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

Channel model for the dual-polarization b-modulated nonlinear frequency-division multiplexing optical transmission systems

Open Access Open Access

Abstract

We consider optical transmission systems based on the nonlinear frequency division multiplexing (NFDM) concept, i.e., the systems employing the nonlinear Fourier transform (NFT) for signal processing and data modulation. Our work specifically addresses the double-polarization (DP) NFDM setup that utilizes the so-called b-modulation, the most efficient NFDM method proposed up-to-date. We extend the previously-developed analytical approach based on the adiabatic perturbation theory for the continuous nonlinear Fourier spectrum (b-coefficient) onto the DP case to obtain the leading order of continuous input-output signal relation, i.e., the asymptotic channel model, for an arbitrary b-modulated DP-NFDM optical communication system. Our main result is in deriving the relatively simple analytical expressions for the power spectral density of the components of effective conditionally Gaussian input-dependent noise emerging inside the nonlinear Fourier domain. We also demonstrate that our analytical expressions are in remarkable agreement with direct numerical results if one extracts the “processing noise” arising due to the imprecision of numerical NFT operations.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The exponentially-growing bandwidth demand in optical networks, which we experience nowadays, is inciting studies aimed at improving the reach and throughput of core fiber-optic communication systems [13]. The nonlinear effects, determined mainly by the Kerr nonlinear silica fiber-media response, are widely accepted as culpable for the optical systems’ efficacy degradation at high signal powers [1,4]. To reduce the effect of nonlinearity backlash and improve the optical systems’ data-transmission capacity, many nonlinearity mitigation techniques have been proposed up to date [3,5]. Among the existing methods, the nonlinear Fourier transform (NFT) based optical signal processing and modulation techniques, and, in particular, the nonlinear frequency division multiplexing (NFDM) as the most efficient constituent of the latter group, have been intensively studied over the last decade [68].

Within the NFDM systems, the data modulation and transmission take place inside the special nonlinear Fourier (NF) domain, where the nonlinear intermodal cross-talk (due to the Kerr effect) between the effective “nonlinear modes” is virtually absent [7], provided that the signal’s evolution is well approximated by an intergable evolutionary equation. When no propagation effects other than the chromatic dispersion and Kerr nonlinearity are present, the signal’s propagation in NFDM systems is unaffected by the fiber nonlinearity capping the channel capacity of conventional systems at high powers. The idea of using the parameters of NF modes rather than those of “traditional” space-time waveforms for data modulation and transmission dates back to the now well-recognized work of Hasegawa and Nyu [9]. The more efficient NFDM variants based on continuous NF spectrum components and efficient modulation formats, which we consider here, were introduced in Refs. [6,10,11]: we notice that it is precisely the utilization of continuous NF spectrum that opened up the possibility for our reaching relatively high data rates in NFDM optical transmission, see, e.g., Ref. [12, Fig. 1] for direct spectral efficiency comparisons. But the systems based on solitonic (discrete NF) modes are also being actively studied [1316], and more sophisticated techniques incorporating both discrete and continuous NF spectrum components have been developed as well [1720], albeit we do not address these here.

 figure: Fig. 1.

Fig. 1. Normalized PSD for $b_1$ (a), $b_2$ (b), and cross-density (c), as the function of the total power factor $b$ for different values of inter-polarization power split ratio $\theta$.

Download Full Size | PDF

Initially, the proof-of-concept NFDM systems predominantly dealt with the scalar optical channel, i.e., with single-polarization signaling. Within such an approach, we operate with the one-component nonlinear Schrödinger equation describing the signal evolution down the fiber and utilize the respective NFT form [6,10,11]. But up to now, the various variants of NFDM have been successfully generalized to the dual-polarization (DP) case [12,13,19,2123]. This generalization became possible because the narrowband light evolution down the single-mode fiber for the DP signal can be well approximated by the integrable version of the Manakov equation; see more details in the next Section. The integrable Manakov equation emerges if we disregard amplified spontaneous emission (ASE) noise, losses, and other higher-order perturbations. The integrability means that the NFT operations associated with the Manakov equation are known [24], so we can explicitly construct the respective DP-NFDM systems based on the DP-NFT signal processing.

It is worthwhile to notice that despite NFDM’s revealing some promising features in combating the nonlinearity-induced signal distortions, it has not so far delivered a truly superlative performance [12], compared to other state-of-the-art optical transmission systems [25]. For example, the latter Ref. reported the spectral efficiency results in the vicinity of 12 bits/s/Hz while Ref. [12] only reported estimates that are 3 times lower (per polarization). Among the NFDM difficulties related to the deviation of the actual transmission channel from the idealized integrable models [7,26], we underline the signal-noise interaction due to inline ASE noise that emerges due to the amplification process; some works name the ASE noise as the main NFDM performance degradation factor [5,27,28]. Yet the signal-noise interplay occurring inside the NF domain for NFDM systems has been studied only for particular single-polarization cases [2731], and there is still a lack of analytical results regarding the noise inside the NF domain for the DP-NFDM, aside from some derivations for a less interesting single-soliton case [3234]. But, naturally, the DP is the most practically important scenario insofar as we do not want to lose the second polarization’s modulation degrees of freedom. We emphasize that doubling the results for the single-polarization case is generally incorrect for the nonlinear signal processing methods to which the NFT belongs. At the same time, the analytical expressions for the channel law are also useful for, e.g., the estimates on the achievable information rates with a mismatched decoding [3539], and, generally, can give us the vision on how to develop more efficient transmission systems [28,40]. Thus, our current work is dedicated to deriving the asymptotic continuous channel model for the DP-NFDM systems, which can help improve their efficiency and explain the physics behind their noise-related performance degradation.

Recently, a specific continuous NFDM-type system, the so-called $b$-modulation [4144], has demonstrated the improved efficiency over previously proposed concepts [12,23,40,45]. In this scheme, we modulate the so-called Jost coefficient $b(\xi )$ [46,47], see Eqs. (6)–(10) below: in contrast to the earlier-proposed scattering coefficient modulation [10], the $b$-modulation allows us to generate the signals having compact support, and the latter helps explicitly control the signal duration and improve system’s efficiency. Therefore, in this paper, we specifically focus on the $b$-modulated DP-NFDM and provide the analytical derivation for the respective continuous (modulation format-independent) channel model in the leading approximation. However, our results can also be readily generalized to the scattering coefficient modulation. Our analytical findings are corroborated by the direct numerical evaluation of the effective NF noise correlation properties for an exemplary $b$-modulated DP-NFDM system. Our results can be seen as the continuation and further extension of single-polarization models from Refs. [28,31,40]. Nonetheless, our findings related to the DP channel are non-trivial and do not boil down to simply applying the single-polarization outcomes to the second polarization.

2. Model, DP-NFT operations, and some NFDM transmission systems details

2.1 Evolutionary equation

We start with writing down the generalized Manakov system, our main model, for the two-component vector electric field envelope $\mathbf {E}=(E_1(z,t),E_2(z,t))$ evolving along the fiber [48] (see also [12,13,21]):

$$\frac{\partial \mathbf{E}}{\partial l} ={-}i \,\frac{\beta_2}{2}\,\frac{\partial^2 \mathbf{E}}{\partial T^2} + i\,\gamma\, \frac{8}{9} \,||\mathbf{E}||^2 \, \mathbf{E} + \mathbf{N},$$
where $i=\sqrt {-1}$, the noise $\mathbf {N}$ is Gaussian delta-correlated process with two uncorrelated complex-valued components $N_i$:
$$\mathbb{E} [ N_{i}(l,T) \,N^*_j (l',T')]= \delta_{ij}\,(N_{ASE}/L) \, \delta(l-l')\,\delta(T-T'),$$

$\mathbb {E}$ is the expectation value, $\delta _{ij}$ is the Kronecker symbol, $||\mathbf {E}||^2$ is the squared vector norm, $T$ is the time in the reference frame co-moving with the envelope, and $l$ is the propagation distance down the fiber. In Eq. (2), $N_{ASE}$ is the noise power spectral density (PSD) per polarization.

There are two different schemes where model Eqs. (1)–(2) are valid. The first one is the ideal distributed Raman amplification case, so the gain-loss term in Eq. (1) is effectively zero. In this case one has $N_{ASE}=\alpha \, L \,h\,\nu \, K_T$, where the factor $K_T \approx 1.13$ at room temperature, $L$ is the total propagation distance, $\nu$ is the carrier frequency, and $h$ is the Plank constant [28].

But this model also applies to the path-averaged approximation of a lumped (erbium-doped fiber amplifiers, EDFA) amplification scheme, where we should substitute the nonlinear coefficient with the path-averaged value $\gamma _{eff}$ [49]:

$$\gamma \to \gamma_{eff}=\gamma[1-\exp(-\alpha L_{a})]/ \alpha L_a,$$
with $L_a$ being the EDFA span length, and $\alpha$ is the attenuation coefficient. Hence, the formula for the noise PSD, in this case, is modified via $N_{ASE}=\big (\exp [\alpha L_a]-1\big )\,N_a\,h\,\nu \,n_{sp}$, with $N_a$ being the number of spans, and $n_{sp}<1$ is the spontaneous emission factor commonly evaluated through the amplifier noise figure [49,50].

In our theoretical analysis below, we do not distinguish between the two aforementioned amplification schemes and assume that Eqs. (1)–(2) describe the true light evolution well enough. However, for the numerical verification of the results given in Section 4, we do consider different implementations.

2.2 Normalization and NFT operations for DP case

Using the standard normalization units (also called soliton units), i.e. adopting some characteristic temporal scale $T_s$ for the time normalization, and then using $L_s=2\,T_s^2/|\beta _2|$ for the propagation distance normalization, and $P_s=(8 \gamma T_s^2/(9 |\beta _2|)^{-1}$ for the power units, we can write the Manakov system down in the standard dimensionless form [13,21,24] for the normalized field variable $q_i=E_i/\sqrt {P_s}$, time $t=T/T_s$, and distance $z = l/L_s$:

$$\begin{aligned} &\frac{\partial q_1}{\partial z} =i\, \frac{\partial^2 q_1}{\partial t^2} +2 \,i\,(|q_1|^2+|q_2|^2) \,q_1+n_1,\\ &\frac{\partial q_2}{\partial z} =i\, \frac{\partial^2 q_2}{\partial t^2} +2\,i (|q_1|^2+|q_2|^2) \,q_2 + n_2,\\ &\mathbb{E } [n_{i}(z,t) \,n^*_j (z',t') ] = \delta_{ij}\,2 D \, \delta(z-z')\,\delta(t-t'),\\ &2 D =\frac{N_{ASE}}{L} \,\frac{L_s}{P_s T_s}, \end{aligned}$$
where we also normalized the noise processes $n_i$, as given by the fourth expression for the normalized PSD intensity $2 D$. Note that system (3) is $z$-reversed compared to the expression from Refs. [13,21,46] and follows the original notations of Manakov [24]. To switch between the notations, one needs to flip the sign of $z$ or complex conjugate the fields.

