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

Self-focusing propagation characteristics of a radially-polarized beam in nonlinear media

Open Access Open Access

Abstract

In this study, an analytical formula for the self-focusing length of a radially polarized beam (RPB) is first derived, which has a similar behavior to the semi-empirical Marburger formula of a Gaussian beam, and is beneficial to quantitatively and qualitatively analyze practical experimental scenarios. However, the relation of the self-focusing length with the states of polarization (SoPs) was evaluated, and it was found that RPB with spatially inhomogeneous SoP at the field cross-section can retain a further self-focusing length compared to a beam with a spatially homogeneous one. The influence of the topological charge on the self-focusing length is explored, which shows that RPB with a low topological charge can achieve a high-power density at a relatively further receiver plane. Therefore, it is demonstrated that the RPB as a laser source not only extends the self-focusing length, but also improves the power density of the target. With the help of RPB, it is possible to realize a controllable self-focusing length and a high target optical power density, which may have potential applications in fine optical manipulation, optical communication, high-power long-range laser atmospheric propagation, and related areas.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

In recent decades, filamentation propagation in transparent media [1], which can generate long plasma channels with very high intensities, has sparked extensive interest owing to its wide application in many fields [25], such as lighting control, remote diagnostics, and LIDAR. The self-focusing length is described as the self-focusing effect that overcomes diffraction to a collapse of the beam on itself [1], which may play a crucial role in the steerable manipulation of the filament length [610]. To date, the known methods for controlling the position of the filamentation domain are as follows: modulating the laser pulse power [6], adjusting the divergence angle of the initial laser [7], launching negatively chirped ultrashort pulses [8,9], and double-lens setup [10]. Recently, the U.S. Naval Research Laboratory evaluated the self-focusing length in atmospheric turbulence using theoretical and numerical methods [11,12].

Since 1961, the vector optical field [13] with a spatially inhomogeneous SoP at the field cross-section has attracted considerable attention, which can be widely used in nonlinear optics [14], optical tweezers [15], optical micromanipulation [16], super-resolution microscopy [17], optical information transmission [18], and atmospheric propagation [1924]. In recent years, optical arrangements containing spatial light modulators have been widely presented [2537]. As a typical type of cylindrical vector beam, the radially polarized beam (RPB) has been studied extensively both theoretically and experimentally owing to its interesting and unique focusing properties [3841]. Although these studies have been extensive and seem comprehensive, nonlinear effects can significantly affect the essence of RPB propagation; particularly, the Kerr effect strongly exists when an intense laser beam is present in nonlinear media.

It is well known that the self-focusing length of a Gaussian beam can be described by the semi-empirical Marburger formula [4243]. However, there is currently no analytical formula for the majority of laser beams (i.e., RPB or Gaussian) to quantitatively characterize the relationship between the self-focusing length and input power, which is worthy of further exploration. The self-focusing length of an RPB may spark extensive interest owing to its wide application in many fields, such as laser filamentation [10], high-power atmospheric propagation [10], optical communications [44], optical coherence encryption [45], and far-field imaging [46]. Thus, the evaluation of the self-focusing length of an RPB has potential application prospects. If a laser beam with a spatially inhomogeneous SoP can be used to control the self-focusing length, it may provide an alternative route for manipulating the filamentation domain.

In this study, the analytical formula of the self-focusing length of the RPB was derived, and the correctness was verified by numerical calculation. It was shown that the self-focusing length is inversely proportional to the input power, which is similar to the performance of the semi-empirical formula of the Gaussian beam. Furthermore, the relationship between the SoP at the field cross-section and the self-focusing length was studied. It was found that the RPB has a relatively long self-focusing length compared to a beam with a spatially homogenous SoP. Additionally, the influence of the topological charge on the self-focusing length was evaluated. Based on numerical calculations, it was found that a low topological charge can not only achieve a long self-focusing length but also obtain a high-power density at the target. With the aid of the RPB, a controllable self-focusing length and high optical power density on the target can be achieved, which may have practical application significance in the fine optical manipulation, optical information transmission, high-power long-range atmospheric propagation, and related areas.

