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

Modal decomposition for few mode fibers using the fractional Fourier system

Open Access Open Access

Abstract

Modal decomposition (MD) has become an indispensable diagnostic tool for optical fibers. A novel MD method using the fractional Fourier system is developed in this paper. Firstly, the existing phase retrieval (PR) algorithm based on the the fractional Fourier transform (FrFT) power spectra is extended to account for the effect of optical vortex. The extended capability can then be employed to reconstruct the phase of modal field to high fidelity in a non-iterative and non-interferometric manner. Combining the reconstructed phase with the measured near-field (NF) intensity, the modal field could be obtained and based upon which the complete MD (involving modal weights and phases) is performed. The validity and reliability of the method are demonstrated through several numerical examples including the noisy signals with different signal-to-noise ratio (SNR) levels.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Modal decomposition (MD) for characterizing the fiber laser beams is an emerging research field [1,2]. This technique is demonstrated to decompose the total field propagating in an optical fiber into sum of individual transverse modes and has become an indispensable diagnostic tool for the development of novel high power fiber laser systems [3], beam quality measurements [4], fiber to fiber coupling processing [5], wavefront reconstruction [6], and adaptive mode control [7], etc.

Recent years, a plethora of MD techniques have been developed in literature [8–18]. Some of the most popular ones are the spatially and spectrally (S2) resolved imaging technique [13], the correlation filter method (CFM) [14], and the numerical analysis method (NAM) [15–18], etc. Each of these MD techniques is based on unique procedures, using numerical or experimental approaches (or both) to access modal properties, and individually possesses outstanding merits. The S2 imaging technique is possible to extract the modal fields from the measurement data instead of being reliant upon theoretical calculated fields as they have been used for the CFM and NAM. So this technique is in particular suitable for characterizing novel or unknown fiber architectures where the actual mode set is unknown. The CFM highlights the promising ability of real-time decomposition of multi-mode beams and has been widely employed in various applications [1,2,4–6]. The NAM stands out by their remarkably low experimental effort [16].

However, the above mentioned techniques still have some limitations. The S2 imaging technique requires the most experimental effort (the broad-band source or tunable laser, optical spectrum analyzer, and two-axis mechanical shifter) as well as the highest measurement time consumption [19]. The resolution power of CFM for measuring the low higher-order mode contents is limited by the sensitivity of the CCD sensor at the measurement wavelength and the ambient noise. The center piece of NAM is the optimization algorithms, such as Gerchberg-Saxton (GS) algorithm [15], Stochastic Parallel Gradient Descent (SPGD) algorithm [17], hybrid genetic algorithm [18], and so on. These optimization algorithms easily suffer from initial value sensitivity and local minima problems.

In this paper, a novel MD method that makes use of the fractional Fourier system is presented. The role of the fractional Fourier system played in MD is to recover the phase of modal field. The most attractable advantages of PR algorithm based on the fractional Fourier system are non-iterative, non-interferometric, and experimentally simple. However, the existing PR algorithm doesn’t cover the effect of the optical vortex [20–22] while the higher-order fiber mode possesses optical vortices intrinsically [14]. Thus, we extend the existing PR algorithm to account for the effect of optical vortex by employing the hidden phase algorithm developed by Fried in adaptive optics area [23]. The extended PR algorithm allows the phase of the modal field to be reconstructed to high fidelity even when the optical vortices exist. Combing the reconstructed phase with the measured NF intensity, a preliminary modal field can be established. Based upon the obtained modal field, the power ratio and relative phase difference of each transverse mode could be determined unambiguously. The core idea behind the proposed method is very similar to the NAM, but the PR algorithm developed in this paper enables the phase to be reconstructed in a non-iterative manner. So the current approach possesses the merits of NAM: low experiment effort and high processing speed, but frees suffering from the challenge of it. The validity and reliability are demonstrated by several numerical examples including noisy signals with different SNR levels.

The outline of this article is as follows. Section 2 and 3 provide the fundamentals on the modal decomposition and PR algorithm using the fractional Fourier system, respectively. Thereafter, the efficacy of the proposed approach is presented in Section 4. Finally, some brief concluding remarks are given in Section 5.

