}, \end{aligned}$$
where we have introduced the JSA neglecting time-ordering corrections
$$J_{1;SI}\left(\omega_{1},\omega_{2}\right)=\frac{\Phi\tau}{2\pi}F\left(\omega_{1},\omega_{2},\omega_{1}+\omega_{2}\right)e^{-\frac{\tau^{2}}{2}\left(\omega_{1}+\omega_{2}-2\omega_{P}\right)^{2}},$$
and the dimensionless quantity
$$\Phi=\gamma_{\text{chan}}^{SIPP}\frac{\hbar\overline{\omega}_{SIPP}N_{P}}{\tau}L,$$
similar to the so-called nonlinear phase shift [109]. Further approximating the phase-matching sinc function as a Gaussian via $\text {sinc}\left (x\right )\approx \exp \left (-sx^{2}\right )$, for $s\approx 0.193$, or, in fact, engineering the nonlinearity along the waveguide to produce a Gaussian PMF [116120], allows us to go further and calculate the Schmidt decomposition of $J_{1;SI}\left (\omega _{1},\omega _{2}\right )$ analytically.

Here we choose to write the Schmidt decomposition of such a JSA as

$$J_{SI}(\omega,\omega')=\sum_{\lambda} \xi_{\lambda} f_{\lambda;S}(\omega) f_{\lambda;I}(\omega'),$$
with Schmidt number
$$K=\frac{\left(\sum_{\lambda} \sinh^2\xi_\lambda\right)^2}{\sum_{\lambda} \sinh^4\xi_\lambda}.$$

Note that the Schmidt number of a separable JSA, $J_{SI}(\omega,\omega ')=\xi _0 f_{0;S}(\omega ) f_{0;I}(\omega ')$, is 1, whereas JSAs that are not separable have $K>1$. Further note that we have chosen to work with unnormalized JSAs. Although this marks a departure from more common usage, we find it useful here for comparing the results of this subsection against those that move beyond the approximations used in deriving $J_{1;SI}\left (\omega _{1},\omega _{2}\right )$. In the low-gain regime, where $K\approx \left (\sum _{\lambda } \xi ^{2}_{\lambda }\right )^2/\sum _{\lambda } \xi ^{4}_{\lambda }$, the Schmidt number is insensitive to a scaling transformation of $J_{SI}$. That is, $J_{SI}$ and $l J_{SI}$ have the same Schmidt number for any $l \neq 0$, though this does not hold in general.

Putting

$$\begin{aligned}\Omega_{1} & = \omega_{1} - \omega_{S},\\ \Omega_{2} & = \omega_{2} - \omega_{I}, \end{aligned}$$
we write
$$\begin{aligned}J_{1;SI}\left(\Omega_{1},\Omega_{2}\right) & =\frac{\Phi\tau}{2\pi}e^{{-}s\left(\Omega_{1}v_{S}^{{-}1}+\Omega_{2}v_{I}^{{-}1}-\left(\Omega_{1}+\Omega_{2}\right)v_{P}^{{-}1}\right)^{2}\frac{L^{2}}{4}}e^{-\frac{\tau^{2}}{2}\left(\Omega_{1}+\Omega_{2}\right)^{2}}\\ & = \frac{\Phi\tau}{2\pi}e^{-\Omega_{1}^{2}\left(s\left(v_{S}^{{-}1}-v_{P}^{{-}1}\right)^{2}\frac{L^{2}}{4}+\frac{\tau^{2}}{2}\right)}e^{-\Omega_{2}^{2}\left(s\left(v_{I}^{{-}1}-v_{P}^{{-}1}\right)^{2}\frac{L^{2}}{4}+\frac{\tau^{2}}{2}\right)}\\ & \quad\times e^{{-}2\Omega_{1}\Omega_{2}\left(s\left(v_{S}^{{-}1}-v_{P}^{{-}1}\right)\left(v_{I}^{{-}1}-v_{P}^{{-}1}\right)\frac{L^{2}}{4}+\frac{\tau^{2}}{2}\right)}, \end{aligned}$$
where we have used the fact that $\omega _{S}+\omega _{I}=2\omega _{P}$, and see that the JSA is separable when
$$s\left(v_{S}^{{-}1}-v_{P}^{{-}1}\right)\left(v_{I}^{{-}1}-v_{P}^{{-}1}\right)\frac{L^{2}}{4}+\frac{\tau^{2}}{2}=0.$$

We note, in passing, that the only square integrable function that is separable and has elliptical (or circular) contours is precisely a two-dimensional Gaussian, and that, moreover, only a combination of a Gaussian PMF and a Gaussian pump can give a separable JSA in the low-gain regime [121,122]. Under this condition (352), and introducing

$$\tau_{J}^{2}=s\left(v_{J}^{{-}1}-v_{P}^{{-}1}\right)^{2}\frac{L^{2}}{4}+\frac{\tau^{2}}{2},$$
we write the final form of the JSA, neglecting GVD, SPM, XPM, and time-ordering corrections, that has been engineered to be separable in the low-gain regime as
$$J_{1;SI}\left(\Omega_{1},\Omega_{2}\right)=\xi_{0}f_{0;S}\left(\Omega_{1}\right)f_{0;I}\left(\Omega_{2}\right),$$
with
$$f_{0;J}\left(\Omega\right)=\sqrt{\frac{\tau_{J}}{\left(\pi/2\right)^{1/2}}}e^{-\Omega^{2}\tau_{J}^{2}}\equiv \phi_{0;J}\left(\Omega\right),$$
the Schmidt functions corresponding to the (only nonzero) Schmidt value
$$\xi_{0}=\frac{\Phi\tau}{2\left(2\pi\tau_{S}\tau_{I}\right)^{1/2}}\equiv \overline{\xi}.$$

We note that this JSA is a two-dimensional Gaussian with elliptical contours, becoming a two-dimensional Gaussian with circular contours for $\tau _{S}=\tau _{I}=\tau$. Using (352) to show that

$$\tau_{S}^{2}+\tau_{I}^{2} = s\left(v_{S}^{{-}1}-v_{I}^{{-}1}\right)^{2}\frac{L^{2}}{4},$$
then allows interpretation of the factor
$$\sqrt{\frac{\tau^{2}}{\tau_{S}\tau_{I}}}=\sqrt{\frac{2\tau_{S}\tau_{I}}{\tau_{S}^{2}+\tau_{I}^{2}}}=\sqrt{\frac{2R}{1+R^{2}}}\le 1,$$
where the aspect ratio
$$R = \frac{\tau_{S}}{\tau_{I}},$$
as measuring the departure from circular contours.

4.3.2 Magnus Expansion

Such an analysis describes the state of generated photons well for small enough $\Phi$. However, as $\Phi$ increases, whether through an increase in $\gamma _{\text {chan}}^{SIPP}$, waveguide length $L$, or pump pulse energy via $N_{P}$, time-ordering corrections must be considered for a correct description. Using the Magnus expansion, discussed in Sec. 3.2, one can show that the leading-order corrections rewrite the JSA as [115,123]

$$J_{SI}\left(\Omega_{1},\Omega_{2}\right)\approx J_{1;SI}\left(\Omega_{1},\Omega_{2}\right) + J_{3;SI}\left(\Omega_{1},\Omega_{2}\right)- i K_{3;SI}\left(\Omega_{1},\Omega_{2}\right),$$
where the terms with subscript 3 are proportional to $\Phi ^{3}$ and the next-order corrections are proportional to $\Phi ^{5}$.

The quantities $J_{3;SI}$ and $K_{3;SI}$ have been calculated previously [115,123,124] and we can write them in closed form when (352) holds as

$$\begin{aligned} J_{3;SI}(\Omega_1,\Omega_2)-i K_{3;SI}(\Omega_1,\Omega_2) & =\overline{\xi}^3 \Bigg(\frac{\phi_{0;S}\left(\Omega_{1}\right)\phi_{0;I}\left(\Omega_{2}\right)-\phi_{1;S}\left(\Omega_{1}\right)\phi_{1;I}\left(\Omega_{2}\right)}{12}\\ & \quad -i\frac{\phi_{0;S}\left(\Omega_{1}\right)\phi_{1;I}\left(\Omega_{2}\right)-\phi_{1;S}\left(\Omega_{1}\right)\phi_{0;I}\left(\Omega_{2}\right)}{4\sqrt{3}}\Bigg), \end{aligned}$$
where we have introduced (recall (355) and see Fig. 10)
$$\phi_{1;J}\left(\Omega\right)=\sqrt{3}\phi_{0;J}\left(\Omega\right)\text{erfi}\left(\sqrt{\frac{2}{3}}\tau_{J}\Omega\right),$$
and the imaginary error function $\text {erfi}\left (z\right )=-i\text {erf}\left (iz\right )$ with $\text {erf}$ the error function, satisfying
$$\int\textrm{d}{\Omega} \ \phi_{i;J}^*(\Omega) \phi_{j;J}(\Omega)=\delta_{i,j}.$$

This allows us to write the JSA up to these corrections as

$$J_{SI}\left(\Omega_{1},\Omega_{2}\right)\approx \textbf{u}_{S}\left(\Omega_{1}\right)\textbf{L}\left[\textbf{u}_{I}\left(\Omega_{2}\right)\right]^{{\dagger}},$$
where
$$\begin{aligned}\textbf{L} & =\left(\begin{array}{cc} \overline{\xi}+\frac{\overline{\xi}^{3}}{12} & -i\frac{\overline{\xi}^{3}}{4\sqrt{3}}\\ i\frac{\overline{\xi}^{3}}{4\sqrt{3}} & -\frac{\overline{\xi}^{3}}{12} \end{array}\right)\\ & =\frac{\overline{\xi}}{2}\mathbb{I}_{2}+\frac{\overline{\xi}^{3}}{4\sqrt{3}}\sigma_{2}+\left(\frac{\overline{\xi}}{2}+\frac{\overline{\xi}^{3}}{12}\right)\sigma_{3}\\&=\frac{\overline{\xi}}{2}\left(\mathbb{I}_{2}+\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\textbf{n}\cdot{\boldsymbol{\sigma}}\right), \end{aligned}$$
with $\mathbb {I}_{2}$ the $2 \times 2$ identity matrix, $\boldsymbol{\sigma }=\left (\sigma _{1}, \sigma _{2}, \sigma _{3}\right )$, $\sigma _{i}$ the Pauli matrices, $\textbf {n}=\left (0,\sin \theta,\cos \theta \right )$, and
$$\textbf{u}_{J}\left(\Omega\right)=\left(\phi_{0;J}\left(\Omega\right),\phi_{1;J}\left(\Omega\right)\right).$$

Note that when $\Phi \ll 1$ the matrix $\textbf {L}\approx \left ( \begin {array}{cc} \overline {\xi } & 0\\ 0 & 0\\ \end {array} \right )$ and we recover the results obtained in (354). More interestingly, past this limit we can obtain the new time-ordering corrected Schmidt values and functions and verify that they are no longer linear functions of $\Phi$ and independent of $\Phi$, respectively.

 figure: Figure 10.

Figure 10. Amplitudes of the orthonormal functions $\phi _{0}(\Omega )$ and $\phi _{1}(\Omega )$ needed to expand the time-ordering corrected JSA of (364).

Download Full Size | PDF

As we have written the JSA as quadratic form (the matrix $\textbf {L}$) in the orthonormal basis $\{\phi _{i;J}\}$, obtaining the Schmidt decomposition of $J_{SI}$ is equivalent to obtaining the SVD of the matrix $\textbf {L}$. In particular, because $\textbf {L}$ is Hermitian, its SVD is related to its eigendecomposition and we can rewrite (364) as

$$J_{SI}\left(\Omega_{1},\Omega_{2}\right)\approx \textbf{u}^\prime_{S}\left(\Omega_{1}\right)\textbf{L}^\prime\left[\textbf{u}^\prime_{I}\left(\Omega_{2}\right)\right]^{{\dagger}},$$
where the diagonal matrix $\textbf {L}^\prime =\textbf {R}^{\dagger }\left (\theta \right )\textbf {L}\textbf {R}\left (\theta \right )=\left ( \begin {array}{cc} \xi _{+} & 0\\ 0 & \xi _{-}\\ \end {array} \right )$ with
$$\xi_{{\pm}}=\frac{\overline{\xi}}{2}\left(1\pm\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right),$$
and $\textbf {u}^{\prime }_{J}\left (\Omega \right )= \textbf {u}_{J}\left (\Omega \right ) \textbf {R}\left (\theta \right )$ with
$$\textbf{R}\left(\theta\right)=\left(\begin{array}{cc} \cos\left(\frac{\theta}{2}\right) & i\sin\left(\frac{\theta}{2}\right)\\ i\sin\left(\frac{\theta}{2}\right) & \cos\left(\frac{\theta}{2}\right) \end{array}\right),\quad\tan(\theta)=\frac{\overline{\xi}^2}{2\sqrt{3}(1+\overline{\xi}^2/6)}.$$

Written in this form, it is easy to see that now the Schmidt values are simply the absolute values of the eigenvalues of $\textbf {L}$ [cf. (356)], $\xi _0=|\xi _+|, \ \xi _1=|\xi _-|$ and the Schmidt functions are the eigenfunctions of $\textbf {L}$

$$\begin{aligned}f_{0;J}\left(\Omega\right) & =\cos\left(\tfrac{\theta}{2}\right)\phi_{0;J}\left(\Omega\right)+i\sin\left(\tfrac{\theta}{2}\right)\phi_{1;J}\left(\Omega\right),\\ f_{1;J}\left(\Omega\right) & =\cos\left(\tfrac{\theta}{2}\right)\phi_{1;J}\left(\Omega\right)+i\sin\left(\tfrac{\theta}{2}\right)\phi_{0;J}\left(\Omega\right). \end{aligned}$$

Thus, the JSA that started as separable in the low-gain regime does not remain so as $\Phi$ increases sufficiently.

4.3.3 Comparisons

We are now in a position to compare the average number of generated pairs per pump pulse

$$\begin{aligned}\left\langle n_{\text{pairs}}\right\rangle & ={\left\langle\textrm{vac}\right\vert}\mathcal{U}^{{\dagger}}\int\text{d}\Omega\,a_{S}^{{\dagger}}\left(\Omega\right)a_{S}\left(\Omega\right)\mathcal{U}\left\vert\text{vac}\right\rangle\\ & ={\left\langle\textrm{vac}\right\vert}\mathcal{U}^{{\dagger}}\int\text{d}\Omega\,a_{I}^{{\dagger}}\left(\Omega\right)a_{I}\left(\Omega\right)\mathcal{U}\left\vert\text{vac}\right\rangle\\ & =\sum_{\lambda}\sinh^{2}\xi_{\lambda}, \end{aligned}$$
and Schmidt number (349), for the approximations in Secs. 4.3.1 and 4.3.2, as well as the full numerical apparatus as $\Phi$ increases. In particular, neglecting time-ordering corrections we have
$$\left\langle n_{\text{pairs}}\right\rangle=\sinh^{2}\overline{\xi},$$
and
$$K=1,$$
whereas including time-ordering corrections up to $\Phi ^3$ we have
$$\left\langle n_{\text{pairs}}\right\rangle=\sinh^{2}\left[\frac{\overline{\xi}}{2}\left(1+\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right] +\sinh^{2}\left[\frac{\overline{\xi}}{2}\left(1-\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right],$$
and
$$K=1+2\frac{\sinh^{2}\left[\frac{\overline{\xi}}{2}\left(1+\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right] \sinh^{2}\left[\frac{\overline{\xi}}{2}\left(1-\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right]}{\sinh^{4} \left[\frac{\overline{\xi}}{2}\left(1+\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right] +\sinh^{4}\left[\frac{\overline{\xi}}{2}\left(1-\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right]}.$$

We plot these, as well as a full numerical solution, as functions of $\Phi$, setting $\tau _{S}=\tau _{I}=\tau$, so that

$$\overline{\xi} = \frac{\Phi}{2\left(2\pi\right)^{1/2}},$$
for convenience, in Figs. 11(a) and 11(b), respectively. Note how the time-uncorrected approximations only remain valid up until $\Phi \approx 2$ or so. We also plot the absolute values of the corresponding JSAs at $\Phi =4$ in Figs. 12(a), 12(b), and 12(c), as well as $\Phi =11$ in Figs. 12(d), 12(e), and 12(f). For large enough $\Phi$ [see Fig. 11(b)], JSAs that appear separable when time-ordering effects are neglected, as in Figs. 12(a) and 12(b), are seen to not actually be so when such effects are included.

 figure: Figure 11.

Figure 11. Comparison of the generated mean photon number per pump pulse $\langle n \rangle$ and Schmidt number $K$ for no time-ordering corrections (Magnus 1), time-ordering corrections up to $\Phi ^{3}$ (Magnus 3), and a full numerical solution.

Download Full Size | PDF

 figure: Figure 12.

Figure 12. Comparison of the absolute values of the generated JSAs $\left \vert J_{SI}(\Omega _{1},\Omega _{2})\right \vert$ with $\Phi =4$ (left column) for (a) no time-ordering corrections, (b) time-ordering corrections up to $\Phi ^{3}$, and (c) a full numerical solution, as well as $\Phi =11$ (right column) for (d) no time-ordering corrections, (e) time-ordering corrections up to $\Phi ^{3}$, and (f) a full numerical solution.

Download Full Size | PDF

5. Heisenberg Equations for Rings

In this section we build on top of Sec. 2 to continue the analysis of ring resonators in the presence of second- or third-order nonlinearities. Integrated optical resonators are some of the most efficient and successful sources of quantum states of light, thanks to the large field enhancements and the possibility to tailor independently the density of states of each mode involved in the nonlinear process, thereby controlling the statistical properties of the generated quantum light [125,126]. A wide range of integrated source geometries have been proposed and analyzed, including Fabry–Perot cavities, ring resonators, photonic crystal cavities, and multi-resonator designs, each with a variety of coupling and pumping schemes.

Here, to provide a simple and intuitive example, we restrict the discussion to a simple ring resonator point-coupled to a channel waveguide following the prescriptions of Sec. 2, and sketched in Fig. 13. Similar derivations can be carried out with more complex resonator or coupler designs [128,129], resulting in analogous conclusions.

 figure: Figure 13.

Figure 13. Schematic representation of an integrated ring resonator of radius $R$ point-coupled to a bus waveguide. The interaction is regulated by the self-coupling and cross-coupling coefficients $\sigma$ and $\kappa$, linked by $\sigma ^2+\kappa ^2=1$ [127].

Download Full Size | PDF

When this idealized structure is pumped by bright classical fields, it proves an efficient, stable, and compact source of NDSV [130] or DSV states [131,132]. In the following, we discuss the generation of such states using second- and third-order nonlinearities in the most common pumping schemes, namely SP-SFWM, DP-SFWM, and SPDC (as illustrated in Fig. 14), and present some simplified cases.

 figure: Figure 14.

Figure 14. Energy-level diagrams for the nonlinear processes considered in this section. From left to right: SP-SFWM (Sec. 5.4), SP-SFWM (Sec. 5.5), and SPDC (Sec. 5.6).

Download Full Size | PDF

5.1 SFWM Hamiltonian

Considering a ring with a third-order nonlinearity, first we can extend Sec. 2.4.2 to include SFWM, the driver of squeezing in such a ring. Starting from the form (57) of the $\boldsymbol{D}(\boldsymbol{r})$ operators, we restrict our attention to three resonant modes, with equally spaced wave vectors, conventionally labeled pump, signal, and idler, with frequencies $\omega _{\text {P}}$, $\omega _{\text {S}}$, and $\omega _{\text {I}}$, respectively. Here we have chosen to reduce the scope of this discussion to a three-resonance system, because it greatly simplifies the description of its dynamics. Indeed, with an unconstrained set of resonances, other processes, usually referred to as spurious, and that involve resonant modes beyond the three of interest here can lead to nonlinear frequency shifts and undesired parametric noise being generated completely or partly in the signal mode, affecting the quantum properties of the photons generated by SFWM [133,134]. Although the restriction to a three-resonance system is a practical expedient to simplify our discussion, it should be noted that a variety of interesting and complex quantum states can be generated across more than three resonances, and often extend over entire frequency combs [135137].

The third-order SFWM contribution to the nonlinear Hamiltonian (141) is

$$\begin{aligned} H_{\text{ring}}^{\text{SFWM}} = & -\frac{1}{4\epsilon_{0}}\left(\frac{\hbar\omega_{\text{P}}}{2\mathcal{L}}\right)\sqrt{\frac{\hbar\omega_{\text{S}}}{2\mathcal{L}}\frac{\hbar\omega_{\text{I}}}{2\mathcal{L}}}\left(\frac{4!}{2!1!1!}\right)\int \textrm{d}\boldsymbol{r}_{{\perp}}d\zeta\ \Gamma_{3}^{ijkl}(\boldsymbol{r}_{{\perp}},\zeta) \\ & \times \left(\textsf{d}_{\text{S}}^{k}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{d}_{\text{I}}^{l}(\boldsymbol{r}_{{\perp}},\zeta)\right)^{*}\textsf{d}_{\text{P}}^{i}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{d}_{\text{P}}^{j}(\boldsymbol{r}_{{\perp}},\zeta)e^{i\Delta\kappa\zeta}c_{\text{S}}^{{\dagger}}c_{\text{I}}^{{\dagger}}c_{\text{P}}c_{\text{P}} + \textrm{H.c.}, \end{aligned}$$
with $\Delta \kappa =2\kappa _{\text {P}}-\kappa _{\text {S}}-\kappa _{\text {I}}$ and we assume $\Delta \kappa =0$ for the resonances are equally spaced in $\kappa$. Similarly to the general case of SPM and XPM, cf. Eqs.  (162) and (168), and for nonlinear processes in channel waveguides, cf. Eqs. (F5), (303), and (F6), we can introduce a general form of the nonlinear coefficient $\gamma _{\text {ring}}^{J_1J_2J_3J_4}$ such that
$$H_{\text{ring}}^{\text{SFWM}}={-}\frac{\hbar^{2}\left(\omega_{\text{P}}^{2}\omega_{\text{S}}\omega_{\text{I}}\right)^{\frac{1}{4}}v_{\text{P}}\sqrt{v_{\text{S}}v_{\text{I}}}\gamma_{\text{ring}}^{\text{SIPP}}}{\mathcal{L}}c_{\text{S}}^{{\dagger}}c_{\text{I}}^{{\dagger}}c_{\text{P}}c_{\text{P}} + \textrm{H.c.},$$
where
$$\begin{aligned} \gamma_{\text{ring}}^{J_1J_2J_3J_4}= & \frac{3\left(\omega_{J_1}\omega_{J_2}\omega_{J_3}\omega_{J_4}\right)^{\frac{1}{4}}\epsilon_{0}}{4\sqrt{v_{J_1}v_{J_2}v_{J_3}v_{J_4}}}\frac{1}{\mathcal{L}}\\ & \times\int \textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta\ \chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}},\zeta)\left(\textsf{e}_{J_3}^{k}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{e}_{J_4}^{l}(\boldsymbol{r}_{{\perp}},\zeta)\right)^{*}\textsf{e}_{J_1}^{i}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{e}_{J_2}^{j}(\boldsymbol{r}_{{\perp}},\zeta). \end{aligned}$$
The expression (379), which is also useful for describing SPM and XPM in different pumping regimes, considers that $\textsf {d}_{J}(\boldsymbol{r}_{\perp },\zeta )=\epsilon _{0}n^{2}(\boldsymbol{r}_{\perp };\omega _{J})\textsf {e}_{J}(\boldsymbol{r}_{\perp },\zeta )$, and uses $n_{J}(\boldsymbol{r}_{\perp };\omega )=\sqrt {\varepsilon _{1}(\boldsymbol{r}_{\perp };\omega _{J})}$. Using the normalization condition (164) and, as usual, introducing characteristic group velocities $\overline {n}_{J}$ and $\overline {\chi }_{3}$ for the indices $n(\boldsymbol{r}_{\perp };\omega _{J})$ and the nonvanishing components of $\chi _{3}^{ijkl}(\boldsymbol{r}_{\perp },\zeta )$, the general nonlinear coefficient $\gamma _{\text {ring}}^{J_1J_2J_3J_4}$ takes the same form of (304)
$$\gamma_{\text{ring}}^{J_1J_2J_3J_4}=\frac{3\left(\omega_{J_1}\omega_{J_2}\omega_{J_3}\omega_{J_4}\right)^{\frac{1}{4}}\overline{\chi}_{3}}{4\epsilon_{0}\sqrt{\overline{n}_{J_1}\overline{n}_{J_2}\overline{n}_{J_3}\overline{n}_{J_4}}c^{2}}\frac{1}{A_{\text{ring }}^{J_1J_2J_3J_4}},$$
with the effective area $A_{\text {ring}}^{J_1J_2J_3J_4}$ defined as
$$\frac{1}{A_{\text{ring}}^{J_1J_2J_3J_4}} = \frac{\mathcal{L}^{{-}1}\int \textrm{d}\boldsymbol{r}_{{\perp}}{\rm d}\zeta\ \left(\chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}},\zeta)/\overline{\chi}_{3}\right)\left(\textsf{e}_{J_3}^{k}(\boldsymbol{r}_{{\perp}},\zeta) \textsf{e}_{J_4}^{l}(\boldsymbol{r}_{{\perp}},\zeta)\right)^{*}\textsf{e}_{J_1}^{i}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{e}_{J_2}^{j} (\boldsymbol{r}_{{\perp}},\zeta)}{\mathcal{N}_{J_{1}}\mathcal{N}_{J_{2}}\mathcal{N}_{J_{3}}\mathcal{N}_{J_{4}}},$$
with
$$\mathcal{N}_{J} = \sqrt{\int\textrm{d}\boldsymbol{r}_{{\perp}}\frac{n(\boldsymbol{r}_{{\perp}};\omega_{J})/\overline{n}_{J}}{v_{g}(\boldsymbol{r}_{{\perp}}; \omega_{J})/v_{J}}\textsf{e}_{J}^{*}(\boldsymbol{r}_{{\perp}},0)\cdot\textsf{e}_{J}(\boldsymbol{r}_{{\perp}},0)}.$$

In the expression of the effective area (381), as well as throughout the rest of the section, we assume that dispersion can be neglected within the frequency range of each ring resonance. Therefore, in the field normalization we can truncate the dispersion to the first derivative of the wave vectors. However, considering nonlinear processes that involve multiple resonances, we do not make any further assumptions on the effect of material and waveguide dispersion on the central frequencies of the resonant modes.

5.2 SPDC Hamiltonian

When the ring is characterized by a second-order nonlinear response, and we now consider the three relevant modes conventionally labeled second harmonic (SH), and fundamentals (F1 and F2) with frequencies $\omega _{\text {SH}}$, $\omega _{\text {F1}}$ and $\omega _{\text {F2}}$, respectively, the SPDC contribution to the nonlinear Hamiltonian (142) is

$${\begin{aligned}H_{\text{ring}}^{\text{SPDC}} = & -\frac{2}{3\epsilon_{0}^{2}}\Big(\frac{3!}{1!1!1!}\Big)\sqrt{\frac{\hbar\omega_{\text{SH}}}{2\mathcal{L}}}\sqrt{\frac{\hbar\omega_{\text{F1}}}{2\mathcal{L}}\frac{\hbar\omega_{\text{F2}}}{2\mathcal{L}}}\\ & \times \int\textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta\ \Gamma_{ijk}^{(2)}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{d}_{\text{SH}}^{i}(\boldsymbol{r}_{{\perp}},\zeta)\left[\textsf{d}_{\text{F1}}^{j}(\boldsymbol{r}_{{\perp}},\zeta)\right]^{*}\left[\textsf{d}_{\text{F2}}^{k}(\boldsymbol{r}_{{\perp}},\zeta)\right]^{*}e^{i\Delta\kappa\zeta}c_{\text{SH}}c_{\text{F1}}^{{\dagger}}c_{\text{F2}}^{{\dagger}} + \textrm{H.c.}\\ \equiv & -\hbar\Lambda_{\text{ring}}^{\text{SPDC}}c_{\text{SH}}c_{\text{F1}}^{{\dagger}}c_{\text{F2}}^{{\dagger}}+\textrm{H.c.}, \end{aligned}}$$
where $\Delta \kappa =\kappa _{\text {SH}}-\kappa _{\text {F1}}-\kappa _{\text {F2}}=0$ in a ring with equally spaced resonances, and we immediately jump to the definition of
$$\begin{aligned} \Lambda_{\text{ring}}^{\text{SPDC}} = & -\frac{1}{3\epsilon_{0}}\Big(\frac{3!}{1!1!1!}\Big)\sqrt{\frac{\hbar\omega_{\text{SH}}}{2\mathcal{L}}}\sqrt{\frac{\hbar\omega_{\text{F1}}}{2\mathcal{L}}}\\ & \times \sqrt{\frac{\hbar\omega_{\text{F2}}}{2\mathcal{L}}}\int \textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta\ \Gamma_{ijk}^{(2)}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{d}_{\text{SH}}^{i}(\boldsymbol{r}_{{\perp}},\zeta)\left[\textsf{d}_{\text{F1}}^{j}(\boldsymbol{r}_{{\perp}},\zeta)\right]^{*}\left[\textsf{d}_{\text{F2}}^{k}(\boldsymbol{r}_{{\perp}},\zeta)\right]^{*}, \end{aligned}$$
as the characteristic coefficient of the SPDC process in the ring.

5.3 General Expression of the Dynamic Equation

Before diving into particular scenarios, it is useful to write down the expression of the Heisenberg equation

$$i\hbar\frac{\textrm{d}\mathcal{O}(t)}{\textrm{d}t}=\left[\mathcal{O}(t),H_{\text{ring}}\right],$$
with $H_{\text {ring}}$ being the nonlinear Hamiltonian for a ring resonator point-coupled to a channel waveguide and $\mathcal {O}(t)$ any of the ring Heisenberg operators. Following (105) and (116), the full Hamiltonian is
$${\begin{aligned}H_{\text{ring}} = & H_{\text{ring}}^{\text{L}}+H_{\text{ring}}^{\text{NL}}\\ = & \sum_{J}\left(\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x)\psi_{J}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x)\frac{\partial\psi_{J}(x)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x)}{\partial x}\psi_{J}(x)\right)\textrm{d}x\right)\\ & +\sum_{J}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J}+\sum_{J}(\hbar\gamma_{J}c_{J}^{{\dagger}}\psi_{J}(0)+\textrm{H.c.})\\ & +\sum_{J}\left(\hbar\omega_{J}\int\psi_{J\text{ph}}^{{\dagger}}(x)\psi_{J\text{ph}}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J\text{ph}}\int\left(\psi_{J\text{ph}}^{{\dagger}}(x)\frac{\partial\psi_{J\text{ph}}(x)}{\partial x}-\frac{\partial\psi_{J\text{ph}}^{{\dagger}}(x)}{\partial x}\psi_{J\text{ph}}(x)\right)\textrm{d}x\right)\\ & +\sum_{J}(\hbar\gamma_{J\text{ph}}c_{J}^{{\dagger}}\psi_{J\text{ph}}(0)+\textrm{H.c.})+H_{\text{ring}}^{\text{NL}}, \end{aligned}}$$
where, of course, the expression of $H_{\text {ring}}^{\text {NL}}$ is case-specific. Using this in (385), we obtain
$$\begin{aligned} i\hbar\frac{d\mathcal{O}(t)}{dt} = & \sum_{J}\hbar\omega_{J}\left[\mathcal{O}(t),c_{J}^{{\dagger}}c_{J}\right] +\sum_{J}(\hbar\gamma_{J}\left[\mathcal{O}(t),c_{J}^{{\dagger}}\right]\psi_{J}(0)+\textrm{H.c.})\\ & +\sum_{J}(\hbar\gamma_{J\text{ph}} \left[\mathcal{O}(t),c_{J}^{{\dagger}}\right]\psi_{J\text{ph}}(0)+\textrm{H.c.})\\ & +\left[\mathcal{O}(t),H_{\text{ring}}^{\text{NL}}\left(\left\{ c_{J}(t)\right\},\left\{ c_{J}^{{\dagger}}(t)\right\};t\right)\right], \end{aligned}$$
where the otherwise implicit dependence of the nonlinear Hamiltonian on the ring operators and time is made explicit. In the limiting case of vanishing nonlinearity, the Heisenberg equation is just (41) and after solving the commutators in (387) we recover (118).

5.4 SP-SFWM

The resonant wavelengths of the idealized three-mode system introduced above are solutions of $\kappa _{m}\mathcal {L}=\frac {2\pi }{\lambda _{m}}n_{\text {P}}(\omega _{m})\mathcal{L}=2\pi m$ with $m\in \mathbb {Z}$ and are equally spaced in $\kappa$. Solutions are solely affected by the ring geometry, the material and the waveguide dispersion. Let the pump be the central resonance, surrounded by the signal and idler resonances, as sketched in Fig. 15. Although here we consider the interaction between nearest-neighbor resonances, any triple can be considered, as long as equally spaced in $\kappa$.

 figure: Figure 15.

Figure 15. Ring resonances involved in a SP-SFWM process, and the associated SPM and XPM wavelength shifts.

Download Full Size | PDF

In Sec. 2.3 we have already introduced the channel field operators (80), the formalism for the channel–ring coupling (105)–(111), and the related derivation of the full coupling coefficients $\overline {\Gamma }_{J}$ including the scattering loss modeled by a phantom channel (112), (121), and (122). It is worth recalling at this point that such an approach is only suitable for high-finesse systems (131), but this is typically verified for integrated photonic structures at frequencies below the bandgap in most common material platforms such as silicon on insulator (SOI), silicon nitride ($\text {Si}_3\text {N}_4$), lithium niobate on insulator (LNOI), aluminum gallium arsenide (AlGaAs), aluminum nitride (AlN), and chalcogenide glasses.

We illuminate the ring with a single, bright light source, where it is implicit that its bandwidth cannot exceed that of the central resonance. We refer to this case as SP-SFWM. From the full Hamiltonian of the system by composition of the linear Hamiltonian (105) and the nonlinear response due to SPM (161), XPM (168), and SFWM (377), we obtain

$${\begin{aligned} H_{\text{ring}}^{\text{SP-SFWM}} = & H_{\text{ring}}^{\text{L}}+H_{\text{ring}}^{\text{SP-SFWM,NL}}\\ = & \sum_{J\in\{\text{P,S,I}\}}\left(\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x)\psi_{J}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x)\frac{\partial\psi_{J}(x)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x)}{\partial x}\psi_{J}(x)\right)\textrm{d}x\right)\\ & + \sum_{J\in\{\text{P,S,I}\}}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J}+\sum_{J}(\hbar\gamma_{J}c_{J}^{{\dagger}}\psi_{J}(0)+\text{H.c.})\\ & + \sum_{J\in\{\text{P,S,I}\}}\left(\hbar\omega_{J}\int\psi_{J\text{ph}}^{{\dagger}}(x)\psi_{J\text{ph}}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J\text{ph}}\right.\\&\times\left.\int\left(\psi_{J\text{ph}}^{{\dagger}}(x)\frac{\partial\psi_{J\text{ph}}(x)}{\partial x}-\frac{\partial\psi_{J\text{ph}}^{{\dagger}}(x)}{\partial x}\psi_{J\text{ph}}(x)\right)\textrm{d}x\right)+\sum_{J}(\hbar\gamma_{J\text{ph}}c_{J}^{{\dagger}}\psi_{J\text{ph}}(0)+\text{H.c.})\\ & + \left(-\hbar\Lambda_{\text{ring}}^{\text{SFWM}}c_{\text{P}}c_{\text{P}}c_{\text{S}}^{{\dagger}}c_{\text{I}}^{{\dagger}}+\text{H.c.}\right)-\hbar\eta_{\text{ring}}^{\text{SPM}}c_{\text{P}}^{{\dagger}}c_{\text{P}}^{{\dagger}}c_{\text{P}}c_{\text{P}}-\hbar\zeta_{\text{ring}}^{\text{XPM}}\left(c_{\text{S}}^{{\dagger}}c_{\text{P}}^{{\dagger}}c_{\text{S}}c_{\text{P}}+c_{\text{I}}^{{\dagger}}c_{\text{P}}^{{\dagger}}c_{\text{I}}c_{\text{P}}\right), \end{aligned}}$$
where using the general definition of the nonlinear coefficient (379)
$$\begin{aligned}\Lambda_{\text{ring}}^{\text{SFWM}} & =\frac{\hbar\left(\omega_{\text{P}}^{2}\omega_{\text{S}}\omega_{\text{I}}\right)^{\frac{1}{4}}v_{\text{P}}\sqrt{v_{\text{S}}v_{\text{I}}}\gamma_{\text{ring}}^{\text{SIPP}}}{\mathcal{L}},\\ \eta_{\text{ring}}^{\text{SPM}} & =\frac{\hbar\omega_{\text{P}}v_{\text{P}}^{2}\gamma_{\text{ring}}^{\textrm{PPPP}}}{2\mathcal{L}},\\ \zeta_{\text{ring}}^{\text{XPM}} & =\frac{2\hbar\omega_{\text{P}}v_{\text{P}}v_{S(I)}\gamma_{\text{ring}}^{\text{SS(II)PP}}}{\mathcal{L}}, \end{aligned}$$
and for the last expression we allow $\omega _{\text {S}}\simeq \omega _{\text {I}}$. It is also worth pointing out that, in most practical cases, $\omega _{\text {P}}\simeq \omega _{\text {S}}\simeq \omega _{\text {I}}$ and, therefore, $\eta _{\text {ring}}^{\text {SPM}}\simeq \frac {1}{2}\Lambda _{\text {ring}}^{\text {SFWM}}$ and $\zeta _{\text {ring}}^{\text {XPM}}\simeq 2\Lambda _{\text {ring}}^{\text {SFWM}}$. From the comparison of the nonlinear Hamiltonian in (388) and the expression (188) it becomes apparent that once the pump operators are replaced by their expectation values, such a process generates a NDSV state in the signal and idler modes.

In the interest of simplifying notation, throughout the rest of this section we drop the superscripts and subscripts unless necessary. Given the full Hamiltonian (388) we can now calculate the Heisenberg equations of motion for the ring operators following (387), and obtain

\begin{align} \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{P}}+i\omega_{\text{P}}-2i\eta c_{\text{P}}^{{\dagger}}(t)c_{\text{P}}(t)\right)c_{\text{P}}(t) & ={-}i\gamma_{\text{P}}^{*}\psi_{\text{\text{P}<}}(0,t)-i\gamma_{\text{Pph}}^{*}\psi_{\text{Pph}<}(0,t)\nonumber\\&\quad+2i\Lambda^{*}c_{\text{P}}^{{\dagger}}(t)c_{\text{S}}(t)c_{\text{I}}(t),\nonumber\\ \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{S}}+i\omega_{\text{S}}-i\zeta c_{\text{P}}^{{\dagger}}(t)c_{\text{P}}(t)\right)c_{\text{S}}(t) & ={-}i\gamma_{\text{S}}^{*}\psi_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\psi_{\text{Sph<}}(0,t)\nonumber\\&\quad+i\Lambda c_{\text{P}}(t)c_{\text{P}}(t)c_{\text{I}}^{{\dagger}}(t),\nonumber\\ \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{I}}+i\omega_{\text{I}}-i\zeta c_{\text{P}}^{{\dagger}}(t)c_{\text{P}}(t)\right)c_{\text{I}}(t), & ={-}i\gamma_{\text{I}}^{*}\psi_{\text{I<}}(0,t)-i\gamma_{\text{Iph}}^{*}\psi_{\text{Iph<}}(0,t)\nonumber\\&\quad+i\Lambda c_{\text{P}}(t)c_{\text{P}}(t)c_{\text{S}}^{{\dagger}}(t), \end{align}
with damping rates $\overline {\Gamma }_{J}=\Gamma _{J}+\Gamma _{J\text {ph}}=\frac {|\gamma _{J}|^{2}}{2v_{J}}+\frac {|\gamma _{J\text {ph}}|^{2}}{2v_{J\text {ph}}}$. Note that within the approximations introduced, for each resonance the coupling constant $\gamma _{J}$ can be related to the channel–ring self-coupling coefficient $\sigma$ using (114). That self-coupling coefficient, which can be different for different resonances $J$, can be computed numerically in a straightforward way, e.g., via finite difference time domain (FDTD) simulations [138].

Let us then remove the rapidly oscillating part of the fields, by defining $\overline {c}_{J}(t)=e^{i\omega _{J}t}c_{J}(t)$ and $$\overline{\psi}_{J\lt}(0,t)=e^{i\omega_{J}t}\psi_{J\lt}(0,t)$$so that the equations of motion become

$$\begin{aligned} \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{P}}-2i\eta\overline{c}_{\text{P}}^{{\dagger}}(t)\overline{c}_{\text{P}}(t)\right)\overline{c}_{\text{P}}(t) & ={-}i\gamma_{\text{P}}^{*}\overline{\psi}_{\text{P}<}(0,t)-i\gamma_{\text{Pph}}^{*}\overline{\psi}_{\text{Pph}<}(0,t)\\&\quad+2i\Lambda^{*}\overline{c}_{\text{P}}^{{\dagger}}(t)\overline{c}_{\text{S}}(t)\overline{c}_{\text{I}}(t)e^{i\Delta_{\text{ring}}t},\\ \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{S}}-i\zeta\overline{c}_{\text{P}}^{{\dagger}}(t)\overline{c}_{\text{P}}(t)\right)\overline{c}_{\text{S}}(t) & ={-}i\gamma_{\text{S}}^{*}\overline{\psi}_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\overline{\psi}_{\text{Sph<}}(0,t)\\&\quad+i\Lambda\overline{c}_{\text{P}}(t)\overline{c}_{\text{P}}(t)\overline{c}_{\text{I}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t},\\ \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{I}}-i\zeta\overline{c}_{\text{P}}^{{\dagger}}(t)\overline{c}_{\text{P}}(t)\right)\overline{c}_{\text{I}}(t) & ={-}i\gamma_{\text{I}}^{*}\overline{\psi}_{\text{I<}}(0,t)-i\gamma_{\text{Iph}}^{*}\overline{\psi}_{\text{Iph<}}(0,t)\\&\quad+i\Lambda\overline{c}_{\text{P}}(t)\overline{c}_{\text{P}}(t)\overline{c}_{\text{S}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}, \end{aligned}$$
where $\Delta _{\text {ring}}=2\omega _{\text {P}}-\omega _{\text {S}}-\omega _{\text {I}}$. Note that commonly $\Delta _{\text {ring}}=0$ because, neglecting the waveguide GVD, the ring resonances are equally spaced in wavelength.

To solve Eqs. (391) let us restrict the discussion to the case of a coherent laser pump, thus treated classically. In this instance, the pump mode operator is replaced by its expectation value

$$\overline{c}_{\text{P}}(t)\rightarrow\left\langle \overline{c}_{\text{P}}(t)\right\rangle e^{i\Delta_{\text{P}}t}\equiv\tilde{\beta}_{\text{P}}(t)e^{i\Delta_{\text{P}}t},$$
where we allow for a possible pump detuning $\Delta _{\text {P}}$ from $\omega _{\text {P}}$. Furthermore, if we assume that the pump is undepleted, then we can safely neglect the term $\overline {c}_{\text {P}}^{\dagger }(t)\overline {c}_{\text {S}}(t)\overline {c}_{\text {I}}(t)e^{i\Delta _{\text {ring}}t}$ in the pump operator equation. The pump equation is now fully self-contained, and reads
$$\left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{P}}+i\Delta_{\text{P}}-2i\eta|\tilde{\beta}_{\text{P}}(t)|^{2}\right)\tilde{\beta}_{\text{P}}(t)={-}i\gamma_{\text{P}}^{*}\tilde{\psi}_{\text{P}<}(0,t),$$
where we defined $\tilde {\psi }_{\text {P}<}(0,t)=\overline {\psi }_{\text {P}<}(0,t)e^{-i\Delta _{\text {P}}t}$ and we have additionally ignored any input contribution form the phantom channel. Within these limits and assumptions, the equation for the pump can be solved independently. Once such solution is found, analytically or numerically, the signal and idler equations can be written
$$\begin{aligned} \left(\frac{\textrm{d}}{\textrm{d}t}+i\Delta_{\text{P}}+\overline{\Gamma}_{\text{S}}-i\zeta\left|\tilde{\beta}_{\text{P}}(t)\right|^{2}\right)\tilde{c}_{\text{S}}(t) & ={-}i\gamma_{\text{S}}^{*}\tilde{\psi}_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\tilde{\psi}_{\text{Sph<}}(0,t)\\&\quad+i\Lambda\left[\tilde{\beta}_{\text{P}}(t)\right]^{2}\tilde{c}_{\text{I}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t},\\ \left(\frac{\textrm{d}}{\textrm{d}t}+i\Delta_{\text{P}}+\overline{\Gamma}_{\text{I}}-i\zeta\left|\tilde{\beta}_{\text{P}}(t)\right|^{2}\right)\tilde{c}_{\text{I}}(t) & ={-}i\gamma_{\text{I}}^{*}\tilde{\psi}_{\text{I<}}(0,t)-i\gamma_{\text{Iph}}^{*}\tilde{\psi}_{\text{Iph<}}(0,t)\\&\quad+i\Lambda\left[\tilde{\beta}_{\text{P}}(t)\right]^{2}\tilde{c}_{\text{S}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}, \end{aligned}$$
where we have defined $\tilde {c}_{\text {S,I}}(t)=e^{-i\Delta _{\text {P}}t}\overline {c}_{\text {S,I}}(t)$, $\tilde {\psi }_{\text {S,I}<}(0,t)=e^{-i\Delta _{\text {P}}t}\overline {\psi }_{\text {S<}}(0,t)$, and $\tilde {\psi }_{\text {Sph<,Iph<}}(0,t)=e^{-i\Delta _{\text {P}}t}\overline {\psi }_{\text {Sph<}}(0,t)$.

We can reformulate the Heisenberg equations of motion (394) for a ring under SP-SFWM in matrix form as

$$\frac{\textrm{d}}{\textrm{d}t}\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{I}}^{{\dagger}}(t) \end{pmatrix}=M_{\text{SP-SFWM}}(t)\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{I}}^{{\dagger}}(t) \end{pmatrix}+D_{\text{SP-SFWM}}(t),$$
where
$$M_{\text{SP-SFWM}}(t)=\begin{pmatrix}-\overline{\Gamma}_{\text{S}}-i\Delta_{\text{P}}+i\zeta\left|\tilde{\beta}_{\text{P}}(t)\right|^{2} & i\Lambda\left[\tilde{\beta}_{\text{P}}(t)\right]^{2}e^{{-}i\Delta_{\text{ring}}t}\\ -i\Lambda^{*}\left[\tilde{\beta}_{\text{P}}^{*}(t)\right]^{2}e^{i\Delta_{\text{ring}}t} & -\overline{\Gamma}_{\text{I}}+i\Delta_{\text{P}}-i\zeta^{*}\left|\tilde{\beta}_{\text{P}}(t)\right|^{2} \end{pmatrix},$$
and
$$D_{\text{SP-SFWM}}(t)=\begin{pmatrix}-i\gamma_{\text{S}}^{*}\tilde{\psi}_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\tilde{\psi}_{\text{Sph<}}(0,t)\\ i\gamma_{\text{I}}\tilde{\psi}_{\text{I<}}^{{\dagger}}(0,t)+i\gamma_{\text{Iph}}\tilde{\psi}_{\text{Iph<}}^{{\dagger}}(0,t) \end{pmatrix}.$$

The solution of (395) is exact within the assumptions introduced, and can, in general, be found numerically. One common approach to solve (395) is to use the Green function formalism, hence identifying a matrix $G(t,t')$ that satisfies $G(t_{0},t_{0})=\mathbb {I}$ and

$$\frac{\textrm{d}}{\textrm{d}t}G(t,t')=M_{\text{SP-SFWM}}(t)G(t,t').$$

The full solution can be eventually written as

$$\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{I}}^{{\dagger}}(t) \end{pmatrix}=G(t_{0},t)\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{I}}^{{\dagger}}(t) \end{pmatrix}+\int_{-\infty}^{t}\textrm{d}t^{\prime}\ \theta(t-t')G\left(t,t^{\prime}\right)D\left(t^{\prime}\right).$$

Once the Green function has been calculated, it can be used to derive the $N$ and $M$ moments of the converted photon distribution, defined as

$$\begin{aligned} N_{\text{S(I)}}(t_{1},t_{2}) & \equiv v_{\text{S(I)}}\left\langle \tilde{\psi}_{\text{S(I)>}}^{{\dagger}}(0,t_{1})\tilde{\psi}_{\text{S(I)>}}(0,t_{2})\right\rangle \\ M_{\text{SI(IS)}}(t_{1},t_{2}) & \equiv\sqrt{v_{\text{S(I)}}v_{\text{I(S)}}}\left\langle\tilde{\psi}_{\text{S(I)>}}(0,t_{1})\tilde{\psi}_{\text{I(S)>}}(0,t_{2})\right\rangle, \end{aligned}$$
where the connection between the $\tilde {\psi }_{J>}(0,t)$ and $\tilde {\psi }_{J<}(0,t)$ operators is provided by
$$\tilde{\psi}_{J>}(0,t)=\tilde{\psi}_{J<}(0,t)-\frac{i\gamma_{J}}{v_{J}}\tilde{c}_{J}(t),$$
and thanks to (399) the expression of the $\tilde {c}_{J}(t)$ is now known. The explicit calculation of the N moment for the signal mode yields
$$\begin{aligned}N_{\text{S}}(t_{1},t_{2}) = & v_{\text{S}}\left\langle \left[\tilde{\psi}_{\text{S<}}^{{\dagger}}(0,t_{1})+\frac{i\gamma_{\text{S}}^{*}}{v_{\text{S}}}\tilde{c}_{\text{S}}^{{\dagger}}(t_{1})\right]\left[\tilde{\psi}_{\text{S<}}(0,t_{2})-\frac{i\gamma_{\text{S}}}{v_{\text{S}}}\tilde{c}_{\text{S}}(t_{2})\right]\right\rangle \\ = & v_{\text{S}}\frac{|\gamma_{\text{S}}|^{2}}{v_{\text{S}}^{2}}\int_{-\infty}^{t_{1}}\textrm{d}t'\ \int_{-\infty}^{t_{2}}\textrm{d}t^{\prime\prime}\ G_{12}^{*}\left(t_{1},t'\right)G_{12}(t_{2},t^{\prime\prime})\\ & \times\left\langle \left({-}i\gamma_{\text{I}}^{*}\tilde{\psi}_{\text{I<}}(0,t')-i\mu_{\text{I}}^{*}\tilde{\psi}_{\text{Iph<}}(0,t')\right)\left(i\gamma_{\text{I}}\tilde{\psi}_{\text{I<}}^{{\dagger}}(0,t^{\prime\prime})+i\mu_{\text{I}}\tilde{\psi}_{\text{Iph<}}^{{\dagger}}(0,t^{\prime\prime})\right)\right\rangle\\ = & 2\Gamma_{\text{S}}2\overline{\Gamma}_{\text{I}}\int_{-\infty}^{t_{1}}\textrm{d}t'\ \int_{-\infty}^{t_{2}}\textrm{d}t^{\prime\prime}\ G_{12}^{*}\left(t_{1},t'\right)G_{12}(t_{2},t^{\prime\prime})\delta(t'-t^{\prime\prime})\\ = & 4\Gamma_{\text{S}}\overline{\Gamma}_{\text{I}}\int \textrm{d}\tau\ G_{12}^{*}\left(t_{1},\tau\right)\theta(t_{1}-\tau)G_{12}(t_{2},\tau)\theta(t_{2}-\tau), \end{aligned}$$
and a similar expression holds for the idler mode. The $M_{\text {SI}}$ moment is
$$\begin{aligned} M_{\text{SI}}(t_{1},t_{2})&= \sqrt{v_{\text{S}}v_{\text{I}}}\left\langle \tilde{\psi}_{\text{S>}}(0,t_{1})\tilde{\psi}_{\text{I>}}(0,t_{2})\right\rangle = -2\sqrt{\eta_{\text{S}}\eta_{\text{I}}\overline{\Gamma}_{\text{S}}\overline{\Gamma}_{\text{I}}}\left(\overline{\Gamma}_{\text{S}}+\overline{\Gamma}_{\text{I}}\right)\\&\quad\times\int \textrm{d}\tau\ \left[G_{21}^{*}(t_{1},\tau)G_{22}^{*}(t_{2},\tau)-G_{22}^{*}(t_{1},\tau)G_{21}^{*}(t_{2},\tau)\right]\theta(t_{2}-t_{1})\theta(t_{1}-\tau)\\ &\quad +2\overline{\Gamma}_{\text{S}}\int \textrm{d}\tau\ G_{11}(t_{1},\tau)\theta(t_{1}-\tau)G_{21}^{*}(t_{2},\tau)\theta(t_{2}-\tau). \end{aligned}$$

As discussed in Sec. 3.9 the $N$ and $M$ moments are instrumental in calculating key quantities such as the average photon number of the squeezed state, its variance and covariance matrix (286).

5.5 DP-SFWM

In a variety of applications, such as quantum sensing or photonic quantum computing, it is preferable to generate DSV rather than NDSV states; yet, the SP approach described in the previous section (where one neglects the bright squeezed state generated in the P mode), is not suitable for the task. Nonetheless, with the same high-finesse resonator and still restricting our discussion to three isolated modes, one can invert the pumping scheme and readily obtain a DSV state. Let us illuminate the ring with two individual pumps, named $\text {P1}$ and $\text {P2}$, addressing the side resonances as prescribed in the previous section and represented in Fig. 16.

 figure: Figure 16.

Figure 16. Ring resonances involved in a DP-SFWM process, and the associated SPM and XPM wavelength shifts.

Download Full Size | PDF

Incidentally, we note that a two-lobed pump where each lobe is in correspondence of a ring resonance is a completely equivalent view of the pumping scheme. We can conveniently rename the three ring resonances as P1, S, and P2, respectively, where it is assumed that the DSV state is generated in the S resonance by SFWM. We refer to this pumping scheme as DP-SFWM.

The full Hamiltonian of the process is

\begin{align} H_{\text{ring}}^{\text{DP-SFWM}} & =H_{\text{ring}}^{\text{L}}+H_{\text{ring}}^{\text{DP-SFWM},\text{NL}}\nonumber\\ &= \sum_{J\in\{\text{S,P1,P2}\}}\left(\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x)\psi_{J}(x)\textrm{d}x\right.\nonumber\\&\quad\left.-\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x)\frac{\partial\psi_{J}(x)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x)}{\partial x}\psi_{J}(x)\right)\textrm{d}x\right)\nonumber\\ & +\sum_{J\in\{\text{S,P1,P2}\}}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J}+\sum_{J\in\{\text{S,P1,P2}\}}(\hbar\gamma_{J}c_{J}^{{\dagger}}\psi_{J}(0)+\text{H.c.})\nonumber\\ & + \sum_{J\in\{\text{S,P1,P2}\}}\left(\hbar\omega_{J}\int\psi_{J\text{ph}}^{{\dagger}}(x)\psi_{J\text{ph}}(x)\textrm{d}x\right.\nonumber\\&\quad\left.-\frac{1}{2}i\hbar v_{J\text{ph}}\int\left(\psi_{J\text{ph}}^{{\dagger}}(x)\frac{\partial\psi_{J\text{ph}}(x)}{\partial x}-\frac{\partial\psi_{J\text{ph}}^{{\dagger}}(x)}{\partial x}\psi_{J\text{ph}}(x)\right)\textrm{d}x\right)\nonumber\\ & +\sum_{J\in\{\text{S,P1,P2}\}}(\hbar\gamma_{J\text{ph}}c_{J}^{{\dagger}}\psi_{J\text{ph}}(0)+\text{H.c.})\nonumber\\ & -\left(\hbar\Lambda_{\text{ring}}^{\text{DP-SFWM}}c_{\text{P1}}c_{\text{P2}}c_{\text{S}}^{{\dagger}}c_{\text{S}}^{{\dagger}}+\text{H.c.}\right)-\hbar\eta_{\text{ring}}^{\text{SPM}}(c_{\text{P1}}^{{\dagger}}c_{\text{P1}}^{{\dagger}}c_{\text{P1}}c_{\text{P1}}+c_{\text{P2}}^{{\dagger}}c_{\text{P2}}^{{\dagger}}c_{\text{P2}}c_{\text{P2}})\nonumber\\ & -\hbar\zeta_{\text{ring}}^{\text{XPM}}\left(c_{\text{S}}^{{\dagger}}c_{\text{P1}}^{{\dagger}}c_{\text{S}}c_{\text{P1}}+c_{\text{S}}^{{\dagger}} c_{\text{P2}}^{{\dagger}}c_{\text{S}}c_{\text{P2}}+c_{\text{P1}}^{{\dagger}}c_{\text{P2}}^{{\dagger}}c_{\text{P1}}c_{\text{P2}}\right), \end{align}
where, now, together with the SFWM term, we have two SPM terms, one for each pump, and three XPM terms, between each pump and the signal mode, and between the two pumps. As one can now appreciate from the comparison of (404) and (174), DP-SFWM leads to the generation of a DSV state in the signal resonance.

As before, in the following we drop the unnecessary subscripts for the sake of clarity. We can proceed with the same reasoning as in the SP-SFWM case and work out the Heisenberg equations of motion, to obtain

$$\begin{aligned} & \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{P1}}-2i\eta\overline{c}_{\text{P1}}^{{\dagger}}(t)\overline{c}_{\text{P1}}(t)-i\zeta\overline{c}_{\text{P2}}^{{\dagger}}(t)\overline{c}_{\text{P2}}(t)\right)\overline{c}_{\text{P1}}(t) \\ & \quad \quad =-i\gamma_{\text{P1}}^{*}\overline{\psi}_{\text{P1<}}(0,t)-i\gamma_{\text{P1ph}}^{*}\overline{\psi}_{\text{P1ph<}}(0,t)+i\Lambda^{*}\overline{c}_{\text{P2}}^{{\dagger}}(t)\overline{c}_{\text{S}}(t)\overline{c}_{\text{S}}(t)e^{i\Delta_{\text{ring}}t},\\ & \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{P2}}-2i\eta\overline{c}_{\text{P2}}^{{\dagger}}(t)\overline{c}_{\text{P2}}(t)-i\zeta\overline{c}_{\text{P1}}^{{\dagger}}(t)\overline{c}_{\text{P1}}(t)\right)\overline{c}_{\text{P2}}(t) \\ & \quad \quad=-i\gamma_{\text{P2}}^{*}\overline{\psi}_{\text{P2<}}(0,t)-i\gamma_{\text{P2ph}}^{*}\overline{\psi}_{\text{P2ph<}}(0,t)+i\Lambda^{*}\overline{c}_{\text{P1}}^{{\dagger}}(t)\overline{c}_{\text{S}}(t)\overline{c}_{\text{S}}(t)e^{i\Delta_{\text{ring}}t},\\ & \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{S}}-i\zeta\overline{c}_{\text{P1}}^{{\dagger}}(t)\overline{c}_{\text{P1}}(t)-i\zeta\overline{c}_{\text{P2}}^{{\dagger}}(t)\overline{c}_{\text{P2}}(t)\right)\overline{c}_{\text{S}}(t) \\ & \quad \quad=-i\gamma_{\text{S}}^{*}\overline{\psi}_{\text{S<}}(0,t)-i\gamma_{S\text{ph}}^{*}\overline{\psi}_{\text{Sph<}}(0,t)+2i\Lambda\overline{c}_{\text{P1}}(t)\overline{c}_{\text{P2}}(t)\overline{c}_{\text{S}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}, \end{aligned}$$
where, now, $\Delta _{\text {ring}}=\omega _{\text {P1}}+\omega _{\text {P2}}-2\omega _{\text {S}}$.