Neglecting the noise in Eqs. (3), and, thus, dealing with the integrable Manakov system, we now introduce the Jost functions (we omit the $z$-dependence further for brevity), the main quantities in determining the nonlinear eigenmodes. The Jost functions for Manakov equations are the special solutions of the following linear scattering problem [24], written for auxiliary vector- or matrix-valued function $v(t)$:

$$\frac{\partial v}{\partial t} = \left( \begin{array}{ccc} -i \lambda & q_1 & q_2 \\ -q_1^* & i\lambda & 0 \\ -q_2^* & 0 & i\lambda \end{array} \right) \,v = M \,v,$$
where $q_i$ are the components of the field, entering the scattering problem as effective “potentials”, $\lambda$ is called the (nonlinear) spectral parameter that plays the role of Fourier frequency for the NFT, and the definition of matrix operator $M$ is obvious from the expression above.

The Jost functions are defined as four vectors/matrices satisfying the following boundary conditions [13,46] at the leading ($t \to \infty$) and trailing ($t \to -\infty$) ends of our sufficiently localized pulse $[q_1(t),q_2(t)]$:

$$\begin{aligned} \phi & \to \left(\begin{array}{c} 1 \\ 0 \\ 0 \end{array} \right) \,e^{{-}i \lambda t} , \quad \bar{\phi} \to \left(\begin{array}{cc} 0 & 0 \\ 1 & 0 \\ 0 & 1 \end{array}\right)\,e^{i \lambda t}, \quad \quad t \to -\infty,\\ \psi & \to \left(\begin{array}{cc} 0 & 0 \\ 1 & 0 \\ 0 & 1 \end{array}\right)\,e^{i \lambda t}, \quad \bar{\psi} \to \left(\begin{array}{c} 1 \\ 0 \\ 0 \end{array} \right) \,e^{{-}i \lambda t}, \quad t \to + \infty. \end{aligned}$$

Note that the Jost functions in the form of three 3-component different vectors are also used [51], but we found it more convenient to present our further derivations in vector-matrix form (5). The relation between these functions is as follows:

$$\phi(t,\lambda) = \psi(t,\lambda)\,b(\lambda) + \bar{\psi}(t,\lambda) a(\lambda), \qquad \bar{\phi}(t,\lambda) = \psi(t, \lambda) \bar{a}(\lambda) +\bar{\psi}(t, \lambda) \,\bar{b}(\lambda),$$
where $a(\lambda )$ is a scalar, $\bar {a}(\lambda )$ is a 2x2 matrix, $b(\lambda )$ is a two-entry column vector $b=(b_1,b_2)^T$, and $\bar {b}(\lambda )$ is a two-entry row vector.

The fundamental $3 \times 3$ matrices $\Phi$ and $\Psi$ of system (4) are obtained by horizontal concatenation of the Jost functions:

$$\Phi = [\phi, \; \bar{\phi}], \qquad \Psi=[ \bar{\psi}, \; \psi].$$

Since the $M$-operator in (4) is skew-Hermitian, it follows that $\Phi ^{\dagger} \Phi$ and $\Psi ^{\dagger} \Psi$ are $t$-independent for real $\lambda$. Indeed since matrices $\Phi$ and $\Psi$ contain column Jost solutions of system (4) they satisfy the matrix form of this equation where the vector $v$ is substituted by the corresponding matrix. Then taking $\Phi$ as an example one immediately obtains:

$$\frac{\partial}{\partial t} \, (\Phi^{\dagger} \Phi) = \frac{\partial \Phi^{\dagger}}{\partial t} \,\Phi + \Phi^{\dagger} \,\frac{\partial \Phi}{\partial t} = \Phi^{\dagger} \,(M^{\dagger} + M) \,\Phi = 0.$$

Next, from boundary conditions (5), one can deduce that both $\Phi$ and $\Psi$ are also unitary for each $t$. From Eqs. (6), we can see that the two matrices are related: $\Phi = \Psi \,S$, where $S$ is the unimodular scattering matrix [46,51] that is also unitary for real $\lambda$. Its explicit form is:

$$S = \left( \begin{array}{ccc} a & \bar{b}_1 & \bar{b}_2\\ b_1 & \bar{a}_{11} & \bar{a}_{12} \\ b_2 & \bar{a}_{21} & \bar{a}_{22} \end{array} \right).$$

This matrix depends on the spectral parameter $\lambda$ (and the propagation distance, $z$) but not on time, $t$.

The most general conventional NF spectrum consists of the two-entry row and column vectors of the left, $\rho$, or right, $\bar \rho$, reflection coefficients, appended with the set of discrete eigenvalues $\lambda _k$ with respective norming constants (spectral amplitudes) $C_k$ (also left and right). The full left, $\Sigma$, and right, $\bar \Sigma$, scattering data sets are written as:

$$\begin{aligned} \Sigma & = \left\{ \rho(\lambda) =b(\lambda) \,a^{{-}1}(\lambda), \;\lambda_k, \; C_k=b(\lambda_k)/a'(\lambda_k), \; a(\lambda_k)=0 \right\} ,\\ \bar{\Sigma} & = \left\{ \bar{\rho}(\lambda) =\bar{b}(\lambda) \,\bar{a}^{{-}1}(\lambda), \; \bar{\lambda}_k, \; \bar{C_k} =\bar{b}(\bar{\lambda_k})\,\bar{\alpha}(\bar{\lambda_k})/\det \bar{a}(\bar{\lambda}_k)' , \; \det \bar{a}(\bar{\lambda_k})=0 \right\}, \end{aligned}$$
where $\alpha (\lambda )$ is the co-factor matrix for $\bar {a}(\lambda )$ and prime means the derivative with respect to $\lambda$.

However, in what follows we will not be dealing with the discrete (solitonic) NF components as these will be absent by the design in $b$-modulated waveforms, and so that the full NF spectrum is given by just either left or right reflection coefficient.

The following symmetries between the two sets of scattering coefficients exist (for real $\lambda$) [46]:

$$\bar{a}^{\dagger}(\lambda) \bar{a}(\lambda) + \bar{b}^{\dagger}(\lambda) \bar{b}(\lambda) = I_2, \quad |a(\lambda)|^2+b^{\dagger}(\lambda) \,b(\lambda) =1, \quad \bar{a}^{\dagger}(\lambda) b(\lambda) + \bar{b}^{\dagger}(\lambda) a(\lambda) =0 .$$
with $I_2$ being a 2$\times$2 unity matrix. The spatial evolution of the scattering data is given by:
$$\begin{aligned} a(\lambda,z) =a(\lambda,0), & \qquad \bar{a}(\lambda,z) = \bar{a}(\lambda,0),\\ b(\lambda, z) = b(\lambda,0) \,e^{4 i \lambda^2 \,z}, & \qquad \bar{b}(\lambda, z) = \bar{b}(\lambda,0) \,e^{{-}4 i \lambda^2 \,z}. \end{aligned}$$

To reduce the number of degrees of freedom, we shall work with the upper set: $a$ and $b$, as in this case $a$ is just a scalar. For the forward NFT operation, we need to solve Eq. (4) associating the solution $v$ with the Jost vector $\phi$, assuming the initial condition on the signal’s $\mathrm {q}(t)$ trailing end (left boundary) $t=T_0$: $v(T_0)=(1, 0, 0)^T \,\exp (-i \lambda T_0)$. So we solve Eqs. (4) using the numerical scheme described in [21] and restore the Jost coefficients at $t=T_1$ via:

$$v(T_1)= \left( \begin{array}{c} a(\lambda)e^{{-}i\lambda T_1} \\ b_1(\lambda) e^{i\lambda T_1} \\ b_2(\lambda) e^{i\lambda T_1} \end{array} \right).$$

For the completeness of our exposition, we discuss the scalar limit for the DP-NFT. This is obtained by taking the limit $q_2 \to 0$. Then the evolution of the third component is decoupled so that $v_3'=i\lambda v_3$, and the solution is determined by the boundary conditions. In particular, it follows that $\phi _3=\bar {\phi }_{31}=\psi _{31}=\bar {\psi }_3=\psi _{12} = \psi _{32} = 0$. Then from Eq. (6) we obtain that $b_2=\bar {a}_{21}=0$, $\bar {a}_{22}=1$. From symmetry relations (9), it also follows that $\bar {a}_{12}=\bar {b_2}=0$.

The inverse NFT, performed at the NFDM system’s transmitter to generate the time-domain shape of our data-bearing signal, produces the initial DP waveform $\mathbf {q}(t,z=0)=[q_1(t,0),q_2(t,0)]$ from the modulated DP NF spectrum (see the schematics in Refs. [18,40]). Recall that in the case considered, the NF spectrum does not contain discrete solitonic components. Then the resulting signal is launched into the optical fiber. At the receiver, located at some desired distance $z=L$, we apply the direct NFT operation retrieving the modulated NF spectrum and removing the accumulated phase rotation by a single-tap channel equalization according to Eq. (10). However, for our exposition here, we do not need to analyze the explicit form of the inverse DP-NFT. For our numerical analysis, to generate the signal, we used the inverse NFT code from Ref. [21].