2. Modal decomposition for step-index fiber (SIF)

Modal fields in weakly guiding SIF can be described to good approximation by linearly polarized (LP) modes [24]. The transverse structure of this mode set {ψi(x, y)} can be derived from the scalar Helmholtz equation. The ith LP mode structure can be factorized using

ψi(r,ϕ)=Rnm(r)Φn(ϕ)
where n and m are integers
Rnm(r)=c{Jn(Umr/r0)Jn(Um)(0r<r0)Kn(Wmr/r0)Kn(Wm)(rr0)andΦn(ϕ)={cos(nϕ)forevenmodessin(nϕ)foroddmodes
respectively. The symbol r0 denotes the fiber core radius and c is chosen so that ψi fulfills the normalization condition. Jn means the nth Bessel function of first kind, whereas Kn means the nth modified Bessel function of second kind. The value Um and Wm can be calculated from the characteristic equations
UmJn+1(Um)Jn(Um)=WmKn+1(Wm)Kn(Wm)andUm2+Wm2=V2
where
Um=2πaλnco2neff2andWm=2πaλneff2ncl2
with V=2πaλnco2ncl2 being the well-known V-parameter of the fiber. λ is the light wavelength. nco is the refractive index of the fiber core, while ncl is the refractive index of the fiber cladding and neff corresponds to an effective index.

An arbitrary fully coherent modal field that is normalized to unit power can be described as the coherent superposition of LP modes.

ψ(x,y)=|ψ(x,y)|exp[iϕ(x,y)]=n=1Ncnψn(x,y)
where ψ(x, y) denotes the modal field and ϕ(x, y) is the phase of it. Due to the orthogonal property of the LP modes, the modal coefficient cn can be given by
cn=ρnexp(iϕn)=ψn*(x,y)ψ(x,y)dxdy
where ρn2 is the modal weight, which fulfills the relation n=1N|cn|2=n=1Nρn2=1, ϕn is the inter-modal phase, and ψn* is the complex conjugate of the ψn. The integrals, unless specified, run from −∞ to +∞.

The difficulty along with Eq. (6) to evaluate modal coefficients stems from the fact that the modal field ψ(x, y) is generally immeasurable except its amplitude. Consequently the PR is indispensable. The PR algorithm employed in this paper can recover the phase from the FrFT power spectra measurement and the detail will be illustrated in the next section.

3. PR algorithm using the fractional Fourier system

Nowadays, The FrFT plays an important role in many branches of optics especially in optical information processing [25]. Its application in optics started with the seminal work of Mendlovic and Ozaktas [26]. There are many ways to define the FrFT, while in the present article the two dimensional FrFT is defined as [27]

ψαx,αy(u,v)=Kαx(x,u)Kαy(y,v)ψ(x,y)dxdy
where ψαx,αy is the FrFT of ψ(x, y) with two independent transformation parameters αq, αq ∈ [0, 2π], q is a placeholder for x and y, Kαq is the FrFT transformation kernel which is given by
Kαq(q,p)=exp{iπσ2sinαq[(q2+p2)cosαq2qp]}σisinaq
where p is a variable for u and v, σ is the scale parameter with the dimension of length. If αx = αy = 0, we have the identity transformation ψ0,0(x, y) = ψ(x, y). When αx=αy=π2 or αx=αy=π2, we have the ordinary Fourier transform or inverse Fourier transform of the modal field, respectively. The properties and numerical simulation method of the FrFT are detailed in [27]. We will not elaborate these topics in the present paper.

Using the phase space formalism, Bastiaans and Wolf derived the relationship between the phase gradient and angular derivative of the FrFT power spectra [20]

gx(x,y)=ϕ(x,y)x=πI0,0(x,y)Iαx,0(x,y)αx|αx=0sgn(xx)δ(yy)dxdy
gy(x,y)=ϕ(x,y)y=πI0,0(x,y)I0,αy(x,y)αy|αy=0sgn(yy)δ(xx)dxdy
where g(x, y) = [gx(x, y), gy(x, y)]T is the phase gradient field, T denotes transpose, sgn labels the sign function, δ is the Dirac delta function, Iαx,αy = |ψαx,αy (x, y)|2 is the FrFT power spectra, and I0,0(x, y) = |ψ(x, y)|2 is the NF intensity.

If two close fractional Fourier power spectra separated by small angle ε are measured, the phase gradient can be approximated to 2nd order of accuracy

gx(x,y)πI0,0(x,y)Iε,0(x,y)Iε,0(x,y)2εsgn(xx)δ(yy)dxdy
gy(x,y)πI0,0(x,y)I0,ε(x,y)I0,ε(x,y)2εsgn(yy)δ(xx)dxdy
In the practical situation, there are tradeoffs among the angle separation ε, the accuracy of the reconstruction result, and noise considerations. Small ε avoids the nonlinearity error, however, small ε also leads to noisy phase reconstructions. Utilizing the numerical technique to solve the Eqs. (11) and (12), the phase gradient could be obtained.

