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

Guided acoustic Brillouin scattering measurements in optical communication fibers

Open Access Open Access

Abstract

Guided acoustic Brillouin (GAWBS) noise is measured using a novel, homodyne measurement technique for four commonly used fibers in long-distance optical transmission systems. The measurements are made with single spans and then shown to be consistent with separate multi-span long-distance measurements. The inverse dependence of the GAWBS noise on the fiber effective area is confirmed by comparing different fibers with the effective area varying between 80 µm2 and 150 µm2. The line broadening effect of the coating is observed, and the correlation between the width of the GAWBS peaks to the acoustic mode profile is confirmed. An extensive model of the GAWBS noise in long-distance fibers is presented, including corrections to some commonly repeated mistakes in previous reports. It is established through the model and verified with the measurements that the depolarized scattering caused by TR2m modes contributes twice as much to the optical noise in the orthogonal polarization to the original source, as it does to the noise in parallel polarization. Using this relationship, the polarized and depolarized contributions to the measured GAWBS noise is separated for the first time. As a result, a direct comparison between the theory and the measured GAWBS noise spectrum is shown for the first time with excellent agreement. It is confirmed that the total GAWBS noise can be calculated from fiber parameters under certain assumptions. It is predicted that the level of depolarized GAWBS noise created by the fiber may depend on the polarization diffusion length, and consequently, possible ways to reduce GAWBS noise are proposed. Using the developed theory, dependence of GAWBS noise on the location of the core is calculated to show that multi-core fibers would have a similar level of GAWBS noise no matter where their cores are positioned.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Optical transmission capacity is limited by several impairments that arise from the optical fibers including loss, dispersion, nonlinearity, polarization-mode dispersion, etc. Recently it has been showed that guided acoustic Brillouin scattering (GAWBS) induces signal-to-noise SNR degradation that is not negligible [1], and also that it cannot be removed or mitigated with digital signal processing effectively [2]. It is important to be able to estimate the capacity of a link at the design stage accurately, therefore it is important to know how much SNR degradation is expected from GAWBS [3]. The scattered light, which becomes the GAWBS noise in the context of optical communication, generated in short lengths of fibers are quite small, and difficult to measure [3]. On the other hand, after long distances, it is difficult to separate it from other noise sources such as fiber nonlinearity contributing to variations in the estimations of the magnitude of the GAWBS noise [1,2]. It was shown that GAWBS depends on the fiber type used, in particular the effective area of the fiber [4,5]. Even though there are a few modern fiber types that are commonly used for optical transmission systems, there has not been reports tabulating the GAWBS in these fibers. For the fibers that were reported, the GAWBS noise was not categorized in detail, for instance the level of noise generated in the parallel and orthogonal polarizations with respect to the original signal [1,5,6]. Since the experimental results were not compared to the theoretical expectations [1], or they were compared with wrong models [5,6], it was not possible to confirm the veracity of the measurements.

In this paper we report the level of GAWBS noise generated in 4 types of commonly used fibers, SMF28-ULL, Vascade 2000, Vascade 3000, Z+130ULL in both polarizations. A novel homodyne measurement technique is presented that can measure the GAWBS noise spectrum in separate polarizations from just a single span and not limited by instrument dynamic range. We also confirm that GAWBS noise scales linearly with transmission distance by comparing single span measurement with loop transmission up to trans-Pacific distance. Furthermore, the inverse dependence of GAWBS noise on the fiber effective area is confirmed [1,4,5]. We report excellent fitting of the measurements with the theory only after making necessary corrections to the model which was originally outlined in [7]. In particular, it is found that the total GAWBS noise and its spectral shape can be calculated quite accurately from fiber parameters. The relationship between the power level of the GAWBS noise in parallel and orthogonal polarizations with respect to the optical carrier is established through theory and verified by measurements for the first time. Through measurements, the correlation between the magnitude of the radial component of the acoustic displacement vector and the broadening of the GAWBS peaks is confirmed. Impact of polarization diffusion due to random fiber birefringence on the depolarized GAWBS noise is determined and a method for reducing GAWBS noise is proposed. Dependence of GAWBS noise on the location of the core within the cladding is calculated to estimate GAWBS noise in multicore fibers. We note that even though for optical communications, being able to distinguish the polarization nature of the GAWBS noise does not make a direct impact, being able to fit the polarization properties of GAWBS to the theoretical expectation improves the confidence in the measurement accuracy.

2. Theory

GAWBS is generated by transverse acoustic modes in the fiber. These acoustic modes are ever present as they are due to the ambient temperature [79]. These modes do not have appreciable longitudinal components, which means they only generate forward scattering. The theoretical treatment follows that of [7,8]. The mode field distribution of the acoustic modes are calculated assuming a cylindrical glass with infinite length, and a free surface, i.e., the polymer coating is ignored. From the acoustic mode distribution, the induced strain is obtained in the form of a tensor. At this point our approach deviates from that of [7]. The strain is related to the impermeability tensor through the photoelastic tensor of fused silica. Both phase and birefringence modulation caused by the acoustic modes are calculated using coupled-mode which leads to the correct fitting of the experimental results. Note that instead of coupled-mode theory it is also possible to use other techniques such as perturbation or variational approaches which leads to the similar conclusions as ours [10]. A second notable deviation from the Shelby’s derivation [7] is in regards to the dependence of GAWBS noise on fiber length for long fibers. Here it will be argued that the linear dependence on the fiber length originates from the fact that optical mode interacts with many uncorrelated acoustic modes with a finite coherence length. Finally the acoustic mode groups that significantly contribute to the GAWBS noise in fibers with a concentric core are identified and the characteristics of the GAWBS noise generated by each mode group is analyzed.

The acoustic modes giving rise to GAWBS can be described by the displacement vector field components [79]

$$ \begin{aligned} U_r\left(r,\phi \right) &= C_{nm}\left\{ A_1 \left\lbrack J_{n+1}\left(\rho\right) +J_{n-1}\left(\rho\right) \right\rbrack + \alpha A_2 \left\lbrack J_{n+1}\left(\alpha \rho\right) - J_{n-1}\left(\alpha \rho\right) \right\rbrack \right\} \Theta_d \left( n \phi \right) , \\ U_{\phi}\left(r,\phi \right) &= C_{nm}\left\{ A_1 \left\lbrack J_{n+1}\left(\rho\right) - J_{n-1}\left(\rho\right) \right\rbrack + \alpha A_2 \left\lbrack J_{n+1}\left(\alpha \rho\right) + J_{n-1}\left(\alpha \rho\right) \right\rbrack \right\} \Psi_d \left( n \phi \right) , \end{aligned}$$
where $\overrightarrow {U} = \left \lbrack U_r, U_{\phi },0 \right \rbrack \sin \left ( \Omega _{nm}t + \zeta \right )$ is the displacement vector field in cylindrical coordinates defined by radial and angular coordinates $r$ and $\phi$ respectively, $C_{nm}$ is the amplitude of the acoustic mode vibrating with the frequency $f_{nm} = \Omega _{nm}/2\pi$, $\rho =\Omega _{nm} r/V_s$ is the normalized radial coordinate, $V_s$ is the speed of the shear sound waves, $\alpha = V_s/V_d$ is the ratio of the speed of shear and longitudinal acoustic waves. The vibration frequencies $f_{nm}$ can be obtained from the solutions of the characteristic equation obtained by requiring traction-free boundary conditions at the glass cladding surface $r = a$. The characteristic equation can be expressed in the form
$$\begin{vmatrix} B_{11} & B_{12} \\ B_{21} & B_{22} \end{vmatrix} =0 ,$$
where $B_{ij}$ are the components of the matrix
$$B = \begin{bmatrix} \left(n^2-1- \frac{y^2}{2} \right) J_n \left( \alpha y \right) & \left(n\left(n^2-1\right)- \frac{y^2}{2} \right) J_n \left( y \right) - \left(n^2-1\right) y J_{n+1}\left( y \right) \\ \left(n-1\right) J_n \left( \alpha y \right) - \alpha y J_{n+1} \left( \alpha y \right) & \left(n\left(n^2-1\right)- \frac{y^2}{2} \right) J_n \left( y \right)+ y J_{n+1}\left( y \right) \end{bmatrix} ,$$
with
$$y = \frac{\Omega a}{V_s} .$$
Equation (3) accepts multiple solutions $\Omega _{nm}$ for mode groups $n$, with $n,m \geq 0$ integers. In Eq. (1), $A_1 = n B_{11}$, $A_2 = B_{12}$. The acoustic waves are doubly degenerate in their angular dependence and without loss of generality, they can be expressed as follows
$$ \begin{aligned} \Theta_d \left(n \phi \right) = \left\{ \begin{array}{c} \cos\left(n \phi \right), \quad d = 0 \\ \sin\left(n \phi \right), \quad d = 1 \\ \end{array} \right. , \qquad \Psi_d \left(n \phi \right) = \left\{ \begin{array}{c} \sin\left(n \phi \right), \quad d = 0 \\ -\cos\left(n \phi \right), \quad d = 1 \\ \end{array} \right. \end{aligned} .$$
In Eq. (5), the upper or lower solution or any combination of them can be chosen in general.

Displacement caused by the acoustic vibrations induces strain in the fiber. The modulating strain in return causes phase and birefringence modulations which results in generation of optical side bands at the same frequency. The induced strain tensor in cylindrical coordinates can be obtained using the following relations:

$$ S_{rr} = \frac{\partial U_r}{\partial r}, \quad S_{\phi \phi} = \frac{1}{r} \frac{\partial U_{\phi}}{\partial \phi} + \frac{U_r}{r}, \quad S_{r \phi} = \frac{1}{2} \left( \frac{1}{r} \frac{\partial U_r}{\partial \phi}+ \frac{\partial U_{\phi}}{\partial r} - \frac{U_{\phi}}{r} \right) , $$
and the strain components in the z–axis is zero. Inserting Eq. (1) into Eq. (6) we get:
$$ \begin{aligned} S_{rr} &= C_{nm}\frac{\Omega_{nm}}{ V_s} s_{rr}(r) \Theta \left( n \phi \right) \sin\left( \Omega_{nm}t + \zeta \right), \\ S_{\phi\phi} & = C_{nm}\frac{\Omega_{nm}}{ V_s} s_{\phi\phi}(r) \Theta \left( n \phi \right) \sin\left( \Omega_{nm}t + \zeta \right), \\ S_{r \phi} & = C_{nm}\frac{\Omega_{nm}}{ V_s} s_{r \phi} (r) \Psi \left( n \phi \right) \sin\left( \Omega_{nm}t + \zeta \right) , \end{aligned}$$
where
$$ \begin{aligned} s_{rr} &= \frac{1}{2}\left\{ A_1 \left\lbrack J_{n-2}\left(\rho\right) -J_{n+2}\left(\rho\right) \right\rbrack + \alpha^2 A_2 \left\lbrack 2 J_{n}\left(\alpha \rho\right) - J_{n-2}\left(\alpha \rho\right) - J_{n+2}\left(\alpha \rho\right) \right\rbrack \right\} ,\\ s_{\phi\phi} &=\frac{1}{\rho}\left\{ A_1 \left\lbrack (n+1)J_{n+1}\left(\rho\right) -(n-1)J_{n-1}\left(\rho\right) \right\rbrack \right. \\ & \left. + \alpha A_2 \left\lbrack (n+1) J_{n+1}\left(\alpha \rho\right) + (n-1)J_{n-1}\left(\alpha \rho\right) \right\rbrack \right\} , \\ s_{r \phi} &= \frac{1}{2\rho}\left\{ -A_1 \left\lbrack (n+1)J_{n+1}\left(\rho\right) +(n-1)J_{n-1}\left(\rho\right) + \frac{\rho}{2} \left\lbrack 2 J_{n}\left(\rho\right) - J_{n+2}\left(\rho\right) - J_{n- 2}\left(\rho\right) \right\rbrack \right\rbrack \right. \\ & + \left. \alpha A_2 \left\lbrack (n-1) J_{n-1}\left(\alpha \rho\right) - (n+1)J_{n+1}\left(\alpha \rho\right)+ \frac{\alpha\rho}{2} \left\lbrack J_{n-2}\left(\alpha\rho\right) - J_{n+ 2}\left(\alpha\rho\right) \right\rbrack \right\rbrack \right\} . \end{aligned}$$
It should be noted that the angular the strain tensor components $S_{rr}$ and $S_{\phi \phi }$ has the same angular dependence as the radial component of the displacement vector $U_r$ and the strain tensor component $S_{r\phi }$ has the same angular dependence as $U_{\phi }$.

So far Shelby’s derivation was followed to calculate the strain from the acoustic modes [7]. The next step is to estimate how the strain is converted to optical modulation which will be presented in more detail. As stated, the variations in the impermeability tensor $\eta _{ij}$ can be obtained from the strain tensor through photoelastic tensor in the reduced notation in Cartesian coordinates as follows [11]:

$$\begin{bmatrix} \Delta\eta_{xx}\left(r,\phi \right) \\ \Delta\eta_{yy}\left(r,\phi \right) \\ \Delta\eta_{zz}\left(r,\phi \right) \\ \Delta\eta_{yz}\left(r,\phi \right) \\ \Delta\eta_{xz}\left(r,\phi \right) \\ \Delta\eta_{xy}\left(r,\phi \right) \end{bmatrix} = \begin{bmatrix} p_{11} & p_{12} & p_{12} & 0 & 0 & 0 \\ p_{12} & p_{11} & p_{12} & 0 & 0 & 0 \\ p_{12} & p_{12} & p_{11} & 0 & 0 & 0 \\ 0 & 0 & 0 & p_{11} - p_{12} & 0 & 0 \\ 0 & 0 & 0 & 0 & p_{11} - p_{12} & 0 \\ 0 & 0 & 0 & 0 & 0 & p_{11} - p_{12} \end{bmatrix} \begin{bmatrix} S_{xx}\left(r,\phi \right) \\ S_{yy}\left(r,\phi \right) \\ S_{zz}\left(r,\phi \right) \\ S_{yz}\left(r,\phi \right) \\ S_{xz}\left(r,\phi \right) \\ S_{xy}\left(r,\phi \right) \end{bmatrix}.$$
The photoelastic tensor in Eq. (9) takes a particularly simple form as fused silica is well approximated as an isotropic solid. Impermeability tensor is related to the dielectric permittivity as $\pmb {\eta } = \epsilon _0 \pmb {\epsilon }^{-1}$ where $\epsilon _0$ is the free space permittivity [12]. Assuming that the variations in dielectric permittivity due to GAWBS $\Delta \pmb {\epsilon }$ is a small perturbation on the undisturbed permittivity $\epsilon$ of the fiber which is assumed to be close to isotropic, i.e., $\pmb {\epsilon } \approx \epsilon + \Delta \pmb {\epsilon }$ and $\pmb {\eta } \approx 1/\epsilon + \Delta \pmb {\eta }$
$$\Delta\eta_{ij}\left(r,\phi,t \right) \approx{-}\epsilon_0 \frac{\Delta \epsilon_{ij} \left( r,\phi,t \right) }{\epsilon^2} \implies \Delta \epsilon_{ij} \approx{-}\Delta\eta_{ij}\left(r,\phi,t \right) n_0^4 \epsilon_0$$
where $n_0 = \sqrt {\epsilon /\epsilon _0}$ is the refractive index of the unperturbed fiber. Aligning the x-axis with $\phi =0$, the relevant strain tensor components can be converted from the cylindrical coordinates to the Cartesian coordinates as follows:
$$ \begin{aligned} S_{xx} &= \cos\left(\phi \right)^2 S_{rr} + \sin\left(\phi \right)^2 S_{\phi\phi} - \sin\left(2\phi \right) S_{r\phi} , \\ S_{yy} &= \sin\left(\phi \right)^2 S_{rr} + \cos\left(\phi \right)^2 S_{\phi\phi} + \sin\left(2\phi \right) S_{r\phi} , \\ S_{xy} &= \sin\left(2\phi \right) \left( S_{rr} - S_{\phi\phi} \right)/2 + \cos\left(2\phi \right)S_{r\phi}, \end{aligned}$$
Combining Eqs. (11) and (9) we obtain
$$ \begin{aligned} \Delta\eta_{xx}(r,\phi,t) &= p_p\left( S_{rr} +S_{\phi\phi} \right) + p_m \left\lbrack \left( S_{rr} -S_{\phi\phi} \right)\cos\left(2\phi \right) - 2S_{r\phi}\sin\left(2\phi \right) \right\rbrack ,\\ \Delta\eta_{yy}(r,\phi,t) &= p_p\left( S_{rr} +S_{\phi\phi} \right) - p_m \left\lbrack \left( S_{rr} -S_{\phi\phi} \right)\cos\left(2\phi \right) - 2S_{r\phi}\sin\left(2\phi \right) \right\rbrack ,\\ \Delta\eta_{xy}(r,\phi,t) &= p_m \left\lbrack \left( S_{rr} -S_{\phi\phi} \right)\sin\left(2\phi \right) + 2S_{r\phi}\cos\left(2\phi \right) \right\rbrack , \end{aligned}$$
where $p_p = \left ( p_{11} + p_{12} \right )/2$, and $p_m = \left ( p_{11} - p_{12} \right )/2$.

Equation (12) describes how the impermeability tensor varies across the fiber cross section due to the acoustic vibrations. In general, such spatial variations in the refractive index distribution would scatter the signal traveling in the fiber into all the modes supported by the optical fiber and radiation modes. In the framework of coupled-mode theory [1316], when the modulations are very small, as in the case of GAWBS, power lost to these scattering would be negligible, and not of interest to us. The non-negligible contribution would be the optical signal coupling into the modes supported by the fiber in the absence of acoustic vibrations. The transverse part of the electric field in the fiber can be expanded in terms of the transverse components of the electric fields of the modes of the undisturbed fiber as follows [13]

$$\vec{E_t}\left(r,\phi,z,t \right) = \sum_{\nu} h_{\nu}\left( z \right) \vec{E}_{\nu,t}\left(r,\phi,t \right).$$
The subscript $t$ denotes the transverse component. The variation of the electric field along the fiber is contained in $h_{\nu }\left ( z \right )$ which is governed by the coupled mode equation given by
$$\frac{\partial h_{\mu}\left(z\right)}{\partial z} ={-}\frac{\alpha}{2}h_{\mu}+ i k_{\mu}h_{\mu} + i \sum_{\nu}\kappa_{\mu\nu}\left( t \right) h_{\nu}\left(z\right), \quad \quad i = \sqrt{-1}$$
where $\alpha$ is the fiber attenuation coefficient, $k_{\mu }$ is the propagation constant of the mode $\mu$ in the absence of GAWBS. The coupling coefficients $\kappa _{\mu \nu }$ caused by GAWBS are given by [13]
$$\kappa_{\mu\nu} ={-}i\frac{\omega}{4P} \int_{0}^{2\pi}\int_{0}^{\infty} \vec{E}^*_{\mu}\left(r,\phi,t \right) \Delta\pmb{\epsilon}\left( r,\phi,t \right) \vec{E}_{\nu}\left(r,\phi,t \right) r \,\mathrm{d}r \mathrm{d}\phi$$
where $\omega$ is the optical carrier frequency, P is used as the normalization factor for the amplitudes of the coupling modes defined as
$$P \delta_{\mu\nu}=\frac{1}{4} \int_{0}^{2\pi}\int_{0}^{\infty}\hat{e}_z\cdot\left\lbrack \vec{E}_{\nu,t} \times \vec{H}^*_{\mu,t} +\vec{E}^*_{\nu,t} \times \vec{H}_{\mu,t} \right\rbrack r \,\mathrm{d}r \mathrm{d}\phi$$
and $\delta _{\mu \nu }$ is the Kronecker delta function. A minor detail to note is that, in Eq. (15) the optical mode field terms are no longer only transverse components. However, for the case of GAWBS this small detail is has no consequence as $\Delta \epsilon _{rz}$, $\Delta \epsilon _{\phi z}$, and $\Delta \epsilon _{zz}$ are assumed to be zero.