2. Theory

We discuss the monochromatic (or continuous) regime, for which the intensity of the laser does not depend on time. The propagation nonlinear dynamics of a high-power laser beam in the Kerr medium were evaluated using nonlinear Schrödinger (NLS) equations. Under the slowly varying amplitude approximation, the NLS equation is [47,48]:

$$2\textrm{i}k\frac{{\partial {\mathbf E}}}{{\partial z}} + {\nabla ^2}{\mathbf E} + \frac{{4{n_2}{k^2}}}{{3{n_0}}}[{2({{\mathbf E}\cdot {{\mathbf E}^ \ast }} ){\mathbf E} + ({{\mathbf E}\cdot {\mathbf E}} ){{\mathbf E}^ \ast }} ]= 0,$$
where ${\mathbf E} = {\mathbf E}({\mathbf r},z)$ is the amplitude of the electric field, the wavenumber related to the wavelength is $k = {{2\pi } / \lambda }$, ${\nabla ^\textrm{2}}$ is the transverse Laplacian, ${n_\textrm{0}}$ (${n_2}$) is the linear (nonlinear) refractive index, $\left\langle \bullet \right\rangle$ denotes the statistical ensemble average, and ${\ast} $ is the complex conjugate.

The SoP distribution of the radially polarized vector field at the initial cross-section is linearly polarized at any position, and the electric field can be expressed as [25,34,36]:

$${\mathbf E}(r,\varphi ,z = 0) = A(r)({\cos \delta {{\vec{{\mathbf e}}}_h} + \sin \delta {{\vec{{\mathbf e}}}_v}} ),$$
where $A({\mathbf r}) = {A_0}({{\mathbf r}/{w_0}} )\exp [{ - {{\mathbf r}^2}/2w_0^2} ]$, ${A_0}$ is the normalized constant, ${w_0}$ is the beam waist associated with the Gaussian beam, $\delta = m\varphi + {\varphi _0}$, $m$ is the topological charge, $\varphi$ is the azimuthal angle, and ${\varphi _0}$ is the initial phase.

With the introduction of polarization, Eq. (1) can be divided into a pair of coupled NLS equations for the two orthogonal components.

$$\begin{array}{l} \textrm{2i}k\frac{{\partial {E_h}}}{{\partial z}} + {\nabla ^2}{E_h} + \frac{{4{n_2}{k^2}}}{{3{n_0}}}[{3{{|{{E_h}} |}^2}{E_h} + 2{{|{{E_v}} |}^2}{E_h} + E_v^2E_h^ \ast } ]= 0\\ \textrm{2i}k\frac{{\partial {E_v}}}{{\partial z}} + {\nabla ^2}{E_v} + \frac{{4{n_2}{k^2}}}{{3{n_0}}}[{3{{|{{E_v}} |}^2}{E_v} + 2{{|{{E_h}} |}^2}{E_v} + E_h^2E_v^ \ast } ]= 0 \end{array}$$
where ${E_h}$ (${E_v}$) is the horizontal or vertical component of an optical field.

Owing to the hollow shape in the initial source plane, the RPB can retain its original shape for long-range propagation, and the electric field can be written as:

$${\mathbf E}(r,z) = {A_z}\left( {\frac{r}{{{w_z}}}} \right)\exp \left( { - \frac{{{r^2}}}{{w_z^2}} + \textrm{i}{r^2}{\varphi_z}} \right)({\cos \delta {{\vec{{\mathbf e}}}_h} + \sin \delta {{\vec{{\mathbf e}}}_v}} )$$