As usual, we assume the pumps to be coherent lasers, thus treated classically. Each pump mode operator is then replaced by its expectation value

$$\overline{c}_{\text{P1(P2)}}(t)\rightarrow\left\langle \overline{c}_{\text{P1(P2)}}(t)\right\rangle e^{i\Delta_{\text{P1(P2)}}t}\equiv\tilde{\beta}_{\text{P1(P2)}}(t)e^{i\Delta_{\text{P1(P2)}}t},$$
once more allowing for optional pump detunings $\Delta _{\text {P1(P2)}}$. With undepleted pumps, neglecting the $\overline {c}_{\text {P1(P2)}}^{\dagger }(t)\overline {c}_{\text {S}}(t)\overline {c}_{\text {S}}(t)e^{i\Delta _{\text {ring}}t}$ terms, we have
$$\begin{aligned}\left(\frac{\textrm{d}}{\textrm{d}t}+i\Delta_{\text{P1}}+\overline{\Gamma}_{\text{P1}}-2i\eta\left|\tilde{\beta}_{\text{P1}}(t)\right|^{2}-i\zeta\left|\tilde{\beta}_{\text{P2}}(t)\right|^{2}\right)\tilde{\beta}_{\text{P1}}(t) & ={-}i\gamma_{\text{P1}}^{*}\tilde{\psi}_{\text{P1<}}(0,t),\\ \left(\frac{\textrm{d}}{\textrm{d}t}+i\Delta_{\text{P2}}+\overline{\Gamma}_{\text{P2}}-2i\eta\left|\tilde{\beta}_{\text{P2}}(t)\right|^{2}-i\zeta\left|\tilde{\beta}_{\text{P1}}(t)\right|^{2}\right)\overline{\beta}_{\text{P2}}(t) & ={-}i\gamma_{\text{P2}}^{*}\tilde{\psi}_{\text{P2<}}(0,t), \end{aligned}$$
where $\tilde {\psi }_{\text {P1(P2)<}}(0,t)=\overline {\psi }_{\text {P1(P2)<}}(0,t)e^{-i\Delta _{\text {P1(P2)}}t}$, and we ignored the input contribution of the phantom channel. The dynamic equation of the signal mode is now
$$\begin{aligned} & \left(\frac{\textrm{d}}{\textrm{d}t}+i\frac{\Delta_{\text{P1}}+\Delta_{\text{P2}}}{2}+\overline{\Gamma}_{\text{S}}-i\zeta\left|\tilde{\beta}_{\text{P1}}(t)\right|^{2}-i\zeta\left|\tilde{\beta}_{\text{P2}}(t)\right|^{2}\right)\tilde{c}_{\text{S}}(t) \\ & \quad \quad =-i\gamma_{\text{S}}^{*}\tilde{\psi}_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\tilde{\psi}_{\text{Sph<}}(0,t)+2i\Lambda\tilde{\beta}_{\text{P1}}(t)\tilde{\beta}_{\text{P2}}(t)\tilde{b}_{\text{S}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}, \end{aligned}$$
where, as for the pump equations, we defined $\tilde {c}_{\text {S}}(t)=e^{-i\frac {\Delta _{\text {P1}}+\Delta _{\text {P2}}}{2}t}\overline {c}_{\text {S}}(t)$, $\tilde {\psi }_{\text {S<}}(0,t)=\overline {\psi }_{\text {S<}}(0,t)e^{-i\frac {\Delta _{\text {P1}}+\Delta _{\text {P2}}}{2}t}$, and $\tilde {\psi }_{\text {Sph<}}(0,t)=\overline {\psi }_{\text {Sph<}}(0,t)e^{-i\frac {\Delta _{\text {P1}}+\Delta _{\text {P2}}}{2}t}$.

The analog of the Heisenberg equations of motion (395) for DP-SFWM is

$$\frac{\textrm{d}}{\textrm{d}t}\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{S}}^{{\dagger}}(t) \end{pmatrix} = M_{\text{DP-SFWM}}(t)\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{S}}^{{\dagger}}(t) \end{pmatrix}+D_{\text{DP-SFWM}}(t),$$
but now
$$\begin{aligned} & M_{\text{DP-SFWM}}(t)= -\overline{\Gamma}_{\text{S}}\mathbb{I} \\ &\quad + \begin{pmatrix} -i\frac{\Delta_{\text{P1}}+\Delta_{\text{P2}}}{2}+i\zeta\left[\left|\tilde{\beta}_{\text{P1}}(t)\right|^{2}+\left|\tilde{\beta}_{\text{P2}}(t)\right|^{2}\right] & 2i\Lambda\tilde{\beta}_{\text{P1}}(t)\tilde{\beta}_{\text{P2}}(t)e^{{-}i\Delta_{\text{ring}}t}\\ -2i\Lambda^{*}\tilde{\beta}_{\text{P1}}^{*}(t)\tilde{\beta}_{\text{P2}}^{*}(t)e^{i\Delta_{\text{ring}}t} & \begin{array}{c}\frac{\Delta_{\text{P1}}+\Delta_{\text{P2}}}{2}-i\zeta^{*}\left[\left|\tilde{\beta}_{\text{P1}}(t)\right|^{2}\right.\left.+\left|\tilde{\beta}_{\text{P2}}(t)\right|^{2}\right]\end{array}\end{pmatrix}, \end{aligned}$$
and
$$D_{\text{DP-SFWM}}(t)=\begin{pmatrix}-i\gamma_{\text{S}}^{*}\tilde{\psi}_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\tilde{\psi}_{\text{Sph<}}(0,t)\\ i\gamma_{\text{S}}\tilde{\psi}_{\text{S<}}^{{\dagger}}(0,t)+i\gamma_{\text{Sph}}\tilde{\psi}_{\text{Sph<}}^{{\dagger}}(0,t) \end{pmatrix}.$$

We can finally calculate the $N$ and $M$ moments of the photon distribution as

$$\begin{aligned}N_{\text{S}}(t_{1},t_{2}) & \equiv v_{\text{S}}\left\langle\tilde{\psi}_{\text{S>}}^{{\dagger}}(0,t_{1})\tilde{\psi}_{\text{S>}}(0,t_{2})\right\rangle, \\ M_{\text{SS}}(t_{1},t_{2}) & \equiv\sqrt{v_{\text{S}}v_{\text{S}}}\left\langle\tilde{\psi}_{\text{S>}}(0,t_{1})\tilde{\psi}_{\text{S>}}(0,t_{2})\right\rangle, \end{aligned}$$
and the output fields are connected to the ring operators by (401). Having solved the Green function equation, the moments can be expressed as
$$\begin{aligned} N_{\text{S}}(t_{1},t_{2}) & =2\Gamma_{\text{S}}2\overline{\Gamma}_{\text{S}}\int_{-\infty}^{t_{1}}\textrm{d}t'\ \int_{-\infty}^{t_{2}}\textrm{d}t^{\prime\prime}\ G_{21}\left(t_{1},t'\right)G_{12}(t_{2},t^{\prime\prime})\delta(t'-t^{\prime\prime})\\ & =4\Gamma_{\text{S}}\overline{\Gamma}_{\text{S}}\int \textrm{d}\tau\ G_{21}\left(t_{1},\tau\right)\theta(t_{1}-\tau)G_{12}(t_{2},\tau)\theta(t_{2}-\tau), \end{aligned}$$
and
$$\begin{aligned} M_{\text{SS}}(t_{1},t_{2})&= v_{\text{S}}\left[\frac{\gamma_S^2}{v_S^2}\int_{-\infty}^{t_2}\textrm{d}t^{\prime\prime}\ G_{12}(t_2,t^{\prime\prime})\delta(t_1-t^{\prime\prime})\right. \\ & \quad -\left.\frac{\gamma_{\text{S}}^{2}}{v_{\text{S}}^{2}}\int_{-\infty}^{t_{1}}\textrm{d}t'\ \int_{-\infty}^{t_{2}}\textrm{d}t^{\prime\prime}\ G_{11}(t_{1},t')G_{12}(t_{2},t^{\prime\prime})\delta(t'-t^{\prime\prime})\left[\frac{|\gamma_{\text{S}}|^{2}}{v_{\text{S}}}+\frac{|\mu_{\text{S}}|^{2}}{v_{\text{S}}}\right]\right]\\ &= -4\frac{\gamma_{\text{S}}^{2}}{2v_{\text{S}}}\overline{\Gamma}_{\text{S}}\int \textrm{d}\tau\ \left[G_{12}(t_{1},\tau)G_{11}(t_{2},\tau)\theta(t_{1}-\tau)\theta(t_{2}-t_{1})\theta(t_{2}-\tau)\right.\\ &\quad +\left.G_{11}(t_{1},\tau)G_{12}(t_{2},\tau)\theta(t_{2}-\tau)\theta(t_{1}-t_{2})\theta(t_{1}-\tau)\right]. \end{aligned}$$

5.5.1 Dynamics of the Cavity and Field Operators for DP-SFWM

So far, our main goal was to express the Heisenberg equations of motion in an analytic form, with little insight on how one could solve them and obtain the full dynamics of the cavity. That, in turn, allows one to deduce relevant physical quantities such as the squeezing level, the average photon numbers in the converted modes, the Schmidt number or the distribution of the Schmidt modes. Unsurprisingly, a general analytical solution of those equations cannot be provided. The pumping scheme adopted, the amplitude and phase profile of the pump(s), the strength of the nonlinear terms involved and the resonances detunings are some of the key parameters that affect the solution in a nontrivial way.

Yet, there exist some particular simplified scenarios where an analytical solution can indeed be written in a closed form. For simplicity, here we focus on the DP-SFWM case, but one can get to similar conclusions for SP-SFWM and nondegenerate SPDC. Furthermore, we imagine that XPM from the pump modes to the signal mode can be neglected and that the pumps are on-resonance (hence, $\Delta _{P1(P2)}=0$). Under these conditions we can rephrase (410) as

$$\begin{aligned}M_{\text{DP-SFWM}}(t) & ={-}\overline{\Gamma}_{\text{S}}\mathcal{\mathbb{I}}_{2}+B(t), \end{aligned}$$
$$\begin{aligned}B(t)= & \left[\begin{array}{lr} 0 & f(t)\\ f^{*}(t) & 0 \end{array}\right], \end{aligned}$$
where $f(t)=2i\Lambda \tilde {\beta }_{\text {P1}}(t)\tilde {\beta }_{\text {P2}}(t)e^{-i\Delta _{\text {ring}}t}$. Ignoring any additional contribution to and from the phantom channel (hence, working with a lossless system), we can express the solution using the Green function formalism as
$$\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{S}}^{{\dagger}}(t) \end{pmatrix}=G(t_{0},t)\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{S}}^{{\dagger}}(t) \end{pmatrix}+\int_{-\infty}^{t}\textrm{d}t^{\prime}\ \theta(t-t')G\left(t,t^{\prime}\right)\begin{pmatrix}-i\gamma_{\text{S}}^{*}\tilde{\psi}_{\text{S<}}(0,t^{\prime})\\ i\gamma_{\text{S}}\tilde{\psi}_{\text{S<}}^{{\dagger}}(0,t^{\prime}) \end{pmatrix}$$
with $G(t_{0},t_{0})=\mathbb {I}_{2}$. The general solution of (417) is
$$\begin{aligned}G\left(t,t^{\prime}\right) & =e^{-\overline{\Gamma}_{\text{S}}\left(t-t^{\prime}\right)}S\left(t,t^{\prime}\right),\\ S\left(t,t^{\prime}\right) & =\mathcal{T}\exp\left(\int_{t^{\prime}}^{t}\textrm{d}t^{\prime\prime}B\left(t^{\prime\prime}\right)\right)\in SU(1,1), \end{aligned}$$
where $\mathcal {T}$ is the time-ordering operator. We can now make an assumption on the temporal profile of the pump in the ring, and take it as real and sufficiently short compared with the photon dwelling time in the cavity, such that $f(t)=f^{*}(t)\simeq \mu \delta (t)$. Under these conditions, the time-ordering correction is superfluous and the solution becomes
$$S\left(t,t^{\prime}\right)=\begin{pmatrix}\cosh\left(\mu\left[\theta(t)-\theta\left(t^{\prime}\right)\right]\right) & \sinh\left(\mu\left[\theta(t)-\theta\left(t^{\prime}\right)\right]\right)\\ \sinh\left(\mu\left[\theta(t)-\theta\left(t^{\prime}\right)\right]\right) & \cosh\left(\mu\left[\theta(t)-\theta\left(t^{\prime}\right)\right]\right) \end{pmatrix}.$$

Finally, using (110) we obtain

$$\begin{aligned}\psi_{>}(t) = & \psi_{<}(t)-2\overline{\Gamma}_{\text{S}}\int_{-\infty}^{\infty}\textrm{d}t^{\prime}\theta\left(t-t^{\prime}\right)e^{-\overline{\Gamma}_{\text{S}}\left(t-t^{\prime}\right)}S_{1,1}\left(t,t^{\prime}\right)\psi_{\text{in }}\left(t^{\prime}\right)\\ & +2\overline{\Gamma}_{\text{S}}\int_{-\infty}^{\infty}\textrm{d}t^{\prime}\theta\left(t-t^{\prime}\right)e^{-\overline{\Gamma}_{\text{S}}\left(t-t^{\prime}\right)}S_{1,2}\left(t,t^{\prime}\right)\psi_{\text{in }}^{{\dagger}}\left(t^{\prime}\right),\\ = & \int \textrm{d}t^{\prime}\left[\delta\left(t-t^{\prime}\right)-\left(2\overline{\Gamma}_{\text{S}}\theta\left(t-t^{\prime}\right)e^{-\overline{\Gamma}_{\text{S}}\left(t-t^{\prime}\right)}+[\cosh\mu-1]f_{0}(t)f_{0}\left({-}t^{\prime}\right)\right)\right]\psi_{\text{in }}\left(t^{\prime}\right)\\ & +\int \textrm{d}t^{\prime}\sinh\mu f_{0}(t)f_{0}({-}t)\psi_{\text{in }}^{{\dagger}}\left(t^{\prime}\right), \end{aligned}$$
where (see Fig. 17)
$$f_{0}(t)=\sqrt{2\overline{\Gamma}_{\text{S}}}e^{-\overline{\Gamma}_{\text{S}}t}\theta(t).$$

Thus, in the idealized case of infinitely short pump pulses and neglecting both loss and XPM, we can analytically express the input–output relation for the squeezed mode and assess the single-Schmidt mode nature of the generated state [128,129].

 figure: Figure 17.

Figure 17. Amplitude of the “ringdown” function $f_{0}(t)$ in (421). Note that we make explicit the fact that the Heaviside function takes the value 1/2 when evaluated at 0.

Download Full Size | PDF

5.5.2 Example: Squeezing Spectrum of a Nanophotonic Molecule

To illustrate the practical applications of the formalism presented in this section, we now consider the real-case scenario of Ref. [133]. The integrated resonator the authors propose is a “photonic molecule” made of two coupled rings, where the additional cavity serves the purpose of suppressing spurious nonlinear processes, allowing for an effective three-mode modeling of the squeezer, precisely as we have so far assumed in all the previous derivations. In this particular instance, the rings are fabricated in silicon nitride and have a $1.5\times 0.8\ \mu {\rm m}^2$ cross section. From the field distribution calculated by eigenmode simulation and the other geometrical parameters reported in Ref. [133] one can extract a nonlinear parameter $\Lambda \sim 5\ {\rm Hz}$. The Hamiltonian of the system is the same as (404), and therefore the Heisenberg equation of motion for the signal mode is (408).

In Ref. [133], the cavity is pumped with two quasi-CW lasers (pulses of constant peak power with duration much greater than the cavity lifetime) of equal intensity, and the squeezing spectrum is measured by balanced homodyne detection [3,139], that consists of interfering the signal with a LO on a 50–50 splitter and recording the difference photocurrent while varying the phase of the LO. Once the $N$ and $M$ moments are calculated as in Eqs. (413) and (414), it is more convenient to work in the frequency domain by taking their Fourier transform $N_{\text {S}}(\Omega, \Omega ')$ and $M_{\text {SS}}(\Omega, \Omega ')$ (given the quasi-CW nature of the pumps). Then, the squeezing spectrum can be expressed as

$$S(\Omega) = 1 + N_{\text{S}}(\Omega, \Omega) + N_{\text{S}}(-\Omega, -\Omega) + 2\Re\big[e^{{-}2i\psi_{\text{LOph}}} M_{\text{SS}}(\Omega, -\Omega)\big],$$
with $\psi _{\text {LOph}}$ the LO phase. This expression then allows to predict the squeezing spectra as a function of pumping (and cavity) parameters, such as the pump power and sideband frequency. We report in Fig. 18 the same conclusions from Ref. [133], where the experimental values are compared to the fitting following (422).

 figure: Figure 18.

Figure 18. Squeezing and anti-squeezing spectra (blue and black, respectively) as a function of (a) the pump power and (b) the sideband frequency, according to (422). Reprinted by permission from Macmillan Publishers Ltd: Zhang et al., Nature Commun. 12, 2233 (2021) [133]. Copyright (2021).

Download Full Size | PDF

5.6 Parametric Downconversion

We can consider one last time the ring introduced previously, and focus on the second-order nonlinear response, specifically on SPDC. To this end, we can select three resonances conventionally labeled SH, F1, and F2, where the energy of the former approximately equals the sum of the energies of the latter, as depicted in Fig. 19. We note that, in contrast to the SFWM processes, it is more demanding to guarantee energy and phase matching for SPDC processes, given the large frequency difference between the second harmonic and fundamental modes and the optical dispersion of commonly used $\chi _2$ materials. However, several techniques to circumvent this issue have been developed, including quasi-phase matching via waveguide periodical poling [140142], modal phase matching to higher-order modes (of either same or opposite polarization) [143,144], and cyclic phase matching [145147].

 figure: Figure 19.

Figure 19. Ring resonances involved in a nondegenerate SPDC process.

Download Full Size | PDF

Pumping the SH resonance with a bright field, the Hamiltonian governing the system becomes

\begin{align} H_{\text{ring}}^{\text{SPDC}} &= H_{\text{ring}}^{\text{L}}+H_{\text{ring}}^{\text{SPDC},\text{NL}}\nonumber\\ &= \sum_{J\in\{\text{F1,F2,SH}\}}\left(\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x)\psi_{J}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x)\frac{\partial\psi_{J}(x)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x)}{\partial x}\psi_{J}(x)\right)\textrm{d}x\right)\nonumber\\ &\quad +\sum_{J\in\{\text{F1,F2,SH}\}}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J}+\sum_{J\in\{\text{F1,F2,SH}\}}(\hbar\gamma_{J}c_{J}^{{\dagger}}\psi_{J}(0)+\text{H.c.})\nonumber\\ &\quad +\sum_{J\in\{\text{F1,F2,SH}\}}\left(\hbar\omega_{J}\int\psi_{J\text{ph}}^{{\dagger}}(x)\psi_{J\text{ph}}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J\text{ph}}\right.\nonumber\\&\quad\times\left.\int\left(\psi_{J\text{ph}}^{{\dagger}}(x)\frac{\partial\psi_{J\text{ph}}(x)}{\partial x}-\frac{\partial\psi_{J\text{ph}}^{{\dagger}}(x)}{\partial x}\psi_{J\text{ph}}(x)\right)\textrm{d}x\right)+\sum_{J\in\{\text{F1,F2,SH}\}}(\hbar\gamma_{J\text{ph}}c_{J}^{{\dagger}}\psi_{J\text{ph}}(0)+\text{H.c.})\nonumber\\\ &\quad -\hbar\Lambda_{\text{ring}}^{\text{SPDC}}c_{\text{SH}}c_{\text{F1}}^{{\dagger}}c_{\text{F2}}^{{\dagger}}+\text{H.c.,} \end{align}
where $\Lambda _{\text {ring}}^{\text {SPDC}}$ is given by (384) and the only nonlinear term is responsible for the destruction of a photon in the pump resonance and the correspondent creation of two photons in the fundamental resonances. As for the SFWM cases, we can drop the unnecessary notation, and express the equations of motion as
$$\begin{aligned}\Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{SH}}+i\omega_{\text{SH}}\Big)c_{\text{SH}}(t) & =i\Lambda^{*}c_{\text{F1}}(t)c_{\text{F2}}(t)-i\gamma_{\text{SH}}^{*}\psi_{SH<}\left(0,t\right)-i\gamma_{\text{SHph}}^{*}\psi_{SHph<}(0,t),\\ \Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{F1}}+i\omega_{\text{F1}}\Big)c_{\text{F1}}(t) & =i\Lambda c_{\text{SH}}(t)c_{\text{F2}}^{{\dagger}}(t)-i\gamma_{\text{F1}}^{*}\psi_{F1<}(0,t)-i\gamma_{\text{F1ph}}^{*}\psi_{F1ph<}(0,t),\\ \Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{F2}}+i\omega_{\text{F2}}\Big)c_{\text{F2}}(t) & =i\Lambda c_{\text{SH}}(t)c_{\text{F1}}^{{\dagger}}(t)-i\gamma_{\text{F2}}^{*}\psi_{F2<}(0,t)-i\gamma_{\text{F2ph}}^{*}\psi_{F2ph<}(0,t), \end{aligned}$$
and in terms of the slowly varying operators $\overline {c}_{J}(t)=e^{i\omega _{J}t}c_{J}(t)$ and $\overline{\psi}_{J\lt}(0,t)=e^{i\omega_{J}t}\psi_{J\lt}(0,t)$,
$$\begin{aligned} \Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{SH}}\Big)\overline{c}_{\text{SH}}(t) & =i\Lambda^{*}\overline{c}_{\text{F1}}(t)\overline{c}_{\text{F2}}(t)e^{i\Delta_{\text{ring}}t}-i\gamma_{\text{SH}}^{*}\overline{\psi}_{SH<}\left(0,t\right)-i\gamma_{\text{SHph}}^{*}\overline{\psi}_{SHph<}(0,t),\\ \Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{F1}}\Big)\overline{c}_{\text{F1}}(t) & =i\Lambda\overline{c}_{\text{SH}}(t)\overline{c}_{\text{F2}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}-i\gamma_{\text{F1}}^{*}\overline{\psi}_{F1<}(0,t)-i\gamma_{\text{F1ph}}^{*}\overline{\psi}_{F1ph<}(0,t),\\ \Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{F2}}\Big)\overline{c}_{\text{F2}}(t) & =i\Lambda\overline{c}_{\text{SH}}(t)\overline{c}_{\text{F1}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}-i\gamma_{\text{F2}}^{*}\overline{\psi}_{F2<}(0,t)-i\gamma_{\text{F2ph}}^{*}\overline{\psi}_{F2ph<}(0,t), \end{aligned}$$
with now $\Delta _{\text {ring}}=\omega _{\text {SH}}-\omega _{\text {F1}}-\omega _{\text {F2}}$. Pumping the SH resonance with an intense pump we can make the substitutions
$$\begin{aligned}\overline{c}_{\text{SH}}(t) & \rightarrow\left\langle \overline{c}_{\text{SH}}(t)\right\rangle e^{i\Delta_{\text{SH}}t}\equiv\beta_{\text{SH}}(t)e^{i\Delta_{\text{SH}}t},\\ \overline{\psi}_{\text{SH}<}(0,t) & \rightarrow\left\langle \overline{\psi}_{\text{SH}<}(0,t)\right\rangle e^{i\Delta_{\text{SH}}t}. \end{aligned}$$

Once more assuming an undepleted pump and ignoring all contributions from the phantom channel, the pump equation becomes

$$\frac{\textrm{d}}{\textrm{d}t}\beta_{\text{SH}}(t)=\left(-\overline{\Gamma}_{\text{SH}}-i\Delta_{\text{SH}}\right)\beta_{\text{SH}}(t)-i\gamma_{\text{SH}}^{*}\left\langle \overline{\psi}_{SH<}(0,t)\right\rangle.$$

Finally, the downconverted photons Heisenberg equation can be expressed in matrix form as

$$\frac{\textrm{d}}{\textrm{d}t}\begin{pmatrix}\tilde{c}_{\text{F1}}(t)\\ \tilde{c}_{\text{F2}}^{{\dagger}}(t) \end{pmatrix}=M_{\text{SPDC}}(t)\begin{pmatrix}\tilde{c}_{\text{F1}}(t)\\ \tilde{c}_{\text{F2}}^{{\dagger}}(t) \end{pmatrix}+D_{\text{SPDC}}(t),$$
where
$$\begin{aligned}M_{\text{SPDC}}(t) & =\begin{pmatrix}-\overline{\Gamma}_{\text{F1}}-i\frac{\Delta_{\text{SH}}-\Delta_{\text{ring}}}{2} & i\Lambda\beta_{\text{SH}}(t)\\ -i\Lambda^{*}\beta_{\text{SH}}^{*}(t) & -\overline{\Gamma}_{\text{F2}}+i\frac{\Delta_{\text{SH}}-\Delta_{\text{ring}}}{2} \end{pmatrix}, \end{aligned}$$
$$\begin{aligned}D_{\text{SPDC}}(t) & =\begin{pmatrix}-i\gamma_{\text{F1}}^{*}\tilde{\psi}_{\text{F1}<}(0,t)-i\gamma_{\text{F1}ph}^{*}\tilde{\psi}_{\text{F1ph}<}(0,t)\\ i\gamma_{\text{F2}}\tilde{\psi}_{\text{F2}<}^{{\dagger}}(0,t)+i\gamma_{\text{F2}ph}\tilde{\psi}_{\text{F2ph}<}^{{\dagger}}(0,t) \end{pmatrix}, \end{aligned}$$
and we have performed the transformation $ {\tilde {c}_{\text {F1},\text {F2}}(t)=e^{-\frac {i}{2}(\Delta _{\text {SH}}-\Delta _{\text {ring}})t}\overline {c}_{\text {F1},\text {F2}}(t)}$. The calculation of the $N$ and $M$ moments follow precisely the same path as for SP-SFWM (400)–(403) with the natural substitutions $\text {S}\rightarrow \text {F1}$ and $\text {I}\rightarrow \text {F2}$.

It is worth noting that the nonlinear Hamiltonian in (423) drives the creation of a NDSV state [cf. (188)] in modes F1 and F2. However, in the limit $\text {F1}\rightarrow \text {F2}\equiv \text {F}$, the solution transitions to a DSV state in the F mode, and all of the apparatus for the calculation of the Heisenberg equations is unchanged, provided that the SPDC nonlinear coefficient is replaced by

$$\Lambda_{\text{ring}}^{\text{SPDC}}\rightarrow2\Lambda_{\text{ring}}^{\text{SPDC}}.$$

6. Connections between Classical and Quantum Nonlinear Optics

In Sec. 3.7 we presented the evolution of a state different from vacuum in the presence of a quadratic Hamiltonian. We saw that even in the presence of a seed field, the joint amplitude appearing in the spontaneous case can act as a response function of the system. In other words, it seems as if there is a deep connection between spontaneous and stimulated nonlinear processes. Starting from these results, we now discuss this relation from a different point of view, with the aim of clarifying the connection between classical and quantum nonlinear optics, thus between those nonlinear phenomena that can be explained in the framework of a classical electromagnetic theory (e.g., second harmonic generation) and those that require the quantization of the electromagnetic field (e.g., SPDC). On the one hand, such a connection is interesting from a fundamental point of view, as it allows one to separate the features of a nonlinear phenomenon that are intrinsically quantum from those that are not. On the other hand, the link between classical and quantum nonlinear optics can also have important practical consequences. First, many systems that have been designed to enhance classical nonlinear processes can inspire new solutions to control and amplify quantum nonlinear interactions. Second, as we show in this section, one can gain important information about the quantum correlations of nonclassical light generated by parametric fluorescence by studying the same system and nonlinear interaction in a regime in which they can both be described classically, with advantages in terms of speed and accuracy.

Here we restrict our analysis to the generation of squeezed light via SPDC or SFWM. As we have seen, both of these processes result in the generation of photon pairs and are associated with either the second- or third-order nonlinear optical response of the system. The generic name with which SPDC and SFWM are often referred to is parametric fluorescence, which suggests a certain analogy with the spontaneous emission of light, for example, occurring in an atomic system. This analogy is quite strong. Indeed, likewise the emission of a photon by an atomic system can be either spontaneous or stimulated by the radiation field, there exist also for SPDC and SFWM two corresponding nonlinear processes, difference frequency generation (DFG) and stimulated four-wave mixing (FWM), in which the emission of photons pairs is stimulated by the presence of an additional field, usually labeled seed [148]. More interestingly, in the limit of a sufficiently weak seed field, thus neglecting SPM and XPM induced by it, both the stimulated and spontaneous emission of pairs are described by the very same Hamiltonian, and the probabilities of spontaneous and stimulated emission of photon pairs are related. Under these hypotheses, the study of the stimulated processes, i.e., DFG and FWM, allows one to gain information about the quantum properties and the generation rate of pairs that would be emitted spontaneously through SPDC or SFWM in the absence of the seed field [149].

6.1 Low-Gain Regime: Generation of Two-Photon States

The connection between stimulated and spontaneous emission of photon pairs is manifest in the context of the generation of two-photon states, i.e., in the low-gain regime, for which one has a small probability of generating a photon pair per pump pulse or, in the case of a CW pump, when such a probability is small within the pump coherence time. As an example, we consider the case of SPDC and DFG in a waveguide in the limit of a CW pump. In this case the photon pair generation rates for spontaneous and stimulated emission can be written as [148]

$$R_\textrm{SPDC,wg}=\frac{P_P}{\mathcal{PA}}\frac{L^{3/2}}{\frac{3}{2}\sqrt{2\pi\left|\beta_2(\omega_P/2)\right|}}$$
and
$$R_\textrm{DFG,wg}=\frac{P_P}{\mathcal{PA}}\frac{L^2}{\frac{1}{2}\hbar\omega_P}P_S.$$

Here we assume phase matching between $\omega _P$ and $\omega _P/2$, with $P_P$ and $P_S$ the pump and seed powers, respectively. Finally, $\mathcal {P}$ is a characteristic power, which depends on the material nonlinearity, $\mathcal {A}$ is the nonlinear effective area (see (F4) in Appendix F), which we assume to be constant over the frequency range of interest here, $L$ is the length of the nonlinear interaction region, and $\beta _2$ is the GVD.

By taking the ratio of (433) and (432), one obtains a simple relation between the two rates

$$\frac{R_\textrm{SPDC,wg}}{R_\textrm{DFG,wg}}=\frac{\hbar\omega_P}{3\sqrt{2\pi\left|\beta_2(\omega_P/2)\right|L}}\frac{1}{P_S},$$
in which SPDC appears as a process stimulated by a certain “vacuum power”
$$P_\textrm{vac,wg}= \frac{\hbar\omega_P}{3\sqrt{2\pi\left|\beta_2(\omega_P/2)\right|L}},$$
which corresponds exactly to the effective power associated with the energy half photon in each of the waveguide longitudinal modes in the phase-matching bandwidth $\Delta \omega _{\mathrm{PM}}=2\sqrt {2.78/\beta _2(\omega _P/2)L}$ per transit time [150]. Analogous results can also be found for SPDC in resonant systems. For example, one can derive the generation rates for SPDC and DFG in the case of CW pumping in ring resonators [148], with
$$R_\textrm{SPDC,ring}=\frac{P_p}{\mathcal{PA}}\frac{4v_F^2v_P}{\pi\omega_p\omega_F}\frac{Q_PQ_F}{R},$$
and
$$R_\textrm{DFG,ring}=\frac{P_p}{\mathcal{PA}}\frac{4v_F^2v_P}{\hbar\pi\omega_p\omega_F^3}\frac{Q_PQ_F^2}{R},$$
where $\omega _P=2\omega _F$, $v_{F(P)}$ and $Q_{F(P)}$ are the group velocity and the quality factor at the fundamental (pump) frequency, respectively, and $R$ is the ring radius. Note that in deriving these formulas, losses due to either scattering or absorption are neglected. By dividing (436) by (437), one obtains
$$\frac{R_\textrm{SPDC,ring}}{R_\textrm{DFG,ring}}=\frac{\hbar\omega_F^2}{8Q_F}\frac{1}{P_S},$$
where one can identify
$$P_\textrm{vac,ring}= \frac{\hbar\omega_F^2}{8Q_F}$$
as the power associated with the vacuum. Similar results can be obtained for photonic crystal structures [151], and for SFWM in either nonresonant or resonant systems [148,152]. In Fig. 20, Azzini et al. [152] demonstrated that, for a given seed power, the ratio of the idler powers generated by stimulated and spontaneous FWM in ring resonators depends only the quality factor, which ultimately determined the vacuum power fluctuations in this system. This confirms that in the case of parametric nonlinear processes, spontaneous emission of photon pairs can be viewed as stimulated by vacuum power fluctuations. In addition, Eqs. (434) and (438), and similar relations, allow one to estimate the efficiency of the spontaneous process from that of the corresponding stimulated one.

 figure: Figure 20.

Figure 20. Ratio between the generated idler photons in spontaneous and stimulated FWM for several ring resonators. Reprinted with permission from [152]. Copyright 2012 Optical Society of America.

Download Full Size | PDF

Liscidini and Sipe [149] went even further and demonstrated that the biphoton wave function describing the spontaneously generated two-photon state acts as the response function characterizing how the seed pulse stimulates the emission of photon pairs in the corresponding classical process. Assuming the same pumping condition of the spontaneous process and a seed pulse sufficiently intense to be treated classically, but sufficiently weak to neglect any effect of SPM and XPM induced by the seed, it follows that

$$\beta_{\sigma,\textbf{k}_2}\approx\sqrt{2}\gamma\sum_{\sigma^{\prime}}\int\textrm{d}\textbf{k}_1\phi_{\sigma^{\prime},\sigma}(\textbf{k}_1,\textbf{k}_2)\beta_{\sigma^{\prime},\textbf{k}_1}^{*},$$
where $\beta _{\sigma,\textbf{k}_2}$ and $\beta _{\sigma ^{\prime },\textbf{k}_1}^{*}$ are the amplitudes of the stimulated and seeded classical fields associated with light exiting the system. Here, $\phi _{\sigma ^{\prime },\sigma }(\textbf{k}_1,\textbf{k}_2)$ and $|\gamma |^2$ are the biphoton wave function and the pair generation rate, were they generated spontaneously. Finally $\sigma,\sigma ^{\prime }$ and $\textbf{k}_1,\textbf{k}_2$ label all the discrete (e.g., polarization) and continuous (e.g., momentum) degrees of freedom, respectively. Starting from the result of (440), one can imagine to characterize the two-photon state that would be generated by parametric fluorescence by performing a stimulated emission tomography (SET), in which the seed is used to mimic one of the photon of the pairs and the stimulated beam is analyzed to reconstruct the biphoton wave function. This has been experimentally demonstrated in the case of polarization entangled photon pairs [153,154] and for two-photon states hyper-entangled in path and polarization [155]. Particularly interesting from a practical point of view is the study of the energy correlations of the generated photon pairs. In this respect, stimulated emission has been used to reconstruct the joint spectral density (JSD) in several systems, with pairs generated either by SPDC [156] or SFWM [157159]. This approach is particularly useful in the case of high-$Q$ resonant systems, in which high resolution is required. As an example, in Fig. 21, we show the comparison between the JSD measured by SET for a silicon microring resonator and the expected theoretical result. Finally, although most of the experiments have been focused on the JSD, there are also cases in which stimulated emission was used to determine the phase of the biphoton wave function [160,161].

 figure: Figure 21.

Figure 21. JSD (a) measured by SET and (b) calculated in the case of photon pairs generated under pulsed pumping in a microring resonator. Reprinted by permission from Macmillan Publishers Ltd: Grassani et al., Sci. Rep. 6, 23564 (2016) [159]. Copyright 2016.

Download Full Size | PDF

6.2 High-Gain Regime: Characterization of Squeezed Light

Although the vast majority of the results regarding the connection between classical and quantum nonlinear processes have been obtained in the low-gain regime, when only a few photon pairs are generated per pump pulse, a similar link is expected to hold also in the high-gain regime, e.g., in the case of the generation of bright squeezed light. Indeed, when effects associated with SPM and XPM induced by the seed field are negligible, stimulated and spontaneous processes are described by the same nonlinear Hamiltonian. This opens to the possibility of characterizing nonclassical light generated in a nonperturbative regime by exploiting the stimulated emission of photon pairs.

Recently, Triginer et al. [92] proposed that cascaded stimulated emission of photon pairs can be used to measure the strength of squeezed light generated by SPDC. Unlike in SET, where the connection between the biphoton wave function and the stimulated idler is direct (see (440)), in this case the authors used two cascaded DFG processes to have a self-reference field, which is made to interfere with the generated squeezed light. From such an interference, it is possible to gain information about the squeezing strength and phase information, which depends on a nontrivial interplay of SPM and XPM associated with the pump field.

7. World beyond Gaussian States: Tensor Networks for the Perplexed

7.1 Curse of Dimensionality in Many-Body Systems and MPSs

In Secs. 3, 4, and 5 we have introduced methods to describe the quantum states generated in quantum nonlinear optics whenever the Heisenberg equations of motion can be linearized. Under these circumstances we can use the methods from Sec. 3. In a nutshell, whenever we can linearize the equations of motion for the field operators we can always write the output of the spontaneous problem as a multimode squeezed state, which in the low-gain limit corresponds to pair production. Although these methods provide useful tools for many problems, they leave open the question of what can be done beyond linearized dynamics and Gaussian state generation. A naïve approach to this problem is to expand every possible time, position or frequency bin into an orthonormal basis, for example, the Fock basis. Thus, a state over $\ell$ bins would be written as

$$|{\Psi}\rangle = \sum_{i_0=0}^{c-1} \cdots \sum_{i_{\ell-1}=0}^{c-1} R_{i_0\ldots i_{\ell-1}} |{i_0,\ldots,i_{\ell-1}}\rangle,$$
where $R_{i_0\ldots i_{\ell -1}}$ is an $\ell$-rank tensor $| {i_0 \ldots i_{\ell -1}}\rangle \equiv | {i_0}\rangle \otimes \cdots \otimes | {i_{\ell -1}}\rangle$ and $| {i_k}\rangle = \tfrac {a_k^{\dagger i}}{\sqrt {i!}} | {\text {vac}}\rangle$ is a Fock state with $i$ photons in the $k$th bin. If we assume a minimal cutoff of one photon per mode ($c=2$) and consider, for example, $\ell = 100$ bins, we arrive at the conclusion that we need to store in memory the $2^{100} \approx 10^{30}$ elements of the tensor $\boldsymbol{R}$. This is simply impractical and seems to put the simulation of these systems beyond any current or future classical (super)computer. Now let us assume that the tensor $\boldsymbol{R}$ can be written in the following so-called MPS [162164] form (as shown schematically in Fig. 22)
$$R_{i_0,\ldots,i_{\ell-1}} =\sum_{\alpha_1=0}^{D_1-1} \ldots \sum_{\alpha_{\ell-1} =0}^{D_{\ell-1} - 1} (S^{[0]})_{i_0,\alpha_1} (S^{[1]})_{i_1,\alpha_2}^{\alpha_1} \ldots (S^{[\ell-2]})_{i_{\ell-2},\alpha_{\ell-1}}^{\alpha_{\ell-2}} (S^{[\ell-1]})_{i_{\ell-1}}^{\alpha_{\ell-1}},$$
where for $n \in \{1,\ldots,\ell -2\}$ the $S^{[n]}$ are rank-three tensors whereas $S^{[0]}$ and $S^{[\ell -1]}$ are rank-two tensors (matrices). For different notations used when describing MPSs, see Appendix H. The indices $i_0,\ldots,i_{\ell -1}$ are called physical indices, whereas the contracted indices $\alpha _k$ in the sum in the last equation are known as auxiliary indices. In general, an arbitrary tensor with $\ell$ indices can always be taken into the matrix product form above (see Appendix H; however, it is not obvious a priori that the expression of (442) is useful. In particular, we do not know the possible values $\{D_k\}$ the auxiliary indices $\{\alpha _k\}$ can take. The maximum of these quantities over all possible auxiliary indices, $D = \max \{D_1,\ldots, D_{\ell -1}\}$, is called the bond dimension of the tensor and, in principle, for an arbitrary tensor this quantity can be exponentially large in the number of physical indices $\ell$. However, if this is not the case and the bond dimension is constant or grows only polynomially in $\ell$, then we can obtain significant savings in representing a tensor similarly to the form of (442). Indeed, if the bond dimension of the matrix product expression in (442) is a constant $D$ then each of the rank-three tensors will have a total number of elements bounded by $D c^2$ and if we have a total of $\ell -2$ rank-three tensors the memory required to represent the tensor $\boldsymbol{R}$ will not be larger than $\ell D c^2$. In practice, the cutoff can be chosen self-consistently, by increasing it until there is convergence.

 figure: Figure 22.

Figure 22. (a) Graphical representation of an arbitrary tensor $R$ with indices $i_0,i_1,\ldots, i_{\ell -1}$. (b) Representation of the same tensor using MPSs with auxiliary indices $\alpha _1,\alpha _2,\ldots,\alpha _{\ell -1}$.

Download Full Size | PDF

If a quantum state can be efficiently represented as an MPS, then one can easily calculate expectation values of local observables such as correlation functions. As recently shown by Yanagimoto et al. [30] one can also extract the density matrix of any Schmidt (or broadband) mode $A_\lambda ^\dagger = \sum _{k=0}^{\ell -1} F_{k,\lambda } a_k^\dagger$ defined as a linear combination of the bare modes $\{a_k \}$.

In the following section we show how to construct a constant bond-dimension MPS representation of a resonant system with an arbitrary Hamiltonian coupled to a waveguide with a linear dispersion relation. Note that very recently it was also shown that MPSs provide an efficient description of quantum states where the nonlinearity is distributed along a dispersive nonlinear waveguide [30]. Finally, note that the results we will derive are a qudit extension of the qubit results derived by Schön et al. [165,166], where they consider “the deterministic generation of entangled multiqubit states by the sequential coupling of an ancillary system to initially uncorrelated qubits” and “characterize all achievable states in terms of classes of matrix-product states”.

7.2 MPSs in Cavity QED

We start by writing a Hamiltonian for an isolated resonance coupled to a waveguide [167]

$$\tilde H = \tilde H_{\text{res}} + \tilde H_{\text{wg}} + \tilde V,$$
where
$$\begin{aligned}\tilde H_{\text{res}} & = \hbar \omega_0 a^\dagger a + \tilde H_{\text{NL}}(a^\dagger,a,t), \end{aligned}$$
$$\begin{aligned}\tilde H_{\text{wg}} & = \hbar \omega_1 \int \textrm{d}x \ \psi^\dagger(x) \psi(x) + \frac{i}{2} \hbar v \int \textrm{d}x \left( \left[\frac{\partial}{\partial x} \psi^\dagger(x) \right] \psi(x) - \text{H.c.} \right), \end{aligned}$$
$$\begin{aligned}\tilde V & = \hbar \gamma \left( \psi^\dagger(0)a + \text{H.c.} \right), \end{aligned}$$
and $\omega _0/\omega _1$ are the frequencies of the resonator/waveguide, $v$ is the group velocity of the waveguide mode and $\gamma ^2/(2v)$ is the decay rate of the resonator into the waveguide. We furthermore assume point coupling between the waveguide and the resonator at position $x=0$. All the calculations in the section are done in the interaction picture, meaning that we will move some trivial time evolution into the operators, whereas the nontrivial part is carried by the state vector. Moreover, we ignore dispersion within each field in the waveguide, as done in Sec. 5.

The nonzero canonical commutation relations of the field and resonance are

$$[\psi(x),\psi^\dagger(x')]=\delta(x-x') \quad \text{and} \quad [a,a^\dagger] = 1.$$

Note that $H_{\text {NL}}$ is a nonlinear Hamiltonian that is written in terms of resonator operators and that is possibly time-dependent. We leave this completely general, but it can be taken to be, for example, that of a classical field with frequency $\omega _c \approx 3\omega _0$ coupled to the resonance of interest; for this case we would write the Hamiltonian

$$\tilde{H}_{\rm NL}(a^\dagger, a, t) = \mu \beta(t) e^{i \omega_c t} a^3 + \mu^* \beta^* e^{{-}i \omega_c t} a^{{\dagger} 3},$$
where $\mu$ is proportional to the coupling between the modes inside the resonator and $\beta (t)$ is the slowly varying amplitude of the field in the classical resonance.

We can now go to an interaction picture rotating at frequency $\omega _1$ by transforming the Hamiltonian using the following unitary (see Appendix C):

$$\mathcal{ U}_1 = \exp\left(- \frac{i}{\hbar} H_1 (t-t_0) \right),\\\ H_1 = \hbar \omega_1\left[a^\dagger a + \int \textrm{d}x \ \psi^\dagger(x) \psi(x) \right],$$
which now gives the Hamiltonian
$$\tilde H = \tilde H_{\text{res}} + \tilde H_{\text{wg}} + \tilde V,$$
where
$$\begin{aligned}\bar H_{\text{res}} & = \hbar \delta a^\dagger a + H_{\text{NL}}(t), \quad \delta = \omega_0 - \omega_1, \end{aligned}$$
$$\begin{aligned}\bar H_{\text{wg}} & = \frac{i}{2} \hbar v \int \textrm{d}x \left( \left[ \frac{\partial}{\partial x} \psi^\dagger(x) \right] \psi(x) - \text{H.c.} \right), \end{aligned}$$
$$\begin{aligned}H_{\text{NL}}(t) & = \tilde{H}_{\text{NL}}({e^{i \omega_1 t}a^\dagger},e^{{-}i \omega_1 t} a,t), \end{aligned}$$
$$\begin{aligned}\bar V & = \tilde V. \end{aligned}$$

Note that by going into the interaction picture with respect to $\omega _1$ the nonlinear Hamiltonian is now evaluated at the time evolving operators $e^{i \omega _1 t}a^\dagger$ and $e^{-i \omega _1 t} a$. Finally, we can go to a second interaction picture with the unitary operation

$$\begin{aligned}\mathcal{U}_2 & = \exp\left( - \frac{i}{\hbar} (t-t_0) \bar{H}_{\text{wg}} \right) \\ & = \exp\left( (t-t_0) \frac{v}{2} \int \textrm{d}x \left\{ \left[ \frac{\partial}{\partial x} \psi^\dagger(x) \right] \psi(x) - \text{H.c.} \right\} \right), \end{aligned}$$
to obtain the final form of the Hamiltonian
$$H =\hbar \delta a^\dagger a + H_{\text{NL}}(t) + \hbar \gamma \left( \psi^\dagger({-}v (t-t_0))a + \text{H.c.} \right).$$

Note that the last term on the right-hand side can be written as

$$\hbar \gamma \left( \psi^\dagger({-}v (t-t_0))a + \text{H.c.} \right)= \hbar \gamma \left[ a \int \textrm{d}x \ \psi^\dagger (x)\delta(x+v(t-t_0)) +\text{H.c.} \right],$$
where it is made explicit that the Hamiltonian is time-dependent but it is only made of Schrödinger operators.

This (interaction-picture) Hamiltonian has a very nice physical interpretation. The evolution of the resonator is given by the first two-terms in (456); then, at a given time $t$, the section of waveguide field located at $x=-v(t-t_0)$ gets to exchange some energy with the resonator by a beam splitter (hopping) interaction mediated by the last term in (456).

Now we are ready to discretize time and space. To this end we introduce $t_j = t_0 + j \Delta t$ and write

$$\begin{aligned}\psi_j & = \frac{1}{\sqrt{\Delta x}} \int_{-(j+1/2)\Delta x}^{-(j-1/2)\Delta x} \textrm{d}x \ \psi(x)& \approx \sqrt{\Delta x} \ \psi(- j \Delta x) = \sqrt{\Delta x} \ \psi(- j v \Delta t). \end{aligned}$$

This operator satisfies $[\psi _j, \psi _k^\dagger ] = \delta _{j,k}$ and we can now write the time-dependent field-resonance Hamiltonian as

$$H_j = \hbar \delta a^\dagger a + H_{\text{NL}}(t_0+j\Delta t) +i \hbar \sqrt{\frac{\gamma v}{\Delta x }} \left( \psi_j^\dagger a -\text{H.c.} \right).$$

The Trotterized time-evolution operator after discretizing time in slices of size Δt and space into $\ell$ slices can be written as

$$\begin{aligned} \mathcal{ U} & = \left[ \mathcal{B}_{0,\ell-1} \mathcal{V}_0(\ell-1) \right] \ldots \left[ \mathcal{B}_{0,0} \mathcal{V}_0(0) \right], \end{aligned}$$
$$\begin{aligned}\mathcal{V}_0(j) & = \exp\left(-\frac{i}{\hbar} \Delta t \left[ \hbar \delta a^\dagger a + H_{\text{NL}}(t_0+j \Delta t ) \right] \right), \end{aligned}$$
$$\begin{aligned}\mathcal{B}_{0,j} & = \exp\left( i \gamma \sqrt{ \frac{\Delta t} {v}} \left( \psi^\dagger_j a + \text{H.c.} \right) \right), \end{aligned}$$
where for each time “slice” ($\Delta t = \Delta x/v$) two things happen as follows:
  • 1. At time $j$ the local unitary $\mathcal {V}_0(j)$ is applied on the resonator mode. Note that we allow for an explicit dependence on time $j$.
  • 2. A beam splitter by rotation angle $\theta = \gamma \sqrt { \frac {\Delta t} {v}}$ is applied between the resonator mode and the waveguide section labeled by the index $j$.
These two sets of operations can be represented by a quantum circuit diagram as shown in Fig. 23. Note that in this diagram the first wire refers to the resonance whereas the rest of the wires represent the different discretized sections of the waveguide. In this representation we are required to carry out beam splitter interactions between non-nearest-neighbor modes. Thus in this representation the interactions between the different wires are nonlocal. To simplify the evaluation of the circuit we now permute the order of the wires, that is, after each beam splitter operation we swap the resonator wire with the corresponding waveguide section [168]. This will simplify the calculation because the interaction between the different modes will now only consist of nearest neighbors beam splitters as shown in Fig. 24. Physically this corresponds to a moving frame in which the field is no longer propagating at its group velocity but instead the resonance is moving in the opposite direction at the same group velocity.

 figure: Figure 23.

Figure 23. Circuit representation of the unitary evolution operator in (460) in a discretization using four modes. Note that in this figure the first wire represents the resonator mode, whereas the rest of the wires represent the different discretized sections of the waveguide. We use vertical lines with dots at the ends to represent the beam splitter operation $\mathcal {B}_{0,j}$ between modes 0 (the resonator) and the $j$th part of the waveguide.

Download Full Size | PDF

 figure: Figure 24.

Figure 24. Circuit representation of the unitary evolution operator in (460) and for a discretization using four modes. Note that in this figure we apply swap operations after each beam splitter. Thus, after the $i$th swap the resonator mode lies in the $i$th wire and, in particular, at the end of the interaction the resonator mode is represented by the last wire. We use vertical lines with dots at the end to represent the beam splitter operation $\mathcal {B}_{i,i+1}$ between wire $i$ and $i+1$ and similarly we use vertical lines with crosses to represent swap operations between wires $i$ and $i+1$.

Download Full Size | PDF

Introducing the following two-wire unitary

$$\mathcal{W}(i+1,i) = \text{SWAP}_{i+1,i} \mathcal{B}_{i+1,i} \mathcal{V}_i(i),$$
we can write the Trotterized evolution operator as
$$\mathcal{U} = \mathcal{W}(\ell, \ell-1) \ldots \mathcal{W}(1,0),$$
with the added convention that the resonator mode starts in the first wire but ends in the last one. We consider a total of $\ell +1$ modes where $\ell$ modes are used for the discretization of the waveguide and the remaining one for the resonance. Now assume that the initial state has product form
$$| \Psi (t_0) \rangle = \bigotimes_{i=0}^\ell | 0_i \rangle,$$
where $| 0_i \rangle$ is the vacuum state in either the resonance or a section of the waveguide (this can be generalized to any product state such as a broadband coherent state). We now want to argue that if one starts with a product state, then the time evolved state has MPS form. To this end we calculate the tensor
$$R_{i_0,i_1,\ldots,r_\ell} = \langle i_0,i_1,\ldots,i_\ell|\mathcal{U}|\Psi \rangle,$$
and show that it has MPS form. We consider a discretization with $\ell =4$ modes and with Fock cutoff $c$ but everything follows in a straightforward manner for an arbitrary number of modes. We thus want to find
$$\begin{aligned}R_{i_0, i_1, i_2, i_3, i_4} & = \langle i_0, i_1, i_2, i_3, i_4|\mathcal{ U}|0_0,0_1,0_2,0_3,0_4 \rangle \\ & = \langle i_0, i_1, i_2, i_3, i_4|W(4,3)W(3,2)W(2,1)W(1,0)|0_0,0_1,0_2,0_3,0_4 \rangle. \end{aligned}$$

We start by writing

$$\langle i_0|W(0,1)|0_0,0_1 \rangle=\sum_{\alpha_1=0}^{c-1} (T^{[0]})^{i_0,\alpha_1}_{0_0,0_1} | \alpha_1 \rangle \leftrightarrow \langle i_0, \alpha_1|W(1,0)|0_0,0_1 \rangle= (T^{[0]})^{i_0,\alpha_1}_{0_0,0_1},$$
and then find
$$R_{i_0, i_1, i_2, i_3, i_4} = \sum_{\alpha_1=0}^{c-1} (T^{[0]})^{i_0,\alpha_1}_{0_0,0_1} \langle i_1, i_2, i_3, i_4|W(4,3)W(3,2)W(2,1)|\alpha_1,0_2,0_3,0_4 \rangle.$$

Next we introduce

$$\langle i_1|W(2,1)|\alpha_1,0_2 \rangle=\sum_{\alpha_2=0}^{c-1} (T^{[1]})^{i_1,\alpha_2}_{\alpha_1,0_2} | \alpha_2 \rangle,$$
to find
$$R_{i_0, i_1, i_2, i_3, i_4} = \sum_{\alpha_1,\alpha_2=0}^{c-1} (T^{[0]})^{i_0,\alpha_1}_{0_0,0_1} (T^{[1]})^{i_1,\alpha_2}_{\alpha_1,0_2} \langle i_2, i_3, i_4|W(4,3)W(3,2)|\alpha_2,0_3,0_4 \rangle.$$

We can proceed like this two more times, defining

$$\begin{aligned}\langle i_2|W(3,2)|\alpha_2,0_3 \rangle & =\sum_{\alpha_3=0}^{c-1} (T^{[2]})^{i_2,\alpha_3}_{\alpha_2,0_3} | \alpha_3 \rangle, \end{aligned}$$
$$\begin{aligned}\langle i_3|W(4,3)|\alpha_3,0_4 \rangle & =\sum_{\alpha_4=0}^{c-1} (T^{[3]})^{i_3,\alpha_4}_{\alpha_3,0_4} | \alpha_4 \rangle, \end{aligned}$$
to finally write the coefficients of the time-evolved ket in MPS form
$$R_{i_0, i_1, i_2, i_3, i_4} = \sum_{\alpha_1,\alpha_2,\alpha_3,\alpha_4=0}^{c-1} (T^{[0]})^{i_0,\alpha_1}_{0_0,0_1} (T^{[1]})^{i_1,\alpha_2}_{\alpha_1,0_2} (T^{[2]})^{i_2,\alpha_3}_{\alpha_2,0_3} (T^{[3]})^{i_3,\alpha_4}_{\alpha_3,0_4} \langle i_4|\alpha_4 \rangle.$$

Defining the simplified notation $(T^{[k]})^{x,y}_{w,0_p} \equiv (S^{[k]})_{x,y}^{w}$ for $0<k< \ell = 4$ (recall that $0_p$ is fixed from the initial ket (465)) and also $(T^{[0]})^{x,y}_{0_0,0_1} \equiv (S^{[0]})_{i_0,\alpha _1}$ and $(S^{[4]})_{i_4}^{j_4} = \langle i_4|\alpha _4 \rangle$ we can finally write the MPS form

$$R_{i_0, i_1, i_2, i_3, i_4} = \sum_{\alpha_1,\alpha_2,\alpha_3,\alpha_4=0}^{c-1} (S^{[0]})_{i_0,\alpha_1} (S^{[1]})_{i_1,\alpha_2}^{\alpha_1} (S^{[2]})_{i_2,\alpha_3}^{\alpha_2} (S^{[3]})_{i_3,i_4}^{\alpha_3} (S^{[4]})^{\alpha_4}_{i_4},$$
which is the MPS form of the ket. Note that for this ket the bond dimension is precisely the Fock cutoff $c$ and, thus, our $(\ell +1)$-mode time-evolved system can be represented in memory with only on the order $\ell c^3$ coefficients, as opposed to the $c^{\ell +1}$ coefficients a naïve Fock representation would entail.

As stated in the previous subsection, once the MPS form of a state is obtained, most useful properties of the state can be obtained in a straightforward manner. The methods to obtain expectation values from MPS states are presented and derived elsewhere [163,164]. We would just like to point out that standard computer libraries already exist for the calculation of expectation values of local observables from MPSs [169172].

8. Summary, Conclusions, and Future Outlook

Since the initial experiments on squeezing carried out more than three decades ago, there has been a constant drive to move quantum optics from a collection of bulk optical elements spanning large tabletops to a single chip-scale device no larger than a fingernail. Such a transition not only reduces the system size and power requirements, but it also offers the promise of scalability and an unprecedented control over the device parameters, and the properties of the generated light. Although advances in the fabrication of integrated photonic devices have accelerated this evolution, they have also ushered in an era of “high-gain” quantum optics, for example, enabling single-photon generation via post-selection involving photon-number resolving detection, efficient photon frequency conversion, and state generation beyond the single-photon pair regime. For these and other high-gain applications, a first-order perturbation theory, the traditional tool of choice, is no longer sufficient to describe the nonlinear interaction. A key theoretical issue is that relevant Hamiltonian terms describing the nonlinear interaction do not commute with themselves at different times. In short, this means that weakly transformed photon amplitudes at early times influence their own transformation at later times.