2.3 Some notes on the NFDM transmission scheme

First, we notice that the details of the NFDM that we consider here are not different from the standard schemes discussed previously, see, e.g., [40] for the analogous single-polarization system; thus, we give below a very brief description, as it is not crucial for our further analysis. In our study, we address the $b$-modulated NFDM communication as is the recent trend [12,13,21,23]. In this approach, the information is imparted on both components of the $b$-coefficient. However, since according to the third identity in (9) the absolute values of both components of the $b$-vector are upper-bounded by unity, the corresponding squeezing transformation is applied [23,40]. Specifically, the data are encoded onto two loading spectra (one for each polarization), as we typically have for the ordinary polarization division-multiplexing:

$$u_j(\lambda) = \sqrt{S} \,\sum_{k=0}^{N_{sc}-1} \,c_{kj} \,\psi_k(\lambda), \qquad j=1, \, 2,$$
where $S$ is the dimensionless power scaling factor, $c_{ki}$ is the amplitude of $k$-th out of $N_{sc}$ nonlinear subcarriers drawn from a standard constellation (say, M-QAM), and $\psi _k(\lambda )$ form the set of orthogonal nonlinear subcarriers (chosen here to be identical for both polarizations, for simplicity). The most popular choice of subcarriers used exclusively until recently were sinc pulses: $\psi _k(\lambda ) = \mathrm {sinc}(\lambda -k)$. But we notice that, recently, Hermite-Gaussian functions have been suggested as a promising alternative for achieving higher spectral efficiency [45].

Next we apply the exponential mapping from the loading spectra to the $b$-coefficient [12,23]:

$$b_{in,j}(\lambda) = \frac{u_j(\lambda)}{\sqrt{|u_1(\lambda)|^2+|u_2(\lambda)|^2}} \,\sqrt{1 - e^{-|u_1(\lambda)|^2-|u_2(\lambda)|^2}}.$$

This is followed by pre-compensation: $b_j(\lambda,0) = b_{in,j}(\lambda )\,\exp (-4 i\,s\,\lambda ^2 L)$, where $L$ is the propagation distance (in normalized units), and $s \in [0,1]$ is the precompensation factor: $s=0$ corresponds to 100% post-compensation, $s=1$ – to 100% pre-compensation. The most popular choice in NFDM is $s=1/2$ corresponding to the halfway pre-compensation [52] that provides the densest data packing. Then, the spectrum is used for the inverse NFT operation to generate the initial data-bearing two-component waveform at the transmitter.

At the receiver, the operations are performed in reversed order. First, the post-compensation is performed: $b_{out,j}(\lambda ) = b_j(\lambda, L) \exp (-4 i \,(1-s)\,\lambda ^2 \,L)$; next, the loading spectra are restored by inverting Eq. (12). Finally, demodulation is performed for each polarization via the cross-correlation with the corresponding orthogonal wave-carriers $\psi _k(\lambda )$.

In what follows, we do not specialize any particular modulation format and present the general analytical expressions; the particular shape of $b_{1,2}(\lambda )$ is used only for the numerical verification of our results.

3. Perturbation theory

The exact theory of noise in the NFT domain is unavailable, and hence no general channel model exists neither for single nor dual polarization cases. However, when the signal-to-noise ratio is large, we can treat the ASE noise as a perturbation and use the perturbation theory methods to develop an approximate channel model. This was done successfully in a single polarization case for $r$-modulated [28] and $b$-modulated system [40]. Here, we aim to generalize these results for the case of two modulations.

3.1 General perturbative framework for the Manakov system

Perturbation theory for Manakov Eq. (3) was considered in several publications [32,5355]. However, the notation of the above references was slightly cumbersome, and the authors there concentrated mostly on single soliton perturbation theory whereas our goal here is the perturbative evolution of the vector $b$-coefficient. Therefore, in order to make the presentation self-consistent, we develop here the perturbation theory for the $b$-coefficient along the lines of the similar treatment by Kaup [56] for single polarization.

We start by applying Eq. (4) to the Jost vector $\phi$ (see its definition in the previous section). Differentiating both sides of Eq. (4) w.r.t. $z$ and defining vector $u = \partial \phi /\partial z$, we obtain the following inhomogeneous system of equations:

$$\frac{\partial u}{\partial t} - M \,u = f, \quad \text{where} \quad f =\left( \begin{array}{c} \frac{\partial q_1}{\partial z}\, \phi_2 + \frac{\partial q_2}{\partial z}\, \phi_3 \\ -\frac{\partial q_1^*}{\partial z} \,\phi_1 \\ -\frac{\partial q_2^*}{\partial z} \,\phi_1 \end{array}\right).$$

The solution of (13) can be sought using the method of variation of parameters: $u = \frac {\partial \phi }{\partial z} = \Phi \varepsilon$, where $\varepsilon (t,z)$ is a 3 by 1 column vector to be determined. Substituting this into (13) we obtain:

$$\begin{aligned}&\frac{\partial \Phi}{\partial t} \,\varepsilon + \Phi \,\frac{\partial \varepsilon}{\partial t} - M \,\Phi \,\varepsilon = \Phi \,\frac{\partial \varepsilon}{\partial t} = f, \\ &\varepsilon(t) = \intop_{-\infty}^t \Phi^{\dagger}(t') f(t') \, dt',\end{aligned}$$
where in the first line we have used the fact that $\Phi$ is the solution of the homogeneous part of Eq. (13) and in the second line we have used the unitarity of $\Phi$ and the boundary conditions (5) which state that at $t= -\infty$ the Jost functions $\phi$ and $\bar {\phi }$ are $z$-independent so that $u(-\infty )=0$. We have also omitted the explicit $z$-dependence of all the functions for brevity.

We are interested in the $z$-evolution of the Jost scattering coefficients $a$ and $b$. Introducing a vertically concatenated $3 \times 1$ vector $c = [a, \, b]$, the first equation in (6) can be written in a compact form as: $\phi = \Psi \,c$. Differentiating this equation w.r.t. $z$ and taking the limit $t \to +\infty$, we obtain:

$$u(+\infty) =\frac{\partial \phi}{\partial z}(+\infty) = \Psi(+\infty) \frac{\partial c}{\partial z} = \Phi(+\infty)\varepsilon(+\infty) = \Phi(+\infty) \,\intop_{-\infty}^{\infty} \Phi^{\dagger}(t) \,f(t) \, dt ,$$
where we have used the definition of $u$ above, the fact that the $\Psi$ matrix is $z$-independent when $t \to +\infty$ and the previously found solution for $\varepsilon$. Using the relation $\Phi = \Psi S$ and the fact that the scattering matrix $S$, Eq. (7), is unitary and $t$-independent, we obtain the final result:
$$\frac{\partial c}{\partial z} = \intop_{-\infty}^\infty \,\Psi^{\dagger} f \, dt,$$
or, written for each component individually,
$$\begin{aligned} \frac{\partial a}{\partial z} & = \intop_{-\infty}^\infty \left\{ \frac{\partial q_1}{\partial z} \,\bar{\psi}_1^* \, \phi_2 - \frac{\partial q_1^*}{\partial z} \, \bar{\psi}_2^* \, \phi_1 + \frac{\partial q_2}{\partial z} \,\bar{\psi}_1^* \, \phi_3 - \frac{\partial q_2^*}{\partial z} \, \bar{\psi}_3^* \, \phi_1 \right\} \, dt,\\ \frac{\partial b_1}{\partial z} & = \intop_{-\infty}^\infty \left\{ \frac{\partial q_1}{\partial z} \,\psi_{11}^* \, \phi_2 - \frac{\partial q_1^*}{\partial z} \, \psi_{21}^* \, \phi_1 + \frac{\partial q_2}{\partial z} \,\psi_{11}^* \, \phi_3 - \frac{\partial q_2^*}{\partial z} \, \psi_{31}^* \, \phi_1 \right\} \, dt,\\ \frac{\partial b_2}{\partial z} & = \intop_{-\infty}^\infty \left\{ \frac{\partial q_1}{\partial z} \,\psi_{12}^* \, \phi_2 - \frac{\partial q_1^*}{\partial z} \, \psi_{22}^* \, \phi_1 + \frac{\partial q_2}{\partial z} \,\psi_{12}^* \, \phi_3 - \frac{\partial q_2^*}{\partial z} \, \psi_{32}^* \, \phi_1 \right\} \, dt. \end{aligned}$$

One can verify that in the scalar limit, the first two equations transform into the scalar perturbation equations given in Kaup [56]. The above results were first obtained by Midrio et al. [53] using a slightly different method. In the perturbative equations for the $b$-coefficient, we need to substitute the r.h.s. of the Manakov Eq. (1) for $\partial q_i/\partial z$ and their complex conjugates. The terms corresponding to the integrable (noiseless) equation must produce the results compatible with the known spatial evolution of the scattering data (10), i.e. $\partial b/\partial z = 4 \,i\,\lambda ^2 \,b$. The noise terms, therefore, provide the perturbative corrections for this evolution.

The final master model to be used in this paper is therefore:

$$\frac{\partial b_j}{\partial z} = 4 i\,\lambda^2\, b_j^2 + \intop_{-\infty}^\infty \left\{ n_1 \,\psi_{1j}^* \, \phi_2 - n_1^* \, \psi_{2j}^* \, \phi_1 + n_2 \,\psi_{1j}^* \, \phi_3 -n_2^* \, \psi_{3j}^* \, \phi_1 \right\} \, dt, \qquad j=1,\,2.$$

An equation for $a$, which is needed when we consider the modulation of the scattering coefficient $\rho (\lambda )$ itself, can be derived similarly:

$$\frac{\partial a}{\partial z} = \intop_{-\infty}^\infty \left\{n_1 \,\bar{\psi}_1^* \, \phi_2 - n_1^* \, \bar{\psi}_2^* \, \phi_1 \right. + \left. n_2 \,\bar{\psi}_1^* \, \phi_3 - n_2^* \, \bar{\psi}_3^* \, \phi_1 \right\} \, dt.$$