By inserting the initial conditions (beam width ${w_0}$, phase ${\varphi _0} = 0$, and amplitude ${A_0} = 1$) and Eq. (4) into Eq. (3), a set of equations for these quantities is obtained as follows:

$$\frac{{\textrm{d}{w_z}}}{{\textrm{d}z}} = \frac{2}{k}{\varphi _z}{A_z},$$
$$\frac{{\textrm{d}{\varphi _z}}}{{\textrm{d}z}} = \frac{1}{{2kw_z^4}} - \frac{{2\varphi _z^2}}{k} + \frac{{2{n_2}kA_z^2}}{{{n_0}w_z^2}},$$
$$\frac{{\textrm{d}{A_z}}}{{\textrm{d}z}} ={-} \frac{2}{k}{\varphi _z}{A_z}.$$

Combining Eqs. (5) and (6), the dynamics of the beam width of a RPB are

$$\frac{{{\textrm{d}^2}{w_z}}}{{\textrm{d}{z^2}}} = \frac{1}{{{k^2}w_z^3}} + \frac{{4{n_2}A_z^2}}{{{n_0}{w_z}}},$$
then, using ${ {({{{\textrm{d}{w_z}} / {\textrm{d}z}}} )} |_{z = 0}} = 0$, Eq. (8) can also be expressed as,
$${\left( {\frac{{\textrm{d}{w_z}}}{{\textrm{d}z}}} \right)^2} + \left( {\frac{4}{{w_z^2}} - \frac{4}{{w_0^2}}} \right) + \frac{{16{n_2}kA_z^2}}{{{n_0}}}\ln \left( {\frac{{{w_z}}}{{{w_0}}}} \right) = 0.$$

With boundary conditions ${w_{z = 0}} = {w_0}$ and ${ {({{{\textrm{d}{w_z}} / {\textrm{d}z}}} )} |_{z = 0}} = 0$, the analytical expression for the beam width is obtained,

$$w_z^2 = w_0^2 + \frac{{{z^2}}}{{{k^2}w_0^2}} + \frac{{4{n_2}A_z^2{z^2}}}{{{n_0}}}\left( {1 - \ln \frac{{{A_z}}}{{{A_0}}}} \right).$$

In physical terms, the evolution of the beam width is determined by the competition between two factors: the diffraction and nonlinearity of the medium. Here, the focusing case with ${n_2} > 0$ is considered, where ${{{A_z}} / {{A_0}}}$ represents the normalized amplitude.

Substituting the critical power of the Gaussian beam ${P_{cr}} = {{{\varepsilon _0}n_0^2\pi c} / {{n_2}{k^2}}}$ and the input power of the RPB ${P_{in}} = {\varepsilon _0}{n_0}\pi c{I_0}w_0^2$ into Eq. (10), the self-focusing length can be expressed as follows:

$${z_f} = \frac{{{z_R}}}{{\sqrt {[{ - 1 + \ln ({{{{A_z}} / {{A_0}}}} )} ]{{{P_{\textrm{in}}}} / {{P_{\textrm{cr}}}}} - 0.25} }}.$$
where the critical power of the RPB should be satisfied [31]: $P_{cr}^R = 2.4({m^2} + 1){P_{cr}}$. The self-focusing length corresponds to the input power and topological charge.

3. Numerical calculation and analysis

3.1 Comparison of analytical and numerical results

To verify the accuracy of the analytical formula for the self-focusing length of the RPB, a comparison of the analytical formula in Eq. (11), and the numerical result should be obtained firstly. In the numerical calculation, the parameters were chosen as follows: wavelength $\lambda = 0.532$ µm, step number $M = \textrm{2}000$, transverse size $8{w_0}$, grid number $N = 512$, Rayleigh length is ${z_R} = \pi w_0^2/\lambda$, linear refractive index of Kerr medium ${n_0} = 1$, nonlinear refractive index ${n_2} = \textrm{5} \times {10^{ - 23}}$ m2/W, initial phase ${\varphi _0}\textrm{ = 0}$, and the range of input power for PRB is $[27{P_{cr}},50{P_{cr}}]$ (or $[68{P_{cr}},91{P_{cr}}]$) with the topological charge $m = 1$ (or $m = 2$). The numerical results were obtained by the fast Fourier transform Runge-Kutta split-step method [49]. In Fig. 1, a comparison of the numerical calculations shows that the analytical and numerical results are in good agreement. This will facilitate the quantitative and qualitative analysis, which may have a significant theoretical basis in optical field collapse, filament propagation, and related areas. It is well known that the self-focusing length for a Gaussian beam with moderate input powers can be described by the semi-empirical Marburger formula [4243], $z = {{0.367{z_R}} / {\sqrt {{{[\sqrt {P/{P_{cr}}} - 0.852]}^2} - 0.0219} }}$, and it is obvious that the self-focusing length for the RPB exhibits a similar phenomenon, that is, the self-focusing length is inversely proportional to the input power. The analytical relation between the self-focusing length and input power is beneficial for the selection of the initial parameters for practical applications, and it is convenient to further analyze the physical characteristics in detail.