Drawing on earlier works, in this tutorial we have shown how to overcome this difficulty, starting from a rigorous first-principles quantization procedure and employing a Trotter–Suzuki expansion of linearized Heisenberg picture dynamics to reduce the problem to the multiplication of matrices with a size proportional to the number of bosonic modes involved. In particular, in Sec. 2 we detailed the quantization of nonlinear optics in integrated photonic structures, paying particular attention to channel waveguides and ring resonators. In Sec. 3 we presented the tools necessary to calculate the output states of a few nonlinear optical processes, whether in the low-gain or high-gain regimes, and whether the initial quantum state is vacuum or not, resulting in a process that is spontaneous or stimulated, respectively. These tools correctly account for loss and material dispersion and, as we showed, they can be readily applied to calculate the results of homodyne and photon-number resolved measurements of the generated states. With these mathematical tools firmly established, we then turned to the particular structures singled out in Sec. 2, going into further detail regarding their application to waveguides in Sec. 4 and ring resonators in Sec. 5 for various SPDC and SFWM processes. In Sec. 6 we discussed the deep connection between spontaneous and stimulated processes, showing how the knowledge of the latter can be used to inform the former. Finally, in Sec. 7 we examined how tensor networks and MPSs can be used to move beyond linearized equations of motion and Gaussian state generation. Throughout this tutorial, we have also emphasized connections to relevant theoretical and experimental works.

It is our hope that the tools provided here will benefit the entire integrated quantum photonics community, leading to devices with still greater complexity and enhanced functionalities than their present-day counterparts, especially as state generation moves beyond the single or few photon pairs regime. Indeed, integrated quantum photonic technologies promise ultimate information security, enhanced measurement precision, and greater computing power than traditional technologies, and their realization is now within reach.

Appendix A: Field Normalization and Dispersion

As a first example we look at a uniform medium in the dispersionless limit, where $\varepsilon _{1}(\boldsymbol {r})=\varepsilon _{1}$. Then the equations (18) for $\boldsymbol {B}_{\alpha }(\boldsymbol {r})$ become simply

$$\begin{aligned} & \boldsymbol{\nabla}\times\left(\boldsymbol{\nabla}\times\boldsymbol{B}_{\alpha}(\boldsymbol{r})\right)=\frac{\omega_{\alpha}^{2}\varepsilon_{1}}{c^{2}}\boldsymbol{B}_{\alpha}(\boldsymbol{r}),\\ & \boldsymbol{\nabla}\cdot\boldsymbol{B}_{\alpha}(\boldsymbol{r})=0. \end{aligned}$$

Solutions satisfying periodic boundary conditions are of the form $\boldsymbol{b}\exp(i\boldsymbol{k}\cdot\boldsymbol{r})$, where the vector $\boldsymbol {b}$ is perpendicular to $\boldsymbol {k}$ and

$$ \omega_{\alpha}=c\left|\boldsymbol{k}\right|/n, $$
where $n=\sqrt {\varepsilon _{1}}$. If the normalization volume $V$ is a cube with each side of length $L$, $V=L^{3}$, the allowed $\boldsymbol {k}$ are of the form
$$\boldsymbol{k}=\frac{2\pi}{L}(m_{x}\boldsymbol{\hat{x}}+m_{y}\boldsymbol{\hat{y}}+m_{z}\boldsymbol{\hat{z}}), $$
where the $m_{i}$ are integers and $\boldsymbol {\hat {x}}$ is a unit vector pointing in the $x$ direction, etc. There are two independent vectors perpendicular to $\boldsymbol {k}$; call them $\boldsymbol {\hat {e}}_{1\boldsymbol {k}}$ and $\boldsymbol {\hat {e}}_{2\boldsymbol {k}}$, and define them such that
$$\boldsymbol{\hat{e}}_{1\boldsymbol{k}}\times\boldsymbol{\hat{e}}_{2\boldsymbol{k}}=\boldsymbol{\hat{k}}. $$

In a uniform medium it is often useful to work with left- and right-handed circularly polarized components of light, which can be identified by the polarization vectors

$$\begin{aligned} & \boldsymbol{\hat{e}}_{L\boldsymbol{k}}={-}\frac{1}{\sqrt{2}}\left(\boldsymbol{\hat{e}}_{1\boldsymbol{k}}+i\boldsymbol{\hat{e}}_{2\boldsymbol{k}}\right),\\ & \boldsymbol{\hat{e}}_{R\boldsymbol{k}}=\frac{1}{\sqrt{2}}\left(\boldsymbol{\hat{e}}_{1\boldsymbol{k}}-i\boldsymbol{\hat{e}}_{2\boldsymbol{k}}\right), \end{aligned}$$
respectively. Using $I$ to denote either $L$ or $R$ (positive or negative helicity respectively), and $\overline {I}$ to equal the opposite helicity, we have
$$\begin{aligned} & \boldsymbol{\hat{e}}_{I\boldsymbol{k}}\cdot\boldsymbol{\hat{e}}_{I\boldsymbol{k}}^{*}=1,\\ & \boldsymbol{\hat{e}}_{I\boldsymbol{k}}\cdot\boldsymbol{\hat{e}}_{\overline{I}\boldsymbol{k}}^{*}=0, \end{aligned}$$
and
$$\begin{aligned} & i\boldsymbol{\hat{k}}\times\boldsymbol{\hat{e}}_{I\boldsymbol{k}}=s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}},\\ & \hat{\boldsymbol{e}}_{I\boldsymbol{k}}\times\boldsymbol{\hat{e}}_{I\boldsymbol{k}}^{*}={-}is_{I}\boldsymbol{\hat{k}}, \end{aligned}$$
where $\boldsymbol{\hat{k}}\equiv{\boldsymbol{k}}/\left|\boldsymbol{k}\right|$ and $s_{L}=1$, $s_{R}=-1$ identifies the helicity. It is convenient to link the definitions of $\boldsymbol {\hat {e}}_{1\boldsymbol {k}}$ and $\boldsymbol {\hat {e}}_{2\boldsymbol {k}}$ to those of $\boldsymbol {\hat {e}}_{1(-\boldsymbol {k})}$ and $\boldsymbol {\hat {e}}_{2(-\boldsymbol {k})}$ according to
$$\begin{aligned} & \boldsymbol{\hat{e}}_{1(-\boldsymbol{k})}=\boldsymbol{\hat{e}}_{1\boldsymbol{k}},\\ & \boldsymbol{\hat{e}}_{2(-\boldsymbol{k})}={-}\boldsymbol{\hat{e}}_{2\boldsymbol{k}}, \end{aligned}$$
for then
$$\begin{aligned} & \boldsymbol{\hat{e}}_{I\boldsymbol{k}}^{*}=\boldsymbol{\hat{e}}_{I(-\boldsymbol{k})},\\ & \boldsymbol{\hat{e}}_{I\boldsymbol{k}}\cdot\boldsymbol{\hat{e}}_{I'(-\boldsymbol{k})}=\delta_{II'}. \end{aligned}$$

We can label the modes $\alpha$ by the pair of indices $I,\boldsymbol {k}$, and seeking $\boldsymbol {b}$ vectors proportional to $s_{I}\boldsymbol {\hat {e}}_{I\boldsymbol {k}}$ the normalization condition (23) means we can take

$$\boldsymbol{B}_{I\boldsymbol{k}}(\boldsymbol{r})=\sqrt{\frac{\mu_{0}}{V}}s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}},$$
whereas from the last of (18) we then find
$$\boldsymbol{D}_{I\boldsymbol{k}}(\boldsymbol{r})=in\sqrt{\frac{\epsilon_{0}}{V}}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}$$
and for the mode frequencies we write
$$\omega_{I\boldsymbol{k}}=\frac{c}{n}\left|\boldsymbol{k}\right|, $$
which, of course, only depend on $\left |\boldsymbol {k}\right |$. Note that the partner mode of a mode of polarization $I$ with propagation wave vector $\boldsymbol {k}$ is the mode of polarization $I$ with propagation wave vector $-\boldsymbol {k}$. Clearly other choices of the modes in addition to (A10) and (A11) are possible; one could multiply both by an overall phase factor or the factor $s_{I}$, or one could work with the linearly polarized basis $\boldsymbol {\hat {e}}_{1\boldsymbol {k}}$ and $\boldsymbol {\hat {e}}_{2\boldsymbol {k}}$ at each $\boldsymbol {k}$. However, choosing (A10) and (A11) in (38) we find
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=i\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar cn\left|\boldsymbol{k}\right|\epsilon_{0}}{2V}}a_{I\boldsymbol{k}}(t)\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar c\left|\boldsymbol{k}\right|\mu_{0}}{2nV}}a_{I\boldsymbol{k}}(t)s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.} \end{aligned}$$

Finally, we consider the passage to an infinite normalization volume. Here the analogs of (49) and (51) in three dimensions are

$$\sum_{\boldsymbol{k}}\rightarrow\frac{\int \textrm{d}\boldsymbol{k}}{\left(\frac{2\pi}{L}\right)^{3}}=\frac{V}{8\pi^{3}}\int {\rm d}\boldsymbol{k}$$
and
$$a_{I\boldsymbol{k}}(t)\rightarrow\sqrt{\frac{8\pi^{3}}{V}}a_{I}(\boldsymbol{k},t).$$

We then have

$$\begin{aligned} & \left[a_{I}(\boldsymbol{k},t),a_{I'}^{{\dagger}}(\boldsymbol{k'},t)\right]=\delta_{II'}\delta(\boldsymbol{k}-\boldsymbol{k'}),\\ & \left[a_{I}(\boldsymbol{k},t),a_{I'}(\boldsymbol{k'},t)\right]=0. \end{aligned}$$

In terms of these variables $a_{I}(\boldsymbol {k},t)$ (A13) become

$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=i\sum_{I}\int\sqrt{\frac{\hbar cn\left|\boldsymbol{k}\right|\epsilon_{0}}{16\pi^{3}}}a_{I}(\boldsymbol{k},t)\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}\textrm{d}\boldsymbol{k}+\text{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I}\int\sqrt{\frac{\hbar c\left|\boldsymbol{k}\right|\mu_{0}}{16\pi^{3}n}}a_{I}(\boldsymbol{k},t)s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}\textrm{d}\boldsymbol{k}+\text{H.c.}, \end{aligned}$$
where we keep the same notation for $\boldsymbol {\hat {e}}_{I\boldsymbol {k}}$ and $\omega _{I\boldsymbol {k}}$ as in (A13), although $\boldsymbol {k}$ now ranges continuously. The Hamiltonian (40) becomes
$$H=\sum_{I}\int\hbar\omega_{I\boldsymbol{k}}a_{I}(\boldsymbol{k},t)a_{I}(\boldsymbol{k,}t)\textrm{d}\boldsymbol{k}.$$

We now turn to the inclusion of dispersion. The consequences of this new factor $v_{p}(\boldsymbol {r};\omega _{\alpha })/v_{g}(\boldsymbol {r};\omega _{\alpha })$ in the normalization condition (67) can be easiest seen by first considering periodic boundary conditions. As in a uniform medium the relative dielectric constant and the velocities are independent of position, we can put $\varepsilon _{1}(\boldsymbol {r};\omega _{I\boldsymbol {k}})\rightarrow \varepsilon _{1}(\omega _{I\boldsymbol {k}})$, $n(\boldsymbol {r},\omega _{I\boldsymbol {k}})\rightarrow n(\omega _{I\boldsymbol {k}})$, $v_{p,g}(\boldsymbol {r};\omega _{I\boldsymbol {k}})\rightarrow v_{p,g}(\omega _{I\boldsymbol {k}})$, and the modes (A10) and (A11) in (38) in the dispersionless limit are replaced by

$$\boldsymbol{B}_{I\boldsymbol{k}}(\boldsymbol{r})=\sqrt{\frac{\mu_{0}v_{g}(\omega_{I\boldsymbol{k}})}{Vv_{p}(\omega_{I\boldsymbol{k}})}}s_{I} \boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}},$$
and
$$\boldsymbol{D}_{I\boldsymbol{k}}(\boldsymbol{r})=in(\omega_{I\boldsymbol{k}})\sqrt{\frac{\epsilon_{0}v_{g}(\omega_{I\boldsymbol{k}})}{Vv_{p} (\omega_{I\boldsymbol{k}})}}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}},$$
where now the mode frequency for a given $\boldsymbol {k}$ is determined by self-consistently solving
$$n(\omega_{I\boldsymbol{k}})\omega_{I\boldsymbol{k}}=c\left|\boldsymbol{k}\right|$$
for $\omega _{I\boldsymbol {k}}$ (compare (A2)), and the field expansions (A17) become
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=i\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|\epsilon_{0}\varepsilon_{1} (\omega_{I\boldsymbol{k}})v_{g}(\omega_{I\boldsymbol{k}})}{2V}}a_{I\boldsymbol{k}}(t)\boldsymbol{\hat{e}}_{I\boldsymbol{k}} e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|\mu_{0}v_{g}(\omega_{I\boldsymbol{k}})}{2V}} a_{I\boldsymbol{k}}(t)s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.} \end{aligned}$$

Again we form the Poynting vector $\boldsymbol {S}(\boldsymbol {r},t)=\boldsymbol {E}(\boldsymbol {r},t)\times \boldsymbol {H}(\boldsymbol {r},t)$; using (63) and (64) we have

$$\begin{aligned} & \boldsymbol{E}(\boldsymbol{r},t)=i\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|v_{g}(\omega_{I\boldsymbol{k}})}{ 2V\epsilon_{0}\varepsilon_{1}(\omega_{I\left|\boldsymbol{k}\right|})}}a_{I\boldsymbol{k}}(t)\boldsymbol{\hat{e}}_{I\boldsymbol{k}} e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.},\\ & \boldsymbol{H}(\boldsymbol{r},t)=\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|v_{g}(\omega_{I\boldsymbol{k}})}{2V\mu_{0}}} a_{I\boldsymbol{k}}(t)s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.}, \end{aligned}$$
and considering a state $\left |\Psi \right \rangle$ with one photon in mode $I,\boldsymbol {k}$,
$$\left|\Psi\right\rangle =a_{I\boldsymbol{k}}^{{\dagger}}(0)\left\vert\text{vac}\right\rangle$$
where $\left |vac\right \rangle$ is the vacuum state, we find
$$\int_{V}\left\langle \boldsymbol{S}(\boldsymbol{r},t)\right\rangle d\boldsymbol{r}=\hbar\omega_{I\boldsymbol{k}}v_{g}(\omega_{I\boldsymbol{k}}), $$
with the group velocity governing the energy flow, as expected.

For reference, we give the field expansion for a uniform infinite medium, the appropriate generalization of (A17):

$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=i\sum_{I}\int\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|\epsilon_{0} \varepsilon_{1}(\omega_{I\boldsymbol{k}})v_{g}(\omega_{I\boldsymbol{k}})}{16\pi^{3}}}a_{I}(\boldsymbol{k},t) \boldsymbol{\hat{e}}_{I}(\boldsymbol{k})e^{i\boldsymbol{k}\cdot\boldsymbol{r}}d\boldsymbol{k}+\text{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I}\int\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|\mu_{0}v_{g}(\omega_{I\boldsymbol{k}})}{16\pi^{3}}} a_{I}(\boldsymbol{k},t)s_{I}\boldsymbol{\hat{e}}_{I}(\boldsymbol{k})e^{i\boldsymbol{k}\cdot\boldsymbol{r}}d\boldsymbol{k}+\text{H.c.}, \end{aligned}$$
where the dispersion relation is given by  (A21), although $\boldsymbol {k}$ now varies continuously, and the commutation relations  (A16) and the form of the Hamiltonian  (A18) are as they were in the limit of no dispersion.

Appendix B: Group Velocity

In this appendix we confirm the expression (73) for the channel mode group velocity. Temporarily introducing

$$\begin{aligned} & \mathcal{B}_{Ik}(\boldsymbol{r})=\boldsymbol{b}_{Ik}(y,z)e^{ikx},\\ & \mathcal{D}_{Ik}(\boldsymbol{r})=\boldsymbol{d}_{Ik}(y,z)e^{ikx}, \end{aligned}$$
we begin with the master equation (recall the first of (66)), which we can write as
$$\boldsymbol{\nabla}\times\left(\frac{\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right) =\frac{\omega_{Ik}^{2}}{c^{2}}\mathcal{B}_{Ik}(\boldsymbol{r}),$$
and take its derivative with respect to $k$. Using
$$\begin{aligned} & \frac{\partial\mathcal{B}_{Ik}(\boldsymbol{r})}{\partial k}=\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}+ix\mathcal{B}_{Ik}(\boldsymbol{r}),\\ & \frac{\partial}{\partial k}\left(\frac{1}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)=\frac{\partial\omega_{Ik}}{\partial k}\frac{\partial}{\partial\omega_{Ik}}\left(\frac{1}{\varepsilon_{1}(y,z;\omega_{Ik})}\right) \end{aligned}$$
and putting
$$ v_{Ik}\equiv\frac{\partial\omega_{Ik}}{\partial k}, $$
the group velocity of the mode at $k$ including both modal and material dispersion effects, we find
$$\begin{aligned} & \boldsymbol{\nabla}\times\left(\boldsymbol{C}_{1}(\boldsymbol{r})+\boldsymbol{C}_{2}(\boldsymbol{r})+v_{Ik}\boldsymbol{C}_{3}(\boldsymbol{r)}\right)\\ & =\frac{2\omega_{Ik}}{c^{2}}v_{Ik}\mathcal{B}_{Ik}(\boldsymbol{r})+\frac{\omega_{Ik}^{2}}{c^{2}}\left(\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}+ix\mathcal{B}_{Ik}(\boldsymbol{r})\right),\end{aligned}$$
where
$$\begin{aligned} & \boldsymbol{C}_{1}(\boldsymbol{r})=\frac{\boldsymbol{\nabla}\times\left(\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}\right)}{\varepsilon_{1}(y,z;\omega_{Ik})},\\ & \boldsymbol{C}_{2}(\boldsymbol{r})=\frac{\boldsymbol{\nabla}\times\left(ix\mathcal{B}_{Ik}(\boldsymbol{r})\right)}{\varepsilon_{1} (y,z;\omega_{Ik})}=\frac{i\boldsymbol{\hat{x}}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})} +\frac{ix\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})},\\ & \boldsymbol{C}_{3}(\boldsymbol{r})=\frac{\partial}{\partial\omega_{Ik}}\left(\frac{1}{\varepsilon_{1}(y,z;\omega_{Ik})}\right) \boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r}). \end{aligned}$$
Now write the curl of $\boldsymbol {C}_{2}(\boldsymbol {r})$ as
$$\boldsymbol{\nabla}\times\boldsymbol{C}_{2}(\boldsymbol{r})=\boldsymbol{\nabla}\times\boldsymbol{C}_{4}(\boldsymbol{r}) +ix\boldsymbol{\nabla}\times\left(\frac{\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right) +\boldsymbol{C}_{6}(\boldsymbol{r}),$$
where
$$\begin{aligned} & \boldsymbol{C}_{4}(\boldsymbol{r})=\frac{i\boldsymbol{\hat{x}}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})},\\ & \boldsymbol{C}_{6}(\boldsymbol{r})=i\hat{\boldsymbol{x}}\times\left(\frac{\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1} (y,z;\omega_{Ik})}\right). \end{aligned}$$
Using the expression  (B7) for $\boldsymbol {\nabla }\times \boldsymbol {C}_{2}(\boldsymbol {r})$ in  (B5), we see that the terms proportional to $x$ cancel out by virtue of the master equation  (B2). Further, we can combine the $\boldsymbol {\nabla }\times \boldsymbol {C}_{1}(\boldsymbol {r})$ term on the left-hand side of  (B5) with the term involving $\partial \boldsymbol {b}_{Ik}(y,z)/\partial k$ on the right-hand side of that equation to write  (B5) as
$$\begin{aligned} & \left(\mathcal{\overline{M}}-\frac{\omega_{Ik}^{2}}{c^{2}}\right)\left(\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}\right)+\boldsymbol{\nabla}\times\boldsymbol{C}_{4}(\boldsymbol{r})+\boldsymbol{C}_{6}(\boldsymbol{r})\\ & =\frac{2\omega_{Ik}}{c^{2}}v_{Ik}\mathcal{B}_{Ik}(\boldsymbol{r})-v_{Ik}\boldsymbol{\nabla}\times\boldsymbol{C}_{3}(\boldsymbol{r}), \end{aligned}$$
where we have introduced the Hermitian operator $\mathcal {\overline {M}}$,
$$\mathcal{\overline{M}}\boldsymbol{g}(\boldsymbol{r})\equiv\boldsymbol{\nabla}\times\left(\frac{\boldsymbol{\nabla}\times\boldsymbol{g} (\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)$$
for any vector function $\boldsymbol {g}(\boldsymbol {r}).$

Now choose a particular $k$, and introduce any length $L$. The derivation presented here can be easily extended to a photonic crystal structure as well, where instead of $\boldsymbol {b}_{Ik}(y,z)$ we would have $\boldsymbol {b}_{Ik}(x;y,z),$ with $\boldsymbol {b}_{Ik}(x;y,z)=\boldsymbol {b}_{Ik}(x+a;y,z)$, where $a$ is the length of the unit cell; in such a case, we would take $L=a$, but in a channel waveguide the choice of $L$ is irrelevant. Then take the scalar product of each side of  (B9) into $\mathcal {B}_{Ik}^{*}(\boldsymbol {r})$ and integrate over all $y$ and $z$, and over $x$ from $-L/2$ to $L/2.$ From the first term on the left-hand side of  (B9), we find

$$\begin{aligned} & \int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\left(\mathcal{\overline{M}}-\frac{\omega_{Ik}^{2}}{c^{2}}\right)\left(\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}\right)\\ & =\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\left(\mathcal{\overline{M}}-\frac{\omega_{Ik}^{2}}{c^{2}}\right)\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\right)\cdot\left(\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}\right)\\ & =0, \end{aligned}$$
where in the second line we have partially integrated and used the fact that the terms in the integrand are the same at $z=-L/2$ and  $z=L/2$, and in the third line we have used the master Eq. (B2). From the second term on the left-hand side of  (B9) we find
$$\begin{aligned} & \int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\boldsymbol{\nabla}\times\boldsymbol{C}_{4}(\boldsymbol{r})\\ & =i\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\boldsymbol{\nabla}\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\right).\left(\frac{\boldsymbol{\hat{x}}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)\\ & ={-}\mu_{0}\omega_{Ik}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{D}_{Ik}^{*}(\boldsymbol{r}).\left(\frac{\boldsymbol{\hat{x}}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)\\ & =\mu_{0}\omega_{Ik}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\hat{x}}\cdot\left(\frac{\mathcal{D}_{Ik}^{*}(\boldsymbol{r})\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right). \end{aligned}$$
In the second line we have used the result
$$\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{U}(\boldsymbol{r})\cdot\left(\boldsymbol{\nabla}\times\boldsymbol{V}(\boldsymbol{r})\right)=\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{V}(\boldsymbol{r})\cdot\left(\boldsymbol{\nabla}\times\boldsymbol{U}(\boldsymbol{r})\right),$$
which holds for functions $\boldsymbol {U}(\boldsymbol {r})$ and $\boldsymbol {V}(\boldsymbol {r})$ that vanish at the infinite limits of integration for $\textrm{d}y$ and $\textrm{d}z$, and such that they yield no net contribution from the remaining limit of integration (for $\textrm{d}x$); in the third line we have used the result
$$\mathcal{D}_{Ik}(\boldsymbol{r})=\frac{i}{\mu_{0}\omega_{Ik}}\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r)},$$
which holds because $\mathcal {D}_{Ik}(\boldsymbol {r})\exp (-i\omega _{Ik}t)$ and $\mathcal {B}_{Ik}(\boldsymbol {r})\exp (-i\omega _{Ik}t)$ satisfy the linear Maxwell equations; in the fourth line we have used
$$\boldsymbol{u}\cdot\left(\boldsymbol{v}\times\boldsymbol{w}\right)=\boldsymbol{v}\cdot\left(\boldsymbol{w}\times\boldsymbol{u}\right),$$
for any three vectors $\boldsymbol {u}$, $\boldsymbol {v}$, and $\boldsymbol {w}$. Finally, from the last term on the right-hand side of (B9) we find
$$\begin{aligned} & \int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\boldsymbol{C}_{6}(\boldsymbol{r})\\ & =i\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\hat{x}}\cdot\left(\left(\frac{\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right) \times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\right)\\ & =\mu_{0}\omega_{Ik}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\hat{x}}\cdot\left(\frac{\mathcal{D}_{Ik}(\boldsymbol{r})\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right), \end{aligned}$$
where in the second line we have used (B15) and in the third line we have used (B14). Thus, from (B9) in all we have
$$\begin{aligned} & \mu_{0}\omega_{Ik}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\hat{x}}\cdot\left(\frac{\mathcal{D}_{Ik}(\boldsymbol{r})\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})+\mathcal{D}_{Ik}^{*} (\boldsymbol{r})\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)\\ & =v_{Ik}\mathcal{T}, \end{aligned}$$
where
$$\begin{aligned} & \mathcal{T}=\frac{2\omega_{Ik}}{c^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{B}_{Ik}(\boldsymbol{r})\\ & -\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\nabla\times\boldsymbol{C}_{3}(\boldsymbol{r}). \end{aligned}$$
Look first at the first term on the right-hand side. As $\mathcal {D}_{Ik}(\boldsymbol {r})\exp (-i\omega _{Ik}t)$ and $\mathcal {B}_{Ik}(\boldsymbol {r})\exp (-i\omega _{Ik}t)$ satisfy the linear Maxwell equations, from Faraday’s law we have
$$\mathcal{B}_{Ik}(\boldsymbol{r})=\frac{1}{i\omega_{Ik}}\boldsymbol{\nabla}\times\left(\frac{\mathcal{D}_{Ik}(\boldsymbol{r})}{\epsilon_{0} \varepsilon_{1}(y,z;\omega_{Ik})}\right),$$
so
$$\begin{aligned} & \frac{2\omega_{Ik}}{c^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{B}_{Ik}(\boldsymbol{r})\\ & =\frac{2}{ic^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\left(\boldsymbol{\nabla}\times\left(\frac{\mathcal{D}_{Ik} (\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\right)\right)\\ & =\frac{2}{ic^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\boldsymbol{\nabla}\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\right)\cdot\left(\frac{\mathcal{D}_{Ik} (\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\right)\\ & =\frac{2\mu_{0}\omega_{Ik}}{c^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\frac{\mathcal{D}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{D}_{Ik}(\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\right), \end{aligned}$$
where in the second expression on the right-hand side we have used  (B13) and in the last expression we have used  (B14). Looking at the second term on the right-hand side of  (B18) we have
$$\begin{aligned} & -\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\nabla\times\boldsymbol{C}_{3}(\boldsymbol{r})\\ & ={-}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\nabla\times\left(\frac{\partial}{\partial\omega_{Ik}}\left(\frac{1}{\varepsilon_{1} (y,z;\omega_{Ik})}\right)\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})\right)\\ & ={-}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\nabla}\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\left(\frac{\partial}{\partial\omega_{Ik}} \left(\frac{1}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})\right)\\ & ={-}\mu_{0}^{2}\omega_{Ik}^{2}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\mathcal{D}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{D}_{Ik}(\boldsymbol{r})\right)\frac{\partial}{\partial\omega_{Ik}} \left(\frac{1}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)\\ & ={-}\frac{2\mu_{0}\omega_{Ik}}{c^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\mathcal{D}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{D}_{Ik}(\boldsymbol{r})\right)\frac{\omega_{Ik}}{2} \frac{\partial}{\partial\omega_{Ik}}\left(\frac{1}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\right) \end{aligned}$$
or
$$\begin{aligned} & -\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\nabla\times\boldsymbol{C}_{3}(\boldsymbol{r})\\ &\quad ={-}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\nabla}\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\left(\frac{\partial}{\partial\omega_{Ik}}\left(\frac{1}{\varepsilon_{1} (y,z;\omega_{Ik})}\right)\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})\right)\\ & ={-}\frac{2\mu_{0}\omega_{Ik}}{c^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\mathcal{D}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{D}_{Ik}(\boldsymbol{r})\right)\frac{\omega_{Ik}}{2}\frac{\partial}{\partial\omega_{Ik}} \left(\frac{1}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\right), \end{aligned}$$
where in the first term on the right-hand side we have used (B13), and in the second we have used (B14). Combining (B20) and (B22) in (B18) we find
$$\mathcal{T}=\frac{2\mu_{0}\omega_{Ik}}{c^{2}}L,$$
where we have used (B1) and (70). Using this in (B17), and using (B1) and (74) we indeed find (73).

Appendix C: Interaction Picture

In this appendix, we show that it is always possible to go to an interaction picture where the terms proportional to $a_i^\dagger a_j$ in (188) can be eliminated if the coefficients $\Delta _{i,j}$ are time independent. We split the Hamiltonian as

$$H(t)= \underbrace{ \hbar \sum_{i,j =1}^\ell \Delta_{i,j} a_i^\dagger a_j}_{{\equiv} H_0} + \underbrace{\hbar i \frac12 \sum_{k,l=1}^\ell \left[ \zeta_{k,l}(t) a_k^{{\dagger} } a_{l}^{{\dagger}} -\text{H.c.} \right] }_{{\equiv} V(t)},$$
To study the dynamics of the problem let us derive the equation of motion obeyed by a modified time evolution operator $\mathcal { U}_I$ related to the original time evolution operator $\mathcal { U}$ via
$$\mathcal{ U}_I(t,t_0)\equiv e^{i H_0 (t-t_0)/\hbar} \mathcal{ U}(t,t_0).$$
Using the Schrödinger equation for $\mathcal { U}$ in (175) we find that
$$i \hbar \frac{\textrm{d}}{\textrm{d}t} \mathcal{ U}_I(t,t_0)=\left( e^{i H_0(t-t_0)/\hbar} V(t) e^{{-}i H_0(t-t_0)/\hbar} \right) \mathcal{ U}_I(t,t_0)=V_I(t) \mathcal{ U}_I(t,t_0).$$
In the last line we introduced the interaction picture Hamiltonian $V_I(t)$. For the explicit form of our Hamiltonian we can first obtain
$$e^{i H_0(t-t_0)/\hbar} \boldsymbol{a} e^{{-}i H_0(t-t_0)/\hbar} = \exp\left( i \boldsymbol{\Delta} (t-t_0) \right) \boldsymbol{a} = \boldsymbol{U}(t)\boldsymbol{a},$$
where $\boldsymbol{U}(t)$ is a time-dependent unitary matrix. With this solution we can write
$$\begin{aligned}V_I(t) = e^{i H_0(t-t_0)/\hbar} V(t) e^{{-}i H_0(t-t_0)/\hbar} & =\hbar i \frac12 \sum_{k,l=1}^\ell \left[ \zeta_{k,l}(t) e^{i H_0(t-t_0)/\hbar} a_k^{{\dagger} } a_{l}^{{\dagger}} e^{{-}i H_0(t-t_0)/\hbar} -\text{H.c.} \right] \end{aligned}$$
$$\begin{aligned}& =\hbar i \frac12 \sum_{i,j,k,l=1}^\ell \left[ \zeta_{k,l}(t) U_{l,i}^*(t) a_i^\dagger U_{k,j}^*(t)a_j^\dagger{-}\text{H.c.} \right] \end{aligned}$$
$$\begin{aligned}& =\hbar i \frac12 \sum_{i,j=1}^\ell \left[ \tilde{\zeta}_{i,j}(t) a_i^\dagger a_j^\dagger{-}\text{H.c.} \right], \end{aligned}$$
where we introduced the symmetric matrix $\tilde {\boldsymbol{\zeta }}(t) = \boldsymbol{U}^\dagger (t) \boldsymbol{\zeta }(t) \boldsymbol{U}^*(t)$. This completes the proof, because now we need to solve for the interaction picture evolution operator which is solely driven by terms of the form $a_i^\dagger a_j^\dagger$.

Appendix D: Symplectic Group

In Sec. 3.4 we showed that the solution of the Heisenberg equations of motion had the form (209), and respected the equal time commutation relations. We can write the canonical commutation relation of the ladder operators more succinctly as

$$[z_i, z_j^\dagger] = Z_{i,j}, \quad \boldsymbol{z} = \begin{pmatrix} \boldsymbol{a} \\ \boldsymbol{a}^\dagger \end{pmatrix}, \quad \boldsymbol{Z} = \begin{pmatrix} \mathbb{I}_\ell & \boldsymbol{0} \\ \boldsymbol{0} & -\mathbb{I}_\ell \end{pmatrix},$$
and then the fact that the Heisenberg transformation induced by $\boldsymbol{K}$ respects these commutation relations is equivalent to
$$\boldsymbol{K} \boldsymbol{Z} \boldsymbol{K}^\dagger{=} \boldsymbol{Z}.$$
To connect this equation to the usual definition of the symplectic group one introduces quadratures
$$\begin{aligned} \boldsymbol{r} \equiv \begin{pmatrix} \boldsymbol{q} \\ \boldsymbol{p} \end{pmatrix} = \sqrt{\hbar}\frac{1}{\sqrt{2}} \begin{pmatrix} \mathbb{I}_\ell & \mathbb{I}_\ell \\ -i \mathbb{I}_\ell & i \mathbb{I}_\ell \end{pmatrix} \begin{pmatrix} \boldsymbol{a} \\ \boldsymbol{a}^\dagger \end{pmatrix} = \sqrt{\hbar}\boldsymbol{R}\boldsymbol{z}. \end{aligned}$$
The canonical commutation relation for the operators in $\boldsymbol{r}$ follows from that in $\boldsymbol{z}$ as
$$[r_i,r_j] = i \hbar \Omega_{i,j} \quad \boldsymbol{\Omega} = \begin{pmatrix} 0 & \mathbb{I}_\ell \\ -\mathbb{I}_\ell & \boldsymbol{0} \end{pmatrix}.$$
Now it is easy to state the transformation for quadratures
$$\boldsymbol{z} \to \boldsymbol{z'} = \boldsymbol{K} \boldsymbol{z} \Longleftrightarrow \boldsymbol{r} \to \boldsymbol{r'} = \boldsymbol{S} \boldsymbol{r},$$
where
$$\begin{aligned} \boldsymbol{S} = \boldsymbol{R} \boldsymbol{K} \boldsymbol{R}^\dagger{=} \begin{pmatrix} \Re(\boldsymbol{V}) + \Re(\boldsymbol{W}) & \Im(\boldsymbol{W}) -\Im(\boldsymbol{V}) \\ \Im(\boldsymbol{V}) + \Im(\boldsymbol{W}) & \Re(\boldsymbol{V}) - \Re(\boldsymbol{W}) \end{pmatrix}. \end{aligned}$$
The condition in  D2 implies that the matrix $\boldsymbol{S}$ needs to satisfy
$$\boldsymbol{S} \boldsymbol{\Omega} \boldsymbol{S}^T = \boldsymbol{\Omega},$$
which implies that $\boldsymbol{S}$ is a symplectic matrix [5,6467].

Appendix E: Form of the Evolution Operator in the High-Gain Regime

At the end of Sec. 3.5 we argued that the evolution operator associated with a general quadratic Hamiltonian can be written as the product of two exponentials

$$\mathcal{U} = \mathcal{U}_1 \mathcal{U}_0, \quad \mathcal{U}_1 = \exp\left[ \frac12 \sum_{k,l=1}^\ell J_{k,l} a^\dagger_{k} a^\dagger_l - \text{H.c.}\right],$$
where $\mathcal {U}_0$ is a unitary operator that satisfies $\mathcal {U}_0 | {\text {vac}}\rangle = | {\text {vac}}\rangle$. In this appendix we determine the form of the generator of $\mathcal {U}_0$. To pin down $\mathcal {U}_0$ we calculate
$$\mathcal{U}_0^\dagger \mathcal{U}_1^\dagger \boldsymbol{a} \mathcal{U}_1 \mathcal{U}_0 = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \cosh r_\lambda \right] \boldsymbol{F}^* \left\{ \mathcal{U}_0^\dagger \boldsymbol{a} \mathcal{U}_0 \right\} + \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \sinh r_\lambda \right] \boldsymbol{F} \left\{ \mathcal{U}_0^\dagger \boldsymbol{a}^\dagger \mathcal{U}_0 \right\},$$
and compare with the solution in (216) which tells us that we want
$$\boldsymbol{F}^* \left\{ \mathcal{U}_0^\dagger \boldsymbol{a} \mathcal{U}_0 \right\} = \boldsymbol{G} \boldsymbol{a} \Longleftrightarrow \boldsymbol{G}^\dagger\boldsymbol{F}^* \left\{ \mathcal{U}_0^\dagger \boldsymbol{a} \mathcal{U}_0 \right\} = \boldsymbol{a},$$
To satisfy this constraint we recall the results from Appendix C and write
$$\mathcal{U}_0 = \exp\left[{-}i \sum_{j,k} \phi_{j,k}a^\dagger_j a_k \right],$$
such that
$$\mathcal{U}_0^\dagger \boldsymbol{a} \mathcal{U}_0 = \exp\left[ - i\boldsymbol{\phi} \right] \boldsymbol{a},$$
and comparing the two equations above tells us that we want $\boldsymbol{G}^\dagger \boldsymbol{F}^* \exp \left [ - i\boldsymbol{\phi } \right ] = \mathbb {I}_{\ell }$ or, equivalently, $\boldsymbol{\phi } = -i \log _e \boldsymbol{G}^\dagger \boldsymbol{F}^*$. Note that since both $\boldsymbol{G}^\dagger$ and $\boldsymbol{F}$ are unitaries their logarithm is $i$ times a Hermitian matrix which upon multiplication by $(-i)$ gives that $\boldsymbol{\phi }$ is a purely Hermitian matrix as it should. Also note that as expected $\mathcal {U}_0$ in (E4) has the vacuum as an eigenket with eigenvalue 1. It is interesting to note that the Heisenberg transformation generated by a general quadratic Hamiltonian requires the exponentiation of two generators. This happens because the exponential map from the Lie algebra $\textbf {sp}(2n, \mathbb {R})$ to the symplectic group Sp$(2n,\mathbb {R})$ is not surjective. However, any element of the group may be generated by the group multiplication of two elements of the group.

Appendix F: Waveguide Hamiltonians

Concerning waveguides in which a second-order nonlinearity is dominant, the process with $2k_{S}=k_{P}$ leads to

$$H_{\text{DSV}}^{\text{SPDC}}={-}\frac{\zeta_{\text{chan}}^{SSP}}{2}\int\textrm{d}x\,\psi_{S}^{{\dagger}}\left(x\right)\psi_{S}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)+\textrm{H.c.},$$
and the process with $k_{S}+k_{I}=k_{P}$ leads to
$$H_{\text{NDSV}}^{\text{SPDC}}={-}\zeta_{\text{chan}}^{SIP}\int\textrm{d}x\,\psi_{S}^{{\dagger}}\left(x\right)\psi_{I}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)+\textrm{H.c.},$$
where
$$\begin{aligned}\zeta_{\text{chan}}^{J_{1}J_{2}J_{3}} & =\frac{2}{\epsilon_{0}}\sqrt{\frac{\hbar^{3}\omega_{J_{1}}\omega_{J_{2}}\omega_{J_{3}}}{2^{3}}}\int\textrm{d}y\textrm{d}z\,\Gamma_{2}^{ijk}\left(\boldsymbol{r}\right)\left[d_{J_{1}k_{J_{1}}}^{i}\left(y,z\right)d_{J_{2}k_{J_{2}}}^{j}\left(y,z\right)\right]^{*} d_{J_{3}k_{J_{3}}}^{k}\left(y,z\right)\\ & =\hbar\sqrt{\frac{\hbar\omega_{J_{1}}\omega_{J_{2}}\omega_{J_{3}}}{2\epsilon_{0}\overline{n}_{J_{1}}\overline{n}_{J_{2}}\overline{n}_{J_{3}} A_{\text{chan}}^{J_{1}J_{2}J_{3}}}}\overline{\chi}_{2}, \end{aligned}$$
with
$$\frac{1}{\sqrt{A_{\text{chan}}^{J_{1}J_{2}J_{3}}}}=\frac{\int\textrm{d}y\textrm{d}z\frac{\chi_{2}^{ijk}\left(y,z\right)}{\overline{\chi}_{2}}\left[e_{J_{1}}^{i}\left(y,z\right)e_{J_{2}}^{j}\left(y,z\right)\right]^{*}e_{J_{3}}^{k}\left(y,z\right)}{\mathcal{N}_{J_{1}}\mathcal{N}_{J_{2}}\mathcal{N}_{J_{3}}\sqrt{c^{3}/\left(v_{J_{1}}v_{J_{2}}v_{J_{3}}\right)}},$$
and we have used (94), (138), and (307).

Concerning waveguides in which a third-order nonlinearity is dominant, the SP DSV process leads to

$$H_{\text{DSV}}^{\text{SP-SFWM}}={-}\frac{\gamma_{\text{chan}}^{PPPP}}{2}\hbar^{2}\omega_{P}v_{P}^{2}\int\textrm{d}x\,\psi_{P}^{{\dagger}}\left(x\right)\psi_{P}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)\psi_{P}\left(x\right),$$
the process with $k_{S}+k_{I}=2k_{P}$ leads to (303), the process with $2k_{S}=k_{P_{1}}+k_{P_{2}}$ leads to
$$\begin{aligned} H_{\text{DSV}}^{\text{DP-SFWM}} & ={-}\gamma_{\text{chan}}^{SSP_{1}P_{2}}\hbar^{2}\overline{\omega}_{SSP_{1}P_{2}}\overline{v}_{SSP_{1}P_{2}}^{2}\int\textrm{d}x\,\psi_{S}^{{\dagger}}\left(x\right)\psi_{S}^{{\dagger}}\left(x\right)\psi_{P_{1}}\left(x\right)\psi_{P_{2}}\left(x\right)+\text{H.c.}\\ & \quad-\frac{\hbar^{2}}{2}\sum_{J}\gamma_{\text{chan}}^{JJJJ}\omega_{J}v_{J}^{2}\int\textrm{d}x\,\psi_{J}^{{\dagger}}\left(x\right)\psi_{J}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)\psi_{J}\left(x\right)\\ & \quad-2\hbar^{2}\sum_{J\neq S}\gamma_{\text{chan}}^{JSJS}\sqrt{\omega_{J}\omega_{S}}v_{J}v_{S}\int\textrm{d}x\,\psi_{J}^{{\dagger}}\left(x\right)\psi_{S}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)\psi_{S}\left(x\right)\\ & \quad-2\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar^{2}\sqrt{\omega_{P_{1}}\omega_{P_{2}}}v_{P_{1}}v_{P_{2}}\int\textrm{d}x\,\psi_{P_{1}}^{{\dagger}}\left(x\right)\psi_{P_{2}}^{{\dagger}}\left(x\right)\psi_{P_{1}}\left(x\right)\psi_{P_{2}}\left(x\right), \end{aligned}$$
and the process with $k_{S}+k_{I}=k_{P_{1}}+k_{P_{2}}$ leads to
$$\begin{aligned} H_{\text{NDSV}}^{\text{DP-SFWM}} & ={-}2\gamma_{\text{chan}}^{SIP_{1}P_{2}}\hbar^{2}\overline{\omega}_{SIP_{1}P_{2}}\overline{v}_{SIP_{1}P_{2}}^{2}\int\textrm{d}x\,\psi_{S}^{{\dagger}}\left(x\right)\psi_{I}^{{\dagger}}\left(x\right)\psi_{P_{1}}\left(x\right)\psi_{P_{2}}\left(x\right)+\textrm{H.c.}\\ & \quad-\frac{\hbar^{2}}{2}\sum_{J}\gamma_{\text{chan}}^{JJJJ}\omega_{J}v_{J}^{2}\int\textrm{d}x\,\psi_{J}^{{\dagger}}\left(x\right)\psi_{J}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)\psi_{J}\left(x\right)\\ & \quad-2\hbar^{2}\sum_{J\neq S}\gamma_{\text{chan}}^{JSJS}\sqrt{\omega_{J}\omega_{S}}v_{J}v_{S}\int\textrm{d}x\,\psi_{J}^{{\dagger}}\left(x\right)\psi_{S}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)\psi_{S}\left(x\right)\\ & \quad-2\hbar^{2}\sum_{J\neq I}\gamma_{\text{chan}}^{JIJI}\sqrt{\omega_{J}\omega_{I}}v_{J}v_{I}\int\textrm{d}x\,\psi_{J}^{{\dagger}}\left(x\right)\psi_{I}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)\psi_{I}\left(x\right)\\ & \quad-2\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar^{2}\sqrt{\omega_{P_{1}}\omega_{P_{2}}}v_{P_{1}}v_{P_{2}}\int\textrm{d}x\,\psi_{P_{1}}^{{\dagger}}\left(x\right)\psi_{P_{2}}^{{\dagger}}\left(x\right)\psi_{P_{1}}\left(x\right)\psi_{P_{2}}\left(x\right), \end{aligned}$$
where we have used (304) and (306).

Appendix G: Heisenberg Equations of Motion for Waveguides

Concerning waveguides in which a second-order nonlinearity is dominant, and using (310) along with $2\omega _{S}=\omega _{P}$, the Heisenberg equations of motion yield

$$\begin{aligned} \left(\frac{\partial}{\partial t}+v_{P}\frac{\partial}{\partial x}-i\frac{v_{P}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle & =0,\\ \left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right) & =\frac{i}{\hbar}\zeta_{\text{chan}}^{SSP}\left(x\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \overline{\psi}_{S}^{{\dagger}}\left(x,t\right), \end{aligned}$$
for the process with $2k_{S}=k_{P}$, and, using $\omega _{S}+\omega _{I}=\omega _{P}$
$$\begin{aligned} \left(\frac{\partial}{\partial t}+v_{P}\frac{\partial}{\partial x}-i\frac{v_{P}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle & =0,\\ \left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right) & =\frac{i}{\hbar}\zeta_{\text{chan}}^{SIP}\left(x\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \overline{\psi}_{I}^{{\dagger}}\left(x,t\right),\\ \left(\frac{\partial}{\partial t}+v_{I}\frac{\partial}{\partial x}-i\frac{v_{I}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{I}\left(x,t\right) & =\frac{i}{\hbar}\zeta_{\text{chan}}^{SIP}\left(x\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \overline{\psi}_{S}^{{\dagger}}\left(x,t\right), \end{aligned}$$
for the process with $k_{S}+k_{I}=k_{P}$. Note how the second of (G1) is of the form of (315), and the second and third of (G2) are of the form of (314).

Concerning waveguides in which a third-order nonlinearity is dominant, and again using (310), the Heisenberg equations of motion yield

$$\begin{aligned} \left(\frac{\partial}{\partial t}+v_{P}\frac{\partial}{\partial x}-i\frac{v_{P}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle & =i\gamma_{\text{chan}}^{PPPP}\hbar\omega_{P}v_{P}^{2}\left|\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle,\\ \left(\frac{\partial}{\partial t}+v_{P}\frac{\partial}{\partial x}-i\frac{v_{P}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\delta\overline{\psi}_{P}\left(x,t\right) & =i\gamma_{\text{chan}}^{PPPP}\hbar\omega_{P}v_{P}^{2}\left[\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle ^{2}\delta\overline{\psi}_{P}^{{\dagger}}\left(x,t\right)\right.\\&\left.+2\left|\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \right|^{2}\delta\overline{\psi}_{P}\left(x,t\right)\right], \end{aligned}$$
where we have put $\overline {\psi }_{P}\left (x,t\right )\rightarrow \left \langle \overline {\psi }_{P}\left (x,t\right )\right \rangle +\delta \overline {\psi }_{P}\left (x,t\right )$, for the SP DSV process. Similarly, using $\omega _{S}+\omega _{I}=2\omega _{P}$, they yield (311) and (312) for the process with $k_{S}+k_{I}=2k_{P}$, using $2\omega _{S}=\omega _{P_{1}}+\omega _{P_{2}}$ they yield
$$\begin{aligned} &\left(\frac{\partial}{\partial t}+v_{P_{1}}\frac{\partial}{\partial x}-i\frac{v_{P_{1}}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \\&\quad =i\gamma_{\text{chan}}^{P_{1}P_{1}P_{1}P_{1}}\hbar\omega_{P_{1}}v_{P_{1}}^{2}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar\sqrt{\omega_{P_{1}}\omega_{P_{2}}}v_{P_{1}}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle,\\ &\left(\frac{\partial}{\partial t}+v_{P_{2}}\frac{\partial}{\partial x}-i\frac{v_{P_{2}}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \\& \quad=i\gamma_{\text{chan}}^{P_{2}P_{2}P_{2}P_{2}}\hbar\omega_{P_{2}}v_{P_{2}}^{2}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar\sqrt{\omega_{P_{1}}\omega_{P_{2}}}v_{P_{1}}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle,\\ &\left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right)\\ &\quad =2i\gamma_{\text{chan}}^{SSP_{1}P_{2}}\hbar\overline{\omega}_{SSP_{1}P_{2}}\bar{v}_{SSP_{1}P_{2}}^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \overline{\psi}_{S}^{{\dagger}}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}SP_{1}S}\hbar\sqrt{\omega_{P_{1}}\omega_{S}}v_{S}v_{P_{1}}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{S}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{P_{2}SP_{2}S}\hbar\sqrt{\omega_{P_{2}}\omega_{S}}v_{S}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{S}\left(x,t\right), \end{aligned}$$
for the process with $2k_{S}=k_{P_{1}}+k_{P_{2}}$, and using $\omega _{S}+\omega _{I}=\omega _{P_{1}}+\omega _{P_{2}}$ they yield
$$\begin{aligned} &\left(\frac{\partial}{\partial t}+v_{P_{1}}\frac{\partial}{\partial x}-i\frac{v_{P_{1}}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \\&\quad =i\gamma_{\text{chan}}^{P_{1}P_{1}P_{1}P_{1}}\hbar\omega_{P_{1}}v_{P_{1}}^{2}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar\overline{\omega}_{P_{1}P_{2}P_{1}P_{2}}v_{P_{1}}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle,\\ &\left(\frac{\partial}{\partial t}+v_{P_{2}}\frac{\partial}{\partial x}-i\frac{v_{P_{2}}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \\&\quad =i\gamma_{\text{chan}}^{P_{2}P_{2}P_{2}P_{2}}\hbar\omega_{P_{2}}v_{P_{2}}^{2}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar\overline{\omega}_{P_{1}P_{2}P_{1}P_{2}}v_{P_{1}}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle,\\ &\left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right)\\ &\quad =2i\gamma_{\text{chan}}^{SIP_{1}P_{2}}\hbar\overline{\omega}_{SIP_{1}P_{2}}\bar{v}_{SIP_{1}P_{2}}^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \overline{\psi}_{I}^{{\dagger}}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{SP_{1}SP_{1}}\hbar\overline{\omega}_{SP_{1}SP_{1}}v_{S}v_{P_{1}}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{S}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{SP_{2}SP_{2}}\hbar\overline{\omega}_{SP_{2}SP_{2}}v_{S}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{S}\left(x,t\right),\\ &\left(\frac{\partial}{\partial t}+v_{I}\frac{\partial}{\partial x}-i\frac{v_{I}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{I}\left(x,t\right)\\ &\quad =2i\gamma_{\text{chan}}^{SIP_{1}P_{2}}\hbar\overline{\omega}_{SIP_{1}P_{2}}\bar{v}_{SIP_{1}P_{2}}^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \overline{\psi}_{S}^{{\dagger}}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}IP_{1}I}\hbar\overline{\omega}_{P_{1}IP_{1}I}v_{P_{1}}v_{I}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{I}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{P_{2}IP_{2}I}\hbar\overline{\omega}_{SP_{2}SP_{2}}v_{P_{2}}v_{I}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{I}\left(x,t\right), \end{aligned}$$
for the process with $k_{S}+k_{I}=k_{P_{1}}+k_{P_{2}}$. Note how the second of  (G3) and the third of  (G4) are both of the form of (315), and the third and fourth of  (G5) are of the form of (314).

Appendix H: Obtaining the Matrix Product Form of an Arbitrary Tensor

Consider an arbitrary tensor with $\ell$ indices $R_{i_0\ldots i_{\ell -1}} \in \mathbb {C}^{c_0 \times c_1 \times \cdots \times c_{\ell -1}}$ where the index $i_k$ can take $c_k$ values. In this appendix we show how one can construct its matrix product representation by leveraging the SVD of matrices, which we now recall. For a $c_0 \times c_1$ rectangular matrix with entries $M_{i j}$ one can construct its SVD as

$$M_{ij} = \sum_{k = 0}^{\min\{c_0,c_1\}-1} U_{ik} \lambda_{kk} V_{kj}.$$
In the last equation one can take $\boldsymbol{U}$ to be a $c_0 \times c_0$ unitary matrix, $\boldsymbol{V}$ to be $c_1 \times c_1$ unitary matrix and $\boldsymbol{\lambda }$ to be a $c_0 \times c_1$ diagonal matrix with nonnegative entries. Equivalently, one can take $\boldsymbol{U}$ to be a rectangular $c_0 \times \min \{c_0,c_1\}$ semi-unitary matrix, $\boldsymbol{V}$ to be $\min \{c_0,c_1\} \times D_1$ semi-unitary matrix and $\boldsymbol{\lambda }$ to be $\min \{c_0,c_1\} \times \min \{c_0,c_1\}$ square nonnegative matrix.

A second useful property of a multidimensional array is that we are free to reshape it as a matrix by grouping indices together. Consider, for example, a four-index tensor $R_{i_0 i_1 i_2 i_3}$ where each index takes two possible values, $i_k \in \{0,1\}$. We can group the last three indices $i_1 i_2 i_3$ into a common index that takes $2^3 = 8$ values. This allows us to consider the tensor $R$ as a $2 \times 8$ matrix. We can emphasize this partition by putting a semicolon between the indices we want to separate writing

$$R_{i_0;i_1 i_2 i_3} \in \mathbb{C}^{2 \times 8}.$$
We can instead group the first and last two indices to now obtain a $2^2 \times 2^2 = 4 \times 4$ matrix that we write as
$$R_{i_0 i_1 ; i_2 i_3} \in \mathbb{C}^{4 \times 4}.$$
Having introduced this useful notation and recalled the SVD for matrices we are now ready to decompose an arbitrary tensor.

To this end, we split our arbitrary input tensor as $R_{i_0;i_1\ldots i_{\ell -1}}$ and then write the singular decomposition for this $c_0 \times c_{1:\ell -1}$ (with $c_{1:\ell -1} = c_1 \times \cdots \times c_{\ell -1}$) matrix

$$R_{i_0;i_1\ldots i_{\ell-1}} = \sum_{\alpha_1} U_{i_0; \alpha_1}^{[0]} \lambda^{[0]}_{\alpha_1 ; \alpha_1} V_{\alpha_1;i_1\ldots i_{\ell-1}}^{[0]},$$
where we do not write explicitly the possible values of the dummy index $\alpha _1$. This can be easily inferred by looking at the dimensions of the two index vectors separated by the semicolon on the right-hand side. For this first step in our derivation $\alpha _1$ can take $\min \{c_0, c_1 \times \cdots \times c_{\ell -1}\}$ values. We can now look at the tensor $V_{\alpha _1,i_1\ldots i_{\ell -1}}$ and apply once more the SVDs grouping the first two indices together:
$$V_{\alpha_1,i_1;i_2\ldots i_{\ell-1}}^{[0]} = \sum_{\alpha_2} U_{\alpha_1,i_1;\alpha_2}^{[1]} \lambda^{[1]}_{\alpha_2;\alpha_2} V_{\alpha_2;i_2\ldots i_{\ell-1}}^{[1]},$$
which we can use to write
$$R_{i_0 i_1\ldots i_{\ell-1}} = \sum_{\alpha_1,\alpha_2 } U_{i_0; \alpha_1}^{[0]} \lambda^{[0]}_{\alpha_1 ; \alpha_1} U_{\alpha_1,i_1;\alpha_2}^{[1]} \lambda^{[1]}_{\alpha_2;\alpha_2} V_{\alpha_2;i_2\ldots i_{\ell-1}}^{[1]}.$$
Now note that the leftover higher-rank tensor $V^{[1]}$ has $\ell -1$ indices, one less than our original tensor $R$. We can continue as before and write the SVD of $V^{[1]}$ separating $\alpha _2$ and $i_2$ from the rest of the indices
$$V_{\alpha_2,i_2;i_3\ldots i_{\ell-1}}^{[1]} = \sum_{\alpha_3} U_{\alpha_2,i_2;\alpha_3}^{[2]} \lambda^{[2]}_{\alpha_3;\alpha_3} V_{\alpha_3;i_3\ldots i_{\ell-1}}^{[2]}.$$
We can then plug this decomposition into H6 to obtain
$$R_{i_0 i_1\ldots i_{\ell-1}} = \sum_{\alpha_1,\alpha_2 } U_{i_0; \alpha_1}^{[0]} \lambda^{[0]}_{\alpha_1 ; \alpha_1} U_{\alpha_1,i_1;\alpha_2}^{[1]} \lambda^{[1]}_{\alpha_2;\alpha_2} V_{\alpha_2;i_2\ldots i_{\ell-1}}^{[1]},$$
where we note that the leftover higher-order tensor $V^{[2]}$ has $\ell -2$ indices. By this point it is easy to guess that in the $k$th step of this recursive decomposition, we will have a tensor $V^{[k]}$ with $\ell -k$ indices $\{\alpha _{k+1},i_{k+1}\ldots,i_{\ell -1}\}$ on which we will apply the SVD to obtain
$$V_{\alpha_{k+1},i_{k+1};i_{k+2}\ldots i_{\ell-1}}^{[k]} = \sum_{\alpha_{k+2}} U^{[k+1]}_{\alpha_{k+1},i_{k+1}; \alpha_{k+2}} \lambda^{[k+1]}_{\alpha_{k+2};\alpha_{k+2}} V^{[k+1]}_{\alpha_{k+2};i_{k+2}\ldots i_{\ell-1}}.$$
Based on this iterative decomposition we finally arrive at the first matrix product form of the tensor $R$,
$$R_{i_0 i_1\ldots i_{\ell-1}} = \sum_{\alpha_1,\alpha_2,\ldots, \alpha_{\ell-1}} U_{i_0; \alpha_1}^{[0]} \lambda^{[0]}_{\alpha_1 ; \alpha_1} U_{\alpha_1,i_1;\alpha_2}^{[1]} \lambda^{[1]}_{\alpha_2;\alpha_2} \ldots U_{\alpha_{\ell-2},i_{\ell-2};\alpha_{\ell-1}}^{[1]} \lambda^{[1]}_{\alpha_{\ell-1};\alpha_{\ell-1}} V_{\alpha_{\ell-1},i_{\ell-1}}$$
which is the form used by Yanagimoto et al. [30]. A second useful form can be obtained by introducing $(S^{[k]})^{\alpha _k}_{i_k,\alpha _{k+1}} \equiv U_{\alpha _k,i_k;\alpha _{k+1}}^{[k]} \lambda ^{[k]}_{\alpha _{k+1};\alpha _{k+1}}$ as used in the main text.

Acknowledgments

N. Q. thanks A. M. Brańczyk, M. D Vidrighin, G. Triginer, A. L. Grimsmo, and I. Dhand for useful discussions and acknowledges support from the Ministère de l’Économie et de l’Innovation du Québec. M. L. acknowledges the support from Ministero dell’Istruzione, dell’Università e della Ricerca (Dipartimenti di Eccellenza Program (2018–2022)). J. E. S. acknowledges support from the Natural Sciences and Engineering Research Council of Canada.