We do not use the latter equation in the current work but provide it for completeness. Equations (15) can be conveniently rewritten as:

$$\frac{\partial b_j}{\partial z} = 4 i\,\lambda^2\, b_j^2 + \Gamma_j(\lambda,z), \qquad j =1, \, 2,$$
where the explicit expressions for $\Gamma _{1,2}(\lambda,z)$ are obvious from comparing Eqs. (15) and (17). Note that we must take the Jost functions entering the aforementioned expressions for $\Gamma _j$, for the unperturbed (integrable) Manakov system; these Jost functions are, therefore, deterministic. It follows then that the vector $\Gamma = (\Gamma _1, \Gamma _2)^T$ is zero mean complex Gaussian process conditioned on the input $b_{1,2}(\lambda,0)$. The covariance and pseudocovariance matrices of this effective nonlinear spectral noise can be obtained using expressions (15). This involves interchanging time integration and ensemble averaging over the progenitor inline ASE noise $n_i$. Since the latter is delta-correlated both in space and time, see Eq. (2), the spatial “whiteness” is carried over to the nonlinear domain while the temporal delta-function removes one temporal integration leading to the following result:
$$\begin{aligned} \mathbb{E} [\Gamma(z,\lambda)\,\Gamma^{\dagger}(z',\lambda')] & = 2D \,\delta(z-z')\, \left( \begin{array}{cc} \bar{A}_1(\lambda, \lambda',z) & \bar{B}(\lambda, \lambda', z) \\ \bar{B}^*(\lambda',\lambda, z) & \bar{A}_2(\lambda, \lambda',z) \end{array} \right) = 2D \,\delta(z-z')\, \bar{\mathcal{C}}(\lambda,\lambda',z),\\ \mathbb{E} [\Gamma(z,\lambda)\,\Gamma^T(z',\lambda')] & = 2D \,\delta(z-z')\, \left( \begin{array}{cc} A_1(\lambda, \lambda',z) & B(\lambda, \lambda', z) \\ B(\lambda',\lambda, z) & A_2(\lambda, \lambda',z) \end{array} \right) =2D \,\delta(z-z')\, \mathcal{C}(\lambda,\lambda',z), \end{aligned}$$
where the elements of the $\mathcal {C}$ and $\bar {\mathcal {C}}$ matrices, $A_j$, $\bar {A}_j$, $B$ and $\bar {B}$, are given by:
$$\begin{aligned} A_j(\lambda,\lambda',z) =\intop_{-\infty}^\infty &\big[ - \psi_{1j}^*(\lambda) \, \phi_2(\lambda) \, \psi_{2j}^*(\lambda') \, \phi_1(\lambda') - \psi_{2j}^*(\lambda) \, \phi_1(\lambda) \,\psi_{1j}^*(\lambda') \, \phi_2(\lambda')\\ &- \psi_{1j}^*(\lambda) \, \phi_3(\lambda) \, \psi_{3j}^*(\lambda') \, \phi_1(\lambda') - \psi_{3j}^*(\lambda) \, \phi_1(\lambda) \,\psi_{1j}^*(\lambda') \, \phi_3(\lambda') \big] \,dt, \end{aligned}$$
$$\begin{aligned} \bar A_j(\lambda,\lambda',z)= \intop_{-\infty}^\infty &\big[ \psi_{1j}^*(\lambda) \, \phi_2(\lambda) \, \psi_{1j}(\lambda') \, \phi_2^*(\lambda') + \psi_{2j}^*(\lambda) \, \phi_1(\lambda) \,\psi_{2j}(\lambda') \, \phi_1^*(\lambda')\\ &+ \psi_{1j}^*(\lambda) \, \phi_3(\lambda) \, \psi_{1j}(\lambda') \, \phi_3^*(\lambda') + \psi_{3j}^*(\lambda) \, \phi_1(\lambda) \,\psi_{3j}(\lambda') \, \phi_1^*(\lambda')\big] \,dt , \end{aligned}$$
$$\begin{aligned} B(\lambda,\lambda',z)= \intop_{-\infty}^\infty &\big[- \psi_{11}^*(\lambda) \, \phi_2(\lambda) \, \psi_{22}^*(\lambda') \, \phi_1(\lambda') - \psi_{21}^*(\lambda) \, \phi_1(\lambda) \,\psi_{12}^*(\lambda') \, \phi_2(\lambda')\\ &- \psi_{11}^*(\lambda) \, \phi_3(\lambda) \, \psi_{32}^*(\lambda') \, \phi_1(\lambda') - \psi_{31}^*(\lambda) \, \phi_1(\lambda) \,\psi_{12}^*(\lambda') \, \phi_3(\lambda') \big] \,dt , \end{aligned}$$
$$\begin{aligned} \bar B(\lambda,\lambda',z) = \intop_{-\infty}^\infty &\big[ \psi_{11}^*(\lambda) \, \phi_2(\lambda) \, \psi_{12}(\lambda') \, \phi_2^*(\lambda') + \psi_{21}^*(\lambda) \, \phi_1(\lambda) \,\psi_{22}(\lambda') \, \phi_1^*(\lambda')\\ &+\psi_{11}^*(\lambda) \, \phi_3(\lambda) \, \psi_{12}(\lambda') \, \phi_3^*(\lambda') + \psi_{31}^*(\lambda) \, \phi_1(\lambda) \,\psi_{32}(\lambda') \, \phi_1^*(\lambda') \big] \, dt . \end{aligned}$$

Again, we have omitted the explicit dependencies of the Jost function components on $z$ and $t$ in the r.h.s. of the expressions above.

3.2 Asymptotic correlation properties of the ASE noise in the NF domain

Equations (15) can be integrated to produce a continuous input-output relation for the vector $b = (b_1, b_2)^T$:

$$\begin{aligned} b_{out}(\lambda) & = b_{in}(\lambda) + N(\lambda),\\ N_j(\lambda) = & \, e^{4 i\,s \,\lambda^2 L} \intop_0^L \,e^{{-}4 i \,\lambda^2 z}\,\Gamma_j(\lambda,z) \, dz, \qquad j=1, \, 2, \end{aligned}$$
where the zero-mean noise vector $N=(N_1,N_2)^T$ represents a complex Gaussian process conditioned on the input $b_{in}$; we have also taken into account the (possible) pre-compensation factor $s$. Note that the channel model (23) includes implicitly the whole equalization block (pre-compensation, fiber propagation, post-compensation) so that in the noise-free case when $N=0$ the output spectrum exactly matches the input.

The correlation properties of the effective NF noise $N_j(\lambda )$ can be readily derived from Eq. (18):

$$\begin{aligned} \mathbb{E} [N(\lambda)\,N^{\dagger}(\lambda')] & =2D\,e^{4 i\, s(\lambda^2- \lambda'^2)\,L} \intop_0^L \, e^{{-}4 i \, s (\lambda^2 - \lambda'^2) \,z } \,\bar{\mathcal{C}}(\lambda, \lambda',z) \, dz ,\\ \mathbb{E} [N(\lambda)\,N^T(\lambda')] & = 2D\,e^{4 i\, s(\lambda^2 + \lambda'^2)\,L} \intop_0^L \, e^{{-}4 i \, s (\lambda^2 + \lambda'^2) \,z } \,\mathcal{C}(\lambda, \lambda',z) \, dz . \end{aligned}$$

Unfortunately, expressions (24) are generally of little practical use, since the Jost functions for the unperturbed problem entering matrices $\mathcal {C}$, $\bar {\mathcal {C}}$, are rarely available in the analytical form. However, at large distances, one can use asymptotic expressions for these coefficients, as was done in [28,40]. The usual method relies on asymptotic expressions first derived by Zakharov and Manakov [57] and can be further elaborated via the steepest descent for Riemann–Hilbert factorization problem, see [40,58]. However, when one is interested only in the leading terms in the large distance asymptotic expansion of the covariance and pseudocovariance matrices, it is possible to use a simplified analysis first applied for the $r$-modulated scalar transmission, see Ref. [28, Supplementary Note III]. In this approach, we assume that the time domain pulse has a finite extent: $q_j(t) \equiv 0$ if $t<T_1$ and $t>T_2$, with arbitrary fields $q_j(t)$, containing no solitonic component in its NFT decomposition, inside the finite interval $[T_1,T_2]$ (as before we omit the $z$-dependence for brevity). Such an assumption complies with the “burst mode” modulation requirement that has to be used for the NFT-based transmission methods to avoid intersymbol interference, as we always process a finite interval. It is also true for the original version of unscaled $b$-modulation by design [41]. It is now convenient to split the integrals in the general expression for the (pseudo) densities $A_j$, $\bar A_j$, $B$, and $\bar B$, into tree subintervals: $t<T_1$, $T_1<t<T_2$, and $t>T_2$. The quantities referring to the first interval, i.e. to the left from the finite $q(t)$ extent, will be marked with the superscript “<”; the quantities for the interior interval, where $q_j(t) \neq 0$, will be indexed with the tilde accent “$\sim$” and the remaining interval to the right from the $q(t)$ extent is marked with “>”. Further, we will need the Fourier transform of the Heaviside unit step function, $H(t)=1$ for $t>0$ and 0 otherwise:

$$H(\omega) = \intop_{-\infty}^{\infty} \!\! dt \, H(t) \, e^{- i \omega t} ={-} \frac{i}{ \omega} + \pi \, \delta(\omega),$$
where the first term is to be understood in the sense of principal value.

Using (5), we immediately obtain:

$$\phi_1^<(\lambda,t) = e^{{-}i \lambda t}, \qquad \phi_{2,3}^<(\lambda,t) = 0,$$
$$\psi_{11}^>{=} \psi^>_{12} = \psi^>_{31} = \psi^>_{22} = 0, \qquad \psi_{21}^>{=} \psi_{32}^>{=} e^{i \lambda t}.$$

We also need opposite end asymptotes for both $\phi$ and $\psi$. These can be obtained via the unitary scattering matrix $S$, Eq. (6):

$$\begin{aligned} \phi^>(\lambda,t) & = a(\lambda) \,\bar{\psi}^>(\lambda, t) + \psi^>(\lambda,t)\,b(\lambda),\\ \phi_1^>(\lambda,t) & = a(\lambda) \, e^{- i \lambda t}, \qquad \phi_{2,3}^>(\lambda,t) = b_{1,2}(\lambda) \, e^{i \lambda t}. \end{aligned}$$

Next, using $\Psi = \Phi \, S^{\dagger}$, we obtain:

$$\begin{aligned} \psi^<(\lambda,t) & = \phi^<(\lambda,t)\,b^{\dagger}(\lambda) + \bar{\phi}^<(\lambda,t)\,\bar{a}^{\dagger}(\lambda),\\ \psi_{1j}^<(\lambda,t) & =b_j^*(\lambda) \, e^{- i \lambda t}, \qquad \psi_{2j}^<(\lambda,t) = \bar{a}^*_{j\,1}(\lambda) \, e^{i \lambda t},\\ \psi_{3j}^<(\lambda,t) & = \bar{a}^*_{j\,2}(\lambda) \, e^{i \lambda t}, \qquad j=1, \, 2. \end{aligned}$$