In [20–22], the authors employed some kind of numerical integral techniques to recover the phase from the phase gradient. While these kind of numerical integral techniques all depend on the integrand being free of singularities. In [14], Kaiser et al show that the higher-order transverse modes possess optical vortices intrinsically. An optical vortex (also known as branch point, phase singularity or screw dislocation) is one type of optical singularity that possesses a spiral wavefront pattern [29]. The spiral wavefront rotates around a point where the light intensity is zero and phase is undefined. According to the direction of rotation, the sign of the optical vortex could be positive or negative. Consequently, the above mentioned numerical integral techniques can not produce results that properly match the actual phase. The least-squares (LS) wavefront reconstruction technique is another standard way to recover the phase from the phase gradient field and has been widely utilized in adaptive optics systems [28]. Unfortunately, the LS techniques still does not work well in the presence of optical vortex [23].

We extend the existing PR algorithm to account for the effect of the optical vortex by employing the hidden phase theorem. This theorem is developed by Fried in adaptive optics area where the optical vortices is caused by the strong atmospheric turbulence [23]. According to the hidden phase theorem, if the modal field possesses optical vortices, the gradient field can be decomposed into two terms: the gradient of the scalar potential and the curl of the Hertz potential [23].

g(x,y)=s(x,y)+×H(x,y)
where ∇ is the gradient operator, ∇× defines the curl of the vector, s is the scalar potential, and H is the Hertz potential which is the contribution of the optical vortices and is along the z direction as [0, 0, h(x, y)]T. h(x, y) is known as the Hertz function.

Under this circumstance, the reconstructed phase ϕrc could be written as

ϕrc(x,y)=ϕls(x,y)+ϕhid(x,y)

The first term on the right-hand side of Eq. (14) is the LS phase. In this paper, we will make use of Zernike modal reconstructor to obtain this term. Recalling the fact that the wavefront can be expanded by the Zernike polynomials [28], the LS phase can be described as

ϕls(x,y)=p=1PapZp(x,y)
where ap is the expansion coefficient of the pth Zernike polynomial Zp.

According to Eqs. (13)(15), the gradient of the LS phase can be described as

ϕls(x,y)x=s(x,y)x=p=1PapZp(x,y)xϕls(x,y)y=s(x,y)y=p=1PapZp(x,y)y
It is possible to formulate a matrix version of Eq. (16) as
[s]=[Z][ap]
where [∂s] = [xs, ys]T, [∂Z] = [xZ, yZ]T, and [ap] is the expansion coefficient column. Because the Hertz potential has no contribution to LS phase [23], the [ap] can be estimated through a LS manner as
[ap]=[Z]+g
where [∂Z]+ = ([∂Z]T [∂Z])−1[∂Z]T is the pseudo-inverse of [∂Z] called Zernike modal reconstructor. Once the g is obtained, the coefficient column as well as the LS phase can also be determined straightforwardly.

The second term on the right-hand side of Eq. (14) is the hidden phase which is given by [23]

ϕhid(x,y)=Im{log[n=1N(xxn)+i(yyn)m=1M(xxm)+i(yym)]}
where the coordinate (xn, yn) represents the nth of N positive optical vortices location and (x′m, y′m) represents the mth of M negative optical vortices location. It is easy to find that the hidden phase is directly related to the locations and signs of the optical vortices. However, these two properties of the optical vortex are difficult to estimate in practical situation unless several complex algorithms are applied [30]. In this article, these two properties will be determined in a straightforward way.

The Hertz function h(x, y) that satisfies to Eq. (13) can be written as [23]

h(x,y)=log[m=1M|(xxm)2+(yym)2|n=1N|(xxn)2+(yyn)2|]
According to Eq. (20), it is evident that that the presence of optical vortices could be visualized as the peaks and valleys of the Hertz function where the peaks correspond to the positive optical vortices and the valleys correspond to the negative ones. Once the Hertz function is obtained, the locations and signs of the optical vortices will also be determined straightforwardly. We will show in the following that the Hertz function can be obtained from the phase gradient measurement.

The Eq. (13) can be expanded and expressed in matrix form as

[gx(x,y)gy(x,y)]=[xyyx][s(x,y)h(x,y)]
It is evident that the Fourier transform of Eq. (21) is
[g¯x(kx,ky)g¯y(kx,ky)]=[i2πkxi2πkyi2πkyi2πkx][s¯(kx,ky)h¯(kx,ky)]
where (kx, ky) = [x(kx, ky), y(kx, ky)]T is the Fourier transform of the gradient field g, kx and ky denote the spatial frequency, (kx, ky) and (kx, ky) are the Fourier transform of the scalar potential s(x, y) and Hertz function h(x, y), respectively. The Matrix inversion, together with the inverse Fourier transform, can yield
h(x,y)=dkxdkykyg¯x(kx,ky)kxg¯y(kx,ky)i2π(kx2+ky2)exp[i2π(kxx+kyy)]
Eq. (23) shows that the Hertz function can also be expressed in the Fourier transform terms of the gradient field g. As we will show in the next section, by using Eqs. (11) and (12), plenty of gradient data could be obtained based on which not only will the vortex structure of the gradient field be manifested clearly, but also could the Hertz function be obtained using Eq. (23). Thenceforth the locations of optical vortices will be recognized in a very convenient way.