Disclosures

The authors declare no conflicts of interest.

Data availability

No data were generated or analyzed in the presented research.

References and Notes

1. S. E. Harris, M. K. Oshman, and R. L. Byer, “Observation of tunable optical parametric fluorescence,” Phys. Rev. Lett. 18, 732–734 (1967). [CrossRef]  

2. D. Magde and H. Mahr, “Study in ammonium dihydrogen phosphate of spontaneous parametric interaction tunable from 4400 to 16 000 Å,” Phys. Rev. Lett. 18, 905–907 (1967). [CrossRef]  

3. A. I. Lvovsky, “Squeezed light,” in Photonics: Scientific Foundations, Technology and Applications (2015), Vol. 1, pp. 121–163. [CrossRef]  

4. U. L. Andersen, T. Gehring, C. Marquardt, and G. Leuchs, “30 years of squeezed light generation,” Phys. Scr. 91, 053001 (2016). [CrossRef]  

5. C. Weedbrook, S. Pirandola, R. García-Patrón, N. J. Cerf, T. C. Ralph, J. H. Shapiro, and S. Lloyd, “Gaussian quantum information,” Rev. Mod. Phys. 84, 621–669 (2012). [CrossRef]  

6. S. L. Braunstein, “Squeezing as an irreducible resource,” Phys. Rev. A 71, 055801 (2005). [CrossRef]  

7. T. Rudolph, “Why I am optimistic about the silicon-photonic route to quantum computing,” APL Photonics 2, 030901 (2017). [CrossRef]  

8. J. E. Bourassa, R. N. Alexander, M. Vasmer, A. Patil, I. Tzitrin, T. Matsuura, D. Su, B. Q. Baragiola, S. Guha, G. Dauphinais, K. K. Sabapathy, N. C. Menicucci, and I. Dhand, “Blueprint for a scalable photonic fault-tolerant quantum computer,” Quantum 5, 392 (2021). [CrossRef]  

9. T. R. Bromley, J. M. Arrazola, S. Jahangiri, J. Izaac, N. Quesada, A. D. Gran, M. Schuld, J. Swinarton, Z. Zabaneh, and N. Killoran, “Applications of near-term photonic quantum computers: software and algorithms,” Quantum Sci. Technol. 5, 034010 (2020). [CrossRef]  

10. M. V. Larsen, C. Chamberland, K. Noh, J. S. Neergaard-Nielsen, and U. L. Andersen, “Fault-tolerant continuous-variable measurement-based quantum computation architecture,” PRX Quantum 2, 030325 (2021). [CrossRef]  

11. M. Kues, C. Reimer, P. Roztocki, L. R. Cortés, S. Sciara, B. Wetzel, Y. Zhang, A. Cino, S. T. Chu, B. E. Little, D. J. Moss, L. Caspani, J. Azaña, and R. Morandotti, “On-chip generation of high-dimensional entangled quantum states and their coherent control,” Nature 546, 622–626 (2017). [CrossRef]  

12. C. Reimer, S. Sciara, P. Roztocki, M. Islam, L. Romero Cortés, Y. Zhang, B. Fischer, S. Loranger, R. Kashyap, A. Cino, S. T. Chu, B. E. Little, D. J. Moss, L. Caspani, W. J. Munro, J. Azaña, M. Kues, and R. Morandotti, “High-dimensional one-way quantum processing implemented on d-level cluster states,” Nat. Phys. 15, 148–153 (2019). [CrossRef]  

13. J. Aasi, J. Abadie, B. Abbott, et al., “Enhanced sensitivity of the LIGO gravitational wave detector by using squeezed states of light,” Nat. Photonics 7, 613–619 (2013). [CrossRef]  

14. E. Polzik, J. Carri, and H. Kimble, “Spectroscopy with squeezed light,” Phys. Rev. Lett. 68, 3020–3023 (1992). [CrossRef]  

15. C. A. Casacio, L. S. Madsen, A. Terrasson, M. Waleed, K. Barnscheidt, B. Hage, M. A. Taylor, and W. P. Bowen, “Quantum-enhanced nonlinear microscopy,” Nature 594, 201–206 (2021). [CrossRef]  

16. P. J. Mosley, J. S. Lundeen, B. J. Smith, P. Wasylczyk, A. B. U’Ren, C. Silberhorn, and I. A. Walmsley, “Heralded generation of ultrafast single photons in pure quantum states,” Phys. Rev. Lett. 100, 133601 (2008). [CrossRef]  

17. H. Takesue and K. Inoue, “Generation of polarization-entangled photon pairs and violation of Bell’s inequality using spontaneous four-wave mixing in a fiber loop,” Phys. Rev. A 70, 031802 (2004). [CrossRef]  

18. G. Harder, V. Ansari, B. Brecht, T. Dirmeier, C. Marquardt, and C. Silberhorn, “An optimized photon pair source for quantum circuits,” Opt. Express 21, 13975–13985 (2013). [CrossRef]  

19. G. Harder, T. J. Bartley, A. E. Lita, S. W. Nam, T. Gerrits, and C. Silberhorn, “Single-mode parametric-down-conversion states with 50 photons as a source for mesoscopic quantum optics,” Phys. Rev. Lett. 116, 143601 (2016). [CrossRef]  

20. M. A. Finger, T. S. Iskhakov, N. Y. Joly, M. V. Chekhova, and P. S. J. Russell, “Raman-free, noble-gas-filled photonic-crystal fiber source for ultrafast, very bright twin-beam squeezed vacuum,” Phys. Rev. Lett. 115, 143602 (2015). [CrossRef]  

21. J.-Y. Chen, Z. Li, Z. Ma, C. Tang, H. Fan, Y. M. Sua, and Y.-P. Huang, “Photon conversion and interaction on chip,”preprint arXiv:2105.00275 (2021).

22. M. Placke and S. Ramelow, “Engineering AlGaAs-on-insulator toward quantum optical applications,” Opt. Lett. 45, 6763–6766 (2020). [CrossRef]  

23. J. B. Surya, X. Guo, C.-L. Zou, and H. X. Tang, “Efficient third-harmonic generation in composite aluminum nitride/silicon nitride microrings,” Optica 5, 103–108 (2018). [CrossRef]  

24. J. Lu, M. Li, C.-L. Zou, A. Al Sayem, and H. X. Tang, “Toward 1% single-photon anharmonicity with periodically poled lithium niobate microring resonators,” Optica 7, 1654–1659 (2020). [CrossRef]  

25. A. Cavanna, J. Hammer, C. Okoth, E. Ortiz-Ricardo, H. Cruz-Ramirez, K. Garay-Palmett, A. B. U’Ren, M. H. Frosz, X. Jiang, N. Y. Joly, and M. V. Chekhova, “Progress toward third-order parametric down-conversion in optical fibers,” Phys. Rev. A 101, 033840 (2020). [CrossRef]  

26. J. Flórez, J. S. Lundeen, and M. V. Chekhova, “Pump depletion in parametric down-conversion with low pump energies,” Opt. Lett. 45, 4264–4267 (2020). [CrossRef]  

27. S. D. Bartlett and B. C. Sanders, “Universal continuous-variable quantum computation: requirement of optical nonlinearity for photon counting,” Phys. Rev. A 65, 042304 (2002). [CrossRef]  

28. S. D. Bartlett and B. C. Sanders, “Efficient classical simulation of optical quantum information circuits,” Phys. Rev. Lett. 89, 207903 (2002). [CrossRef]  

29. U. Chabaud, D. Markham, and F. Grosshans, “Stellar representation of non-Gaussian quantum states,” Phys. Rev. Lett. 124, 063605 (2020). [CrossRef]  

30. R. Yanagimoto, E. Ng, L. G. Wright, T. Onodera, and H. Mabuchi, “Efficient simulation of ultrafast quantum nonlinear optics with matrix product states,” Optica 8, 1306–1315 (2021). [CrossRef]  

31. J. E. Bourassa, N. Quesada, I. Tzitrin, A. Száva, T. Isacsson, J. Izaac, K. K. Sabapathy, G. Dauphinais, and I. Dhand, “Fast simulation of bosonic qubits via Gaussian functions in phase space,” PRX Quantum 2, 040315 (2021). [CrossRef]  

32. M. Walschaers, “Non-Gaussian quantum states and where to find them,” PRX Quantum 2, 030204 (2021). [CrossRef]  

33. Y.-X. Huang, M. Li, K. Lin, Y.-L. Zhang, G.-C. Guo, and C.-L. Zou, “Classical-to-quantum transition in multimode nonlinear systems with strong photon-photon coupling,” preprint arXiv:2111.09557 (2021).

34. R. Yanagimoto, E. Ng, A. Yamamura, T. Onodera, L. G. Wright, M. Jankowski, M. Fejer, P. L. McMahon, and H. Mabuchi, “Onset of non-Gaussian quantum physics in pulsed squeezing with mesoscopic fields,” preprint arXiv:2111.13799 (2021).

35. Y.-X. Huang, M. Li, Z.-J. Chen, X.-B. Zou, G.-C. Guo, and C.-L. Zou, “Quantum effects beyond mean-field treatment in quantum optics,” preprint arXiv:2111.14636 (2021).

36. W. Xing and T. Ralph, “Pump depletion in parametric amplification,”preprint arXiv:2201.01372 (2022).

37. N. Quesada and J. E. Sipe, “Why you should not use the electric field to quantize in nonlinear optics,” Opt. Lett. 42, 3443 (2017). [CrossRef]  

38. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University Press, 1995).

39. M. Born and L. Infeld, “On the quantization of the new field equations I,” Proc. R. Soc. Lond. A 147, 522–546 (1934). [CrossRef]  

40. P. D. Drummond and M. Hillery, The Quantum Theory of Nonlinear Optics (Cambridge University Press, 2014).

41. R. Loudon, The Quantum Theory of Light (Oxford University Press, 2000).

42. D. N. Klyshko, Photons and Nonlinear Optics (Routledge, 2018).

43. C. C. Gerry and P. Knight, Introductory Quantum Optics (Cambridge University Press, Cambridge, UK; New York, 2005).

44. M. Born and E. Wolf, Principles of Optics: Electromagnetic Theory of Propagation, Interference and Diffraction of Light, 7th ed. (Cambridge University Press, Cambridge; New York, 1999).

45. M. Hillery and L. D. Mlodinow, “Semiclassical expansion for nonlinear dielectric media,” Phys. Rev. A 31, 797–806 (1985). [CrossRef]  

46. M. Hillery and L. Mlodinow, “Quantized fields in a nonlinear dielectric medium: a microscopic approach,” Phys. Rev. A 55, 678–689 (1997). [CrossRef]  

47. P. Drummond and M. Hillery, “Quantum theory of dispersive electromagnetic modes,” Phys. Rev. A 59, 691–707 (1999). [CrossRef]  

48. M. Hillery and L. D. Mlodinow, “Quantization of electrodynamics in nonlinear dielectric media,” Phys. Rev. A 30, 1860–1865 (1984). [CrossRef]  

49. I. Abram and E. Cohen, “Quantum theory for light propagation in a nonlinear effective medium,” Phys. Rev. A 44, 500–517 (1991). [CrossRef]  

50. J. Sipe, N. A. Bhat, P. Chak, and S. Pereira, “Effective field theory for the nonlinear optical properties of photonic crystals,” Phys. Rev. E 69, 016604 (2004). [CrossRef]  

51. M. G. Raymer, “Quantum theory of light in a dispersive structured linear dielectric: a macroscopic Hamiltonian tutorial treatment,” J. Mod. Opt. 67, 196–212 (2020). [CrossRef]  

52. J. D. Joannopoulos, S. G. Johnson, J. N. Winn, and R. D. Meade, Photonic Crystals (Princeton University Press, 2011).

53. J. Sipe and M. Steel, “A Hamiltonian treatment of stimulated Brillouin scattering in nanoscale integrated waveguides,” New J. Phys. 18, 045004 (2016). [CrossRef]  

54. W. Xie, L. Chang, H. Shu, J. C. Norman, J. D. Peters, X. Wang, and J. E. Bowers, “Ultrahigh-q algaas-on-insulator microresonators for integrated nonlinear photonics,” Opt. Express 28, 32894–32906 (2020). [CrossRef]  

55. J. D. Jackson, Classical Electrodynamics (Wiley, 1998).

56. N. A. Bhat and J. Sipe, “Hamiltonian treatment of the electromagnetic field in dispersive and absorptive structured media,” Phys. Rev. A 73, 063808 (2006). [CrossRef]  

57. A. Judge, M. Steel, J. Sipe, and C. De Sterke, “Canonical quantization of macroscopic electrodynamics in a linear, inhomogeneous magnetoelectric medium,” Phys. Rev. A 87, 033824 (2013). [CrossRef]  

58. J. Sipe, “Photons in dispersive dielectrics,” J. Opt. A: Pure Appl. Opt. 11, 114006 (2009). [CrossRef]  

59. J. Heebner, R. Grover, and T. Ibrahim, Optical Microresonators Theory, Fabrication and Applications (Springer, 2008).

60. Z. Vernon and J. E. Sipe, “Spontaneous four-wave mixing in lossy microring resonators,” Phys. Rev. A 91, 053802 (2015). [CrossRef]  

61. P. Chak, S. Pereira, and J. Sipe, “Coupled-mode theory for periodic side-coupled microcavity and photonic crystal structures,” Phys. Rev. B 73, 035105 (2006). [CrossRef]  

62. M. Liscidini, L. Helt, and J. Sipe, “Asymptotic fields for a Hamiltonian treatment of nonlinear electromagnetic phenomena,” Phys. Rev. A 85, 013833 (2012). [CrossRef]  

63. R. W. Boyd, Nonlinear Optics (Academic Press, 2008).

64. A. Serafini, Quantum Continuous Variables: A Primer of Theoretical Methods (CRC Press, 2017).

65. R. Simon, E. C. G. Sudarshan, and N. Mukunda, “Gaussian pure states in quantum mechanics and the symplectic group,” Phys. Rev. A 37, 3028–3038 (1988). [CrossRef]  

66. B. Dutta, N. Mukunda, and R. Simon, “The real symplectic groups in quantum mechanics and optics,” Pramana 45, 471–497 (1995). [CrossRef]  

67. G. Adesso, S. Ragy, and A. R. Lee, “Continuous variable quantum information: Gaussian states and beyond,” Open Syst. Inf. Dyn. 21, 1440001 (2014). [CrossRef]  

68. R. Shankar, Principles of Quantum Mechanics (Springer Science Business Media, 2012).

69. J. J. Sakurai, Modern Quantum Mechanics (Addison-Wesley, 2011).

70. S. Barnett and P. M. Radmore, Methods in Theoretical Quantum Optics, Vol. 15 (Oxford University Press, 2002).

71. A. B. Klimov and S. M. Chumakov, A Group-theoretical Approach to Quantum Optics: Models of Atom–Field Interactions (John Wiley & Sons, 2009).

72. F. J. Dyson, “The radiation theories of Tomonaga, Schwinger, and Feynman,” Phys. Rev. 75, 486–502 (1949). [CrossRef]  

73. S. Blanes, F. Casas, J. Oteo, and J. Ros, “The magnus expansion and some of its applications,” Phys. Rep. 470, 151–238 (2009). [CrossRef]  

74. W. Magnus, “On the exponential solution of differential equations for a linear operator,” Comm. Pure Appl. Math. 7, 649–673 (1954). [CrossRef]  

75. M. Suzuki, “Generalized Trotter’s formula and systematic approximants of exponential operators and inner derivations with applications to many-body problems,” Commun. Math. Phys. 51, 183–190 (1976). [CrossRef]  

76. H. F. Trotter, “On the product of semi-groups of operators,” Proc. Am. Math. Soc. 10, 545–551 (1959). [CrossRef]  

77. In the case where Δij is not time-dependent, this can always be assumed without loss of generality as shown in Appendix C.

78. R. A. Horn and C. R. Johnson, Matrix Analysis (Cambridge University Press, 2012).

79. N. Killoran, J. Izaac, N. Quesada, V. Bergholm, M. Amy, and C. Weedbrook, “Strawberry Fields: asoftware platform for photonic quantum computing,” Quantum 3, 129129 (2019). [CrossRef]  

80. B. Gupt, J. Izaac, and N. Quesada, “The Walrus: a library for the calculation of Hafnians, Hermite polynomials and Gaussian boson sampling,” JOSS 4, 1705 (2019). [CrossRef]  

81. A. Lvovsky, W. Wasilewski, and K. Banaszek, “Decomposing a pulsed optical parametric amplifier into independent squeezers,” J. Mod. Opt. 54, 721–733 (2007). [CrossRef]  

82. W. Wasilewski, A. I. Lvovsky, K. Banaszek, and C. Radzewicz, “Pulsed squeezed light: simultaneous squeezing of multiple modes,” Phys. Rev. A 73, 063819 (2006). [CrossRef]  

83. T. Lipfert, D. B. Horoshko, G. Patera, and M. I. Kolobov, “Bloch-Messiah decomposition and magnus expansion for parametric down-conversion with monochromatic pump,” Phys. Rev. A 98, 013815 (2018). [CrossRef]  

84. A. Gatti, R. Zambrini, M. San Miguel, and L. A. Lugiato, “Multiphoton multimode polarization entanglement in parametric down-conversion,” Phys. Rev. A 68, 053807 (2003). [CrossRef]  

85. A. K. Ekert and P. L. Knight, “Relationship between semiclassical and quantum-mechanical input–output theories of optical response,” Phys. Rev. A 43, 3934–3938 (1991). [CrossRef]  

86. A. Christ, B. Brecht, W. Mauerer, and C. Silberhorn, “Theory of quantum frequency conversion and type-II parametric down-conversion in the high-gain regime,” New J. Phys. 15, 053038 (2013). [CrossRef]  

87. K. E. Cahill and R. J. Glauber, “Ordered expansions in boson amplitude operators,” Phys. Rev. 177, 1857–1881 (1969). [CrossRef]  

88. Note that this equation does not imply that $\mathcal {U} = \exp \left \{ \tfrac 12 \sum _{k, l=1}^\ell J_{kl} a^\dagger _{k} a^\dagger _l - \text {H.c.}\right \}$; in particular, we can postmultiply the exponential by any unitary operator that satisfies $\mathcal {U}_{0} | {\text {vac}}\rangle = | {\text {vac}}\rangle$ and still satisfy Eq. (237). The form of the operator $\mathcal {U}_{0}$ is described in Appendix E.

89. The diagonal and unitary matrices obtained from the Takagi decomposition can be related to the diagonal and orthogonal matrices obtained by performing a Bloch–Messiah decomposition of the symplectic matrix that diagonalizes the covariance matrix of the pure state in Eq. (237). This symplectic matrix can be obtained by doing a Williamson decomposition of the covariance matrix of the quantum state as described by, e.g., Serafini [64].

90. A. M. Brańczyk, T. M. Stace, and T. Ralph, “Time ordering in spontaneous parametric down-conversion,” in AIP Conference Proceedings, Vol. 1363 (2011), pp. 335–338.

91. W. P. Grice and I. A. Walmsley, “Spectral information and distinguishability in type-II down-conversion with a broadband pump,” Phys. Rev. A 56, 1627–1634 (1997). [CrossRef]  

92. G. Triginer, M. D. Vidrighin, N. Quesada, A. Eckstein, M. Moore, W. S. Kolthammer, J. E. Sipe, and I. A. Walmsley, “Understanding high-gain twin-beam sources using cascaded stimulated emission,” Phys. Rev. X 10, 031063 (2020). [CrossRef]  

93. V. Bargmann, “On a Hilbert space of analytic functions and an associated integral transform part I,” Comm. Pure Appl. Math. 14, 187–214 (1961). [CrossRef]  

94. I. E. Segal and G. W. Mackey, Mathematical Problems of Relativistic Physics, Vol. 2 (American Mathematical Soc., 1963).

95. R. García-Patrón, J. J. Renema, and V. Shchesnovich, “Simulating boson sampling in lossy architectures,” Quantum 3, 169 (2019). [CrossRef]  

96. D. J. Brod and M. Oszmaniec, “Classical simulation of linear optics subject to nonuniform losses,” Quantum 4, 267 (2020). [CrossRef]  

97. O. F. Thomas, W. McCutcheon, and D. P. McCutcheon, “A general framework for multimode Gaussian quantum optics and photo-detection: application to Hong–Ou–Mandel interference with filtered heralded single photon sources,” APL Photonics 6, 040801 (2021). [CrossRef]  

98. K. Laiho, T. Dirmeier, M. Schmidt, S. Reitzenstein, and C. Marquardt, “Measuring higher-order photon correlations of faint quantum light: a short review,” Phys. Lett. A 435, 128059 (2022). [CrossRef]  

99. A. Christ, K. Laiho, A. Eckstein, K. N. Cassemiro, and C. Silberhorn, “Probing multimode squeezing with correlation functions,” New J. Phys. 13, 033027 (2011). [CrossRef]  

100. C. S. Hamilton, R. Kruse, L. Sansoni, S. Barkhofen, C. Silberhorn, and I. Jex, “Gaussian boson sampling,” Phys. Rev. Lett. 119, 170501 (2017). [CrossRef]  

101. R. Kruse, C. S. Hamilton, L. Sansoni, S. Barkhofen, C. Silberhorn, and I. Jex, “Detailed study of Gaussian boson sampling,” Phys. Rev. A 100, 032326 (2019). [CrossRef]  

102. N. Quesada, J. M. Arrazola, and N. Killoran, “Gaussian boson sampling using threshold detectors,” Phys. Rev. A 98, 062322 (2018). [CrossRef]  

103. A. Deshpande, A. Mehta, T. Vincent, N. Quesada, M. Hinsche, M. Ioannou, L. Madsen, J. Lavoie, H. Qi, J. Eisert, D. Hangleiter, B. Fefferman, and I. Dhand, “Quantum computational advantage via high-dimensional Gaussian boson sampling,” Sci. Adv. 8, eabi7894 (2022). [CrossRef]  

104. N. Quesada and J. M. Arrazola, “Exact simulation of Gaussian boson sampling in polynomial space and exponential time,” Phys. Rev. Res. 2, 023005 (2020). [CrossRef]  

105. H.-S. Zhong, L.-C. Peng, Y. Li, Y. Hu, W. Li, J. Qin, D. Wu, W. Zhang, H. Li, L. Zhang, Z. Wang, L. You, X. Jiang, L. Li, N.-L. Liu, J. P. Dowling, C.-Y. Lu, and J.-W. Pan, “Experimental Gaussian boson sampling,” Sci. Bull. 64, 511–515 (2019). [CrossRef]  

106. H.-S. Zhong, H. Wang, Y.-H. Deng, et al., “Quantum computational advantage using photons,” Science 370, 1460–1463 (2020). [CrossRef]  

107. L. G. Helt and N. Quesada, “Degenerate squeezing in waveguides: a unified theoretical approach,” JPhys Photonics 2, 035001 (2020). [CrossRef]  

108. N. Quesada, G. Triginer, M. D. Vidrighin, and J. E. Sipe, “Theory of high-gain twin-beam generation in waveguides: from Maxwell’s equations to efficient simulation,” Phys. Rev. A 102, 033519 (2020). [CrossRef]  

109. G. P. Agrawal, Nonlinear Fiber Optics, 4th ed. (Academic Press, Burlington, MA, 2007).

110. M.-D. Vidrighin, “Quantum optical measurements for practical estimation and information thermodynamics,” Ph.D. thesis, Imperial College London (2017).

111. C. McKinstrie and M. Marhic, “Quadrature and number fluctuations produced by parametric devices driven by pulsed pumps,” Opt. Express 21, 19437–19466 (2013). [CrossRef]  

112. D. B. Horoshko, L. La Volpe, F. Arzani, N. Treps, C. Fabre, and M. I. Kolobov, “Bloch–Messiah reduction for twin beams of light,” Phys. Rev. A 100, 013837 (2019). [CrossRef]  

113. M. I. Kolobov, “The spatial behavior of nonclassical light,” Rev. Mod. Phys. 71, 1539–1589 (1999). [CrossRef]  

114. Z. Yang, M. Liscidini, and J. E. Sipe, “Spontaneous parametric down-conversion in waveguides: a backward heisenberg picture approach,” Phys. Rev. A 77, 033808 (2008). [CrossRef]  

115. N. Quesada and J. E. Sipe, “Effects of time ordering in quantum nonlinear optics,” Phys. Rev. A 90, 063840 (2014). [CrossRef]  

116. A. M. Brańczyk, A. Fedrizzi, T. M. Stace, T. C. Ralph, and A. G. White, “Engineered optical nonlinearity for quantum light sources,” Opt. Express 19, 55–65 (2011). [CrossRef]  

117. P. B. Dixon, J. H. Shapiro, and F. N. C. Wong, “Spectral engineering by Gaussian phase-matching for quantum photonics,” Opt. Express 21, 5879–5890 (2013). [CrossRef]  

118. J.-L. Tambasco, A. Boes, L. G. Helt, M. J. Steel, and A. Mitchell, “Domain engineering algorithm for practical and effective photon sources,” Opt. Express 24, 19616–19626 (2016). [CrossRef]  

119. A. Dosseva, L. Cincio, and A. M. Brańczyk, “Shaping the joint spectrum of down-converted photons through optimized custom poling,” Phys. Rev. A 93, 013801 (2016). [CrossRef]  

120. F. Graffitti, D. Kundys, D. T. Reid, A. M. Brańczyk, and A. Fedrizzi, “Pure down-conversion photons through sub-coherence-length domain engineering,” Quantum Sci. Technol. 2, 035001 (2017). [CrossRef]  

121. A. B. U’Ren, C. Silberhorn, R. Erdmann, K. Banaszek, W. P. Grice, I. A. Walmsley, and M. G. Raymer, “Generation of pure-state single-photon wavepackets by conditional preparation based on spontaneous parametric downconversion,” arXiv preprint quant-ph/0611019 (2006).

122. N. Quesada and A. M. Brańczyk, “Gaussian functions are optimal for waveguided nonlinear-quantum-optical processes,” Phys. Rev. A 98, 043813 (2018). [CrossRef]  

123. N. Quesada and J. E. Sipe, “Time-ordering effects in the generation of entangled photons using nonlinear optical processes,” Phys. Rev. Lett. 114, 093903 (2015). [CrossRef]  

124. N. Quesada, “Very nonlinear quantum optics,” Ph.D. thesis, University of Toronto (2015).

125. D. G. Rabus, Integrated Ring Resonators (Springer, 2007).

126. Z. Vernon, “Microresonators for nonlinear quantum optics,” Ph.D. thesis, University of Toronto (2017).

127. A. Yariv and P. Yeh, Photonics: Optical Electronics in Modern Communications (Oxford University Press, 2007).

128. Z. Vernon, M. Menotti, C. Tison, J. Steidle, M. Fanto, P. Thomas, S. Preble, A. Smith, P. Alsing, M. Liscidini, and J. E. Sipe, “Truly unentangled photon pairs without spectral filtering,” Opt. Lett. 42, 3638–3641 (2017). [CrossRef]  

129. J. B. Christensen, J. G. Koefoed, K. Rottwitt, and C. McKinstrie, “Engineering spectrally unentangled photon pairs from nonlinear microring resonators by pump manipulation,” Opt. Lett. 43, 859–862 (2018). [CrossRef]  

130. V. D. Vaidya, B. Morrison, L. Helt, R. Shahrokshahi, D. Mahler, M. Collins, K. Tan, J. Lavoie, A. Repingon, M. Menotti, N. Quesada, R. C. Pooser, A. E. Lita, T. Gerrits, S. W. Nam, and Z. Vernon, “Broadband quadrature-squeezed vacuum and nonclassical photon number correlations from a nanophotonic device,” Sci. Adv. 6, eaba9186 (2020). [CrossRef]  

131. Z. Vernon, N. Quesada, M. Liscidini, B. Morrison, M. Menotti, K. Tan, and J. Sipe, “Scalable squeezed-light source for continuous-variable quantum sampling,” Phys. Rev. Appl. 12, 064024 (2019). [CrossRef]  

132. Y. Zhao, Y. Okawachi, J. K. Jang, X. Ji, M. Lipson, and A. L. Gaeta, “Near-degenerate quadrature-squeezed vacuum generation on a silicon-nitride chip,” Phys. Rev. Lett. 124, 193601 (2020). [CrossRef]  

133. Y. Zhang, M. Menotti, K. Tan, V. Vaidya, D. Mahler, L. Helt, L. Zatti, M. Liscidini, B. Morrison, and Z. Vernon, “Squeezed light from a nanophotonic molecule,” Nat. Commun. 12, 1–6 (2021). [CrossRef]  

134. H. Seifoory, Z. Vernon, D. Mahler, M. Menotti, Y. Zhang, and J. E. Sipe, “Degenerate squeezing in a dual-pumped integrated microresonator: parasitic processes and their suppression,”preprint arXiv:2109.03298 (2021).

135. M. Chen, N. C. Menicucci, and O. Pfister, “Experimental realization of multipartite entanglement of 60 modes of a quantum optical frequency comb,” Phys. Rev. Lett. 112, 120505 (2014). [CrossRef]  

136. Y. K. Chembo, “Quantum dynamics of Kerr optical frequency combs below and above threshold: spontaneous four-wave mixing, entanglement, and squeezed states of light,” Phys. Rev. A 93, 033820 (2016). [CrossRef]  

137. C. Reimer, M. Kues, P. Roztocki, B. Wetzel, F. Grazioso, B. E. Little, S. T. Chu, T. Johnston, Y. Bromberg, L. Caspani, D. J. Moss, and R. Morandotti, “Generation of multiphoton entangled quantum states by means of integrated frequency combs,” Science 351, 1176–1180 (2016). [CrossRef]  

138. S. Hagness, D. Rafizadeh, S. Ho, and A. Taflove, “FDTD microcavity simulations: design and experimental realization of waveguide-coupled single-mode ring and whispering-gallery-mode disk resonators,” J. Lightwave Technol. 15, 2154–2165 (1997). [CrossRef]  

139. H. P. Yuen and V. W. Chan, “Noise in homodyne and heterodyne detection,” Opt. Lett. 8, 177–179 (1983). [CrossRef]  

140. R. Wolf, Y. Jia, S. Bonaus, C. S. Werner, S. J. Herr, I. Breunig, K. Buse, and H. Zappe, “Quasi-phase-matched nonlinear optical frequency conversion in on-chip whispering galleries,” Optica 5, 872–875 (2018). [CrossRef]  

141. J.-Y. Chen, Z.-H. Ma, Y. M. Sua, Z. Li, C. Tang, and Y.-P. Huang, “Ultra-efficient frequency conversion in quasi-phase-matched lithium niobate microrings,” Optica 6, 1244–1245 (2019). [CrossRef]  

142. J. Lu, J. B. Surya, X. Liu, A. W. Bruch, Z. Gong, Y. Xu, and H. X. Tang, “Periodically poled thin-film lithium niobate microring resonators with a second-harmonic generation efficiency of 250,000%/w,” Optica 6, 1455–1460 (2019). [CrossRef]  

143. R. Luo, Y. He, H. Liang, M. Li, J. Ling, and Q. Lin, “Optical parametric generation in a lithium niobate microring with modal phase matching,” Phys. Rev. Appl. 11, 034026 (2019). [CrossRef]  

144. R. Luo, H. Jiang, S. Rogers, H. Liang, Y. He, and Q. Lin, “On-chip second-harmonic generation and broadband parametric down-conversion in a lithium niobate microresonator,” Opt. Express 25, 24531–24539 (2017). [CrossRef]  

145. Z. Yang, P. Chak, A. D. Bristow, H. M. van Driel, R. Iyer, J. S. Aitchison, A. L. Smirl, and J. Sipe, “Enhanced second-harmonic generation in AlGaAs microring resonators,” Opt. Lett. 32, 826–828 (2007). [CrossRef]  

146. J. Lin, Y. Xu, J. Ni, M. Wang, Z. Fang, L. Qiao, W. Fang, and Y. Cheng, “Phase-matched second-harmonic generation in an on-chip LiNbO3 microresonator,” Phys. Rev. Appl. 6, 014002 (2016). [CrossRef]  

147. G. Lin, J. U. Fürst, D. V. Strekalov, and N. Yu, “Wide-range cyclic phase matching and second harmonic generation in whispering gallery resonators,” Appl. Phys. Lett. 103, 181107 (2013). [CrossRef]  

148. L. G. Helt, M. Liscidini, and J. E. Sipe, “How does it scale? Comparing quantum and classical nonlinear optical processes in integrated devices,” J. Opt. Soc. Am. B 29, 2199–2212 (2012). [CrossRef]  

149. M. Liscidini and J. Sipe, “Stimulated emission tomography,” Phys. Rev. Lett. 111, 193602 (2013). [CrossRef]  

150. A. De Rossi, V. Berger, M. Calligaro, G. Leo, V. Ortiz, and X. Marcadet, “Parametric fluorescence in oxidized aluminum gallium arsenide waveguides,” Appl. Phys. Lett. 79, 3758–3760 (2001). [CrossRef]  

151. V. Introini, M. J. Steel, J. E. Sipe, L. G. Helt, and M. Liscidini, “Spontaneous parametric down conversion in a doubly resonant one-dimensional photonic crystal,” Opt. Lett. 45, 1244 (2020). [CrossRef]  

152. S. Azzini, D. Grassani, M. Galli, L. C. Andreani, M. Sorel, M. J. Strain, L. G. Helt, J. E. Sipe, M. Liscidini, and D. Bajoni, “From classical four-wave mixing to parametric fluorescence in silicon microring resonators,” Opt. Lett. 37, 3807 (2012). [CrossRef]  

153. L. A. Rozema, C. Wang, D. H. Mahler, A. Hayat, A. M. Steinberg, J. E. Sipe, and M. Liscidini, “Characterizing an entangled-photon source with classical detectors and measurements,” Optica 2, 430 (2015). [CrossRef]  

154. B. Fang, M. Liscidini, J. E. Sipe, and V. O. Lorenz, “Multidimensional characterization of an entangled photon-pair source via stimulated emission tomography,” Opt. Express 24, 10013 (2016). [CrossRef]  

155. M. A. Ciampini, A. Geraldi, V. Cimini, C. Macchiavello, J. E. Sipe, M. Liscidini, and P. Mataloni, “Stimulated emission tomography: beyond polarization,” Opt. Lett. 44, 41 (2019). [CrossRef]  

156. A. Eckstein, G. Boucher, A. Lemaître, P. Filloux, I. Favero, G. Leo, J. E. Sipe, M. Liscidini, and S. Ducci, “High-resolution spectral characterization of two photon states via classical measurements,” Laser Photonics Rev. 8, L76–L80 (2014). [CrossRef]  

157. B. Fang, O. Cohen, M. Liscidini, J. E. Sipe, and V. O. Lorenz, “Fast and highly resolved capture of the joint spectral density of photon pairs,” Optica 1, 281 (2014). [CrossRef]  

158. I. Jizan, L. G. Helt, C. Xiong, M. J. Collins, D.-Y. Choi, C. Joon Chae, M. Liscidini, M. J. Steel, B. J. Eggleton, and A. S. Clark, “Bi-photon spectral correlation measurements from a silicon nanowire in the quantum and classical regimes,” Sci. Rep. 5, 12557 (2015). [CrossRef]  

159. D. Grassani, A. Simbula, S. Pirotta, M. Galli, M. Menotti, N. C. Harris, T. Baehr-Jones, M. Hochberg, C. Galland, M. Liscidini, and D. Bajoni, “Energy correlations of photon pairs generated by a silicon microring resonator probed by Stimulated Four Wave Mixing,” Sci. Rep. 6, 23564 (2016). [CrossRef]  

160. I. Jizan, B. Bell, L. G. Helt, A. C. Bedoya, C. Xiong, and B. J. Eggleton, “Phase-sensitive tomography of the joint spectral amplitude of photon pair sources,” Opt. Lett. 41, 4803–4806 (2016). [CrossRef]  

161. M. Borghi, “Phase-resolved joint spectra tomography of a ring resonator photon pair source using a silicon photonic chip,” Opt. Express 28, 7442 (2020). [CrossRef]  

162. U. Schollwöck, “The density-matrix renormalization group in the age of matrix product states,” Ann. Phys. 326, 96–192 (2011). [CrossRef]  

163. R. Orús, “A practical introduction to tensor networks: matrix product states and projected entangled pair states,” Ann. Phys. 349, 117–158 (2014). [CrossRef]  

164. S. Montangero, Introduction to Tensor Network Methods (Springer, 2018).

165. C. Schön, E. Solano, F. Verstraete, J. I. Cirac, and M. M. Wolf, “Sequential generation of entangled multiqubit states,” Phys. Rev. Lett. 95, 110503 (2005). [CrossRef]  

166. C. Schön, K. Hammerer, M. M. Wolf, J. I. Cirac, and E. Solano, “Sequential generation of matrix-product states in cavity QED,” Phys. Rev. A 75, 032311 (2007). [CrossRef]  

167. L. G. Helt, Z. Yang, M. Liscidini, and J. Sipe, “Spontaneous four-wave mixing in microring resonators,” Opt. Lett. 35, 3006–3008 (2010). [CrossRef]  

168. I. Dhand, M. Engelkemeier, L. Sansoni, S. Barkhofen, C. Silberhorn, and M. B. Plenio, “Proposal for quantum simulation via all-optically-generated tensor network states,” Phys. Rev. Lett. 120, 130501 (2018). [CrossRef]  

169. C. Roberts, A. Milsted, M. Ganahl, A. Zalcman, B. Fontaine, Y. Zou, J. Hidary, G. Vidal, and S. Leichenauer, “Tensornetwork: a library for physics and machine learning,”preprint arXiv:1905.01330 (2019).

170. M. Fishman, S. R. White, and E. M. Stoudenmire, “The ITensor software library for tensor network calculations,”preprint arXiv:2007.14822 (2020).

171. D. Jaschke, M. L. Wall, and L. D. Carr, “Open source matrix product states: opening ways to simulate entangled many-body quantum systems in one dimension,” Comput. Phys. Commun. 225, 59–91 (2018). [CrossRef]  

172. B. Bauer, L. D. Carr, H. G. Evertz, A. Feiguin, J. Freire, S. Fuchs, L. Gamper, J. Gukelberger, E. Gull, S. Guertler, A. Hehn, R. Igarashi, S. V. Isakov, D. Koop, P. N. Ma, P. Mates, H. Matsuo, O. Parcollet, G. Pawlowski, J. D. Picon, L. Pollet, E. Santos, V. W. Scarola, U. Schollwöck, C. Silva, B. Surer, S. Todo, S. Trebst, M. Troyer, M. L. Wall, P. Werner, and S. Wessel, “The Alps project release 2.0: open source software for strongly correlated systems,” J. Stat. Mech.: Theory Exp. 2011, P05001 (2011). [CrossRef]  

aop-14-3-291-i001

Nicolás Quesada is an Assistant Professor in the Engineering Physics department at Polytechnique Montréal and holds an MEI research chair in quantum photonics. His group develops the theory and computational tools underpinning the next generation of nonclassical bright light sources and devices needed for building fault-tolerant quantum computers, quantum communication networks, and quantum sensors. Prior to Polytechnique, he worked at Xanadu Quantum Technologies as a lead developer of Strawberry Fields and The Walrus software libraries and led their theoretical efforts to demonstrate photonic quantum advantage. Before joining industry, he was an NSERC postdoctoral fellow and, as a PhD student, a Vanier and Stoicheff scholar at the University of Toronto.

aop-14-3-291-i002

Lukas G. Helt is a quantum photonics engineer and software developer at Xanadu Quantum Technologies. He received his PhD degree in physics in 2013 from the University of Toronto (Canada). Prior to joining Xanadu he was project leader of the CUDOS Nonlinear Quantum Photonics project while holding a research fellowship at Macquarie University (Australia).

aop-14-3-291-i003

Matteo Menotti received his PhD degree in Physics from the University of Pavia (Italy) in 2018 with a dissertation on the generation of nonclassical states of light in integrated devices. He has been a quantum photonics engineer at Xanadu Quantum Technologies since 2018, where he works on the design and simulation of integrated sources of squeezed states for photonic quantum computation.

aop-14-3-291-i004

Marco Liscidini is Associate Professor at the Department of Physics of the University of Pavia (Italy), and he serves as technical advisor to Xanadu Quantum Technologies. He received the PhD degree in physics from the University of Pavia (Italy) in 2006. From 2007 to 2009, he was a Post-Doctoral Fellow at the Department of Physics of the University of Toronto (Canada). His research activity focuses on the theoretical study and modeling of the light–matter interaction in micro- and nanostructures. He works in several areas of photonics, including classical and quantum nonlinear optics, photonic crystal structures, and optical sensing. His theoretical research activity is in strong collaboration with experimental groups and in the framework of national, European, US, and Canadian research programs. He is a Fellow of Optica.

aop-14-3-291-i005

John E. Sipe is a Professor of Physics at the University of Toronto, where he has been on the academic staff since 1981; he is a theorist who interacts strongly with experimental groups. His research interests are in quantum optics and condensed matter physics, particularly on topics such as quantum interference processes in solids, quantum nonlinear optics in integrated photonic structures, and foundational issues in the description of the interaction of light with matter. He is a Fellow of Optica, of the American Physical Society, and of the Royal Society of Canada.

Data availability

No data were generated or analyzed in the presented research.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (24)

Figure 1.
Figure 1. Schematic representation of an isolated ring resonator. The inset shows the typical field profile of the fundamental transverse-electric mode supported by the bent waveguide. In the simulation the waveguide is 800 nm wide and 400 nm high, is made of silicon nitride fully clad in silica, and has a bending radius of 10 $\mu$ m. The simulation is performed at a wavelength of 1550 nm. The contour plot accounts for the $E_{\zeta }$ component, whereas the arrow field represents $\boldsymbol{E}_{\boldsymbol{r}_\perp }$
Figure 2.
Figure 2. Schematic representation of a channel waveguide. (a) Typical field intensity profile of the fundamental transverse-electric mode, computed by an eigenmode solver. The waveguide and simulation parameters are the same as in Fig. 1, but with no curvature. By comparison with Fig. 1 it is possible to appreciate the different field distribution produced by the waveguide curvature. The contour plot accounts for the $E_{x}$ component, whereas the arrow field represents $\boldsymbol{E}_{y,z}$ . (b) Scanning electron microscopy (SEM) image of a typical high index contrast waveguide cross section. Reprinted from [54]. Copyright 2020 Optical Society of America.
Figure 3.
Figure 3. Schematic representation of a point-coupled ring resonator. The coupling is modeled by a simple $\begin {pmatrix} \sigma & i\kappa \\ i\kappa & \sigma \\ \end {pmatrix}$ matrix, where $(\sigma, \kappa ) \in \mathbb {R}$ are the self-coupling and cross-coupling coefficients, respectively. The coupler is assumed lossless, hence yielding $\sigma ^2 + \kappa ^2 = 1$ .
Figure 4.
Figure 4. Typical transmission of a point-coupled ring resonator. Here $\mathscr {R}(J)$ represents the wave vector range associated with each resonance $J$ , centered at $k_J$ .
Figure 5.
Figure 5. Enlarged view of the point coupler of Fig. 3 between a channel and a ring, with particular emphasis on the conventional coordinate systems used in the two waveguide.
Figure 6.
Figure 6. Schematic representation of a point-coupled ring resonator with and additional fictitious coupler and waveguide, referred to as “phantom channel.”
Figure 7.
Figure 7. Wigner functions of the vacuum state and a squeezed vacuum state.
Figure 8.
Figure 8. Block diagram describing how to calculate the state generated by a quadratic bosonic Hamiltonian.
Figure 9.
Figure 9. Experimental and theoretical prediction for the blocks of the transfer function $\boldsymbol{K}^{b,c}$ for nondegenerate squeezing in a periodically poled potassium titanyl phosphate SPDC source. The labels $U_{s,s}, U_{i,s}, U_{s,i}, U_{i,i}$ in the figure correspond to $V_{b,b}, W_{c,b}, W_{b,c}, W_{c,c}$ in the notation used in this paper. Figure 9 reprinted with permission from Triginer et al., Phys. Rev. X 10, 031063 (2019) [92]. Copyright 2019 by the American Physical Society.
Figure 10.
Figure 10. Amplitudes of the orthonormal functions $\phi _{0}(\Omega )$ and $\phi _{1}(\Omega )$ needed to expand the time-ordering corrected JSA of (364).
Figure 11.
Figure 11. Comparison of the generated mean photon number per pump pulse $\langle n \rangle$ and Schmidt number $K$ for no time-ordering corrections (Magnus 1), time-ordering corrections up to $\Phi ^{3}$ (Magnus 3), and a full numerical solution.
Figure 12.
Figure 12. Comparison of the absolute values of the generated JSAs $\left \vert J_{SI}(\Omega _{1},\Omega _{2})\right \vert$ with $\Phi =4$ (left column) for (a) no time-ordering corrections, (b) time-ordering corrections up to $\Phi ^{3}$ , and (c) a full numerical solution, as well as $\Phi =11$ (right column) for (d) no time-ordering corrections, (e) time-ordering corrections up to $\Phi ^{3}$ , and (f) a full numerical solution.
Figure 13.
Figure 13. Schematic representation of an integrated ring resonator of radius $R$ point-coupled to a bus waveguide. The interaction is regulated by the self-coupling and cross-coupling coefficients $\sigma$ and $\kappa$ , linked by $\sigma ^2+\kappa ^2=1$ [127].
Figure 14.
Figure 14. Energy-level diagrams for the nonlinear processes considered in this section. From left to right: SP-SFWM (Sec. 5.4), SP-SFWM (Sec. 5.5), and SPDC (Sec. 5.6).
Figure 15.
Figure 15. Ring resonances involved in a SP-SFWM process, and the associated SPM and XPM wavelength shifts.
Figure 16.
Figure 16. Ring resonances involved in a DP-SFWM process, and the associated SPM and XPM wavelength shifts.
Figure 17.
Figure 17. Amplitude of the “ringdown” function $f_{0}(t)$ in (421). Note that we make explicit the fact that the Heaviside function takes the value 1/2 when evaluated at 0.
Figure 18.
Figure 18. Squeezing and anti-squeezing spectra (blue and black, respectively) as a function of (a) the pump power and (b) the sideband frequency, according to (422). Reprinted by permission from Macmillan Publishers Ltd: Zhang et al., Nature Commun. 12, 2233 (2021) [133]. Copyright (2021).
Figure 19.
Figure 19. Ring resonances involved in a nondegenerate SPDC process.
Figure 20.
Figure 20. Ratio between the generated idler photons in spontaneous and stimulated FWM for several ring resonators. Reprinted with permission from [152]. Copyright 2012 Optical Society of America.
Figure 21.
Figure 21. JSD (a) measured by SET and (b) calculated in the case of photon pairs generated under pulsed pumping in a microring resonator. Reprinted by permission from Macmillan Publishers Ltd: Grassani et al., Sci. Rep. 6, 23564 (2016) [159]. Copyright 2016.
Figure 22.
Figure 22. (a) Graphical representation of an arbitrary tensor $R$ with indices $i_0,i_1,\ldots, i_{\ell -1}$ . (b) Representation of the same tensor using MPSs with auxiliary indices $\alpha _1,\alpha _2,\ldots,\alpha _{\ell -1}$ .
Figure 23.
Figure 23. Circuit representation of the unitary evolution operator in (460) in a discretization using four modes. Note that in this figure the first wire represents the resonator mode, whereas the rest of the wires represent the different discretized sections of the waveguide. We use vertical lines with dots at the ends to represent the beam splitter operation $\mathcal {B}_{0,j}$ between modes 0 (the resonator) and the $j$ th part of the waveguide.
Figure 24.
Figure 24. Circuit representation of the unitary evolution operator in (460) and for a discretization using four modes. Note that in this figure we apply swap operations after each beam splitter. Thus, after the $i$ th swap the resonator mode lies in the $i$ th wire and, in particular, at the end of the interaction the resonator mode is represented by the last wire. We use vertical lines with dots at the end to represent the beam splitter operation $\mathcal {B}_{i,i+1}$ between wire $i$ and $i+1$ and similarly we use vertical lines with crosses to represent swap operations between wires $i$ and $i+1$ .

Equations (565)

Equations on this page are rendered with MathJax. Learn more.