Under the weakly guiding approximation which holds for long-distance optical fibers, the LP modes approximate the electrical field very well satisfying

$$\vec{H}_{\mu,t} = c n_0 \epsilon_0 \hat{e}_z \times \vec{E}_{\mu,t}$$
where $c$ is the speed of light in vacuum. Combining Eqs. (10), (16), and (17) the mode coupling equation Eq. (15) simplifies to
$$\kappa_{\mu\nu} ={-}i\frac{k_0 n_0^3}{2} \frac{\int_{0}^{2\pi}\int_{0}^{\infty} \vec{E}^*_{\mu}\left(r,\phi,t \right) \Delta\pmb{\eta}\left( r,\phi,t \right) \vec{E}_{\nu}\left(r,\phi,t \right) r \,\mathrm{d}r \mathrm{d}\phi } {\int_{0}^{2\pi}\int_{0}^{\infty} \vec{E}^*_{\mu,t}\left(r,\phi,t \right) \cdot \vec{E}_{\nu,t}\left(r,\phi,t \right) r \,\mathrm{d}r \mathrm{d}\phi},$$
with $k_0 = \omega /c$.

So far the Eqs. (1)–(16) are quite general in terms of the optical mode profile. For instance they can be used to estimate the scattering efficiency for single-mode, multimode or multicore fibers. From here on, the analysis will be limited to the special case of single-mode fibers. As single-mode fibers support two polarization modes, the coupled mode Eq. (14) reduces to a pair of equations that can be written in the Jones vector notation as follows [17]:

$$ \begin{aligned} \frac{\partial| h(z) \rangle}{\partial z} &= -\frac{\alpha}{2}|h(z)\rangle + i \bar{k} |h(z)\rangle + i\Delta k \sigma_1|h(z)\rangle + i\bar{\kappa}(t) |h(z)\rangle + i \overrightarrow{\Delta\kappa}(t) \cdot\overrightarrow{\sigma} |h(z)\rangle, \\ | h(z) \rangle &= \begin{bmatrix} h_{x}(z) \\ h_{y}(z) \end{bmatrix}, \quad \sigma_1= \begin{bmatrix} 1 & 0 \\ 0 & -1 \end{bmatrix}, \, \sigma_2= \begin{bmatrix} 0 & 1 \\ 1 & 0 \end{bmatrix}, \, \sigma_3= \begin{bmatrix} 0 & -i \\ i & 0 \end{bmatrix} , \end{aligned} $$
where $\bar {k} = (k_x + k_y)/2$ is the average propagation constant, $\Delta {k} = (k_x - k_y)/2$ is the fiber birefringence in the absence of GAWBS, $\bar {\kappa } = (\kappa _{xx} + \kappa _{yy})/2$ and $\overrightarrow {\Delta \kappa } = [(\kappa _{xx} - \kappa _{yy})/2, \kappa _{xy},0]$ are the isotropic and anisotropic part of the coupling coefficients induced by the GAWBS, $\sigma _1$, $\sigma _2$, $\sigma _3$ are the 2x2 Pauli matrices, and $\overrightarrow {\sigma }=[\sigma _1, \sigma _2 ,\sigma _3]$. The “$\cdot$" stands for the dot product, yielding $\overrightarrow {\Delta \kappa }(t) \cdot \overrightarrow {\sigma } = \Delta \kappa _1 \sigma _1 + \Delta \kappa _2 \sigma _2 + \Delta \kappa _3 \sigma _3$ [17]. The isotropic part of the GAWBS induces pure phase modulation whereas the anisotropic portion induces birefringence with a birefringence vector given by $\overrightarrow {\Delta \kappa }$ which will be referred to as GAWBS birefringence vector. In the absence of fiber birefringence, in Stokes space this rotation due to GAWBS can be expressed as:
$$\frac{\partial \overrightarrow{h}}{\partial z} ={-} 2\overrightarrow{\Delta\kappa} \times \overrightarrow{h} ,$$
where $\overrightarrow {h} = \langle h | \overrightarrow {\sigma } | h\rangle$ is the Stokes vector representation of the Jones vector $|h \rangle$ [17] and “$\times$" stands for cross product. Equation (20) shows that GAWBS induces maximum rate of rotation of the optical field when the Stokes vector of the optical field $\overrightarrow {h}$ is orthogonal to $\overrightarrow {\Delta \kappa }$. It also shows that there is no rotation when these two vectors are either parallel or anti-parallel in Stokes space, corresponding to parallel and orthogonal in Jones Space. This can be easily shown by noting that the Jones vector $|h\rangle$ is the eigenvector of the birefringence matrix as $\overrightarrow {\Delta \kappa }.\overrightarrow {\sigma } |h\rangle = \pm |\Delta \kappa ||h\rangle$ [17]. Therefore, per Eq. (19), the birefringence term does not induce rotation but pure phase modulation. For any case in between these extremes, the birefringence terms will induce both phase modulation and polarization rotation depending on the angle between Stokes vectors $\overrightarrow {h}$ and $\overrightarrow {\Delta \kappa }$. It should be noted that the x- and the y-axis are defined with respect to the angle $\phi$ in Eq. (1) that determines the orientation of the angular features of the acoustic waves. For the sake of simplicity, in the rest of the paper, it will be assumed that the Jones vector $| h(z) \rangle$ is normalized such that its magnitude is equal to the optical power, i.e., $P = \langle h| h \rangle = \sqrt {\overrightarrow {h}\cdot \overrightarrow {h}}$.

Given a short piece of fiber with known birefringence and GAWBS parameters, Eq. (19) is well defined and it can be integrated. So far the the variation of the acoustic modes along the fiber was ignored. In fact, $U_z$ is explicitly set to zero, and it is also implicitly assumed that the acoustic mode oscillates coherently along the entire fiber length as $\sin \left ( \Omega _{nm}t + \zeta \right )$ with a fixed phase of $\zeta$. These are only approximations that are justified by the fact that the acoustic modes that contribute to GAWBS extend much longer in the fiber compared to the fiber diameter. In reality, it is expected that the GAWBS modes separated by long distances in the fiber would be oscillating with independent and random phases. Here this random variation along the length will be modeled as the fiber consisting of short lengths that are much longer than the fiber diameter but short enough that the acoustic mode would be completely coherent within. At the same time it will be assumed that the acoustic modes would be completely independent from the modes in other sections. In other words, $\zeta$ will be treated as a random variable uniformly distributed between $(-\pi ,\pi )$ that is constant within the mode coherence length $l_c$, but completely independent between different sections satisfying

$$ \langle \zeta(z_1) \zeta(z_2) \rangle = \left\{ \begin{array}{c} \frac{\pi^2}{3}, \quad |z_1-z_2| \le l_c \\ 0, \quad |z_1-z_2| > l_c \\ \end{array} \right.$$
Incorporating random variations in the acoustic mode phase and the birefringence along the fiber, Eq. (19) can be put in the following stochastic form
$$ \frac{\partial| h(z) \rangle}{\partial z} = \left( -\frac{\alpha}{2} + i \bar{k} + i\overrightarrow{b}(z)\cdot\overrightarrow{\sigma} \right)|h(z)\rangle + \left( i\bar{\kappa} + i \overrightarrow{\Delta\kappa} \cdot\overrightarrow{\sigma} \right)|h(z)\rangle \sin\left( \Omega_{nm}t + \zeta(z) \right) ,$$
where the vector $\overrightarrow {b}$ is the random fiber birefringence vector that randomly rotates the polarization. Since GAWBS is a very weak effect, the optical field propagates largely unaffected by the scattering therefore a perturbative formal solution of the following form can be sought
$$|h(z)\rangle = e^{-\frac{\alpha z}{2}+i\bar{k}z}U(z)\left(|h_0\rangle + |\delta h(z)\rangle \right), \quad \frac{\partial U(z)}{\partial z} = i\overrightarrow{b}\cdot\overrightarrow{\sigma}U(z) ,$$
which amounts to adopting the reference frame of optical field where $|h_0\rangle$ is the undepleted optical carrier, $|\delta h\rangle$ is the scattered field generated by GAWBS, $U(z)$ is the unitary matrix describing the rotation of the field due to fiber birefringence. Inserting the formal solution in Eq. (23) into Eq. (22) and ignoring the smaller terms, the equation for the GAWBS field can be expressed as
$$ \begin{aligned} \frac{\partial |h_0\rangle}{\partial z} &\approx 0, \\ \frac{\partial |\delta h(z)\rangle}{\partial z} &\approx i\left(\bar{\kappa} + \overrightarrow{\Delta\kappa}(z)\cdot\overrightarrow{\sigma}' \right) |h_0 \rangle \sin\left( \Omega_{nm}t + \zeta(z) \right), \end{aligned}$$
where $\overrightarrow {\Delta \kappa }'(z)\cdot \overrightarrow {\sigma } = U(z) \overrightarrow {\Delta \kappa }(z)\cdot \overrightarrow {\sigma } U^\dagger (z)$. As the matrix $U(z)$ is a random unitary matrix, $\overrightarrow {\Delta \kappa }'(z)$ is a randomly rotated version of $\overrightarrow {\Delta \kappa }$ at a distance z [17]. Since the equation is expressed in the rotating reference frame of the optical carrier due to the fiber birefringence the GAWBS term is rotating with the same statistics. This allows to describe this random rotation in terms of the polarization diffusion length $l_p$ defined as the distance the where the average of Stokes vector over an ensemble decays by a factor of natural logarithm, i.e., $\langle \overrightarrow {\Delta \kappa }'(z) \rangle = \langle \overrightarrow {\Delta \kappa }'(0) \rangle \exp (-z/l_p)$. However, for the sake of simplicity, the polarization rotation will be modeled similar to the case of acoustic mode phase variation where $\overrightarrow {\Delta \kappa }'(z)$ will be assumed to be constant over a section of length $l_p$ and completely random orientation at a different section, i.e., $\langle \overrightarrow {\Delta \kappa }'(z_1) \cdot \overrightarrow {\Delta \kappa }'(z_2) \rangle = \Delta \kappa ^2$ for $|z_1 -z_2| \le l_p$ and $\langle \overrightarrow {\Delta \kappa }'(z_1)\cdot \overrightarrow {\Delta \kappa }'(z_2) \rangle = 0$ for $|z_1 -z_2| > l_p$. For typical transmission fibers that are nominally non-birefringent, the polarization coherence length is on the order of 10–100 m [18].

After a distance $L$ that is much larger than both correlation lengths $l_c$, and $l_p$ the GAWBS field can be approximated as a discrete sum in the following form

$$|\delta h(L)\rangle = \Delta z\sum_{q=1}^{N} i\left( \bar{\kappa} + \overrightarrow{\Delta\kappa}'_q\cdot \overrightarrow{\sigma} \right)|h_0\rangle \sin\left( \Omega_{nm}t + \zeta_q \right) ,$$
where $\Delta z = L/N$. The GAWBS noise power defined as $P_G(L) = \langle \langle \delta h(L) |\delta h(L) \rangle \rangle$ where the outer brackets stand for combined ensemble and time average can be expressed as
$$\begin{aligned} P_G(L) &= \Delta z^2\sum_{q=1}^{N}\sum_{p=1}^{N} \left[ \bar{\kappa}^2 P_0 + \langle \overrightarrow{\Delta\kappa}'_q \cdot \overrightarrow{\Delta\kappa}'_p \rangle P_0 + \bar{\kappa} \langle \left(\overrightarrow{\Delta\kappa}'_q + \overrightarrow{\Delta\kappa}'_p \right) \cdot \overrightarrow{P_0} \rangle \right. \\ &\left. + i \langle \left( \overrightarrow{\Delta\kappa}'_q \times \overrightarrow{\Delta\kappa}'_p \right)\cdot \overrightarrow{P_0} \rangle \right] \langle \sin\left( \Omega_{nm}t + \zeta_q \right) \sin\left( \Omega_{nm}t + \zeta_p \right) \rangle , \end{aligned}$$
where $P_0 = \langle \langle h_0 | h_0 \rangle \rangle$, $\overrightarrow {P_0} = \langle h_0 |\overrightarrow {\sigma } |h_0 \rangle$ and it is assumed that the variation in the fiber birefringence and the acoustic mode phases are independent processes. In deriving Eq. (26) the identity $(\overrightarrow {a}\cdot \overrightarrow {\sigma })(\overrightarrow {b}\cdot \overrightarrow {\sigma }) = \overrightarrow {a}\cdot \overrightarrow {b}I + (\overrightarrow {a}\times \overrightarrow {b})\cdot \overrightarrow {\sigma }$ is used, where $I$ is 2x2 identity matrix [17]. The third and fourth terms in the summation in Eq. (26) would vanish as the average of the dot product between GAWBS birefringence vector and the Stokes vector of the optical field would average to zero. The first term is independent of rotation due to fiber birefringence and its accumulation would only depend on the correlation length of the acoustic mode phase variation, but the second term would depend on both the phase variation and also on the polarization rotations. In the two extreme cases where either $l_c \ll l_p$, or $l_p \ll l_c$ Eq. (26) simplifies to
$$ P_G(L) = \left\{ \begin{array}{c} \frac{P_0L}{2} \left(\bar{\kappa}^2 + \Delta \kappa^2 \right)l_c , \quad l_c \ll l_p \\ \frac{P_0L}{2} \left(\bar{\kappa}^2 + \Delta \kappa^2 \frac{l_p}{l_c} \right)l_c , \quad l_c \gg l_p \\ \end{array} \right. .$$
Equation (27) shows that the GAWBS noise power grows linearly with fiber length $L$ as expected. This linear dependence was forced by Shelby by arguing dimensional consistency requirement [7]. Here we arrive at the same conclusion as a consequence of the short correlation length of the acoustic modes responsible for the GAWBS scattering compared to the lengths of interest, i.e. $l_c \ll L$. However, we now have the $l_c$ term entering in the expression that affects the magnitude of the GAWBS noise power. Within $l_c$, the noise power grows quadratically with distance, but the noise power from each section add linearly. Therefore, larger the $l_c$ larger the noise power for the same total length. However, it will be shown later that under the assumptions that GAWBS is generated by ambient temperature and that $l_c$ is equivalent to the average acoustic mode length, the dependence on $l_c$ will be eliminated. What is new is that when $l_p \ll l_c$, effective correlation length for the growth of the GAWBS noise due to the birefringence term reduces to $l_p$ instead of $l_c$. It will be shown based on measurements that for typical transmission fibers $l_c \ll l_p$ holds.

When measuring GAWBS, the polarization direction of the optical carrier makes the most natural reference frame. In this reference frame, the GAWBS noise can be separated into two parts; the part that has the same polarization as the carrier, and the part that is orthogonal. These two will be referred to as the parallel and orthogonal GAWBS noise given by $P_{G\parallel } = \langle |\langle \hat {e}_{\parallel }|\delta h \rangle |^2 \rangle$ and $P_{G\perp }= \langle |\langle \hat {e}_{\perp }|\delta h \rangle |^2 \rangle$, respectively, where $|\hat {e}_{\parallel } \rangle$ and $|\hat {e}_{\perp } \rangle$ are the unit Jones vectors that are parallel and perpendicular to the carrier, i.e., $|\hat {e}_{\parallel } \rangle = |\ h_0 \rangle / \sqrt {P_0}$ and $\langle \hat {e}_{\perp }| \hat {e}_{\parallel }\rangle =0$ with $\langle \hat {e}_{\perp }| \hat {e}_{\perp }\rangle =1$. An expression for the parallel and orthogonal GAWBS noise power after a distance $L$ can be obtained similar to Eq. (27) under the same assumptions as

$$ \begin{aligned} P_{G\parallel}(L) &= \Delta z^2P_0\sum_{q=1}^{N}\sum_{p=1}^{N} \left\{ \left[\bar{\kappa}^2 + \langle \left( \overrightarrow{\Delta\kappa}'_q\cdot\hat{p} \right)\left( \overrightarrow{\Delta\kappa}'_p\cdot\hat{p} \right) \rangle + \bar{\kappa} \langle\left(\overrightarrow{\Delta\kappa}'_q +\overrightarrow{\Delta\kappa}'_p \right)\cdot\hat{p} \rangle \right] \right.\\ &\left. \langle \sin\left( \Omega_{nm}t + \zeta_q \right) \sin\left( \Omega_{nm}t + \zeta_p \right) \rangle \right\} , \\ P_{G\perp}(L) &= \Delta z^2P_0\sum_{q=1}^{N}\sum_{p=1}^{N} \left\{ \left[\langle \left( \overrightarrow{\Delta\kappa}'_q\cdot \overrightarrow{\Delta\kappa}'_p \right) \rangle - \langle \left( \overrightarrow{\Delta\kappa}'_q\cdot\hat{p} \right)\left( \overrightarrow{\Delta\kappa}'_p\cdot\hat{p} \right) \rangle + i\langle\left(\overrightarrow{\Delta\kappa}'_q \times \overrightarrow{\Delta\kappa}'_p \right)\cdot\hat{p} \rangle \right] \right.\\ &\left. \langle \sin\left( \Omega_{nm}t + \zeta_q \right) \sin\left( \Omega_{nm}t + \zeta_p \right) \rangle \right\}. \end{aligned}$$
In deriving Eq. (29) the identity $|\hat {e}_{\parallel } \rangle \langle \hat {e}_{\parallel } |$ $= (I + \hat {p}\cdot \overrightarrow {\sigma } )/2$ was used, where $\hat {p} = \overrightarrow {P_0}/P_0$ is the unit Stokes vector along the Stokes vector of the carrier $\overrightarrow {P_0}$, and the fact that orthogonal Jones vectors are anti-parallel in the Stokes space, i.e., $\hat {p}_{\perp } = -\hat {p}$, and $\hat {p}_{\perp } = \langle \hat {e}_{\perp } |\overrightarrow {\sigma } | \hat {e}_{\perp } \rangle$ [17].