 figure: Fig. 1.

Fig. 1. ${{{z_f}} / {{z_R}}}$ as a function of the relative input laser power ${P / {{P_{cr}}}}$ for RPB.

Download Full Size | PDF

3.2 Self-focusing length with SOP

The self-focusing length can be adjusted by the input power, which is a common method for manipulating the position of the filamentation domain [6]. In practical experimental scenarios, as an alternative or combination regimen to input power manipulation, achieving a controllable self-focusing length remains a great challenge, which may significantly broaden the manipulation dimensionality of the self-focusing length. Owing to the intrinsic nature of light, polarization is important to serve as a degree of freedom for manipulating light. The self-focusing property of the RPB under fixed input power was evaluated with the aid of SoP. The input power $P = 30{P_{cr}}$ is considered as an example, and the extension to other input powers is straightforward because of the deduction of the analytical expression of the self-focusing length for RPB (Eq. (11)). When the initial optical field cross-section has a spatially homogeneous SoP, that is, a hollow Gaussian beam (HGB), there is no analytical expression for the self-focusing length because of the variable propagation shape. Therefore, the numerical calculation mentioned in Section 3.1 is used. In Fig. 2, it is shown that the self-focusing length can be extended to 5.44 times the Rayleigh length, and that of the HGB is only 1.58 times the Rayleigh length, which may have a practical application significance for increasing the self-focusing length when the laser beam has a spatially inhomogeneous SoP at the field cross-section, and it may be a potential feasible solution to fill the shortage up with the method of the input power as well.

 figure: Fig. 2.

Fig. 2. Intensity distribution located at various propagation planes for (a) RPB and (b) HGB, where the corresponding propagation dynamics are (c) and (d), respectively.

Download Full Size | PDF

3.3 Self-focusing length with topological charge

In this section, the influence of the topological charge on the self-focusing length is evaluated using Stokes parameters. Herein, the topological charges were selected as $m = 1$ and $m = 2$. At the source plane $z = 0$ with the input power $P = 30{P_{cr}}$, the Stokes parameters for the case of $m = 1$ are shown in Fig. 3(a); as the propagation distance increases, the Stokes parameters are redistributed and the hollow beam converges into four filaments at the location of $z = 5.44{z_R}$ in Fig. 3(b). Similarly, for the $m = 2$ case, the Stokes parameters at the source plane are shown in Fig. 3(c), which requires a much higher input power to collapse, that is, $P = 90{P_{cr}}$, and converges to eight filaments at the plane of $z = 1.30{z_R}$ in Fig. 3(d). In physics, the axial symmetry of the light-matter nonlinear interaction is broken by the designable symmetrical polarization structure at the initial optical field cross-section and the self-focusing property of the propagation medium; there is no circular polarization (${S_3} = 0$) at the source plane in Fig. 3(a) and Fig. 3(c), while the circular polarization states occur on the receiver plane in Fig. 3(b) and Fig. 3(d) because the Stokes parameter ${S_3}$ is not equal to zero. It is clear that the hybrid polarization spatial structure is formed during the propagation process, so the symmetry is broken by refractive index change and self-focusing spots appear subsequently. As shown in Figs. 3(b) and (d), it is observed that a lower topological charge induces a longer self-focusing length; thus, the topological charge can also be regarded as a controllable parameter to manipulate the length of self-focusing. Additionally, the effective radius of the filaments of $m = 1$ is smaller than that of $m = 2$, which indicates that the power density of the lower topological charge is higher. It is possible that an RPB with a low topological charge as a laser source may be of great advantage for achieving long-range and high-power laser atmospheric propagation.

 figure: Fig. 3.