$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})\boldsymbol{E}(\boldsymbol{r},t),\\ & \boldsymbol{B}(\boldsymbol{r},t)=\mu_{0}\boldsymbol{H}(\boldsymbol{r},t), \end{aligned}$$
$${H}_{\textrm{ring}}^{\textrm{L}}=\sum_{J}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J},$$
$$\boldsymbol{E}(\boldsymbol{r})=\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2\mathcal{L}}}c_{J}\mathsf{e}_{J}(\boldsymbol{r}_{{\perp}})e^{i\kappa_{J}\zeta}+\textrm{H.c.},$$
$$d\mathfrak{h}=\boldsymbol{H}\cdot \textrm{d}\boldsymbol{B}+\boldsymbol{E}\cdot \textrm{d}\boldsymbol{D},$$
$$\mathfrak{h}(\boldsymbol{r},t)=\frac{1}{2\mu_{0}}\boldsymbol{B}(\boldsymbol{r},t)\cdot\boldsymbol{B}(\boldsymbol{r},t) +\frac{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}{2}\boldsymbol{E}(\boldsymbol{r},t)\cdot\boldsymbol{E}(\boldsymbol{r},t),$$
$${H}_{\textrm{ring}}={H}_{\textrm{ring}}^{\textrm{L}}+{H}_{\textrm{ring}}^{\textrm{NL}},$$
$$\begin{aligned}& D(\boldsymbol{r},t)=\epsilon_{0}\sum_{n=1}^{N}\varepsilon_{n}(\boldsymbol{r})E^{n}(\boldsymbol{r},t),\\ & B(\boldsymbol{r},t)=\mu_{0}H(\boldsymbol{r},t), \end{aligned}$$
$$\mathfrak{h}(\boldsymbol{r},t)=\frac{1}{2\mu_{0}}B^{2}(\boldsymbol{r},t)+\epsilon_{0}\sum_{n=1}^{N}\frac{n}{n+1}\varepsilon_{n}(\boldsymbol{r})E^{n+1}(\boldsymbol{r},t).$$
$${H}_{\textrm{ring}}^{\textrm{NL}}=\frac{3}{4}\epsilon_{0}\int\varepsilon_{3}(\boldsymbol{r})E^{4}(\boldsymbol{r})\textrm{d}\boldsymbol{r}.$$
$$\frac{\textrm{d}{c_{P}}}{\textrm{d}t}={-}i\omega_{P}c_{P}-\frac{9i}{\hbar}\left(\frac{\hbar\omega_{P}}{2\mathcal{L}}\right)^{2}Kc_{P}^{{\dagger}}c_{P}c_{P},$$
$$K=\epsilon_{0}\int\varepsilon_{3}\left(\boldsymbol{r}\right)\left[\mathsf{e}_{P}^{*}\left(\boldsymbol{r}_{{\perp}}\right)\right]^{2} \left[\mathsf{e}_{P}\left(\boldsymbol{r}_{{\perp}}\right)\right]^{2}\textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta.$$
$$\omega_{P}\rightarrow\omega_{P}+\frac{9}{\hbar}\left(\frac{\hbar\omega_{P}}{2\mathcal{L}}\right)^{2}Kc_{P}^{{\dagger}}c_{P}.$$
$$\omega_{P}\rightarrow\omega_{P}-\frac{3}{\hbar}\left(\frac{\hbar\omega_{P}}{2\mathcal{L}}\right)^{2}Kc_{P}^{{\dagger}}c_{P},$$
$$\frac{\partial\boldsymbol{B}(\boldsymbol{r},t)}{\partial t}=\frac{1}{i\hbar}\left[\boldsymbol{B}(\boldsymbol{r},t),{H}_{\textrm{ring}}^{\textrm{L}}\right]+\frac{1}{i\hbar}\left[\boldsymbol{B}(\boldsymbol{r},t),{H}_{\textrm{ring}}^{\textrm{NL}}\right].$$
$$\textrm{d}\mathfrak{L=}-\boldsymbol{D}\cdot \textrm{d}\boldsymbol{E}+\boldsymbol{H}\cdot \textrm{d}\boldsymbol{B}.$$
$$\begin{aligned} & \frac{\partial\boldsymbol{B}(\boldsymbol{r},t)}{\partial t}={-}\boldsymbol{\nabla}\times\boldsymbol{E}(\boldsymbol{r},t),\\ & \frac{\partial\boldsymbol{D}(\boldsymbol{r},t)}{\partial t}=\boldsymbol{\nabla}\times\boldsymbol{H}(\boldsymbol{r},t),\\ & \boldsymbol{\nabla}\cdot\boldsymbol{B}(\boldsymbol{r},t)=0,\\ & \boldsymbol{\nabla}\cdot\boldsymbol{D}(\boldsymbol{r},t)=0 \end{aligned}$$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\boldsymbol{D}_{\alpha}(\boldsymbol{r})e^{{-}i\omega_{\alpha}t}+\textrm{c.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\boldsymbol{B}_{\alpha}(\boldsymbol{r})e^{{-}i\omega_{\alpha}t}+\textrm{c.c.}, \end{aligned}$$
$$\begin{aligned} & \boldsymbol{\nabla}\times\left[\frac{\boldsymbol{\nabla}\times\boldsymbol{B}_{\alpha}(\boldsymbol{r})}{\varepsilon_{1}(\boldsymbol{r})}\right]=\frac{\omega_{\alpha}^{2}}{c^{2}}\boldsymbol{B}_{\alpha}(\boldsymbol{r}),\\ & \boldsymbol{\nabla}\cdot\boldsymbol{B}_{\alpha}(\boldsymbol{r})=0,\\ & \boldsymbol{D}_{\alpha}(\boldsymbol{r})=\frac{i}{\omega_{\alpha}\mu_{0}}\boldsymbol{\nabla}\times\boldsymbol{B}_{\alpha}(\boldsymbol{r}). \end{aligned}$$
$$\left(\boldsymbol{D}_{\alpha'}\left(\boldsymbol{r}\right),\boldsymbol{B}_{\alpha'}\left(\boldsymbol{r}\right)\right)= \left(-\boldsymbol{D}_{\alpha}^{*}\left(\boldsymbol{r}\right),\boldsymbol{B}_{\alpha}^{*}\left(\boldsymbol{r}\right)\right)$$
$$\mathcal{M}(\cdots)\equiv\boldsymbol{\nabla}\times\left[\frac{\boldsymbol{\nabla}\times(\cdots)}{\varepsilon_{1}(\boldsymbol{r})}\right]$$
$$\int_{V}\boldsymbol{A}^{*}(\boldsymbol{r})\cdot\mathcal{M}(\boldsymbol{C}(\boldsymbol{r}))\textrm{d}\boldsymbol{r}=\int_{V}\left(\mathcal{M}(\boldsymbol{A}(\boldsymbol{r}))\right)^{*}\cdot\boldsymbol{C}(\boldsymbol{r})\textrm{d}\boldsymbol{r},$$
$$\begin{aligned} & \int_{V}\boldsymbol{B}_{\alpha}^{*}(\boldsymbol{r})\cdot\boldsymbol{B}_{\beta}(\boldsymbol{r})\textrm{d}\boldsymbol{r}=0,\\ & \omega_{\alpha}\neq\omega_{\beta}, \end{aligned}$$
$$\int_{V}\frac{\boldsymbol{B}_{\alpha}^{*}(\boldsymbol{r})\cdot\boldsymbol{B}_{\alpha}(\boldsymbol{r})}{\mu_{0}}\textrm{d}\boldsymbol{r}=1,$$
$$\int_{V}\frac{\boldsymbol{D}_{\alpha}^{*}(\boldsymbol{r})\cdot\boldsymbol{D}_{\alpha}(\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}\textrm{d}\boldsymbol{r}=1.$$
$$\left\{ f,g\right\} =\sum_{k}\left(\frac{\partial f}{\partial q_{k}}\frac{\partial g}{\partial p_{k}}-\frac{\partial f}{\partial p_{k}}\frac{\partial g}{\partial q_{k}}\right),$$
$$\begin{aligned} & \left\{ f,g\right\} ={-}\left\{ g,f\right\},\\ & \left\{ f+g,b\right\} =\left\{ f,b\right\} +\left\{ g,b\right\},\\ & \left\{ fg,b\right\} =f\left\{ g,b\right\} +g\left\{ f,b\right\},\\ & \left\{ f,\left\{ g,b\right\} \right\} +\left\{ g,\left\{ b,f\right\} \right\} +\left\{ b,\left\{ f,g\right\} \right\} =0. \end{aligned}$$
$$\begin{aligned} & \left\{ q_{i},q_{j}\right\} =\left\{ p_{i},p_{j}\right\} =0,\\ & \left\{ q_{i},p_{j}\right\} =\delta_{ij}. \end{aligned}$$
$$\frac{\textrm{d}f}{\textrm{d}t}=\left\{ f,H\right\}.$$
$$\left\{ J_{x},J_{y}\right\} =J_{z},$$
$$\textrm{H}^{\textrm{L}}=\int_{V}\left(\frac{\boldsymbol{D}(\boldsymbol{r},t)\cdot\boldsymbol{D}(\boldsymbol{r},t)}{2\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})} +\frac{\boldsymbol{B}(\boldsymbol{r},t)\cdot\boldsymbol{B}(\boldsymbol{r},t)}{2\mu_{0}}\right)\textrm{d}\boldsymbol{r},$$
$$\begin{aligned} & \left\{ D^{i}(\boldsymbol{r},t),D^{j}(\boldsymbol{r'},t)\right\} =\left\{ B^{i}(\boldsymbol{r},t),B^{j}(\boldsymbol{r'},t)\right\} =0,\\ & \left\{ D^{i}(\boldsymbol{r},t),B^{j}(\boldsymbol{r'},t)\right\} =\epsilon^{ilj}\frac{\partial}{\partial r^{l}}\delta(\boldsymbol{r}-\boldsymbol{r'}), \end{aligned}$$
$$\begin{aligned}& \frac{\partial\boldsymbol{D}(\boldsymbol{r},t)}{\partial t}=\left\{ \boldsymbol{D}(\boldsymbol{r},t),\textrm{H}^{\textrm{L}}\right\} =\frac{1}{\mu_{0}}\boldsymbol{\nabla}\times\boldsymbol{B}(\boldsymbol{r},t),\\ & \frac{\partial\boldsymbol{B}(\boldsymbol{r},t)}{\partial t}=\left\{ \boldsymbol{B}(\boldsymbol{r},t),\textrm{H}^{\textrm{L}}\right\} ={-}\boldsymbol{\nabla}\times\left(\frac{\boldsymbol{D}(\boldsymbol{r},t)}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}\right). \end{aligned}$$
$$\left\{ \;\right\} \Rightarrow\frac{1}{i\hbar}\left[\;\right],$$
$$\begin{aligned} & \left[D^{i}(\boldsymbol{r},t),D^{j}(\boldsymbol{r'},t)\right]=\left[B^{i}(\boldsymbol{r},t),B^{j}(\boldsymbol{r'},t)\right]=0,\\ & \left[D^{i}(\boldsymbol{r},t),B^{j}(\boldsymbol{r'},t)\right]=i\hbar\epsilon^{ilj}\frac{\partial}{\partial r^{l}}\delta(\boldsymbol{r}-\boldsymbol{r'}), \end{aligned}$$
$$\begin{aligned}& \boldsymbol{D}(\boldsymbol{r},t)=\sum_{\alpha}\mathcal{C}_{\alpha}^{(1)}(t)\boldsymbol{D}_{\alpha}(\boldsymbol{r}),\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{\alpha}\mathcal{C}_{\alpha}^{(2)}(t)\boldsymbol{B}_{\alpha}(\boldsymbol{r}), \end{aligned}$$
$$\begin{aligned}& \mathcal{C}_{\alpha'}^{(1)}(t)={-}\left(\mathcal{C}_{\alpha}^{(1)}(t)\right)^{{\dagger}},\\ & \mathcal{C}_{\alpha'}^{(2)}(t)={+}\left(\mathcal{C}_{\alpha}^{(2)}(t)\right)^{{\dagger}}, \end{aligned}$$
$$\begin{aligned}& \mathcal{C}_{\alpha}^{(1)}(t)=\sqrt{\frac{\hbar\omega_{\alpha}}{2}}\left(a_{\alpha}(t)-a_{\alpha'}^{{\dagger}}(t)\right),\\ & \mathcal{C}_{\alpha}^{(2)}(t)=\sqrt{\frac{\hbar\omega_{\alpha}}{2}}\left(a_{\alpha}(t)+a_{\alpha'}^{{\dagger}}(t)\right), \end{aligned}$$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\sum_{\alpha}\sqrt{\frac{\hbar\omega_{\alpha}}{2}}\left(a_{\alpha}(t)\boldsymbol{D}_{\alpha}(\boldsymbol{r})+a_{\alpha}^{{\dagger}}(t)\boldsymbol{D}_{\alpha}^{*}(\boldsymbol{r})\right),\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{\alpha}\sqrt{\frac{\hbar\omega_{\alpha}}{2}}\left(a_{\alpha}(t)\boldsymbol{B}_{\alpha}(\boldsymbol{r})+a_{\alpha}^{{\dagger}}(t)\boldsymbol{B}_{\alpha}^{*}(\boldsymbol{r})\right), \end{aligned}$$
$$\begin{aligned}& \left[a_{\alpha}(t),a_{\beta}(t)\right]=0,\\ & \left[a_{\alpha}(t),a_{\beta}^{{\dagger}}(t)\right]=\delta_{\alpha\beta}, \end{aligned}$$
$$\begin{aligned} \textrm{H}^{\textrm{L}} & =\sum_{\alpha}\frac{\hbar\omega_{\alpha}}{2}\left(a_{\alpha}^{{\dagger}}(t)a_{\alpha}(t)+a_{\alpha}(t)a_{\alpha}^{{\dagger}}(t)\right)\\ & =\sum_{\alpha}\hbar\omega_{\alpha}\left(a_{\alpha}^{{\dagger}}(t)a_{\alpha}(t)+\frac{1}{2}\right)\\ & \rightarrow\sum_{\alpha}\hbar\omega_{\alpha}a_{\alpha}^{{\dagger}}(t)a_{\alpha}(t), \end{aligned}$$
$$i\hbar\frac{\textrm{d}\mathcal{O}(t)}{\textrm{d}t}=\left[\mathcal{O}(t),\textrm{H}^{\textrm{L}}\right],$$
$$\frac{\textrm{d}a_{\alpha}(t)}{\textrm{d}t}={-}i\omega_{\alpha}a_{\alpha}(t),$$
$$\begin{aligned} & \boldsymbol{D}_{Ik}(\boldsymbol{r})=\frac{\boldsymbol{d}_{Ik}(y,z)}{\sqrt{L}}e^{ikx},\\ & \boldsymbol{B}_{Ik}(\boldsymbol{r})=\frac{\boldsymbol{b}_{Ik}(y,z)}{\sqrt{L}}e^{ikx}, \end{aligned}$$
$$\int\frac{\boldsymbol{b}_{Ik}^{*}(y,z)\cdot\boldsymbol{b}_{Ik}(y,z)}{\mu_{0}}\textrm{d}y\textrm{d}z=1,$$
$$\int\frac{\boldsymbol{d}_{Ik}^{*}(y,z)\cdot\boldsymbol{d}_{Ik}(y,z)}{\epsilon_{0}\varepsilon_{1}(y,z)}\textrm{d}y\textrm{d}z=1,$$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\sum_{I,k}\sqrt{\frac{\hbar\omega_{Ik}}{2L}}a_{Ik}(t)\boldsymbol{d}_{Ik}(y,z)e^{ikx}+\textrm{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I,k}\sqrt{\frac{\hbar\omega_{Ik}}{2L}}a_{Ik}(t)\boldsymbol{b}_{Ik}(y,z)e^{ikx}+\textrm{H.c.}, \end{aligned}$$
$$\begin{aligned} & \left[a_{Ik}(t),a_{I'k'}(t)\right]=0,\\ & \left[a_{Ik}(t),a_{I'k'}^{{\dagger}}(t)\right]=\delta_{II'}\delta_{kk'}, \end{aligned}$$
$$\textrm{H}^{\textrm{L}}=\sum_{I,k}\hbar\omega_{Ik}a_{Ik}^{{\dagger}}(t)a_{Ik}(t).$$
$$\sum_{k}\rightarrow\frac{\int \textrm{d}k}{\left(\frac{2\pi}{L}\right)}=\frac{L}{2\pi}\int \textrm{d}k.$$
$$\begin{aligned}& \sum_{k'}\left[a_{Ik}(t),a_{I'k'}^{{\dagger}}(t)\right]=\delta_{II'},\\ & \sum_{k'}\left[a_{Ik}(t),a_{I'k'}(t)\right]=0. \end{aligned}$$
$$a_{Ik}(t)\rightarrow\sqrt{\frac{2\pi}{L}}a_{I}(k,t),$$
$$\begin{aligned}& \int\left[a_{I}(k,t),a_{I}^{{\dagger}}(k',t)\right]\textrm{d}k'=\delta_{II'},\\ & \int\left[a_{I}(k,t),a_{I}(k',t)\right]\textrm{d}k'=0, \end{aligned}$$
$$\begin{aligned} & \left[a_{I}(k,t),a_{I'}(k',t)\right]=0,\\ & \left[a_{I}(k,t),a_{I'}^{{\dagger}}(k',t)\right]=\delta_{II'}\delta(k-k'). \end{aligned}$$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\sum_{I}\int\sqrt{\frac{\hbar\omega_{Ik}}{4\pi}}a_{I}(k,t)\boldsymbol{d}_{Ik}(y,z)e^{ikx}\textrm{d}k+\textrm{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I}\int\sqrt{\frac{\hbar\omega_{Ik}}{4\pi}}a_{I}(k,t)\boldsymbol{b}_{Ik}(y,z)e^{ikx}\textrm{d}k+\textrm{H.c.}, \end{aligned}$$
$$\textrm{H}_{\textrm{chan}}^{\textrm{L}}=\sum_{I}\int\hbar\omega_{Ik}a_{I}^{{\dagger}}(k,t)a_{I}(k,t)\textrm{d}k.$$
$$\begin{aligned} & \boldsymbol{D}_{J}(\boldsymbol{r})=\frac{\textsf{d}_{J}(\boldsymbol{r}_{{\perp}};\zeta)}{\sqrt{\mathcal{L}}}e^{i\kappa_{J}\zeta},\\ & \boldsymbol{B}_{J}(\boldsymbol{r})=\frac{\textsf{b}_{J}(\boldsymbol{r}_{{\perp}};\zeta)}{\sqrt{\mathcal{L}}}e^{i\kappa_{J}\zeta}, \end{aligned}$$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2\mathcal{L}}}c_{J}(t)\textsf{d}_{J}(\boldsymbol{r}_{{\perp}},\zeta)e^{i\kappa_{J}\zeta}+\textrm{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2\mathcal{L}}}c_{J}(t)\textsf{b}_{J}(\boldsymbol{r}_{{\perp}},\zeta)e^{i\kappa_{J}\zeta}+\textrm{H.c.}, \end{aligned}$$
$$\begin{aligned} & \left[c_{J}(t),c_{J'}(t)\right]=0,\\ & \left[c_{J}(t),c_{J'}^{{\dagger}}(t)\right]=\delta_{JJ'}, \end{aligned}$$
$${H}_{\textrm{ring}}^{\textrm{L}}=\sum_{J}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J},$$
$$\int\frac{\textsf{b}_{J}^{*}(\boldsymbol{r}_{{\perp}};0)\cdot\textsf{b}_{J}(\boldsymbol{r}_{{\perp}};0)}{\mu_{0}}\textrm{d}\boldsymbol{r_{{\perp}}}=1,$$
$$\int\frac{\textsf{d}_{J}^{*}(\boldsymbol{r}_{{\perp}};0)\cdot\textsf{d}_{J}(\boldsymbol{r}_{{\perp}};0)}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r_{{\perp}}})}\textrm{d}\boldsymbol{r}_{{\perp}}=1.$$
$$\begin{aligned} & \textsf{d}_{J}(\rho,z;0)\approx\boldsymbol{d}_{I_{\alpha}k_{\alpha}}(R-\rho,z),\\ & \textsf{b}_{J}(\rho,z;0)\approx\boldsymbol{b}_{I_{\alpha}k_{\alpha}}(R-\rho,z), \end{aligned}$$
$$\begin{aligned} & \boldsymbol{E}(\boldsymbol{r},t)=\boldsymbol{E}(\boldsymbol{r})e^{{-}i\omega t}+\textrm{c.c.},\\ & \boldsymbol{D}(\boldsymbol{r},t)=\boldsymbol{D}(\boldsymbol{r})e^{{-}i\omega t}+\textrm{c.c.}, \end{aligned}$$
$$\boldsymbol{D}(\boldsymbol{r})=\epsilon_{0}\varepsilon_{1}(\boldsymbol{r};\omega)\boldsymbol{E}(\boldsymbol{r}).$$
$$\begin{aligned} & \boldsymbol{D}_{\alpha}(\boldsymbol{r})=\epsilon_{0}\varepsilon_{1}(\boldsymbol{r},\omega_{\alpha})\boldsymbol{E}_{\alpha}(\boldsymbol{r}),\\ & \boldsymbol{B}_{\alpha}(\boldsymbol{r})=\mu_{0}\boldsymbol{H}_{\alpha}(\boldsymbol{r}), \end{aligned}$$
$$\begin{aligned} & \boldsymbol{\nabla}\times\left[\frac{\boldsymbol{\nabla}\times\boldsymbol{B}_{\alpha}(\boldsymbol{r})}{\varepsilon_{1}(\boldsymbol{r};\omega_{\alpha})}\right]=\frac{\omega_{\alpha}^{2}}{c^{2}}\boldsymbol{B}_{\alpha}(\boldsymbol{r}),\\ & \boldsymbol{\nabla}\cdot\boldsymbol{B}_{\alpha}(\boldsymbol{r})=0,\\ & \boldsymbol{D}_{\alpha}(\boldsymbol{r})=\frac{i}{\omega_{\alpha}\mu_{0}}\boldsymbol{\nabla}\times\boldsymbol{B}_{\alpha}(\boldsymbol{r}). \end{aligned}$$
$$\int_{V}\frac{\boldsymbol{D}_{\alpha}^{*}(\boldsymbol{r})\cdot\boldsymbol{D}_{\alpha}(\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r};\omega_{\alpha})} \frac{v_{p}(\boldsymbol{r};\omega_{\alpha})}{v_{g}(\boldsymbol{r};\omega_{\alpha})}\textrm{d}\boldsymbol{r}=1$$
$$v_{p}(\boldsymbol{r};\omega)=\frac{c}{n(\boldsymbol{r};\omega)}$$
$$v_{g}(\boldsymbol{r};\omega)=\frac{v_{p}(\boldsymbol{r};\omega)}{1+\frac{\omega}{n(\boldsymbol{r};\omega)}\frac{\partial n(\boldsymbol{r};\omega)}{\partial\omega}}$$
$$\int\frac{\boldsymbol{d}_{Ik}^{*}(y,z)\cdot\boldsymbol{d}_{Ik}(y,z)}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\frac{v_{p}(y,z;\omega_{Ik})}{v_{g}(y,z;\omega_{Ik})}\textrm{d}y\textrm{d}z=1.$$
$$\left|\Psi\right\rangle =a_{Ik}^{{\dagger}}(0)\left\vert\text{vac}\right\rangle$$
$$\int\left\langle \Psi|\boldsymbol{S}(\boldsymbol{r},t)|\Psi\right\rangle \cdot\boldsymbol{\hat{x}}\textrm{d}y\textrm{d}z=\frac{\hbar\omega_{Ik}}{L}v_{Ik},$$
$$v_{Ik}\equiv\frac{\partial\omega_{Ik}}{\partial k}=\frac{1}{2}\int\left(\boldsymbol{e}_{Ik}(y,z)\times\boldsymbol{h}_{Ik}^{*}(y,z)+\boldsymbol{e}_{Ik}^{*}(y,z)\times\boldsymbol{h}_{Ik}(y,z)\right)\cdot\boldsymbol{\hat{x}}\textrm{d}y\textrm{d}z$$
$$\begin{aligned} & \boldsymbol{e}_{Ik}(y,z)=\frac{\boldsymbol{d}_{Ik}(y,z)}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})},\\ & \boldsymbol{h}_{Ik}(y,z)=\frac{\boldsymbol{b}_{Ik}(y,z)}{\mu_{0}}, \end{aligned}$$
$$\int\frac{\textsf{d}_{J}^{*}(\boldsymbol{r}_{{\perp}};0)\cdot\textsf{d}_{J}(\boldsymbol{r}_{{\perp}};0)}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r_{{\perp}}};\omega_{J})} \frac{v_{p}(\boldsymbol{r}_{{\perp}};\omega_{J})}{v_{g}(\boldsymbol{r}_{{\perp}};\omega_{J})}\textrm{d}\boldsymbol{r}_{{\perp}}=1.$$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=\sum_{J}\int_{k\in\mathscr{R}(J)}\sqrt{\frac{\hbar\omega_{Jk}}{4\pi}}a_{J}(k,t)\boldsymbol{d}_{Jk}(y,z)e^{ikx}\textrm{d}k+\textrm{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{J}\int_{k\in\mathscr{R}(J)}\sqrt{\frac{\hbar\omega_{Jk}}{4\pi}}a_{J}(k,t)\boldsymbol{b}_{Jk}(y,z)e^{ikx}\textrm{d}k+\textrm{H.c.}, \end{aligned}$$
$$\begin{aligned} & \left[a_{J}(k,t),a_{J'}(k',t)\right]=0,\\ & \left[a_{J}(k,t),a_{J'}^{{\dagger}}(k',t)\right]=\delta_{JJ'}\delta(k-k'), \end{aligned}$$
$$\textrm{H}_{\textrm{chan}}^{\textrm{L}}=\sum_{J}H_{J}^{\text{L}},$$
$$H_{J}^{\text{L}}=\int_{k\in\mathscr{R}(J)}\hbar\omega_{Jk}a_{J}^{{\dagger}}(k,t)a_{J}(k,t)\textrm{d}k.$$
$$\psi_{J}(x,t)\equiv\int_{k\in\mathscr{R}(J)}\frac{\textrm{d}k}{\sqrt{2\pi}}a_{J}(k,t)e^{i(k-k_{J})x},$$
$$\begin{aligned} \frac{\partial\psi_{J}(x,t)}{\partial t} & ={-}i\int_{k\in\mathscr{R}(J)}\frac{\textrm{d}k}{\sqrt{2\pi}}\omega_{Jk}a_{J}(k,t)e^{i(k-k_{J})x}\\ & ={-}i\int_{k\in\mathscr{R}(J)}\frac{\textrm{d}k}{\sqrt{2\pi}}\left(\omega_{J}+v_{J}(k-k_{J})+\frac{1}{2}v'_{J}(k-k_{J})^{2}+\cdots\right)a_{J}(k,t)e^{i(k-k_{J})x}, \end{aligned}$$
$$v'_{Jk}=\frac{\partial^{2}\omega_{Jk}}{\partial k^{2}}.$$
$$\frac{\partial\psi_{J}(x,t)}{\partial t}={-}i\omega_{J}\psi_{J}(x,t)-v_{J}\frac{\partial\psi_{J}(x,t)}{\partial x}+\frac{1}{2}iv'_{J}\frac{\partial^{2}\psi_{J}(x,t)}{\partial x^{2}}+\cdots.$$
$$\frac{\partial\psi_{J}(x,t)}{\partial t}={-}i\omega_{J}\psi_{J}(x,t)-v_{J}\frac{\partial\psi_{J}(x,t)}{\partial x}.$$
$$\left[\psi_{J}(x,t),\psi_{J'}(x',t)\right]=0,$$
$$\left[\psi_{J}(x,t),\psi_{J'}^{{\dagger}}(x',t)\right]=0,$$
$$\left[\psi_{J}(x,t),\psi_{J}^{{\dagger}}(x',t)\right]=\int_{k\in\mathscr{R}(J)}\frac{\textrm{d}k}{2\pi}e^{i(k-k_{J})(x-x')},$$
$$\left[\psi_{J}(x,t),\psi_{J}^{{\dagger}}(x',t)\right]=\frac{1}{2\pi}\frac{\sin\left(\Delta\kappa\frac{x-x'}{2}\right)}{\frac{x-x'}{2}}.$$
$$\left[\psi_{J}(x,t),\psi_{J}^{{\dagger}}(x',t)\right]\approx\delta(x-x'),$$
$$\begin{aligned} H_{J}^{\text{L}} & =\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x,t)\psi_{J}(x,t)\textrm{d}x\\ & \quad -\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x,t)\frac{\partial\psi_{J}(x,t)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x,t)}{\partial x}\psi_{J}(x,t)\right)\textrm{d}x, \end{aligned}$$
$$\begin{aligned} & \left[\psi_{J}(x,t),\psi_{J'}^{{\dagger}}(x',t)\right]=\delta_{JJ'}\delta(x-x'),\\ & \left[\psi_{J}(x,t),\psi_{J'}(x',t)\right]=0, \end{aligned}$$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)\approx\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2}}\boldsymbol{d}_{J}(y,z)\psi_{J}(x,t)e^{ik_{J}x}+\textrm{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)\approx\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2}}\boldsymbol{b}_{J}(y,z)\psi_{J}(x,t)e^{ik_{J}x}+\textrm{H.c.} \end{aligned}$$
$$\begin{aligned} & \boldsymbol{E}(\boldsymbol{r},t)\approx\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2}}\boldsymbol{e}_{J}(y,z)\psi_{J}(x,t)e^{ik_{J}x}+\textrm{H.c.},\\ & \boldsymbol{H}(\boldsymbol{r},t)\approx\sum_{J}\sqrt{\frac{\hbar\omega_{J}}{2}}\boldsymbol{h}_{J}(y,z)\psi_{J}(x,t)e^{ik_{J}x}+\textrm{H.c.}, \end{aligned}$$
$$\begin{aligned} & \boldsymbol{e}_{J}(y,z)=\frac{\boldsymbol{d}_{J}(y,z)}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{J})},\\ & \boldsymbol{h}_{J}(y,z)=\frac{\boldsymbol{b}_{J}(y,z)}{\mu_{0}}, \end{aligned}$$
$$P(x,t)=\int \textrm{d}y\textrm{d}z\boldsymbol{S}(\boldsymbol{r},t)\cdot\boldsymbol{\hat{x}},$$
$$P(x,t)\rightarrow\sum_{J}\hbar\omega_{J}v_{J}\psi_{J}^{{\dagger}}(x,t)\psi_{J}(x,t),$$
$$\psi_{J}(x,\omega)=\int\frac{\textrm{d}t}{\sqrt{2\pi}}\psi_{J}(x,t)e^{i\omega t}, \psi_{r}(\zeta,\omega)=\int\frac{\textrm{d}t}{\sqrt{2\pi}}\psi_{r}(\zeta,t)e^{i\omega t},$$
$$\begin{aligned} \begin{pmatrix} \psi_{J}(0^{+},\omega) \\ \psi_{r}(0^{+},\omega) \end{pmatrix} = \left( \begin{array}{cc} \ \sigma & i\kappa \\ i\kappa & \sigma \\ \end{array} \right) \begin{pmatrix} \psi_{J}(0^{-},\omega) \\ \psi_{r}(0^{-},\omega) \end{pmatrix}, \end{aligned}$$
$$\sigma^{2}+\kappa^{2}=1.$$
$$\psi_{r}(0^{-},\omega)=\psi_{r}(0^{+},\omega) e^{ik\mathcal{L}},$$
$$ \begin{aligned}k &=\frac{2\pi n_{J}}{\mathcal{L}}+(\omega-\omega_{J}) \left (\frac{dk}{d\omega} \right)_{\omega_{J}}+\cdots\\ &= \frac{2\pi n_{J}}{\mathcal{L}}+\frac{\omega-\omega_{J}}{v_{J}}+\cdots. \end{aligned}$$
$$\eta=\left(\frac{\omega-\omega_J}{v_J} \right) \mathcal{L}.$$
$$\psi_{J}(0^{+},\omega)=\left( \frac{\sigma-e^{i\eta}}{1-\sigma e^{i\eta}} \right) \psi_{J}(0^{-},\omega),$$
$$\psi_{J}(0^{+},\omega) \approx \frac{\omega-\omega_{J}-i \left( \frac{(1-\sigma) v_{J}}{ \mathcal{L}} \right)} {\omega-\omega_{J}+i \left( \frac{(1-\sigma) v_{J}}{ \mathcal{L}} \right)} \psi_{J}(0^{-},\omega).$$
$$\begin{aligned} \textrm{H}^{\textrm{L}} & =\sum_{J}\left(\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x)\psi_{J}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x)\frac{\partial\psi_{J}(x)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x)}{\partial x}\psi_{J}(x)\right)\textrm{d}x\right)\\ & \quad +\sum_{J}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J}+\sum_{J}\left(\hbar\gamma_{J}^*c_{J}^{{\dagger}}\psi_{J}\left(0\right)+\textrm{H.c.}\right). \end{aligned}$$
$$\begin{aligned} & \left(\frac{\partial}{\partial t}+v_{J}\frac{\partial}{\partial x}+i\omega_{J}\right)\psi_{J}(x,t)={-}i\gamma_{J}c_{J}(t)\delta(x),\\ & \left(\frac{d}{dt}+i\omega_{J}\right)c_{J}(t)={-}i\gamma_{J}^{*}\psi_{J}(0,t). \end{aligned}$$
$$\psi_{J}(x,t)=\psi_{J}(x-v_Jt,0)e^{{-}i\omega_{J}t}-\frac{i\gamma_{J}}{v_{J}}c_{J}(t-\tfrac{x}{v_{J}})\left[\theta(x)-\theta(x-v_{J}t)\right] e^{{-}i\omega_{J}x/v_{J}}.$$
$$\psi_{J}(0^{{\pm}},t)=\lim_{x\rightarrow0^{{\pm}}}\psi_{J}(x,t),$$
$$\psi_{J}(0,t)\rightarrow\frac{1}{2}\left(\psi_{J}(0^{-},t)+\psi_{J}(0^{+},t)\right);$$
$$\psi_{J}(0^{+},t)=\psi_{J}(0^{-},t)-\frac{i\gamma_{J}}{v_{J}}c_{J}(t),$$
$$\left(\frac{\textrm{d}}{\textrm{d}t}+\Gamma_{J}+i\omega_{J}\right)c_{J}(t)={-}i\gamma_{J}^{*}\psi_{J}(0^{-},t),$$
$$\Gamma_{J}=\frac{\left|\gamma_{J}\right|^{2}}{2v_{J}},$$
$$\psi_{J}(0^{+},\omega)=\left[\frac{\omega_{J}-\omega+i\Gamma_{J}}{\omega_{J}-\omega-i\Gamma_{J}}\right]\psi_{J}(0^{-},\omega),$$
$$\frac{\left|\gamma_{J}\right|^{2}}{2v_{J}}=\frac{(1-\sigma)v_J}{\mathcal{L}}.$$
$$\textrm{H}^{\textrm{L}}\rightarrow \textrm{H}^{\textrm{L}}+{\textrm{H}^{\textrm{phantom}}},$$
$$\begin{aligned} {\textrm{H}^{\textrm{phantom}}} & =\sum_{J}\left(\hbar\omega_{J}\int\psi_{J\text{ph}}^{{\dagger}}(x)\psi_{J\text{ph}}(x)\textrm{d}x \right.\\&\quad\left.-\frac{1}{2}i\hbar v_{J\text{ph}}\int\left(\psi_{J\text{ph}}^{{\dagger}}(x)\frac{\partial\psi_{J\text{ph}}(x)}{\partial x}-\frac{\partial\psi_{J\text{ph}}^{{\dagger}}(x)}{\partial x}\psi_{J\text{ph}}(x)\right)\textrm{d}x\right)\\ & \quad +\sum_{J}(\hbar \gamma_{J\text{ph}}c_{J}^{{\dagger}}\psi_{J\text{ph}}(0)+\textrm{H.c.}). \end{aligned}$$
$$\begin{aligned}& \left[\psi_{J\text{ph}}(x,t),\psi_{J'\text{ph}}^{{\dagger}}(x',t)\right]=\delta_{JJ'}\delta(x-x'),\\ & \left[\psi_{J\text{ph}}(x,t),\psi_{J'\text{ph}}(x',t)\right]=0, \end{aligned}$$
$$\begin{aligned} & \left(\frac{\partial}{\partial t}+v_{J}\frac{\partial}{\partial x}+i\omega_{J}\right)\psi_{J}(x,t)={-}i\gamma_{J}c_{J}(t)\delta(x),\\ & \left(\frac{\partial}{\partial t}+v_{J\text{ph}}\frac{\partial}{\partial x}+i\omega_{J}\right)\psi_{J\text{ph}}(x,t)={-}i\gamma_{J\text{ph}}c_{J}(t)\delta(x),\\ & \left(\frac{\textrm{d}}{\textrm{d}t}+i\omega_{J}\right)c_{J}(t)={-}i\gamma_{J}^{*}\psi_{J}(0,t)-i\gamma_{J\text{ph}}^{*}\psi_{J\text{ph}}(0,t). \end{aligned}$$
$$\begin{aligned}& \psi_{J\text{ph}}(0^{{\pm}},\omega)=\int\frac{\textrm{d}t}{\sqrt{2\pi}}\psi_{J\text{ph}}(0^{{\pm}},t)e^{i\omega t},\\ & c_{J}(\omega)=\int\frac{\textrm{d}t}{\sqrt{2\pi}}c_{J}(t)e^{i\omega t}, \end{aligned}$$
$$c_{J}(\omega)={-} \frac{\gamma_{J}^{*}\psi_{J}(0^{-},\omega)+\gamma_{J\text{ph}}^{*}\psi_{J\text{ph}}(0^{-},\omega)}{\omega_{J}-\omega-i\overline{\Gamma}_{J}},$$
$$\overline{\Gamma}_{J}=\Gamma_{J}+\Gamma_{J\text{ph}},$$
$$\Gamma_{J\text{ph}}=\frac{\left|\gamma_{J\text{ph}}\right|^{2}}{2v_{J\text{ph}}},$$
$$|\left\langle c_{J}(\omega)\right\rangle|^{2}=\frac{|\gamma_{J}^{*}\left\langle\psi_{J}(0^{-},\omega)\right\rangle|^{2}}{(\omega_{J}-\omega)^{2}+\overline{\Gamma}_{J}^{2}}.$$
$$Q_{J}=\frac{\omega_{J}}{\Delta\omega_{J}}=\frac{\omega_{J}}{2\overline{\Gamma}_{J}}.$$
$$\frac{1}{Q_{J}}=\frac{1}{Q_{J\text{int}}}+\frac{1}{Q_{J\text{ext}}},$$
$$Q_{J\text{ext}}=\frac{\omega_{J}}{2{\Gamma}_{J}}$$
$$Q_{J\text{int}}=\frac{\omega_{J}}{2{\Gamma}_{J\text{ph}}}.$$
$$\psi_{J}(0^{+},\omega)=\left[\frac{\omega_{J}-\omega+i(\Gamma_{J}-\Gamma_{J\text{ph}})}{\omega_{J}-\omega-i\overline{\Gamma}_{J}}\right],\,\psi_{J}(0^{-},\omega) +\left[\frac{2i\Gamma_{J\text{ph}}}{\omega_{J}-\omega-i\overline{\Gamma}_{J}}\right]\psi_{J\text{ph}}(0^{-},\omega)$$
$$\left\langle \psi_{J}(0^{+},\omega)\right\rangle =\left[\frac{\omega_{J}-\omega+i(\Gamma_{J}-\Gamma_{J\text{ph}})}{\omega_{J}-\omega-i\overline{\Gamma}_{J}}\right]\left\langle \psi_{J}(0^{-},\omega)\right\rangle,$$
$$\delta\omega_{J}=v_{J}\delta\kappa_{J}=v_J\frac{2\pi}{\mathcal{L}},$$
$$\mathcal{F}=\frac{\delta\omega_{J}}{\Delta\omega_{J}}=\frac{\pi v_{J}}{\overline{\Gamma}_{J} \mathcal{L}}.$$
$$\left\vert\text{FE}_\text{crit}\left(\omega_J\right)\right\vert^{2}\simeq\frac{\mathcal{F}}{\pi},$$
$$\left\vert\text{FE}_\text{over}\left(\omega_J\right)\right\vert^{2}\simeq\frac{2}{\pi}\mathcal{F}.$$
$$P_{\text{NL}}^{i}(\boldsymbol{r},t)=\epsilon_{0}\chi_{2}^{ijk}(\boldsymbol{r})E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t) +\epsilon_{0}\chi_{3}^{ijkl}(\boldsymbol{r})E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t)E^{l}(\boldsymbol{r},t)+\cdots.$$
$$D^{i}(\boldsymbol{r},t)=\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})E^{i}(\boldsymbol{r},t)+\epsilon_{0}\chi_{2}^{ijk}(\boldsymbol{r})E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t) +\epsilon_{0}\chi_{3}^{ijkl}(\boldsymbol{r})E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t)E^{l}(\boldsymbol{r},t)+\cdots.$$
$$E^{i}(\boldsymbol{r},t)=\frac{D^{i}(\boldsymbol{r},t)}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}-\frac{\chi_{2}^{ijk}(\boldsymbol{r})}{\varepsilon_{1}(\boldsymbol{r})} E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t)-\frac{\chi_{3}^{ijkl}(\boldsymbol{r})}{\varepsilon_{1}(\boldsymbol{r})}E^{j}(\boldsymbol{r},t)E^{k}(\boldsymbol{r},t)E^{l}(\boldsymbol{r},t)+\cdots.$$
$$E^{i}(\boldsymbol{r},t)=\frac{D^{j}(\boldsymbol{r},t)}{\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}-\frac{\Gamma_{2}^{ijk}(\boldsymbol{r})}{\epsilon_{0}}D^{j} (\boldsymbol{r},t)D^{k}(\boldsymbol{r},t)-\frac{\Gamma_{3}^{ijkl}(\boldsymbol{r})}{\epsilon_{0}}D^{j}(\boldsymbol{r},t)D^{k}(\boldsymbol{r},t)D^{l}(\boldsymbol{r},t)+\cdots,$$
$$\begin{aligned} & \Gamma_{2}^{ijk}(\boldsymbol{r})=\frac{\chi_{2}^{ijk}(\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}^{3}(\boldsymbol{r})},\\ & \Gamma_{3}^{ijkl}(\boldsymbol{r})=\frac{\chi_{3}^{ijkl}(\boldsymbol{r})}{\epsilon_{0}^{2}\varepsilon_{1}^{4}(\boldsymbol{r})}-2\frac{\chi_{2}^{ijm}(\boldsymbol{r}) \chi_{2}^{mkl}(\boldsymbol{r})}{\epsilon_{0}^{2}\varepsilon_{1}^{5}(\boldsymbol{r})}, \end{aligned}$$
$$\begin{aligned}H & =\int\left(\frac{\boldsymbol{D}(\boldsymbol{r},t)\cdot\boldsymbol{D}(\boldsymbol{r},t)}{2\epsilon_{0}\varepsilon_{1}(\boldsymbol{r})}+\frac{\boldsymbol{B}(\boldsymbol{r},t)\cdot\boldsymbol{B}(\boldsymbol{r},t)}{2\mu_{0}}\right)\textrm{d}\boldsymbol{r}\\ & \quad -\frac{1}{3\epsilon_{0}}\int\Gamma_{2}^{ijk}(\boldsymbol{r})D^{i}(\boldsymbol{r},t)D^{j}(\boldsymbol{r},t)D^{k}(\boldsymbol{r},t)\textrm{d}\boldsymbol{r}\\ & \quad -\frac{1}{4\epsilon_{0}}\int\Gamma_{3}^{ijkl}(\boldsymbol{r})D^{i}(\boldsymbol{r},t)D^{j}(\boldsymbol{r},t)D^{k}(\boldsymbol{r},t)D^{l}(\boldsymbol{r},t)\textrm{d}\boldsymbol{r}+\cdots, \end{aligned}$$
$$\begin{aligned}& \frac{\partial\boldsymbol{D}(\boldsymbol{r},t)}{\partial t}=\frac{1}{\mu_{0}}\boldsymbol{\nabla}\times\boldsymbol{B}(\boldsymbol{r},t),\\ & \frac{\partial\boldsymbol{B}(\boldsymbol{r},t)}{\partial t}={-}\boldsymbol{\nabla}\times\boldsymbol{E}(\boldsymbol{r},t), \end{aligned}$$
$$H=\sum_{\alpha}\hbar\omega_{\alpha}c_{\alpha}^{{\dagger}}c_{\alpha}+{\textrm{H}^{\textrm{NL}}},$$
$$\begin{aligned} {\textrm{H}^{\textrm{NL}}} & ={-}\frac{1}{3\epsilon_{0}}\int\Gamma_{2}^{ijk}(\boldsymbol{r}):\left(D^{i}(\boldsymbol{r})D^{j}(\boldsymbol{r})D^{k}(\boldsymbol{r})\right):\textrm{d}\boldsymbol{r}\\ & \quad -\frac{1}{4\epsilon_{0}}\int\Gamma_{3}^{ijkl}(\boldsymbol{r}):\left(D^{i}(\boldsymbol{r})D^{j}(\boldsymbol{r})D^{k}(\boldsymbol{r})D^{l}(\boldsymbol{r})\right):\textrm{d}\boldsymbol{r}+\cdots. \end{aligned}$$
$$\begin{aligned} {H}_{\textrm{chan}}^{\textrm{SPM}} & ={-}\frac{1}{4\epsilon_{0}}\left(\frac{\hbar\omega_{P}}{2}\right)^{2}\frac{4!}{2!2!}\left(\int \textrm{d}y\textrm{d}z\Gamma_{3}^{ijkl}(y,z)\left(d_{P}^{i}(y,z)d_{P}^{j}(y,z)\right)^{*}d_{P}^{k}(y,z)d_{P}^{l}(y,z)\right)\\ & \quad\times\int \textrm{d}x\psi_{P}^{{\dagger}}(x)\psi_{P}^{{\dagger}}(x)\psi_{P}(x)\psi_{P}(x), \end{aligned}$$
$${H}_{\textrm{chan}}^{\textrm{SPM}} ={-}{\gamma_{\textrm{chan}}^{\textrm{SPM}}}\frac{\hbar^{2}\omega_{P}}{2}v_{P}^{2}\int \textrm{d}x\psi_{P}^{{\dagger}}(x)\psi_{P}^{{\dagger}}(x)\psi_{P}(x)\psi_{P}(x),$$
$${\gamma_{\textrm{chan}}^{\textrm{SPM}}}=\frac{3\omega_{P}\epsilon_{0}}{4v_{P}^{2}}\int \textrm{d}y\textrm{d}z\chi_{3}^{ijkl}(y,z)\left(e_{P}^{i}(y,z)e_{P}^{j}(y,z)\right)^{*}e_{P}^{k}(y,z)e_{P}^{l}(y,z).$$
$$\frac{\partial\psi_{P}(x,t)}{\partial t}={-}i\omega_{P}\psi_{P}(x,t)-v_{P}\frac{\partial\psi_{P}(x,t)}{\partial x}+i{\gamma_{\textrm{chan}}^{\textrm{SPM}}}\hbar\omega_{P}v_{P}^{2}\psi_{P}^{{\dagger}}(x,t)\psi_{P}(x,t)\psi_{P}(x,t).$$
$$\begin{aligned} & \psi_{P}(x,t)\rightarrow\phi_{P}(x,t),\\ & \psi_{P}^{{\dagger}}(x,t)\rightarrow\phi_{P}^{*}(x,t), \end{aligned}$$
$$\begin{aligned} \frac{\textrm{d}\phi_{P}(x)}{\textrm{d}x} & =i{\gamma_{\textrm{chan}}^{\textrm{SPM}}}\hbar\omega_{P}v_{P}\left|\phi_{P}(x)\right|^{2}\phi_{P}(x)\\ & =i{\gamma_{\textrm{chan}}^{\textrm{SPM}}} P_{P}(x)\phi_{P}(x), \end{aligned}$$
$$\epsilon_{0}c\int\frac{n(y,z;\omega_{P})}{v_{g}(y,z;\omega_{P})}\boldsymbol{e}_{P}^{*}(y,z)\cdot\boldsymbol{e}_{P}(y,z)\textrm{d}y\textrm{d}z=1,$$
$${\gamma_{\textrm{chan}}^{\textrm{SPM}}}=\frac{3\omega_{P}\overline{\chi}_{3}}{4\epsilon_{0}\overline{n}_{P}^{2}c^{2}}\frac{1}{{\textrm{A}_{\textrm{chan}}^{PPPP}}},$$
$$\frac{1}{{\textrm{A}_{\textrm{chan}}^{PPPP}}}=\frac{\int \textrm{d}y\textrm{d}z\left(\chi_{3}^{ijkl}(y,z)/\overline{\chi}_{3}\right)\left(e_{P}^{i}(y,z)e_{P}^{j}(y,z)\right)^{*}e_{P}^{k}(y,z)e_{P}^{l} (y,z)}{\left(\int\frac{n(y,z;\omega_{P})/\overline{n}_{P}}{v_{g}(y,z;\omega_{P})/v_{P}}\boldsymbol{e}_{P}^{*}(y,z)\cdot\boldsymbol{e}_{P}(y,z)\textrm{d}y\textrm{d}z\right)^{2}}.$$
$$n_{2}=\frac{3\overline{\chi}_{3}}{4\epsilon_{0}c\overline{n}_{P}^{2}},$$
$${\gamma_{\textrm{chan}}^{\textrm{SPM}}}=\frac{\omega_{P}}{c}n_{2}\frac{1}{{\textrm{A}_{\textrm{chan}}^{PPPP}}},$$
$$\frac{1}{{\textrm{A}_{\textrm{chan}}^{PPPP}}}\approx\frac{\int\left|e_{P}(y,z)\right|^{4}\textrm{d}y\textrm{d}z}{\left(\int\left|e_{P}(y,z)\right|^{2}\textrm{d}y\textrm{d}z\right)^{2}},$$
$$\begin{aligned} {H}_{\textrm{chan}}^{\textrm{XPM}} & ={-}\frac{1}{4\epsilon_{0}}\left(\frac{\hbar\omega_{P}}{2}\right)\left(\frac{\hbar\omega_{S}}{2}\right)4!\left(\int \textrm{d}y\textrm{d}z\Gamma_{3}^{ijkl}(y,z)\left(d_{P}^{i}(y,z)d_{S}^{j}(y,z)\right)^{*}d_{P}^{k}(y,z)d_{S}^{l}(y,z)\right)\\ & \quad\times\int \textrm{d}x\psi_{P}^{{\dagger}}(x)\psi_{S}^{{\dagger}}(x)\psi_{P}(x)\psi_{S}(x), \end{aligned}$$
$${\gamma_{\text{chan}}^{\text{XPM}}}=\frac{3\epsilon_{0}\omega_{S}}{4v_{P}v_{S}}\int \textrm{d}y\textrm{d}z\chi_{3}^{ijkl}(y,z)\left(e_{P}^{i}(y,z)e_{S}^{j}(y,z)\right)^{*}e_{P}^{k}(y,z)e_{S}^{l}(y,z),$$
$${H}_{\textrm{chan}}^{\textrm{XPM}}={-}2{\gamma_{\text{chan}}^{\text{XPM}}}\hbar^{2}\omega_{P}v_{P}v_{S}\int \textrm{d}x\psi_{P}^{{\dagger}}(x)\psi_{S}^{{\dagger}}(x)\psi_{P}(x)\psi_{S}(x).$$
$$\begin{aligned}\frac{\textrm{d}\phi_{S}(x)}{\textrm{d}x} & =2i{\gamma_{\text{chan}}^{\text{XPM}}}\hbar\omega_{P}v_{P}\left|\phi_{P}(x)\right|^{2}\phi_{S}(x)\\ & =2i{\gamma_{\text{chan}}^{\text{XPM}}} P_{P}(x)\phi_{S}(x), \end{aligned}$$
$${\gamma_{\text{chan}}^{\text{XPM}}}=\frac{3\omega_{S}\overline{\chi}_{3}}{4\epsilon_{0}\overline{n}_{P}\overline{n}_{S}c^{2}}\frac{1}{{\textrm{A}_{\textrm{chan}}^{PSPS}}},$$
$$\frac{1}{{\textrm{A}_{\textrm{chan}}^{PSPS}}}=\frac{\int \textrm{d}y\textrm{d}z\left(\chi_{3}^{ijkl}(y,z)/\overline{\chi}_{3}\right)\left(e_{P}^{i}(y,z)e_{S}^{j}(y,z)\right)^{*}e_{P}^{k}(y,z)e_{S}^{l} (y,z)}{\left(\int\frac{n(y,z;\omega_{P})/\overline{n}_{P}}{v_{g}(y,z;\omega_{P})/v_{P}}\boldsymbol{e}_{P}^{*}(y,z)\cdot\boldsymbol{e}_{P}(y,z)\textrm{d}y\textrm{d}z\right)\left(\int\frac{n(y,z;\omega_{S})/\overline{n}_{S}}{v_{g}(y,z;\omega_{S})/v_{S}}\boldsymbol{e}_{S}^{*}(y,z)\cdot\boldsymbol{e}_{S}(y,z)\textrm{d}y\textrm{d}z\right)},$$
$${\begin{aligned} H_{\text{ring}}^{\text{SPM}} & ={-}\frac{1}{4\epsilon_{0}}\left(\frac{\hbar\omega_{P}}{2\mathcal{L}}\right)^{2}\frac{4!}{2!2!}\left(\int \textrm{d}\boldsymbol{r}_{{\perp}}d\zeta\Gamma_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})\left(\textsf{d}_{P}^{i}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{d}_{P}^{j} (\boldsymbol{r}_{{\perp}};\zeta)\right)^{*}\textsf{d}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{d}_{P}^{l}(\boldsymbol{r}_{{\perp}};\zeta)\right)\\ & \quad\times c_{P}^{{\dagger}}c_{P}^{{\dagger}}c_{P}c_{P}. \end{aligned}}$$
$$H_{\text{ring}}^{\text{SPM}}={-}\frac{\hbar^{2}\omega_{P}v_{P}^{2}{\gamma_{\textrm{ring}}^{\textrm{SPM}}}}{2\mathcal{L}}c_{P}^{{\dagger}}c_{P}^{{\dagger}}c_{P}c_{P},$$
$${\gamma_{\textrm{ring}}^{\textrm{SPM}}}=\frac{3\omega_{P}\epsilon_{0}}{4v_{P}^{2}}\frac{1}{\mathcal{L}}\int\chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})\left(\textsf{e}_{P}^{i} (\boldsymbol{r}_{{\perp}};\zeta)\textsf{e}_{P}^{j}(\boldsymbol{r}_{{\perp}};\zeta)\right)^{*}\textsf{e}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta) \textsf{e}_{P}^{l}(\boldsymbol{r}_{{\perp}};\zeta)\textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta,$$
$$\epsilon_{0}c\int\frac{n(\boldsymbol{r}_{{\perp}};\omega_{P})}{v_{g}(\boldsymbol{r}_{{\perp}};\omega_{P})}\textsf{e}_{P}^{*}(\boldsymbol{r}_{{\perp}};0) \cdot\textsf{e}_{P}(\boldsymbol{r}_{{\perp}};0)\textrm{d}\boldsymbol{r}_{{\perp}}=1,$$
$${\gamma_{\textrm{ring}}^{\textrm{SPM}}}=\frac{3\omega_{P}\overline{\chi}_{3}}{4\epsilon_{0}\overline{n}_{P}^{2}c^{2}}\frac{1}{{\textrm{A}_{\textrm{ring}}^{PPPP}}},$$
$$\frac{1}{{\textrm{A}_{\textrm{ring}}^{PPPP}}}=\frac{\mathcal{L}^{{-}1}\int\left(\chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})/\overline{\chi}_{3}\right)\left(\textsf{e}_{P}^{i}(\boldsymbol{r}_{{\perp}};\zeta) \textsf{e}_{P}^{j}(\boldsymbol{r}_{{\perp}};\zeta)\right)^{*}\textsf{e}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{e}_{P}^{l}(\boldsymbol{r}_{{\perp}};\zeta) \textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta}{\left(\int\frac{n(\boldsymbol{r}_{{\perp}};\omega_{P})/\overline{n}_{P}}{v_{g}(\boldsymbol{r}_{{\perp}};\omega_{P})/v_{P}}\textsf{e}_{P}^{*} (\boldsymbol{r}_{{\perp}};0)\cdot\textsf{e}_{P}(\boldsymbol{r}_{{\perp}};0)\textrm{d}\boldsymbol{r}_{{\perp}}\right)^{2}}.$$
$${\begin{aligned}H_{\text{ring}}^{\text{XPM}} & ={-}\frac{1}{4\epsilon_{0}}\left(\frac{\hbar\omega_{P}}{2\mathcal{L}}\right)\left(\frac{\hbar\omega_{S}}{2\mathcal{L}}\right)4!\left(\int \textrm{d}\boldsymbol{r}_{{\perp}}d\zeta\Gamma_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})\left(\textsf{d}_{P}^{i}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{d}_{S}^{j}(\boldsymbol{r}_{{\perp}};\zeta) \right)^{*}\textsf{d}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{d}_{S}^{l}(\boldsymbol{r}_{{\perp}};\zeta)\right)\\ & \quad\times c_{P}^{{\dagger}}c_{S}^{{\dagger}}c_{P}c_{S}. \end{aligned}}$$
$$H_{\text{ring}}^{\text{XPM}}={-}\frac{2{\gamma_{\textrm{ring}}^{\textrm{XPM}}}\hbar^{2}\omega_{P}v_{P}v_{S}}{\mathcal{L}}c_{P}^{{\dagger}}c_{S}^{{\dagger}}c_{P}c_{S},$$
$${\gamma_{\textrm{ring}}^{\textrm{XPM}}}=\frac{3\omega_{S}\epsilon_{0}}{4v_{P}v_{S}}\frac{1}{\mathcal{L}}\int\chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})\left(\textsf{e}_{P}^{i}(\boldsymbol{r}_{{\perp}};\zeta) \textsf{e}_{S}^{j}(\boldsymbol{r}_{{\perp}};\zeta)\right)^{*}\textsf{e}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{e}_{S}^{l}(\boldsymbol{r}_{{\perp}};\zeta)\textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta,$$
$${\gamma_{\textrm{ring}}^{\textrm{XPM}}}=\frac{3\omega_{S}\overline{\chi}_{3}}{4\epsilon_{0}\overline{n}_{P}\overline{n}_{S}c^{2}}\frac{1}{{\textrm{A}_{\textrm{ring}}^{\textrm{PSPS}}}},$$
$$\frac{1}{{\textrm{A}_{\textrm{ring}}^{PSPS}}}=\frac{\mathcal{L}^{{-}1}\int\left(\chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}})/\overline{\chi}_{3}\right)\!\left(\textsf{e}_{P}^{i}(\boldsymbol{r}_{{\perp}};\zeta) \textsf{e}_{S}^{j}(\boldsymbol{r}_{{\perp}};\zeta)\right)^{*}\textsf{e}_{P}^{k}(\boldsymbol{r}_{{\perp}};\zeta)\textsf{e}_{S}^{l}(\boldsymbol{r}_{{\perp}};\zeta) \textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta}{\left(\int\frac{n(\boldsymbol{r}_{{\perp}};\omega_{P})/\overline{n}_{P}}{v_{g}(\boldsymbol{r}_{{\perp}};\omega_{P})/v_{P}} \textsf{e}_{P}^{*}(\boldsymbol{r}_{{\perp}};0)\cdot\textsf{e}_{P}(\boldsymbol{r}_{{\perp}};0)\textrm{d}\boldsymbol{r}_{{\perp}}\right)\!\left(\int\frac{n(\boldsymbol{r}_{{\perp}}; \omega_{S})/\overline{n}_{S}}{v_{g}(\boldsymbol{r}_{{\perp}};\omega_{S})/v_{S}}\textsf{e}_{S}^{*}(\boldsymbol{r}_{{\perp}};0)\cdot\textsf{e}_{S} (\boldsymbol{r}_{{\perp}};0)\textrm{d}\boldsymbol{r}_{{\perp}}\right)}.$$
$${\gamma_{\textrm{ring}}^{\textrm{SPM}}}=\frac{3\omega_{J}}{4v_{J}^{2}\mathcal{L}}K,$$
$$H_{\text{ring}}^{\text{SPM}}={-}\frac{3}{2}\left(\frac{\hbar\omega_{J}}{2\mathcal{L}}\right)^{2}Kc_{J}^{{\dagger}}c_{J}^{{\dagger}}c_{J}c_{J},$$
$$H = i \frac{\hbar g}{2}(e^{i \phi} a^{{\dagger} 2} - e^{{-}i \phi} a^2 ),$$
$$i \hbar \frac{\textrm{d}}{\textrm{d}t} \mathcal{U}(t,t_{\text{i}}) = H(t) \mathcal{U}(t,t_{\text{i}}),$$
$$\mathcal{U}(t_{\text{f}},t_{\text{i}}) = \exp\left( - i \frac{H}{\hbar} (t_{\text{f}}-t_{\text{i}}) \right) = \exp\left(\tfrac{r}{2}\left[e^{i \phi} a^{{\dagger} 2} - e^{{-}i \phi} a^2 \right] \right), r = g (t_{\text{f}}-t_{\text{i}}).$$
$$|{\Psi(t_{\text{f}})} = \exp\left(\tfrac{r}{2}\left[ e^{i \phi} a^{{\dagger} 2} - e^{{-}i \phi} a^2 \right] \right)|0 \rangle = \frac{1}{\sqrt{\cosh r}}\sum_{n=0}^\infty e^{i n\phi} \tanh^n r \frac{\sqrt{2n!}}{2^n n!} |2n \rangle,$$
$$|{n} = \frac{a^{{\dagger} n} }{ \sqrt{n!}} |{0},$$
$$\begin{aligned}\exp\left(\tfrac{r}{2}\left[ e^{i \phi} a^{{\dagger} 2} - e^{{-}i \phi} a^2 \right] \right) & = \exp\left( \tfrac{1}{2} e^{i \phi } \tanh r \ a^{{\dagger} 2}\right) \\ & \quad\times \exp\left( - \left[ a^\dagger a+\tfrac{1}{2} \right] \ln \cosh r \right)\\ & \quad\times \exp\left( -\tfrac{1}{2} e^{{-}i \phi } \tanh r \ a^{ 2}\right). \end{aligned}$$
$$\begin{aligned} \langle{a}_{t_{\text{f}}}\rangle & = \langle{\Psi(t_{\text{f}})\rangle |a|\Psi(t_{\text{f}})} = 0, \\ n_{t_{\text{f}}} = \langle{a^\dagger a}\rangle_{t_{\text{f}}} & = \langle{\Psi(t_{\text{f}})|a^\dagger a|\Psi(t_{\text{f}})}\rangle = \sinh^2 r,\\ m_{t_{\text{f}}} =\langle{a^2}\rangle_{t_{\text{f}}} & = \langle{\Psi(t_{\text{f}})|a^2|\Psi(t_{\text{f}})}\rangle = \tfrac{1}{2}e^{i \phi} \sinh 2r,\\ \langle{a^\dagger a a^\dagger a}\rangle_{t_\text{f}} & = 3 n_{t_\text{f}}^2+ 2n_{t_\text{f}}. \end{aligned}$$
$$q = \sqrt{\frac{\hbar}{2}} \left(a + a^\dagger \right), \quad p ={-}i \sqrt{\frac{\hbar}{2}} \left(a - a^\dagger \right)$$
$$\begin{aligned}\langle{q}\rangle_{t_{\text{f}}} & = \langle{p}\rangle_{t_{\text{f}}} = 0, \\ \langle{q^2}\rangle_{t_{\text{f}}} & = \tfrac{\hbar}{2}\left(2n_{t_{\text{f}}}+1+ m_{t_{\text{f}}}+m_{t_{\text{f}}}^* \right) = \frac{\hbar}{2} \left(\sinh (2 r) \cos (\phi )+2 \sinh ^2(r)+1 \right),\\ \langle{p^2}\rangle_{t_{\text{f}}} & = \tfrac{\hbar}{2}\left(2n_{t_{\text{f}}}+1- m_{t_{\text{f}}}-m_{t_{\text{f}}}^* \right)=\frac{\hbar}{2} \left(-\sinh (2 r) \cos (\phi )+2 \sinh ^2(r)+1 \right),\\ \frac{\langle{qp+pq}\rangle_{t_{\text{f}}}}{2} & = i \tfrac{\hbar}{2} (m^* - m) =\frac{\hbar}{2} \sin(\phi) \sinh(2r). \end{aligned}$$