As expected the isotropic term only contributes to the parallel GAWBS and none to the orthogonal GAWBS, where as the anisotropic term contributes to both $P_{G\parallel }$ and $P_{G\perp }$. The new non-vanishing term that appears in Eq. (29) is $\langle \left ( \overrightarrow {\Delta \kappa }'_q\cdot \hat {p} \right )\left ( \overrightarrow {\Delta \kappa }'_p\cdot \hat {p} \right ) \rangle$, which can alternatively be expressed as $\Delta \kappa ^2\langle \cos \left ( \theta _q \right )\cos \left ( \theta _p \right )\rangle$ where $\theta$ is the angle between the GAWBS birefringence vector $\overrightarrow {\Delta \kappa }$ and $\hat {p}$. Using the case of $l_c \ll l_p$ as an example, and setting $\Delta z = l_c$, Eq. (29) can be expressed as

$$ \begin{aligned} P_{G\parallel}(L) &= l_c^2 P_0\sum_{q=1}^{N}\sum_{p=1}^{N} \left\{ \left[\bar{\kappa}^2 +\Delta\kappa^2\langle\cos\left( \theta_q \right)\cos\left( \theta_p \right)\rangle + \bar{\kappa} \Delta\kappa \langle\cos\left( \theta_q \right) + \cos\left( \theta_p \right) \rangle \right] \right.\\ &\left. \langle \sin\left( \Omega_{nm}t + \zeta_q \right) \sin\left( \Omega_{nm}t + \zeta_p \right) \rangle \right\} , \\ P_{G\perp}(L) &= l_c^2 P_0\sum_{q=1}^{N}\sum_{p=1}^{N} \left\{ \left[\langle \left( \overrightarrow{\Delta\kappa}'_q\cdot \overrightarrow{\Delta\kappa}'_p \right) \rangle - \Delta\kappa^2\langle\cos\left( \theta_q \right)\cos\left( \theta_p \right)\rangle + i\langle\left(\overrightarrow{\Delta\kappa}'_q \times \overrightarrow{\Delta\kappa}'_p \right)\cdot\hat{p} \rangle \right] \right.\\ &\left. \langle \sin\left( \Omega_{nm}t + \zeta_q \right) \sin\left( \Omega_{nm}t + \zeta_p \right) \rangle \right\} . \end{aligned}$$
The assumption $l_c \ll l_p$ implies $\theta _p \approx \theta _q$, and $\overrightarrow {\Delta \kappa }'_q \approx \overrightarrow {\Delta \kappa }'_p$ therefore $\langle \cos \left ( \theta _q \right )\cos \left ( \theta _p \right )\rangle \approx \langle \cos \left ( \theta _p \right )^2\rangle$, $\langle \left ( \overrightarrow {\Delta \kappa }'_q\cdot \overrightarrow {\Delta \kappa }'_p \right ) \rangle \approx \Delta \kappa ^2$ and $\overrightarrow {\Delta \kappa }'_q \times \overrightarrow {\Delta \kappa }'_p \approx 0$. Furthermore $\langle \cos \left ( \theta _p \right )^2\rangle = 1/3$ and $\langle \cos \left ( \theta _p \right )\rangle = 0$ would hold as the GAWBS birefringence vector $\overrightarrow {\Delta \kappa }$ is expected to be distributed uniformly on the Poincare sphere due to random polarization rotation in the reference frame of the carrier Stokes vector. It should be noted that, to satisfy the condition of uniform distribution on the Poincare sphere, $\cos \left ( \theta _p \right )$ is uniformly distributed between -1 and 1, rather than $\theta _p$ uniformly distributed between 0 and $2\pi$, resulting in $\langle \cos \left ( \theta _p \right )^2\rangle = 1/3$ rather than $\langle \cos \left ( \theta _p \right )^2\rangle = 1/2$. Considering $\langle \sin \left ( \Omega _{nm}t + \zeta _q \right ) \sin \left ( \Omega _{nm}t + \zeta _p \right ) \rangle = \delta _{pq}/2$ Eq. (29) can be simplified to
$$ \begin{aligned} P_{G\parallel}(L) &= \frac{P_0 L}{2}\left(\bar{\kappa}^2 + \frac{1}{3}\Delta \kappa^2 \right)l_c , \\ P_{G\perp}(L) &= \frac{P_0 L}{2}\left(\frac{2}{3}\Delta \kappa^2 \right)l_c, \end{aligned}$$
where $L = Nl_c$. For the other extreme case where $l_p \ll l_c$, Eq. (29) can be put in the form
$$ \begin{aligned} P_{G\parallel}(L) &= l_c^2 P_0\sum_{u=1}^{N}\sum_{v=1}^{N}\bar{\kappa}^2 \langle \sin\left( \Omega_{nm}t + \zeta_u \right) \sin\left( \Omega_{nm}t+ \zeta_v \right) \rangle \\ &+ l_p^2 P_0\sum_{q=1}^{M}\sum_{p=1}^{M} \left\{ \left[\bar{\kappa}^2 +\Delta\kappa^2\langle\cos\left( \theta_q \right)\cos\left( \theta_p \right)\rangle + \bar{\kappa} \Delta\kappa \langle\cos\left( \theta_q \right) + \cos\left( \theta_p \right) \rangle \right] \right.\\ &\left. \langle \sin\left( \Omega_{nm}t + \zeta_q \right) \sin\left( \Omega_{nm}t + \zeta_p \right) \rangle \right\} , \\ P_{G\perp}(L) &= l_p^2 P_0\sum_{q=1}^{M}\sum_{p=1}^{M} \left\{ \left[\langle \left( \overrightarrow{\Delta\kappa}'_q\cdot \overrightarrow{\Delta\kappa}'_p \right) \rangle - \Delta\kappa^2\langle\cos\left( \theta_q \right)\cos\left( \theta_p \right)\rangle + i\langle\left(\overrightarrow{\Delta\kappa}'_q \times \overrightarrow{\Delta\kappa}'_p \right)\cdot\hat{p} \rangle \right] \right.\\ &\left. \langle \sin\left( \Omega_{nm}t + \zeta_q \right) \sin\left( \Omega_{nm}t + \zeta_p \right) \rangle \right\} , \end{aligned}$$
where $L = N l_c = M l_p$. In this case $l_p \ll l_c$ implies $\zeta _p \approx \zeta _q$ and hence $\langle \sin \left ( \Omega _{nm}t + \zeta _q \right ) \sin \left ( \Omega _{nm}t + \zeta _p \right ) \rangle =1/2$. Using $\langle \overrightarrow {\Delta \kappa }'_q\cdot \overrightarrow {\Delta \kappa }'_p \rangle = \Delta \kappa ^2\delta _{pq}$, and $\langle \cos \left (\theta _q \right ) \cos \left (\theta _p \right )\rangle = \delta _{pq}/3$ Eq. (31) reduces to
$$ \begin{aligned} P_{G\parallel}(L) &= \frac{P_0 L}{2}\left(\bar{\kappa}^2l_c + \frac{1}{3}\Delta \kappa^2l_p \right) ,\\ P_{G\perp}(L) &= \frac{P_0 L}{2}\left(\frac{2}{3}\Delta \kappa^2 \right)l_p . \end{aligned}$$
Keeping in mind the assumption that this is only true when either $l_c \ll l_p$ or $l_c \gg l_p$, Eqs. (30) and (32) can be combined together as
$$ \begin{aligned} P_{G\parallel}(L) &= \frac{P_0 L}{2}\left(\bar{\kappa}^2 + \frac{\min[l_c,l_p]}{3l_c}\Delta \kappa^2 \right)l_c , \\ P_{G\perp}(L) &= \frac{P_0 L}{2}\left(\frac{2\min[l_c,l_p]}{3l_c}\Delta \kappa^2 \right)l_c. \end{aligned}$$
Comparing Eq. (33) with Eq. (27) shows that $P_G = P_{G\parallel } + P_{G\perp }$ as expected.

Now that the GAWBS noise is obtained in terms of the GAWBS coupling coefficients and fiber length in Eqs. (27) and (33), it can be related to the acoustic strain. The GAWBS terms in Eq. (33), namely, $\bar {\kappa }$ and $\overrightarrow {\Delta \kappa }$ can be expressed in terms of overlap integrals between the GAWBS strain components and the square of optical mode field by combining Eqs. (7), (12), and (18) as

$$ \begin{aligned} \bar{\kappa}& = \kappa_0 p_p \iint_A\Theta_d\left(n\phi \right) \left( s_{rr} +s_{\phi\phi} \right) \bar{f} \left( r,\phi \right)^2 r\,\mathrm{d}r \mathrm{d}\phi , \\ \delta\kappa & = \kappa_0 p_m \iint_A \left[\Theta_d\left(n\phi \right)\cos\left(2\phi \right) \left( s_{rr} -s_{\phi\phi} \right)-2 \Psi_d\left(n\phi \right)\sin\left(2\phi \right) s_{r\phi} \right] \bar{f} \left( r, \phi \right)^2 r\,\mathrm{d}r \mathrm{d}\phi , \\ \kappa_{xy} & = \kappa_0 p_m \iint_A \left[\Theta_d\left(n\phi \right)\sin\left(2\phi \right) \left( s_{rr} -s_{\phi\phi} \right) + 2 \Psi_d\left(n\phi \right)\cos\left(2\phi \right) s_{r\phi} \right]\bar{f} \left( r, \phi \right)^2 r\,\mathrm{d}r \mathrm{d}\phi , \end{aligned}$$
where $\delta \kappa = (\kappa _{xx} - \kappa _{yy})/2$, $\kappa _0 = \frac {k_0 n_0^3}{2} \frac {C_{nm}\Omega _{nm}}{ V_s}$. It was assumed implicitly that the single mode core is circularly symmetric therefore both polarization modes have the same spatial distribution given by $f\left (r,\phi \right )$, and $\bar {f}\left (r,\phi \right )^2 \equiv f\left (r,\phi \right )^2/\iint _Af\left (r,\phi \right )^2r\,\mathrm {d}r \mathrm {d}\phi$ is the corresponding normalized distribution. The integration domain $A$ stands for the fiber cross-section. Since we assume circular symmetry for the core, it may seem that the $\phi$ dependence in $f\left (r,\phi \right )$ may be redundant. However, the coordinate frame that is adopted is centered at the center of the fiber. If the core is not concentric with the fiber, as in multi-core fibers, or due to residual core-clad concentricity error, $f\left (r,\phi \right )$ is not circularly symmetric in the chosen coordinate system. Therefore, Eq. (34) can be used to calculate GAWBS noise for cores that are not at the center of the fiber or to estimate the impact of core-clad concentricity error on GAWBS noise [19].

In the further special case of the core being concentric with the fiber, the mode field distribution becomes circularly symmetric and no longer has $\phi$ dependence. In this case the radial and angular integrals in Eq. (34) can be separated as

$$ \begin{aligned} \bar{\kappa}& = \frac{\kappa_0 p_p}{2\pi} \int_{0}^{a}\left( s_{rr} +s_{\phi\phi} \right) \bar{f} \left( r\right)^2 r\,\mathrm{d}r \int_{0}^{2\pi}\Theta_d\left(n\phi \right) \mathrm{d}\phi , \\ \delta\kappa & = \frac{\kappa_0 p_m}{2\pi} \int_{0}^{a}\left( s_{rr} -s_{\phi\phi} \right)\bar{f} \left( r \right)^2 r\,\mathrm{d}r \int_{0}^{2\pi}\Theta_d\left(n\phi \right)\cos\left(2\phi \right)\phi \\ -&2\frac{\kappa_0 p_m}{2\pi}\int_{0}^{a}s_{r\phi} \bar{f} \left( r \right)^2 r\,\mathrm{d}r \int_{0}^{2\pi}\Psi_d\left(n\phi \right)\sin\left(2\phi \right) \mathrm{d}\phi , \\ \kappa_{xy} & = \frac{\kappa_0 p_m}{2\pi} \int_{0}^{a}\left( s_{rr} -s_{\phi\phi} \right)\bar{f} \left( r \right)^2 r\,\mathrm{d}r \int_{0}^{2\pi}\Theta_d\left(n\phi \right)\sin\left(2\phi \right)\phi \\ +&2\frac{\kappa_0 p_m}{2\pi}\int_{0}^{a}s_{r\phi} \bar{f} \left( r \right)^2 r\,\mathrm{d}r \int_{0}^{2\pi}\Psi_d\left(n\phi \right)\cos\left(2\phi \right) \mathrm{d}\phi , \end{aligned} $$
where $\bar {f}\left (r\right )^2 \equiv f \left (r\right )^2/\int _{0}^{\infty }f\left ( r\right )^2 r \,\mathrm {d}r$, and the angular integral is calculated and taken out. The angular integrals in Eq. (35) determine which acoustic modes can cause coupling between the two polarization components, and as such act as selection rules. The equations are written in a general form where the angular dependence can have both upper, $d=0$ and lover solutions $d=1$, as shown in Eq. (5). First thing to notice is that all angular integrals vanish unless $n=0$, or $n=2$. Moreover, when $n=0$, but for the lower solution in Eq. (5), i.e., $\Theta \left (0\right ) = 0$ and $\Psi \left (0\right ) = -1$, all the coupling coefficients still vanish. This choice of angular dependence corresponds to pure shear acoustic modes as $U_r=0$ in Eq. (1). On the other hand, for the case of the upper solution, i.e., $\Theta \left (0\right ) = 1$ and $\Psi \left (0\right ) = 0$ corresponds to pure radial dilation modes as $U_{\phi }=0$. These acoustic modes are denoted as R$_{0m}$ modes. For R$_{0m}$ modes the coupling coefficients do not vanish and carrying out the angular integrals in Eq. (35) they can be reduced to
$$ \begin{aligned} d = 0: \quad &\bar{\kappa} = \kappa_0 p_p \int_{0}^{\infty}\left(s_{rr} +s_{\phi\phi} \right) \bar{f} \left( r\right)^2 r \,\mathrm{d}r, \quad &\delta\kappa = 0, \quad &\kappa_{xy} = 0 , \\ d = 1: \quad &\bar{\kappa} = 0, \quad &\delta\kappa = 0, \quad &\kappa_{xy} = 0 . \end{aligned} $$

In the case of $n=0$, optical field experiences phase modulation only since the phase shift is the same for both polarizations, therefore no birefringence i.e., $\delta \kappa =0$, and the cross coupling term $\kappa _{xy}=0$. Therefore, if a polarized field is incident, the GAWBS peaks created by $n=0$ mode group create only GAWBS peaks that are also polarized and in the same polarization as the incident optical field. As a result these GAWBS side bands created by the R$_{0m}$ modes are referred to as the polarized GAWBS peaks, or polarized GAWBS noise.

The remaining case where the coupling coefficients does not vanish is for $n=2$. These modes are mixed acoustic modes having both radial and shear (torsional) components, hence denoted as $TR_{2m}$ modes. For $n=2$, the upper and lower solutions in Eq. (5) are simply rotated versions of one another by $45^{\mathrm {o}}$. The upper solution, i.e. $d = 0$, has two isoclines at angles $\phi =0^{\mathrm {o}}, 90^{\mathrm {o}}$, whereas for the lower solution, $d=1$, the two isoclines are at angles $\phi =\pm 45^{\mathrm {o}}$. Again by carrying out the angular integrals in Eq. (35), the coupling coefficients for the case of $TR_{2m}$ modes can be expressed as

$$ \begin{aligned} d = 0: \quad &\bar{\kappa} = 0, \quad &\delta\kappa = \frac{\kappa_0}{2}p_m\int_{0}^{\infty}\left(s_{rr} - s_{\phi\phi} - 2s_{r\phi}\right) \bar{f} \left( r\right)^2 r \,\mathrm{d}r, \quad &\kappa_{xy} = 0 , \\ d = 1: \quad &\bar{\kappa} = 0, \quad &\kappa_{xy} = \frac{\kappa_0}{2}p_m\int_{0}^{\infty}\left(s_{rr} - s_{\phi\phi} - 2s_{r\phi}\right) \bar{f} \left( r\right)^2 r \,\mathrm{d}r, \quad &\delta \kappa = 0 . \end{aligned} $$
Equation (37) shows that for the two degenerate cases $d=0$, and $d=1$, the contribution from $TR_{2m}$ modes are either through $\delta \kappa$ or $\kappa _{xy}$ depending on the relative orientations of the isoclines of the acoustic mode and the incident polarization of the optical field. However, since on average the total noise power is determined by $\bar {\kappa }^2$ and $\Delta \kappa ^2 = \delta \kappa ^2 + \kappa _{xy}^2$ according to Eq. (27), both degenerate cases contribute equally as expected. Since both degenerate modes are equally present their contributions should be summed to find the GAWBS noise contribution of $TR_{2m}$ modes. Therefore, for the case of $TR_{2m}$ modes $\Delta \kappa ^2$ simplifies to
$$\Delta\kappa^2 = \frac{\kappa_0^2p_m^2}{2} \left[ \int_{0}^{\infty}\left(s_{rr} - s_{\phi\phi} - 2s_{r\phi}\right) \bar{f} \left( r\right)^2 r \,\mathrm{d}r\right]^2 .$$
According to Eq. (37), $TR_{2m}$ modes only contribute to the birefringence term $\Delta \kappa$ and none to the isotropic term $\bar {\kappa }$. Therefore, $TR_{2m}$ modes are solely responsible for the creation of optical noise in the orthogonal polarization with respect to the optical carrier. Since $TR_{2m}$ modes contribute to both parallel and orthogonal GAWBS, the total noise generated by these modes are referred to as the depolarized GAWBS [7] noise as opposed to the polarized noise generated by $R_{0m}$ modes. Previously, the total GAWBS noise was considered as the sum of parallel $P_{G\parallel }$ and orthogonal contributions $P_{G\perp }$ which is the most natural way to separate it from a measurement point of view. However, since for the special case of single-mode fiber with a circularly symmetric core at the center of the fiber there are only two mode groups that generates GAWBS noise, an alternative way to analyze it is to separate into parts based on their physical origin. Instead of calling them $R_{0m}$ and $TR_{2m}$ contributions, it was preferred to name them as the polarized GAWBS noise created by $R_{0m}$ modes denoted as$P_{GP}$ and depolarized GAWBS noise generated by $TR_{2m}$ modes denoted as $P_{GD}$ [7]. According to Eq. (33) these four parameters are related as
$$ \begin{aligned} P_{G\parallel} &= P_{GP} + \frac{1}{3}P_{GD} \\ P_{G\perp} &= \frac{2}{3}P_{GD} \quad \quad \quad \implies \quad P_{GP} = P_{G\parallel} - \frac{1}{2}P_{G\perp} \\ & \qquad \qquad\qquad \qquad \qquad \qquad P_{GD} = \frac{3}{2}P_{G\perp} \end{aligned} .$$
Equation (39) shows that even if it may not be possible to measure the polarized and depolarized GAWBS directly, they can be obtained from the parallel and orthogonal measurements.