Now we can calculate the partial contributions to the densities. We start from the region to the left from the $q_j(t)$ extent, substituting Eqs. (26), (27) into expressions (19)–(22), we obtained:

$$A_j^<(\lambda,\lambda',z) = 0,$$
$$\bar{A}_j^<(\lambda,\lambda',z) = \left[\bar{a}_{j1}(\lambda)\,\bar{a}^*_{j1}(\lambda') + \bar{a}_{j2}(\lambda)\,\bar{a}^*_{j2}(\lambda')\right] \,e^{{-}2i\,(\lambda-\lambda')T_1}\,\left( \frac{i}{2(\lambda-\lambda')} +\frac{\pi}{2}\,\delta(\lambda-\lambda')\right) ,$$
$$B^<(\lambda,\lambda',z) =0,$$
$$\bar{B}_j^<(\lambda,\lambda',z) = \left[\bar{a}_{11}(\lambda)\,\bar{a}^*_{21}(\lambda') + \bar{a}_{12}(\lambda)\,\bar{a}^*_{22}(\lambda')\right] \,e^{{-}2i\,(\lambda-\lambda')T_1}\,\left( \frac{i}{2(\lambda-\lambda')} +\frac{\pi}{2}\,\delta(\lambda-\lambda')\right) .$$

A similar analysis applies to the right region, $t>T_2$:

$$A_j^>(\lambda,\lambda',z) = 0,$$
$$\bar{A}_j^>(\lambda,\lambda',z) = a(\lambda)\,a^*(\lambda')\, e^{{-}2i\,(\lambda-\lambda')T_2} \left[ -\frac{i}{2(\lambda-\lambda')} +\frac{\pi}{2} \,\delta(\lambda-\lambda') \right],$$
$$B^>(\lambda,\lambda',z) =\bar{B}^>(\lambda,\lambda',z)=0.$$

Apart from these terms, there are the contributions from the interval $[T_1,T_2]$: we designate these as $\tilde {A}_j$, $\bar {\tilde {A}}_j$, $\tilde {B}$ and $\bar {\tilde {B}}$. The total values of the densities can be obtained by summing up the individual contributions from left, right, and center subintervals.

The crucial observation first made in Ref. [28], is that the covariance and pseudocovariance matrices (24) contain spatial integral of highly oscillating factors. Therefore, at large distances $z$, the leading order contribution to the covariance matrix comes from the values $\lambda \approx \lambda '$. Hence, the contributions from the intermediate region $[T_1,T_2]$ as well as the principle value terms of the densities only contribute $O(1)$ terms, and the main contribution comes only from the Dirac-delta terms in the boundary regions. For the pseudocovariance, as we have shown above, $\mathcal {C}^\lessgtr =0$, and the contribution of the intermediate region is again of the order $O(1)$.

Therefore, keeping only the leading-order diagonal Dirac-delta terms, we obtain:

$$\bar{\mathcal{C}} (\lambda,\lambda',z) = \left( \begin{array}{cc} \sum_{i=1}^2|\bar{a}_{1i}(\lambda)|^2+|a(\lambda)|^2 & \sum_{i=1}^2 \bar{a}_{1i}(\lambda)\,\bar{a}^*_{2i}(\lambda) \\ & \\ \sum_{i=1}^2\bar{a}^*_{1i}(\lambda)\,\bar{a}_{2i}(\lambda) & \sum_{i=1}^2|\bar{a}_{2i}(\lambda)|^2+|a(\lambda)|^2 \end{array} \right)\times \frac{\pi}{2}\,\delta(\lambda-\lambda') .$$

Finally, from the unitarity of the $S$ matrix (7) it follows that $|b_i|^2+|\bar {a}_{i1}|^2 +|\bar {a}_{i2}|=1$, $\bar {a}_{11}\,\bar {a}^*_{21}+\bar {a}_{12}\,\bar {a}^*_{22} + b_1\,b_2^*=0$ and $|a|^2+|b_1|^2+b_2|^2 =1$. This allows us to arrive at the following remarkably simple answer:

$$\mathbb{E} [N(\lambda)\,N^{\dagger}(\lambda')] = n_{ASE} \, B(\lambda)\, \delta(\lambda-\lambda') , \qquad \mathbb{E} [N(\lambda)\,N^T(\lambda')] =0 ,$$
$$B = \left( \begin{array}{cc} 1-|b_1(\lambda)|^2-|b_2(\lambda)|^2/2 & - \frac{1}{2} b_1(\lambda)\, b_2^*(\lambda)\\ & \\ - \frac{1}{2} b^*_1(\lambda)\, b_2(\lambda) & 1-|b_2(\lambda)|^2-|b_1(\lambda)|^2/2 \end{array} \right),$$
where $n_{ASE} = 2 \pi D L$ is the noise PSD in the linear case in the dimensionless units.

The above is the main result of our work. When only one polarization is used, $b_2=0$, one recovers the results for the scalar case given by the upper left corners of the covariance and pseudocovariance matrices:

$$\mathbb{E}[N(\lambda)^* N(\lambda')] = n_{ASE}(1-|b_1(\lambda)|^2)\,\delta(\lambda-\lambda'), \qquad \mathbb{E}[N(\lambda) N(\lambda')] = 0,$$
obtained earlier in Ref. [40].

On the other hand, interestingly, the noise in an initially “empty” polarization is squeezed: $B_{22}=1-|b_1(\lambda )|^2/2<1$. The maximum squeezing amount is 50%.

In the general case, we can see that the contribution of the cross-polarization to the noise PSD reduction is half of the self-term. Also, due to the constraint $|b_1|^2+|b_2|^2 \leq 1$, the normalized PSDs $B_{11}$ and $B_{22}$ are both positive as required.

4. Numerical verification

In order to perform the numerical verification of the above result, we used two popular transmission schemes corresponding to ideal distributed Raman amplification and lumped EDFA [1] with the parameters given in Table 1: the fiber parameters there are typical for the standard single mode fiber in the C-band regime [7].

Tables Icon

Table 1. The parameters of the single-mode fiber link used for the numerical verification of our results.

For the signal propagation in the fiber we employed a standard split-step Fourier method [59] with a fixed step size of 0.2 km; for the direct and reciprocal NFT code we used the algorithms of Ref. [21].

4.1 Rectangular spectra

For simplicity, in the first set of simulations we consider the case where the initial components of the $b$ vector are the rectangular spectral shapes of constant real-valued amplitudes $b_1$ and $b_2$, localized in the interval $\xi \in [-2,2]$, which roughly corresponds to the two-sided linear bandwidth $W \approx 4/(\pi T_s) =12$ GHz. Since we must have $|b_1|^2 + |b_2|^2 \leq 1$, it makes sense to use the trigonometric parametrization: $b_1 = b\cos \theta$, $b_2 = b\sin \theta$, where $0 \leq b \leq 1$ is the measure of the total power in two polarizations, and $0 \leq \theta \leq \pi /2$ characterizes the power splitting ratio between them.

We have performed a series of Monte-Carlo runs for the distributed Raman case, and the results are depicted in Fig. 1, where we plot the computed normalized PSDs of the individual components $B_{11}$ and $B_{12}$ as well as the cross-density $\mathrm {Re}(B_{12})$, see Eq. (32), vs. the total power factor $b$ for different power split ratios parameterized by $\theta$. One can readily observe a very good agreement between theory and numerics for the self-PSDs, Fig. 1(a), (b), and slight discrepancy for the cross-correlation, Fig. 1(c). The latter is more pronounced at large input powers, and we attribute it to two factors. First, the absolute value of the leading order contribution to $B_{12}$ is relatively low, and, thus, the next orders of the $B_{12}$-expansion (that we did not account for in the final analytical formula) contribute a visually distinguishable amount. Then, we also have the contribution from the processing noise [31,40], i.e. from the final accuracy of the numerical NFT routines used in the simulations. In the next subsection, we provide a brief description of how one can “subtract” the processing noise contribution and extract the ASE-related PSD.

Observe also that in the case of a single polarization, $\theta =0$, the PSD of the noise in the second polarization is reduced up to the maximum level of half the linear density, Fig. 1(b), as discussed at the end of the previous section and in full correspondence with our analytical predictions.

4.2 NFDM spectra

Next, we analyze the performance of our theoretical model in more realistic NFDM settings. The basic setup of b-modulated NFDM systems for both single a dual polarization was described in Section 2.3 and can be also found in multiple references (see e.g. [12,18,23,40]).

In this more applicable scenario, we note that we ought to extract the ‘’processing noise” from the overall result to obtain the correspondence with the analytics. The procedure was first suggested in [31]. Briefly, the extraction procedure works as follows: we assume that the total signal distortion at the receiver is the sum of ASE and the processing noise that are assumed to be statistically independent zero mean processes. This means that their variances (which are properly normalized nonlinear spectral densities and pseudo-densities studied here) should add up. Therefore to extract the ASE-only contribution to the signal which is the subject of the present theory we have performed a separate set of simulations where at each power level the inline ASE noise was artificially switched off in the propagation model. In this simulation, we randomly generated different sequences of input symbols (see below) by keeping the overall power scaling factor fixed. The resulting densities obtained from this noise-free run were then subtracted from the results of full simulation (at the same level of input power) to provide ASE only contribution.

We used 128 subcarriers with 32-QAM pseudorandom constellation points and a sinc carrier shape for both polarizations: $\psi _n= \mathrm {sinc}(\lambda -n)$. We also used half-way pre-compensation, $s=1/2$. The scaling of the NFT parameters corresponded to lumped EDFA amplification case as discussed in Section 2.1 with the parameters from Table 1. The corresponding results are shown in Fig. 2.

 figure: Fig. 2.

Fig. 2. Normalized PSD for: $b_1$ (a), (d); $b_2$ (b), (e), and the real part of the cross-density (c), (f), as the function of the total power of NFDM burst. Top row: full simulation, bottom row: after the subtraction of processing noise.

