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

Fabrication, characterization, and theoretical analysis of controlled disorder in the core of optical fibers

Open Access Open Access

Abstract

We present results of experimental and theoretical studies of polarization-resolved light transmission through optical fiber with disorder generated in its germanium-doped core via UV radiation transmitted through a diffuser. In samples longer than a certain characteristic length, the power transmitted with preserved polarization is observed to be distributed over all forward-propagating modes, as evidenced by the Rayleigh negative exponential distribution of the near-field intensity at the output surface of the fiber. Furthermore, the transmitted power becomes also equally distributed over both polarizations. To describe the optical properties of the fibers with the experimentally induced disorder, a theoretical model based on coupled-mode theory is developed. The obtained analytical expression for the correlation function describing spatial properties of the disorder shows that it is highly anisotropic. Our calculations demonstrate that this experimentally controllable anisotropy can lead to suppression of the radiative leakage of the propagating modes, so that intermode coupling becomes the dominant scattering process. The obtained theoretical expressions for the polarization-resolved transmission fit very well with the experimental data, and the information extracted from the fit shows that radiative leakage is indeed small. The reported technique provides an easy way to fabricate different configurations of controlled disorder in optical fibers suitable for such applications as random fiber lasers.

© 2011 Optical Society of America

1. Introduction

In recent years, there has been a considerable interest in optical disordered media. This is largely due to the new functionalities brought about when dis order is introduced into homogeneous and periodic systems. A random laser [1], where laser action is ensured by coherent feedback in disordered structures such as powders or porous crystals, is a striking example. In Ref. [2], the advantages of disordered systems in wireless communications of high information capacity have been shown. It has also been reported [3] that the disorder induced in nonlinear crystals can greatly improve the efficiency of operation of nonlinear optical devices. It appears that disordered media open numerous possibilities for applications in sensors, nanophotonics, and, more generally, in various light transmission systems.

Localization of electromagnetic radiation in strongly disordered random media has attracted great interest from both fundamental and practical points of view [4]. Studied in the optical as well as in the microwave spectral regions, the phenomenon of localization depends on the dimensionality of the system. In particular, in surface- [5, 6, 7] and volume- [8] disordered waveguides, it leads to arbitrarily small transmission, which diminishes exponentially with the length of the system. This disorder-induced confinement can be employed in such an application as a laser.

Disorder-induced confinement has been shown to lead to unusual, while at the same time, useful properties in photonic-crystal waveguides [9] and in optical fibers [10, 11, 12, 13, 14, 15]. An optical fiber is an extremely promising experimental system for random lasing applications [1]: Lizárraga et al. [15] reported coherent random lasing on randomly distributed Bragg gratings in single-mode optical fibers, whereas Turitsyn et al. [14] demonstrated an incoherent random lasing.

In this report we present experiments on the fabrication of random variations of the refractive index throughout the core of a Ge-doped multimode optical fiber, whose parameters can be controlled in our experimental setup. The characteristics of the disorder created are evaluated from an analysis of the intensity distribution of the near fields at the output of the fiber and by the analysis of the speckle size dependence of the total intensity of the transmitted light. The experimental results are compared and agreement is found, with the predictions of the coupled-mode theory, which is adapted to the particular type of volume disorder considered in this work. We show that by varying the correlation size of the disorder, scattering sufficiently strong for achieving the complete mixing of the forward-propagating modes can be achieved in a centimeter-length segment of multimode fiber. We also demonstrate that disorder with strongly anisotropic correlation function can lead to a dramatic suppression of radiative losses, so that coupling between modes becomes dominant. Thus, the scattering is much more efficient compared to the weak scattering of the material impurities as in, e.g., Ref. [14], and, unlike Bragg gratings, it is broadband in the propagation constant of a mode or the frequency of the light. The above properties of our system make it a promising candidate for fabrication of a compact multimode random fiber laser. This can be achieved by sandwiching the disordered segment of the fiber between two Bragg gratings, which would provide feedback.

The paper consists of the following sections. In Section 2, the experimental setup used for the fabrication of the fiber samples with disorder is described. In Section 3, we find the experimental and numerical results on the intensity distribution of light emerging from an optical fiber with different scales of the disorder. In Section 4, theoretical analysis of the optical properties of a fiber with speckled perturbations of the refractive index in its core is presented. Finally, discussion and outlook are presented in Section 5.

2. Fabrication of the Disorder

The experimental setup utilized for the fabrication of disorder in optical fibers is schematically depicted in Fig. 1. In our experiments, we employed a step-index optical fiber (PS1250/1500 of Fibercore) sensitized by Ge. The main parameters of the fiber are a core diameter of 7.66μm, cladding diameter of 125μm, and numerical aperture (NA) of 0.13, with the refractive indices of the core and the cladding being 1.463 and 1.457, respectively. The cutoff wavelength of the fiber with these parameters is about 1200nm. The disorder was introduced in the Ge-doped fiber core by exposing it to UV light from an intracavity frequency-doubled argon-ion laser (244nm) that passed through a cylindrical lens and a diffuser, creating, in this way, a speckle pattern in a plane parallel to the fiber axis. The light beam generated by the UV laser was initially expanded by a cylindrical lens with a focal length of 12cm in order to form an elliptical spot with desired dimensions at the diffuser plane. The beam transmitted through the diffuser was used for exposing the photosensitive fiber. Speckle, as the strongly fluctuating, grainy intensity pattern resulting from the interference of randomly scattered coherent waves, resulted in fluctuations of the illuminating UV intensity in the fiber core. An expression for the size of a speckle, Eqs. (6, 7, 8) is derived in Section 4 below. It depends on the distance between the diffuser and the fiber axis, D, the size of the illuminated region in the diffuser plane Lx,z, and the wavelength of the recording UV light λUV. Variations of D in the range 28mm and of Lx,z in the range of 810mm allowed us to obtain an average speckle size along the fiber axis between 200 and 600nm.