The remaining term left in determining the coupling coefficients in Eqs. (36) and (37) are $C_{nm}$. These terms can be determined following the approach of Shelby [7] which invokes the argument of equipartition theorem. The total energy taken by each acoustic mode should be equal to $k_{B} T$ where $k_{B}$ is the Boltzmann constant, and $T$ is the absolute temperature. The energy of an acoustic mode covering the extent of the fiber cross section and a length $l_m$ is given by

$$\mathcal{E}_{nm} = \int_0^{l_m}\int_0^{2\pi}\int_0^{a}\frac{1}{2}\rho_0\Omega_{nm}^2\left(U_{r\mbox{-}nm}^2+U_{\phi\mbox{-}nm}^2\right)r \,\mathrm{d}r\mathrm{d}\phi\mathrm{d}z = k_{B} T ,$$
where $\rho _0$ is the density of the glass. Inserting Eq. (1) into Eq. (40) $C_{0m}$ and $C_{2m}$ can be related to the fiber parameters as
$$ \begin{aligned} C_{nm}^2 = \frac{2 k_B T}{\chi \pi l_m \rho_0 a^2 \Omega_{nm}^2 D_{nm}}, \quad \chi = \left\{ \begin{array}{c} 2, \, n=0 \\ 1, \, n \ne 0 \\ \end{array} \right. , \end{aligned}$$
with the term $D_{nm}$ is given by
$$ \begin{aligned} D_{nm} &= A_1^2 \left\lbrack J_{n-1}\left(y\right)^2 + J_{n+1}\left( y \right)^2 - J_{n}\left(y \right) \left\lbrack J_{n+2}\left(y\right) + J_{n-2}\left(y\right) \right\rbrack \right\rbrack \\ &+ \alpha^2 A_2^2 \left\lbrack J_{n-1}\left(\alpha y\right)^2 + J_{n+1}\left(\alpha y\right)^2 - J_{n}\left(\alpha y\right) \left\lbrack J_{n+2}\left(\alpha y\right) + J_{n-2}\left(\alpha y\right) \right\rbrack \right\rbrack \\ & + A_1 A_2 \frac{2\alpha}{y\left(\alpha^2 -1 \right)} \left\lbrack \frac{2 n \left(\alpha^2 -1 \right)}{\alpha} J_{n}\left(y \right) J_{n}\left(\alpha y \right) + J_{n-2}\left(y \right) J_{n-1}\left(\alpha y \right) + J_{n+2}\left(y \right) J_{n+1}\left(\alpha y \right) \right. \\ & \left. -\alpha \left\lbrack J_{n-1}\left(y \right) J_{n-2}\left(\alpha y \right) + J_{n+1}\left(y \right) J_{n+2}\left(\alpha y \right) \right\rbrack \right\rbrack , \end{aligned}$$
where $y = \Omega _{nm}a/V_s$. Putting together Eqs. (33), (36), (41), and (42), polarized GAWBS power generated by R$_{0m}$ modes can be obtained in terms of fiber parameters as follows
$$ \begin{aligned} P_{GP}\left(L,\Omega_{0m}\right) &= P_0 \frac{k_B T n_0^6 p_p^2 k_0^2 L}{4\pi \rho_0 a^2 V_d^2 } \left(\frac{l_c}{l_m}\right) \left\lbrack \frac{S\left(\Omega_{0m}\right)^2 }{ J_1 \left( \frac{\Omega_{0m}a}{V_d}\right) - J_0 \left( \frac{\Omega_{0m}a}{V_d} \right)J_2 \left( \frac{\Omega_{0m}a}{V_d} \right) } \right\rbrack, \\ S\left(\Omega_{0m}\right) &=\int_0^a J_0\left(\frac{\Omega_{0m}r}{V_d }\right) \bar{f} \left( r\right)^2 r \,\mathrm{d}r , \end{aligned}$$
Similarly for the GAWBS noise generated by the TR$_{2m}$ modes we get
$$ \begin{aligned} P_{GD}\left(L,\Omega_{2m}\right) &= P_0 \frac{k_B T n_0^6 p_m^2 k_0^2 L}{2\pi \rho_0 a^2 V_s^2 } \left(\frac{\min[l_c,l_p]}{l_m}\right)\left\lbrack \frac{ S\left(\Omega_{2m}\right)^2 }{D_{2m} } \right\rbrack , \\ S\left(\Omega_{2m}\right) &=\int_0^a \left\lbrack A_1 J_0\left(\frac{\Omega_{2m}r}{V_s }\right) - \frac{V_s^2}{V_d^2} A_2 J_0\left(\frac{ \Omega_{2m}r}{V_d }\right) \right\rbrack \bar{f} \left( r\right)^2 r \,\mathrm{d}r , \end{aligned}$$
where we used the fact that the $s_{rr} + s_{\phi \phi }$, $s_{rr} - s_{\phi \phi } - 2s_{r\phi }$ and $D_{0m}$ simplifies to
$$ \begin{aligned} s_{rr} + s_{\phi\phi} &= 2\alpha^2 A_2 J_{n}\left(\frac{\Omega_{nm}r}{V_d}\right), \\ s_{rr} - s_{\phi\phi} - 2 s_{r\phi} &= 2\left\lbrack A_1 J_{n-2}\left(\frac{\Omega_{nm}r}{V_s}\right) - \alpha^2 A_2 J_{n-2}\left(\frac{\Omega_{nm}r}{V_d}\right) \right\rbrack \\ D_{0m} &= 2\left(\alpha A_2\right)^2\left\lbrack J_1 \left( \frac{\Omega_{0m}a}{V_d} \right)- J_0 \left( \frac{\Omega_{0m}a}{V_d} \right)J_2 \left( \frac{\Omega_{0m}a}{V_d} \right) \right\rbrack . \end{aligned}$$
In deriving Eq. (44) the doubly degenerate nature of the TR$_{2m}$ modes is taken into account as described in Eq. (38).

The factor $S\left (\Omega _{nm}\right )$ in Eqs. (43) and (44) is the resulting overlap integral between the acoustic strain and the optical mode field squared in our derivation. In [7] several steps connecting the acoustic strain to the effective index modulation were skipped, and an unfortunate naming convention was used to describe the mode intensity profile as “The profile of the guided optical mode". This resulted in several papers following the derivation in [7] to conclude incorrectly that the overlap integral is between the acoustic mode profile and the optical mode-field profile, instead of the optical intensity profile [46,20]. Interestingly, others [10,15,21,22] that were not directly interested in GAWBS noise for long distance communication fibers, did not have this misunderstanding.

Equations (43) and (44) show explicit dependence of GAWBS noise power on the temperature which is expected as the acoustic modes causing GAWBS are driven by the ambient temperature. Therefore the measurements carried out at one temperature can be translated to an environment at a different temperature. A nontrivial example is that GAWBS noise for fibers in submarine cables at the bottom of the sea would have lower GAWBS noise compared to measured values at room temperature, e.g., 0.25 dB lower at 4 $^{\circ }$C compared to 20 $^{\circ }$C.

As mentioned earlier, the vibration frequencies of the acoustic modes can be found out by solving the characteristic equation given by Eq. (2). For the case of $n=2$, the two modes corresponding to the upper and lower solutions of the Eq. (5) are degenerate and have the same frequencies as the solution, however, this is not the case for the $n=0$ modes. The pure shear modes and the radial modes R$_{0m}$ have different solutions. For the case of $n=0$ the Eq. (2) factors into a product as follows

$$|B| = \left\lbrack \frac{a^2 \Omega^2}{2V_s^2}J_{2}\left(\frac{a \Omega}{V_s} \right) \right\rbrack \left\lbrack \frac{a^2 \Omega^2}{2V_s^2}J_{0}\left(\frac{a \Omega}{V_d}\right) - \frac{a \Omega}{V_d}J_{1}\left(\frac{a \Omega}{V_d} \right) \right\rbrack = 0,$$
where the roots of the first parenthesis yield the frequencies of the pure shear acoustic modes, and the roots of the parenthesis on the right corresponds to the frequencies $\Omega _{0m}$ of the R$_{0m}$ modes supported by the fiber.

In Eqs. (43) and (44), the GAWBS noise power is expressed in terms of fiber parameters that can be easily measured except for the length scales $l_m$ and $l_c$ which determine the average mode length, and the average mode coherence length, respectively. It is reasonable to assume that these two length scales are the same, as the acoustic modes would be coherent within the mode length and vice versa. Therefore, it will be assumed without further proof that $l_m =l_c$ and their ratio in these equations will be omitted. Furthermore, it will be shown that for the fibers measured for this paper, a good agreement is observed for the condition $l_c \ll l_p$. The remaining parameters can be divided into a group that depends on the waveguiding parameters of the fiber, e.g., core diameter, cut-off wavelength which is contained in the overlap integral term $\Gamma$, and those parameters that does not play a direct role on determining the optical mode profile, such as properties of the glass, cladding diameter, absolute temperature etc.. The overlap integral shows that a large overlap between strain from a particular acoustic mode and optical mode-squared creates a larger GAWBS noise power created at the frequency of that particular mode. It is worth noting that an important piece missing in the analysis is the impact of the polymer coating which plays a crucial role in mechanical reliability of optical fibers.

Equations (43) and (44) can be used to calculate the magnitude of GAWBS peaks generated by individual acoustic modes with vibration frequencies $\Omega _{nm}$. More often than not it is the relative power of the GAWBS noise with respect to the carrier power $P_0$ that matters, therefore when appropriate the results in the rest of the paper will be presented in terms of GAWBS scattering efficiency defined as $\gamma _{G}\left (\Omega _{nm} \right ) = P_{G}\left (\Omega _{nm} \right ) /P_0$. The same notation used to describe the polarization of the GAWBS noise power will be applied to the GAWBS efficiency, i.e., $\gamma _{Gx}\left (\Omega _{nm} \right ) = P_{Gx}\left (\Omega _{nm} \right ) /P_0$ where $x = \parallel , \perp , P, D$ depending on whether the GAWBS noise is parallel or orthogonal to the carrier, polarized or depolarized, respectively. In defining the spectral features of the GAWBS noise, the single side spectrum convention is adopted here as per Eq. (24) the spectrum is symmetric, and a time-domain average is implemented in deriving Eqs. (27) and (33).

For optical communications, what matters is the sum of GAWBS noise power with respect to the carrier from all the GAWBS modes, as this sum determines the signal to noise ratio due to GAWBS noise only. Connecting the notation of [3] to this paper, the total GAWBS efficiency is defined as

$$\Gamma_{GAWBS} = \sum_{n,m=0}\gamma_{G}\left(L,\Omega_{nm}\right), \quad \Gamma_{GAWBS,x} = \sum_{n,m=0}\gamma_{Gx}\left(L,\Omega_{nm}\right),$$
where, again, $x = \parallel , \perp , P, D$ is the descriptor for the polarization.

Inspecting Eqs. (43) and (44), it may appear that there is a strong dependence on the fiber cladding radius $a$. Comparing the GAWBS peak powers for $a=62.5~\mu m$ and $a=40~\mu m$ as shown in Fig. 1(a), indeed the magnitude of GAWBS peaks generated by individual modes scales inversely with cladding radius $a$ however, the frequency spacing between GAWBS modes is reduced in proportion to $a$, resulting in the total GAWBS noise to remain the same. It is easy to see that the frequency spacing between modes reduces as cladding radius increases since the mode frequencies are roughly determined by the round-trip time of sound waves between the cladding air interface. The total GAWBS noise power is plotted in Fig. 1(b) with respect to the cladding radius, for polarized only ($\Gamma _{GAWBS,P}$), depolarized ($\Gamma _{GAWBS,D}$) only and their sum ($\Gamma _{GAWBS}$). The total GAWBS noise is virtually independent of fiber radius. Another point worth mentioning is that the total depolarized GAWBS noise power generated by TR$_{2m}$ modes is lower than the polarized GAWBS power generated by the R$_{0m}$ by only $\approx 2.5$ dB and should not be neglected as suggested in [4]. The parameters used to calculate the curves in Fig. 1 and the rest of the paper are listed in Table 1 unless specified otherwise. Moreover, the optical mode profile is assumed to be the fundamental mode of a weakly guiding circularly symmetric core given by [23]

$$f\left( r \right) = \left\{ \begin{array}{@{}c@{}} \frac{J_0\left( \frac{p r}{w_c} \right)}{J_0\left( p \right)}, \, r \leq w_c \\ \frac{K_0\left( \frac{q r}{w_c} \right)}{K_0\left( q \right)}, \, r > w_c \\ \end{array} \right. \quad p = w_c k_0\sqrt{n_c^2 - n_0^2}, \quad q = w_c k_0\sqrt{n_0^2 - n_{cl}^2 } ,$$
where $w_c$ is the core radius, $n_c$ and $n_{cl}$ are the core and cladding refractive indices. Even though the mode profile is described in terms of the parameters $w_c$, $n_c$ and $n_{cl}$ in Eq. (48), it can also be determined by providing the effective area $A_{eff}$ and the cutoff wavelength $\lambda _c$ under the weakly guiding assumption. The latter parameters will be used in the rest of the paper as they are more likely to be quoted in fiber specifications, where the effective area is defined as [24]
$$A_{eff} = 2\pi \frac{\left\lbrack \int_{0}^{\infty} f \left( r\right)^2 r \,\mathrm{d}r \right\rbrack^2}{\int_{0}^{\infty} f \left( r\right)^4 r \,\mathrm{d}r}.$$

 figure: Fig. 1.

Fig. 1. (a) GAWBS peaks generated by R$_{0m}$ modes (filled markers), and peaks generated in the perpendicular polarization by TR$_{2m}$ modes (empty markers), for fiber cladding radius $a = 40 \mu m$ (blue) and $a = 62.5 \mu m$ (yellow) (b) Polarized (blue), depolarized (yellow) and total (green) GAWBS noise power as a function of cladding radius $a$.

Download Full Size | PDF

Tables Icon

Table 1. Parameters used for calculating GAWBS noise

3. Measurements

GAWBS noise is a sizable noise source after accumulating transoceanic distances [1,3]. Therefore it is important to measure the level of this noise accurately since it affects the expected capacity of a new transmission system. Even though it is not negligible after long distances GAWBS noise is still quite small after short lengths of fiber and therefore not easy to measure accurately. Earlier papers used measurement set ups that used expensive and very stable free-space lasers with high powers and short pieces of fibers of the order of several meters [7,25]. One solution is setting up a recirculating loop to emulate long distance transmission where the GAWBS noise accumulates sufficiently [1]. However, amplified spontaneous emission (ASE) noise and nonlinear noise accumulates at the same rate as GAWBS noise [3], therefore, separating the GAWBS from other noise sources becomes an issue. The signal to ASE noise ratio in the case of single-span measurement would be less limited by nonlinear noise sources such as modulational instability, and stimulated Brillouin scattering. Moreover, setting up a stable recirculating loop is cumbersome and requires multiple spans of fiber which is not practical and may not be cost effective for measuring many types of fibers. Others were able to measure GAWBS noise in lengths similar to typical span lengths, $\sim$80 km using low-phase noise lasers with a heterodyne setup using a single photo-diode and an electric spectrum analyzer or a digitizer [5,26] with sufficient signal to noise ratio. What is missing in these reports so far is the portion of the GAWBS noise orthogonal to the input signal. In the nomenclature of this paper, $\gamma _{G\parallel }$ is reported, however, $\gamma _{G\perp }$ which makes up about a third of the total noise is missing. It should be noted that even though impact of GAWBS noise is difficult to mitigate using digital signal processing, contribution of the depolarized GAWBS is the most difficult part as typically the equalizers for polarization rotation are much slower than the phase equalizers. Therefore, it is important to know how much of the GAWBS noise is pure phase noise and how much is polarization noise and properly include both.

Our measurement set up shown in Fig. 2 is similar to the heterodyne technique, though we use a homodyne setup. The other major differences are that we use a polarization diversity receiver to simultaneously measure all the polarization components of the GAWBS noise, and that we use a secondary laser to automatically calibrate the noise power. A tunable laser ITLA1 with a nominal linewidth of 100 kHz is split into two parts. One part combines with a second ITLA and goes through the fiber under test (FUT). After the FUT they are amplified, and a 50 GHz filter centered around the laser frequency removes ASE noise. After the filter the lasers go into the signal port of a phase and polarization diversity hybrid. ITLA1 in the remaining arm of the polarization maintaining coupler is connected to the LO port of the hybrid. Frequency of ITLA2 is detuned from that of ITLA1 by $\sim$1 GHz, and its power is reduced by 29 dB compared to the ITLA1. After the four photodetectors that receives the in-phase and out-of-phase beating between the two orthogonal components of the hybrid inputs, a real-time sampling scope samples the signal at 5Gsa/s. After the photodetectors band-pass filters (BPFs) rejects noise outside of the range between 20 MHz to 1000 MHz to avoid ASE noise folding back in to the Nyquist band, and also to reject the high carrier power. High extinction ratio polarization maintaining fibers and couplers are used to combine the two lasers to make sure they remain in the same polarization.

 figure: Fig. 2.

Fig. 2. Homodyne measurement setup. ITLA: integrated tunable laser, PM:polarization maintaining, EDFA: erbium-doped fiber, PC: polarization controller, LO: local oscillator PD: photo-diode, BPD: band-pass filter, ADC: analog-to-digital converter. Polarization maintaining paths and components are shown in blue. A and B show the input and output output points for the fiber under test (FUT).

Download Full Size | PDF

This measurement set up has several advantages compared to the previous methods. First, because the carrier is rejected by the BPFs, the dynamic range of the receiver is not wasted on capturing the carrier, and the instrument noise can be reduced. However, without the carrier, it may require indirect methods to relate the measured GAWBS power level to that of the carrier. This is the main purpose of the second ITLA which is placed just outside of the spectral window that contain the majority of the GAWBS noise power but still within the pass-band of the BPFs. Power level of GAWBS noise is measured simultaneously with ITLA2 whose power with respect to the first ITLA is already known. The second advantage is that GAWBS noise in both polarizations, in other words, both parallel and orthogonal contributions can be measured simultaneously. Since the second ITLA have the same polarization as the first ITLA, the GAWBS noise that is parallel and orthogonal to the original carrier can easily be distinguished. A third advantage of this set up is that since it is a homodyne measurement, the impact of laser phase noise and laser drift is reduced, and therefore the measurement can be carried out with simpler and cheaper lasers. Compared to the method that emulates long-distance transmission with recirculating loops, the measurement set up is much more simple, allowing for quickly characterizing different fiber types of short lengths while avoiding interference from ASE noise, or nonlinear noise. On the other hand, since the fiber length is on the shorter side, the laser phase noise is larger than GAWBS level especially in the low frequency range, and receiver noise floor limits the measurement at the high frequency side. Both of these noise sources needs to be carefully measured and removed. Fortunately, it is straightforward to characterize both in back-to-back measurements compared to nonlinear noise after long transmission distances.

Some of the fibers measured are listed in Table 2 and they will be referred to using the abbreviations in the Table for brevity. Figure 3 shows the spectra before (red) and after (black) SMF28_80 in the polarization aligned with that of the carrier, i.e, ITLA1. The BPFs are replaced by high-pass filters (HPFs) with their cutoff frequency at 1200 MHz to show how small the GAWBS peaks are compared to the carrier. As expected the GAWBS peaks appear symmetrically around the carrier which are pointed out in the Fig. 3. The carrier power launched into the FUT is 4.5 dBm. With the HPFs instead of BPF, the carrier power needs to be reduced by close to 12 dB in order not to exceed the dynamic range of the ADC. As a result, the receiver dynamic range is wasted on capturing the carrier power, while pushing up the noise floor to obscure the GAWBS spectrum. To identify the noise sources, a back-to-back (BTB) measurement is taken where the FUT is removed. The fringes that appear in the BTB measurement close to the carrier follows closely the noise envelope measured after the FUT in the window close to the carrier confirming that in that region the noise is dominated by phase noise and not fiber nonlinearity. Moreover, the noise floor away from the carrier converge for both cases showing that without the BPF the measurement is limited by receiver noise floor away from the carrier. The scope also has spurious tones that shows up in the spectra as sharp peaks which can be identified in the BTB measurement and removed as well.

 figure: Fig. 3.

Fig. 3. Spectrum in the polarization parallel to that of the carrier without the BPFs after 80 km of SMF-28 (black) and in back-to-back configuration (red).