Combing the reconstructed phase ϕrc with the NF intensity I0,0, a preliminary modal field can be established by ψrc(p)(x,y)=I0,0(x,y)exp(iϕrc(x,y)). Based upon Eq. (6), the modal weights ρn2 and modal phases ϕn could be obtained. The estimated modal weights may have a sub-normalized energy, i.e. nρn2<1. We compensated for this by renormalizing at the end of MD.

In order to filter out any information that cannot be represented as a superposition of the fiber modes, the modal field is reconstructed again from the obtained modal coefficients using Eq. (5). To quantify the accuracy of the final reconstructed modal field, the correlation coefficient between it and the original field ψ will be defined as:

C=|ψ(x,y)ψrc*(x,y)dxdyψ(x,y)ψ*(x,y)dxdyψrc(x,y)ψrc*(x,y)dxdy|
where ψrc denotes the final reconstructed modal field. It is noted that in the case of perfect agreement between ψrc and ψ, the correlation coefficient will be 1.

4. Numerical examples

In the following, we will demonstrate the validity and reliability of the MD method derived above through numerical simulations. As a way of example, an arbitrary linearly polarized modal field of the few mode step-index fiber (core diameter = 20μm, V-parameter = 4.7) operated at λ = 1μm will be considered and about 6 LP modes (LP01, LP02, LP11e, LP11o, LP21e, and LP21o) could be guided in it. Two different modal combinations are demonstrated in Table 1 and the numerical dimension of the modal field is 128 × 128. Last but not least, The first 70 Zernike polynomial terms are used to establish the Zernike modal reconstructor. Fig. 1(a) shows an example experimental setup that measures the FrFT power spectra and implements the MD. The FrFT system performs the transformation depending on the input fractional orders and the output FrFT power spectra are registered by a CCD camera which is connected to a PC. The customized software can output the modal contents based on the input FrFT power spectra. We accomplish this experimental setup using the numerical simulation method and the flowchart of the numerical simulation code is illustrated in Fig. 1(b). It should be noted that we do not demonstrate the detailed 2D FrFT optical setup in Fig. 1(a) because that isn’t the emphasis of the present work. But the reader can consult [31] where a flexible optical system able to perform the 2D FrFT is given in.

Tables Icon

Table 1. Two Different Modal Combinations.

 figure: Fig. 1

Fig. 1 (a) schematic for the modal decomposition using the fractional Fourier systems; (b) flowchart of the numerical simulation code.

Download Full Size | PDF

4.1. Example1

In the first numerical example, the input modal field is generated following the modal combination 1 and no noise effect is considered. Choosing ε = 0.3 rad and measuring five FrFT power spectra: I0,0, Iε,0, Iε,0, I0,ε, and I0,−ε, the two dimensional x, y component of the phase gradient g can be obtained using Eqs. (11) and (12). Figs. 2(a) and 2(b) depict these two components respectively. Due to the high density of gradient data, only a zoomed portion of whole phase gradient vector is illustrated in Fig. 2(c).

 figure: Fig. 2

Fig. 2 The two dimensional phase gradient and phase obtained using the FrFT power spectra. The input modal field is generated according to the modal combination 1 and no noise effect is considered. (a), (b) the x and y component of the phase gradient g, respectively; (c) a zoomed portion of the phase gradient vector; (d) the reconstructed phase obtained by Zernike modal reconstructor.

Download Full Size | PDF

In this example, the modal field does not possess any optical vortex. This conclusion is evident from the sketch in Fig. 2(c) that the phase gradient vector shows no spiral pattern. The reason may be that the fundamental mode dominates the modal combination. Under this circumstance, the Zernike modal reconstructor can furnish the required phase. Fig. 2(d) illustrates the reconstructed phase. As the Zernike polynomials are orthogonal over the unit disk, only the phase over the whole unit disk can be obtained.

Combining the reconstructed phase with the measured NF intensity I0,0, the preliminary modal field is obtained. And then the complete MD is performed. In order to make sure that the energy conservation is compromised, the modal weights are renormalized after the MD. The revised modal weights and phases are compared with the actual results in Figs. 3(a) and 3(b) respectively. We can see that the agreement is very high.

 figure: Fig. 3

Fig. 3 The comparison between the MD results and the actual results. (a) modal weights; (b) modal phases.