The length of each segment with the fabricated disorder was 12mm. The experimental geometry allowed us to record the segments with lengths up to 5cm. In order to achieve disorder with similar statistical parameters in each segment, the same exposure time was used for all segments, namely, about 10min at a mean power of the UV laser of about 100mW. We observed experimentally that after this exposure, the intensity distribution of the output probe light at the fiber output did not change. Every next segment with a random distribution of the refractive index was recorded directly after the preceding one. The total lengths of the fabricated disordered part (Ls) were 2, 4, 6, 8, 10, and 15cm.

After forming the disordered segment, we launched the probe beam of the He–Ne laser operated at λ=543nm into the fiber, and detected the image of the output intensity distribution by a CCD camera (ST-402ME SBIG). The selected wavelength 543nm of the probe beam ensured a low mode- number propagation regime, and corresponded to the sensitivity range of the CCD camera quite well. The light emerging from the fiber passed through the microscope objective ×100, which imaged the output end of the fiber on the CCD camera. In front of the CCD camera there was a polarizer utilized for characterization of the transmitted light.

3. Experimental Results

The resulting V parameter of the utilized fibers was 5.8171 at the probe wavelength, and the expected number of the guided LP modes is N=20. By varying the angle of incidence of the probe beam, different combinations of modes were excited and the corresponding near-field transmitted intensity was recorded. It appears that these measurements can be made quite reliably. Indeed, (i) the light polarization was preserved in the straight fiber without disorder and (ii) the ambient temperature was controlled by a special air conditioning system that excluded the fluctuation of the parameters of the fiber samples during measurements. At the input of the optical fiber, the polarized light goes through a half-wave plate and a linear polarizer. The output light was detected separately for both polarizations: (i) after passing through a polarizer of the same orientation as at the input (pp polarization) or (ii) perpendicularly polarized (ps polarization). We analyzed the output light of each polarization independently. The polarization extinction ratio of the laser source and the fiber output was measured in the linear transmission regime.

Examples of the intensity distribution of the light emerging from the fiber, obtained for different realizations of the disorder and for different angles of the incident beam with disordered segments of the fiber of 1 (a) and 2cm (b) length, are presented in Fig. 2. The left column presents results of pp-polarization measurements, and the right column presents results of ps-polarization measurements. Different realizations were obtained by slightly bending the disordered part of the fiber.

In Fig. 3, the ensemble-averaged intensities of the output light measured experimentally as functions of the length of the disordered parts of the fiber are presented. The averaging was performed over ten realizations. The solid and dashed curves are the fit with the theoretical expression Eqs. (24, 25) obtained from Eq. (19) in the Section 4. The theoretical and experimental results show excellent agreement.

4. Coupled-Mode Theory in Fibers with Speckled Perturbations of the Refractive Index

As was shown in Section 3, the random fluctuations of the refractive index imprinted in the core of the photosensitive fiber resulted in the mixing of different forward-propagating modes. To describe this process and to obtain the characteristic (mixing) length of the disordered segment of fiber, we employ the coupled-power method developed by Marcuse [16]. However, because the disorder induced by the speckle pattern (see Section 3) does not allow a factorization of the refractive index modulations into a product of a function of the transverse coordinates and a function of the longitudinal coordinate δn(x,y,z)δn(x,y)×f(z), the original derivation is not applicable. The goal of this section is to obtain a system of coupled-power equations applicable to the experimentally induced disorder. In the process of derivation we verify that coupling between the forward and backward propagating modes is negligible. We also give detailed estimates of the radiative loss due to scattering into the nonguiding modes. We show that because of the highly asymmetric correlation function of the disorder, the radiative loss is greatly reduced, so it becomes comparable to the coupling coefficients between guided modes.

4A. Statistical Properties of the Disorder

To begin our analysis, we need to obtain the statistical properties of the disorder, specifically, the two-point correlator of the fluctuations of the dielectric function δε(r)δε(r), where the angular brackets denote averaging over different realizations of disorder. Here we defined the fluctuation of the dielectric function δε(r)=ε(r)ε(r), which has the property δε(r)=0. We make an assumption that in the process of exposure to the ultraviolet (UV) radiation, the material in the fiber core remains in a linear regime, i.e.,