Fig. 3. Stokes parameters with different topological charges $m = 1$ (i.e., (a) $P = 30{P_{cr}}$ at the source plane $z = 0$, (b)$P = 30{P_{cr}}$ at the location of the self-focusing length $z = 5.44{z_R}$) and $m = 2$ (i.e., (c) $P = 90{P_{cr}}$ at the source plane $z = 0$, (d) $P = 90{P_{cr}}$ at the location of the self-focusing length $z = 1.30{z_R}$.

Download Full Size | PDF

4. Conclusion and discussion

In summary, an analytical expression of the self-focusing length of the RPB with respect to the input power was obtained. It was demonstrated that the self-focusing length with the input power has a similar performance to the semi-empirical Marburger formula for a Gaussian beam. Furthermore, the influence of the SoPs and topological charge on the self-focusing length were evaluated. It was found that the self-focusing length could be controlled by adjusting the SoP at the initial cross-sectional optical field and topological charge. Additionally, two characteristics of the RPB should be mentioned: (1) the self-focusing length of the RPB appears at a relatively long-range propagation distance compared with the HGB; (2) for a lower topological charge, the target optical power density of the RPB is larger than that of a higher topological charge. Therefore, the RPB as a laser source may not only extend the self-focusing length but also improve the power density of the target. These analytical and numerical findings may open a new window for realizing applications such as the multi-manipulation of optical fields, optical information transmission, and high-power long-range atmospheric propagation.

Funding

National Key Research and Development Program of China (2019YFA0705000); National Natural Science Foundation of China (11804234, 11903062, 11974218, 12192254, 91750201); Innovation Group of Jinan (2018GXRC010); Local Science and Technology Development Project of the Central Government (YDZX20203700001766).

Disclosures

The authors declare that they have no conflicts of interest.

Data availability

The 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. A. Couairon and A. Mysyrowicz, “Femtosecond filamentation in transparent media,” Phys. Rep. 441(2-4), 47–189 (2007). [CrossRef]  

2. J. Kasparian, M. Rodriguez, G. Méjean, J. Yu, E. Salmon, H. Wille, R. Bourayou, S. Frey, Y. B. André, A. Mysyrowicz, R. Sauerbrey, J. P. Wolf, and L. Wöste, “White-Light Filaments for Atmospheric Analysis,” Science 301(5629), 61–64 (2003). [CrossRef]  

3. A. Houard, C. D’Amico, Y. Liu, Y. B. André, M. Franco, B. Prade, E. Salmon, P. Pierlot, L. M. Cléon, and A. Mysyrowicz, “High current permanent discharges in air induced by femtosecond laser filamentation,” Appl. Phys. Lett. 90(17), 171501 (2007). [CrossRef]  

4. C. D’Amico, A. Houard, M. Franco, B. Prade, A. Mysyrowicz, A. Couairon, and V. T. Tikhonchuk, “Conical Forward THz Emission from Femtosecond-Laser-Beam Filamentation in Air,” Phys. Rev. Lett. 98(23), 235002 (2007). [CrossRef]  

5. C. D’Amico, A. Houard, S. Akturk, Y. Liu, J. Le Bloas, M. Franco, B. Prade, A. Couairon, V. T. Tikhonchuk, and A. Mysyrowicz, “Forward THz radiation emission by femtosecond filamentation in gases: theory and experiment,” New J. Phys. 10(1), 013015 (2008). [CrossRef]  