Download Full Size | PDF

Tables Icon

Table 2. Fibers and some of their parameters that are measured.

Figure 4 shows the measured parallel GAWBS noise before and after various stages of post-processing. First, data stream from the 4 scope channels are combined to obtained the total field in two polarizations. Then the spectrum is calculated with a frequency interval of 9.5 kHz. The polarization state is determined using the ITLA2 as the reference. At this point the blue line in Fig. 4 is obtained which is dubbed the "raw spectrum". After removing the receiver noise floor, the yellow line is obtained. The frequency dependent channel response of the scope which is measured separately using a flat ASE source, is removed from the spectrum resulting in the green curve. Finally the laser noise (red curve), which is also characterized separately is removed yielding the purple line which we consider the final result. Since the GAWBS spectrum is symmetric around the carrier, for simplicity, the spectra are shown after being folded onto the positive side. Figure 4(b) is same as Fig. 4(a) except that it is plotted at an effective frequency resolution of 503 kHz for better clarity.

 figure: Fig. 4.

Fig. 4. The parallel GAWBS scattering efficiency for SMF28_80 after various post-processing steps at a frequency resolution of (a) 9.5 kHz and (b) 503 kHz.

Download Full Size | PDF

The processing steps for obtaining the orthogonal GAWBS noise is similar to the case of parallel GAWBS noise except that laser noise removal step can be skipped as the polarization of the laser phase noise is well aligned with the carrier as shown in Fig. 5. A residual of ITLA2 power is left over due to imperfect alignment of the polarization angle.

 figure: Fig. 5.

Fig. 5. Orthogonal GAWBS noise scattering efficiency measured for SMF28_80 at various stages of processing at a frequency resolution of (a) 9.5 kHz and (b) 503 kHz.

Download Full Size | PDF

For the case of parallel GAWBS, the measurement is dominated by the laser noise in the frequency range close to the carrier, and by the scope noise floor away from the carrier. For SMF28_80 GAWBS peaks could actually still be distinguished were it not for the ITLA2, and the BPFs, however, at that point their power level is already close to 20 dB below the maximum. Similarly the very first peak at 30.9 MHz cannot be accurately characterized as a small error in the estimation of the laser phase noise can distort the magnitude and shape of this peak significantly. Nevertheless, contribution of this peak to the total GAWBS noise is small enough that the impact of error in estimation of this peak is negligible. Similarly, an error in estimation of the scope noise floor can affect the accurate estimation of the GAWBS peaks furthest from the carrier, however since the main GAWBS peaks are more than 20 dB higher than the noise floor, the impact of this error can also be neglected. Similarly, for the case of orthogonal GAWBS noise, the impact of an error in the noise floor estimation can significantly alter the shape of the GAWBS spectrum away from the carrier, however its impact on the total GAWBS noise is small.

GAWBS noise is measured for four different fiber types described in Table 2. The parallel GAWBS noise measured for these four fiber types are shown in Fig. 6. Since the lengths of the measured fibers are different, and the GAWBS noise scales linearly with the fiber length, all measurements are scaled to 1000 km for easier comparison. The measurements for the all fiber types look quite similar with periodic GAWBS peaks, except that the GAWBS peaks are smaller, and they taper off at larger frequencies faster for fibers with larger effective areas. Figure 7 shows the measurement results for the orthogonal GAWBS noise.

 figure: Fig. 6.

Fig. 6. Parallel GAWBS noise scattering efficiency measured for four different fiber types extrapolated to 1000 km.

Download Full Size | PDF

 figure: Fig. 7.

Fig. 7. Orthogonal GAWBS noise scattering efficiency measured for four different fiber types extrapolated to 1000 km.

Download Full Size | PDF

The four curves in Figs. 6 and 7 are plotted together in Figs. 8(a) and 8(b), respectively, for easier visual comparison. It can be seen that the GAWBS peak frequencies align pretty well for all four fiber types. This is expected as all the fibers are pure silica-core fibers that have the same nominal cladding diameter of 125 $\mu$m. Nevertheless, a small shift in frequency spacing is still noticeable especially at higher frequencies which can be attributed to small variations in the cladding diameters [25].

 figure: Fig. 8.

Fig. 8. (a) Curves in Fig. 6 plotted together. (b) Curves in Fig. 7 plotted together.

Download Full Size | PDF

At this point the total GAWBS noise power $\Gamma _{GAWBS}$ can be calculated by integrating GAWBS noise spectrum. The integration window is chosen so that the lower limit includes the first prominent peak which occurs at about 30 MHz for the parallel GAWBS and at 20 MHz for the orthogonal GAWBS, and the highest limit includes all the peaks that are larger than -10 dB below the highest GAWBS peak. In general it is found that the higher frequency peaks lower than this threshold contribute less than 2% of the total noise power. As an example, Fig. 9 shows GAWBS noise and the accumulated noise as a percentage of the total noise from left to right for SMF28_80. The cumulative power is obtained by performing a rolling sum starting from the shorter frequencies. The green curve shows the frequency window over which the GAWBS noise is integrated for this particular case. The last two GAWBS peaks in the case of parallel GAWBS in Fig. 9(a), and the last three peaks in the orthogonal case in Fig. 9(b) contribute less than 1% of the total noise.

 figure: Fig. 9.

Fig. 9. GAWBS scattering efficiency on the left axis and the cumulative noise as a percentage of the total sum on the right axis obtained by a rolling sum of the spectrum form left to right for parallel (a) and orthogonal (b) contributions.

Download Full Size | PDF

Using the procedure outlined above, the total GAWBS noise calculated for the four fiber types are tabulated in Table 3. Since GAWBS noise power is calculated in reference to the carrier power, the tabulated values can be considered the inverse of signal-to-noise ratio (SNR) due to GAWBS noise only. It has been reported that the GAWBS noise is inversely correlated with the fiber effective area [4,5] which is confirmed with our measurement of fibers with different effective areas.

Tables Icon

Table 3. Total GAWBS noise power measured relative to the carrier power for four types of fiber broken down into parallel and orthogonal contributions.

The black markers in Fig. 10 shows the total GAWBS noise for the four fiber types scaled to 1000 km, as a function of their effective areas. The measurement results are fitted with a linear curve (red) using mean-squared regression resulting in a coefficient of determination $R^2 > 0.99$ which confirms the expected relationship between GAWBS noise and effective area [4,5]. Furthermore, the expected GAWBS noise is calculated using Eqs. (43) and (44) with the parameters in Table 1 where the effective area and length are replaced by the values in Table 2, and $p_{11} = 0.115$, and $p_{12} = 0.266$ are used for a better fit. The result from the developed theory is shown with the blue curve in Fig. 10. The agreement with the measurements and the theoretical estimation is very good, even though the theory does not take into account the impact of the coating, and the same parameters such as refractive index, and speed of sound are assumed for all the fibers.

 figure: Fig. 10.

Fig. 10. Total GAWBS noise measured (black diamonds) for four fiber types including both parallel and orthogonal contributions plotted as a function of their effective area. Linear fitting (red) and theoretical estimation (blue) area also included. For theory part, $p_{11} = 0.115$, and $p_{12} = 0.266$ are used for a better fit.

Download Full Size | PDF

GAWBS noise is also measured for various lengths of Vascade Ex2000 shown in Table 2, where the 40-km-long Ex2_40 is obtained by splicing the four 10-km-long fibers. Figure 11 shows the measured GAWBS noise as a function of their lengths. As expected [1,4,7], the GAWBS noise scales linearly with their length. A linear fitting using mean-squared regression resulted in a coefficient of determination of $R^2 = 0.999, 0.992$ and $0.998$ for the parallel, orthogonal and total GAWBS noise, respectively. The results show that this measurement technique can be used to estimate the GAWBS noise for fiber lengths as short as 10 km. However, a larger variation in the measurement is observed for the 10 km fibers especially for the smaller orthogonal contribution.

 figure: Fig. 11.

Fig. 11. Total measured GAWBS noise and its break down into the parallel and orthogonal contributions as a function of their lengths for Vascade Ex2000 fibers, i.e. Ex2_10_1, Ex2_10_2, Ex2_10_3, Ex2_10_4, Ex2_40, Ex2_48, and Ex2_84. The Ex2_40 is obtained by splicing the four 10-km pieces.

Download Full Size | PDF

So far, GAWBS measurement results are presented for fiber lengths that are on the order of a span length or shorter. Even though the linear dependence on fiber length is confirmed by measurement results presented in Fig. 11, it is worth confirming this dependency in a long distance, multi-span transmission set up. To compare the single span measurements with multi-span measurements, we replaced the FUT in Fig. 2 with the circulating loop shown in Fig. 12(a). At the transmitter, a flat ASE spectrum spanning the C-band is prepared to emulate transmission spectrum. The ITLAs for the GAWBS measurement are inserted in a 150 GHz wide gap centered at 1547.7 nm. The loop consists of 8 spans of Ex2_84 with an average span loss of 13.4 dB and a total loop length of 675 km. Though the optimum transmission power would require the EDFA output to be above 18 dBm, it was found that the GAWBS spectrum is significantly distorted by cross-phase modulation (XPM) at these power levels. In order to minimize the impact of XPM, the EDFA output power is reduced to 13 dBm. The reduced power level reduces the SNR due to accumulated ASE from the EDFAs. At shorter span lengths the ITLA powers can be kept high to increase SNR, however at longer distances, beyond 4 loops, ITLA power is limited by four-wave mixing (FWM). Figure 12(b) shows the received spectra after 1, 4, 8, 12, and 16 loops. The ITLA power required to keep FWM at a negligible level limited the received optical SNR (OSNR), defined as the signal power divided by the noise power within 0.1 nm bandwidth, to 28.9, 17.7, 12.9, 8.8, 6.7 dB for the distances shown in Fig. 12(b). It was found that FWM noise grew faster than linear in the narrow frequency window around the carrier which made GAWBS measurements in long distances less accurate.

 figure: Fig. 12.

Fig. 12. (a) Recirculating loop setup. SW:switch, CPL:coupler, WSS: wavelength selective switch. (b) Received spectra after 1, 4, 8, 12, and 16 loops.

Download Full Size | PDF

Figure 13 shows the parallel and orthogonal GAWBS noise spectra measured after various distances using the recirculating loop setup and compared to the single-span measurement (dashed black). All spectra are are scaled to 1000 km assuming linear dependence on distance. As expected, multi-span measurements agree very well with the single-span measurement. At shorter distances, less than 4 loops, a high level of OSNR can be maintained while avoiding FWM. At longer distances, the OSNR has to be reduced low enough to avoid FWM that the spectra becomes more noisy especially away from the carrier where the GAWBS power is lower. At long distances the impact of laser linewidth can be neglected, even in the region close to the carrier. That the measurements at long distances agree well with the single-span measurements confirm that the impact of laser linewidth was properly taken care of in the single-span measurements.

 figure: Fig. 13.

Fig. 13. GAWBS noise scattering efficiency after various distances in the loop setting compared to single span (dashed black). Even though measured at different distances, all measurement results are scaled to 1000 km for easier comparison (a) Parallel GAWBS noise. (b) Orthogonal GAWBS noise.

Download Full Size | PDF

Total GAWBS noise is calculated by integrating the GAWBS spectrum for the loop measurements similar to the single-span measurements. Figure 14 shows the total GAWBS noise as a function of transmission distance both in absolute value and also as scaled to 1000 km. The shortest distance point is the single-span measurement for comparison. As expected, total GAWBS noise power depends linearly on the transmission distance. The red diamonds show the comparison where all the measurements are scaled back to 1000 km. The loop measurement results starts deviating from the single-span measurements beyond 4 loops due to the low OSNR, however, overall all the measurements are within 0.3 dB of the single-span measurement. Interestingly, the magnitude of the total measured GAWBS noise scaled at 1000 km which is 29.9 dB is consistent with the value estimated based on inverse generalized signal to noise ratio method [27].

 figure: Fig. 14.

Fig. 14. Total GAWBS noise scattering ratio as a function of transmission distance (blue). The same measurement is scaled to 1000 km and replotted on the right axis (red).

Download Full Size | PDF

4. Comparison of theory and measurements

In Section 2, it was claimed that the magnitude and the spectral shape of the GAWBS noise can be modeled using Eqs. (43) and (44). Another conclusion was that according to Eq. (39) TR$_{2m}$ modes create GAWBS peaks in the parallel polarization that is half of the magnitude created in the orthogonal direction. In this section it will be shown that these and other theoretical estimations match very well with the measurements. First the relationship between the parallel and orthogonal GAWBS contributions from the TR$_{2m}$ modes is analyzed.

Figure 15 shows both parallel and orthogonal GAWBS noise spectra measured for the SMF28_80 fiber in the same graph up to 470 MHz. As explained above, R$_{0m}$ acoustic modes generate GAWBS noise only in the parallel polarization however, TR$_{2m}$ modes generate GAWBS noise in both polarizations. Some of the TR$_{2m}$ modes overlap with the R$_{0m}$ modes and therefore they are difficult to identify in the parallel noise, however, others, especially in the frequency window not too far from the carrier can be clearly identified as it can be seen in Fig. 15. For some of the modes such as TR$_{24}$ and TR$_{29}$ that are located in the valleys of the polarized GAWBS noise, ratio of the power in the orthogonal peak and the parallel peak is very close to 2 confirming the prediction of Eq. (39).

 figure: Fig. 15.

Fig. 15. GAWBS noise spectra in the parallel (blue) and orthogonal polarizations (red). First ten GAWBS peaks originating from the R$_{0m}$ modes, and some of the TR$_{2m}$ modes that can be distinguished in the parallel polarizations are pointed out by arrows.

Download Full Size | PDF

The measured GAWBS noise is separated into parallel and orthogonal contributions, as they can be measured directly. By subtracting half of the orthogonal GAWBS noise from the parallel GAWBS noise in accordance with Eq. (39), the measurements can instead be separated into polarized and depolarized contributions which are generated by R$_{0m}$ and TR$_{2m}$ modes, respectively. Figure 16 shows the polarized GAWBS contribution obtained in this fashion for the four fiber types. To the best of our knowledge, this is the first time the measured polarized GAWBS noise spectrum is reported. It should be noted that even though a factor of 2 is predicted between the parallel and orthogonal parts of the depolarized contribution, factors of 2.05, 2.08, 2.13, and 2.13 were found to be a better fit for the fibers SMF28_80, Ex2_40, ZPL_60 and Ex3_60, respectively, which we attribute to measurement uncertainties. The depolarized contribution is not shown as it is simply the orthogonal contribution multiplied by a factor of 1.5. It can be seen that the polarized contribution is simply the peaks generated by the R$_{0m}$ peaks which are spaced quite periodically beyond the first couple of peaks as expected. In the case of Ex3_60, there are small but noticeable peaks remaining in between the R$_{0m}$ peaks beyond 200 MHz, and to a lesser degree for SMF28_80. These peaks do not correspond to GAWBS peaks originating from TR$_{2m}$ modes and therefore they do not contradict the conclusion of Eq. (39). It is presumed that these peaks originate from $TR_{1m}$ modes which can be noticeable when the condition of axial symmetry which was assumed in the theory section is broken, even for small core-cladding concentricity error [19].

 figure: Fig. 16.

Fig. 16. Polarized GAWBS noise spectra for the four fiber types.

Download Full Size | PDF

In order to compare the measurement results with Eqs. (43) and (44) the GAWBS noise power under each GAWBS peaks should be integrated since the equations only predict the total power created by each acoustic mode as acoustic damping is not taken into account. It was reported that polymer coating causes broadening of the GAWBS peaks with a Lorentzian shape [25,28]. To reduce the uncertainty of figuring out how much power is contained in each peak, GAWBS peaks are fitted with Lorentzian shapes defined as

$$f_L\left(f; A_p, \Delta f, f_0, \right) = \frac{A_p}{1 + \left\lbrack \frac{ \left(f - f_0 \right) } {\Delta f} \right\rbrack^2 },$$
where $A_p$, $\Delta f$, and $f_0$ are the peak value, width and the center frequency. The multi-peak fitting in linear scale is obtained with the following procedure. First the highest peak is found which determines the $f_0$. This peak is fitted using a least-square fitting algorithm with the width $\Delta f$ and the height of the Lorentzian shape $A_p$ as free parameters. Once a fitting to this single peak is obtained, it is subtracted from the measured spectrum. At this point algorithm goes back to the top to find the highest remaining peak for fitting until a predetermined number of peaks are fitted.

Figure 17 shows the multi-peak fitting obtained for polarized GAWBS noise measured for SMF28_80. Indeed, all eighteen peaks in the measurement window are very well fit with Lorentzian shape. The black solid curve in Fig. 17 shows the sum of the Lorentzian fits to the peaks. It fits very well to the measured spectrum except for the frequency range very close to the carrier. Note that the highest peaks i.e., ones from 100 MHz to 500 MHz, have a large peak-to-dip ratio. Therefore the Lorentzian fitting is dominated by the peaks and not the valleys, and even then the valley portion of the sum fits the measured spectrum very well. This shows that the valley portions are dominated by GAWBS noise and not other noise sources. This simultaneous good fitting of the peaks and valleys can be used as a sanity check on the measurement post-processing steps. To further illustrate this point, Fig. 18 shows the impact of an incorrect estimation of the laser noise on fitting the measurement results especially at the valleys. The red curve in Fig. 18 shows the estimated polarized GAWBS noise when the actual measured laser phase noise was used, and the black line shows the fitting from the multi-peak Lorentzian fitting, which are the same as in Fig. 17. Then, the laser phase noise is artificially increased or decreased by a large factor of 0.5 dB, in other words an error is introduced in the estimation of the laser phase noise resulting in the blue and yellow curves, respectively. Such a large error in laser phase noise estimation would change the total GAWBS noise estimation by 0.5 dB, and -0.2 dB, respectively. The error in the laser phase noise estimation affects the valleys significantly, however, the peaks dictate the actual fitting. The Lorentzian fitting could in principle be used as a secondary confirmation of correct estimation of the laser phase noise. It should be noted however that this is not a fool-proof way to confirm that the laser phase noise and other noise sources are removed correctly, as there are several assumptions made. The strictest of these assumptions is that all the GAWBS peaks have Lorentzian shapes which may not necessarily hold for peaks close to the carrier which also happen to be where the impact of laser phase noise is the largest. It was suggested [25] that the GAWBS peaks closest to the carrier would have the largest impact from the polymer coating since the radial component of the R$_{0m}$ modes, $U_r$, have the largest amplitude at the cladding-coating boundary. Figure 19(a) shows the normalized amplitude at the edge of the cladding as a function of the mode frequencies. Even though the dependence we obtained is not exactly the same as the one reported in [25], it indeed shows that the first peak has by far the largest magnitude. As such, it is reasonable to expect that the impact of the coating would be the strongest for this peak. At a first glance this seems to be contradicting our measurements as the first peak in our measurements has the lowest broadening, which can be seen in Fig. 19(b) which shows a close-up view of Fig. 17. Since the analysis presented in this paper does not model the impact of the coating, we can only speculate that due to stronger interaction with the coating in this frequency range, it is possible that the GAWBS peak may be strongly distorted and may have broader features that may be lost in the noise background. It is also evident in Fig. 19(b) that the first peak could not be fit very well with Lorentzian. Therefore it remains critical that the laser phase noise is characterized as accurately as possible to avoid introducing errors in this part of the GAWBS noise spectrum. The polarized GAWBS is calculated and the Lorentzian fits are performed for the rest of the measured fibers. The results are plotted in Fig. (20) which shows that the GAWBS peaks are fit very well with Lorentzian shapes for all the fibers measured.

 figure: Fig. 17.