Download Full Size | PDF

One can see that since both $b_1$ and $b_2$ have the same statistical distribution of the input symbols, their average PSDs are identical (we checked that the relative error does not exceed $5 \times 10^{-3}$). Next, one can clearly see from the top row figures that the noise densities and cross densities begin to grow with power at around $-5$ dBm, contrary to the prediction of the theory. This is exactly the manifestation of the processing noise not related to the ASE but rather to the numerical accuracy of the NFT-related DSP, burst tails truncation, insufficient sampling rate, etc. [31,40]. When we extract the processing noise, see the lower row of panels in Fig. 2, the agreement with the theory becomes almost perfect for the PSDs, while the cross-density if almost negligible in all the cases (note that the scale of the panes (c) and (f) vertical axis is reduced compared to other figures).

5. Conclusion

In our paper, we derived a channel model for NFDM b-modulated system in the presence of the inline ASE noise, exploring the previously unattended case of double polarization modulation. The aim of our study was to derive analytically and verify the respective channel model numerically, i.e., to study the effective noise emerging in the NF domain due to the progenitor ASE noise. In fact, our work extends the previously-developed perturbative approach for the derivation of noise statistics onto the DP case. We note that our analytical results refer to the leading order of continuous input-output signal relation but are universal and applicable to any DP $b$-modulation type; the expressions for the reflection coefficient modulations, requiring our considering the $a$-coefficient, can also be obtained using the first equation from the set (14), but the b-modulation concept seems to be more appealing in terms of achievable efficiency.

We want to emphasize the non-trivial findings of our work here. In particular, we demonstrated that the contribution of the second polarization to the diagonal elements of the covariance matrix in the DP case amounts to just half of the signal’s own polarization contribution. It means, in particular, that if we set the coefficient $b$ in one polarization close to 1 and to 0 in the second polarization, it would bring about the appearance of the non-zero noise in the second polarization, while the noise effect in the first polarization will dwindle. Our results were verified through direct Monte-Carlo simulations: it was shown that the components of the covariance matrix for the effective input-dependent noise in the NF domain are in very good agreement with direct numerical results. The deviations of the theory and the numerical data for the off-diagonal covariance matrix elements are attributed to the contributions of the high-order expansion terms and the effects of processing noise; when we extract the latter, the correspondence becomes almost perfect.

This paper follows our previous works [28,31,40] devoted to the channel model derivation for the NFDM systems that use the continuous NF spectrum modulation. However, we note that the case where the solitary modes (the discrete NF spectrum) coexist with the continuous NF spectrum components, as in Refs. [1820], still requires a separate study: the noise-perturbed evolutionary equations for the NF spectral quantities, given in Ref. [60], have to be used together with the respective Jost functions that account for the coexisting NF spectra, and the analytical expressions for the asymptotics of the latter functions seem to be unavailable at the moment. Finally, specifically for the DP case, the master Manakov model (1) represents only an approximation and does not take into account the effects of polarization mode dispersion (PMD) or polarization-dependent loss. The implications of these effects on DP-NFDM transmission are still poorly studied but we mention that a recent publication [61] does consider this subject and offers an effective scheme for channel equalization in this case.

Funding

Leverhulme Trust (RP-2018-063); Israel Science Foundation (466/18).

Acknowledgements

SAD was supported by the Israel Science Foundation (grant No. 466/18). JEP acknowledges Leverhulme Trust Project RP-2018-063.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon a reasonable request.

References

1. R.-J. Essiambre, G. Kramer, P. J. Winzer, G. J. Foschini, and B. Goebel, “Capacity limits of optical fiber networks,” J. Lightwave Technol. 28(4), 662–701 (2010). [CrossRef]  

2. E. Agrell, M. Karlsson, A. R. Chraplyvy, D. J. Richardson, P. M. Krummrich, P. Winzer, K. Roberts, J. K. Fischer, S. J. Savory, B. J. Eggleton, M. Secondini, F. R. Kschischang, A. Lord, J. Prat, I. Tomkos, J. E. Bowers, A. Srinivasan, M. Brandt-Pearce, and N. Gisin, “Roadmap of optical communications,” J. Opt. 18(6), 063002 (2016). [CrossRef]  

3. P. J. Winzer, D. T. Neilson, and A. R. Chraplyvy, “Fiber-optic transmission and networking: the previous 20 and the next 20 years,” Opt. Express 26(18), 24190–24239 (2018). [CrossRef]  

4. M. Karlsson, N. Alic, S. Chandrsekhar, and A. Mecozzi, “Feature issue introduction: Nonlinearity mitigation for coherent transmission systems,” Opt. Express 25(4), 4552–4553 (2017). [CrossRef]  

5. J. C. Cartledge, F. P. Guiomar, F. R. Kschischang, G. Liga, and M. P. Yankov, “Digital signal processing for fiber nonlinearities,” Opt. Express 25(3), 1916–1936 (2017). [CrossRef]  

6. M. I. Yousefi and F. R. Kschischang, “Information transmission using the nonlinear Fourier transform, Part i–iii,” IEEE Trans. Inf. Theory 60(7), 4312–4328 (2014). [CrossRef]  

7. S. K. Turitsyn, J. E. Prilepsky, S. T. Le, S. Wahls, L. L. Frumin, M. Kamalian, and S. A. Derevyanko, “Nonlinear Fourier transform for optical data processing and transmission: advances and perspectives,” Optica 4(3), 307 (2017). [CrossRef]  

8. L. Xi, J. Wei, and W. Zhang, “Applications of machine learning on nonlinear frequency division multiplexing optic-fiber communication systems,” in 2021 IEEE 9th International Conference on Information, Communication and Networks (ICICN), (2021), pp. 190–194.

9. A. Hasegawa and T. Nyu, “Eigenvalue communication,” J. Lightwave Technol. 11(3), 395–399 (1993). [CrossRef]  

10. J. E. Prilepsky, S. A. Derevyanko, K. J. Blow, I. Gabitov, and S. K. Turitsyn, “Nonlinear inverse synthesis and eigenvalue division multiplexing in optical fiber channels,” Phys. Rev. Lett. 113(1), 013901 (2014). [CrossRef]  

11. S. T. Le, J. E. Prilepsky, and S. K. Turitsyn, “Nonlinear inverse synthesis for high spectral efficiency transmission in optical fibers,” Opt. Express 22(22), 26720–26741 (2014). [CrossRef]  

12. X. Yangzhang, V. Aref, S. T. Le, H. Buelow, D. Lavery, and P. Bayvel, “Dual-Polarization Non-Linear Frequency-Division Multiplexed Transmission With b-Modulation,” J. Lightwave Technol. 37(6), 1570–1578 (2019). [CrossRef]  

13. S. Gaiarin, A. M. Perego, E. P. da Silva, F. Da Ros, and D. Zibar, “Dual-polarization nonlinear Fourier transform-based optical communication system,” Optica 5(3), 263–270 (2018). [CrossRef]  

14. G. Zhou, L. Sun, C. Lu, and A. P. T. Lau, “Multi-Symbol Digital Signal Processing Techniques for Discrete Eigenvalue Transmissions Based on Nonlinear Fourier Transform,” J. Lightwave Technol. 39(17), 5459–5467 (2021). [CrossRef]  

15. G. Zhou, T. Gui, T. Chan, C. Lu, A. P. T. Lau, and P. Wai, “Signal Processing Techniques for Nonlinear Fourier Transform Systems,” in Optical Fiber Communication Conference (OFC) 2019, (Optical Society of America, 2019), p. M2H.5.

16. S. Wahls, “Shortening solitons for fiber-optic transmission,” in 2021 17th International Symposium on Wireless Communication Systems (ISWCS), (2021), pp. 1–6.

17. V. Aref, S. T. Le, and H. Buelow, “Modulation over nonlinear Fourier spectrum: Continuous and discrete spectrum,” J. Lightwave Technol. 36(6), 1289–1295 (2018). [CrossRef]  

18. S. Le, V. Aref, and H. Buelow, “Nonlinear signal multiplexing for communication beyond the Kerr nonlinearity limit,” Nat. Photonics 11(9), 570–576 (2017). [CrossRef]  

19. F. Da Ros, S. Civelli, S. Gaiarin, E. P. da Silva, N. De Renzis, M. Secondini, and D. Zibar, “Dual-polarization NFDM transmission with continuous and discrete spectral modulation,” J. Lightwave Technol. 37(10), 2335–2343 (2019). [CrossRef]  

20. A. Vasylchenkova, J. E. Prylepsky, N. B. Chichkov, and S. K. Turitsyn, “Multieigenvalue communication paired with b-modulation,” in 45th European Conference on Optical Communication (ECOC 2019), (2019), pp. 1–4.

21. J. W. Goossens, M. I. Yousefi, Y. Jaouan, and H. Hafermann, “Polarization-division multiplexing based on the nonlinear Fourier transform,” Opt. Express 25(22), 26437–26452 (2017). [CrossRef]  

22. S. Civelli, S. K. Turitsyn, M. Secondini, and J. E. Prilepsky, “Polarization-multiplexed nonlinear inverse synthesis with standard and reduced-complexity NFT processing,” Opt. Express 26(13), 17360–17377 (2018). [CrossRef]  

23. X. Yangzhang, V. Aref, S. T. Le, H. Buelow, D. Lavery, and P. Bayvel, “Experimental Demonstration of Dual-polarisation NFDM Transmission with b-Modulation,” IEEE Photonics Technol. Lett. 31(11), 885–888 (2019). [CrossRef]  

24. S. Manakov, “On the theory of two-dimensional stationary self-focusing of electromagnetic waves,” Sov. Phys. JETP 36, 248–253 (1974).