Download Full Size | PDF

As described in section 3, the modal field is reconstructed again from the obtained modal weights and phases using Eq. (5). This procedure is very necessary since it can filter out any information that can not be represented as the superposition of the transverse modes. Fig. 4 depicts the comparison between the reconstructed NF intensity and phase with the original ones. We can see that the reconstructed NF intensity and phase of modal field are visually almost the same as the original ones. The correlation coefficient between the reconstructed modal field and the original one is 0.9998. The relative root-mean-square (RMS) error of the reconstructed NF intensity and phase are 8.65 × 10−3, 4.27 × 10−2, respectively.

 figure: Fig. 4

Fig. 4 The comparison between the reconstructed results and the original ones. (a) the original NF intensity; (b) the reconstructed NF intensity; (c) the discrepancy between (a) and (b); (d) the original phase of modal field; (e) the reconstructed phase; (f) the discrepancy between (d) and (e).

Download Full Size | PDF

4.2. Example2

In the second numerical example, the input modal field is generated following the modal combination 2 and also no noise effect is considered. In this combination, the power ratio of the fundamental mode is decreased while the corresponding higher-order mode contents are increased. The values of ε are the same as example 1. The two dimensional x, y component of the phase gradient g are demonstrated in Figs. 5(a) and 5(b), respectively. Because of the same reason as example 1, only a zoomed portion of the gradient vector is shown in Fig. 5(c) where two spiral patterns are clearly evident, which means that the modal field possesses two optical vortices. The counter-clockwise one marked by plus corresponds to the positive optical vortex and the clockwise one marked by minus corresponds to the negative optical vortex.

 figure: Fig. 5

Fig. 5 The two dimensional phase gradient obtained using the FrFT power spectra. The input modal field is generated according to the modal combination 2 and no noise effect is considered. (a), (b) the x and y component of the phase gradient g, respectively; (c) a zoomed portion of the phase gradient vector. Evidently, it has two spiral patterns where plus marks the positive one and minus marks the negative one.

Download Full Size | PDF

The Hertz function obtained from Eq. (23) is given in Fig. 6(a). It is apparent that the positive and negative optical vortices are visualized as peaks and valleys of the Hertz function which permits the locations and signs of optical vortex to be determined in a very convenient way. Figs. 6(b) and 6(c) provide the LS phase and hidden phase respectively. The LS phase is reconstructed by the Zernike modal reconstructor too and the hidden phase is obtained by Eq. (19) where a 2π phase discontinuity is quite obvious, which cannot be reconstructed by the Zernike modal reconstructor.

 figure: Fig. 6

Fig. 6 (a) the Hertz function which obviously demonstrate a peak and a valley indicating the positive and negative optical vortex, respectively; (b) the LS phase reconstructed by Zernike modal reconstructor; (c) the hidden phase obtained by Eq. (19).

Download Full Size | PDF

Combining the LS phase and the hidden phase with the measured NF intensity I0,0, the preliminary modal field is obtained. The further process is the same as example 1. The complete MD is performed. The obtained modal weights and phases are compared with the actual results in Figs. 7(a) and 7(b) respectively. The agreement is also quite high.

 figure: Fig. 7

Fig. 7 The comparison between the MD results and actual results. (a) modal weights; (b) modal phases.

Download Full Size | PDF

The modal field is reconstructed again from the obtained modal weights and phases. Fig. 8 demonstrates the comparison between the reconstructed NF intensity and phase with the original ones. We can see again that the reconstructed NF intensity and phase are visually almost the same as the original ones. The correlation coefficient (CC), the relative RMS error of the reconstructed NF intensity (RMS_I) and phase (RMS_P) are shown in Table 2. In comparison, The results where the hidden phase algorithm is not employed also illustrated in it. From Table 2, we can conclude that the extended PR algorithm outperforms the existing one in our numerical example.

 figure: Fig. 8

Fig. 8 The comparison between the reconstructed results and the original ones. (a) the original NF intensity; (b) the reconstructed NF intensity; (c) the discrepancy between (a) and (b); (d) the original phase of modal field; (e) the reconstructed phase; (f) the discrepancy between (d) and (e).

Download Full Size | PDF

Tables Icon

Table 2. The Comparison of the Correlation Coefficient and the Relative RMS Errors between the Extended PR Algorithm and the Existing One.

4.3. Example3

In this example, the proposed MD method is tested under several different SNR levels. When the FrFT power spectra are recorded by CCD-based optical imaging systems, the noise is generally inevitable. In the present paper, we mainly incorporate the photon noise and readout noise (RON) into the FrFT power spectra. The photon noise can be well-approximated by the Poisson random numbers, while the RON can be well-approximated by the zero-mean Gaussian random numbers with standard deviation (STD) σr. The input modal field is generated following the modal combination 1. The values ε are the same as in example 1. But the five FrFT power spectra: I0,0, Iε,0, Iε,0, I0,ε, and I0,−ε, are contaminated by RON with σr = 3e and photon noise. For each SNR level, the numerical simulation is repeated 100 times, thereafter several important statistical data for evaluation the method, such as: the average correlation coefficient (ACC), the average relative RMS error of reconstructed NF intensity (ARMS_I), and their STDs (STD_C and STD_I, respectively) are recorded. All these simulation results are given in Table 3.