Fig. 17. The polarized GAWBS spectrum measured for SMF28_80 (solid red) fitted with Lorentzian peaks (dashed lines). Solid black line shows the sum of all the individual Lorentzian fittings.

Download Full Size | PDF

 figure: Fig. 18.

Fig. 18. Red and black curves are the same as in Fig. (17). The polarized GAWBS noise is re-processed with the laser phase noise artificially shifted up (blue) and shifted down (yellow) by 0.5 dB.

Download Full Size | PDF

 figure: Fig. 19.

Fig. 19. (a) Normalized amplitude of the radial component of the displacement vector $U_r$ for the R$_{0m}$ acoustic modes at the cladding-coating boundary as a function of the mode frequency. (b) A zoomed-in view of Fig. 17.

Download Full Size | PDF

 figure: Fig. 20.

Fig. 20. (a) Normalized amplitude of the radial component of the displacement vector $U_r$ for the R$_{0m}$ acoustic modes at the cladding-coating boundary as a function of the mode frequency. (b) A zoomed-in view of Fig. 20.

Download Full Size | PDF

It was proposed that GAWBS peaks could be used to estimate the fiber cladding diameter [28,29]. This can be seen from Eq. (4). Even though the major contribution to the broadening of the GAWBS peaks is expected to be from the damping from the coating [25,28], a variation in the cladding diameter can cause additional broadening. For small variations it can cause the width of GAWBS peaks to increase linearly with the peak frequency as outlined in Eq. (51) [25]

$$f_{0m} = \frac{V_s y_{0m}}{2 \pi a} \quad \Rightarrow \quad \lvert \Delta f_{0m} \rvert \propto \frac{\Delta a V_s y_{0m}}{2 \pi a^2} = \frac{\Delta a}{a} f_{0m},$$
where, $\Delta a$ stands for the random variations in the cladding radius along the fiber.

To determine whether an estimate of the cladding diameter variation can be obtained from the polarized GAWBS measurements, widths of the GAWBS peaks are plotted with respect to their peak frequencies for the four fiber types in Fig. 21. The variation in the width is far from linear in general, especially for the first few peaks where laser noise, or strong coupling to the polymer coating might be dominant, and the last peaks where the peaks are closer to the noise floor. However, between 100 MHz and 570 MHz, a linear trend is observable for all fiber types. Based on Eq. (51), variation in cladding size can be estimated from the slope of peak linewidth as a function of peak frequency. The slope is estimated by fitting in this region with a linear line using a least-mean square algorithm. The fitting lines are also included in Fig. 21 with dashed lines in the same color corresponding to the fibers. Table 4 tabulates the estimated cladding diameter variation for the four fiber types based on the fitting shown in Fig. 21 and Eq. (51). Even though the variation in the cladding diameter is typically not included in the data sheet for most fibers, our estimation is within reason considering that the specification sheet for SMF28_80 reports an uncertainty in the cladding diameter as $125 \pm 0.7 \mu$m [30]. It should be noted that the variation in the peak linewidth can also be induced by cladding ellipticity. Another source of error can be the variation in the degree of coupling between the acoustic modes and the polymer coating which could be correlated with the magnitude of the radial displacement at the cladding-coating boundary as noted in Fig. 19.

 figure: Fig. 21.

Fig. 21. Width of the Lorentzian fitting $\Delta f$ as a function of GAWBS peak frequency for four fiber types (solid circles). For each fiber, a linear fit between 100 MHz and 560 MHz is included (dashed).

Download Full Size | PDF

Tables Icon

Table 4. Estimated variation in the cladding radius $\Delta a$ and the corresponding coefficient of determination $R^2$ obtained for the linear fitting.

Similar to the case of polarized GAWBS, depolarized GAWBS noise can also be fitted with Lorentzian shape. However, the TR$_{2m}$ modes responsible for the depolarized GAWBS are much less separated, and many modes are quite close to degeneracy. Moreover, some of the modes have small overlap with the optical mode and do not create GAWBS noise as efficiently. Therefore, more ambiguity remains in mapping the peaks to the TR$_{2m}$ modes compared to the R$_{0m}$ modes. Nevertheless, peaks that could be distinguished are fitted and shown in Fig. 22. It shows that the standalone peaks are very well fit, and even some of the slightly overlapping peaks are well resolved into two Lorentzian peaks. Similar to the polarized case, the fitting is not as good close to the carrier and at the edge.

 figure: Fig. 22.

Fig. 22. depolarized GAWBS noise spectra for four fiber types (solid red) including the individual Lorentzian fittings (dashed) and their sum (solid black).

Download Full Size | PDF

Comparison of Fig. 16 with Fig. 7 shows that width of the peaks are not as uniform but fluctuates. This is due to larger variation in the coupling between the TR$_{2m}$ modes and the polymer coating compared to R$_{0m}$ modes. As mentioned above the level of coupling to the polymer coating layer and therefore broadening of the peaks are correlated with how strong the radial component of the displacement vector $U_r$ at the edge of the cladding [25]. For R$_{0m}$ modes, the dependence of magnitude of $U_r$ at the cladding edge on the mode frequencies is quite smooth and monotonically decreasing for increasing frequency, which is not the case for the TR$_{2m}$ modes. The magnitude of $U_r$ at the cladding edge for R$_{0m}$ and TR$_{2m}$ modes are compared in Fig. 23(a) without normalizing, but using the same fiber parameters. For R$_{0m}$ modes the magnitude almost flattens out beyond 100 MHz although it never stops decreasing. In comparison, for TR$_{2m}$ modes, there is one set of modes following along with the R$_{0m}$ modes and another at a lower level. This is very well correlated with the width of the GAWBS peaks as shown in Fig. 23(b). Beyond 100 MHz, the linewidth for GAWBS peaks generated by R$_{0m}$ modes almost flattens, while the linewidth for the TR$_{2m}$ modes are split into two rails, i.e. the upper rail above 4 MHz and the lower below 3 MHz, well correlated with the amplitude of the $U_r$ for the corresponding mode. Moreover, a similar level of radial amplitude at the cladding causes a similar level of broadening for both R$_{0m}$ and TR$_{2m}$ peaks. This observation of clear correlation strengthens the claim that coupling to the polymer coating causes broadening of the GAWBS peaks. The radial amplitude for both upper and lower rail of the TR$_{2m}$ modes keeps reducing with increasing frequency, however, peak widths increase slightly showing that the variation in the cladding diameter might be larger than estimated in Table 4. It should be noted that below 100 MHz where the coupling is the strongest, the peak widths are reduced sharply which shows in this region the coupling is affecting the peaks beyond just broadening them. As explained above, some of the TR$_{2m}$ modes are missing in Fig. 23(b) since they could not be resolved, or have too small overlap with the optical mode, and therefore cold not be measured.

 figure: Fig. 23.

Fig. 23. (a) Radial component of the displacement vector $U_r$ for both R$_{0m}$ modes (red diamonds) and TR$_{2m}$ modes (blue circles) at the cladding-coating boundary as a function of the mode frequency. (b) Width of the Lorentzian fitting $\Delta f$ as a function GAWBS peak frequency for both polarized (red diamond) and depolarized (blue circles) GAWBS noise.

Download Full Size | PDF

Now that the GAWBS peaks are fit with Lorentzian shapes, the noise power under each peak can be integrated to find the GAWBS scattering efficiency per GAWBS mode which can be compared directly with Eqs. (43) and (44). These two equations can be split into three parts each. The first is the coherence length terms in the parenthesis which is set to unity as it is assumed that $l_m = l_c$ and $l_c \ll l_p$. The second is the ratio in the square brackets which contains all the frequency dependence of the GAWBS scattering. This term is also the only term where the optical mode field enters into the equation through the mode coupling terms. The third part is the collection of terms in the first ratio related to the glass parameters and fundamental constants, namely, $k_B$, $T$, $n_0$, $k_0$, $L$, $\rho _0$, $a$, $V_d$, $V_s$, $p_p$ and $p_m$. Some of these terms are well known such as the optical frequency and fiber length. Others like acoustic and photo-acoustic glass parameters such as $V_d$, $V_s$, $p_p$ and $p_m$ are reported in the literature though with small variations [7,9,29,3133]. Optical mode field distribution has to be approximated based on known parameters of the fibers such as the effective area, and under the weakly-guiding simple-step index approximations. Rather than determining the most reasonable parameters and inserting them into Eqs. (43) and (44) to compare with the measurements, the best fitting parameters to the measurements will be determined, and then they will be compared with the specs, or the reported parameters. It will be left to the readers to decide how close the best fitting parameters are to the expected values.

The best fitting parameters for the measurements can be found out in a certain order. First, the frequency of the polarized and depolarized GAWBS modes obtained from measurements can be used together with Eqs. (2), (3), (4), and (46) to determine best fitting $V_d/a$ and $V_s/a$. Even though these two parameters can be fit jointly to the polarized and depolarized peaks simultaneously, it is found that it is easier to first obtain $V_d/a$ from the polarized peaks with $\alpha \approx 0.63$ as a starting point, and then using the depolarized peaks to obtain $V_s/a$ and iterating only a few times works very well. This works since the R$_{0m}$ modes depend much more strongly on $V_d$ being pure radial modes. $V_d/a$ and $V_s/a$ cannot be resolved further into $V_d$, $V_s$, and $a$. Table 5 shows the best fitting values for $V_d/a$ and $V_s/a$ for our measurements.

Tables Icon

Table 5. Parameters used for obtaining Fig. 24 in addition to those in Table 1 and Table 2

Next, the best fitting optical mode-field distribution can be determined. In Eqs. (43) and (44) the only terms that determine the frequency dependence are in the last term in the square brackets, namely the overlap factor between the optical and acoustic modes. With $V_d/a$ and $V_s/a$ fixed, this term depends now only on shape of the optical mode. Even though the optical mode profile can have a complicated structure, it was found that the simple LP$_{01}$ mode distribution under weakly guiding approximation as described in Eq. (48) provides a good fit to the measurement. As shown in Eq. (48), the mode field distribution can be determined from the refractive index profile of the fiber which were not available to the authors. Alternatively, effective area together with the fiber cut-off wavelength $\lambda _c$ can be used to uniquely determine the mode field distribution. Using the nominal effective areas of the fibers listed in Table 2 leaves the cut-off wavelength $\lambda _c$ as the only fitting parameter. Since at this stage only the frequency dependence is sufficient to find the best fitting optical mode parameters, the fitting can be done by normalizing both the measurement and theory to the maximum peak. Table 5 shows the best fitting $\lambda _c$ for the four fibers using the procedure above. Though the $\lambda _c$ values are reasonable for the rest of the fibers, the best fitting value for the ZPL_60 is not, since it implies multimodedness when it is safely single-moded in the measured wavelength. This is an artifact of the simplified optical mode model used in Eq. (48), which assumes a simple step-index profile whereas ZPL_60’s index profile is not [34]. In fact, it has a relatively smaller mode-field diameter for the same effective area compared to a fiber with a simple step-index profile [34].

Once the optical mode parameters are fixed, the parameters in the first ratio in Eqs. (43) and (44) can be determined. Assuming the uncertainty in the optical wavelength and fiber length are negligible, the following best fitting factors for the polarized and depolarized cases can be determined by fitting theory and the measurement in the least mean square sense:

$$F_G = \frac{n_0^6 p_p^2}{\rho_0 a^4 } \qquad F_{G\perp} = \frac{n_0^6 p_m^2}{\rho_0 a^4}$$
Note that, $F_G$ and $F_{G\perp }$ are obtained by replacing $V_d$ and $V_s$ by $a V_d/a$ and $a V_s/a$ in Eqs. (43) and (44) which reduces the level of uncertainty. The best fitting factors for the four fibers are listed in Table 5. The Table shows that the measured fibers have similar fitting parameters which is reasonable as the fibers are of similar type, i.e., silica core fibers. The parameters in Table 1 is an example of set of parameters that approximately satisfies the best fitting values in Table 5. As an example, assuming the nominal cladding diameter of $a=125$ $\mu$m, the average value of $V_d = 5834$ m/s and $V_s=3637$ m/s are quite similar to values reported for silica-core fibers with varying degrees of F dopant in the cladding: $V_d = \textrm{5738-5824}$ m/s and $V_s=\textrm{3615-3669}$ m/s [35]. Values reported by others [7,9,33] are still within 3%.

Finally all the best fitting parameters can be inserted into Eqs. (43) and (44) and compared with the measured GAWBS spectra. Figure 24 shows the measured (black diamond) and calculated (green star) polarized and orthogonal GAWBS power as a function of GAWBS peak frequency for the measured fibers using parameters in Tables 1, 2, and 5. The fitting between measurement and theory is very good especially for the polarized GAWBS. The largest discrepancy is for the first GAWBS peak where the impact of the damping from the coating is the largest. The fitting for the orthogonal GAWBS is not as good, and the major factor is that GAWBS peaks overlap and not all can be fit with Lorentzian peaks individually. Some of these peaks that could not be resolved in the measurements are marked with gray markers.

 figure: Fig. 24.

Fig. 24. GAWBS peak power as a function of peak frequency for the polarized case the figures on the left, and for the case of orthogonal GAWBS on the right for the measured fiber types. The black diamonds are the measured peaks, and the green stars show the values calculated. Gray stars show the peaks that that were expected in theory but could not be resolved accurately in the measurement.

Download Full Size | PDF

Considering that the first peak in the polarized GAWBS measurement could not be well approximated by a Lorentzian, and that the overlapping peaks in the orthogonal GAWBS could not be well separated, it is worth checking how well the total GAWBS noise can be estimated using theory. After all, $\Gamma _{GAWBS}$ is what matters in determining signal transmission quality degradation due to GAWBS. Since total GAWBS noise is obtained by integrating the GAWBS noise spectrum, additional errors arising from imperfections due to Lorentzian fitting, or inaccurate assumptions on the optical mode distribution can be eliminated. Figure 25 shows the integrated GAWBS noise power, as well its break down into the polarized and depolarized components. The x-axis corresponds to the effective areas of the measured fiber types. Figure 25(a) shows when the best fitting photo-elastic coefficient values $p_{11}=0.115$ and $p_{12}=0.266$ are used for calculation. These values are consistent with previously reported values listed in Table 6 even though none of them were measured under the same measurement conditions as ours, i.e., either different glass composition or a different wavelength. To see how much discrepancy would be introduced between measurement and theory by using previously reported photo-elastic coefficients Figs. 25(b) and 25(c) are made with those values. The rest of the parameters are same as in Tables 1, 2, and 5. It can be seen that using the model developed in Section 2 the total GAWBS noise can be estimated within 0.5 dB for fibers with effective areas varying between 80–150 $\mu$m$^2$ using previously reported photo-elastic coefficient values. We conclude that Eqs. (43) and (44) can be used to closely estimate the total GAWBS noise.

 figure: Fig. 25.

Fig. 25. Integrated GAWBS noise in measured polarized (blue diamond), depolarized (red diamond) and their total (black diamond) vs corresponding calculated GAWBS noise (dashed lines with corresponding colors) for best fitting photo-elastic coefficient (a). (b)–(d) Same as (a) but the photo-elastic coefficients from Table 6 are used for calculation.

Download Full Size | PDF

Tables Icon

Table 6. Previously reported photo-elastic coefficients vs this work

Considering the good fit between the theory and the measurements, we can circle back and revisit some of the assumptions and simplifications made. One simplification made was to ignore the impact of the polymer coating which clearly induces significant damping on the acoustic modes as shown in Figs. 21 and 23. It is interesting that the measured GAWBS noise still agrees well with the calculated GAWBS noise. This agreement suggests that even though the acoustic modes couple to the coating, its impact is limited to broadening but the total power remains unchanged as dictated by the equipartition principle in Eq. (40).

Another assumption was that $l_c \ll l_p$. Equations (33) and (39) show that the the ratio $l_p/l_c$ affects the ratio of $\Gamma _{GAWBS,D}/\Gamma _{GAWBS,P}$. The simultaneous good fitting for both $\Gamma _{GAWBS,D}$, and $\Gamma _{GAWBS,P}$ in Fig. 25 confirms this assumption for the measured fibers. Regardless of the fiber type, the difference between the polarized and depolarized GAWBS noise is about 2.4 dB. For a fiber with $l_p \ll l_c$, Eq. (32) predicts $\Gamma _{GAWBS} \approx \Gamma _{GAWBS,P}$ which corresponds to a $\approx 2$ dB reduction in GAWBS total GAWBS noise per Fig. 25. If we consider the example of a long-distance transmission link with a fiber like EX3_60 with span length of 60 km and operation close to the optimum power, a 2 dB reduction in the GAWBS noise would correspond to $\approx 0.2$ dB improvement in the received generalized signal-to-noise ratio [1,3,37]. In this simple estimation, the nonlinear noise is approximated to be half of the amplified spontaneous emission noise, and the optimum power is assumed to be $\approx 1$ dBm per 50 GHz for a fully occupied C-band transmission.

Though the acoustic mode correlation length $l_c$ may not be known, there are well-established techniques for reducing the polarization diffusion length $l_p$ such as maintaining a long birefringence correlation length while drawing the fiber [18,3840]. Alternatively, increasing local birefringence is another way to reduce the polarization coherence length [18,38]. These methods may not have been preferred in the past as it would lead to higher polarization-mode dispersion [41]. However, modern coherent receivers can handle large levels of differential group delay [3].

Most of the theory laid out in Section 2 focused on the special case of single-mode single core fibers with the cores at the center of the cladding since this is the case for most typical transmission fibers. Recently multicore fibers have been considered for long-distance optical transmission as they can dramatically increase the number of cores that can be fit in a submarine cable [42,43]. It was reported recently that the multicore fibers have similar GAWBS noise levels as single-core fibers [6] with otherwise similar parameters. Since the measurement was reported for a particular multicore fiber design, it remains to be determined whether magnitude of total GAWBS noise has a dependence on the location of the cores.

The main difference in the case of multicore fibers is that the core is no longer at the cladding center. No matter how many cores a fiber has or how they are distributed, due to the circular symmetry of the fiber it would suffice to estimate the GAWBS noise dependence as a function of the radial offset of the core with respect to the fiber center. The main difference in the calculation of GAWBS is that, even if the mode field in a circular core would not have any angular dependence in the reference frame centered at the core, since the core is offset from the cladding center, in the coordinate system adopted in this paper, the mode-field distribution would have angular dependence. Therefore, in the case of multi-core fiber we need to revert to the two dimensional integrals in Eq. (34) instead of Eq. (35) where the angular and radial parts were separated. The loss of this symmetry means that the selection rules summarized in Eqs. (36) and (37) which limits GAWBS scattering to only modes with $n=0$, or $n=2$ no longer applies. In general many more acoustic modes contribute to GAWBS however each with smaller amounts.