δε(r)δε(r)=δε2|A(r)A*(r)|2|A(r)|2|A(r)|2δε2|μ(r,r)|2,
where A(r) are statistically uniform complex field amplitudes of the UV light scattered by the diffuser. The amplitudes can be computed in the paraxial approximation with the help of the Fresnel diffraction integral, which propagates the fields delta-correlated in the plane of the diffuser; the procedure is described in Subsection 4.D of Ref. [17]. In our problem, we are interested in δε(r)δε(r) as a function of all three spatial coordinates, including both those perpendicular (x and z axes) and parallel (y axis) to the direction of the UV illumination. In the geometry considered, it is impossible to obtain such an expression in a compact form. To proceed, we assume that
μ(r,r)μ(rr)μ(xx,0,zz)μ(0,yy,0).
In this expression the first factor describes the correlation in the plane perpendicular to the UV propagation, whereas the second factor describes the depth of the speckle. The expressions for these functions can now be computed with the knowledge that the Gaussian UV laser beam is spread out by the cylindrical lens to cover the spot
I(x˜,z˜)exp[x˜2/Lx2z˜2/Lz2],
where x˜, z˜ denote the coordinates in the plane of the diffuser. The intensity distribution in Eq. (3) allows one to compute the Fresnel integrals [17], which define the correlation functions μ(xx,0,zz) and μ(0,yy,0) in Eq. (2). Performing the integrations, we obtain
|μ(xx,0,zz)|2=exp[(xxSx)2]×exp[(zzSz)2],
|μ(0,yy,0)|2=1(1+[πLx2λUVD2(yy)]2)1/2(1+[πLz2λUVD2(yy)]2)1/21(1+[yySy2]2)1/2.
The length Si was introduced to describe the spatial dimensions of the speckles:
Sx=λUVD2πncoreLx0.15λUVDLx,
Sy=3λUVD2πncoreLz20.38λUVD2Lz2,
Sz=λUVD2πncoreLz0.15λUVDLz,
where D denotes the distance from the diffuser to the fiber core during the exposure; all dimensions are scaled by the refractive index of the core; and LzLx is assumed in Eq. (5). Finally, by substituting Eqs. (4, 5) into Eq. (1), we obtain the desired expression for the second-order statistics of disorder introduced in imprinting the speckle pattern in the core of the photosensitive optical fiber:
δε(r)δε(r)δε2exp[(xxSx)2]1[1+(yySy2)2]1/2exp[(zzSz)2].
The parameter δε2=2ncoreΔnUV is related to the change in the refractive index ΔnUV due to the UV irradiation. We note that the above approxi mate expression remains valid for |yy|Sy. For |yy|Sy the factor omitted in Eq. (5) has to be also included to ensure that the function is normalizable.

4B. Derivation of Coupled-Power Equations

We begin our derivation of a system of coupled-power equations by expressing the electric field in terms of the linearly x- and y-polarized modes in the weakly guiding step-index fiber without disorder:

E(r)νcν(z)ei(ωtβνz)(Et,ν(x,y)+e^zEz,ν(x,y)).
Here the summation runs over all modes ν of the fiber, including the odd and even modes of both x (odd ν’s) and y polarizations (even ν’s), assumed to be normalized as
βν[Et,ν(x,y)·Et,ν(x,y)]dxdy=δνν,
where δνν is the Kronecker symbol. Equation (10) contains contributions from only forward-propagating modes. In Subsection 4C we will support this assumption by showing that the coupling coefficients into the backpropagating modes is negligible.

Further, in Eq. (10) both the transverse Et,ν(x,y) and the longitudinal e^zEz,ν(x,y) components of the individual modes are retained despite the smallness of the latter. As will be seen below, retaining the longitudinal components is crucial because it gives the dominant contribution to the coupling between the modes with the orthogonal polarizations. βν is the propagation constant of the νth mode, and cν(z) is its amplitude at position z along the fiber.

Following [16], we obtain the coupled amplitude equation

dcν(z)dz=νKνν(z)cν(z)ei(βνβν)z,
where
Kνν(z)=ω22c2δε(r)[Et,ν(x,y)·Et,ν(x,y)+Ez,ν(x,y)Ez,ν(x,y)]dxdy
are the amplitude coupling coefficients. The system of equations Eq. (12) can be used to obtain the solution for a particular realization of the random function δε(r). The ensemble-averaged information can be obtained by defining the power in each mode as Pν=|cν|2, which satisfies the evolution equation:
dPνdz=cν*dcνdz+c.c.,
where c.c. stands for the complex conjugate. We proceed by substituting Eqs. (12, 13) into Eq. (14). Evaluation of the ensemble average requires the following two assumptions. Pν(z) is assumed to vary on scales much larger than that of the disorder Szλ. This assumption is easily satisfied because the magnitude of the refractive index fluctuations is small—ΔnUV1. The experimental data in Fig. 3 further corroborate this assertion.

At this point, our derivation departs from that of Marcuse [16]. To evaluate |Kνν(z)|2, instead of the stringent requirement that the function describing the disorder in the refractive index can be factorized as δn(x,y,z)δn(x,y)×f(z), we use a much weaker assumption that z dependence is factorizable in δε(r)δε(r). Indeed, the multiplicative property of the correlation in the speckle in Eq. (9) that separates the dependencies on the transverse (x and y) and the longitudinal (z) coordinates, enables one to complete the derivation of the system of coupled-power equations

dPνdz=νhνν(PνPν),
with the power coupling coefficients given by the following expression
hνν=δε2ω4πlog(2)SxSySzc4eSz2|βνβν|2/4×[Et,ν(x,y)·Et,ν(x,y)+Ez,ν(x,y)Ez,ν(x,y)]2dxdy.
In obtaining Eq. (16) we approximated exp[(xx)2/Sx2]×[1+(yy)2/Sy2]1/2 by the product of two delta functions 4π1/2log(2)SxSyδ(xx)δ(yy) with the coefficients chosen so that both pairs of functions enclose identical area. This approximation is justified fairly well in our case because Sx,y are smaller than the characteristic scale, a, of the field variation in the transverse direction for all guided modes. In case of the function that describes y dependence, the full expression Eq. (5) was used to obtain the normalization, and the correction terms logarithmic in Lx/Lz were omitted in the result.