$${\begin{aligned}V_{t_{\text{f}}} =\begin{pmatrix} \langle{q}\rangle_{t_{\text{f}}}^2 & \frac{\langle{qp+pq}\rangle_{t_{\text{f}}}}{2} \\ \frac{\langle{qp+pq}\rangle_{t_{\text{f}}}}{2} & \langle{p^2}\rangle_{t_{\text{f}}} \end{pmatrix} = \frac{\hbar}{2} R(\phi/2) \begin{pmatrix} e^{2r} & 0 \\ 0 & e^{{-}2r} \end{pmatrix} R(\phi/2)^T, \quad R(\phi) = \begin{pmatrix} \cos \phi & - \sin \phi \\ \sin \phi & \cos \phi \end{pmatrix}. \end{aligned}}$$
$$\frac{\textrm{d}}{\textrm{d}t} O(t) = \frac{i}{\hbar} [H(t),O(t)],$$
$$\frac{\textrm{d}}{\textrm{d}t} a = \frac{i}{\hbar} [H,O] = g e^{i \phi} a^\dagger, \frac{\textrm{d}}{\textrm{d}t} a^\dagger{=} \frac{i}{\hbar} [H,O] = g e^{{-}i \phi} a,$$
$$\begin{aligned}\frac{\textrm{d}}{\textrm{d}t} \begin{pmatrix} a \\ a^\dagger \end{pmatrix} & ={-}i \underbrace{\begin{pmatrix} 0 & i g e^{i \phi}\\ i g e^{{-}i \phi} & 0 \end{pmatrix} }_{{\equiv} \boldsymbol{A}}\begin{pmatrix} a \\ a^\dagger \end{pmatrix}. \end{aligned}$$
$${\begin{aligned}\begin{pmatrix} a(t_{\text{f}}) \\ a^\dagger(t_{\text{f}}) \end{pmatrix} = \exp({-}i \boldsymbol{A}(t_{\text{f}}-t_{\text{i}})) \begin{pmatrix} a(t_{\text{i}}) \\ a^\dagger(t_{\text{i}}) \end{pmatrix} = \left( \begin{array}{cc} \cosh g (t_{\text{f}} - t_{\text{i}}) & e^{i \phi } \sinh g (t_{\text{f}} - t_{\text{i}}) \\ e^{{-}i \phi } \sinh g(t_{\text{f}} - t_{\text{i}}) & \cosh g (t_{\text{f}} - t_{\text{i}}) \\ \end{array} \right) \begin{pmatrix} a(t_{\text{i}}) \\ a^\dagger(t_{\text{i}}) \end{pmatrix}, \end{aligned}}$$
$$H(t) = \hbar \left\{ \sum_{i,j =1}^\ell \Delta_{ij}(t) a_i^\dagger a_j + i \frac12 \sum_{k,l=1}^\ell \left[ \zeta_{kl}(t) a_k^{{\dagger} } a_{l}^{{\dagger}} -\text{H.c.} \right] \right\},$$
$$[a_i,a_j] = [a_i^\dagger,a_j^\dagger] = 0, \quad [a_i,a_j^\dagger] = \delta_{ij},$$
$$\mathcal{U}(t,t_{\text{i}}) = \mathcal{T} \exp\left(-\frac{i}{\hbar} \int_{t_{\text{i}}}^t \textrm{d}t' H(t') \right),$$
$$\mathcal{T}\left[H(t_1) H(t_2) \ldots H(t_n) \right] = H(t_{i_1}) H(t_{i_2}) \ldots H(t_{i_n}),$$
$$\mathcal{U}(t_{\text{f}},t_{\text{i}}) = \mathbb{I}+({-}i)\int_{t_{\text{i}}}^{t_{\text{f}}} \textrm{d}t' \frac{ H(t')}{\hbar}+({-}i)^2\int_{t_{\text{i}}}^{t_{\text{f}}} \textrm{d}t' \int_{t_{\text{i}}}^{t'} \textrm{d}t^{\prime\prime} \frac{H(t') H(t^{\prime\prime})}{\hbar^2}+\cdots ,$$
$$\begin{aligned}\mathcal{U}(t_{\text{f}},t_{\text{i}}) & =\exp\big( \Omega_1(t_{\text{f}},t_{\text{i}})+ \Omega_2(t_{\text{f}},t_{\text{i}})+ \Omega_3(t_{\text{f}},t_{\text{i}})+\cdots \big), \end{aligned}$$
$$\begin{aligned}\Omega_1(t_{\text{f}},t_{\text{i}}) & ={-}\frac{i}{\hbar} \int_{t_{\text{i}}}^{t_{\text{f}}} \textrm{d}t H(t), \end{aligned}$$
$$\begin{aligned} \Omega_2(t_{\text{f}},t_{\text{i}}) & =\frac{({-}i)^2}{2 \hbar^2}\int_{t_{\text{i}}}^{t_{\text{f}}} \textrm{d}t \int_{t_{\text{i}}} ^{t} \textrm{d}t' \left[ H(t), H(t') \right], \end{aligned}$$
$$\begin{aligned} \Omega_3(t_{\text{f}},t_{\text{i}}) & =\frac{({-}i)^3}{6 \hbar^3}\int_{t_{\text{i}}}^{t_{\text{f}}} \textrm{d}t \int_{t_{\text{i}}} ^{t} \textrm{d}t' \int_{t_{\text{i}}} ^{t'} \textrm{d}t^{\prime\prime} \Big( \left[ H(t),\left[ H(t'), H(t^{\prime\prime}) \right] \right]\\ & \quad +\left[\left[ H(t), H(t') \right], H(t^{\prime\prime}) \right] \Big). \end{aligned}$$
$$\mathcal{U}(t_{\text{f}},t_{\text{i}}) = \exp\left({-}i \frac{H(t_{N})}{\hbar} \Delta t \right) \exp\left({-}i \frac{H(t_{N-1})}{\hbar} \Delta t \right) \ldots \exp\left({-}i \frac{H(t_{\text{i}})}{\hbar} \Delta t \right),$$
$$\begin{aligned} |{\overline{\Psi}(t_\text{f})}\rangle & \approx \overline{\mathcal{U}}(t_\text{f}, t_\text{i})\, |{\text{vac}}\rangle, \end{aligned}$$
$$\begin{aligned}\overline{\mathcal{U}}(t_\text{f}, t_\text{i}) & = \exp\left(-\frac{i}{\hbar} \int_{t_{\text{i}}}^{t_\text{f}} \textrm{d}t' H(t') \right) = \exp\left[ \tfrac{1}{2}\sum_{k,l} \overline{\mathit{J}}_{kl} a_l^\dagger a_k^\dagger{-} \text{H.c.} \right], \end{aligned}$$
$$\overline{A}_\lambda^\dagger{=} \sum_{k=1}^\ell a_k^\dagger \overline{F}_{k\lambda} \Longleftrightarrow a_k^\dagger{=} \sum_{i=1}^\ell \overline{F}^*_{k\lambda} \overline{A}_{\lambda}^\dagger,$$
$$[\overline{A}_{\lambda}, \overline{A}_{\lambda'}] = 0, \quad [\overline{A}_{\lambda}, \overline{A}_{\lambda'}^\dagger] = \delta_{\lambda,\lambda'}$$
$$\begin{aligned}|{\overline{\Psi}(t_\text{f})}\rangle & = \exp\left[\tfrac12 \sum_{k,l=1}^\ell \overline{J}_{kl} a^\dagger_{k} a^\dagger_l - \text{H.c.}\right] |{\text{vac}}\rangle = \exp \left[ \tfrac12 \sum_{\lambda=1}^\ell \overline{\mathit{r}}_\lambda \overline{A}_\lambda^{{\dagger} 2} -\text{H.c.} \right] |{\text{vac}}\rangle \\ & = \bigotimes_{\lambda=1}^\ell \exp \left[ \tfrac12 \overline{\mathit{r}}_\lambda \overline{A}_\lambda^{{\dagger} 2} -\text{H.c.} \right] |{\text{vac}}\rangle. \end{aligned}$$
$$|{\overline{\Psi}(t_\text{f})}\rangle = \bigotimes_{\lambda=1}^\ell \left[ \frac{1}{\sqrt{\cosh \overline{\mathit{r}}_\lambda}}\sum_{n=0}^\infty \tanh^n \overline{\mathit{r}}_\lambda \frac{\sqrt{(2n)!}}{2^n n!} |2n \rangle_\lambda \right],$$
$$\begin{aligned}a_i(t_{\text{f}}) & = \overline{\mathcal{U}}(t_\text{f}, t_\text{i})^\dagger a_i( t_\text{i}) \overline{\mathcal{U}}(t_\text{f}, t_\text{i}) = \sum_{j=1}^\ell \overline{V}_{ij} a_{j}(t_{\text{i}}) + \sum_{j=1}^\ell \overline{W}_{ij} a^\dagger _{j} (t_{\text{i}}), \end{aligned}$$
$$\begin{aligned}a_i^\dagger(t_{\text{f}}) & = \overline{\mathcal{U}}(t_\text{f}, t_\text{i})^\dagger a_i^\dagger( t_\text{i}) \overline{\mathcal{U}}(t_\text{f}, t_\text{i}) = \sum_{j=1}^\ell \overline{V}^*_{ij} a^\dagger_{j}(t_{\text{i}}) + \sum_{j=1}^\ell \overline{W}^*_{ij} a _{j}(t_{\text{i}}), \end{aligned}$$
$$\overline{\boldsymbol{V}} = \overline{\boldsymbol{F}} \left[ \oplus_{\lambda=1}^\ell \cosh \overline{\mathit{r}}_{\lambda} \right] \overline{\boldsymbol{F}}^\dagger, \quad \overline{\boldsymbol{W}} = \overline{\boldsymbol{F}} \left[ \oplus_{\lambda=1}^\ell \sinh \overline{\mathit{r}}_{\lambda} \right] \overline{\boldsymbol{F}}^T.$$
$$\begin{aligned} \frac{\textrm{d}}{\textrm{d}t} \begin{pmatrix} \boldsymbol{a} \\ \boldsymbol{a}^\dagger \end{pmatrix} & ={-}i \underbrace{\begin{pmatrix} - \boldsymbol{\Delta}(t) & i \boldsymbol{\zeta}(t) \\ i \boldsymbol{\zeta}^*(t) & \boldsymbol{\Delta}(t) \end{pmatrix} }_{{\equiv} \boldsymbol{A}(t)}\begin{pmatrix} \boldsymbol{a} \\ \boldsymbol{a}^\dagger \end{pmatrix}, \end{aligned}$$
$$\begin{aligned}\boldsymbol{a} = \begin{pmatrix} a_1 \\ \vdots \\ a_\ell \end{pmatrix}, \quad \boldsymbol{a}^\dagger{=} \begin{pmatrix} a^\dagger_1 \\ \vdots \\ a^\dagger_\ell \end{pmatrix}. \end{aligned}$$
$${\begin{aligned} \begin{pmatrix} \boldsymbol{a}(t_{\text{f}}) \\ \boldsymbol{a}^\dagger(t_{\text{f}}) \end{pmatrix} = \mathcal{U}^\dagger(t_{\text{f}},t_{\text{i}}) \begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix} \mathcal{U}(t_{\text{f}},t_{\text{i}})= \mathcal{T} \exp\left( - i \int_{t_{\text{i}}}^{_{\text{f}}} \textrm{d}t'\boldsymbol{A}(t') \right) \begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix} =\boldsymbol{K}(t_{\text{f}},t_{\text{i}})\begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix}. \end{aligned}}$$
$$\boldsymbol{K}(t_{\text{f}},t_{\text{i}}) \approx \exp\left({-}i \Delta t \boldsymbol{A}(t_{N}) \right) \exp\left({-}i \Delta t \boldsymbol{A}(t_{N-1}) \right) \ldots \exp\left({-}i \Delta t \boldsymbol{A}(t_2) \right) \exp\left({-}i \Delta t \boldsymbol{A}(t_{\text{i}}) \right),$$
$$\begin{aligned}\boldsymbol{K} = \begin{pmatrix} \boldsymbol{V} & \boldsymbol{W} \\ \boldsymbol{W}^* & \boldsymbol{V}^* \end{pmatrix}, \end{aligned}$$
$$\begin{aligned} a_i(t_{\text{f}}) & = \mathcal{U}^\dagger a_i \mathcal{U} = \sum_{j=1}^\ell V_{ij} a_{j}(t_{\text{i}}) + \sum_{j=1}^\ell W_{ij} a^\dagger _{j} (t_{\text{i}}), \end{aligned}$$
$$\begin{aligned} a_i^\dagger(t_{\text{f}}) & = \mathcal{U}^\dagger a_i^\dagger \mathcal{U} = \sum_{j=1}^\ell V^*_{ij} a^\dagger_{j}(t_{\text{i}}) + \sum_{j=1}^\ell W^*_{ij} a _{j}(t_{\text{i}}). \end{aligned}$$
$$ [a_i(t_{\text{f}}), a_j(t_{\text{f}})] = 0 \Leftrightarrow \boldsymbol{W} \boldsymbol{V}^T = \boldsymbol{V} \boldsymbol{W}^T, $$
$$ [a_i(t_{\text{f}}), a^\dagger_j(t_{\text{f}})] = \delta_{ij} \Leftrightarrow \boldsymbol{V}^* \boldsymbol{V}^T - \boldsymbol{W}^* \boldsymbol{W}^T = \mathbb{I}_M. $$
$$\boldsymbol{V} = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \cosh r_{\lambda} \right] \boldsymbol{G}, \quad \boldsymbol{W} = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \sinh r_{\lambda} \right] \boldsymbol{G}^*,$$
$$\boldsymbol{M} = \boldsymbol{M}^T = \boldsymbol{W} \boldsymbol{V}^T$$
$$M_{ij} = \langle{a_i(t_{\text{f}})a_j(t_{\text{f}})}\rangle_{|{\text{vac}}\rangle} = \rangle{\text{vac}}|\mathcal{U}^\dagger a_i a_j \mathcal{U} |{\text{vac}}\rangle = \left( \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \tfrac{1}{2}\sinh 2r_\lambda \right] \boldsymbol{F}^T \right)_{ij}.$$
$$\boldsymbol{N} = \boldsymbol{N}^\dagger{=} \boldsymbol{W}^* \boldsymbol{W}^T$$
$$N_{ij} = \langle{a_i^\dagger(t_{\text{f}})a_j(t_{\text{f}})}\rangle_{|{\text{vac}}\rangle} = \rangle{\text{vac}}|\mathcal{U}^\dagger a_i^\dagger a_j \mathcal{U} |{\text{vac}}\rangle = \left( \boldsymbol{F}^* \left[ \oplus_{\lambda=1}^\ell \sinh^2r_\lambda \right] \boldsymbol{F}^T \right)_{ij}.$$
$$\begin{aligned} \boldsymbol{K} = \begin{pmatrix} \boldsymbol{F} & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{F}^* \end{pmatrix} \begin{pmatrix} \oplus_{\lambda=1}^\ell \cosh r_{\lambda} & \oplus_{\lambda=1}^\ell \sinh r_{\lambda} \\ \oplus_{\lambda=1}^\ell \sinh r_{\lambda} & \oplus_{\lambda=1}^\ell \cosh r_{\lambda} \end{pmatrix} \begin{pmatrix} \boldsymbol{G} & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{G}^* \end{pmatrix}, \end{aligned}$$
$$\begin{aligned} \boldsymbol{K}^{{-}1} = \begin{pmatrix} \boldsymbol{G}^\dagger & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{G}^T \end{pmatrix} \begin{pmatrix} \oplus_{\lambda=1}^\ell \cosh r_{\lambda} & -\oplus_{\lambda=1}^\ell \sinh r_{\lambda} \\ -\oplus_{\lambda=1}^\ell \sinh r_{\lambda} & \oplus_{\lambda=1}^\ell \cosh r_{\lambda} \end{pmatrix} \begin{pmatrix} \boldsymbol{F}^\dagger & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{F}^T \end{pmatrix}= \begin{pmatrix} \tilde{\boldsymbol{V}} & \tilde{\boldsymbol{W}} \\ \tilde{\boldsymbol{W}}^* & \tilde{\boldsymbol{V}}^* \end{pmatrix}, \end{aligned}$$
$$\tilde{\boldsymbol{V}} = \boldsymbol{G}^\dagger \left[ \oplus_{\lambda=1}^\ell \cosh r_\lambda \right] \boldsymbol{F}^\dagger{=} \boldsymbol{V}^\dagger, \quad \tilde{\boldsymbol{W}} ={-}\boldsymbol{G}^\dagger \left[ \oplus_{\lambda=1}^\ell \sinh r_\lambda \right] \boldsymbol{F}^T ={-} \boldsymbol{W}^T,$$
$$\begin{aligned} \mathcal{U}(t_{\text{f}},t_{\text{i}}) \begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix} \mathcal{U}^\dagger(t_{\text{f}},t_{\text{i}}) = \boldsymbol{K}^{{-}1} \begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix} = \begin{pmatrix} \tilde{\boldsymbol{V}} & \tilde{\boldsymbol{W}} \\ \tilde{\boldsymbol{W}}^* & \tilde{\boldsymbol{V}}^* \end{pmatrix} \begin{pmatrix} \boldsymbol{a}(t_{\text{i}}) \\ \boldsymbol{a}^\dagger(t_{\text{i}}) \end{pmatrix}. \end{aligned}$$
$$\chi_{\rho}(\boldsymbol{\alpha}) = \text{tr}\left( \mathcal{D}(\boldsymbol{\alpha}) \rho \right),$$
$$\begin{aligned} \mathcal{D}(\boldsymbol{\alpha}) & = \exp\left[ \boldsymbol{\alpha}^T \boldsymbol{a}^\dagger{-} \boldsymbol{\alpha}^\dagger \boldsymbol{a} \right] = \exp\left[ \boldsymbol{\alpha}^T \boldsymbol{a}^\dagger \right] \exp\left[ - \boldsymbol{\alpha}^\dagger \boldsymbol{a} \right] \exp[ -\tfrac12 \|\boldsymbol{\alpha} \|^2], \end{aligned}$$
$$\begin{aligned} \boldsymbol{\alpha} & = (\alpha_1,\ldots,\alpha_M)^T \in \mathbb{C}^\ell, \end{aligned}$$
$$\chi_{|{\text{vac}}\rangle}(\boldsymbol{\alpha}) = \exp[ -\tfrac12 \|\boldsymbol{\alpha} \|^2],$$
$$\begin{aligned} \chi_{\mathcal{U}|{\text{vac}}\rangle}(\boldsymbol{\alpha}) & = \text{tr}\left( \mathcal{D}(\boldsymbol{\alpha}) \mathcal{U}|{\text{vac}}\rangle \rangle{\text{vac}}| \mathcal{U}^\dagger \right) \\ & = \text{tr}\left(\mathcal{U}^\dagger \mathcal{D}(\boldsymbol{\alpha}) \mathcal{U}|{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \text{tr}\left(\mathcal{U}^\dagger \exp\left[ \boldsymbol{\alpha}^T \boldsymbol{a}^\dagger{-} \boldsymbol{\alpha}^\dagger \boldsymbol{a} \right]\mathcal{U}|{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \text{tr}\left( \exp\left[ \boldsymbol{\alpha}^T \mathcal{U}^\dagger \boldsymbol{a}^\dagger \mathcal{U} - \boldsymbol{\alpha}^\dagger \mathcal{U}^\dagger \boldsymbol{a} \mathcal{U}\right]|{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \text{tr}\left( \exp\left[ \boldsymbol{\alpha}^T \left\{ \boldsymbol{V}^*\boldsymbol{a}^\dagger{+} \boldsymbol{W}^* \boldsymbol{a} \right\} - \boldsymbol{\alpha}^\dagger \left\{ \boldsymbol{V}\boldsymbol{a} + \boldsymbol{W} \boldsymbol{a}^\dagger \right\}\right]|{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \text{tr}\left( \exp\left[ \left\{ \boldsymbol{\alpha}^T \boldsymbol{V}^* - \boldsymbol{\alpha}^\dagger \boldsymbol{W} \right\} \boldsymbol{a}^\dagger{-} \left\{ \boldsymbol{\alpha}^\dagger\boldsymbol{V} - \boldsymbol{\alpha}^T \boldsymbol{W}^* \right\} \boldsymbol{a} \right]|{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \text{tr}\left( \mathcal{D}( \boldsymbol{V}^\dagger \boldsymbol{\alpha} - \boldsymbol{W}^T \boldsymbol{\alpha}^* ) |{\text{vac}}\rangle \rangle{\text{vac}}| \right)\\ & = \exp[ -\tfrac12 \| \boldsymbol{\xi} \|^2], \quad \boldsymbol{\xi} = \boldsymbol{V}^\dagger \boldsymbol{\alpha} - \boldsymbol{W}^T \boldsymbol{\alpha}^*. \end{aligned}$$
$$\| \boldsymbol{\xi}\|^2 = \boldsymbol{\xi}^\dagger \boldsymbol{\xi} = \boldsymbol{\alpha}^\dagger \boldsymbol{V} \boldsymbol{V}^\dagger \boldsymbol{\alpha} - \boldsymbol{\alpha}^\dagger \boldsymbol{V} \boldsymbol{W}^T \boldsymbol{\alpha}^* - \boldsymbol{\alpha}^T \boldsymbol{W}^* \boldsymbol{V}^\dagger \boldsymbol{\alpha} + \boldsymbol{\alpha}^T \boldsymbol{W}^* \boldsymbol{W}^T \boldsymbol{\alpha}^*,$$
$$ \boldsymbol{V} \boldsymbol{V}^\dagger = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \cosh^2 r_{\lambda} \right] \boldsymbol{F}^\dagger{=} \mathbb{I}_{M}+ \boldsymbol{N}^*, $$
$$ \boldsymbol{V} \boldsymbol{W}^T = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \tfrac12 \sinh 2r_{\lambda} \right] \boldsymbol{F}^T = \left( \boldsymbol{W}^* \boldsymbol{V}^\dagger \right)^\dagger{=} \boldsymbol{M}, $$
$$ \boldsymbol{W}^* \boldsymbol{W}^T = \boldsymbol{F}^* \left[ \oplus_{\lambda=1}^\ell \sinh^2 r_{\lambda} \right] \boldsymbol{F}^T = \boldsymbol{N}, $$
$$\chi_{\mathcal{U}|{\text{vac}}\rangle}(\boldsymbol{\alpha}) = \exp\left[- \tfrac12 \left\{ \|\boldsymbol{\alpha}\|^2 + \boldsymbol{\alpha}^\dagger \boldsymbol{N}^* \boldsymbol{\alpha} - \boldsymbol{\alpha}^\dagger \boldsymbol{M} \boldsymbol{\alpha}^* - \boldsymbol{\alpha}^T \boldsymbol{M}^* \boldsymbol{\alpha} + \boldsymbol{\alpha}^T \boldsymbol{N} \boldsymbol{\alpha}^* \right\}\right].$$
$$ |{\Psi(t_{\text{f}})}\rangle \equiv \mathcal{U}|{\text{vac}}\rangle = \exp\left[ \frac12 \sum_{k,l=1}^\ell J_{kl} a^\dagger_{k} a^\dagger_l - \text{H.c.}\right]|{\text{vac}}\rangle, $$
$$ \boldsymbol{J} = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell r_{\lambda} \right] \boldsymbol{F}^T = \boldsymbol{J}^T, $$
$$[b_i, b_j] = [c_i,c_j] = [b_i,c_j] = [b_i,c_j^\dagger] = 0, \quad [b_i,b_j^\dagger] = [c_i,c_j^\dagger] = \delta_{ij}.$$
$$H(t) = \hbar \left\{\sum_{i,j=1}^{\ell} \Delta^{b}_{ij}(t) b_i^\dagger b_j + \sum_{i,j=1}^{\ell} \Delta^{c}_{ij}(t) c_i^\dagger c_j + i \left[ \sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_c} \zeta^{b,c}_{ij}(t) b_i^\dagger c_j^\dagger{-} \text{H.c.} \right] \right\}.$$
$${\begin{aligned}\boldsymbol{a} = \begin{pmatrix} \boldsymbol{b} \\ \boldsymbol{c} \end{pmatrix}, \;\; \boldsymbol{b} = \begin{pmatrix} b_1 \\ \vdots \\ b_{\ell_b} \end{pmatrix}, \quad \boldsymbol{c} = \begin{pmatrix} c_1 \\ \vdots \\ c_{\ell_c} \end{pmatrix}, \;\; \boldsymbol{\Delta}(t) = \begin{pmatrix} \boldsymbol{\Delta}_b(t) & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{\Delta}_c(t) \end{pmatrix}, \;\; \boldsymbol{\zeta}(t) = \begin{pmatrix} \boldsymbol{0} & \boldsymbol{\zeta}^{b,c}(t)\\ [\boldsymbol{\zeta}^{b,c}(t)]^T & \boldsymbol{0} \end{pmatrix}. \end{aligned}}$$
$$ |{\overline{\Psi}(t_\text{f})}\rangle \approx \overline{\mathcal{U}}(t_\text{f}, t_\text{i}) |{\text{vac}}\rangle $$
$$ \overline{\mathcal{U}}(t_\text{f}, t_\text{i}) = \exp\left(-\frac{i}{\hbar} \int_{t_{\text{i}}}^{t_\text{f}} \textrm{d}t' H(t') \right) = \exp\left[ \sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_c} \overline{J}^{b,c}_{ij} b_i^\dagger c_j^\dagger{-} \text{H.c.}\right], $$
$$\overline{\boldsymbol{J}}^{b,c} = \overline{\boldsymbol{F}}^{b} \ \overline{\boldsymbol{D}} \ [\overline{\boldsymbol{F}}^{c} ]^T,$$
$$ \overline{B}_\lambda^\dagger{=} \sum_{k=1}^{\ell_b} b_k^\dagger \overline{F}^b_{k,\lambda} \Longleftrightarrow b_k^\dagger{=} \sum_{i=1}^{\ell_b} [\overline{F}^b]^*_{k,\lambda} \overline{B}_{\lambda}^\dagger, $$
$$ \overline{C}_\lambda^\dagger{=} \sum_{k=1}^{\ell_c} c_k^\dagger \overline{F}^c_{k,\lambda} \Longleftrightarrow c_k^\dagger{=} \sum_{i=1}^{\ell_c} [\overline{F}^c]^*_{k,\lambda} \overline{C}_{\lambda}^\dagger. $$
$$\begin{aligned} |{\overline{\Psi}(t_{\text{f}})}\rangle& = \exp\left[ \sum_{\lambda=1}^{\ell_{\rm min}} \overline{r}_{\lambda} \overline{B}_{\lambda}^\dagger \overline{C}_{\lambda}^\dagger{-} \text{H.c.}\right] |{\text{vac}}\rangle = \bigotimes_{\lambda=1}^{\ell_{\rm min}} \exp\left[ \overline{r}_{\lambda} \overline{B}_{\lambda}^\dagger \overline{C}_{\lambda}^\dagger{-} \text{H.c.}\right] |{\text{vac}}\rangle\\& = \bigotimes_{\lambda=1}^{\ell_{\rm min}} \left[ \frac{1}{\cosh \overline{r}_\lambda} \sum_{n=0}^\infty \tanh^n \overline{r}_\lambda |{n,n}\rangle_{\lambda} \right],\end{aligned}$$
$$\begin{aligned} \exp(r e^{i \phi} b^\dagger c^\dagger{-}\text{H.c.}) = & \exp\left( e^{i \phi} \tanh r \ c^{{\dagger} } b^\dagger\right) \\ & \times \exp\left( - \left[ c^\dagger c+ b^\dagger b +1\right] \ln \cosh r \right)\\ & \times \exp\left( - e^{{-}i \phi} \tanh r \ c b\right), \end{aligned}$$
$$|{n,n}\rangle_{\lambda} = \frac{\overline{B}_\lambda^{{\dagger} n} \overline{C}_\lambda^{{\dagger} n}}{n!} |{\text{vac}}\rangle.$$
$$ b_i(t_{\text{f}}) = \overline{\mathcal{U}}^\dagger b_i(t_{\text{i}}) \overline{\mathcal{U}} = \sum_{j=1}^\ell \overline{V}^{b,b}_{ij} b_{j}(t_{\text{i}}) + \sum_{j=1}^\ell \overline{W}^{b,c}_{ij} c^\dagger _{j} (t_{\text{i}}), $$
$$ c_i^\dagger(t_{\text{f}}) = \overline{\mathcal{U}}^\dagger c_i^\dagger(t_{\text{i}}) \overline{\mathcal{U}} = \sum_{j=1}^\ell \left[\overline{V}^{(c,c)}_{ij}\right]^* c^\dagger_{j}(t_{\text{i}}) + \sum_{j=1}^\ell \left[\overline{W}^{c,b}_{ij}\right]^* b _{j}(t_{\text{i}}), $$
$$ \overline{\boldsymbol{V}}^{b,b} = \overline{\boldsymbol{F}}^{b} \left[ \oplus_{\lambda=1}^\ell \cosh \overline{r}_{\lambda} \right] \left[\overline{\boldsymbol{F}}^{b} \right]^\dagger, \quad \overline{\boldsymbol{V}}^{(c,c)} = \overline{\boldsymbol{F}}^{c} \left[ \oplus_{\lambda=1}^\ell \cosh \overline{r}_{\lambda} \right] \left[\overline{\boldsymbol{F}}^{c} \right]^\dagger, $$
$$ \overline{\boldsymbol{W}}^{b,c} = \overline{\boldsymbol{F}}^{b} \left[ \oplus_{\lambda=1}^\ell \sinh \overline{r}_{\lambda} \right] \left[\overline{\boldsymbol{F}}^{c} \right]^T, \quad \overline{\boldsymbol{W}}^{c,b} = \overline{\boldsymbol{F}}^{c} \left[ \oplus_{\lambda=1}^\ell \sinh \overline{r}_{\lambda} \right] \left[\overline{\boldsymbol{F}}^{b} \right]^T = \left[ \overline{\boldsymbol{W}}^{b,c} \right]^T. $$
$$\overline{\boldsymbol{V}}^{b,b} \approx \mathbb{I}_{\ell_b}, \overline{\boldsymbol{V}}^{c,c} \approx \mathbb{I}_{\ell_c} \quad \text{and} \quad \overline{\boldsymbol{W}}^{b,c} = \left[\overline{\boldsymbol{W}}^{c,b} \right]^T \approx \overline{\boldsymbol{J}}^{b,c}.$$
$$\begin{aligned}\frac{\textrm{d}}{\textrm{d}t} \begin{pmatrix} \boldsymbol{b} \\ \boldsymbol{c}^\dagger \end{pmatrix} ={-}i \underbrace{\begin{pmatrix} - \boldsymbol{\Delta}_b(t) & i \boldsymbol{\zeta}^{b,c}(t) \\ i [\boldsymbol{\zeta}^{b,c}(t)]^\dagger & \boldsymbol{\Delta}_c(t) \end{pmatrix} }_{{\equiv} \boldsymbol{A}^{b,c}(t)} \begin{pmatrix} \boldsymbol{b} \\ \boldsymbol{c}^\dagger \end{pmatrix}. \end{aligned}$$
$$\begin{aligned}\begin{pmatrix} \boldsymbol{b}(t_{\text{f}}) \\ \boldsymbol{c}^\dagger(t_{\text{f}}) \end{pmatrix} = & \mathcal{T}\exp\left[ - i \int_{t_{\text{i}}}^{t_\text{f}} \textrm{d}t \boldsymbol{A}^{b,c}(t) \right] \begin{pmatrix} \boldsymbol{b}(t_{\text{i}}) \\ \boldsymbol{c}^\dagger(t_{\text{i}}) \end{pmatrix} = \begin{pmatrix} \boldsymbol{V}^{b,b} & \boldsymbol{W}^{b,c} \\ (\boldsymbol{W}^{c,b})^* & (\boldsymbol{V}^{c,c})^* \end{pmatrix} \begin{pmatrix} \boldsymbol{b}(t_{\text{i}}) \\ \boldsymbol{c}^\dagger(t_{\text{i}}) \end{pmatrix} \\ = & \boldsymbol{K}^{b,c}(t_{\text{f}},t_{\text{i}}) \begin{pmatrix} \boldsymbol{b}(t_{\text{i}}) \\ \boldsymbol{c}^\dagger(t_{\text{i}}) \end{pmatrix}. \end{aligned}$$
$$\begin{aligned}N^b_{ij} = \rangle{\text{vac}}| b_i^\dagger(t_{\text{f}}) b_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger b_i^\dagger b_j \mathcal{U} |{\text{vac}}\rangle, \end{aligned}$$
$$\begin{aligned}N^c_{ij} = \rangle{\text{vac}}| c_i^\dagger(t_{\text{f}}) c_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger c_i^\dagger c_j \mathcal{U} |{\text{vac}}\rangle, \end{aligned}$$
$$\begin{aligned}M^{b,c}_{ij} = \rangle{\text{vac}}| b_i(t_{\text{f}}) c_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger b_i c_j \mathcal{U} |{\text{vac}}\rangle, \end{aligned}$$
$$\begin{aligned}N^{b,c}_{ij} = \rangle{\text{vac}}| b_i^\dagger(t_{\text{f}}) c_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger b_i^\dagger c_j \mathcal{U} |{\text{vac}}\rangle = 0, \end{aligned}$$
$$\begin{aligned}M^{b}_{ij} = \rangle{\text{vac}}| b_i(t_{\text{f}}) b_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger b_i b_j \mathcal{U} |{\text{vac}}\rangle =0, \end{aligned}$$
$$\begin{aligned}M^{c}_{ij} = \rangle{\text{vac}}| c_i(t_{\text{f}}) c_j(t_{\text{f}}) |{\text{vac}}\rangle = \rangle{\text{vac}}| \mathcal{U}^\dagger c_i c_j \mathcal{U} |{\text{vac}}\rangle =0. \end{aligned}$$
$$\begin{aligned} \boldsymbol{N} = \begin{pmatrix} \boldsymbol{N}^b & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{N}^c \end{pmatrix}, \quad \boldsymbol{M} = \begin{pmatrix} \boldsymbol{0} & \boldsymbol{M}^{b,c} \\ [\boldsymbol{M}^{b,c}]^T & \boldsymbol{0} \end{pmatrix}. \end{aligned}$$
$$\boldsymbol{M}^{b,c} = {\boldsymbol{F}}^{b} \ {\boldsymbol{D}} \ [{\boldsymbol{F}}^{c} ]^T,$$
$$|{\Psi(t_{\text{f}})}\rangle = \mathcal{U} |{\text{vac}}\rangle = \exp\left[ \sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_c} {J}^{b,c}_{ij} b_i^\dagger c_j^\dagger{-} \text{H.c.}\right] |{\text{vac}}\rangle,$$
$$|{\Psi(t_{\text{i}})}\rangle = f(a_i^\dagger) |{\text{vac}}\rangle,$$
$$|{1_{\boldsymbol{\alpha}}}\rangle = \sum_{i=1}^\ell \alpha_i a_i^\dagger |{\text{vac}}\rangle, \quad |{\boldsymbol{\alpha}}\rangle = \mathcal{D}(\boldsymbol{\alpha})|{\text{vac}}\rangle = e^{-\tfrac12 \| \boldsymbol{\alpha}\|^2} \exp\left[ \sum_{i=1}^\ell \alpha_i a_i^\dagger \right] |{\text{vac}}\rangle.$$
$$|{\Psi(t_{\text{f}})}\rangle = \mathcal{U}|{\Psi(t_{\text{i}})}\rangle = \mathcal{U} f(a_i^\dagger)\mathcal{U}^\dagger \mathcal{U}|{\text{vac}}\rangle,$$
$$\mathcal{U} f(a_i^\dagger)\mathcal{U}^\dagger{=} f(\mathcal{U} a_i^\dagger \mathcal{U}^\dagger) = f\left(\sum_{k=1}^\ell \tilde{V}^*_{ik} a_k^\dagger{+} \sum_{k=1}^\ell \tilde{W}^*_{ik} a_k \right),$$
$$\mathcal{U}|{1_{\boldsymbol{\alpha}}}\rangle = \left(\sum_{k=1}^\ell \tilde{V}^*_{ik} a_k^\dagger{+} \sum_{k=1}^\ell \tilde{W}^*_{ik} a_k \right) \mathcal{U} |{\text{vac}}\rangle,$$
$$\mathcal{U}|{\boldsymbol{\alpha}}\rangle = \mathcal{D}\left( \boldsymbol{V}\boldsymbol{\alpha} + \boldsymbol{W} \boldsymbol{\alpha}^* \right) \mathcal{U} |{\text{vac}}\rangle$$
$$|{\boldsymbol{\nu}}\rangle = \mathcal{D}_c(\boldsymbol{\nu}) |{\text{vac}}\rangle = \exp\left[ \boldsymbol{\nu}^T \boldsymbol{c}^\dagger{-} \boldsymbol{\nu}^\dagger \boldsymbol{c} \right] |{\text{vac}}\rangle.$$
$$\begin{aligned}\overline{\mathcal{U}} |{\boldsymbol{\nu}}\rangle & = \overline{\mathcal{U}} \exp\left[ \boldsymbol{\nu}^T \boldsymbol{c}^\dagger{-} \boldsymbol{\nu}^\dagger \boldsymbol{c} \right] \overline{\mathcal{U}}^\dagger \overline{\mathcal{U}} |{\text{vac}}\rangle \\ & =\exp\left[ \boldsymbol{\nu}^T \overline{\mathcal{U}} \boldsymbol{c}^\dagger \overline{\mathcal{U}}^\dagger{-} \boldsymbol{\nu}^\dagger \overline{\mathcal{U}}\boldsymbol{c}\overline{\mathcal{U}}^\dagger \right] \overline{\mathcal{U}} |{\text{vac}}\rangle. \end{aligned}$$
$$\begin{aligned} \overline{\mathcal{U}} |{\boldsymbol{\nu}}\rangle & =\exp\left[ \boldsymbol{\nu}^T \left(\boldsymbol{c}^\dagger{-} [\overline{\boldsymbol{J}}^{c,b}]^* \boldsymbol{b} \right) - \boldsymbol{\nu}^\dagger \left(\boldsymbol{c} - \overline{\boldsymbol{J}}^{c,b} \boldsymbol{b}^\dagger \right) \right] \overline{\mathcal{U}} |{\text{vac}}\rangle \nonumber\\ & =\exp\left[ \boldsymbol{\nu}^T \boldsymbol{c}^\dagger{-} \boldsymbol{\nu}^\dagger \boldsymbol{c} \right] \exp\left[ \boldsymbol{\nu}^\dagger \overline{\boldsymbol{J}}^{c,b} \boldsymbol{b}^\dagger{-} \boldsymbol{\nu}^T [\overline{\boldsymbol{J}}^{c,b}]^* \boldsymbol{b} \right] \overline{\mathcal{U}} |{\text{vac}}\rangle \\ & = \mathcal{D}_{\boldsymbol{c}}(\boldsymbol{\nu}) \mathcal{D}_{\boldsymbol{b}}( [\overline{\boldsymbol{J}}^{c,b}]^T \boldsymbol{\nu}^*) \overline{\mathcal{U}} |{\text{vac}}\rangle\nonumber\\ & = \mathcal{D}_{\boldsymbol{c}}(\boldsymbol{\nu}) \mathcal{D}_{\boldsymbol{b}}( \overline{\boldsymbol{J}}^{b,c} \boldsymbol{\nu}^*) \exp\left[ \sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_c} \overline{J}^{b,c}_{ij} b_i^\dagger c_j^\dagger{-} \text{H.c.}\right] |{\text{vac}}\rangle. \end{aligned}$$
$$\boldsymbol{a}(t_\text{f}) = \mathcal{U}_{\text{BS}}^\dagger \boldsymbol{a}(t_{\text{i}}) \mathcal{U}_{\text{BS}} = \boldsymbol{L} \boldsymbol{a}(t_{\text{i}}) + \sqrt{\mathbb{I}_\ell - \boldsymbol{L} \boldsymbol{L}^\dagger } \boldsymbol{e},$$
$$\boldsymbol{a}(t_\text{f}) = \mathcal{U}_{\text{BS}}^\dagger \boldsymbol{a}(t_{\text{i}}) \mathcal{U}_{\text{BS}} = \boldsymbol{U} \boldsymbol{a}(t_{\text{i}}).$$
$$\begin{aligned}M_{ij} (t_{\text{f}}) & = \text{tr}\left( a_i a_j \mathcal{U}_{\text{BS}} \left[ \rho_{S} \otimes |{\text{vac}_e}\rangle \rangle{\text{vac}_e}| \right] \mathcal{U}_{\text{BS}}^\dagger \right) \\ & =\text{tr}\left( \mathcal{U}_{\text{BS}}^\dagger a_i a_j \mathcal{U}_{\text{BS}} \left[ \rho_{S} \otimes |{\text{vac}_e}\rangle \rangle{\text{vac}_e}| \right] \right)\\ & =\text{tr}\left( \left[\boldsymbol{L} \boldsymbol{a}(t_{\text{i}}) + \sqrt{\mathbb{I}_\ell - \boldsymbol{L} \boldsymbol{L}^\dagger } \ \boldsymbol{e}\right]_i \left[\boldsymbol{L} \boldsymbol{a}(t_{\text{i}}) + \sqrt{\mathbb{I}_\ell - \boldsymbol{L} \boldsymbol{L}^\dagger } \ \boldsymbol{e}\right]_j \left[ \rho_{S} \otimes |{\text{vac}_e}\rangle \rangle{\text{vac}_e}| \right] \right)\\ & =\text{tr}\left( \left[\boldsymbol{L} \boldsymbol{a}(t_{\text{i}}) \right]_i \left[\boldsymbol{L} \boldsymbol{a}(t_{\text{i}}) \right]_j \left[ \rho_{S} \otimes |{\text{vac}_e}\rangle \rangle{\text{vac}_e}| \right] \right)\\ & =\text{tr}\left( \sum_{k=1}^\ell L_{ik} {a}_k(t_{\text{i}}) \sum_{l=1}^\ell L_{jl} {a}_l(t_{\text{i}}) \rho_{S} \right) = \left[ \boldsymbol{L} \boldsymbol{M}(t_{\text{i}}) \boldsymbol{L}^T \right]_{ij}, \end{aligned}$$
$$\boldsymbol{N}(t_{\text{f}}) = \boldsymbol{L}^* \boldsymbol{N}(t_{\text{i}}) \boldsymbol{L}^T.$$
$$\boldsymbol{N}(t_{\text{f}}) =\eta \boldsymbol{N}(t_{\text{i}}), \quad \boldsymbol{M}(t_{\text{f}}) = \eta \boldsymbol{M}(t_{\text{i}}),$$
$$N_{\boldsymbol{a}} = \sum_{i=1}^\ell a_i^\dagger a_i.$$
$$N_{\boldsymbol{a}} = \sum_{\lambda=1}^\ell A_\lambda^\dagger A_\lambda,$$
$${\displaystyle\langle{N_{\boldsymbol{a}}}\rangle = \sum_{\lambda=1}^{\ell} n_\lambda, \quad \text{Var}(N_{\boldsymbol{a}}) = \sum_{\lambda=1}^{\ell} \text{Var}( A^\dagger_\lambda A_\lambda) = \sum_{\lambda=1}^{\ell} 2 n_\lambda (1+n_\lambda), \quad n_\lambda = \langle{A^\dagger_\lambda A_\lambda}\rangle = \sinh^2 r_\lambda,}$$
$$\begin{aligned}\langle{N_{\boldsymbol{b}}}\rangle & = \langle{N_{\boldsymbol{c}}}\rangle = \sum_{\lambda=1}^{\min\{\ell_b, \ell_c\}} n_\lambda, \end{aligned}$$
$$\begin{aligned}\text{Var}(N_{\boldsymbol{b}}) & = \text{Var}(N_{\boldsymbol{c}})\\& = \sum_{\lambda=1}^{\min\{\ell_b, \ell_c\}} \text{Var}( B^\dagger_\lambda B_\lambda) = \sum_{\lambda=1}^{\min\{\ell_b, \ell_c\}} n_\lambda (1+n_\lambda), \quad n_\lambda = \langle{A^\dagger_\lambda A_\lambda}\rangle = \sinh^2 r_\lambda. \end{aligned}$$
$$\left\langle \prod_{i=1}^\ell a_i^{{\dagger} m_i } \prod_{i=1}^\ell a_i^{n_i } \right\rangle.$$
$$\chi^{N}_{\rho}(\boldsymbol{\alpha}) = \chi(\boldsymbol{\alpha}) e^{\tfrac12 \|\boldsymbol{\alpha}\|^2} = \left\langle \exp\left[ \boldsymbol{\alpha}^T \boldsymbol{a}^\dagger \right] \exp\left[ - \boldsymbol{\alpha}^\dagger \boldsymbol{a} \right] \right\rangle_\rho,$$
$$\left\langle \prod_{i=1}^\ell a_i^{{\dagger} m_i } \prod_{i=1}^\ell a_i^{n_i } \right\rangle = \left. ({-}1)^{\sum_{i=1}^\ell n_i}\frac{\partial^{m_1}}{\partial \alpha_1^{m_1}} \ldots \frac{\partial^{m_\ell}}{\partial \alpha_\ell^{m_\ell}} \frac{\partial^{n_1}}{\partial \alpha_1^{*n_1}} \ldots \frac{\partial^{n_\ell}}{\partial \alpha_\ell^{* n_\ell}} \chi^{N}_\rho(\boldsymbol{\alpha}) \right|_{\boldsymbol{\alpha}=\boldsymbol{\alpha}^* = 0}.$$
$$\begin{aligned} \langle{n_i}\rangle & = N_{i,i}, \end{aligned}$$
$$\begin{aligned}\text{Var}(n_i) & = \langle{n_i^2}\rangle - \langle{n_i}\rangle^2 = |M_{i,i}|^2 + N_{i,i}(N_{i,i}+1), \end{aligned}$$
$$\begin{aligned}\text{Cov}(n_i,n_j) & = |M_{ij}|^2 + |N_{ij}|^2, \end{aligned}$$
$$\begin{aligned}\text{Var}(n_i-n_j) & = \text{Var}(n_i) + \text{Var}(n_j) - 2 \ \text{Cov}(n_i,n_j) \\ & = |M_{i,i}|^2 + N_{i,i}(N_{i,i}+1) + |M_{j,j}|^2 + N_{j,j}(N_{j,j}+1) - 2 \left( |M_{ij}|^2 + |N_{ij}|^2 \right),\\ \end{aligned}$$
$$\begin{aligned}\text{Var}(N) & = \sum_{i=1}^\ell \text{Var}(n_i) + 2 \sum_{i<j}^\ell \text{Cov}(n_i,n_j), \end{aligned}$$
$$g^{(2)}_{\boldsymbol{b}} = \frac{\langle{N_{\boldsymbol{b}}^2 - N_{\boldsymbol{b}}}\rangle}{\langle{N_{\boldsymbol{b}}}\rangle^2} = \frac{\sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_b} \langle{b_i^\dagger b_j^\dagger b_i b_j}\rangle }{\left( \sum_{i=1}^{\ell_b} \langle{b_i^\dagger b_i}\rangle\right)^2},$$
$$g^{(1,1)}_{\boldsymbol{b},\boldsymbol{c}} = \frac{\langle{N_{\boldsymbol{b}} N_{\boldsymbol{c}}}\rangle}{\langle{N_{\boldsymbol{b}}}\rangle \langle{N_{\boldsymbol{c}}}\rangle} = \frac{\sum_{i=1}^{\ell_b} \sum_{j=1}^{\ell_c} \langle{b_i^\dagger c_j^\dagger b_i c_j}\rangle }{\left( \sum_{i=1}^{\ell_b} \langle{b_i^\dagger b_i}\rangle\right) \left( \sum_{i=1}^{\ell_c} \langle{c_i^\dagger c_i}\rangle\right)}.$$
$$g^{(2)}_{\boldsymbol{b}} = g^{(2)}_{\boldsymbol{c}} = 1 + \frac{1}{K}, \text{ where } K = \frac{ \left[ \sum_{\lambda=1}^{\min(\ell_c,\ell_b)} \langle{A_\lambda^\dagger A_\lambda}\rangle \right]^2 }{\sum_{\lambda=1}^{\min(\ell_c,\ell_b)} \langle{A_\lambda^\dagger A_\lambda}\rangle^2}.$$
$$g^{(1,1)}_{\boldsymbol{b},\boldsymbol{c}} = 1 + \frac{1}{K} + \frac{1}{\langle{N}\rangle},$$
$$\langle{N_{\boldsymbol{b}/\boldsymbol{c}}}\rangle = \eta_{\boldsymbol{b}/\boldsymbol{c}} \langle{N}\rangle,$$
$$\eta_{\boldsymbol{b}/\boldsymbol{c}} = \langle{N_{\boldsymbol{b}/\boldsymbol{c}}}\rangle \left( g^{(2)}_{\boldsymbol{b}/\boldsymbol{c}} - g^{(1,1)}_{\boldsymbol{b},\boldsymbol{c}} \right),$$
$$X_{\phi} = e^{i \phi} \sum_{i=1}^\ell \alpha_i^* a_i + \text{H.c.}$$
$$V_{\phi} = \|\boldsymbol{\alpha}\|^2 + 2\boldsymbol{\alpha}^T \boldsymbol{N} \boldsymbol{\alpha}^* + \left( e^{2i\phi} \boldsymbol{\alpha}^\dagger \boldsymbol{M} \boldsymbol{\alpha}^* +\text{c.c.} \right).$$
$$V_{\phi}^{\rm max/min} = e^{{\pm} 2 r_\lambda},$$
$$X_{\phi} = e^{i \phi} \left[ \sum_{i=1}^{\ell_b} \mu_i^* b_i + \sum_{j=1}^{\ell_c} \nu_j^* c_j\right] + \text{H.c.},$$
$$V_{\phi} = \|\boldsymbol{\nu}\|^2 + \|\boldsymbol{\mu}\|^2 + 2\boldsymbol{\mu}^T \boldsymbol{N}^{b} \boldsymbol{\mu}^* + 2 \boldsymbol{\nu}^T \boldsymbol{N}^{c} \boldsymbol{\nu}^* +\left[ 2 e^{2i \phi} \boldsymbol{\mu}^\dagger \boldsymbol{M}^{b,c} \boldsymbol{\nu}^* + \text{c.c.} \right].$$
$$\begin{aligned} \textrm{H}^{\textrm{L}} & =\sum_{J}\int\textrm{d}x\,\hbar\left\{ \omega_{J}\psi_{J}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)+\frac{i}{2}v_{J}\left[\frac{\partial\psi_{J}^{{\dagger}}\left(x\right)}{\partial x}\psi_{J}\left(x\right)-\psi_{J}^{{\dagger}}\left(x\right)\frac{\partial\psi_{J}\left(x\right)}{\partial x}\right]\right.\\ & \left.\quad+\frac{v_{J}^{\prime}}{2}\frac{\partial\psi_{J}^{{\dagger}}\left(x\right)}{\partial x}\frac{\partial\psi_{J}\left(x\right)}{\partial x}+\cdots\right\}. \end{aligned}$$
$$\begin{aligned} H_{\text{NDSV}}^{\text{SP-SFWM}} & ={-}\gamma_{\text{chan}}^{SIPP}\hbar^{2}\overline{\omega}_{SIPP}\overline{v}_{SIPP}^{2}\int\textrm{d}{x}\,\psi_{S}^{{\dagger}}\left(x\right)\psi_{I}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)\psi_{P}\left(x\right)+\textrm{H.c.}\\ & \quad-\frac{\gamma_{\text{chan}}^{PPPP}}{2}\hbar^{2}\omega_{P}v_{P}^{2}\int\textrm{d}{x}\,\psi_{P}^{{\dagger}}\left(x\right)\psi_{P}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)\psi_{P}\left(x\right)\\ & \quad-2\gamma_{\text{chan}}^{PSPS}\hbar^{2}\sqrt{\omega_{P}\omega_{S}}v_{P}v_{S}\int\textrm{d}{x}\,\psi_{P}^{{\dagger}}\left(x\right)\psi_{S}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)\psi_{S}\left(x\right)\\ & \quad-2\gamma_{\text{chan}}^{PIPI}\hbar^{2}\sqrt{\omega_{P}\omega_{I}}v_{P}v_{I}\int\textrm{d}{x}\,\psi_{P}^{{\dagger}}\left(x\right)\psi_{I}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)\psi_{I}\left(x\right), \end{aligned}$$
$$\gamma_{\text{chan}}^{J_{1}J_{2}J_{3}J_{4}} = \frac{3\overline{\omega}_{J_{1}J_{2}J_{3}J_{4}}\overline{\chi}_{3}}{4\epsilon_{0}\left(\overline{n}_{J_{1}}\overline{n}_{J_{2}}\overline{n}_{J_{3}}\overline{n}_{J_{4}}\right)^{1/2}c^{2}}\frac{1}{A_{\text{chan}}^{J_{1}J_{2}J_{3}J_{4}}},$$
$$\frac{1}{A_{\text{chan}}^{J_{1}J_{2}J_{3}J_{4}}}=\frac{\int\textrm{d}{y}\textrm{d}{z}\frac{\chi^{ijkl}_{3}\left(y,z\right)}{\overline{\chi}_{3}}\left[e_{J_{1}}^{i}\left(y,z\right)e_{J_{2}}^{j}\left(y,z\right)\right]^{*}e_{J_{3}}^{k}\left(y,z\right)e_{J_{4}}^{l}\left(y,z\right)}{\mathcal{N}_{J_{1}}\mathcal{N}_{J_{2}}\mathcal{N}_{J_{3}}\mathcal{N}_{J_{4}}},$$
$$\overline{\omega}_{J_{1}J_{2}J_{3}J_{4}}=\left(\omega_{J_{1}}\omega_{J_{2}}\omega_{J_{3}}\omega_{J_{4}}\right)^{1/4}, \quad\overline{v}_{J_{1}J_{2}J_{3}J_{4}}=\left(v_{J_{1}}v_{J_{2}}v_{J_{3}}v_{J_{4}}\right)^{1/4},$$
$$\mathcal{N}_{J}=\sqrt{\int\text{d}y\text{d}z\frac{n\left(y,z;\omega_{J}\right)/\overline{n}_{J}}{v_{g}\left(y,z;\omega_{J}\right)/v_{J}}\boldsymbol{e}_{J}^{*}\left(y,z\right)\cdot\boldsymbol{e}_{J}\left(y,z\right)}.$$
$$\gamma_{\text{chan}}^{\text{SPM}}=\gamma_{\text{chan}}^{PPPP},$$
$$\gamma_{\text{chan}}^{\text{XPM}}=\sqrt{\frac{\omega_S}{\omega_P}}\gamma_{\text{chan}}^{PSPS}.$$
$$\overline{\psi}_{J}\left(x,t\right)=e^{i\omega_{J}t}\psi_{J}\left(x,t\right),$$
$$\left(\frac{\partial}{\partial t}+v_{P}\frac{\partial}{\partial x}-i\frac{v_{P}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle =i\gamma_{\text{chan}}^{PPPP}\hbar\omega_{P}v_{P}^{2}\left|\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle,$$
$$\begin{aligned} \left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right) & =i\gamma_{\text{chan}}^{SIPP}\hbar\overline{\omega}_{SIPP}\overline{v}_{SIPP}^{2}\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle ^{2}\overline{\psi}_{I}^{{\dagger}}\left(x,t\right)\\ & \quad+2i\gamma_{\text{chan}}^{PSPS}\hbar\sqrt{\omega_{P}\omega_{S}}v_{P}v_{S}\left|\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{S}\left(x,t\right),\\ \left(\frac{\partial}{\partial t}+v_{I}\frac{\partial}{\partial x}-i\frac{v_{I}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{I}\left(x,t\right) & =i\gamma_{\text{chan}}^{SIPP}\hbar\overline{\omega}_{SIPP}\overline{v}_{SIPP}^{2}\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle ^{2}\overline{\psi}_{S}^{{\dagger}}\left(x,t\right)\\ & \quad+2i\gamma_{\text{chan}}^{PIPI}\hbar\sqrt{\omega_{P}\omega_{I}}v_{P}v_{I}\left|\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{I}\left(x,t\right), \end{aligned}$$
$$\psi_{P}\left(x,t\right)\rightarrow\left\langle\psi_{P}\left(x,t\right)\right\rangle.$$
$$\begin{aligned} \left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right) & =i\mathcal{\tilde{S}}\left(x,t\right)\overline{\psi}_{I}^{{\dagger}}\left(x,t\right)+2i\mathcal{\tilde{M}_{S}}\left(x,t\right)\overline{\psi}_{S}\left(x,t\right),\\ \left(\frac{\partial}{\partial t}+v_{I}\frac{\partial}{\partial x}-i\frac{v_{I}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{I}\left(x,t\right) & =i\mathcal{\tilde{S}}\left(x,t\right)\overline{\psi}_{S}^{{\dagger}}\left(x,t\right)+2i\mathcal{\tilde{M}_{I}}\left(x,t\right)\overline{\psi}_{I}\left(x,t\right). \end{aligned}$$
$$\left(\frac{\partial}{\partial t}+v\frac{\partial}{\partial x}-i\frac{v^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}\left(x,t\right)=i\tilde{\mathcal{S}}\left(x,t\right)\overline{\psi}^{{\dagger}}\left(x,t\right)+2i\tilde{\mathcal{M}}\left(x,t\right)\overline{\psi}\left(x,t\right),$$
$$\begin{aligned}\omega_{J}\left(\kappa\right) & =v_{J}\kappa+\frac{v_{J}^\prime}{2}\kappa^{2},\\ \tilde{\mathcal{S}}\left(x,t\right) & =\int\textrm{d}{\kappa}\frac{e^{i\kappa x}}{\sqrt{2\pi}}\mathcal{S}\left(\kappa,t\right),\\ \tilde{\mathcal{M}_{J}}\left(x,t\right) & =\int\textrm{d}{\kappa}\frac{e^{i\kappa x}}{\sqrt{2\pi}}\mathcal{M}_{J}\left(\kappa,t\right), \end{aligned}$$
$$\overline{\psi}_{J}\left(x,t\right)=\int\textrm{d}{\kappa}\frac{e^{i\kappa x}}{\sqrt{2\pi}}b_{J}\left(\kappa,t\right),$$
$${\begin{aligned} \left[\frac{\partial}{\partial t}+i\omega_{S}\left(\kappa\right)\right]b_{S}\left(\kappa,t\right) & =i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{S}\left(\kappa+\kappa^{\prime},t\right)b_{I}^{{\dagger}}\left(\kappa^{\prime},t\right)+2i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{M}_{S}\left(\kappa-\kappa^{\prime},t\right)b_{S}\left(\kappa^{\prime},t\right),\\ \left[\frac{\partial}{\partial t}+i\omega_{I}\left(\kappa\right)\right]b_{I}\left(\kappa,t\right) & =i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{S}\left(\kappa+\kappa^{\prime},t\right)b_{S}^{{\dagger}}\left(\kappa^{\prime},t\right)+2i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{M}_{I}\left(\kappa-\kappa^{\prime},t\right)b_{I}\left(\kappa^{\prime},t\right), \end{aligned}}$$
$${\displaystyle\left[\frac{\partial}{\partial t}+i\omega\left(\kappa\right)\right]b\left(\kappa,t\right)=i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{S}\left(\kappa+\kappa^{\prime},t\right)b^{{\dagger}}\left(\kappa^{\prime},t\right)+2i\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{M}\left(\kappa-\kappa^{\prime},t\right)b\left(\kappa^{\prime},t\right),}$$
$$\begin{aligned}\int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{S}\left(\kappa+\kappa^{\prime},t\right)b_{J}^{{\dagger}}\left(\kappa^{\prime},t\right) & \approx\sum_{j^{\prime}}\sqrt{\frac{\Delta\kappa}{2\pi}}\mathcal{S}\left(\kappa_{j}+\kappa_{j^{\prime}},t\right)a_{J;j^{\prime}}^{{\dagger}}\left(t\right),\\ \int\frac{\textrm{d}{\kappa}^{\prime}}{\sqrt{2\pi}}\mathcal{M}_{J}\left(\kappa-\kappa^{\prime},t\right)b_{J}\left(\kappa^{\prime},t\right) & \approx\sum_{j^{\prime}}\sqrt{\frac{\Delta\kappa}{2\pi}}\mathcal{M}_{J}\left(\kappa_{j}-\kappa_{j^{\prime}},t\right)a_{J;j^{\prime}}\left(t\right), \end{aligned}$$
$$\textbf{A}^{J_{1}J_{2}}\left(t\right)=\left(\begin{array}{cc} \boldsymbol{\Delta}_{J_{1}}\left(t\right) & \boldsymbol{\zeta}\left(t\right)\\ -\boldsymbol{\zeta}^{*}\left(t\right) & -\boldsymbol{\Delta}_{J_{2}}^{*}\left(t\right) \end{array}\right),$$