Tables Icon

Table 3. Summary of the Numerical Simulation Results under Different SNR Level.

From the results given in Table 3, we can conclude that the proposed MD method can work quite well for the SNR higher than 20dB and the uncertainty of the results are also very small. Below this value, the degradation of the reconstructed result is significant. In the future, we will explore the more effective noise suppression algorithm to improve the ability of the proposed MD technique in the presence of noise.

5. Conclusion

To conclude, we have introduced a novel MD technique which can process the modal content of few mode fibers very conveniently and quickly. The core of this method is the extended PR algorithm using the fractional Fourier system. The numerical simulation results show that the discussed MD method produces quite good result no matter whether the optical vortex exists or not. The noise sensitivity for the proposed technique is also analyzed. It is shown that the present MD method can still work well when the SNR level is as low as 20dB in our numerical example. The main advantages of the proposed method are its non-iterative, non-interferometric manner, and low experimental effort. This technique can be employed in multiple areas to improve the MD ability.

References

1. A. Forbes, A. Dudley, and M. McLaren, “Creation and detection of optical modes with spatial light modulators,” Adv. Opt. Photonics 8, 200–227 (2016). [CrossRef]  

2. D. Flamm, C. Schulze, D. Naidoo, S. Schröter, A. Forbes, and M. Duparré, “All-digital holographic tool for mode excitation and analysis in optical fibers,” J. Light. Technol. 31, 1023–1032 (2013). [CrossRef]  

3. F. Stutzki, F. Jansen, T. Eidam, A. Steinmetz, C. Jauregui, J. Limpert, and A. Tünnermann, “High average power large-pitch fiber amplifier with robust single-mode operation,” Opt. Lett. 36, 689–691 (2011). [CrossRef]   [PubMed]  

4. D. Flamm, C. Schulze, R. Brüning, O. Schmidt, A. Forbes, and M. Duparré, “Fast M2 measurement for fiber beams based on modal analysis,” Appl. Opt. 51, 987–993 (2012). [CrossRef]   [PubMed]  

5. D. Flamm, K. C. Hou, P. Gelszinnis, C. Schulze, S. Schröter, and M. Duparré, “Modal characterization of fiber-to-fiber coupling processes,” Opt. Lett. 38, 2128–2130 (2013). [CrossRef]   [PubMed]  

6. C. Schulze, D. Naidoo, D. Flamm, O. A. Schmidt, A. Forbes, and M. Duparré, “Wavefront reconstruction by modal decomposition,” Opt. Express 20, 19714–19725 (2012). [CrossRef]   [PubMed]  

7. L. Huang, J. Leng, P. Zhou, S. Guo, H. Lü, and X. Cheng, “Adapitve mode control of a few mode fiber by real-time mode decomposition,” Opt. Express 23, 28082–28090 (2015). [CrossRef]   [PubMed]  

8. D. Soh, J. Nilsson, L. Sun, S. Baek, C. Codemard, Y. Jeong, and V. Philippov, “Modal power decomposition of beam intensity profiles into linearly polarized modes of multimode optical fibers,” J. Opt. Soc. Am. A 21, 1241–1250 (2004). [CrossRef]  

9. H. Zhou, Q. Zhu, W. Liang, G. Zhu, Y. Xue, S. Chen, L. Shen, M. Liu, J. Dong, and X. Zhang, “Mode measurement of few-mode fibers by mode-frequency mapping,” Opt. Lett. 43, 1435–1438 (2018). [CrossRef]   [PubMed]  

10. F. Fatemi and G. Beadie, “Rapid complex mode decomposition of vector beams by common path interferometry,” Opt. Express 21, 32291–32305 (2013). [CrossRef]  

11. M. Lyu, Z. Lin, G. Li, and G. Situ, “Fast modal decomposition for optical fibers using digital holography,” Sci. Reports 7, 1–9 (2017).

12. M. Paurisse, L. Lévèque, M. Hanna, F. Druon, and P. Georges, “Complete measurement of fiber modal content by wavefront analysis,” Opt. Express 20, 4074–4084 (2012). [CrossRef]   [PubMed]  

13. J. W. Nicholson, A. D. Yablon, and S. Ghalmi, “Spatially and Spectrally resolved imaging of modal content in large-mode-area fibers,” Opt. Express 16, 7233–7243 (2008). [CrossRef]   [PubMed]  

14. T. Kaiser, D. Flamm, S. Schröter, and M. Duparré, “Complete modal decomposition for optical fibers using CGH-based correlation filters,” Opt. Express 17, 9347–9356 (2009). [CrossRef]   [PubMed]  