25. M. Shtaif, C. Antonelli, A. Mecozzi, and X. Chen, “Challenges in estimating the information capacity of the fiber-optic channel,” Proc. IEEE 110(11), 1655–1678 (2022). [CrossRef]  

26. X. Yangzhang, D. Lavery, P. Bayvel, and M. I. Yousefi, “Impact of perturbations on nonlinear frequency-division multiplexing,” J. Lightwave Technol. 36(2), 485–494 (2018). [CrossRef]  

27. S. Civelli, E. Forestieri, and M. Secondini, “Mitigating the Impact of Noise on Nonlinear Frequency Division Multiplexing,” Appl. Sci. 10(24), 9099 (2020). [CrossRef]  

28. S. A. Derevyanko, J. E. Prilepsky, and S. K. Turitsyn, “Capacity estimates for optical transmission based on the nonlinear Fourier transform,” Nat. Commun. 7(1), 12710 (2016). [CrossRef]  

29. S. Civelli, E. Forestieri, and M. Secondini, “Why noise and dispersion may seriously hamper nonlinear frequency-division multiplexing,” IEEE Photonics Technol. Lett. 29(16), 1332–1335 (2017). [CrossRef]  

30. I. Tavakkolnia and M. Safari, “Capacity analysis of signaling on the continuous spectrum of nonlinear optical fibers,” J. Lightwave Technol. 35(11), 2086–2097 (2017). [CrossRef]  

31. M. Pankratova, A. Vasylchenkova, S. A. Derevyanko, N. B. Chichkov, and J. E. Prilepsky, “Signal-noise interaction in optical-fiber communication systems employing nonlinear frequency-division multiplexing,” Phys. Rev. Appl. 13(5), 054021 (2020). [CrossRef]  

32. S. A. Derevyanko, J. E. Prilepsky, and D. A. Yakushev, “Statistics of a noise-driven Manakov soliton,” J. Phys. A: Math. Gen. 39(6), 1297–1309 (2006). [CrossRef]  

33. E. Doktorov and I. Kuten, “Suppression of timing jitter for optical solitons in fibers,” Nonlinear Phenomena in Complex Systems 4, 1–5 (2001).

34. N. A. Shevchenko, S. A. Derevyanko, J. E. Prilepsky, A. Alvarado, P. Bayvel, and S. K. Turitsyn, “Capacity lower bounds of the noncentral chi-channel with applications to soliton amplitude modulation,” IEEE Trans. Commun. 66(7), 2978–2993 (2018). [CrossRef]  

35. D. Arnold, H. Loeliger, P. Vontobel, A. Kavcic, and W. Zeng, “Simulation-based computation of information rates for channels with memory,” IEEE Trans. Inf. Theory 52(8), 3498–3508 (2006). [CrossRef]  

36. M. Secondini, E. Forestieri, and G. Prati, “Achievable information rates in nonlinear wdm fiber-optic systems with arbitrary modilation formats and dispersion maps,” J. Lightwave Technol. 31(23), 3839–3852 (2013). [CrossRef]  

37. A. Alvarado, T. Fehenberge, B. Che, and F. M. J. Willems, “Achievable information rates for fiber optics: Applications and computations,” J. Lightwave Technol. 36(2), 424–439 (2018). [CrossRef]  

38. M. Baniasadi, Y. Chen, and M. Safari, “On the capacity of b-modulated nonlinear frequency division multiplexing,” in 2022 IEEE International Symposium on Information Theory (ISIT), (IEEE, 2022), pp. 1933–1938.

39. Y. Chen, M. Baniasadi, and M. Safari, “Optimal shaping gains for continuous spectrum nonlinear frequency signalling over long fibers,” in 2022 IEEE International Symposium on Information Theory (ISIT), (IEEE, 2022), pp. 1927–1932.

40. S. A. Derevyanko, M. Balogun, O. Aluf, D. Shepelsky, and J. E. Prilepsky, “Channel model and the achievable information rates of the optical nonlinear frequency division-multiplexed systems employing continuous b-modulation,” Opt. Express 29(5), 6384–6406 (2021). [CrossRef]  

41. S. Wahls, “Generation of time-limited signals in the nonlinear fourier domain via b-modulation,” in 2017 European Conference on Optical Communication (ECOC), (2017), pp. 1–3.

42. T. Gui, G. Zhou, C. Lu, A. P. T. Lau, and S. Wahls, “Nonlinear frequency division multiplexing with b-modulation: Shifting the energy barrier,” Opt. Express 26(21), 27978–27990 (2018). [CrossRef]  

43. S. Chimmalgi and S. Wahls, “Bounds on the transmit power of b-modulated NFDM systems in anomalous dispersion fiber,” Entropy 22(6), 639 (2020). [CrossRef]  

44. D. Shepelsky, A. Vasylchenkova, J. E. Prilepsky, and I. Karpenko, “Nonlinear Fourier Spectrum Characterization of Time-Limited Signals,” IEEE Trans. Commun. 68(5), 3024–3032 (2020). [CrossRef]  

45. M. Balogun and S. Derevyanko, “Hermite-Gaussian Nonlinear Spectral Carriers for Optical Communication Systems Employing the Nonlinear Fourier Transform,” IEEE Commun. Lett. 26(1), 109–112 (2022). [CrossRef]  

46. M. Ablowitz, B. Prinari, and A. Trubatch, Discrete and Continuous Nonlinear Schrodinger systems (Cambridge University Press, 2003).

47. V. E. Zakharov and A. B. Shabat, “Exact theory of two-dimensional self-focusing and one-dimensional self-modulation of waves in nonlinear media,” Soviet Physics JETP 34, 62–69 (1972).

48. R. J. Essiambre, R. Tkach, and R. Ryf, “Fiber nonlinearity and capacity: Single mode and multimode fibers,” in Optical Fiber Telecommunications, Vol. VIB: Systems and Networks, I. Kaminow, T. Li, and A. Willner, eds. (Academic Press, 2013).

49. S. T. Le, J. E. Prilepsky, and S. K. Turitsyn, “Nonlinear inverse synthesis technique for optical links with lumped amplification,” Opt. Express 23(7), 8317–8328 (2015). [CrossRef]  

50. M. Kamalian, J. E. Prilepsky, S. T. Le, and S. K. Turitsyn, “On the design of NFT-based communication systems with lumped amplification,” J. Lightwave Technol. 35(24), 5464–5472 (2017). [CrossRef]  

51. J. A. E. Rocha, Short-time asymptotic analysis of the Manakov system, PhD Thesis (The University of Arizona, 2006).

52. I. Tavakkolnia and M. Safari, “Dispersion pre-compensation for NFT-based optical fiber communication systems,” in 2016 Conference on Lasers and Electro-Optics (CLEO), (IEEE, 2016), pp. 1–2.

53. M. Midrio, S. Wabnitz, and P. Franco, “Perturbation theory for coupled nonlinear Schrödinger equations,” Phys. Rev. E 54(5), 5743–5751 (1996). [CrossRef]  

54. T. Lakoba and D. Kaup, “Perturbation theory for the Manakov soliton and its applications to pulse propagation in randomly birefringent fibers,” Phys. Rev. E 56(5), 6147–6165 (1997). [CrossRef]  

55. V. Shchesnovich and E. Doktorov, “Perturbation theory for solitons of the Manakov system,” Phys. Rev. E 55(6), 7626–7635 (1997). [CrossRef]  

56. D. Kaup, “A perturbation expansion for the Zakharov–Shabat inverse scattering transform,” SIAM J. Appl. Math. 31(1), 121–133 (1976). [CrossRef]  

57. V. E. Zakharov and S. V. Manakov, “Asymptotic behavior of non-linear wave systems integrated by the inverse scattering method,” Sov. Phys. JETP 44, 106–112 (1976).

58. P. Deift and X. Zhou, “A steepest descent method for oscillatory Riemann-Hilbert problems. Asymptotics for the MKdV equation,” Annals Math. 137(2), 295–368 (1993). [CrossRef]  

59. G. P. Agrawal, Fiber-optic communication systems, vol. 222 (John Wiley & Sons, 2012).

60. J. E. Prilepsky, S. A. Derevyanko, and S. K. Turitsyn, “Nonlinear Fourier Transform-Based Optical Transmission: Methods for Capacity Estimation,” in Optical Communication Systems: Limits and Possibilities, A. Ellis and M. Sorokina, eds. (Jenny Stanford Publishing, 2019), chap. 7, pp. 243–272.

61. X. Zhang, L. Xi, J. Wei, S. Du, W. Zhang, J. Li, and X. Zhang, “Nonlinear frequency domain PMD modeling and equalization for nonlinear frequency division multiplexing transmission,” Opt. Express 29(18), 28190–28201 (2021). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon a reasonable request.

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 (2)

Fig. 1.
Fig. 1. Normalized PSD for $b_1$ (a), $b_2$ (b), and cross-density (c), as the function of the total power factor $b$ for different values of inter-polarization power split ratio $\theta$.
Fig. 2.
Fig. 2. Normalized PSD for: $b_1$ (a), (d); $b_2$ (b), (e), and the real part of the cross-density (c), (f), as the function of the total power of NFDM burst. Top row: full simulation, bottom row: after the subtraction of processing noise.

Tables (1)

Tables Icon

Table 1. The parameters of the single-mode fiber link used for the numerical verification of our results.

Equations (47)