$$\begin{aligned}\Delta_{J;jj^{\prime}}\left(t\right) & ={-}\frac{\omega_{J}\left(\kappa_{j}\right)}{\Delta\kappa}\delta_{jj^{\prime}}+2\sqrt{\frac{\Delta\kappa}{2\pi}}\mathcal{M}_{J}\left(\kappa_{j}-\kappa_{j^{\prime}},t\right),\\ \zeta_{jj^{\prime}}\left(t\right) & =\sqrt{\frac{\Delta\kappa}{2\pi}}\mathcal{S}\left(\kappa_{j}+\kappa_{j^{\prime}},t\right), \end{aligned}$$
$$\frac{\partial}{\partial t}\left(\begin{array}{c} \textbf{a}_{S}\left(t\right)\\ \textbf{a}_{I}^{{\dagger}}\left(t\right) \end{array}\right)=i\textbf{A}^{SI}\left(t\right)\left(\begin{array}{c} \textbf{a}_{S}\left(t\right)\\ \textbf{a}_{I}^{{\dagger}}\left(t\right) \end{array}\right),$$
$$\frac{\partial}{\partial t}\left(\begin{array}{c} \textbf{a}\left(t\right)\\ \textbf{a}^{{\dagger}}\left(t\right) \end{array}\right)=i\textbf{A}\left(t\right)\left(\begin{array}{c} \textbf{a}\left(t\right)\\ \textbf{a}^{{\dagger}}\left(t\right) \end{array}\right),$$
$$\left(\begin{array}{c} \textbf{a}_{J_{1}}\left(t+\Delta t\right)\\ \textbf{a}_{J_{2}}^{{\dagger}}\left(t+\Delta t\right) \end{array}\right)=\textbf{K}^{J_{1}J_{2}}(t)\left(\begin{array}{c} \textbf{a}_{J_{1}}\left(t\right)\\ \textbf{a}_{J_{2}}^{{\dagger}}\left(t\right) \end{array}\right),$$
$$\textbf{K}^{J_{1}J_{2}}(t)=\exp\left[i\Delta t\textbf{A}^{J_{1}J_{2}}\left(t\right)\right]=\left(\begin{array}{cc} \textbf{V}_{J_{1}}(t) & \textbf{W}_{J_{1}J_{2}}(t)\\ \textbf{W}_{J_{2}J_{1}}^{*}(t) & \textbf{V}_{J_{2}}^{*}(t) \end{array}\right).$$
$$\begin{aligned} \textbf{K}^{J_{1}J_{2}}(t,t_0) = \prod_{j=1}^\ell \textbf{K}^{J_{1}J_{2}}(t_j) = \left(\begin{array}{cc} \textbf{V}_{J_{1}}(t,t_0) & \textbf{W}_{J_{1}J_{2}}(t,t_0)\\ \textbf{W}_{J_{2}J_{1}}^{*}(t,t_0) & \textbf{V}_{J_{2}}^{*}(t,t_0) \end{array}\right), \end{aligned}$$
$$\begin{aligned}\left(\begin{array}{c} \textbf{a}_{J_{1}}\left(t\right)\\ \textbf{a}_{J_{2}}^{{\dagger}}\left(t\right) \end{array}\right) = \textbf{K}^{J_{1}J_{2}}(t,t_0) \left(\begin{array}{c} \textbf{a}_{J_{1}}\left(t_{0}\right)\\ \textbf{a}_{J_{2}}^{{\dagger}}\left(t_{0}\right) \end{array}\right), \end{aligned}$$
$$\begin{aligned}b_{S}(\kappa,t) & = \int \textrm{d}\kappa'\, \mathcal{V}_{S}(\kappa,\kappa';t,t_0) b_{S}(\kappa',t_0)+ \int \textrm{d}\kappa'\, \mathcal{W}_{SI}(\kappa,\kappa';t,t_0) b_{I}^\dagger(\kappa',t_0)\\ b_{I}^\dagger(\kappa,t) & = \int \textrm{d}\kappa'\, \mathcal{W}_{IS}^{*}(\kappa,\kappa';t,t_0) b_{S}(\kappa',t_0) + \int \textrm{d}\kappa'\, \mathcal{V}_{I}^{*}(\kappa,\kappa';t,t_0) b_{I}^\dagger(\kappa',t_0) \end{aligned}$$
$$b(\kappa,t) = \int \textrm{d}\kappa'\, \mathcal{V}(\kappa,\kappa';t,t_0) b(\kappa',t_0)+ \int \textrm{d}\kappa'\, \mathcal{W}(\kappa,\kappa';t,t_0) b^\dagger(\kappa',t_0),$$
$$\begin{aligned} \mathcal{V}_{J_{1}J_{2}}(\kappa_j,\kappa_{j'};t,t_0) & = V_{J_{1}J_{2};j,j'}(t,t_0)/\Delta \kappa,\\ \mathcal{W}_{J_{1}J_{2}}(\kappa_j,\kappa_{j'};t,t_0) & = W_{J_{1}J_{2};j,j'}(t,t_0)/\Delta \kappa. \end{aligned}$$
$$\begin{aligned} N_{J;i,j}\left(t\right) & =\left\langle a_{J;i}^{{\dagger}}\left(t\right)a_{J;j}\left(t\right)\right\rangle,\\ M_{J_{1}J_{2};i,j}\left(t\right) & =\left\langle a_{J_{1};i}\left(t\right)a_{J_{2};j}\left(t\right)\right\rangle. \end{aligned}$$
$$\textbf{N}_{J}(t+\Delta t) = \eta_{J} \textbf{N}_{J}(t), \quad \textbf{M}_{J_{1}J_{2}}(t+\Delta t) = \sqrt{\eta_{J_{1}}\eta_{J_{2}}} \textbf{M}_{J_{1}J_{2}}(t),$$
$$\eta_{J} = \exp(-\rho_{J} \Delta t),$$
$$\begin{aligned}\textrm{H}^{\textrm{L}} & = \sum_{J=P,S,I}\int\textrm{d}{k}\,\hbar\omega_{Jk}a_{J}^{{\dagger}}\left(k\right)a_{J}\left(k\right) \end{aligned}$$
$$\begin{aligned}{\textrm{H}^{\textrm{NL}}} & ={-}\frac{\gamma_{\text{chan}}^{SIPP}\hbar^{2}\overline{\omega}_{SIPP}\overline{v}_{SIPP}^{2}}{4\pi^{2}}\\ & \quad\times\int\textrm{d}{k}_{1}\textrm{d}{k}_{2}\textrm{d}{k}_{3}\textrm{d}{k}_{4}\textrm{d}{x}\,e^{{-}i\left(k_{1}+k_{2}-k_{3}-k_{4}\right)x}a_{S}^{{\dagger}}\left(k_{1}\right)a_{I}^{{\dagger}}\left(k_{2}\right) a_{P}\left(k_{3}\right)a_{P}\left(k_{4}\right)+\textrm{H.c.}, \end{aligned}$$
$$\begin{aligned}H_{I}\left(t\right) & ={-}\frac{\gamma_{\text{chan}}^{SIPP}\hbar^{2}\overline{\omega}_{SIPP}}{4\pi^{2}}\int\textrm{d}{\omega}_{1}\textrm{d}{\omega}_{2}\textrm{d}{\omega}_{3}\textrm{d}{\omega}_{4}\int\textrm{d}{x}\,e^{i\left(\omega_{1}+\omega_{2}-\omega_{3}-\omega_{4}\right)t}\\ & \quad \times e^{{-}i\left[k_{S}\left(\omega_{1}\right)+k_{I}\left(\omega_{2}\right)-k_{P}\left(\omega_{3}\right)-k_{P}\left(\omega_{4}\right)\right]x}a_{S}^{{\dagger}}\left(\omega_{1}\right)a_{I}^{{\dagger}}\left(\omega_{2}\right)a_{P}\left(\omega_{3}\right)a_{P}\left(\omega_{4}\right)+\textrm{H.c.} \end{aligned}$$
$$k_{J}\left(\omega\right)=k_{J}+\frac{\omega-\omega_{J}}{v_{J}},$$
$$a_{P}\left(\omega\right)\rightarrow\left\langle a_{P}\left(\omega\right)\right\rangle =\sqrt{\frac{N_{P}\tau}{\left(\pi/2\right)^{1/2}}}e^{-\tau^{2}\left(\omega-\omega_{P}\right)^{2}},$$
$$\omega_{{\pm}}=\omega_{3}\pm\omega_{4},$$
$$\begin{aligned}& \int\textrm{d}{\omega_{3}}\textrm{d}{\omega_{4}}\,e^{{-}i\left(\omega_{3}+\omega_{4}\right)t}e^{i\frac{\omega_{3}+\omega_{4}-2\omega_{P}}{v_{P}}x}e^{-\tau^{2}\left(\omega_{3}-\omega_{P}\right)^{2}}e^{-\tau^{2}\left(\omega_{4}-\omega_{P}\right)^{2}}\\ = & \int\textrm{d}{\omega_{+}}\,e^{{-}i\omega_{+}t}e^{i\frac{\omega_{+}-2\omega_{P}}{v_{P}}x}e^{-\frac{\tau^{2}}{2}\left(\omega_{+}-2\omega_{P}\right)^{2}}\int\frac{\textrm{d}{\omega_{-}}}{2}e^{-\frac{\tau^{2}}{2}\omega_{-}^{2}}\\ = & \int\textrm{d}{\omega_{+}}\,e^{{-}i\omega_{+}t}e^{i\frac{\omega_{+}-2\omega_{P}}{v_{P}}x}e^{-\frac{\tau^{2}}{2}\left(\omega_{+}-2\omega_{P}\right)^{2}}\frac{\left(\pi/2\right)^{1/2}}{\tau}, \end{aligned}$$
$$\begin{aligned} H_{I}\left(t\right) & ={-}\frac{\gamma_{\text{chan}}^{SIPP}\hbar^{2}\overline{\omega}_{SIPP}L N_{P}}{4\pi^{2}}\int\text{d}\omega_{1}\textrm{d}{\omega_{2}}\textrm{d}{\omega_{+}}e^{i\left(\omega_{1}+\omega_{2}-\omega_{+}\right)t}F\left(\omega_{1},\omega_{2},\omega_{+}\right)\\ & \quad\times e^{-\frac{\tau^{2}}{2}\left(\omega_{+}-2\omega_{P}\right)^{2}}a_{S}^{{\dagger}}\left(\omega_{1}\right)a_{I}^{{\dagger}}\left(\omega_{2}\right)+\textrm{H.c.}, \end{aligned}$$
$$\begin{aligned}F\left(\omega_{1},\omega_{2},\omega_{+}\right) & =\int_{{-}L/2}^{L/2}\frac{\textrm{d}{x}}{L}e^{{-}i\left(\frac{\omega_{1}-\omega_{S}}{v_{S}}+\frac{\omega_{2}-\omega_{I}}{v_{I}}-\frac{\omega_{+}-2\omega_{P}}{v_{P}}\right)x}\\ & =\text{sinc}\left[\left(\frac{\omega_{1}-\omega_{S}}{v_{S}}+\frac{\omega_{2}-\omega_{I}}{v_{I}}-\frac{\omega_{+}-2\omega_{P}}{v_{P}}\right)\frac{L}{2}\right], \end{aligned}$$
$$\mathcal{U}=\mathcal{T}\exp{\left(-\frac{i}{\hbar}\int_{-\infty}^{\infty} \textrm{d}{t^{\prime}}H_{I}\left(t^{\prime}\right)\right)},$$
$$\begin{aligned} \mathcal{U} & =\exp{\left(-\frac{i}{\hbar}\int_{-\infty}^{\infty} \textrm{d}{t^{\prime}}H_{I}\left(t^{\prime}\right)\right)}\\ & =\exp{\left(i\int\textrm{d}{\omega_{1}}\textrm{d}{\omega_{2}}\,J_{1;SI}\left(\omega_{1},\omega_{2}\right)a_{S}^{{\dagger}}\left(\omega_{1}\right)a_{I}^{{\dagger}}\left(\omega_{2}\right)+\textrm{H.c.}\right)}, \end{aligned}$$
$$J_{1;SI}\left(\omega_{1},\omega_{2}\right)=\frac{\Phi\tau}{2\pi}F\left(\omega_{1},\omega_{2},\omega_{1}+\omega_{2}\right)e^{-\frac{\tau^{2}}{2}\left(\omega_{1}+\omega_{2}-2\omega_{P}\right)^{2}},$$
$$\Phi=\gamma_{\text{chan}}^{SIPP}\frac{\hbar\overline{\omega}_{SIPP}N_{P}}{\tau}L,$$
$$J_{SI}(\omega,\omega')=\sum_{\lambda} \xi_{\lambda} f_{\lambda;S}(\omega) f_{\lambda;I}(\omega'),$$
$$K=\frac{\left(\sum_{\lambda} \sinh^2\xi_\lambda\right)^2}{\sum_{\lambda} \sinh^4\xi_\lambda}.$$
$$\begin{aligned}\Omega_{1} & = \omega_{1} - \omega_{S},\\ \Omega_{2} & = \omega_{2} - \omega_{I}, \end{aligned}$$
$$\begin{aligned}J_{1;SI}\left(\Omega_{1},\Omega_{2}\right) & =\frac{\Phi\tau}{2\pi}e^{{-}s\left(\Omega_{1}v_{S}^{{-}1}+\Omega_{2}v_{I}^{{-}1}-\left(\Omega_{1}+\Omega_{2}\right)v_{P}^{{-}1}\right)^{2}\frac{L^{2}}{4}}e^{-\frac{\tau^{2}}{2}\left(\Omega_{1}+\Omega_{2}\right)^{2}}\\ & = \frac{\Phi\tau}{2\pi}e^{-\Omega_{1}^{2}\left(s\left(v_{S}^{{-}1}-v_{P}^{{-}1}\right)^{2}\frac{L^{2}}{4}+\frac{\tau^{2}}{2}\right)}e^{-\Omega_{2}^{2}\left(s\left(v_{I}^{{-}1}-v_{P}^{{-}1}\right)^{2}\frac{L^{2}}{4}+\frac{\tau^{2}}{2}\right)}\\ & \quad\times e^{{-}2\Omega_{1}\Omega_{2}\left(s\left(v_{S}^{{-}1}-v_{P}^{{-}1}\right)\left(v_{I}^{{-}1}-v_{P}^{{-}1}\right)\frac{L^{2}}{4}+\frac{\tau^{2}}{2}\right)}, \end{aligned}$$
$$s\left(v_{S}^{{-}1}-v_{P}^{{-}1}\right)\left(v_{I}^{{-}1}-v_{P}^{{-}1}\right)\frac{L^{2}}{4}+\frac{\tau^{2}}{2}=0.$$
$$\tau_{J}^{2}=s\left(v_{J}^{{-}1}-v_{P}^{{-}1}\right)^{2}\frac{L^{2}}{4}+\frac{\tau^{2}}{2},$$
$$J_{1;SI}\left(\Omega_{1},\Omega_{2}\right)=\xi_{0}f_{0;S}\left(\Omega_{1}\right)f_{0;I}\left(\Omega_{2}\right),$$
$$f_{0;J}\left(\Omega\right)=\sqrt{\frac{\tau_{J}}{\left(\pi/2\right)^{1/2}}}e^{-\Omega^{2}\tau_{J}^{2}}\equiv \phi_{0;J}\left(\Omega\right),$$
$$\xi_{0}=\frac{\Phi\tau}{2\left(2\pi\tau_{S}\tau_{I}\right)^{1/2}}\equiv \overline{\xi}.$$
$$\tau_{S}^{2}+\tau_{I}^{2} = s\left(v_{S}^{{-}1}-v_{I}^{{-}1}\right)^{2}\frac{L^{2}}{4},$$
$$\sqrt{\frac{\tau^{2}}{\tau_{S}\tau_{I}}}=\sqrt{\frac{2\tau_{S}\tau_{I}}{\tau_{S}^{2}+\tau_{I}^{2}}}=\sqrt{\frac{2R}{1+R^{2}}}\le 1,$$
$$R = \frac{\tau_{S}}{\tau_{I}},$$
$$J_{SI}\left(\Omega_{1},\Omega_{2}\right)\approx J_{1;SI}\left(\Omega_{1},\Omega_{2}\right) + J_{3;SI}\left(\Omega_{1},\Omega_{2}\right)- i K_{3;SI}\left(\Omega_{1},\Omega_{2}\right),$$
$$\begin{aligned} J_{3;SI}(\Omega_1,\Omega_2)-i K_{3;SI}(\Omega_1,\Omega_2) & =\overline{\xi}^3 \Bigg(\frac{\phi_{0;S}\left(\Omega_{1}\right)\phi_{0;I}\left(\Omega_{2}\right)-\phi_{1;S}\left(\Omega_{1}\right)\phi_{1;I}\left(\Omega_{2}\right)}{12}\\ & \quad -i\frac{\phi_{0;S}\left(\Omega_{1}\right)\phi_{1;I}\left(\Omega_{2}\right)-\phi_{1;S}\left(\Omega_{1}\right)\phi_{0;I}\left(\Omega_{2}\right)}{4\sqrt{3}}\Bigg), \end{aligned}$$
$$\phi_{1;J}\left(\Omega\right)=\sqrt{3}\phi_{0;J}\left(\Omega\right)\text{erfi}\left(\sqrt{\frac{2}{3}}\tau_{J}\Omega\right),$$
$$\int\textrm{d}{\Omega} \ \phi_{i;J}^*(\Omega) \phi_{j;J}(\Omega)=\delta_{i,j}.$$
$$J_{SI}\left(\Omega_{1},\Omega_{2}\right)\approx \textbf{u}_{S}\left(\Omega_{1}\right)\textbf{L}\left[\textbf{u}_{I}\left(\Omega_{2}\right)\right]^{{\dagger}},$$
$$\begin{aligned}\textbf{L} & =\left(\begin{array}{cc} \overline{\xi}+\frac{\overline{\xi}^{3}}{12} & -i\frac{\overline{\xi}^{3}}{4\sqrt{3}}\\ i\frac{\overline{\xi}^{3}}{4\sqrt{3}} & -\frac{\overline{\xi}^{3}}{12} \end{array}\right)\\ & =\frac{\overline{\xi}}{2}\mathbb{I}_{2}+\frac{\overline{\xi}^{3}}{4\sqrt{3}}\sigma_{2}+\left(\frac{\overline{\xi}}{2}+\frac{\overline{\xi}^{3}}{12}\right)\sigma_{3}\\&=\frac{\overline{\xi}}{2}\left(\mathbb{I}_{2}+\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\textbf{n}\cdot{\boldsymbol{\sigma}}\right), \end{aligned}$$
$$\textbf{u}_{J}\left(\Omega\right)=\left(\phi_{0;J}\left(\Omega\right),\phi_{1;J}\left(\Omega\right)\right).$$
$$J_{SI}\left(\Omega_{1},\Omega_{2}\right)\approx \textbf{u}^\prime_{S}\left(\Omega_{1}\right)\textbf{L}^\prime\left[\textbf{u}^\prime_{I}\left(\Omega_{2}\right)\right]^{{\dagger}},$$
$$\xi_{{\pm}}=\frac{\overline{\xi}}{2}\left(1\pm\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right),$$
$$\textbf{R}\left(\theta\right)=\left(\begin{array}{cc} \cos\left(\frac{\theta}{2}\right) & i\sin\left(\frac{\theta}{2}\right)\\ i\sin\left(\frac{\theta}{2}\right) & \cos\left(\frac{\theta}{2}\right) \end{array}\right),\quad\tan(\theta)=\frac{\overline{\xi}^2}{2\sqrt{3}(1+\overline{\xi}^2/6)}.$$
$$\begin{aligned}f_{0;J}\left(\Omega\right) & =\cos\left(\tfrac{\theta}{2}\right)\phi_{0;J}\left(\Omega\right)+i\sin\left(\tfrac{\theta}{2}\right)\phi_{1;J}\left(\Omega\right),\\ f_{1;J}\left(\Omega\right) & =\cos\left(\tfrac{\theta}{2}\right)\phi_{1;J}\left(\Omega\right)+i\sin\left(\tfrac{\theta}{2}\right)\phi_{0;J}\left(\Omega\right). \end{aligned}$$
$$\begin{aligned}\left\langle n_{\text{pairs}}\right\rangle & ={\left\langle\textrm{vac}\right\vert}\mathcal{U}^{{\dagger}}\int\text{d}\Omega\,a_{S}^{{\dagger}}\left(\Omega\right)a_{S}\left(\Omega\right)\mathcal{U}\left\vert\text{vac}\right\rangle\\ & ={\left\langle\textrm{vac}\right\vert}\mathcal{U}^{{\dagger}}\int\text{d}\Omega\,a_{I}^{{\dagger}}\left(\Omega\right)a_{I}\left(\Omega\right)\mathcal{U}\left\vert\text{vac}\right\rangle\\ & =\sum_{\lambda}\sinh^{2}\xi_{\lambda}, \end{aligned}$$
$$\left\langle n_{\text{pairs}}\right\rangle=\sinh^{2}\overline{\xi},$$
$$K=1,$$
$$\left\langle n_{\text{pairs}}\right\rangle=\sinh^{2}\left[\frac{\overline{\xi}}{2}\left(1+\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right] +\sinh^{2}\left[\frac{\overline{\xi}}{2}\left(1-\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right],$$
$$K=1+2\frac{\sinh^{2}\left[\frac{\overline{\xi}}{2}\left(1+\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right] \sinh^{2}\left[\frac{\overline{\xi}}{2}\left(1-\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right]}{\sinh^{4} \left[\frac{\overline{\xi}}{2}\left(1+\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right] +\sinh^{4}\left[\frac{\overline{\xi}}{2}\left(1-\sqrt{1+\frac{\overline{\xi}^{2}}{3}+\frac{\overline{\xi}^{4}}{9}}\right)\right]}.$$
$$\overline{\xi} = \frac{\Phi}{2\left(2\pi\right)^{1/2}},$$
$$\begin{aligned} H_{\text{ring}}^{\text{SFWM}} = & -\frac{1}{4\epsilon_{0}}\left(\frac{\hbar\omega_{\text{P}}}{2\mathcal{L}}\right)\sqrt{\frac{\hbar\omega_{\text{S}}}{2\mathcal{L}}\frac{\hbar\omega_{\text{I}}}{2\mathcal{L}}}\left(\frac{4!}{2!1!1!}\right)\int \textrm{d}\boldsymbol{r}_{{\perp}}d\zeta\ \Gamma_{3}^{ijkl}(\boldsymbol{r}_{{\perp}},\zeta) \\ & \times \left(\textsf{d}_{\text{S}}^{k}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{d}_{\text{I}}^{l}(\boldsymbol{r}_{{\perp}},\zeta)\right)^{*}\textsf{d}_{\text{P}}^{i}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{d}_{\text{P}}^{j}(\boldsymbol{r}_{{\perp}},\zeta)e^{i\Delta\kappa\zeta}c_{\text{S}}^{{\dagger}}c_{\text{I}}^{{\dagger}}c_{\text{P}}c_{\text{P}} + \textrm{H.c.}, \end{aligned}$$
$$H_{\text{ring}}^{\text{SFWM}}={-}\frac{\hbar^{2}\left(\omega_{\text{P}}^{2}\omega_{\text{S}}\omega_{\text{I}}\right)^{\frac{1}{4}}v_{\text{P}}\sqrt{v_{\text{S}}v_{\text{I}}}\gamma_{\text{ring}}^{\text{SIPP}}}{\mathcal{L}}c_{\text{S}}^{{\dagger}}c_{\text{I}}^{{\dagger}}c_{\text{P}}c_{\text{P}} + \textrm{H.c.},$$
$$\begin{aligned} \gamma_{\text{ring}}^{J_1J_2J_3J_4}= & \frac{3\left(\omega_{J_1}\omega_{J_2}\omega_{J_3}\omega_{J_4}\right)^{\frac{1}{4}}\epsilon_{0}}{4\sqrt{v_{J_1}v_{J_2}v_{J_3}v_{J_4}}}\frac{1}{\mathcal{L}}\\ & \times\int \textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta\ \chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}},\zeta)\left(\textsf{e}_{J_3}^{k}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{e}_{J_4}^{l}(\boldsymbol{r}_{{\perp}},\zeta)\right)^{*}\textsf{e}_{J_1}^{i}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{e}_{J_2}^{j}(\boldsymbol{r}_{{\perp}},\zeta). \end{aligned}$$
$$\gamma_{\text{ring}}^{J_1J_2J_3J_4}=\frac{3\left(\omega_{J_1}\omega_{J_2}\omega_{J_3}\omega_{J_4}\right)^{\frac{1}{4}}\overline{\chi}_{3}}{4\epsilon_{0}\sqrt{\overline{n}_{J_1}\overline{n}_{J_2}\overline{n}_{J_3}\overline{n}_{J_4}}c^{2}}\frac{1}{A_{\text{ring }}^{J_1J_2J_3J_4}},$$
$$\frac{1}{A_{\text{ring}}^{J_1J_2J_3J_4}} = \frac{\mathcal{L}^{{-}1}\int \textrm{d}\boldsymbol{r}_{{\perp}}{\rm d}\zeta\ \left(\chi_{3}^{ijkl}(\boldsymbol{r}_{{\perp}},\zeta)/\overline{\chi}_{3}\right)\left(\textsf{e}_{J_3}^{k}(\boldsymbol{r}_{{\perp}},\zeta) \textsf{e}_{J_4}^{l}(\boldsymbol{r}_{{\perp}},\zeta)\right)^{*}\textsf{e}_{J_1}^{i}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{e}_{J_2}^{j} (\boldsymbol{r}_{{\perp}},\zeta)}{\mathcal{N}_{J_{1}}\mathcal{N}_{J_{2}}\mathcal{N}_{J_{3}}\mathcal{N}_{J_{4}}},$$
$$\mathcal{N}_{J} = \sqrt{\int\textrm{d}\boldsymbol{r}_{{\perp}}\frac{n(\boldsymbol{r}_{{\perp}};\omega_{J})/\overline{n}_{J}}{v_{g}(\boldsymbol{r}_{{\perp}}; \omega_{J})/v_{J}}\textsf{e}_{J}^{*}(\boldsymbol{r}_{{\perp}},0)\cdot\textsf{e}_{J}(\boldsymbol{r}_{{\perp}},0)}.$$
$${\begin{aligned}H_{\text{ring}}^{\text{SPDC}} = & -\frac{2}{3\epsilon_{0}^{2}}\Big(\frac{3!}{1!1!1!}\Big)\sqrt{\frac{\hbar\omega_{\text{SH}}}{2\mathcal{L}}}\sqrt{\frac{\hbar\omega_{\text{F1}}}{2\mathcal{L}}\frac{\hbar\omega_{\text{F2}}}{2\mathcal{L}}}\\ & \times \int\textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta\ \Gamma_{ijk}^{(2)}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{d}_{\text{SH}}^{i}(\boldsymbol{r}_{{\perp}},\zeta)\left[\textsf{d}_{\text{F1}}^{j}(\boldsymbol{r}_{{\perp}},\zeta)\right]^{*}\left[\textsf{d}_{\text{F2}}^{k}(\boldsymbol{r}_{{\perp}},\zeta)\right]^{*}e^{i\Delta\kappa\zeta}c_{\text{SH}}c_{\text{F1}}^{{\dagger}}c_{\text{F2}}^{{\dagger}} + \textrm{H.c.}\\ \equiv & -\hbar\Lambda_{\text{ring}}^{\text{SPDC}}c_{\text{SH}}c_{\text{F1}}^{{\dagger}}c_{\text{F2}}^{{\dagger}}+\textrm{H.c.}, \end{aligned}}$$
$$\begin{aligned} \Lambda_{\text{ring}}^{\text{SPDC}} = & -\frac{1}{3\epsilon_{0}}\Big(\frac{3!}{1!1!1!}\Big)\sqrt{\frac{\hbar\omega_{\text{SH}}}{2\mathcal{L}}}\sqrt{\frac{\hbar\omega_{\text{F1}}}{2\mathcal{L}}}\\ & \times \sqrt{\frac{\hbar\omega_{\text{F2}}}{2\mathcal{L}}}\int \textrm{d}\boldsymbol{r}_{{\perp}}\textrm{d}\zeta\ \Gamma_{ijk}^{(2)}(\boldsymbol{r}_{{\perp}},\zeta)\textsf{d}_{\text{SH}}^{i}(\boldsymbol{r}_{{\perp}},\zeta)\left[\textsf{d}_{\text{F1}}^{j}(\boldsymbol{r}_{{\perp}},\zeta)\right]^{*}\left[\textsf{d}_{\text{F2}}^{k}(\boldsymbol{r}_{{\perp}},\zeta)\right]^{*}, \end{aligned}$$
$$i\hbar\frac{\textrm{d}\mathcal{O}(t)}{\textrm{d}t}=\left[\mathcal{O}(t),H_{\text{ring}}\right],$$
$${\begin{aligned}H_{\text{ring}} = & H_{\text{ring}}^{\text{L}}+H_{\text{ring}}^{\text{NL}}\\ = & \sum_{J}\left(\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x)\psi_{J}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x)\frac{\partial\psi_{J}(x)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x)}{\partial x}\psi_{J}(x)\right)\textrm{d}x\right)\\ & +\sum_{J}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J}+\sum_{J}(\hbar\gamma_{J}c_{J}^{{\dagger}}\psi_{J}(0)+\textrm{H.c.})\\ & +\sum_{J}\left(\hbar\omega_{J}\int\psi_{J\text{ph}}^{{\dagger}}(x)\psi_{J\text{ph}}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J\text{ph}}\int\left(\psi_{J\text{ph}}^{{\dagger}}(x)\frac{\partial\psi_{J\text{ph}}(x)}{\partial x}-\frac{\partial\psi_{J\text{ph}}^{{\dagger}}(x)}{\partial x}\psi_{J\text{ph}}(x)\right)\textrm{d}x\right)\\ & +\sum_{J}(\hbar\gamma_{J\text{ph}}c_{J}^{{\dagger}}\psi_{J\text{ph}}(0)+\textrm{H.c.})+H_{\text{ring}}^{\text{NL}}, \end{aligned}}$$
$$\begin{aligned} i\hbar\frac{d\mathcal{O}(t)}{dt} = & \sum_{J}\hbar\omega_{J}\left[\mathcal{O}(t),c_{J}^{{\dagger}}c_{J}\right] +\sum_{J}(\hbar\gamma_{J}\left[\mathcal{O}(t),c_{J}^{{\dagger}}\right]\psi_{J}(0)+\textrm{H.c.})\\ & +\sum_{J}(\hbar\gamma_{J\text{ph}} \left[\mathcal{O}(t),c_{J}^{{\dagger}}\right]\psi_{J\text{ph}}(0)+\textrm{H.c.})\\ & +\left[\mathcal{O}(t),H_{\text{ring}}^{\text{NL}}\left(\left\{ c_{J}(t)\right\},\left\{ c_{J}^{{\dagger}}(t)\right\};t\right)\right], \end{aligned}$$
$${\begin{aligned} H_{\text{ring}}^{\text{SP-SFWM}} = & H_{\text{ring}}^{\text{L}}+H_{\text{ring}}^{\text{SP-SFWM,NL}}\\ = & \sum_{J\in\{\text{P,S,I}\}}\left(\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x)\psi_{J}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x)\frac{\partial\psi_{J}(x)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x)}{\partial x}\psi_{J}(x)\right)\textrm{d}x\right)\\ & + \sum_{J\in\{\text{P,S,I}\}}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J}+\sum_{J}(\hbar\gamma_{J}c_{J}^{{\dagger}}\psi_{J}(0)+\text{H.c.})\\ & + \sum_{J\in\{\text{P,S,I}\}}\left(\hbar\omega_{J}\int\psi_{J\text{ph}}^{{\dagger}}(x)\psi_{J\text{ph}}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J\text{ph}}\right.\\&\times\left.\int\left(\psi_{J\text{ph}}^{{\dagger}}(x)\frac{\partial\psi_{J\text{ph}}(x)}{\partial x}-\frac{\partial\psi_{J\text{ph}}^{{\dagger}}(x)}{\partial x}\psi_{J\text{ph}}(x)\right)\textrm{d}x\right)+\sum_{J}(\hbar\gamma_{J\text{ph}}c_{J}^{{\dagger}}\psi_{J\text{ph}}(0)+\text{H.c.})\\ & + \left(-\hbar\Lambda_{\text{ring}}^{\text{SFWM}}c_{\text{P}}c_{\text{P}}c_{\text{S}}^{{\dagger}}c_{\text{I}}^{{\dagger}}+\text{H.c.}\right)-\hbar\eta_{\text{ring}}^{\text{SPM}}c_{\text{P}}^{{\dagger}}c_{\text{P}}^{{\dagger}}c_{\text{P}}c_{\text{P}}-\hbar\zeta_{\text{ring}}^{\text{XPM}}\left(c_{\text{S}}^{{\dagger}}c_{\text{P}}^{{\dagger}}c_{\text{S}}c_{\text{P}}+c_{\text{I}}^{{\dagger}}c_{\text{P}}^{{\dagger}}c_{\text{I}}c_{\text{P}}\right), \end{aligned}}$$
$$\begin{aligned}\Lambda_{\text{ring}}^{\text{SFWM}} & =\frac{\hbar\left(\omega_{\text{P}}^{2}\omega_{\text{S}}\omega_{\text{I}}\right)^{\frac{1}{4}}v_{\text{P}}\sqrt{v_{\text{S}}v_{\text{I}}}\gamma_{\text{ring}}^{\text{SIPP}}}{\mathcal{L}},\\ \eta_{\text{ring}}^{\text{SPM}} & =\frac{\hbar\omega_{\text{P}}v_{\text{P}}^{2}\gamma_{\text{ring}}^{\textrm{PPPP}}}{2\mathcal{L}},\\ \zeta_{\text{ring}}^{\text{XPM}} & =\frac{2\hbar\omega_{\text{P}}v_{\text{P}}v_{S(I)}\gamma_{\text{ring}}^{\text{SS(II)PP}}}{\mathcal{L}}, \end{aligned}$$
\begin{align} \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{P}}+i\omega_{\text{P}}-2i\eta c_{\text{P}}^{{\dagger}}(t)c_{\text{P}}(t)\right)c_{\text{P}}(t) & ={-}i\gamma_{\text{P}}^{*}\psi_{\text{\text{P}<}}(0,t)-i\gamma_{\text{Pph}}^{*}\psi_{\text{Pph}<}(0,t)\nonumber\\&\quad+2i\Lambda^{*}c_{\text{P}}^{{\dagger}}(t)c_{\text{S}}(t)c_{\text{I}}(t),\nonumber\\ \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{S}}+i\omega_{\text{S}}-i\zeta c_{\text{P}}^{{\dagger}}(t)c_{\text{P}}(t)\right)c_{\text{S}}(t) & ={-}i\gamma_{\text{S}}^{*}\psi_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\psi_{\text{Sph<}}(0,t)\nonumber\\&\quad+i\Lambda c_{\text{P}}(t)c_{\text{P}}(t)c_{\text{I}}^{{\dagger}}(t),\nonumber\\ \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{I}}+i\omega_{\text{I}}-i\zeta c_{\text{P}}^{{\dagger}}(t)c_{\text{P}}(t)\right)c_{\text{I}}(t), & ={-}i\gamma_{\text{I}}^{*}\psi_{\text{I<}}(0,t)-i\gamma_{\text{Iph}}^{*}\psi_{\text{Iph<}}(0,t)\nonumber\\&\quad+i\Lambda c_{\text{P}}(t)c_{\text{P}}(t)c_{\text{S}}^{{\dagger}}(t), \end{align}
$$\begin{aligned} \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{P}}-2i\eta\overline{c}_{\text{P}}^{{\dagger}}(t)\overline{c}_{\text{P}}(t)\right)\overline{c}_{\text{P}}(t) & ={-}i\gamma_{\text{P}}^{*}\overline{\psi}_{\text{P}<}(0,t)-i\gamma_{\text{Pph}}^{*}\overline{\psi}_{\text{Pph}<}(0,t)\\&\quad+2i\Lambda^{*}\overline{c}_{\text{P}}^{{\dagger}}(t)\overline{c}_{\text{S}}(t)\overline{c}_{\text{I}}(t)e^{i\Delta_{\text{ring}}t},\\ \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{S}}-i\zeta\overline{c}_{\text{P}}^{{\dagger}}(t)\overline{c}_{\text{P}}(t)\right)\overline{c}_{\text{S}}(t) & ={-}i\gamma_{\text{S}}^{*}\overline{\psi}_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\overline{\psi}_{\text{Sph<}}(0,t)\\&\quad+i\Lambda\overline{c}_{\text{P}}(t)\overline{c}_{\text{P}}(t)\overline{c}_{\text{I}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t},\\ \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{I}}-i\zeta\overline{c}_{\text{P}}^{{\dagger}}(t)\overline{c}_{\text{P}}(t)\right)\overline{c}_{\text{I}}(t) & ={-}i\gamma_{\text{I}}^{*}\overline{\psi}_{\text{I<}}(0,t)-i\gamma_{\text{Iph}}^{*}\overline{\psi}_{\text{Iph<}}(0,t)\\&\quad+i\Lambda\overline{c}_{\text{P}}(t)\overline{c}_{\text{P}}(t)\overline{c}_{\text{S}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}, \end{aligned}$$
$$\overline{c}_{\text{P}}(t)\rightarrow\left\langle \overline{c}_{\text{P}}(t)\right\rangle e^{i\Delta_{\text{P}}t}\equiv\tilde{\beta}_{\text{P}}(t)e^{i\Delta_{\text{P}}t},$$
$$\left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{P}}+i\Delta_{\text{P}}-2i\eta|\tilde{\beta}_{\text{P}}(t)|^{2}\right)\tilde{\beta}_{\text{P}}(t)={-}i\gamma_{\text{P}}^{*}\tilde{\psi}_{\text{P}<}(0,t),$$
$$\begin{aligned} \left(\frac{\textrm{d}}{\textrm{d}t}+i\Delta_{\text{P}}+\overline{\Gamma}_{\text{S}}-i\zeta\left|\tilde{\beta}_{\text{P}}(t)\right|^{2}\right)\tilde{c}_{\text{S}}(t) & ={-}i\gamma_{\text{S}}^{*}\tilde{\psi}_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\tilde{\psi}_{\text{Sph<}}(0,t)\\&\quad+i\Lambda\left[\tilde{\beta}_{\text{P}}(t)\right]^{2}\tilde{c}_{\text{I}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t},\\ \left(\frac{\textrm{d}}{\textrm{d}t}+i\Delta_{\text{P}}+\overline{\Gamma}_{\text{I}}-i\zeta\left|\tilde{\beta}_{\text{P}}(t)\right|^{2}\right)\tilde{c}_{\text{I}}(t) & ={-}i\gamma_{\text{I}}^{*}\tilde{\psi}_{\text{I<}}(0,t)-i\gamma_{\text{Iph}}^{*}\tilde{\psi}_{\text{Iph<}}(0,t)\\&\quad+i\Lambda\left[\tilde{\beta}_{\text{P}}(t)\right]^{2}\tilde{c}_{\text{S}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}, \end{aligned}$$
$$\frac{\textrm{d}}{\textrm{d}t}\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{I}}^{{\dagger}}(t) \end{pmatrix}=M_{\text{SP-SFWM}}(t)\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{I}}^{{\dagger}}(t) \end{pmatrix}+D_{\text{SP-SFWM}}(t),$$
$$M_{\text{SP-SFWM}}(t)=\begin{pmatrix}-\overline{\Gamma}_{\text{S}}-i\Delta_{\text{P}}+i\zeta\left|\tilde{\beta}_{\text{P}}(t)\right|^{2} & i\Lambda\left[\tilde{\beta}_{\text{P}}(t)\right]^{2}e^{{-}i\Delta_{\text{ring}}t}\\ -i\Lambda^{*}\left[\tilde{\beta}_{\text{P}}^{*}(t)\right]^{2}e^{i\Delta_{\text{ring}}t} & -\overline{\Gamma}_{\text{I}}+i\Delta_{\text{P}}-i\zeta^{*}\left|\tilde{\beta}_{\text{P}}(t)\right|^{2} \end{pmatrix},$$
$$D_{\text{SP-SFWM}}(t)=\begin{pmatrix}-i\gamma_{\text{S}}^{*}\tilde{\psi}_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\tilde{\psi}_{\text{Sph<}}(0,t)\\ i\gamma_{\text{I}}\tilde{\psi}_{\text{I<}}^{{\dagger}}(0,t)+i\gamma_{\text{Iph}}\tilde{\psi}_{\text{Iph<}}^{{\dagger}}(0,t) \end{pmatrix}.$$
$$\frac{\textrm{d}}{\textrm{d}t}G(t,t')=M_{\text{SP-SFWM}}(t)G(t,t').$$
$$\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{I}}^{{\dagger}}(t) \end{pmatrix}=G(t_{0},t)\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{I}}^{{\dagger}}(t) \end{pmatrix}+\int_{-\infty}^{t}\textrm{d}t^{\prime}\ \theta(t-t')G\left(t,t^{\prime}\right)D\left(t^{\prime}\right).$$
$$\begin{aligned} N_{\text{S(I)}}(t_{1},t_{2}) & \equiv v_{\text{S(I)}}\left\langle \tilde{\psi}_{\text{S(I)>}}^{{\dagger}}(0,t_{1})\tilde{\psi}_{\text{S(I)>}}(0,t_{2})\right\rangle \\ M_{\text{SI(IS)}}(t_{1},t_{2}) & \equiv\sqrt{v_{\text{S(I)}}v_{\text{I(S)}}}\left\langle\tilde{\psi}_{\text{S(I)>}}(0,t_{1})\tilde{\psi}_{\text{I(S)>}}(0,t_{2})\right\rangle, \end{aligned}$$
$$\tilde{\psi}_{J>}(0,t)=\tilde{\psi}_{J<}(0,t)-\frac{i\gamma_{J}}{v_{J}}\tilde{c}_{J}(t),$$
$$\begin{aligned}N_{\text{S}}(t_{1},t_{2}) = & v_{\text{S}}\left\langle \left[\tilde{\psi}_{\text{S<}}^{{\dagger}}(0,t_{1})+\frac{i\gamma_{\text{S}}^{*}}{v_{\text{S}}}\tilde{c}_{\text{S}}^{{\dagger}}(t_{1})\right]\left[\tilde{\psi}_{\text{S<}}(0,t_{2})-\frac{i\gamma_{\text{S}}}{v_{\text{S}}}\tilde{c}_{\text{S}}(t_{2})\right]\right\rangle \\ = & v_{\text{S}}\frac{|\gamma_{\text{S}}|^{2}}{v_{\text{S}}^{2}}\int_{-\infty}^{t_{1}}\textrm{d}t'\ \int_{-\infty}^{t_{2}}\textrm{d}t^{\prime\prime}\ G_{12}^{*}\left(t_{1},t'\right)G_{12}(t_{2},t^{\prime\prime})\\ & \times\left\langle \left({-}i\gamma_{\text{I}}^{*}\tilde{\psi}_{\text{I<}}(0,t')-i\mu_{\text{I}}^{*}\tilde{\psi}_{\text{Iph<}}(0,t')\right)\left(i\gamma_{\text{I}}\tilde{\psi}_{\text{I<}}^{{\dagger}}(0,t^{\prime\prime})+i\mu_{\text{I}}\tilde{\psi}_{\text{Iph<}}^{{\dagger}}(0,t^{\prime\prime})\right)\right\rangle\\ = & 2\Gamma_{\text{S}}2\overline{\Gamma}_{\text{I}}\int_{-\infty}^{t_{1}}\textrm{d}t'\ \int_{-\infty}^{t_{2}}\textrm{d}t^{\prime\prime}\ G_{12}^{*}\left(t_{1},t'\right)G_{12}(t_{2},t^{\prime\prime})\delta(t'-t^{\prime\prime})\\ = & 4\Gamma_{\text{S}}\overline{\Gamma}_{\text{I}}\int \textrm{d}\tau\ G_{12}^{*}\left(t_{1},\tau\right)\theta(t_{1}-\tau)G_{12}(t_{2},\tau)\theta(t_{2}-\tau), \end{aligned}$$
$$\begin{aligned} M_{\text{SI}}(t_{1},t_{2})&= \sqrt{v_{\text{S}}v_{\text{I}}}\left\langle \tilde{\psi}_{\text{S>}}(0,t_{1})\tilde{\psi}_{\text{I>}}(0,t_{2})\right\rangle = -2\sqrt{\eta_{\text{S}}\eta_{\text{I}}\overline{\Gamma}_{\text{S}}\overline{\Gamma}_{\text{I}}}\left(\overline{\Gamma}_{\text{S}}+\overline{\Gamma}_{\text{I}}\right)\\&\quad\times\int \textrm{d}\tau\ \left[G_{21}^{*}(t_{1},\tau)G_{22}^{*}(t_{2},\tau)-G_{22}^{*}(t_{1},\tau)G_{21}^{*}(t_{2},\tau)\right]\theta(t_{2}-t_{1})\theta(t_{1}-\tau)\\ &\quad +2\overline{\Gamma}_{\text{S}}\int \textrm{d}\tau\ G_{11}(t_{1},\tau)\theta(t_{1}-\tau)G_{21}^{*}(t_{2},\tau)\theta(t_{2}-\tau). \end{aligned}$$
\begin{align} H_{\text{ring}}^{\text{DP-SFWM}} & =H_{\text{ring}}^{\text{L}}+H_{\text{ring}}^{\text{DP-SFWM},\text{NL}}\nonumber\\ &= \sum_{J\in\{\text{S,P1,P2}\}}\left(\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x)\psi_{J}(x)\textrm{d}x\right.\nonumber\\&\quad\left.-\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x)\frac{\partial\psi_{J}(x)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x)}{\partial x}\psi_{J}(x)\right)\textrm{d}x\right)\nonumber\\ & +\sum_{J\in\{\text{S,P1,P2}\}}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J}+\sum_{J\in\{\text{S,P1,P2}\}}(\hbar\gamma_{J}c_{J}^{{\dagger}}\psi_{J}(0)+\text{H.c.})\nonumber\\ & + \sum_{J\in\{\text{S,P1,P2}\}}\left(\hbar\omega_{J}\int\psi_{J\text{ph}}^{{\dagger}}(x)\psi_{J\text{ph}}(x)\textrm{d}x\right.\nonumber\\&\quad\left.-\frac{1}{2}i\hbar v_{J\text{ph}}\int\left(\psi_{J\text{ph}}^{{\dagger}}(x)\frac{\partial\psi_{J\text{ph}}(x)}{\partial x}-\frac{\partial\psi_{J\text{ph}}^{{\dagger}}(x)}{\partial x}\psi_{J\text{ph}}(x)\right)\textrm{d}x\right)\nonumber\\ & +\sum_{J\in\{\text{S,P1,P2}\}}(\hbar\gamma_{J\text{ph}}c_{J}^{{\dagger}}\psi_{J\text{ph}}(0)+\text{H.c.})\nonumber\\ & -\left(\hbar\Lambda_{\text{ring}}^{\text{DP-SFWM}}c_{\text{P1}}c_{\text{P2}}c_{\text{S}}^{{\dagger}}c_{\text{S}}^{{\dagger}}+\text{H.c.}\right)-\hbar\eta_{\text{ring}}^{\text{SPM}}(c_{\text{P1}}^{{\dagger}}c_{\text{P1}}^{{\dagger}}c_{\text{P1}}c_{\text{P1}}+c_{\text{P2}}^{{\dagger}}c_{\text{P2}}^{{\dagger}}c_{\text{P2}}c_{\text{P2}})\nonumber\\ & -\hbar\zeta_{\text{ring}}^{\text{XPM}}\left(c_{\text{S}}^{{\dagger}}c_{\text{P1}}^{{\dagger}}c_{\text{S}}c_{\text{P1}}+c_{\text{S}}^{{\dagger}} c_{\text{P2}}^{{\dagger}}c_{\text{S}}c_{\text{P2}}+c_{\text{P1}}^{{\dagger}}c_{\text{P2}}^{{\dagger}}c_{\text{P1}}c_{\text{P2}}\right), \end{align}
$$\begin{aligned} & \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{P1}}-2i\eta\overline{c}_{\text{P1}}^{{\dagger}}(t)\overline{c}_{\text{P1}}(t)-i\zeta\overline{c}_{\text{P2}}^{{\dagger}}(t)\overline{c}_{\text{P2}}(t)\right)\overline{c}_{\text{P1}}(t) \\ & \quad \quad =-i\gamma_{\text{P1}}^{*}\overline{\psi}_{\text{P1<}}(0,t)-i\gamma_{\text{P1ph}}^{*}\overline{\psi}_{\text{P1ph<}}(0,t)+i\Lambda^{*}\overline{c}_{\text{P2}}^{{\dagger}}(t)\overline{c}_{\text{S}}(t)\overline{c}_{\text{S}}(t)e^{i\Delta_{\text{ring}}t},\\ & \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{P2}}-2i\eta\overline{c}_{\text{P2}}^{{\dagger}}(t)\overline{c}_{\text{P2}}(t)-i\zeta\overline{c}_{\text{P1}}^{{\dagger}}(t)\overline{c}_{\text{P1}}(t)\right)\overline{c}_{\text{P2}}(t) \\ & \quad \quad=-i\gamma_{\text{P2}}^{*}\overline{\psi}_{\text{P2<}}(0,t)-i\gamma_{\text{P2ph}}^{*}\overline{\psi}_{\text{P2ph<}}(0,t)+i\Lambda^{*}\overline{c}_{\text{P1}}^{{\dagger}}(t)\overline{c}_{\text{S}}(t)\overline{c}_{\text{S}}(t)e^{i\Delta_{\text{ring}}t},\\ & \left(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{S}}-i\zeta\overline{c}_{\text{P1}}^{{\dagger}}(t)\overline{c}_{\text{P1}}(t)-i\zeta\overline{c}_{\text{P2}}^{{\dagger}}(t)\overline{c}_{\text{P2}}(t)\right)\overline{c}_{\text{S}}(t) \\ & \quad \quad=-i\gamma_{\text{S}}^{*}\overline{\psi}_{\text{S<}}(0,t)-i\gamma_{S\text{ph}}^{*}\overline{\psi}_{\text{Sph<}}(0,t)+2i\Lambda\overline{c}_{\text{P1}}(t)\overline{c}_{\text{P2}}(t)\overline{c}_{\text{S}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}, \end{aligned}$$
$$\overline{c}_{\text{P1(P2)}}(t)\rightarrow\left\langle \overline{c}_{\text{P1(P2)}}(t)\right\rangle e^{i\Delta_{\text{P1(P2)}}t}\equiv\tilde{\beta}_{\text{P1(P2)}}(t)e^{i\Delta_{\text{P1(P2)}}t},$$
$$\begin{aligned}\left(\frac{\textrm{d}}{\textrm{d}t}+i\Delta_{\text{P1}}+\overline{\Gamma}_{\text{P1}}-2i\eta\left|\tilde{\beta}_{\text{P1}}(t)\right|^{2}-i\zeta\left|\tilde{\beta}_{\text{P2}}(t)\right|^{2}\right)\tilde{\beta}_{\text{P1}}(t) & ={-}i\gamma_{\text{P1}}^{*}\tilde{\psi}_{\text{P1<}}(0,t),\\ \left(\frac{\textrm{d}}{\textrm{d}t}+i\Delta_{\text{P2}}+\overline{\Gamma}_{\text{P2}}-2i\eta\left|\tilde{\beta}_{\text{P2}}(t)\right|^{2}-i\zeta\left|\tilde{\beta}_{\text{P1}}(t)\right|^{2}\right)\overline{\beta}_{\text{P2}}(t) & ={-}i\gamma_{\text{P2}}^{*}\tilde{\psi}_{\text{P2<}}(0,t), \end{aligned}$$
$$\begin{aligned} & \left(\frac{\textrm{d}}{\textrm{d}t}+i\frac{\Delta_{\text{P1}}+\Delta_{\text{P2}}}{2}+\overline{\Gamma}_{\text{S}}-i\zeta\left|\tilde{\beta}_{\text{P1}}(t)\right|^{2}-i\zeta\left|\tilde{\beta}_{\text{P2}}(t)\right|^{2}\right)\tilde{c}_{\text{S}}(t) \\ & \quad \quad =-i\gamma_{\text{S}}^{*}\tilde{\psi}_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\tilde{\psi}_{\text{Sph<}}(0,t)+2i\Lambda\tilde{\beta}_{\text{P1}}(t)\tilde{\beta}_{\text{P2}}(t)\tilde{b}_{\text{S}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}, \end{aligned}$$
$$\frac{\textrm{d}}{\textrm{d}t}\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{S}}^{{\dagger}}(t) \end{pmatrix} = M_{\text{DP-SFWM}}(t)\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{S}}^{{\dagger}}(t) \end{pmatrix}+D_{\text{DP-SFWM}}(t),$$
$$\begin{aligned} & M_{\text{DP-SFWM}}(t)= -\overline{\Gamma}_{\text{S}}\mathbb{I} \\ &\quad + \begin{pmatrix} -i\frac{\Delta_{\text{P1}}+\Delta_{\text{P2}}}{2}+i\zeta\left[\left|\tilde{\beta}_{\text{P1}}(t)\right|^{2}+\left|\tilde{\beta}_{\text{P2}}(t)\right|^{2}\right] & 2i\Lambda\tilde{\beta}_{\text{P1}}(t)\tilde{\beta}_{\text{P2}}(t)e^{{-}i\Delta_{\text{ring}}t}\\ -2i\Lambda^{*}\tilde{\beta}_{\text{P1}}^{*}(t)\tilde{\beta}_{\text{P2}}^{*}(t)e^{i\Delta_{\text{ring}}t} & \begin{array}{c}\frac{\Delta_{\text{P1}}+\Delta_{\text{P2}}}{2}-i\zeta^{*}\left[\left|\tilde{\beta}_{\text{P1}}(t)\right|^{2}\right.\left.+\left|\tilde{\beta}_{\text{P2}}(t)\right|^{2}\right]\end{array}\end{pmatrix}, \end{aligned}$$
$$D_{\text{DP-SFWM}}(t)=\begin{pmatrix}-i\gamma_{\text{S}}^{*}\tilde{\psi}_{\text{S<}}(0,t)-i\gamma_{\text{Sph}}^{*}\tilde{\psi}_{\text{Sph<}}(0,t)\\ i\gamma_{\text{S}}\tilde{\psi}_{\text{S<}}^{{\dagger}}(0,t)+i\gamma_{\text{Sph}}\tilde{\psi}_{\text{Sph<}}^{{\dagger}}(0,t) \end{pmatrix}.$$
$$\begin{aligned}N_{\text{S}}(t_{1},t_{2}) & \equiv v_{\text{S}}\left\langle\tilde{\psi}_{\text{S>}}^{{\dagger}}(0,t_{1})\tilde{\psi}_{\text{S>}}(0,t_{2})\right\rangle, \\ M_{\text{SS}}(t_{1},t_{2}) & \equiv\sqrt{v_{\text{S}}v_{\text{S}}}\left\langle\tilde{\psi}_{\text{S>}}(0,t_{1})\tilde{\psi}_{\text{S>}}(0,t_{2})\right\rangle, \end{aligned}$$
$$\begin{aligned} N_{\text{S}}(t_{1},t_{2}) & =2\Gamma_{\text{S}}2\overline{\Gamma}_{\text{S}}\int_{-\infty}^{t_{1}}\textrm{d}t'\ \int_{-\infty}^{t_{2}}\textrm{d}t^{\prime\prime}\ G_{21}\left(t_{1},t'\right)G_{12}(t_{2},t^{\prime\prime})\delta(t'-t^{\prime\prime})\\ & =4\Gamma_{\text{S}}\overline{\Gamma}_{\text{S}}\int \textrm{d}\tau\ G_{21}\left(t_{1},\tau\right)\theta(t_{1}-\tau)G_{12}(t_{2},\tau)\theta(t_{2}-\tau), \end{aligned}$$
$$\begin{aligned} M_{\text{SS}}(t_{1},t_{2})&= v_{\text{S}}\left[\frac{\gamma_S^2}{v_S^2}\int_{-\infty}^{t_2}\textrm{d}t^{\prime\prime}\ G_{12}(t_2,t^{\prime\prime})\delta(t_1-t^{\prime\prime})\right. \\ & \quad -\left.\frac{\gamma_{\text{S}}^{2}}{v_{\text{S}}^{2}}\int_{-\infty}^{t_{1}}\textrm{d}t'\ \int_{-\infty}^{t_{2}}\textrm{d}t^{\prime\prime}\ G_{11}(t_{1},t')G_{12}(t_{2},t^{\prime\prime})\delta(t'-t^{\prime\prime})\left[\frac{|\gamma_{\text{S}}|^{2}}{v_{\text{S}}}+\frac{|\mu_{\text{S}}|^{2}}{v_{\text{S}}}\right]\right]\\ &= -4\frac{\gamma_{\text{S}}^{2}}{2v_{\text{S}}}\overline{\Gamma}_{\text{S}}\int \textrm{d}\tau\ \left[G_{12}(t_{1},\tau)G_{11}(t_{2},\tau)\theta(t_{1}-\tau)\theta(t_{2}-t_{1})\theta(t_{2}-\tau)\right.\\ &\quad +\left.G_{11}(t_{1},\tau)G_{12}(t_{2},\tau)\theta(t_{2}-\tau)\theta(t_{1}-t_{2})\theta(t_{1}-\tau)\right]. \end{aligned}$$
$$\begin{aligned}M_{\text{DP-SFWM}}(t) & ={-}\overline{\Gamma}_{\text{S}}\mathcal{\mathbb{I}}_{2}+B(t), \end{aligned}$$
$$\begin{aligned}B(t)= & \left[\begin{array}{lr} 0 & f(t)\\ f^{*}(t) & 0 \end{array}\right], \end{aligned}$$
$$\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{S}}^{{\dagger}}(t) \end{pmatrix}=G(t_{0},t)\begin{pmatrix}\tilde{c}_{\text{S}}(t)\\ \tilde{c}_{\text{S}}^{{\dagger}}(t) \end{pmatrix}+\int_{-\infty}^{t}\textrm{d}t^{\prime}\ \theta(t-t')G\left(t,t^{\prime}\right)\begin{pmatrix}-i\gamma_{\text{S}}^{*}\tilde{\psi}_{\text{S<}}(0,t^{\prime})\\ i\gamma_{\text{S}}\tilde{\psi}_{\text{S<}}^{{\dagger}}(0,t^{\prime}) \end{pmatrix}$$
$$\begin{aligned}G\left(t,t^{\prime}\right) & =e^{-\overline{\Gamma}_{\text{S}}\left(t-t^{\prime}\right)}S\left(t,t^{\prime}\right),\\ S\left(t,t^{\prime}\right) & =\mathcal{T}\exp\left(\int_{t^{\prime}}^{t}\textrm{d}t^{\prime\prime}B\left(t^{\prime\prime}\right)\right)\in SU(1,1), \end{aligned}$$
$$S\left(t,t^{\prime}\right)=\begin{pmatrix}\cosh\left(\mu\left[\theta(t)-\theta\left(t^{\prime}\right)\right]\right) & \sinh\left(\mu\left[\theta(t)-\theta\left(t^{\prime}\right)\right]\right)\\ \sinh\left(\mu\left[\theta(t)-\theta\left(t^{\prime}\right)\right]\right) & \cosh\left(\mu\left[\theta(t)-\theta\left(t^{\prime}\right)\right]\right) \end{pmatrix}.$$
$$\begin{aligned}\psi_{>}(t) = & \psi_{<}(t)-2\overline{\Gamma}_{\text{S}}\int_{-\infty}^{\infty}\textrm{d}t^{\prime}\theta\left(t-t^{\prime}\right)e^{-\overline{\Gamma}_{\text{S}}\left(t-t^{\prime}\right)}S_{1,1}\left(t,t^{\prime}\right)\psi_{\text{in }}\left(t^{\prime}\right)\\ & +2\overline{\Gamma}_{\text{S}}\int_{-\infty}^{\infty}\textrm{d}t^{\prime}\theta\left(t-t^{\prime}\right)e^{-\overline{\Gamma}_{\text{S}}\left(t-t^{\prime}\right)}S_{1,2}\left(t,t^{\prime}\right)\psi_{\text{in }}^{{\dagger}}\left(t^{\prime}\right),\\ = & \int \textrm{d}t^{\prime}\left[\delta\left(t-t^{\prime}\right)-\left(2\overline{\Gamma}_{\text{S}}\theta\left(t-t^{\prime}\right)e^{-\overline{\Gamma}_{\text{S}}\left(t-t^{\prime}\right)}+[\cosh\mu-1]f_{0}(t)f_{0}\left({-}t^{\prime}\right)\right)\right]\psi_{\text{in }}\left(t^{\prime}\right)\\ & +\int \textrm{d}t^{\prime}\sinh\mu f_{0}(t)f_{0}({-}t)\psi_{\text{in }}^{{\dagger}}\left(t^{\prime}\right), \end{aligned}$$
$$f_{0}(t)=\sqrt{2\overline{\Gamma}_{\text{S}}}e^{-\overline{\Gamma}_{\text{S}}t}\theta(t).$$
$$S(\Omega) = 1 + N_{\text{S}}(\Omega, \Omega) + N_{\text{S}}(-\Omega, -\Omega) + 2\Re\big[e^{{-}2i\psi_{\text{LOph}}} M_{\text{SS}}(\Omega, -\Omega)\big],$$
\begin{align} H_{\text{ring}}^{\text{SPDC}} &= H_{\text{ring}}^{\text{L}}+H_{\text{ring}}^{\text{SPDC},\text{NL}}\nonumber\\ &= \sum_{J\in\{\text{F1,F2,SH}\}}\left(\hbar\omega_{J}\int\psi_{J}^{{\dagger}}(x)\psi_{J}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J}\int\left(\psi_{J}^{{\dagger}}(x)\frac{\partial\psi_{J}(x)}{\partial x}-\frac{\partial\psi_{J}^{{\dagger}}(x)}{\partial x}\psi_{J}(x)\right)\textrm{d}x\right)\nonumber\\ &\quad +\sum_{J\in\{\text{F1,F2,SH}\}}\hbar\omega_{J}c_{J}^{{\dagger}}c_{J}+\sum_{J\in\{\text{F1,F2,SH}\}}(\hbar\gamma_{J}c_{J}^{{\dagger}}\psi_{J}(0)+\text{H.c.})\nonumber\\ &\quad +\sum_{J\in\{\text{F1,F2,SH}\}}\left(\hbar\omega_{J}\int\psi_{J\text{ph}}^{{\dagger}}(x)\psi_{J\text{ph}}(x)\textrm{d}x-\frac{1}{2}i\hbar v_{J\text{ph}}\right.\nonumber\\&\quad\times\left.\int\left(\psi_{J\text{ph}}^{{\dagger}}(x)\frac{\partial\psi_{J\text{ph}}(x)}{\partial x}-\frac{\partial\psi_{J\text{ph}}^{{\dagger}}(x)}{\partial x}\psi_{J\text{ph}}(x)\right)\textrm{d}x\right)+\sum_{J\in\{\text{F1,F2,SH}\}}(\hbar\gamma_{J\text{ph}}c_{J}^{{\dagger}}\psi_{J\text{ph}}(0)+\text{H.c.})\nonumber\\\ &\quad -\hbar\Lambda_{\text{ring}}^{\text{SPDC}}c_{\text{SH}}c_{\text{F1}}^{{\dagger}}c_{\text{F2}}^{{\dagger}}+\text{H.c.,} \end{align}