6. A. Houard, M. Franco, B. Prade, A. Durécu, L. Lombard, P. Bourdon, O. Vasseur, B. Fleury, C. Robert, V. Michau, A. Couairon, and A. Mysyrowicz, “Femtosecond filamentation in turbulent air,” Phys. Rev. A 78(3), 033804 (2008). [CrossRef]  

7. Z. Jin, J. Zhang, M. Xu, X. Lu, Y. Li, Z. Wang, Z. Wei, X. Yuan, and W. Yu, “Control of filamentation induced by femtosecond laser pulses propagating in air,” Opt. Express 13(25), 10424–10430 (2005). [CrossRef]  

8. L. Woste, C. Wedekind, H. Wille, P. Rairoux, B. Stein, S. Nikolov, C. Werner, S. Niedermeier, F. Ronnenberger, H. Schillinger, and R. Sauerbrey, “Femtosecond atmospheric lamp,” Laser und Optoelektronik 29, 51–53 (1997). [CrossRef]  

9. P. Sprangle, J.R. Penano, and B. Hafizi, “Propagation of intense short laser pulses in the atmosphere,” Phys. Rev. E 66, 046418 (2002). [CrossRef]  

10. G. Fibich, Y. Sivan, Y. Ehrlich, E. Louzon, M. Fraenkel, S. Eisenmann, Y. Katzir, and A. Zigler, “Control of the collapse distance in atmospheric propagation,” Opt. Express 14(12), 4946–4957 (2006). [CrossRef]  

11. L. B. Stotts, J. Peñano, and V. J. Urick, “Engineering equation for filamentation self-focusing collapse distance in atmospheric turbulence,” Opt. Express 27(11), 15159–15171 (2019). [CrossRef]  

12. G. DiComo, M. Helle, D. Kaganovich, A. Schmitt-Sody, J. Elle, and J. Peñano, “Nonlinear self-channeling of high-power lasers through controlled atmospheric turbulence,” J. Opt. Soc. Am. B 37(3), 797–803 (2020). [CrossRef]  

13. E. Snitzer, “Cylindrical Dielectric Waveguide Modes,” J. Opt. Soc. Am. 51(5), 491–498 (1961). [CrossRef]  

14. A. Ciattoni, B. Crosignani, P. Porto, and A. Yariv, “Azimuthally polarized spatial dark solitons: exact solutions of Maxwell's equations in a Kerr medium,” Phys. Rev. Lett. 94(7), 073902 (2005). [CrossRef]  

15. H. Kawauchi, K. Yonezawa, Y. Kozawa, and S. Sato, “Calculation of optical trapping forces on a dielectric sphere in the ray optics regime produced by a radially polarized laser beam,” Opt. Lett. 32(13), 1839–1841 (2007). [CrossRef]  

16. K. Lou, S. Qian, X. Wang, Y. Li, B. Gu, C. Tu, and H. Wang, “Two-dimensional microstructures induced by femtosecond vector light fields on silicon,” Opt. Express 20(1), 120–127 (2012). [CrossRef]  

17. X. Xie, Y. Chen, K. Yang, and J. Zhou, “Harnessing the point-spread function for high-resolution far-field optical microscopy,” Phys. Rev. Lett. 113(26), 263901 (2014). [CrossRef]  

18. X. Gao, Y. Pan, M. Zhao, G. Zhang, Y. Zhang, C. Tu, Y. Li, and H. Wang, “Focusing behavior of the fractal vector optical fields designed by fractal lattice growth model,” Opt. Express 26(2), 1597–1614 (2018). [CrossRef]  

19. Y. Cai, Q. Lin, H. T. Eyyuboğlu, and Y. Baykal, “Average irradiance and polarization properties of a radially or azimuthally polarized beam in a turbulent atmosphere,” Opt. Express 16(11), 7665–7673 (2008). [CrossRef]  

20. W. Chen, J. W. Haus, and Q. Zhan, “Propagation of vector vortex beams through a turbulent atmosphere,” Opt. Express 17(20), 17829–17836 (2009). [CrossRef]  