Combining Eqs. (33), (34), (41), and (42), the GAWBS noise is calculated as a function of the radial offset r$_0$ between the cladding center and the core center, assuming the parameters of SMF28_80. Figure 26 summarizes the calculation results. The inset in Fig. 26(a) shows the schematic of a four-core fiber with the definition of the radial and angular offset parameters r$_0$ and $\theta _0$. Figure 26(a) shows how much different modes contribute to GAWBS noise depending on the radial offset. For the case of $r_0=0$, it can be seen that only modes $n=0$, and $n=2$ corresponding to R$_{0m}$ and TR$_{2m}$ contribute as explained above in detail. As the core offset increases, contribution from these two mode subsides as contribution from other modes increases. To make sure all the contributions are accounted for modes up to $n=50$ are calculated, even though it can be seen that no more than first 40 modes would be necessary. Figure 26(b) plots the same data, but this time it is plotted as a function of the offset for the first 9 modes. It shows that when the core is closer to the fiber center, only a few modes contribute the majority, and when it is farther more mores contribute evenly.

 figure: Fig. 26.

Fig. 26. Estimated GAWBS noise for a multicore fiber with the parameters of an SMF28_80 scaled to 1000 km. (a) Integrated GAWBS noise originating from different acoustic modes as a function of the mode number n for different values of radial offset of the core r$_0$ normalized to cladding radius a. The inset shows schematically definition of the radial and angular core offset r$_0$ and $\theta _0$, respectively. (b) Dependence of how much different acoustic modes contribute to GAWBS as a function of r$_0$/a. (c) Integrated GAWBS noise as a function of radial core offset with the blue, red and black lines showing the polarized, depolarized and total noise. (d) Dependence of total GAWBS noise originating from TR$_{2m}$ mode as a function of radial core offset. The GAWBS noise is separated into contribution from the degenerate mode with d=0 (blue) and d=1 (yellow) and their total (green).

Download Full Size | PDF

Figure 26(c) shows the major result of this calculation for the multicore fibers, that is, total GAWBS noise as a function of the radial offset and its break down into the polarized and depolarized components. It shows that the total GAWBS noise does not depend on the location of the cores, which agrees with the reported experimental result [6]. Interestingly, neither the polarized or the depolarized contributions have an appreciable dependency on the core offset. Even though no dependence on the angular offset is expected, it is still calculated and plotted in Fig. 26(d) as a sanity check. It shows the GAWBS noise originating from TR$_{2m}$ modes, however it is split into the contribution from the two cases of degeneracy $d=0$ and $d=1$, in other words the upper and lower solutions in Eq. (5). It shows that depending on the location of the core the overlap may oscillate with individual degenerate modes, but their sum is constant as expected. Since $n=2$ which corresponds to 2 isoclines in the acoustic modes as described in Eq. (1), 4 peaks per full rotations are observed as it should be the case.

5. Conclusion

Measurements of GAWBS noise for commonly used optical fibers types in long-distance optical communications were confirmed with well fitting theory. In particular, that the GAWBS noise scales inversely with the effective area. Furthermore, it is shown that the GAWBS scales linearly with transmission distance by measurements done on different span lengths, as well as using a recirculating loop. In theory this dependence was explained by introducing mode coherence length $l_c$. It is shown that due to GAWBS being a small and linear effect, and that the relative orientation of the polarization state of the optical field and the GAWBS mode is uniformly distributed along the fiber there is a constant ratio between the parallel and orthogonal contributions from the depolarized GAWBS noise. This result was used to separate the polarized and depolarized contributions to the measured GAWBS noise for the first time. This enabled us for the first time to make a direct comparison of measurement with theory. Furthermore, it was shown that the depolarized GAWBS noise can be suppressed by reducing the polarization diffusion length $l_p$ with respect to the $l_c$. It is pointed out that $l_p$ can be reduced by increasing the local birefringence or by reducing birefringence coherence length. Finally, the developed theory is used to show that multi-core fibers are most likely to have similar GAWBS noise level as single-core fibers with otherwise similar parameters.

Acknowledgments

The authors would like to thank Corning Inc. for fiber samples and Dr. John Downie, Dr. Dan Nguyen, and Dr. Paulo Dainese for insightful discussions.

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 reasonable request.

References

1. M. A. Bolshtyansky, J.-X. Cai, C. R. Davidson, M. Mazurczyk, D. Wang, M. Paskov, O. V. Sinkin, D. G. Foursa, and A. N. Pilipetskii, “Impact of spontaneous guided acoustic-wave brillouin scattering on long-haul transmission,” in Optical Fiber Communication Conference, (Optical Society of America, 2018), pp. M4B–3.

2. M. Paskov, M. A. Bolshtyansky, J.-X. Cai, C. R. Davidson, D. G. Foursa, and A. N. Pilipetskii, “Observation and compensation of guided acoustic-wave brillouin scattering in modulated channels,” in 2019 Optical Fiber Communications Conference and Exhibition (OFC), (IEEE, 2019), pp. 1–3.

3. E. R. Hartling, A. Pilipetskii, D. Evans, E. Mateo, M. Salsi, P. Pecci, and P. Mehta, “Design, acceptance and capacity of subsea open cables,” J. Lightwave Technol. 39(3), 742–756 (2021). [CrossRef]  

4. P. Serena, F. Poli, A. Bononi, and J.-C. Antona, “Scattering efficiency of thermally excited gawbs in fibres for optical communications,” in 45th European Conference on Optical Communication (ECOC 2019), (IET, 2019), pp. 1–4.

5. N. Takefushi, M. Yoshida, K. Kasai, T. Hirooka, and M. Nakazawa, “Theoretical and experimental analyses of gawbs phase noise in various optical fibers for digital coherent transmission,” Opt. Express 28(3), 2873–2883 (2020). [CrossRef]  

6. N. Takefushi, M. Yoshida, K. Kasai, T. Hirooka, and M. Nakazawa, “Experimental and theoretical analyses of gawbs phase noise in multi-core fiber for digital coherent transmission,” in Optical Fiber Communication Conference, (Optical Society of America, 2020), pp. T4J–3.

7. R. Shelby, M. Levenson, and P. Bayer, “Guided acoustic-wave brillouin scattering,” Phys. Rev. B 31(8), 5244–5252 (1985). [CrossRef]  

8. E. Sittig and G. Coquin, “Visualization of plane-strain vibration modes of a long cylinder capable of producing sound radiation,” J. Acoust. Soc. Am. 48(5B), 1150–1159 (1970). [CrossRef]  

9. R. N. Thurston, “Elastic waves in rods and clad rods,” J. Acoust. Soc. Am. 64(1), 1–37 (1978). [CrossRef]  

10. O. Florez, P. F. Jarschel, Y. A. Espinel, C. Cordeiro, T. M. Alegre, G. S. Wiederhecker, and P. Dainese, “Brillouin scattering self-cancellation,” Nat. Commun. 7(1), 11759–8 (2016). [CrossRef]  

11. J. F. Nye, Physical properties of crystals: their representation by tensors and matrices (Oxford University Press, 1985), chap. 9, pp. 150–168.

12. B. E. Saleh and M. C. Teich, Fundamentals of photonics (John Wiley & Son, 2019), pp. 216–217.

13. D. Marcuse, “Coupled-mode theory for anisotropic optical waveguides,” The Bell Syst. Tech. J. 54(6), 985–995 (1975). [CrossRef]  

14. R. Ulrich and A. Simon, “Polarization optics of twisted single-mode fibers,” Appl. Opt. 18(13), 2241–2251 (1979). [CrossRef]  

15. H. E. Engan, “Analysis of polarization-mode coupling by acoustic torsional waves in optical fibers,” J. Opt. Soc. Am. A 13(1), 112–118 (1996). [CrossRef]  

16. L. Palmieri and A. Galtarossa, “Coupling effects among degenerate modes in multimode optical fibers,” IEEE Photonics J. 6(6), 1–8 (2014). [CrossRef]  

17. J. Gordon and H. Kogelnik, “Pmd fundamentals: Polarization mode dispersion in optical fibers,” Proc. Natl. Acad. Sci. 97(9), 4541–4550 (2000). [CrossRef]  

18. D. Marcuse, C. Manyuk, and P. K. A. Wai, “Application of the manakov-pmd equation to studies of signal propagation in optical fibers with randomly varying birefringence,” J. Lightwave Technol. 15(9), 1735–1746 (1997). [CrossRef]  

19. F. Yaman, K. Nakamura, E. Mateo, S. Fujisawa, H. G. Batshon, T. Inoue, and Y. Inada, “Estimation of core-cladding concentricity error from gawbs noise spectrum,” in Optical Fiber Communication (OFC 2021), (OSA, 2021), p. M3C.6.

20. N. Nishizawa, S. Kume, M. Mori, T. Goto, and A. Miyauchi, “Characteristics of guided acoustic wave brillouin scattering in polarization maintaining fibers,” Opt. Rev. 3(1), 29–33 (1996). [CrossRef]  

21. J.-C. Beugnot, T. Sylvestre, H. Maillotte, G. Mélin, and V. Laude, “Guided acoustic wave brillouin scattering in photonic crystal fibers,” Opt. Lett. 32(1), 17–19 (2007). [CrossRef]  

22. P. Dainese, P. S. J. Russell, G. S. Wiederhecker, N. Joly, H. L. Fragnito, V. Laude, and A. Khelif, “Raman-like light scattering from acoustic phonons in photonic crystal fiber,” Opt. Express 14(9), 4141–4150 (2006). [CrossRef]  

23. D. Gloge, “Weakly guiding fibers,” Appl. Opt. 10(10), 2252–2258 (1971). [CrossRef]  

24. ITU-T, “Transmission systems and media, digital systems and networks,” Recommendation G.650.1 Series G, Geneva (2020).

25. A. J. Poustie, “Bandwidth and mode intensities of guided acoustic-wave brillouin scattering in optical fibers,” J. Opt. Soc. Am. B 10(4), 691–696 (1993). [CrossRef]  

26. T. Matsui, K. Nakajima, and F. Yamamoto, “Guided acoustic-wave brillouin scattering characteristics of few-mode fiber,” Appl. Opt. 54(19), 6093–6097 (2015). [CrossRef]  

27. J.-C. Antona, A. C. Meseguer, S. Dupont, R. Garuz, P. Plantady, A. Calsat, and V. Letellier, “Analysis of 34 to 101gbaud submarine transmissions and performance prediction models,” in Optical Fiber Communication Conference, (Optical Society of America, 2020), pp. T4I–3.

28. Y. Tanaka and K. Ogusu, “Tensile-strain coefficient of resonance frequency of depolarized guided acoustic-wave brillouin scattering,” IEEE Photonics Technol. Lett. 11(7), 865–867 (1999). [CrossRef]  

29. M. Ohashi, N. Shibata, and K. Shirakai, “Fibre diameter estimation based on guided acoustic wave brillouin scattering,” Electron. Lett. 28(10), 900–902 (1992). [CrossRef]  

30. “Smf-28 ull optical fiber product information sheet,”.

31. A. Bertholds and R. Dandliker, “Determination of the individual strain-optic coefficients in single-mode optical fibres,” J. Lightwave Technol. 6(1), 17–20 (1988). [CrossRef]  

32. X. Roselló-Mechó, M. Delgado-Pinar, A. Díez, and M. Andrés, “Measurement of pockels’ coefficients and demonstration of the anisotropy of the elasto-optic effect in optical fibers under axial strain,” Opt. Lett. 41(13), 2934–2937 (2016). [CrossRef]  

33. P.-C. Law, A. Croteau, and P. D. Dragic, “Acoustic coefficients of p 2 o 5-doped silica fiber: the strain-optic and strain-acoustic coefficients,” Opt. Mater. Express 2(4), 391–404 (2012). [CrossRef]  

34. K. Morita, Y. Yamamoto, T. Hasegawa, Y. Honma, K. Sohma, and T. Fujii, “Ultralow-loss large-core fiber for submarine cables,” SEI Tech. Rev. 85, 15–18 (2017).

35. K. Shiraki and M. Ohashi, “Sound velocity measurement based on guided acoustic-wave brillouin scattering,” IEEE Photonics Technol. Lett. 4(10), 1177–1180 (1992). [CrossRef]  

36. N. Borrelli and R. Miller, “Determination of the individual strain-optic coefficients of glass by an ultrasonic technique,” Appl. Opt. 7(5), 745–750 (1968). [CrossRef]  

37. V. V. Ivanov, J. D. Downie, and S. Makovejs, “Modeling of guided acoustic waveguide brillouin scattering impact in long-haul fiber optic transmission systems,” in 2020 European Conference on Optical Communications (ECOC), (IEEE, 2020), pp. 1–4.

38. I. P. Kaminow, “Polarization in optical fibers,” IEEE J. Quantum Electron. 17(1), 15–22 (1981). [CrossRef]  

39. A. Barlow, J. Ramskov-Hansen, and D. Payne, “Birefringence and polarization mode-dispersion in spun single-mode fibers,” Appl. Opt. 20(17), 2962–2968 (1981). [CrossRef]  

40. M. Li and D. Nolan, “Fiber spin-profile designs for producing fibers with low polarization mode dispersion,” Opt. Lett. 23(21), 1659–1661 (1998). [CrossRef]  

41. C. D. Poole, “Statistical treatment of polarization dispersion in single-mode fiber,” Opt. Lett. 13(8), 687–689 (1988). [CrossRef]  

42. E. Mateo, Y. Inada, T. Ogata, S. Mikami, V. Kamalov, and V. Vusirikala, “Capacity limits of submarine cables,” in Proc. SubOptic, (2016), pp. 1–3.

43. J. D. Downie, “Maximum cable capacity in submarine systems with power feed constraints and implications for sdm requirements,” in 2017 European Conference on Optical Communication (ECOC), (IEEE, 2017), pp. 1–3.

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 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 (26)

Fig. 1.
Fig. 1. (a) GAWBS peaks generated by R $_{0m}$ modes (filled markers), and peaks generated in the perpendicular polarization by TR $_{2m}$ modes (empty markers), for fiber cladding radius $a = 40 \mu m$ (blue) and $a = 62.5 \mu m$ (yellow) (b) Polarized (blue), depolarized (yellow) and total (green) GAWBS noise power as a function of cladding radius $a$ .
Fig. 2.
Fig. 2. Homodyne measurement setup. ITLA: integrated tunable laser, PM:polarization maintaining, EDFA: erbium-doped fiber, PC: polarization controller, LO: local oscillator PD: photo-diode, BPD: band-pass filter, ADC: analog-to-digital converter. Polarization maintaining paths and components are shown in blue. A and B show the input and output output points for the fiber under test (FUT).
Fig. 3.
Fig. 3. Spectrum in the polarization parallel to that of the carrier without the BPFs after 80 km of SMF-28 (black) and in back-to-back configuration (red).
Fig. 4.
Fig. 4. The parallel GAWBS scattering efficiency for SMF28_80 after various post-processing steps at a frequency resolution of (a) 9.5 kHz and (b) 503 kHz.
Fig. 5.
Fig. 5. Orthogonal GAWBS noise scattering efficiency measured for SMF28_80 at various stages of processing at a frequency resolution of (a) 9.5 kHz and (b) 503 kHz.
Fig. 6.
Fig. 6. Parallel GAWBS noise scattering efficiency measured for four different fiber types extrapolated to 1000 km.
Fig. 7.
Fig. 7. Orthogonal GAWBS noise scattering efficiency measured for four different fiber types extrapolated to 1000 km.
Fig. 8.
Fig. 8. (a) Curves in Fig. 6 plotted together. (b) Curves in Fig. 7 plotted together.
Fig. 9.
Fig. 9. GAWBS scattering efficiency on the left axis and the cumulative noise as a percentage of the total sum on the right axis obtained by a rolling sum of the spectrum form left to right for parallel (a) and orthogonal (b) contributions.
Fig. 10.
Fig. 10. Total GAWBS noise measured (black diamonds) for four fiber types including both parallel and orthogonal contributions plotted as a function of their effective area. Linear fitting (red) and theoretical estimation (blue) area also included. For theory part, $p_{11} = 0.115$ , and $p_{12} = 0.266$ are used for a better fit.
Fig. 11.
Fig. 11. Total measured GAWBS noise and its break down into the parallel and orthogonal contributions as a function of their lengths for Vascade Ex2000 fibers, i.e. Ex2_10_1, Ex2_10_2, Ex2_10_3, Ex2_10_4, Ex2_40, Ex2_48, and Ex2_84. The Ex2_40 is obtained by splicing the four 10-km pieces.
Fig. 12.
Fig. 12. (a) Recirculating loop setup. SW:switch, CPL:coupler, WSS: wavelength selective switch. (b) Received spectra after 1, 4, 8, 12, and 16 loops.
Fig. 13.
Fig. 13. GAWBS noise scattering efficiency after various distances in the loop setting compared to single span (dashed black). Even though measured at different distances, all measurement results are scaled to 1000 km for easier comparison (a) Parallel GAWBS noise. (b) Orthogonal GAWBS noise.
Fig. 14.
Fig. 14. Total GAWBS noise scattering ratio as a function of transmission distance (blue). The same measurement is scaled to 1000 km and replotted on the right axis (red).
Fig. 15.
Fig. 15. GAWBS noise spectra in the parallel (blue) and orthogonal polarizations (red). First ten GAWBS peaks originating from the R $_{0m}$ modes, and some of the TR $_{2m}$ modes that can be distinguished in the parallel polarizations are pointed out by arrows.
Fig. 16.
Fig. 16. Polarized GAWBS noise spectra for the four fiber types.
Fig. 17.
Fig. 17. The polarized GAWBS spectrum measured for SMF28_80 (solid red) fitted with Lorentzian peaks (dashed lines). Solid black line shows the sum of all the individual Lorentzian fittings.
Fig. 18.
Fig. 18. Red and black curves are the same as in Fig. (17). The polarized GAWBS noise is re-processed with the laser phase noise artificially shifted up (blue) and shifted down (yellow) by 0.5 dB.
Fig. 19.
Fig. 19. (a) Normalized amplitude of the radial component of the displacement vector $U_r$ for the R $_{0m}$ acoustic modes at the cladding-coating boundary as a function of the mode frequency. (b) A zoomed-in view of Fig. 17.
Fig. 20.
Fig. 20. (a) Normalized amplitude of the radial component of the displacement vector $U_r$ for the R $_{0m}$ acoustic modes at the cladding-coating boundary as a function of the mode frequency. (b) A zoomed-in view of Fig. 20.
Fig. 21.
Fig. 21. Width of the Lorentzian fitting $\Delta f$ as a function of GAWBS peak frequency for four fiber types (solid circles). For each fiber, a linear fit between 100 MHz and 560 MHz is included (dashed).
Fig. 22.
Fig. 22. depolarized GAWBS noise spectra for four fiber types (solid red) including the individual Lorentzian fittings (dashed) and their sum (solid black).
Fig. 23.
Fig. 23. (a) Radial component of the displacement vector $U_r$ for both R $_{0m}$ modes (red diamonds) and TR $_{2m}$ modes (blue circles) at the cladding-coating boundary as a function of the mode frequency. (b) Width of the Lorentzian fitting $\Delta f$ as a function GAWBS peak frequency for both polarized (red diamond) and depolarized (blue circles) GAWBS noise.
Fig. 24.
Fig. 24. GAWBS peak power as a function of peak frequency for the polarized case the figures on the left, and for the case of orthogonal GAWBS on the right for the measured fiber types. The black diamonds are the measured peaks, and the green stars show the values calculated. Gray stars show the peaks that that were expected in theory but could not be resolved accurately in the measurement.
Fig. 25.
Fig. 25. Integrated GAWBS noise in measured polarized (blue diamond), depolarized (red diamond) and their total (black diamond) vs corresponding calculated GAWBS noise (dashed lines with corresponding colors) for best fitting photo-elastic coefficient (a). (b)–(d) Same as (a) but the photo-elastic coefficients from Table 6 are used for calculation.
Fig. 26.
Fig. 26. Estimated GAWBS noise for a multicore fiber with the parameters of an SMF28_80 scaled to 1000 km. (a) Integrated GAWBS noise originating from different acoustic modes as a function of the mode number n for different values of radial offset of the core r $_0$ normalized to cladding radius a. The inset shows schematically definition of the radial and angular core offset r $_0$ and $\theta _0$ , respectively. (b) Dependence of how much different acoustic modes contribute to GAWBS as a function of r $_0$ /a. (c) Integrated GAWBS noise as a function of radial core offset with the blue, red and black lines showing the polarized, depolarized and total noise. (d) Dependence of total GAWBS noise originating from TR $_{2m}$ mode as a function of radial core offset. The GAWBS noise is separated into contribution from the degenerate mode with d=0 (blue) and d=1 (yellow) and their total (green).