$$\begin{aligned}\Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{SH}}+i\omega_{\text{SH}}\Big)c_{\text{SH}}(t) & =i\Lambda^{*}c_{\text{F1}}(t)c_{\text{F2}}(t)-i\gamma_{\text{SH}}^{*}\psi_{SH<}\left(0,t\right)-i\gamma_{\text{SHph}}^{*}\psi_{SHph<}(0,t),\\ \Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{F1}}+i\omega_{\text{F1}}\Big)c_{\text{F1}}(t) & =i\Lambda c_{\text{SH}}(t)c_{\text{F2}}^{{\dagger}}(t)-i\gamma_{\text{F1}}^{*}\psi_{F1<}(0,t)-i\gamma_{\text{F1ph}}^{*}\psi_{F1ph<}(0,t),\\ \Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{F2}}+i\omega_{\text{F2}}\Big)c_{\text{F2}}(t) & =i\Lambda c_{\text{SH}}(t)c_{\text{F1}}^{{\dagger}}(t)-i\gamma_{\text{F2}}^{*}\psi_{F2<}(0,t)-i\gamma_{\text{F2ph}}^{*}\psi_{F2ph<}(0,t), \end{aligned}$$
$$\begin{aligned} \Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{SH}}\Big)\overline{c}_{\text{SH}}(t) & =i\Lambda^{*}\overline{c}_{\text{F1}}(t)\overline{c}_{\text{F2}}(t)e^{i\Delta_{\text{ring}}t}-i\gamma_{\text{SH}}^{*}\overline{\psi}_{SH<}\left(0,t\right)-i\gamma_{\text{SHph}}^{*}\overline{\psi}_{SHph<}(0,t),\\ \Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{F1}}\Big)\overline{c}_{\text{F1}}(t) & =i\Lambda\overline{c}_{\text{SH}}(t)\overline{c}_{\text{F2}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}-i\gamma_{\text{F1}}^{*}\overline{\psi}_{F1<}(0,t)-i\gamma_{\text{F1ph}}^{*}\overline{\psi}_{F1ph<}(0,t),\\ \Big(\frac{\textrm{d}}{\textrm{d}t}+\overline{\Gamma}_{\text{F2}}\Big)\overline{c}_{\text{F2}}(t) & =i\Lambda\overline{c}_{\text{SH}}(t)\overline{c}_{\text{F1}}^{{\dagger}}(t)e^{{-}i\Delta_{\text{ring}}t}-i\gamma_{\text{F2}}^{*}\overline{\psi}_{F2<}(0,t)-i\gamma_{\text{F2ph}}^{*}\overline{\psi}_{F2ph<}(0,t), \end{aligned}$$
$$\begin{aligned}\overline{c}_{\text{SH}}(t) & \rightarrow\left\langle \overline{c}_{\text{SH}}(t)\right\rangle e^{i\Delta_{\text{SH}}t}\equiv\beta_{\text{SH}}(t)e^{i\Delta_{\text{SH}}t},\\ \overline{\psi}_{\text{SH}<}(0,t) & \rightarrow\left\langle \overline{\psi}_{\text{SH}<}(0,t)\right\rangle e^{i\Delta_{\text{SH}}t}. \end{aligned}$$
$$\frac{\textrm{d}}{\textrm{d}t}\beta_{\text{SH}}(t)=\left(-\overline{\Gamma}_{\text{SH}}-i\Delta_{\text{SH}}\right)\beta_{\text{SH}}(t)-i\gamma_{\text{SH}}^{*}\left\langle \overline{\psi}_{SH<}(0,t)\right\rangle.$$
$$\frac{\textrm{d}}{\textrm{d}t}\begin{pmatrix}\tilde{c}_{\text{F1}}(t)\\ \tilde{c}_{\text{F2}}^{{\dagger}}(t) \end{pmatrix}=M_{\text{SPDC}}(t)\begin{pmatrix}\tilde{c}_{\text{F1}}(t)\\ \tilde{c}_{\text{F2}}^{{\dagger}}(t) \end{pmatrix}+D_{\text{SPDC}}(t),$$
$$\begin{aligned}M_{\text{SPDC}}(t) & =\begin{pmatrix}-\overline{\Gamma}_{\text{F1}}-i\frac{\Delta_{\text{SH}}-\Delta_{\text{ring}}}{2} & i\Lambda\beta_{\text{SH}}(t)\\ -i\Lambda^{*}\beta_{\text{SH}}^{*}(t) & -\overline{\Gamma}_{\text{F2}}+i\frac{\Delta_{\text{SH}}-\Delta_{\text{ring}}}{2} \end{pmatrix}, \end{aligned}$$
$$\begin{aligned}D_{\text{SPDC}}(t) & =\begin{pmatrix}-i\gamma_{\text{F1}}^{*}\tilde{\psi}_{\text{F1}<}(0,t)-i\gamma_{\text{F1}ph}^{*}\tilde{\psi}_{\text{F1ph}<}(0,t)\\ i\gamma_{\text{F2}}\tilde{\psi}_{\text{F2}<}^{{\dagger}}(0,t)+i\gamma_{\text{F2}ph}\tilde{\psi}_{\text{F2ph}<}^{{\dagger}}(0,t) \end{pmatrix}, \end{aligned}$$
$$\Lambda_{\text{ring}}^{\text{SPDC}}\rightarrow2\Lambda_{\text{ring}}^{\text{SPDC}}.$$
$$R_\textrm{SPDC,wg}=\frac{P_P}{\mathcal{PA}}\frac{L^{3/2}}{\frac{3}{2}\sqrt{2\pi\left|\beta_2(\omega_P/2)\right|}}$$
$$R_\textrm{DFG,wg}=\frac{P_P}{\mathcal{PA}}\frac{L^2}{\frac{1}{2}\hbar\omega_P}P_S.$$
$$\frac{R_\textrm{SPDC,wg}}{R_\textrm{DFG,wg}}=\frac{\hbar\omega_P}{3\sqrt{2\pi\left|\beta_2(\omega_P/2)\right|L}}\frac{1}{P_S},$$
$$P_\textrm{vac,wg}= \frac{\hbar\omega_P}{3\sqrt{2\pi\left|\beta_2(\omega_P/2)\right|L}},$$
$$R_\textrm{SPDC,ring}=\frac{P_p}{\mathcal{PA}}\frac{4v_F^2v_P}{\pi\omega_p\omega_F}\frac{Q_PQ_F}{R},$$
$$R_\textrm{DFG,ring}=\frac{P_p}{\mathcal{PA}}\frac{4v_F^2v_P}{\hbar\pi\omega_p\omega_F^3}\frac{Q_PQ_F^2}{R},$$
$$\frac{R_\textrm{SPDC,ring}}{R_\textrm{DFG,ring}}=\frac{\hbar\omega_F^2}{8Q_F}\frac{1}{P_S},$$
$$P_\textrm{vac,ring}= \frac{\hbar\omega_F^2}{8Q_F}$$
$$\beta_{\sigma,\textbf{k}_2}\approx\sqrt{2}\gamma\sum_{\sigma^{\prime}}\int\textrm{d}\textbf{k}_1\phi_{\sigma^{\prime},\sigma}(\textbf{k}_1,\textbf{k}_2)\beta_{\sigma^{\prime},\textbf{k}_1}^{*},$$
$$|{\Psi}\rangle = \sum_{i_0=0}^{c-1} \cdots \sum_{i_{\ell-1}=0}^{c-1} R_{i_0\ldots i_{\ell-1}} |{i_0,\ldots,i_{\ell-1}}\rangle,$$
$$R_{i_0,\ldots,i_{\ell-1}} =\sum_{\alpha_1=0}^{D_1-1} \ldots \sum_{\alpha_{\ell-1} =0}^{D_{\ell-1} - 1} (S^{[0]})_{i_0,\alpha_1} (S^{[1]})_{i_1,\alpha_2}^{\alpha_1} \ldots (S^{[\ell-2]})_{i_{\ell-2},\alpha_{\ell-1}}^{\alpha_{\ell-2}} (S^{[\ell-1]})_{i_{\ell-1}}^{\alpha_{\ell-1}},$$
$$\tilde H = \tilde H_{\text{res}} + \tilde H_{\text{wg}} + \tilde V,$$
$$\begin{aligned}\tilde H_{\text{res}} & = \hbar \omega_0 a^\dagger a + \tilde H_{\text{NL}}(a^\dagger,a,t), \end{aligned}$$
$$\begin{aligned}\tilde H_{\text{wg}} & = \hbar \omega_1 \int \textrm{d}x \ \psi^\dagger(x) \psi(x) + \frac{i}{2} \hbar v \int \textrm{d}x \left( \left[\frac{\partial}{\partial x} \psi^\dagger(x) \right] \psi(x) - \text{H.c.} \right), \end{aligned}$$
$$\begin{aligned}\tilde V & = \hbar \gamma \left( \psi^\dagger(0)a + \text{H.c.} \right), \end{aligned}$$
$$[\psi(x),\psi^\dagger(x')]=\delta(x-x') \quad \text{and} \quad [a,a^\dagger] = 1.$$
$$\tilde{H}_{\rm NL}(a^\dagger, a, t) = \mu \beta(t) e^{i \omega_c t} a^3 + \mu^* \beta^* e^{{-}i \omega_c t} a^{{\dagger} 3},$$
$$\mathcal{ U}_1 = \exp\left(- \frac{i}{\hbar} H_1 (t-t_0) \right),\\\ H_1 = \hbar \omega_1\left[a^\dagger a + \int \textrm{d}x \ \psi^\dagger(x) \psi(x) \right],$$
$$\tilde H = \tilde H_{\text{res}} + \tilde H_{\text{wg}} + \tilde V,$$
$$\begin{aligned}\bar H_{\text{res}} & = \hbar \delta a^\dagger a + H_{\text{NL}}(t), \quad \delta = \omega_0 - \omega_1, \end{aligned}$$
$$\begin{aligned}\bar H_{\text{wg}} & = \frac{i}{2} \hbar v \int \textrm{d}x \left( \left[ \frac{\partial}{\partial x} \psi^\dagger(x) \right] \psi(x) - \text{H.c.} \right), \end{aligned}$$
$$\begin{aligned}H_{\text{NL}}(t) & = \tilde{H}_{\text{NL}}({e^{i \omega_1 t}a^\dagger},e^{{-}i \omega_1 t} a,t), \end{aligned}$$
$$\begin{aligned}\bar V & = \tilde V. \end{aligned}$$
$$\begin{aligned}\mathcal{U}_2 & = \exp\left( - \frac{i}{\hbar} (t-t_0) \bar{H}_{\text{wg}} \right) \\ & = \exp\left( (t-t_0) \frac{v}{2} \int \textrm{d}x \left\{ \left[ \frac{\partial}{\partial x} \psi^\dagger(x) \right] \psi(x) - \text{H.c.} \right\} \right), \end{aligned}$$
$$H =\hbar \delta a^\dagger a + H_{\text{NL}}(t) + \hbar \gamma \left( \psi^\dagger({-}v (t-t_0))a + \text{H.c.} \right).$$
$$\hbar \gamma \left( \psi^\dagger({-}v (t-t_0))a + \text{H.c.} \right)= \hbar \gamma \left[ a \int \textrm{d}x \ \psi^\dagger (x)\delta(x+v(t-t_0)) +\text{H.c.} \right],$$
$$\begin{aligned}\psi_j & = \frac{1}{\sqrt{\Delta x}} \int_{-(j+1/2)\Delta x}^{-(j-1/2)\Delta x} \textrm{d}x \ \psi(x)& \approx \sqrt{\Delta x} \ \psi(- j \Delta x) = \sqrt{\Delta x} \ \psi(- j v \Delta t). \end{aligned}$$
$$H_j = \hbar \delta a^\dagger a + H_{\text{NL}}(t_0+j\Delta t) +i \hbar \sqrt{\frac{\gamma v}{\Delta x }} \left( \psi_j^\dagger a -\text{H.c.} \right).$$
$$\begin{aligned} \mathcal{ U} & = \left[ \mathcal{B}_{0,\ell-1} \mathcal{V}_0(\ell-1) \right] \ldots \left[ \mathcal{B}_{0,0} \mathcal{V}_0(0) \right], \end{aligned}$$
$$\begin{aligned}\mathcal{V}_0(j) & = \exp\left(-\frac{i}{\hbar} \Delta t \left[ \hbar \delta a^\dagger a + H_{\text{NL}}(t_0+j \Delta t ) \right] \right), \end{aligned}$$
$$\begin{aligned}\mathcal{B}_{0,j} & = \exp\left( i \gamma \sqrt{ \frac{\Delta t} {v}} \left( \psi^\dagger_j a + \text{H.c.} \right) \right), \end{aligned}$$
$$\mathcal{W}(i+1,i) = \text{SWAP}_{i+1,i} \mathcal{B}_{i+1,i} \mathcal{V}_i(i),$$
$$\mathcal{U} = \mathcal{W}(\ell, \ell-1) \ldots \mathcal{W}(1,0),$$
$$| \Psi (t_0) \rangle = \bigotimes_{i=0}^\ell | 0_i \rangle,$$
$$R_{i_0,i_1,\ldots,r_\ell} = \langle i_0,i_1,\ldots,i_\ell|\mathcal{U}|\Psi \rangle,$$
$$\begin{aligned}R_{i_0, i_1, i_2, i_3, i_4} & = \langle i_0, i_1, i_2, i_3, i_4|\mathcal{ U}|0_0,0_1,0_2,0_3,0_4 \rangle \\ & = \langle i_0, i_1, i_2, i_3, i_4|W(4,3)W(3,2)W(2,1)W(1,0)|0_0,0_1,0_2,0_3,0_4 \rangle. \end{aligned}$$
$$\langle i_0|W(0,1)|0_0,0_1 \rangle=\sum_{\alpha_1=0}^{c-1} (T^{[0]})^{i_0,\alpha_1}_{0_0,0_1} | \alpha_1 \rangle \leftrightarrow \langle i_0, \alpha_1|W(1,0)|0_0,0_1 \rangle= (T^{[0]})^{i_0,\alpha_1}_{0_0,0_1},$$
$$R_{i_0, i_1, i_2, i_3, i_4} = \sum_{\alpha_1=0}^{c-1} (T^{[0]})^{i_0,\alpha_1}_{0_0,0_1} \langle i_1, i_2, i_3, i_4|W(4,3)W(3,2)W(2,1)|\alpha_1,0_2,0_3,0_4 \rangle.$$
$$\langle i_1|W(2,1)|\alpha_1,0_2 \rangle=\sum_{\alpha_2=0}^{c-1} (T^{[1]})^{i_1,\alpha_2}_{\alpha_1,0_2} | \alpha_2 \rangle,$$
$$R_{i_0, i_1, i_2, i_3, i_4} = \sum_{\alpha_1,\alpha_2=0}^{c-1} (T^{[0]})^{i_0,\alpha_1}_{0_0,0_1} (T^{[1]})^{i_1,\alpha_2}_{\alpha_1,0_2} \langle i_2, i_3, i_4|W(4,3)W(3,2)|\alpha_2,0_3,0_4 \rangle.$$
$$\begin{aligned}\langle i_2|W(3,2)|\alpha_2,0_3 \rangle & =\sum_{\alpha_3=0}^{c-1} (T^{[2]})^{i_2,\alpha_3}_{\alpha_2,0_3} | \alpha_3 \rangle, \end{aligned}$$
$$\begin{aligned}\langle i_3|W(4,3)|\alpha_3,0_4 \rangle & =\sum_{\alpha_4=0}^{c-1} (T^{[3]})^{i_3,\alpha_4}_{\alpha_3,0_4} | \alpha_4 \rangle, \end{aligned}$$
$$R_{i_0, i_1, i_2, i_3, i_4} = \sum_{\alpha_1,\alpha_2,\alpha_3,\alpha_4=0}^{c-1} (T^{[0]})^{i_0,\alpha_1}_{0_0,0_1} (T^{[1]})^{i_1,\alpha_2}_{\alpha_1,0_2} (T^{[2]})^{i_2,\alpha_3}_{\alpha_2,0_3} (T^{[3]})^{i_3,\alpha_4}_{\alpha_3,0_4} \langle i_4|\alpha_4 \rangle.$$
$$R_{i_0, i_1, i_2, i_3, i_4} = \sum_{\alpha_1,\alpha_2,\alpha_3,\alpha_4=0}^{c-1} (S^{[0]})_{i_0,\alpha_1} (S^{[1]})_{i_1,\alpha_2}^{\alpha_1} (S^{[2]})_{i_2,\alpha_3}^{\alpha_2} (S^{[3]})_{i_3,i_4}^{\alpha_3} (S^{[4]})^{\alpha_4}_{i_4},$$
$$\begin{aligned} & \boldsymbol{\nabla}\times\left(\boldsymbol{\nabla}\times\boldsymbol{B}_{\alpha}(\boldsymbol{r})\right)=\frac{\omega_{\alpha}^{2}\varepsilon_{1}}{c^{2}}\boldsymbol{B}_{\alpha}(\boldsymbol{r}),\\ & \boldsymbol{\nabla}\cdot\boldsymbol{B}_{\alpha}(\boldsymbol{r})=0. \end{aligned}$$
$$ \omega_{\alpha}=c\left|\boldsymbol{k}\right|/n, $$
$$\boldsymbol{k}=\frac{2\pi}{L}(m_{x}\boldsymbol{\hat{x}}+m_{y}\boldsymbol{\hat{y}}+m_{z}\boldsymbol{\hat{z}}), $$
$$\boldsymbol{\hat{e}}_{1\boldsymbol{k}}\times\boldsymbol{\hat{e}}_{2\boldsymbol{k}}=\boldsymbol{\hat{k}}. $$
$$\begin{aligned} & \boldsymbol{\hat{e}}_{L\boldsymbol{k}}={-}\frac{1}{\sqrt{2}}\left(\boldsymbol{\hat{e}}_{1\boldsymbol{k}}+i\boldsymbol{\hat{e}}_{2\boldsymbol{k}}\right),\\ & \boldsymbol{\hat{e}}_{R\boldsymbol{k}}=\frac{1}{\sqrt{2}}\left(\boldsymbol{\hat{e}}_{1\boldsymbol{k}}-i\boldsymbol{\hat{e}}_{2\boldsymbol{k}}\right), \end{aligned}$$
$$\begin{aligned} & \boldsymbol{\hat{e}}_{I\boldsymbol{k}}\cdot\boldsymbol{\hat{e}}_{I\boldsymbol{k}}^{*}=1,\\ & \boldsymbol{\hat{e}}_{I\boldsymbol{k}}\cdot\boldsymbol{\hat{e}}_{\overline{I}\boldsymbol{k}}^{*}=0, \end{aligned}$$
$$\begin{aligned} & i\boldsymbol{\hat{k}}\times\boldsymbol{\hat{e}}_{I\boldsymbol{k}}=s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}},\\ & \hat{\boldsymbol{e}}_{I\boldsymbol{k}}\times\boldsymbol{\hat{e}}_{I\boldsymbol{k}}^{*}={-}is_{I}\boldsymbol{\hat{k}}, \end{aligned}$$
$$\begin{aligned} & \boldsymbol{\hat{e}}_{1(-\boldsymbol{k})}=\boldsymbol{\hat{e}}_{1\boldsymbol{k}},\\ & \boldsymbol{\hat{e}}_{2(-\boldsymbol{k})}={-}\boldsymbol{\hat{e}}_{2\boldsymbol{k}}, \end{aligned}$$
$$\begin{aligned} & \boldsymbol{\hat{e}}_{I\boldsymbol{k}}^{*}=\boldsymbol{\hat{e}}_{I(-\boldsymbol{k})},\\ & \boldsymbol{\hat{e}}_{I\boldsymbol{k}}\cdot\boldsymbol{\hat{e}}_{I'(-\boldsymbol{k})}=\delta_{II'}. \end{aligned}$$
$$\boldsymbol{B}_{I\boldsymbol{k}}(\boldsymbol{r})=\sqrt{\frac{\mu_{0}}{V}}s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}},$$
$$\boldsymbol{D}_{I\boldsymbol{k}}(\boldsymbol{r})=in\sqrt{\frac{\epsilon_{0}}{V}}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}$$
$$\omega_{I\boldsymbol{k}}=\frac{c}{n}\left|\boldsymbol{k}\right|, $$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=i\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar cn\left|\boldsymbol{k}\right|\epsilon_{0}}{2V}}a_{I\boldsymbol{k}}(t)\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar c\left|\boldsymbol{k}\right|\mu_{0}}{2nV}}a_{I\boldsymbol{k}}(t)s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.} \end{aligned}$$
$$\sum_{\boldsymbol{k}}\rightarrow\frac{\int \textrm{d}\boldsymbol{k}}{\left(\frac{2\pi}{L}\right)^{3}}=\frac{V}{8\pi^{3}}\int {\rm d}\boldsymbol{k}$$
$$a_{I\boldsymbol{k}}(t)\rightarrow\sqrt{\frac{8\pi^{3}}{V}}a_{I}(\boldsymbol{k},t).$$
$$\begin{aligned} & \left[a_{I}(\boldsymbol{k},t),a_{I'}^{{\dagger}}(\boldsymbol{k'},t)\right]=\delta_{II'}\delta(\boldsymbol{k}-\boldsymbol{k'}),\\ & \left[a_{I}(\boldsymbol{k},t),a_{I'}(\boldsymbol{k'},t)\right]=0. \end{aligned}$$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=i\sum_{I}\int\sqrt{\frac{\hbar cn\left|\boldsymbol{k}\right|\epsilon_{0}}{16\pi^{3}}}a_{I}(\boldsymbol{k},t)\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}\textrm{d}\boldsymbol{k}+\text{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I}\int\sqrt{\frac{\hbar c\left|\boldsymbol{k}\right|\mu_{0}}{16\pi^{3}n}}a_{I}(\boldsymbol{k},t)s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}\textrm{d}\boldsymbol{k}+\text{H.c.}, \end{aligned}$$
$$H=\sum_{I}\int\hbar\omega_{I\boldsymbol{k}}a_{I}(\boldsymbol{k},t)a_{I}(\boldsymbol{k,}t)\textrm{d}\boldsymbol{k}.$$
$$\boldsymbol{B}_{I\boldsymbol{k}}(\boldsymbol{r})=\sqrt{\frac{\mu_{0}v_{g}(\omega_{I\boldsymbol{k}})}{Vv_{p}(\omega_{I\boldsymbol{k}})}}s_{I} \boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}},$$
$$\boldsymbol{D}_{I\boldsymbol{k}}(\boldsymbol{r})=in(\omega_{I\boldsymbol{k}})\sqrt{\frac{\epsilon_{0}v_{g}(\omega_{I\boldsymbol{k}})}{Vv_{p} (\omega_{I\boldsymbol{k}})}}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}},$$
$$n(\omega_{I\boldsymbol{k}})\omega_{I\boldsymbol{k}}=c\left|\boldsymbol{k}\right|$$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=i\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|\epsilon_{0}\varepsilon_{1} (\omega_{I\boldsymbol{k}})v_{g}(\omega_{I\boldsymbol{k}})}{2V}}a_{I\boldsymbol{k}}(t)\boldsymbol{\hat{e}}_{I\boldsymbol{k}} e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|\mu_{0}v_{g}(\omega_{I\boldsymbol{k}})}{2V}} a_{I\boldsymbol{k}}(t)s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.} \end{aligned}$$
$$\begin{aligned} & \boldsymbol{E}(\boldsymbol{r},t)=i\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|v_{g}(\omega_{I\boldsymbol{k}})}{ 2V\epsilon_{0}\varepsilon_{1}(\omega_{I\left|\boldsymbol{k}\right|})}}a_{I\boldsymbol{k}}(t)\boldsymbol{\hat{e}}_{I\boldsymbol{k}} e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.},\\ & \boldsymbol{H}(\boldsymbol{r},t)=\sum_{I,\boldsymbol{k}}\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|v_{g}(\omega_{I\boldsymbol{k}})}{2V\mu_{0}}} a_{I\boldsymbol{k}}(t)s_{I}\boldsymbol{\hat{e}}_{I\boldsymbol{k}}e^{i\boldsymbol{k}\cdot\boldsymbol{r}}+\text{H.c.}, \end{aligned}$$
$$\left|\Psi\right\rangle =a_{I\boldsymbol{k}}^{{\dagger}}(0)\left\vert\text{vac}\right\rangle$$
$$\int_{V}\left\langle \boldsymbol{S}(\boldsymbol{r},t)\right\rangle d\boldsymbol{r}=\hbar\omega_{I\boldsymbol{k}}v_{g}(\omega_{I\boldsymbol{k}}), $$
$$\begin{aligned} & \boldsymbol{D}(\boldsymbol{r},t)=i\sum_{I}\int\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|\epsilon_{0} \varepsilon_{1}(\omega_{I\boldsymbol{k}})v_{g}(\omega_{I\boldsymbol{k}})}{16\pi^{3}}}a_{I}(\boldsymbol{k},t) \boldsymbol{\hat{e}}_{I}(\boldsymbol{k})e^{i\boldsymbol{k}\cdot\boldsymbol{r}}d\boldsymbol{k}+\text{H.c.},\\ & \boldsymbol{B}(\boldsymbol{r},t)=\sum_{I}\int\sqrt{\frac{\hbar\left|\boldsymbol{k}\right|\mu_{0}v_{g}(\omega_{I\boldsymbol{k}})}{16\pi^{3}}} a_{I}(\boldsymbol{k},t)s_{I}\boldsymbol{\hat{e}}_{I}(\boldsymbol{k})e^{i\boldsymbol{k}\cdot\boldsymbol{r}}d\boldsymbol{k}+\text{H.c.}, \end{aligned}$$
$$\begin{aligned} & \mathcal{B}_{Ik}(\boldsymbol{r})=\boldsymbol{b}_{Ik}(y,z)e^{ikx},\\ & \mathcal{D}_{Ik}(\boldsymbol{r})=\boldsymbol{d}_{Ik}(y,z)e^{ikx}, \end{aligned}$$
$$\boldsymbol{\nabla}\times\left(\frac{\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right) =\frac{\omega_{Ik}^{2}}{c^{2}}\mathcal{B}_{Ik}(\boldsymbol{r}),$$
$$\begin{aligned} & \frac{\partial\mathcal{B}_{Ik}(\boldsymbol{r})}{\partial k}=\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}+ix\mathcal{B}_{Ik}(\boldsymbol{r}),\\ & \frac{\partial}{\partial k}\left(\frac{1}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)=\frac{\partial\omega_{Ik}}{\partial k}\frac{\partial}{\partial\omega_{Ik}}\left(\frac{1}{\varepsilon_{1}(y,z;\omega_{Ik})}\right) \end{aligned}$$
$$ v_{Ik}\equiv\frac{\partial\omega_{Ik}}{\partial k}, $$
$$\begin{aligned} & \boldsymbol{\nabla}\times\left(\boldsymbol{C}_{1}(\boldsymbol{r})+\boldsymbol{C}_{2}(\boldsymbol{r})+v_{Ik}\boldsymbol{C}_{3}(\boldsymbol{r)}\right)\\ & =\frac{2\omega_{Ik}}{c^{2}}v_{Ik}\mathcal{B}_{Ik}(\boldsymbol{r})+\frac{\omega_{Ik}^{2}}{c^{2}}\left(\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}+ix\mathcal{B}_{Ik}(\boldsymbol{r})\right),\end{aligned}$$
$$\begin{aligned} & \boldsymbol{C}_{1}(\boldsymbol{r})=\frac{\boldsymbol{\nabla}\times\left(\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}\right)}{\varepsilon_{1}(y,z;\omega_{Ik})},\\ & \boldsymbol{C}_{2}(\boldsymbol{r})=\frac{\boldsymbol{\nabla}\times\left(ix\mathcal{B}_{Ik}(\boldsymbol{r})\right)}{\varepsilon_{1} (y,z;\omega_{Ik})}=\frac{i\boldsymbol{\hat{x}}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})} +\frac{ix\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})},\\ & \boldsymbol{C}_{3}(\boldsymbol{r})=\frac{\partial}{\partial\omega_{Ik}}\left(\frac{1}{\varepsilon_{1}(y,z;\omega_{Ik})}\right) \boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r}). \end{aligned}$$
$$\boldsymbol{\nabla}\times\boldsymbol{C}_{2}(\boldsymbol{r})=\boldsymbol{\nabla}\times\boldsymbol{C}_{4}(\boldsymbol{r}) +ix\boldsymbol{\nabla}\times\left(\frac{\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right) +\boldsymbol{C}_{6}(\boldsymbol{r}),$$
$$\begin{aligned} & \boldsymbol{C}_{4}(\boldsymbol{r})=\frac{i\boldsymbol{\hat{x}}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})},\\ & \boldsymbol{C}_{6}(\boldsymbol{r})=i\hat{\boldsymbol{x}}\times\left(\frac{\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1} (y,z;\omega_{Ik})}\right). \end{aligned}$$
$$\begin{aligned} & \left(\mathcal{\overline{M}}-\frac{\omega_{Ik}^{2}}{c^{2}}\right)\left(\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}\right)+\boldsymbol{\nabla}\times\boldsymbol{C}_{4}(\boldsymbol{r})+\boldsymbol{C}_{6}(\boldsymbol{r})\\ & =\frac{2\omega_{Ik}}{c^{2}}v_{Ik}\mathcal{B}_{Ik}(\boldsymbol{r})-v_{Ik}\boldsymbol{\nabla}\times\boldsymbol{C}_{3}(\boldsymbol{r}), \end{aligned}$$
$$\mathcal{\overline{M}}\boldsymbol{g}(\boldsymbol{r})\equiv\boldsymbol{\nabla}\times\left(\frac{\boldsymbol{\nabla}\times\boldsymbol{g} (\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)$$
$$\begin{aligned} & \int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\left(\mathcal{\overline{M}}-\frac{\omega_{Ik}^{2}}{c^{2}}\right)\left(\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}\right)\\ & =\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\left(\mathcal{\overline{M}}-\frac{\omega_{Ik}^{2}}{c^{2}}\right)\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\right)\cdot\left(\frac{\partial\boldsymbol{b}_{Ik}(y,z)}{\partial k}e^{ikx}\right)\\ & =0, \end{aligned}$$
$$\begin{aligned} & \int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\boldsymbol{\nabla}\times\boldsymbol{C}_{4}(\boldsymbol{r})\\ & =i\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\boldsymbol{\nabla}\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\right).\left(\frac{\boldsymbol{\hat{x}}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)\\ & ={-}\mu_{0}\omega_{Ik}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{D}_{Ik}^{*}(\boldsymbol{r}).\left(\frac{\boldsymbol{\hat{x}}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)\\ & =\mu_{0}\omega_{Ik}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\hat{x}}\cdot\left(\frac{\mathcal{D}_{Ik}^{*}(\boldsymbol{r})\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right). \end{aligned}$$
$$\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{U}(\boldsymbol{r})\cdot\left(\boldsymbol{\nabla}\times\boldsymbol{V}(\boldsymbol{r})\right)=\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{V}(\boldsymbol{r})\cdot\left(\boldsymbol{\nabla}\times\boldsymbol{U}(\boldsymbol{r})\right),$$
$$\mathcal{D}_{Ik}(\boldsymbol{r})=\frac{i}{\mu_{0}\omega_{Ik}}\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r)},$$
$$\boldsymbol{u}\cdot\left(\boldsymbol{v}\times\boldsymbol{w}\right)=\boldsymbol{v}\cdot\left(\boldsymbol{w}\times\boldsymbol{u}\right),$$
$$\begin{aligned} & \int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\boldsymbol{C}_{6}(\boldsymbol{r})\\ & =i\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\hat{x}}\cdot\left(\left(\frac{\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right) \times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\right)\\ & =\mu_{0}\omega_{Ik}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\hat{x}}\cdot\left(\frac{\mathcal{D}_{Ik}(\boldsymbol{r})\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right), \end{aligned}$$
$$\begin{aligned} & \mu_{0}\omega_{Ik}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\hat{x}}\cdot\left(\frac{\mathcal{D}_{Ik}(\boldsymbol{r})\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})+\mathcal{D}_{Ik}^{*} (\boldsymbol{r})\times\mathcal{B}_{Ik}(\boldsymbol{r})}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)\\ & =v_{Ik}\mathcal{T}, \end{aligned}$$
$$\begin{aligned} & \mathcal{T}=\frac{2\omega_{Ik}}{c^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{B}_{Ik}(\boldsymbol{r})\\ & -\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\nabla\times\boldsymbol{C}_{3}(\boldsymbol{r}). \end{aligned}$$
$$\mathcal{B}_{Ik}(\boldsymbol{r})=\frac{1}{i\omega_{Ik}}\boldsymbol{\nabla}\times\left(\frac{\mathcal{D}_{Ik}(\boldsymbol{r})}{\epsilon_{0} \varepsilon_{1}(y,z;\omega_{Ik})}\right),$$
$$\begin{aligned} & \frac{2\omega_{Ik}}{c^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{B}_{Ik}(\boldsymbol{r})\\ & =\frac{2}{ic^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\left(\boldsymbol{\nabla}\times\left(\frac{\mathcal{D}_{Ik} (\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\right)\right)\\ & =\frac{2}{ic^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\boldsymbol{\nabla}\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\right)\cdot\left(\frac{\mathcal{D}_{Ik} (\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\right)\\ & =\frac{2\mu_{0}\omega_{Ik}}{c^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\frac{\mathcal{D}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{D}_{Ik}(\boldsymbol{r})}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\right), \end{aligned}$$
$$\begin{aligned} & -\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\nabla\times\boldsymbol{C}_{3}(\boldsymbol{r})\\ & ={-}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\nabla\times\left(\frac{\partial}{\partial\omega_{Ik}}\left(\frac{1}{\varepsilon_{1} (y,z;\omega_{Ik})}\right)\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})\right)\\ & ={-}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\nabla}\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\left(\frac{\partial}{\partial\omega_{Ik}} \left(\frac{1}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})\right)\\ & ={-}\mu_{0}^{2}\omega_{Ik}^{2}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\mathcal{D}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{D}_{Ik}(\boldsymbol{r})\right)\frac{\partial}{\partial\omega_{Ik}} \left(\frac{1}{\varepsilon_{1}(y,z;\omega_{Ik})}\right)\\ & ={-}\frac{2\mu_{0}\omega_{Ik}}{c^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\mathcal{D}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{D}_{Ik}(\boldsymbol{r})\right)\frac{\omega_{Ik}}{2} \frac{\partial}{\partial\omega_{Ik}}\left(\frac{1}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\right) \end{aligned}$$
$$\begin{aligned} & -\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\nabla\times\boldsymbol{C}_{3}(\boldsymbol{r})\\ &\quad ={-}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\boldsymbol{\nabla}\times\mathcal{B}_{Ik}^{*}(\boldsymbol{r})\cdot\left(\frac{\partial}{\partial\omega_{Ik}}\left(\frac{1}{\varepsilon_{1} (y,z;\omega_{Ik})}\right)\boldsymbol{\nabla}\times\mathcal{B}_{Ik}(\boldsymbol{r})\right)\\ & ={-}\frac{2\mu_{0}\omega_{Ik}}{c^{2}}\int_{{-}L/2}^{L/2}\textrm{d}x\int \textrm{d}y\textrm{d}z\left(\mathcal{D}_{Ik}^{*}(\boldsymbol{r})\cdot\mathcal{D}_{Ik}(\boldsymbol{r})\right)\frac{\omega_{Ik}}{2}\frac{\partial}{\partial\omega_{Ik}} \left(\frac{1}{\epsilon_{0}\varepsilon_{1}(y,z;\omega_{Ik})}\right), \end{aligned}$$
$$\mathcal{T}=\frac{2\mu_{0}\omega_{Ik}}{c^{2}}L,$$
$$H(t)= \underbrace{ \hbar \sum_{i,j =1}^\ell \Delta_{i,j} a_i^\dagger a_j}_{{\equiv} H_0} + \underbrace{\hbar i \frac12 \sum_{k,l=1}^\ell \left[ \zeta_{k,l}(t) a_k^{{\dagger} } a_{l}^{{\dagger}} -\text{H.c.} \right] }_{{\equiv} V(t)},$$
$$\mathcal{ U}_I(t,t_0)\equiv e^{i H_0 (t-t_0)/\hbar} \mathcal{ U}(t,t_0).$$
$$i \hbar \frac{\textrm{d}}{\textrm{d}t} \mathcal{ U}_I(t,t_0)=\left( e^{i H_0(t-t_0)/\hbar} V(t) e^{{-}i H_0(t-t_0)/\hbar} \right) \mathcal{ U}_I(t,t_0)=V_I(t) \mathcal{ U}_I(t,t_0).$$
$$e^{i H_0(t-t_0)/\hbar} \boldsymbol{a} e^{{-}i H_0(t-t_0)/\hbar} = \exp\left( i \boldsymbol{\Delta} (t-t_0) \right) \boldsymbol{a} = \boldsymbol{U}(t)\boldsymbol{a},$$
$$\begin{aligned}V_I(t) = e^{i H_0(t-t_0)/\hbar} V(t) e^{{-}i H_0(t-t_0)/\hbar} & =\hbar i \frac12 \sum_{k,l=1}^\ell \left[ \zeta_{k,l}(t) e^{i H_0(t-t_0)/\hbar} a_k^{{\dagger} } a_{l}^{{\dagger}} e^{{-}i H_0(t-t_0)/\hbar} -\text{H.c.} \right] \end{aligned}$$
$$\begin{aligned}& =\hbar i \frac12 \sum_{i,j,k,l=1}^\ell \left[ \zeta_{k,l}(t) U_{l,i}^*(t) a_i^\dagger U_{k,j}^*(t)a_j^\dagger{-}\text{H.c.} \right] \end{aligned}$$
$$\begin{aligned}& =\hbar i \frac12 \sum_{i,j=1}^\ell \left[ \tilde{\zeta}_{i,j}(t) a_i^\dagger a_j^\dagger{-}\text{H.c.} \right], \end{aligned}$$
$$[z_i, z_j^\dagger] = Z_{i,j}, \quad \boldsymbol{z} = \begin{pmatrix} \boldsymbol{a} \\ \boldsymbol{a}^\dagger \end{pmatrix}, \quad \boldsymbol{Z} = \begin{pmatrix} \mathbb{I}_\ell & \boldsymbol{0} \\ \boldsymbol{0} & -\mathbb{I}_\ell \end{pmatrix},$$
$$\boldsymbol{K} \boldsymbol{Z} \boldsymbol{K}^\dagger{=} \boldsymbol{Z}.$$
$$\begin{aligned} \boldsymbol{r} \equiv \begin{pmatrix} \boldsymbol{q} \\ \boldsymbol{p} \end{pmatrix} = \sqrt{\hbar}\frac{1}{\sqrt{2}} \begin{pmatrix} \mathbb{I}_\ell & \mathbb{I}_\ell \\ -i \mathbb{I}_\ell & i \mathbb{I}_\ell \end{pmatrix} \begin{pmatrix} \boldsymbol{a} \\ \boldsymbol{a}^\dagger \end{pmatrix} = \sqrt{\hbar}\boldsymbol{R}\boldsymbol{z}. \end{aligned}$$
$$[r_i,r_j] = i \hbar \Omega_{i,j} \quad \boldsymbol{\Omega} = \begin{pmatrix} 0 & \mathbb{I}_\ell \\ -\mathbb{I}_\ell & \boldsymbol{0} \end{pmatrix}.$$
$$\boldsymbol{z} \to \boldsymbol{z'} = \boldsymbol{K} \boldsymbol{z} \Longleftrightarrow \boldsymbol{r} \to \boldsymbol{r'} = \boldsymbol{S} \boldsymbol{r},$$
$$\begin{aligned} \boldsymbol{S} = \boldsymbol{R} \boldsymbol{K} \boldsymbol{R}^\dagger{=} \begin{pmatrix} \Re(\boldsymbol{V}) + \Re(\boldsymbol{W}) & \Im(\boldsymbol{W}) -\Im(\boldsymbol{V}) \\ \Im(\boldsymbol{V}) + \Im(\boldsymbol{W}) & \Re(\boldsymbol{V}) - \Re(\boldsymbol{W}) \end{pmatrix}. \end{aligned}$$
$$\boldsymbol{S} \boldsymbol{\Omega} \boldsymbol{S}^T = \boldsymbol{\Omega},$$
$$\mathcal{U} = \mathcal{U}_1 \mathcal{U}_0, \quad \mathcal{U}_1 = \exp\left[ \frac12 \sum_{k,l=1}^\ell J_{k,l} a^\dagger_{k} a^\dagger_l - \text{H.c.}\right],$$
$$\mathcal{U}_0^\dagger \mathcal{U}_1^\dagger \boldsymbol{a} \mathcal{U}_1 \mathcal{U}_0 = \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \cosh r_\lambda \right] \boldsymbol{F}^* \left\{ \mathcal{U}_0^\dagger \boldsymbol{a} \mathcal{U}_0 \right\} + \boldsymbol{F} \left[ \oplus_{\lambda=1}^\ell \sinh r_\lambda \right] \boldsymbol{F} \left\{ \mathcal{U}_0^\dagger \boldsymbol{a}^\dagger \mathcal{U}_0 \right\},$$
$$\boldsymbol{F}^* \left\{ \mathcal{U}_0^\dagger \boldsymbol{a} \mathcal{U}_0 \right\} = \boldsymbol{G} \boldsymbol{a} \Longleftrightarrow \boldsymbol{G}^\dagger\boldsymbol{F}^* \left\{ \mathcal{U}_0^\dagger \boldsymbol{a} \mathcal{U}_0 \right\} = \boldsymbol{a},$$
$$\mathcal{U}_0 = \exp\left[{-}i \sum_{j,k} \phi_{j,k}a^\dagger_j a_k \right],$$
$$\mathcal{U}_0^\dagger \boldsymbol{a} \mathcal{U}_0 = \exp\left[ - i\boldsymbol{\phi} \right] \boldsymbol{a},$$
$$H_{\text{DSV}}^{\text{SPDC}}={-}\frac{\zeta_{\text{chan}}^{SSP}}{2}\int\textrm{d}x\,\psi_{S}^{{\dagger}}\left(x\right)\psi_{S}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)+\textrm{H.c.},$$
$$H_{\text{NDSV}}^{\text{SPDC}}={-}\zeta_{\text{chan}}^{SIP}\int\textrm{d}x\,\psi_{S}^{{\dagger}}\left(x\right)\psi_{I}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)+\textrm{H.c.},$$
$$\begin{aligned}\zeta_{\text{chan}}^{J_{1}J_{2}J_{3}} & =\frac{2}{\epsilon_{0}}\sqrt{\frac{\hbar^{3}\omega_{J_{1}}\omega_{J_{2}}\omega_{J_{3}}}{2^{3}}}\int\textrm{d}y\textrm{d}z\,\Gamma_{2}^{ijk}\left(\boldsymbol{r}\right)\left[d_{J_{1}k_{J_{1}}}^{i}\left(y,z\right)d_{J_{2}k_{J_{2}}}^{j}\left(y,z\right)\right]^{*} d_{J_{3}k_{J_{3}}}^{k}\left(y,z\right)\\ & =\hbar\sqrt{\frac{\hbar\omega_{J_{1}}\omega_{J_{2}}\omega_{J_{3}}}{2\epsilon_{0}\overline{n}_{J_{1}}\overline{n}_{J_{2}}\overline{n}_{J_{3}} A_{\text{chan}}^{J_{1}J_{2}J_{3}}}}\overline{\chi}_{2}, \end{aligned}$$
$$\frac{1}{\sqrt{A_{\text{chan}}^{J_{1}J_{2}J_{3}}}}=\frac{\int\textrm{d}y\textrm{d}z\frac{\chi_{2}^{ijk}\left(y,z\right)}{\overline{\chi}_{2}}\left[e_{J_{1}}^{i}\left(y,z\right)e_{J_{2}}^{j}\left(y,z\right)\right]^{*}e_{J_{3}}^{k}\left(y,z\right)}{\mathcal{N}_{J_{1}}\mathcal{N}_{J_{2}}\mathcal{N}_{J_{3}}\sqrt{c^{3}/\left(v_{J_{1}}v_{J_{2}}v_{J_{3}}\right)}},$$
$$H_{\text{DSV}}^{\text{SP-SFWM}}={-}\frac{\gamma_{\text{chan}}^{PPPP}}{2}\hbar^{2}\omega_{P}v_{P}^{2}\int\textrm{d}x\,\psi_{P}^{{\dagger}}\left(x\right)\psi_{P}^{{\dagger}}\left(x\right)\psi_{P}\left(x\right)\psi_{P}\left(x\right),$$
$$\begin{aligned} H_{\text{DSV}}^{\text{DP-SFWM}} & ={-}\gamma_{\text{chan}}^{SSP_{1}P_{2}}\hbar^{2}\overline{\omega}_{SSP_{1}P_{2}}\overline{v}_{SSP_{1}P_{2}}^{2}\int\textrm{d}x\,\psi_{S}^{{\dagger}}\left(x\right)\psi_{S}^{{\dagger}}\left(x\right)\psi_{P_{1}}\left(x\right)\psi_{P_{2}}\left(x\right)+\text{H.c.}\\ & \quad-\frac{\hbar^{2}}{2}\sum_{J}\gamma_{\text{chan}}^{JJJJ}\omega_{J}v_{J}^{2}\int\textrm{d}x\,\psi_{J}^{{\dagger}}\left(x\right)\psi_{J}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)\psi_{J}\left(x\right)\\ & \quad-2\hbar^{2}\sum_{J\neq S}\gamma_{\text{chan}}^{JSJS}\sqrt{\omega_{J}\omega_{S}}v_{J}v_{S}\int\textrm{d}x\,\psi_{J}^{{\dagger}}\left(x\right)\psi_{S}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)\psi_{S}\left(x\right)\\ & \quad-2\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar^{2}\sqrt{\omega_{P_{1}}\omega_{P_{2}}}v_{P_{1}}v_{P_{2}}\int\textrm{d}x\,\psi_{P_{1}}^{{\dagger}}\left(x\right)\psi_{P_{2}}^{{\dagger}}\left(x\right)\psi_{P_{1}}\left(x\right)\psi_{P_{2}}\left(x\right), \end{aligned}$$
$$\begin{aligned} H_{\text{NDSV}}^{\text{DP-SFWM}} & ={-}2\gamma_{\text{chan}}^{SIP_{1}P_{2}}\hbar^{2}\overline{\omega}_{SIP_{1}P_{2}}\overline{v}_{SIP_{1}P_{2}}^{2}\int\textrm{d}x\,\psi_{S}^{{\dagger}}\left(x\right)\psi_{I}^{{\dagger}}\left(x\right)\psi_{P_{1}}\left(x\right)\psi_{P_{2}}\left(x\right)+\textrm{H.c.}\\ & \quad-\frac{\hbar^{2}}{2}\sum_{J}\gamma_{\text{chan}}^{JJJJ}\omega_{J}v_{J}^{2}\int\textrm{d}x\,\psi_{J}^{{\dagger}}\left(x\right)\psi_{J}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)\psi_{J}\left(x\right)\\ & \quad-2\hbar^{2}\sum_{J\neq S}\gamma_{\text{chan}}^{JSJS}\sqrt{\omega_{J}\omega_{S}}v_{J}v_{S}\int\textrm{d}x\,\psi_{J}^{{\dagger}}\left(x\right)\psi_{S}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)\psi_{S}\left(x\right)\\ & \quad-2\hbar^{2}\sum_{J\neq I}\gamma_{\text{chan}}^{JIJI}\sqrt{\omega_{J}\omega_{I}}v_{J}v_{I}\int\textrm{d}x\,\psi_{J}^{{\dagger}}\left(x\right)\psi_{I}^{{\dagger}}\left(x\right)\psi_{J}\left(x\right)\psi_{I}\left(x\right)\\ & \quad-2\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar^{2}\sqrt{\omega_{P_{1}}\omega_{P_{2}}}v_{P_{1}}v_{P_{2}}\int\textrm{d}x\,\psi_{P_{1}}^{{\dagger}}\left(x\right)\psi_{P_{2}}^{{\dagger}}\left(x\right)\psi_{P_{1}}\left(x\right)\psi_{P_{2}}\left(x\right), \end{aligned}$$
$$\begin{aligned} \left(\frac{\partial}{\partial t}+v_{P}\frac{\partial}{\partial x}-i\frac{v_{P}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle & =0,\\ \left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right) & =\frac{i}{\hbar}\zeta_{\text{chan}}^{SSP}\left(x\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \overline{\psi}_{S}^{{\dagger}}\left(x,t\right), \end{aligned}$$
$$\begin{aligned} \left(\frac{\partial}{\partial t}+v_{P}\frac{\partial}{\partial x}-i\frac{v_{P}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle & =0,\\ \left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right) & =\frac{i}{\hbar}\zeta_{\text{chan}}^{SIP}\left(x\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \overline{\psi}_{I}^{{\dagger}}\left(x,t\right),\\ \left(\frac{\partial}{\partial t}+v_{I}\frac{\partial}{\partial x}-i\frac{v_{I}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{I}\left(x,t\right) & =\frac{i}{\hbar}\zeta_{\text{chan}}^{SIP}\left(x\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \overline{\psi}_{S}^{{\dagger}}\left(x,t\right), \end{aligned}$$
$$\begin{aligned} \left(\frac{\partial}{\partial t}+v_{P}\frac{\partial}{\partial x}-i\frac{v_{P}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle & =i\gamma_{\text{chan}}^{PPPP}\hbar\omega_{P}v_{P}^{2}\left|\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle,\\ \left(\frac{\partial}{\partial t}+v_{P}\frac{\partial}{\partial x}-i\frac{v_{P}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\delta\overline{\psi}_{P}\left(x,t\right) & =i\gamma_{\text{chan}}^{PPPP}\hbar\omega_{P}v_{P}^{2}\left[\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle ^{2}\delta\overline{\psi}_{P}^{{\dagger}}\left(x,t\right)\right.\\&\left.+2\left|\left\langle \overline{\psi}_{P}\left(x,t\right)\right\rangle \right|^{2}\delta\overline{\psi}_{P}\left(x,t\right)\right], \end{aligned}$$
$$\begin{aligned} &\left(\frac{\partial}{\partial t}+v_{P_{1}}\frac{\partial}{\partial x}-i\frac{v_{P_{1}}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \\&\quad =i\gamma_{\text{chan}}^{P_{1}P_{1}P_{1}P_{1}}\hbar\omega_{P_{1}}v_{P_{1}}^{2}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar\sqrt{\omega_{P_{1}}\omega_{P_{2}}}v_{P_{1}}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle,\\ &\left(\frac{\partial}{\partial t}+v_{P_{2}}\frac{\partial}{\partial x}-i\frac{v_{P_{2}}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \\& \quad=i\gamma_{\text{chan}}^{P_{2}P_{2}P_{2}P_{2}}\hbar\omega_{P_{2}}v_{P_{2}}^{2}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar\sqrt{\omega_{P_{1}}\omega_{P_{2}}}v_{P_{1}}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle,\\ &\left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right)\\ &\quad =2i\gamma_{\text{chan}}^{SSP_{1}P_{2}}\hbar\overline{\omega}_{SSP_{1}P_{2}}\bar{v}_{SSP_{1}P_{2}}^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \overline{\psi}_{S}^{{\dagger}}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}SP_{1}S}\hbar\sqrt{\omega_{P_{1}}\omega_{S}}v_{S}v_{P_{1}}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{S}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{P_{2}SP_{2}S}\hbar\sqrt{\omega_{P_{2}}\omega_{S}}v_{S}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{S}\left(x,t\right), \end{aligned}$$
$$\begin{aligned} &\left(\frac{\partial}{\partial t}+v_{P_{1}}\frac{\partial}{\partial x}-i\frac{v_{P_{1}}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \\&\quad =i\gamma_{\text{chan}}^{P_{1}P_{1}P_{1}P_{1}}\hbar\omega_{P_{1}}v_{P_{1}}^{2}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar\overline{\omega}_{P_{1}P_{2}P_{1}P_{2}}v_{P_{1}}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle,\\ &\left(\frac{\partial}{\partial t}+v_{P_{2}}\frac{\partial}{\partial x}-i\frac{v_{P_{2}}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \\&\quad =i\gamma_{\text{chan}}^{P_{2}P_{2}P_{2}P_{2}}\hbar\omega_{P_{2}}v_{P_{2}}^{2}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}P_{2}P_{1}P_{2}}\hbar\overline{\omega}_{P_{1}P_{2}P_{1}P_{2}}v_{P_{1}}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle,\\ &\left(\frac{\partial}{\partial t}+v_{S}\frac{\partial}{\partial x}-i\frac{v_{S}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{S}\left(x,t\right)\\ &\quad =2i\gamma_{\text{chan}}^{SIP_{1}P_{2}}\hbar\overline{\omega}_{SIP_{1}P_{2}}\bar{v}_{SIP_{1}P_{2}}^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \overline{\psi}_{I}^{{\dagger}}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{SP_{1}SP_{1}}\hbar\overline{\omega}_{SP_{1}SP_{1}}v_{S}v_{P_{1}}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{S}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{SP_{2}SP_{2}}\hbar\overline{\omega}_{SP_{2}SP_{2}}v_{S}v_{P_{2}}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{S}\left(x,t\right),\\ &\left(\frac{\partial}{\partial t}+v_{I}\frac{\partial}{\partial x}-i\frac{v_{I}^{\prime}}{2}\frac{\partial^{2}}{\partial x^{2}}\right)\overline{\psi}_{I}\left(x,t\right)\\ &\quad =2i\gamma_{\text{chan}}^{SIP_{1}P_{2}}\hbar\overline{\omega}_{SIP_{1}P_{2}}\bar{v}_{SIP_{1}P_{2}}^{2}\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \overline{\psi}_{S}^{{\dagger}}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{P_{1}IP_{1}I}\hbar\overline{\omega}_{P_{1}IP_{1}I}v_{P_{1}}v_{I}\left|\left\langle \overline{\psi}_{P_{1}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{I}\left(x,t\right)\\ & \qquad+2i\gamma_{\text{chan}}^{P_{2}IP_{2}I}\hbar\overline{\omega}_{SP_{2}SP_{2}}v_{P_{2}}v_{I}\left|\left\langle \overline{\psi}_{P_{2}}\left(x,t\right)\right\rangle \right|^{2}\overline{\psi}_{I}\left(x,t\right), \end{aligned}$$
$$M_{ij} = \sum_{k = 0}^{\min\{c_0,c_1\}-1} U_{ik} \lambda_{kk} V_{kj}.$$
$$R_{i_0;i_1 i_2 i_3} \in \mathbb{C}^{2 \times 8}.$$
$$R_{i_0 i_1 ; i_2 i_3} \in \mathbb{C}^{4 \times 4}.$$
$$R_{i_0;i_1\ldots i_{\ell-1}} = \sum_{\alpha_1} U_{i_0; \alpha_1}^{[0]} \lambda^{[0]}_{\alpha_1 ; \alpha_1} V_{\alpha_1;i_1\ldots i_{\ell-1}}^{[0]},$$
$$V_{\alpha_1,i_1;i_2\ldots i_{\ell-1}}^{[0]} = \sum_{\alpha_2} U_{\alpha_1,i_1;\alpha_2}^{[1]} \lambda^{[1]}_{\alpha_2;\alpha_2} V_{\alpha_2;i_2\ldots i_{\ell-1}}^{[1]},$$
$$R_{i_0 i_1\ldots i_{\ell-1}} = \sum_{\alpha_1,\alpha_2 } U_{i_0; \alpha_1}^{[0]} \lambda^{[0]}_{\alpha_1 ; \alpha_1} U_{\alpha_1,i_1;\alpha_2}^{[1]} \lambda^{[1]}_{\alpha_2;\alpha_2} V_{\alpha_2;i_2\ldots i_{\ell-1}}^{[1]}.$$
$$V_{\alpha_2,i_2;i_3\ldots i_{\ell-1}}^{[1]} = \sum_{\alpha_3} U_{\alpha_2,i_2;\alpha_3}^{[2]} \lambda^{[2]}_{\alpha_3;\alpha_3} V_{\alpha_3;i_3\ldots i_{\ell-1}}^{[2]}.$$
$$R_{i_0 i_1\ldots i_{\ell-1}} = \sum_{\alpha_1,\alpha_2 } U_{i_0; \alpha_1}^{[0]} \lambda^{[0]}_{\alpha_1 ; \alpha_1} U_{\alpha_1,i_1;\alpha_2}^{[1]} \lambda^{[1]}_{\alpha_2;\alpha_2} V_{\alpha_2;i_2\ldots i_{\ell-1}}^{[1]},$$
$$V_{\alpha_{k+1},i_{k+1};i_{k+2}\ldots i_{\ell-1}}^{[k]} = \sum_{\alpha_{k+2}} U^{[k+1]}_{\alpha_{k+1},i_{k+1}; \alpha_{k+2}} \lambda^{[k+1]}_{\alpha_{k+2};\alpha_{k+2}} V^{[k+1]}_{\alpha_{k+2};i_{k+2}\ldots i_{\ell-1}}.$$
$$R_{i_0 i_1\ldots i_{\ell-1}} = \sum_{\alpha_1,\alpha_2,\ldots, \alpha_{\ell-1}} U_{i_0; \alpha_1}^{[0]} \lambda^{[0]}_{\alpha_1 ; \alpha_1} U_{\alpha_1,i_1;\alpha_2}^{[1]} \lambda^{[1]}_{\alpha_2;\alpha_2} \ldots U_{\alpha_{\ell-2},i_{\ell-2};\alpha_{\ell-1}}^{[1]} \lambda^{[1]}_{\alpha_{\ell-1};\alpha_{\ell-1}} V_{\alpha_{\ell-1},i_{\ell-1}}$$
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.