4C. Efficiency of Backscattering

In the process of derivation of the coupled power equations, Eqs. (15, 16), we neglected the possibility of scattering from a forward-propagating mode into one of the backward-propagating modes. This is an important process which, if efficient, can give rise to the phenomenon of Anderson localization, which originates in the studies of mesoscopic systems in condensed matter physics [18]. Multiple scattering and interference of the forward- and backward-propagating waves can suppress transmission and lead to an exponential decay of the transmission coefficient. This dependence may appear similar to that observed in Fig. 3.

To estimate the efficiency of the backscattering process in our system, we compute the forward-to-backward coupling coefficients. The derivation follows the steps similar to those in Subsection 4B, with the final expression for hνν+, being given by the formula similar to Eq. (16) with an exception that the exp[Sz2|βνβν|2/4]1 factor is replaced by exp[Sz2|βν+βν|2/4]1. One can see that this difference proves to be extremely important because |βνβν||βν+βν|2ncore×(2π/λ) and Szλ in our fibers.

The above estimate shows that the back scattering mechanism is, indeed, strongly suppressed in the considered system as it was assumed in the previous section. As a consequence, we do not expect our system to exhibit the phenomenon of Anderson localization.

4D. Radiative Losses

Optical fiber with unwanted or purposefully introduced, as in our case, modulations of the refractive index are invariably susceptible to the radiative losses. Indeed, the index nonuniformity couples the modes guided in the core of the fiber to the nonguided modes that extend into the cladding and are effectively lost. Even the fibers of the highest quality suffer from radiative loss from Rayleigh scattering on molecular inclusions introduced in its fabrication process [19]. The consequence of this loss is the exponential decay of the power in a mode Pν(z)exp[ανz]. Unlike the losses suffered in waveguides with rough surfaces, the radiative loss in the volume- disordered fibers, such as fibers with molecular defects, should not exhibit a strong dependence on the mode index ν. Because the fibers studied in this work are of the latter kind, we will assume ανα hereafter.

In Chap. 4 of Ref. [16], Marcuse has derived an expression for α in the case of Rayleigh scattering. It is interesting to note that under quite general conditions, the ratio between coupling coefficients and the scattering loss appears to be independent of the disorder parameters [20]

αhνν23πk02ncore2A,
where k0=2π/λ and A is the area of the fiber core. One can easily see that the above estimate gives α/hνν1 for a step-index fiber with (ncorencladding)/ncore1. Evaluating this ratio for our system gives a number on the order of a thousand. Although the above estimate is made under the assumption of Rayleigh scattering, it may still be applicable in our case. This is because the Rayleigh criterion involves not only the smallness of the scatterer compared to the wavelength of light but also the difference between its refractive index and that of the surrounding [21]. Below, we expose a flaw in this logic and show that Eq. (17) is not applicable to our system and that, instead, αhνν

Unlike a deterministic scattering off a single particle, the scattering in a random system has to properly account for the exact autocorrelation function given in our system by Eq. (9). The combined effect for a group of scatters can be greatly diminished if the phases of the partial waves are sufficiently random. Quantitatively, this effect is described [16] by the following integral

αI=dΩΔk(e^scat·e^z)2duxduyduzδε(r)δε(r+u)exp[iΔk·u].
Here, Δkncorek0(e^scate^z) defines the change of wavevector after scattering and dΩΔk denotes the solid angle integration over all possible scattering directions.

The Rayleigh approximation in Eq. (18) amounts to assuming that disorder is correlated in the vol ume Lcorr3 much less than λ3, which results in exp[iΔk·u]1. In the optical fibers with photoinduced disorder considered in our work, this assumption is no longer valid. Thus, the Rayleigh result I=(4π/3)δε2Lcorr3 needs to be reevaluated for the correlator Eq. (9) we obtained in Subsection 4A.

Calculation of the triple integral in Eq. (18) is facilitated by the fact that δε(r)δε(r+u) is factorizable into three functions, each of which depend only on one spatial variable. The integrals over ux and uz give rise to πSx,zexp[(Δkx,zSx,z/2)2]. The integral over uy does not give, in general, a compact expression. However, in a special case when Lx=Lz, it leads to a simple expression that illuminates the general tendency: πSyexp[ΔkySy]. Inspection of all three integrals shows that the result of the triple integral in Eq. (18) is a function that is very strongly peaked around |Δk|=0. Therefore, the remaining integration over solid angles should produce a result much smaller than 4π/3 predicted for the isotropic (Rayleigh) scattering. To complete our calculation of the absorption coefficient α we perform the integral over dΩ in Eq. (18) numerically and report the results in Fig. 5.

4E. Solution of Coupled-Power Equations

The system of coupled-power Eqs. (15) obtained in Subsection 4B did not account for loss. This omission can be rectified by a phenomenological correction due Marcuse [16]

dPνdz=αPν+νhνν(PνPν).
Such a treatment of loss can be rigorously justified in the case when the such a loss is independent of the mode index [22]. As already mentioned in the preceding section, this is a reasonable assumption for the volume-disordered fibers that we also adopt here.