Tables (6)

Tables Icon

Table 1. Parameters used for calculating GAWBS noise

Tables Icon

Table 2. Fibers and some of their parameters that are measured.

Tables Icon

Table 3. Total GAWBS noise power measured relative to the carrier power for four types of fiber broken down into parallel and orthogonal contributions.

Tables Icon

Table 4. Estimated variation in the cladding radius Δ a and the corresponding coefficient of determination R 2 obtained for the linear fitting.

Tables Icon

Table 5. Parameters used for obtaining Fig. 24 in addition to those in Table 1 and Table 2

Tables Icon

Table 6. Previously reported photo-elastic coefficients vs this work

Equations (52)

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

U r ( r , ϕ ) = C n m { A 1 [ J n + 1 ( ρ ) + J n 1 ( ρ ) ] + α A 2 [ J n + 1 ( α ρ ) J n 1 ( α ρ ) ] } Θ d ( n ϕ ) , U ϕ ( r , ϕ ) = C n m { A 1 [ J n + 1 ( ρ ) J n 1 ( ρ ) ] + α A 2 [ J n + 1 ( α ρ ) + J n 1 ( α ρ ) ] } Ψ d ( n ϕ ) ,
| B 11 B 12 B 21 B 22 | = 0 ,
B = [ ( n 2 1 y 2 2 ) J n ( α y ) ( n ( n 2 1 ) y 2 2 ) J n ( y ) ( n 2 1 ) y J n + 1 ( y ) ( n 1 ) J n ( α y ) α y J n + 1 ( α y ) ( n ( n 2 1 ) y 2 2 ) J n ( y ) + y J n + 1 ( y ) ] ,
y = Ω a V s .
Θ d ( n ϕ ) = { cos ( n ϕ ) , d = 0 sin ( n ϕ ) , d = 1 , Ψ d ( n ϕ ) = { sin ( n ϕ ) , d = 0 cos ( n ϕ ) , d = 1 .
S r r = U r r , S ϕ ϕ = 1 r U ϕ ϕ + U r r , S r ϕ = 1 2 ( 1 r U r ϕ + U ϕ r U ϕ r ) ,
S r r = C n m Ω n m V s s r r ( r ) Θ ( n ϕ ) sin ( Ω n m t + ζ ) , S ϕ ϕ = C n m Ω n m V s s ϕ ϕ ( r ) Θ ( n ϕ ) sin ( Ω n m t + ζ ) , S r ϕ = C n m Ω n m V s s r ϕ ( r ) Ψ ( n ϕ ) sin ( Ω n m t + ζ ) ,
s r r = 1 2 { A 1 [ J n 2 ( ρ ) J n + 2 ( ρ ) ] + α 2 A 2 [ 2 J n ( α ρ ) J n 2 ( α ρ ) J n + 2 ( α ρ ) ] } , s ϕ ϕ = 1 ρ { A 1 [ ( n + 1 ) J n + 1 ( ρ ) ( n 1 ) J n 1 ( ρ ) ] + α A 2 [ ( n + 1 ) J n + 1 ( α ρ ) + ( n 1 ) J n 1 ( α ρ ) ] } , s r ϕ = 1 2 ρ { A 1 [ ( n + 1 ) J n + 1 ( ρ ) + ( n 1 ) J n 1 ( ρ ) + ρ 2 [ 2 J n ( ρ ) J n + 2 ( ρ ) J n 2 ( ρ ) ] ] + α A 2 [ ( n 1 ) J n 1 ( α ρ ) ( n + 1 ) J n + 1 ( α ρ ) + α ρ 2 [ J n 2 ( α ρ ) J n + 2 ( α ρ ) ] ] } .
[ Δ η x x ( r , ϕ ) Δ η y y ( r , ϕ ) Δ η z z ( r , ϕ ) Δ η y z ( r , ϕ ) Δ η x z ( r , ϕ ) Δ η x y ( r , ϕ ) ] = [ p 11 p 12 p 12 0 0 0 p 12 p 11 p 12 0 0 0 p 12 p 12 p 11 0 0 0 0 0 0 p 11 p 12 0 0 0 0 0 0 p 11 p 12 0 0 0 0 0 0 p 11 p 12 ] [ S x x ( r , ϕ ) S y y ( r , ϕ ) S z z ( r , ϕ ) S y z ( r , ϕ ) S x z ( r , ϕ ) S x y ( r , ϕ ) ] .
Δ η i j ( r , ϕ , t ) ϵ 0 Δ ϵ i j ( r , ϕ , t ) ϵ 2 Δ ϵ i j Δ η i j ( r , ϕ , t ) n 0 4 ϵ 0
S x x = cos ( ϕ ) 2 S r r + sin ( ϕ ) 2 S ϕ ϕ sin ( 2 ϕ ) S r ϕ , S y y = sin ( ϕ ) 2 S r r + cos ( ϕ ) 2 S ϕ ϕ + sin ( 2 ϕ ) S r ϕ , S x y = sin ( 2 ϕ ) ( S r r S ϕ ϕ ) / 2 + cos ( 2 ϕ ) S r ϕ ,
Δ η x x ( r , ϕ , t ) = p p ( S r r + S ϕ ϕ ) + p m [ ( S r r S ϕ ϕ ) cos ( 2 ϕ ) 2 S r ϕ sin ( 2 ϕ ) ] , Δ η y y ( r , ϕ , t ) = p p ( S r r + S ϕ ϕ ) p m [ ( S r r S ϕ ϕ ) cos ( 2 ϕ ) 2 S r ϕ sin ( 2 ϕ ) ] , Δ η x y ( r , ϕ , t ) = p m [ ( S r r S ϕ ϕ ) sin ( 2 ϕ ) + 2 S r ϕ cos ( 2 ϕ ) ] ,
E t ( r , ϕ , z , t ) = ν h ν ( z ) E ν , t ( r , ϕ , t ) .
h μ ( z ) z = α 2 h μ + i k μ h μ + i ν κ μ ν ( t ) h ν ( z ) , i = 1
κ μ ν = i ω 4 P 0 2 π 0 E μ ( r , ϕ , t ) Δ ϵ ϵ ( r , ϕ , t ) E ν ( r , ϕ , t ) r d r d ϕ
P δ μ ν = 1 4 0 2 π 0 e ^ z [ E ν , t × H μ , t + E ν , t × H μ , t ] r d r d ϕ
H μ , t = c n 0 ϵ 0 e ^ z × E μ , t
κ μ ν = i k 0 n 0 3 2 0 2 π 0 E μ ( r , ϕ , t ) Δ η η ( r , ϕ , t ) E ν ( r , ϕ , t ) r d r d ϕ 0 2 π 0 E μ , t ( r , ϕ , t ) E ν , t ( r , ϕ , t ) r d r d ϕ ,
| h ( z ) z = α 2 | h ( z ) + i k ¯ | h ( z ) + i Δ k σ 1 | h ( z ) + i κ ¯ ( t ) | h ( z ) + i Δ κ ( t ) σ | h ( z ) , | h ( z ) = [ h x ( z ) h y ( z ) ] , σ 1 = [ 1 0 0 1 ] , σ 2 = [ 0 1 1 0 ] , σ 3 = [ 0 i i 0 ] ,
h z = 2 Δ κ × h ,
ζ ( z 1 ) ζ ( z 2 ) = { π 2 3 , | z 1 z 2 | l c 0 , | z 1 z 2 | > l c
| h ( z ) z = ( α 2 + i k ¯ + i b ( z ) σ ) | h ( z ) + ( i κ ¯ + i Δ κ σ ) | h ( z ) sin ( Ω n m t + ζ ( z ) ) ,
| h ( z ) = e α z 2 + i k ¯ z U ( z ) ( | h 0 + | δ h ( z ) ) , U ( z ) z = i b σ U ( z ) ,
| h 0 z 0 , | δ h ( z ) z i ( κ ¯ + Δ κ ( z ) σ ) | h 0 sin ( Ω n m t + ζ ( z ) ) ,
| δ h ( L ) = Δ z q = 1 N i ( κ ¯ + Δ κ q σ ) | h 0 sin ( Ω n m t + ζ q ) ,
P G ( L ) = Δ z 2 q = 1 N p = 1 N [ κ ¯ 2 P 0 + Δ κ q Δ κ p P 0 + κ ¯ ( Δ κ q + Δ κ p ) P 0 + i ( Δ κ q × Δ κ p ) P 0 ] sin ( Ω n m t + ζ q ) sin ( Ω n m t + ζ p ) ,
P G ( L ) = { P 0 L 2 ( κ ¯ 2 + Δ κ 2 ) l c , l c l p P 0 L 2 ( κ ¯ 2 + Δ κ 2 l p l c ) l c , l c l p .
P G ( L ) = Δ z 2 P 0 q = 1 N p = 1 N { [ κ ¯ 2 + ( Δ κ q p ^ ) ( Δ κ p p ^ ) + κ ¯ ( Δ κ q + Δ κ p ) p ^ ] sin ( Ω n m t + ζ q ) sin ( Ω n m t + ζ p ) } , P G ( L ) = Δ z 2 P 0 q = 1 N p = 1 N { [ ( Δ κ q Δ κ p ) ( Δ κ q p ^ ) ( Δ κ p p ^ ) + i ( Δ κ q × Δ κ p ) p ^ ] sin ( Ω n m t + ζ q ) sin ( Ω n m t + ζ p ) } .
P G ( L ) = l c 2 P 0 q = 1 N p = 1 N { [ κ ¯ 2 + Δ κ 2 cos ( θ q ) cos ( θ p ) + κ ¯ Δ κ cos ( θ q ) + cos ( θ p ) ] sin ( Ω n m t + ζ q ) sin ( Ω n m t + ζ p ) } , P G ( L ) = l c 2 P 0 q = 1 N p = 1 N { [ ( Δ κ q Δ κ p ) Δ κ 2 cos ( θ q ) cos ( θ p ) + i ( Δ κ q × Δ κ p ) p ^ ] sin ( Ω n m t + ζ q ) sin ( Ω n m t + ζ p ) } .
P G ( L ) = P 0 L 2 ( κ ¯ 2 + 1 3 Δ κ 2 ) l c , P G ( L ) = P 0 L 2 ( 2 3 Δ κ 2 ) l c ,
P G ( L ) = l c 2 P 0 u = 1 N v = 1 N κ ¯ 2 sin ( Ω n m t + ζ u ) sin ( Ω n m t + ζ v ) + l p 2 P 0 q = 1 M p = 1 M { [ κ ¯ 2 + Δ κ 2 cos ( θ q ) cos ( θ p ) + κ ¯ Δ κ cos ( θ q ) + cos ( θ p ) ] sin ( Ω n m t + ζ q ) sin ( Ω n m t + ζ p ) } , P G ( L ) = l p 2 P 0 q = 1 M p = 1 M { [ ( Δ κ q Δ κ p ) Δ κ 2 cos ( θ q ) cos ( θ p ) + i ( Δ κ q × Δ κ p ) p ^ ] sin ( Ω n m t + ζ q ) sin ( Ω n m t + ζ p ) } ,
P G ( L ) = P 0 L 2 ( κ ¯ 2 l c + 1 3 Δ κ 2 l p ) , P G ( L ) = P 0 L 2 ( 2 3 Δ κ 2 ) l p .
P G ( L ) = P 0 L 2 ( κ ¯ 2 + min [ l c , l p ] 3 l c Δ κ 2 ) l c , P G ( L ) = P 0 L 2 ( 2 min [ l c , l p ] 3 l c Δ κ 2 ) l c .
κ ¯ = κ 0 p p A Θ d ( n ϕ ) ( s r r + s ϕ ϕ ) f ¯ ( r , ϕ ) 2 r d r d ϕ , δ κ = κ 0 p m A [ Θ d ( n ϕ ) cos ( 2 ϕ ) ( s r r s ϕ ϕ ) 2 Ψ d ( n ϕ ) sin ( 2 ϕ ) s r ϕ ] f ¯ ( r , ϕ ) 2 r d r d ϕ , κ x y = κ 0 p m A [ Θ d ( n ϕ ) sin ( 2 ϕ ) ( s r r s ϕ ϕ ) + 2 Ψ d ( n ϕ ) cos ( 2 ϕ ) s r ϕ ] f ¯ ( r , ϕ ) 2 r d r d ϕ ,
κ ¯ = κ 0 p p 2 π 0 a ( s r r + s ϕ ϕ ) f ¯ ( r ) 2 r d r 0 2 π Θ d ( n ϕ ) d ϕ , δ κ = κ 0 p m 2 π 0 a ( s r r s ϕ ϕ ) f ¯ ( r ) 2 r d r 0 2 π Θ d ( n ϕ ) cos ( 2 ϕ ) ϕ 2 κ 0 p m 2 π 0 a s r ϕ f ¯ ( r ) 2 r d r 0 2 π Ψ d ( n ϕ ) sin ( 2 ϕ ) d ϕ , κ x y = κ 0 p m 2 π 0 a ( s r r s ϕ ϕ ) f ¯ ( r ) 2 r d r 0 2 π Θ d ( n ϕ ) sin ( 2 ϕ ) ϕ + 2 κ 0 p m 2 π 0 a s r ϕ f ¯ ( r ) 2 r d r 0 2 π Ψ d ( n ϕ ) cos ( 2 ϕ ) d ϕ ,
d = 0 : κ ¯ = κ 0 p p 0 ( s r r + s ϕ ϕ ) f ¯ ( r ) 2 r d r , δ κ = 0 , κ x y = 0 , d = 1 : κ ¯ = 0 , δ κ = 0 , κ x y = 0 .
d = 0 : κ ¯ = 0 , δ κ = κ 0 2 p m 0 ( s r r s ϕ ϕ 2 s r ϕ ) f ¯ ( r ) 2 r d r , κ x y = 0 , d = 1 : κ ¯ = 0 , κ x y = κ 0 2 p m 0 ( s r r s ϕ ϕ 2 s r ϕ ) f ¯ ( r ) 2 r d r , δ κ = 0 .
Δ κ 2 = κ 0 2 p m 2 2 [ 0 ( s r r s ϕ ϕ 2 s r ϕ ) f ¯ ( r ) 2 r d r ] 2 .
P G = P G P + 1 3 P G D P G = 2 3 P G D P G P = P G 1 2 P G P G D = 3 2 P G .
E n m = 0 l m 0 2 π 0 a 1 2 ρ 0 Ω n m 2 ( U r - n m 2 + U ϕ - n m 2 ) r d r d ϕ d z = k B T ,
C n m 2 = 2 k B T χ π l m ρ 0 a 2 Ω n m 2 D n m , χ = { 2 , n = 0 1 , n 0 ,
D n m = A 1 2 [ J n 1 ( y ) 2 + J n + 1 ( y ) 2 J n ( y ) [ J n + 2 ( y ) + J n 2 ( y ) ] ] + α 2 A 2 2 [ J n 1 ( α y ) 2 + J n + 1 ( α y ) 2 J n ( α y ) [ J n + 2 ( α y ) + J n 2 ( α y ) ] ] + A 1 A 2 2 α y ( α 2 1 ) [ 2 n ( α 2 1 ) α J n ( y ) J n ( α y ) + J n 2 ( y ) J n 1 ( α y ) + J n + 2 ( y ) J n + 1 ( α y ) α [ J n 1 ( y ) J n 2 ( α y ) + J n + 1 ( y ) J n + 2 ( α y ) ] ] ,
P G P ( L , Ω 0 m ) = P 0 k B T n 0 6 p p 2 k 0 2 L 4 π ρ 0 a 2 V d 2 ( l c l m ) [ S ( Ω 0 m ) 2 J 1 ( Ω 0 m a V d ) J 0 ( Ω 0 m a V d ) J 2 ( Ω 0 m a V d ) ] , S ( Ω 0 m ) = 0 a J 0 ( Ω 0 m r V d ) f ¯ ( r ) 2 r d r ,
P G D ( L , Ω 2 m ) = P 0 k B T n 0 6 p m 2 k 0 2 L 2 π ρ 0 a 2 V s 2 ( min [ l c , l p ] l m ) [ S ( Ω 2 m ) 2 D 2 m ] , S ( Ω 2 m ) = 0 a [ A 1 J 0 ( Ω 2 m r V s ) V s 2 V d 2 A 2 J 0 ( Ω 2 m r V d ) ] f ¯ ( r ) 2 r d r ,
s r r + s ϕ ϕ = 2 α 2 A 2 J n ( Ω n m r V d ) , s r r s ϕ ϕ 2 s r ϕ = 2 [ A 1 J n 2 ( Ω n m r V s ) α 2 A 2 J n 2 ( Ω n m r V d ) ] D 0 m = 2 ( α A 2 ) 2 [ J 1 ( Ω 0 m a V d ) J 0 ( Ω 0 m a V d ) J 2 ( Ω 0 m a V d ) ] .
| B | = [ a 2 Ω 2 2 V s 2 J 2 ( a Ω V s ) ] [ a 2 Ω 2 2 V s 2 J 0 ( a Ω V d ) a Ω V d J 1 ( a Ω V d ) ] = 0 ,
Γ G A W B S = n , m = 0 γ G ( L , Ω n m ) , Γ G A W B S , x = n , m = 0 γ G x ( L , Ω n m ) ,
f ( r ) = { J 0 ( p r w c ) J 0 ( p ) , r w c K 0 ( q r w c ) K 0 ( q ) , r > w c p = w c k 0 n c 2 n 0 2 , q = w c k 0 n 0 2 n c l 2 ,
A e f f = 2 π [ 0 f ( r ) 2 r d r ] 2 0 f ( r ) 4 r d r .
f L ( f ; A p , Δ f , f 0 , ) = A p 1 + [ ( f f 0 ) Δ f ] 2 ,
f 0 m = V s y 0 m 2 π a | Δ f 0 m | Δ a V s y 0 m 2 π a 2 = Δ a a f 0 m ,
F G = n 0 6 p p 2 ρ 0 a 4 F G = n 0 6 p m 2 ρ 0 a 4
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.