15. O. Shapira, A. F. Abouraddy, J. D. Joannopoulos, and Y. Fink, “Complete modal decomposition for optical waveguides,” Phys. Rev. Lett. 94, 143902 (2005). [CrossRef]   [PubMed]  

16. R. Brüning, P. Gelszinnis, C. Schulzes, D. Flamm, and M. Duparré, “Comparative analysis of numerical methods for the mode analysis of laser beams,” Appl. Opt. 52, 7769–7777 (2013). [CrossRef]   [PubMed]  

17. L. Huang, J. Leng, P. Zhou, S. Guo, H. Lü, and X. Cheng, “Real-time mode decomposition for few-mode fiber based on numerical method,” Opt. Express 23, 4620–4629 (2015). [CrossRef]   [PubMed]  

18. L. Li, J. Leng, P. Zhou, and J. Chen, “Multimode fiber modal decomposition based on hybrid genetic global optimization,” Opt. Express 25, 19680–19690 (2015). [CrossRef]  

19. J. W. Nicholson, A. D. Yablon, J. M. Fini, and M. D. Mermelstein, “Measuring the modal content of large-mode-area fibers,” IEEE J. Sel. Top. Quantum Electron. 15, 61–70 (2009). [CrossRef]  

20. M. J. Bastiaans and K. B. Wolf, “Phase reconstruction from intensity measurements in linear systems,” J. Opt. Soc. Am. A 20, 2046–2049 (2003). [CrossRef]  

21. T. Alieva, M. J. Bastiaans, and L. Stankovic, “Signal reconstruction from two close fractional Fourier power spectra,” IEEE Trans. Signal Process. 51, 112–123 (2003). [CrossRef]  

22. U. Gopinathan, G. Situ, T. J. Naughton, and J. T. Sheridan, “Noninterferometric phase retrieval using a fractional Fourier system,” J. Opt. Soc. Am. A 25, 108–115 (2008). [CrossRef]  

23. D. L. Fried, “Branch point problem in adaptive optics,” J. Opt. Soc. Am. A 15, 2759–2768 (1998). [CrossRef]  

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

25. T. Alieva, M. J. Bastiaans, and M. L. Calvo, “Fractional Fourier transforms in optical information processing,” EURASIP J. on Appl. Signal Process. 10, 1–22 (2005).

26. D. Mendlovic and H. M. Ozaktas, “Fractional Fourier transforms and their optical implementation,” J. Opt. Soc. Am. A 10, 1875–1881 (1993). [CrossRef]  

27. H. M. Ozaktas, Z. Zalevsky, and M. A. Kutay, The Fractional Fourier Transform with Application in Optics and Signal Processing (John Wiley & Sons, 2001).

28. R. K. Tyson, Principles of Adaptive Optics, 2nd ed. (Academic Press, 1998).

29. M. J. Padgett, “Orbital angular momentum 25 years on,” Opt. Express 25, 11265–11274 (2017). [CrossRef]   [PubMed]  

30. J. Luo, H. Huang, Y. Matsui, H. Toyoda, T. Inoue, and J. Bai, “High-order optical vortex position detection using a Shack-Hartmann wavefront sensor,” Opt. Express 23, 8706–8719 (2015). [CrossRef]   [PubMed]  

31. J. A. Rodrigo, T. Alieva, and M. Calvo, “Programmable two-dimensional optical fractional Fourier processor,” Opt. Express 17, 4976–4983 (2009). [CrossRef]   [PubMed]  

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 (8)

Fig. 1
Fig. 1 (a) schematic for the modal decomposition using the fractional Fourier systems; (b) flowchart of the numerical simulation code.
Fig. 2
Fig. 2 The two dimensional phase gradient and phase obtained using the FrFT power spectra. The input modal field is generated according to the modal combination 1 and no noise effect is considered. (a), (b) the x and y component of the phase gradient g, respectively; (c) a zoomed portion of the phase gradient vector; (d) the reconstructed phase obtained by Zernike modal reconstructor.
Fig. 3
Fig. 3 The comparison between the MD results and the actual results. (a) modal weights; (b) modal phases.
Fig. 4
Fig. 4 The comparison between the reconstructed results and the original ones. (a) the original NF intensity; (b) the reconstructed NF intensity; (c) the discrepancy between (a) and (b); (d) the original phase of modal field; (e) the reconstructed phase; (f) the discrepancy between (d) and (e).
Fig. 5
Fig. 5 The two dimensional phase gradient obtained using the FrFT power spectra. The input modal field is generated according to the modal combination 2 and no noise effect is considered. (a), (b) the x and y component of the phase gradient g, respectively; (c) a zoomed portion of the phase gradient vector. Evidently, it has two spiral patterns where plus marks the positive one and minus marks the negative one.
Fig. 6
Fig. 6 (a) the Hertz function which obviously demonstrate a peak and a valley indicating the positive and negative optical vortex, respectively; (b) the LS phase reconstructed by Zernike modal reconstructor; (c) the hidden phase obtained by Eq. (19).
Fig. 7
Fig. 7 The comparison between the MD results and actual results. (a) modal weights; (b) modal phases.
Fig. 8
Fig. 8 The comparison between the reconstructed results and the original ones. (a) the original NF intensity; (b) the reconstructed NF intensity; (c) the discrepancy between (a) and (b); (d) the original phase of modal field; (e) the reconstructed phase; (f) the discrepancy between (d) and (e).