Solution of Eqs. (19) proceeds with two steps. First, the effect of the radiative loss is factored out with substitution

Pν(z)=Pν(lossless)(z)×exp[αz],
which reduces Eqs. (19) back to Eqs. (15) satisfied now by Pν(lossless)(z).

In the second step, the solution for Pν(lossless)(z) is obtained by the following ansatz

Pν(lossless)(z)=Aνexp[σz],
where σn and the corresponding set of Aν(n) are to be determined by substitution of Eq. (21) into Eq. (15). Here σn are the eigenvalues of the secular equation
det[hννδνντhντ+σ]=0,
arranged in increasing order. The overall solution for Pν(z) takes the form
Pν(z)=eαz×[ncnAν(n)eσnz],with cn=[nAν(n)Pν(0)].
Because the effect of radiative loss has been factored out in Eq. (20), the conservation of the total power for Pν(lossless) requires σ10 and, subsequently, Aν(1)=const=1/N leads to uniform distribution of the power over all modes. The knowledge of σn allows estimation of the characteristic lengths of the disordered region of the fiber beyond which such an asymptotic state is achieved, (xx)σ31, and for cross-polarized modes, (xy)σ21. Assuming that the fiber is excited with some mode combination (with total input power equal to unity) of the same polarization, which we assume to be x for definitiveness, and recalling mode numbering convention in Eq. (10), we obtain
P(x)(z)ν=0N/21P2ν+1eαz×12[1+eσ2z],
P(y)(z)ν=1N/2P2νeαz×12[1eσ2z].
The above equations have the following properties. Without loss and polarization coupling, P(x)(z)=1 reflects power conservation. In the presence of absorption P(x)(z)+P(y)(z)=exp[αz] exhibits attenuation due to the radiative losses.

In the case when coupling between two orthogonal subsets of LP modes of the fiber is weak, σ2ασ3, Eq. (23) yields P(y)(z)0, P(x)(z)exp[αz] and the system reaches the state when the power is equally distributed only over (N/2) modes with initially excited polarization, x:

P2ν+1(z)eαz×[(P2ν+1(0)2N)eσ3z+2N],P2ν(z)0.
The above analysis shows that the redistribution of the power carried by the forward-propagating modes can be detected by making the following observations:
  • Making polarization-resolved measurement of the light intensity at the output surface of the fiber and averaging it over several disorder configurations should show that the intensity profile approaches the limit. Alternatively, the conclusion that a perfect mixing (in a statistical sense, i.e., Pν(z)const) indeed occurs in our experimental system can also be tested through measurements of the distribution of the near-field intensity at different spatial locations for just one realization of disorder. A random sum of different modes of the fiber νcνEt,ν(x,y) with Pνconst is expected [17] to result in the Rayleigh negative exponential distribution of the intensity. In an optical fiber, however, the coefficients cν are not completely random because the total power carried by all modes is constrained by ν|cν(z)|2=exp[αz]. This constraint, similar to the power conservation in the lossless fibers [23], makes the distribution deviate slightly from the Rayleigh form. As we can see from Fig. 4, the agreement between theory and experiment is very good, whereas the level of precision of the experimental data does not allow distin guishing between the two theoretical functions— unconstrained and constrained random sums of all modes of the fiber.
  • The dependence of P(x)(z=Ls),P(y)(z=Ls) on the length of the disordered segment of the fiber, Ls, is expected to be described by Eqs. (24, 25). We observe that the power carried by a particular mode ν, Pν=βν[Et,ν(x,y)·Et,ν(x,y)]dxdy is equal to the product of nearly ν-independent βνn1k0 and the field intensity integrated over the surface of the fiber I. Therefore, even in the case of superposition of several modes with the same polarization, the area- integrated intensity at the output facet I(x,y)=|νcνEt,ν(x,y)(x,y)|2dxdy=ν|cν|2|Et,ν(x,y)(x,y)|2dxdy(1/n1k0)νPν(x,y) is proportional to P(x)(z) and P(y)(z) given by Eqs. (24, 25). The outcome of the fit by these expressions to the experimental data in Fig. 3 allows one to extract the characteristic mixing length (xy)=σ21 and loss coefficient α.

Approximate expressions for the mixing lengths can be obtained in a compact analytic form by taking into account the fact that both the transverse Et,ν(x,y) and the longitudinal e^zEz,ν(x,y) modal profiles are spread out over the entire core of the fiber. This observation together with Eq. (11) allows one to estimate

[Et,ν(x,y)·Et,ν(x,y)+Ez,ν(x,y)Ez,ν(x,y)]2dxdy,
{[ncore2(ω2/c2)πa2]1xx,yy(NA/2)4[ncore2(ω2/c2)πa2]1xy,yx.
where a is the radius of the fiber core. In the second case of the cross-polarized modes, we also used the fact that the amplitude of the Ez,ν(x,y) component is a factor NA/2 smaller compared to the amplitude of the transverse fields. The approximations in Eq. (28), σ2,3h22,33 and Sz|βνβν|1 allow us to obtain our final result in a closed analytical form
(xx)1σ3Δn2ncoreπω2SxSySzc2a2=Δnncore4λUV3D4λ2a2LxLz3,(xy)1σ2mixing(xx)1(NA2)4,
where Eqs. (6, 7, 8) were used.