21. C. Wei, D. Wu, C. Liang, F. Wang, and Y. Cai, “Experimental verification of significant reduction of turbulence-induced scintillation in a full Poincaré beam,” Opt. Express 23(19), 24331–24341 (2015). [CrossRef]  

22. M. Cox, C. Guzmán, M. Lavery, D. Versfeld, and A. Forbes, “On the resilience of scalar and vector vortex modes in turbulence,” Opt. Express 24(16), 18105–18113 (2016). [CrossRef]  

23. M. Lavery, C. Peuntinger, K. Günthner, P. Banzer, D. Elser, R. W. Boyd, M. Padgett, C. Marquardt, and G. Leuchs, “Free-space propagation of high-dimensional structured optical fields in an urban environment,” Sci. Adv. 3(10), e1700552 (2017). [CrossRef]  

24. J. Yu, X. Zhu, F. Wang, D. Wei, G. Gbur, and Y. Cai, “Experimental study of reducing beam wander by modulating the coherence structure of structured light beams,” Opt. Lett. 44(17), 4371–4374 (2019). [CrossRef]  

25. W. Han, Y. Yang, W. Cheng, and Q. Zhan, “Vectorial optical field generator for the creation of arbitrarily complex fields,” Opt. Express 21(18), 20692–20706 (2013). [CrossRef]  

26. W. Han, W. Cheng, and Q. Zhan, “Design and alignment strategies of 4f systems used in the vectorial optical field generator,” Appl. Opt. 54(9), 2275–2278 (2015). [CrossRef]  

27. C. Wan and Q. Zhan, “Generation of exotic optical polarization Möbius strips,” Opt. Express 27(8), 11516–11524 (2019). [CrossRef]  

28. S. Stirling, S. Hend, A. Kamel, Q. Zhan, and F. Andrew, “General design principle for structured light lasers,” Opt. Express 28(23), 35006–35017 (2020). [CrossRef]  

29. S. Liu, S. Qi, Y. Zhang, P. Li, D. Wu, L. Han, and J. Zhao, “Highly efficient generation of arbitrary vector beams with tunable polarization, phase, and amplitude,” Photonics Res. 6(4), 228–233 (2018). [CrossRef]  

30. C. Ma, J. Di, Y. Zhang, P. Li, F. Xiao, K. Liu, X. Bai, and J. Zhao, “Reconstruction of structured laser beams through a multimode fiber based on digital optical phase conjugation,” Opt. Lett. 43(14), 3333–3336 (2018). [CrossRef]  

31. C. Ma, J. Di, J. Dou, P. Li, F. Xiao, K. Liu, X. Bai, and J. Zhao, “Structured light beams created through a multimode fiber via virtual Fourier filtering based on digital optical phase conjugation,” Appl. Opt. 59(3), 701–705 (2020). [CrossRef]  

32. Z. Chen, T. Zeng, B. Qian, and J. Ding, “Complete shaping of optical vector beams,” Opt. Express 23(14), 17701–17710 (2015). [CrossRef]  

33. L. Li, C. Chang, C. Yuan, S. Feng, S. Nie, Z. Ren, H. Wang, and J. Ding, “High efficiency generation of tunable ellipse perfect vector beams,” Photonics Res. 6(12), 1116–1123 (2018). [CrossRef]  

34. X. Wang, Y. Li, J. Chen, C. Guo, J. Ding, and H Wang, “A new type of vector fields with hybrid states of polarization,” Opt. Express 18(10), 10786–10795 (2010). [CrossRef]  

35. S. Li, Y. Li, X. Wang, L. Kong, K. Lou, C. Tu, Y. Tian, and H. Wang, “Taming the Collapse of Optical Fields,” Sci. Rep. 2(1), 1007 (2012). [CrossRef]  

36. R. Chen, L. Zhong, K. Chew, T. Zhao, and X. Zhang, “Collapse dynamics of a vector vortex optical field with inhomogeneous states of polarization,” Laser Phys. 25(7), 075401 (2015). [CrossRef]  