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

E l = i β 2 2 2 E T 2 + i γ 8 9 | | E | | 2 E + N ,
E [ N i ( l , T ) N j ( l , T ) ] = δ i j ( N A S E / L ) δ ( l l ) δ ( T T ) ,
γ γ e f f = γ [ 1 exp ( α L a ) ] / α L a ,
q 1 z = i 2 q 1 t 2 + 2 i ( | q 1 | 2 + | q 2 | 2 ) q 1 + n 1 , q 2 z = i 2 q 2 t 2 + 2 i ( | q 1 | 2 + | q 2 | 2 ) q 2 + n 2 , E [ n i ( z , t ) n j ( z , t ) ] = δ i j 2 D δ ( z z ) δ ( t t ) , 2 D = N A S E L L s P s T s ,
v t = ( i λ q 1 q 2 q 1 i λ 0 q 2 0 i λ ) v = M v ,
ϕ ( 1 0 0 ) e i λ t , ϕ ¯ ( 0 0 1 0 0 1 ) e i λ t , t , ψ ( 0 0 1 0 0 1 ) e i λ t , ψ ¯ ( 1 0 0 ) e i λ t , t + .
ϕ ( t , λ ) = ψ ( t , λ ) b ( λ ) + ψ ¯ ( t , λ ) a ( λ ) , ϕ ¯ ( t , λ ) = ψ ( t , λ ) a ¯ ( λ ) + ψ ¯ ( t , λ ) b ¯ ( λ ) ,
Φ = [ ϕ , ϕ ¯ ] , Ψ = [ ψ ¯ , ψ ] .
t ( Φ Φ ) = Φ t Φ + Φ Φ t = Φ ( M + M ) Φ = 0.
S = ( a b ¯ 1 b ¯ 2 b 1 a ¯ 11 a ¯ 12 b 2 a ¯ 21 a ¯ 22 ) .
Σ = { ρ ( λ ) = b ( λ ) a 1 ( λ ) , λ k , C k = b ( λ k ) / a ( λ k ) , a ( λ k ) = 0 } , Σ ¯ = { ρ ¯ ( λ ) = b ¯ ( λ ) a ¯ 1 ( λ ) , λ ¯ k , C k ¯ = b ¯ ( λ k ¯ ) α ¯ ( λ k ¯ ) / det a ¯ ( λ ¯ k ) , det a ¯ ( λ k ¯ ) = 0 } ,
a ¯ ( λ ) a ¯ ( λ ) + b ¯ ( λ ) b ¯ ( λ ) = I 2 , | a ( λ ) | 2 + b ( λ ) b ( λ ) = 1 , a ¯ ( λ ) b ( λ ) + b ¯ ( λ ) a ( λ ) = 0 .
a ( λ , z ) = a ( λ , 0 ) , a ¯ ( λ , z ) = a ¯ ( λ , 0 ) , b ( λ , z ) = b ( λ , 0 ) e 4 i λ 2 z , b ¯ ( λ , z ) = b ¯ ( λ , 0 ) e 4 i λ 2 z .
v ( T 1 ) = ( a ( λ ) e i λ T 1 b 1 ( λ ) e i λ T 1 b 2 ( λ ) e i λ T 1 ) .
u j ( λ ) = S k = 0 N s c 1 c k j ψ k ( λ ) , j = 1 , 2 ,
b i n , j ( λ ) = u j ( λ ) | u 1 ( λ ) | 2 + | u 2 ( λ ) | 2 1 e | u 1 ( λ ) | 2 | u 2 ( λ ) | 2 .
u t M u = f , where f = ( q 1 z ϕ 2 + q 2 z ϕ 3 q 1 z ϕ 1 q 2 z ϕ 1 ) .
Φ t ε + Φ ε t M Φ ε = Φ ε t = f , ε ( t ) = t Φ ( t ) f ( t ) d t ,
u ( + ) = ϕ z ( + ) = Ψ ( + ) c z = Φ ( + ) ε ( + ) = Φ ( + ) Φ ( t ) f ( t ) d t ,
c z = Ψ f d t ,
a z = { q 1 z ψ ¯ 1 ϕ 2 q 1 z ψ ¯ 2 ϕ 1 + q 2 z ψ ¯ 1 ϕ 3 q 2 z ψ ¯ 3 ϕ 1 } d t , b 1 z = { q 1 z ψ 11 ϕ 2 q 1 z ψ 21 ϕ 1 + q 2 z ψ 11 ϕ 3 q 2 z ψ 31 ϕ 1 } d t , b 2 z = { q 1 z ψ 12 ϕ 2 q 1 z ψ 22 ϕ 1 + q 2 z ψ 12 ϕ 3 q 2 z ψ 32 ϕ 1 } d t .
b j z = 4 i λ 2 b j 2 + { n 1 ψ 1 j ϕ 2 n 1 ψ 2 j ϕ 1 + n 2 ψ 1 j ϕ 3 n 2 ψ 3 j ϕ 1 } d t , j = 1 , 2.
a z = { n 1 ψ ¯ 1 ϕ 2 n 1 ψ ¯ 2 ϕ 1 + n 2 ψ ¯ 1 ϕ 3 n 2 ψ ¯ 3 ϕ 1 } d t .
b j z = 4 i λ 2 b j 2 + Γ j ( λ , z ) , j = 1 , 2 ,
E [ Γ ( z , λ ) Γ ( z , λ ) ] = 2 D δ ( z z ) ( A ¯ 1 ( λ , λ , z ) B ¯ ( λ , λ , z ) B ¯ ( λ , λ , z ) A ¯ 2 ( λ , λ , z ) ) = 2 D δ ( z z ) C ¯ ( λ , λ , z ) , E [ Γ ( z , λ ) Γ T ( z , λ ) ] = 2 D δ ( z z ) ( A 1 ( λ , λ , z ) B ( λ , λ , z ) B ( λ , λ , z ) A 2 ( λ , λ , z ) ) = 2 D δ ( z z ) C ( λ , λ , z ) ,
A j ( λ , λ , z ) = [ ψ 1 j ( λ ) ϕ 2 ( λ ) ψ 2 j ( λ ) ϕ 1 ( λ ) ψ 2 j ( λ ) ϕ 1 ( λ ) ψ 1 j ( λ ) ϕ 2 ( λ ) ψ 1 j ( λ ) ϕ 3 ( λ ) ψ 3 j ( λ ) ϕ 1 ( λ ) ψ 3 j ( λ ) ϕ 1 ( λ ) ψ 1 j ( λ ) ϕ 3 ( λ ) ] d t ,
A ¯ j ( λ , λ , z ) = [ ψ 1 j ( λ ) ϕ 2 ( λ ) ψ 1 j ( λ ) ϕ 2 ( λ ) + ψ 2 j ( λ ) ϕ 1 ( λ ) ψ 2 j ( λ ) ϕ 1 ( λ ) + ψ 1 j ( λ ) ϕ 3 ( λ ) ψ 1 j ( λ ) ϕ 3 ( λ ) + ψ 3 j ( λ ) ϕ 1 ( λ ) ψ 3 j ( λ ) ϕ 1 ( λ ) ] d t ,
B ( λ , λ , z ) = [ ψ 11 ( λ ) ϕ 2 ( λ ) ψ 22 ( λ ) ϕ 1 ( λ ) ψ 21 ( λ ) ϕ 1 ( λ ) ψ 12 ( λ ) ϕ 2 ( λ ) ψ 11 ( λ ) ϕ 3 ( λ ) ψ 32 ( λ ) ϕ 1 ( λ ) ψ 31 ( λ ) ϕ 1 ( λ ) ψ 12 ( λ ) ϕ 3 ( λ ) ] d t ,
B ¯ ( λ , λ , z ) = [ ψ 11 ( λ ) ϕ 2 ( λ ) ψ 12 ( λ ) ϕ 2 ( λ ) + ψ 21 ( λ ) ϕ 1 ( λ ) ψ 22 ( λ ) ϕ 1 ( λ ) + ψ 11 ( λ ) ϕ 3 ( λ ) ψ 12 ( λ ) ϕ 3 ( λ ) + ψ 31 ( λ ) ϕ 1 ( λ ) ψ 32 ( λ ) ϕ 1 ( λ ) ] d t .
b o u t ( λ ) = b i n ( λ ) + N ( λ ) , N j ( λ ) = e 4 i s λ 2 L 0 L e 4 i λ 2 z Γ j ( λ , z ) d z , j = 1 , 2 ,
E [ N ( λ ) N ( λ ) ] = 2 D e 4 i s ( λ 2 λ 2 ) L 0 L e 4 i s ( λ 2 λ 2 ) z C ¯ ( λ , λ , z ) d z , E [ N ( λ ) N T ( λ ) ] = 2 D e 4 i s ( λ 2 + λ 2 ) L 0 L e 4 i s ( λ 2 + λ 2 ) z C ( λ , λ , z ) d z .
H ( ω ) = d t H ( t ) e i ω t = i ω + π δ ( ω ) ,
ϕ 1 < ( λ , t ) = e i λ t , ϕ 2 , 3 < ( λ , t ) = 0 ,
ψ 11 > = ψ 12 > = ψ 31 > = ψ 22 > = 0 , ψ 21 > = ψ 32 > = e i λ t .
ϕ > ( λ , t ) = a ( λ ) ψ ¯ > ( λ , t ) + ψ > ( λ , t ) b ( λ ) , ϕ 1 > ( λ , t ) = a ( λ ) e i λ t , ϕ 2 , 3 > ( λ , t ) = b 1 , 2 ( λ ) e i λ t .
ψ < ( λ , t ) = ϕ < ( λ , t ) b ( λ ) + ϕ ¯ < ( λ , t ) a ¯ ( λ ) , ψ 1 j < ( λ , t ) = b j ( λ ) e i λ t , ψ 2 j < ( λ , t ) = a ¯ j 1 ( λ ) e i λ t , ψ 3 j < ( λ , t ) = a ¯ j 2 ( λ ) e i λ t , j = 1 , 2.
A j < ( λ , λ , z ) = 0 ,
A ¯ j < ( λ , λ , z ) = [ a ¯ j 1 ( λ ) a ¯ j 1 ( λ ) + a ¯ j 2 ( λ ) a ¯ j 2 ( λ ) ] e 2 i ( λ λ ) T 1 ( i 2 ( λ λ ) + π 2 δ ( λ λ ) ) ,
B < ( λ , λ , z ) = 0 ,
B ¯ j < ( λ , λ , z ) = [ a ¯ 11 ( λ ) a ¯ 21 ( λ ) + a ¯ 12 ( λ ) a ¯ 22 ( λ ) ] e 2 i ( λ λ ) T 1 ( i 2 ( λ λ ) + π 2 δ ( λ λ ) ) .
A j > ( λ , λ , z ) = 0 ,
A ¯ j > ( λ , λ , z ) = a ( λ ) a ( λ ) e 2 i ( λ λ ) T 2 [ i 2 ( λ λ ) + π 2 δ ( λ λ ) ] ,
B > ( λ , λ , z ) = B ¯ > ( λ , λ , z ) = 0.
C ¯ ( λ , λ , z ) = ( i = 1 2 | a ¯ 1 i ( λ ) | 2 + | a ( λ ) | 2 i = 1 2 a ¯ 1 i ( λ ) a ¯ 2 i ( λ ) i = 1 2 a ¯ 1 i ( λ ) a ¯ 2 i ( λ ) i = 1 2 | a ¯ 2 i ( λ ) | 2 + | a ( λ ) | 2 ) × π 2 δ ( λ λ ) .
E [ N ( λ ) N ( λ ) ] = n A S E B ( λ ) δ ( λ λ ) , E [ N ( λ ) N T ( λ ) ] = 0 ,
B = ( 1 | b 1 ( λ ) | 2 | b 2 ( λ ) | 2 / 2 1 2 b 1 ( λ ) b 2 ( λ ) 1 2 b 1 ( λ ) b 2 ( λ ) 1 | b 2 ( λ ) | 2 | b 1 ( λ ) | 2 / 2 ) ,
E [ N ( λ ) N ( λ ) ] = n A S E ( 1 | b 1 ( λ ) | 2 ) δ ( λ λ ) , E [ N ( λ ) N ( λ ) ] = 0 ,
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.