In Fig. 5a we plot the dependence of the speckle size as a function of the distance D between the diffuser and the fiber core. It is clear that tuning this parameter allows one to widely tune the characteristic size of the prepared disorder. This is an attractive feature of the fabrication technique described in Section 2.

In Fig. 5b three expressions for (xx) obtained in this section are compared. As expected, for a small speckle size (small D), the approximation of Eq. (9) by a product of delta functions appears to be justified and gives a quite accurate result when compared to the direct numerical evaluation of hνν.

Numerical evaluation of the exact expressions in Eq. (16, 22) with the experimentally rele vant parameters (Lx=0.5mm, Lz=3mm and D=0.5cm) yields σ3=(xx)10.15cm1, σ2=(xy)13×106cm1, and α0.015cm1.

5. Conclusion

We have studied the transmission of light through a volume-disordered multimode optical fiber. The disorder was introduced in the germanium-doped core of the fiber via UV radiation transmitted through a diffuser. The disorder generated in an optical fiber can be controlled by the experimental conditions, and it is determined by the speckle size and the value of the induced difference in the refractive index. The measurement of the transmission as a function of the length of the disordered section demonstrates the uniform distribution of the power over all forward-propagating modes beyond Ls=15cm. For long sections of a disordered fiber, the experimentally measured distribution of the near-field intensity at the output surface of the fiber is well described by the Rayleigh negative exponential function. The presented technique provides an easy way to fabricate different configurations of controlled disorder in optical fibers suitable for applications as a coherent and incoherent random fiber laser. Although the specific type of disorder studied in our work leads to mixing of only forward-propagating modes, the feedback necessary to produce laser action can be achieved by surrounding the disordered fiber with Bragg gratings.

Analysis of Fig. 3 shows that the power transfer into the cross-polarized modes occurs quite efficiently with (xy)(xx). This differs from the predictions of the coupled-mode theory developed in Section 4 that gives (xy)(xx). We attribute this enhanced cross polarization coupling to the (i) birefringence effect induced by the bending of the fiber, and (ii) strongly anisotropic disorder pattern defined by Eq. (9). Indeed, although no polarization mixing is observed in the blank fibers (before the disorder is introduced), to generate the statistical ensemble of different realization, the 30cm long fiber sample was displaced in the lateral directions, while both of the sample’s ends were fixed by fiber clips. As a result of fiber bending and tension, a pronounced birefringence was induced. For our experimental condition, we estimate the minimum radius of the bending as 250cm, which gives birefringence of Δn4×105 [24]. Formally, the induced birefringence enables coupling via the transverse components of the modes’ field that is expected to remove the small factor (NA/2)4, which leads to the (xy)(xx) condition in Eq. (28). The effect of induced birefringence will be reported in a separate publication.

We thank A. A. Maradudin and E. R. Méndez for stimulating discussions and constructive comments to the manuscript. N. P. Puente and E. I. Chaikina would like to acknowledge support by Consejo Nacional de Ciencia y Tecnología (Mexico), under grant UCM-42127. The work at Missouri University of Science & Technology was supported by the University of Missouri Research Board and by National Science Foundation grant DMR-0704981.

 figure: Fig. 1

Fig. 1 Experimental setup.

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 Examples of the output intensity distribution observed in some realizations with the disordered part of fiber (a) 1cm and (b) 2cm. The left column in each figure presents the pp-polarized distribution, and the right column presents the ps-polarized distribution. The angles of incidence are 0°, 2°, and 5° from the top to the bottom images.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Experimentally measured total co- (open symbols) and cross-polarized (solid symbols) transmission as a function of the length of the disordered part of the fiber for different polarizations of the transmitted beam. The circles correspond to an angle of incidence of 0°, the triangles to 2°, and the squares to 5°. Solid and dashed curves represent the theoretical fit with Eqs. (24, 25) with parameters α=0.064cm1, σ2=0.1917cm1.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 The distributions, which correspond to an unconstrained random sum (shown as a dashed curve) and to a constrained random sum (shown with the solid curve) of all modes of the fiber, are compared to the experimentally observed distributions of the near-field intensity measured in co- (circles) and cross-polarized (squares) channels in a sample with L=8cm. The thin symbols correspond to an angle of incidence of 2°, and the bold symbols correspond to an angle of incidence of 5°.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Panel (a) plots the size of the speckle defined by Eqs. (6, 7, 8) with Lx=0.3mm and Lz=2mm as a function of the distance between the diffuser and the core of the photosensitive fiber. Panel (b) compares the values of the characteristic length (xx)1σ3 after which all forward-propagating modes with one polarization become equally populated. It is found numerically from Eq. (16) without (solid curve) and with (circles) the delta function approximation to the order of magnitude estimate (squares) in Eq. (29). Panel (c) compares the amplitude of the radiative loss rate computed from Eq. (18) to the intermode coupling rate σ3. The plot shows that for the disorder patterns generated with D>4mm, the coupling becomes the dominant effect. This conclusion is borne out by the experimental results in Fig. 3.