Tables (3)

Tables Icon

Table 1 Two Different Modal Combinations.

Tables Icon

Table 2 The Comparison of the Correlation Coefficient and the Relative RMS Errors between the Extended PR Algorithm and the Existing One.

Tables Icon

Table 3 Summary of the Numerical Simulation Results under Different SNR Level.

Equations (24)

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

ψ i ( r , ϕ ) = R n m ( r ) Φ n ( ϕ )
R n m ( r ) = c { J n ( U m r / r 0 ) J n ( U m ) ( 0 r < r 0 ) K n ( W m r / r 0 ) K n ( W m ) ( r r 0 ) and Φ n ( ϕ ) = { cos ( n ϕ ) for even modes sin ( n ϕ ) for odd modes
U m J n + 1 ( U m ) J n ( U m ) = W m K n + 1 ( W m ) K n ( W m ) and U m 2 + W m 2 = V 2
U m = 2 π a λ n co 2 n eff 2 and W m = 2 π a λ n eff 2 n cl 2
ψ ( x , y ) = | ψ ( x , y ) | exp [ i ϕ ( x , y ) ] = n = 1 N c n ψ n ( x , y )
c n = ρ n exp ( i ϕ n ) = ψ n * ( x , y ) ψ ( x , y ) d x d y
ψ α x , α y ( u , v ) = K α x ( x , u ) K α y ( y , v ) ψ ( x , y ) d x d y
K α q ( q , p ) = exp { i π σ 2 sin α q [ ( q 2 + p 2 ) cos α q 2 q p ] } σ i sin a q
g x ( x , y ) = ϕ ( x , y ) x = π I 0 , 0 ( x , y ) I α x , 0 ( x , y ) α x | α x = 0 sgn ( x x ) δ ( y y ) d x d y
g y ( x , y ) = ϕ ( x , y ) y = π I 0 , 0 ( x , y ) I 0 , α y ( x , y ) α y | α y = 0 sgn ( y y ) δ ( x x ) d x d y
g x ( x , y ) π I 0 , 0 ( x , y ) I ε , 0 ( x , y ) I ε , 0 ( x , y ) 2 ε sgn ( x x ) δ ( y y ) d x d y
g y ( x , y ) π I 0 , 0 ( x , y ) I 0 , ε ( x , y ) I 0 , ε ( x , y ) 2 ε sgn ( y y ) δ ( x x ) d x d y
g ( x , y ) = s ( x , y ) + × H ( x , y )
ϕ rc ( x , y ) = ϕ ls ( x , y ) + ϕ hid ( x , y )
ϕ ls ( x , y ) = p = 1 P a p Z p ( x , y )
ϕ ls ( x , y ) x = s ( x , y ) x = p = 1 P a p Z p ( x , y ) x ϕ ls ( x , y ) y = s ( x , y ) y = p = 1 P a p Z p ( x , y ) y
[ s ] = [ Z ] [ a p ]
[ a p ] = [ Z ] + g
ϕ hid ( x , y ) = Im { log [ n = 1 N ( x x n ) + i ( y y n ) m = 1 M ( x x m ) + i ( y y m ) ] }
h ( x , y ) = log [ m = 1 M | ( x x m ) 2 + ( y y m ) 2 | n = 1 N | ( x x n ) 2 + ( y y n ) 2 | ]
[ g x ( x , y ) g y ( x , y ) ] = [ x y y x ] [ s ( x , y ) h ( x , y ) ]
[ g ¯ x ( k x , k y ) g ¯ y ( k x , k y ) ] = [ i 2 π k x i 2 π k y i 2 π k y i 2 π k x ] [ s ¯ ( k x , k y ) h ¯ ( k x , k y ) ]
h ( x , y ) = d k x d k y k y g ¯ x ( k x , k y ) k x g ¯ y ( k x , k y ) i 2 π ( k x 2 + k y 2 ) exp [ i 2 π ( k x x + k y y ) ]
C = | ψ ( x , y ) ψ rc * ( x , y ) d x d y ψ ( x , y ) ψ * ( x , y ) d x d y ψ rc ( x , y ) ψ rc * ( x , y ) d x d y |
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.