Download Full Size | PDF

1. H. Cao, “Review on latest developments in random lasers with coherent feedback,” J. Phys. A 38, 10497–10535 (2005). [CrossRef]  

2. S. H. Simon, A. L. Moustakas, M. Stoytchev, and H. Safar, “Communication in a disordered world,” Phys. Today 54(9), 38–43 (2001). [CrossRef]  

3. S. E. Skipetrov, “Disorder is the new order,” Nature 432, 285–286 (2004). [CrossRef]   [PubMed]  

4. A. Lagendijk, B. van Tiggelen, and D. Wiersma, “Fifty years of Anderson localization,” Phys. Today 62(8), 24–29 (2009). [CrossRef]  

5. J. A. Sánchez-Gil, V. D. Freilikher, A. A. Maradudin, and I. Yurkevich, “Reflection and transmission of waves in surface disordered waveguides,” Phys. Rev. B 59, 5915–5925 (1999). [CrossRef]  

6. E. I. Chaikina, S. Stepanov, A. G. Navarrete, E. R. Méndez, and T. A. Leskova, “Formation of angular power profile via ballistic light transport in multi-mode optical fiber with corrugated surface,” Phys. Rev. B 71, 085419 (2005). [CrossRef]  

7. F. Bass, V. Freilikher, and I. Fuks, “Propagation in statistically irregular waveguides—part I: average field,” IEEE Trans. Antennas Propagat. 22, 278–288 (1974). [CrossRef]  

8. A. A. Chabanov, M. Stoytchev, and A. Z. Genack, “Statistical signatures of photon localization,” Nature 404, 850–853 (2000). [CrossRef]   [PubMed]  

9. J. Topolancik, F. Vollmer, and B. Ilic, “Random high-Q cavities in disordered photonic crystal waveguides,” Appl. Phys. Lett. 91, 201102 (2007). [CrossRef]  

10. O. Shapira and B. Fischer, “Localization of light in a random-grating array in a single-mode fiber,” J. Opt. Soc. Am. B 22, 2542–2552 (2005). [CrossRef]  

11. C. Lu, J. Cui, and Y. Cui, “Reflection spectra of fiber Bragg gratings with random fluctuations,” Opt. Fiber Technol. 14, 97–101 (2008). [CrossRef]  

12. C. J. S. Matos, L. de S. Menezes, A. M. Brito-Silva, M. A. Martinez-Gámez, A. S. L. Gomes, and C. B. de Araújo, “Random laser action in the core of a photonic crystal fiber,” Opt. Photon. News 19(12), 27–27 (2008). [CrossRef]  

13. M. Gagné and R. Kashyap, “Demonstration of a 3mW threshold Er-doped random fiber laser based on a unique fiber Bragg grating,” Opt. Express 17, 19067–19074 (2009). [CrossRef]  

14. S. K. Turitsyn, S. A. Babin, A. E. El-Taher, P. Harper, D. V. Churkin, S. I. Kablukov, J. D. Ania-Castñón, V. Karalekas, and E. V. Podivilov, “Random distributed feedback fibre laser,” Nature Photon. 4, 231–235 (2010). [CrossRef]  

15. N. Lizárraga, N. P. Puente, E. I. Chaikina, T. A. Leskova, and E. R. Méndez, “Single-mode Er-doped fiber random laser with distributed Bragg grating feedback,” Opt. Express 17, 395–404 (2009). [CrossRef]   [PubMed]  

16. D. Marcuse, Theory of Dielectric Optical Waveguides (Academic, 1974).

17. J. W. Goodman, Speckle Phenomena in Optics: Theory and Applications (Roberts, 2007).

18. P. W. Anderson, “Absence of diffusion in certain random lattices,” Phys. Rev. 109, 1492–1505 (1958). [CrossRef]  

19. D. Marcuse, “Rayleigh scattering and the impulse response of optical fiber,” Bell Syst. Tech. J. 53, 705–715 (1974).

20. B. Crosignani, A. Saar, and A. Yariv, “Coherent backscattering and localization in a single-mode fiber with random imperfections,” Phys. Rev. A 43, 3168–3171 (1991). [CrossRef]   [PubMed]  

21. H. C. van de Hulst, Light Scattering by Small Particles (Dover, 1981).

22. D. Marcuse, “Coupled power equations for lossy fibers,” Appl. Opt. 17, 3232–3237 (1978). [CrossRef]   [PubMed]  

23. J. C. Dainty, “Recent developments,” in Laser Speckle and Related Phenomena, J. C. Dainty, ed. (Springer, 1984).

24. Luc B. Jeunhomme, Single-Mode Fiber Optics (Marcel Dekker, 1990).

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 (5)

Fig. 1
Fig. 1 Experimental setup.
Fig. 2
Fig. 2 Examples of the output intensity distribution observed in some realizations with the disordered part of fiber (a)  1 cm and (b)  2 cm . The left column in each figure presents the pp-polarized distribution, and the right column presents the ps-polarized distribution. The angles of incidence are 0 ° , 2 ° , and 5 ° from the top to the bottom images.
Fig. 3
Fig. 3 Experimentally measured total co- (open symbols) and cross-polarized (solid symbols) transmission as a function of the length of the disordered part of the fiber for different polarizations of the transmitted beam. The circles correspond to an angle of incidence of 0 ° , the triangles to 2 ° , and the squares to 5 ° . Solid and dashed curves represent the theoretical fit with Eqs. (24, 25) with parameters α = 0.064 cm 1 , σ 2 = 0.1917 cm 1 .
Fig. 4
Fig. 4 The distributions, which correspond to an unconstrained random sum (shown as a dashed curve) and to a constrained random sum (shown with the solid curve) of all modes of the fiber, are compared to the experimentally observed distributions of the near-field intensity measured in co- (circles) and cross-polarized (squares) channels in a sample with L = 8 cm . The thin symbols correspond to an angle of incidence of 2 ° , and the bold symbols correspond to an angle of incidence of 5 ° .
Fig. 5
Fig. 5 Panel (a) plots the size of the speckle defined by Eqs. (6, 7, 8) with L x = 0.3 mm and L z = 2 mm as a function of the distance between the diffuser and the core of the photosensitive fiber. Panel (b) compares the values of the characteristic length ( x x ) 1 σ 3 after which all forward-propagating modes with one polarization become equally populated. It is found numerically from Eq. (16) without (solid curve) and with (circles) the delta function approximation to the order of magnitude estimate (squares) in Eq. (29). Panel (c) compares the amplitude of the radiative loss rate computed from Eq. (18) to the intermode coupling rate σ 3 . The plot shows that for the disorder patterns generated with D > 4 mm , the coupling becomes the dominant effect. This conclusion is borne out by the experimental results in Fig. 3.

Equations (29)

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

δ ε ( r ) δ ε ( r ) = δ ε 2 | A ( r ) A * ( r ) | 2 | A ( r ) | 2 | A ( r ) | 2 δ ε 2 | μ ( r , r ) | 2 ,
μ ( r , r ) μ ( r r ) μ ( x x , 0 , z z ) μ ( 0 , y y , 0 ) .
I ( x ˜ , z ˜ ) exp [ x ˜ 2 / L x 2 z ˜ 2 / L z 2 ] ,
| μ ( x x , 0 , z z ) | 2 = exp [ ( x x S x ) 2 ] × exp [ ( z z S z ) 2 ] ,
| μ ( 0 , y y , 0 ) | 2 = 1 ( 1 + [ π L x 2 λ UV D 2 ( y y ) ] 2 ) 1 / 2 ( 1 + [ π L z 2 λ UV D 2 ( y y ) ] 2 ) 1 / 2 1 ( 1 + [ y y S y 2 ] 2 ) 1 / 2 .
S x = λ UV D 2 π n core L x 0.15 λ UV D L x ,
S y = 3 λ UV D 2 π n core L z 2 0.38 λ UV D 2 L z 2 ,
S z = λ UV D 2 π n core L z 0.15 λ UV D L z ,
δ ε ( r ) δ ε ( r ) δ ε 2 exp [ ( x x S x ) 2 ] 1 [ 1 + ( y y S y 2 ) 2 ] 1 / 2 exp [ ( z z S z ) 2 ] .
E ( r ) ν c ν ( z ) e i ( ω t β ν z ) ( E t , ν ( x , y ) + e ^ z E z , ν ( x , y ) ) .
β ν [ E t , ν ( x , y ) · E t , ν ( x , y ) ] d x d y = δ ν ν ,
d c ν ( z ) d z = ν K ν ν ( z ) c ν ( z ) e i ( β ν β ν ) z ,
K ν ν ( z ) = ω 2 2 c 2 δ ε ( r ) [ E t , ν ( x , y ) · E t , ν ( x , y ) + E z , ν ( x , y ) E z , ν ( x , y ) ] d x d y
d P ν d z = c ν * d c ν d z + c.c. ,
d P ν d z = ν h ν ν ( P ν P ν ) ,
h ν ν = δ ε 2 ω 4 π log ( 2 ) S x S y S z c 4 e S z 2 | β ν β ν | 2 / 4 × [ E t , ν ( x , y ) · E t , ν ( x , y ) + E z , ν ( x , y ) E z , ν ( x , y ) ] 2 d x d y .
α h ν ν 2 3 π k 0 2 n core 2 A ,
α I = d Ω Δ k ( e ^ scat · e ^ z ) 2 d u x d u y d u z δ ε ( r ) δ ε ( r + u ) exp [ i Δ k · u ] .
d P ν d z = α P ν + ν h ν ν ( P ν P ν ) .
P ν ( z ) = P ν ( lossless ) ( z ) × exp [ α z ] ,
P ν ( lossless ) ( z ) = A ν exp [ σ z ] ,
det [ h ν ν δ ν ν τ h ν τ + σ ] = 0 ,
P ν ( z ) = e α z × [ n c n A ν ( n ) e σ n z ] , with c n = [ n A ν ( n ) P ν ( 0 ) ] .
P ( x ) ( z ) ν = 0 N / 2 1 P 2 ν + 1 e α z × 1 2 [ 1 + e σ 2 z ] ,
P ( y ) ( z ) ν = 1 N / 2 P 2 ν e α z × 1 2 [ 1 e σ 2 z ] .
P 2 ν + 1 ( z ) e α z × [ ( P 2 ν + 1 ( 0 ) 2 N ) e σ 3 z + 2 N ] , P 2 ν ( z ) 0 .
[ E t , ν ( x , y ) · E t , ν ( x , y ) + E z , ν ( x , y ) E z , ν ( x , y ) ] 2 d x d y ,
{ [ n core 2 ( ω 2 / c 2 ) π a 2 ] 1 x x , y y ( NA / 2 ) 4 [ n core 2 ( ω 2 / c 2 ) π a 2 ] 1 x y , y x .
( x x ) 1 σ 3 Δ n 2 n core π ω 2 S x S y S z c 2 a 2 = Δ n n core 4 λ UV 3 D 4 λ 2 a 2 L x L z 3 , ( x y ) 1 σ 2 mixing ( x x ) 1 ( NA 2 ) 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.