37. S. Li, Z. Ren, L. Kong, S. Qian, C. Tu, Y. Li, and H. Wang, “Unveiling stability of multiple filamentation caused by axial symmetry breaking of polarization,” Photonics Res. 4(5), B29–B34 (2016). [CrossRef]  

38. G. Wu, F. Wang, and Y. Cai, “Coherence and polarization properties of a radially polarized beam with variable spatial coherence,” Opt. Express 20(27), 28301–28318 (2012). [CrossRef]  

39. F. Wang, Y. Cai, Y. Dong, and O. Korotkova, “Experimental generation of a radially polarized beam with controllable spatial coherence,” Appl. Phys. Lett. 100(5), 051108 (2012). [CrossRef]  

40. S. Zhu, J. Wang, X. Liu, Y. Cai, and Z. Li, “Generation of arbitrary radially polarized array beams by manipulating correlation structure,” Appl. Phys. Lett. 109(16), 161904 (2016). [CrossRef]  

41. C. Ping, C. Liang, F. Wang, and Y. Cai, “Radially polarized multi-Gaussian Schell-model beam and its tight focusing properties,” Opt. Express 25(26), 32475–32490 (2017). [CrossRef]  

42. J. Marburger and E. Dawes, “Dynamical formation of a small-scale filament,” Phys. Rev. Lett. 21(8), 556–558 (1968). [CrossRef]  

43. J. Peñano, B. Hafizi, A. Ting, and M. Helle, “Theoretical and numerical investigation of filament onset distance in atmospheric turbulence,” J. Opt. Soc. Am. B 31(5), 963–971 (2014). [CrossRef]  

44. J. Yu, X. Zhu, S. Lin, F. Wang, G. Gbur, and Y. Cai, “Vector partially coherent beams with prescribed non-uniform correlation structure,” Opt. Lett. 45(13), 3824–3827 (2020). [CrossRef]  

45. D. Peng, Z. Huang, Y. Liu, Y. Chen, F. Wang, S. A. Ponomarenko, and Y. Cai, “Optical coherence encryption with structured random light,” PhotoniX 2(1), 6 (2021). [CrossRef]  

46. Y. Liu, Y. Chen, F. Wang, Y. Cai, C. Liang, and O. Korotkova, “Robust far-field imaging by spatial coherence engineering,” Opto-Electron. Adv. 4(12), 210027 (2021). [CrossRef]  

47. C. Menyuk, “Stability of solitons in birefringent optical fibers. II. Arbitrary amplitudes,” J. Opt. Soc. Am. B 5(2), 392–402 (1988). [CrossRef]  

48. Q. Park and H. J. Shin, “Painlevé analysis of the coupled nonlinear Schrödinger equation for polarized optical waves in an isotropic medium,” Phys. Rev. E 59(2), 2373–2379 (1999). [CrossRef]  

49. A. Dubietis, E. Gaizauskas, G. Tamŏsauskas, and P. Di Trapani, “Light filaments without self-channeling,” Phys. Rev. Lett. 92(25), 253903 (2004). [CrossRef]  

Data availability

The 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 (3)

Fig. 1.
Fig. 1. ${{{z_f}} / {{z_R}}}$ as a function of the relative input laser power ${P / {{P_{cr}}}}$ for RPB.
Fig. 2.
Fig. 2. Intensity distribution located at various propagation planes for (a) RPB and (b) HGB, where the corresponding propagation dynamics are (c) and (d), respectively.
Fig. 3.
Fig. 3. Stokes parameters with different topological charges $m = 1$ (i.e., (a) $P = 30{P_{cr}}$ at the source plane $z = 0$, (b)$P = 30{P_{cr}}$ at the location of the self-focusing length $z = 5.44{z_R}$) and $m = 2$ (i.e., (c) $P = 90{P_{cr}}$ at the source plane $z = 0$, (d) $P = 90{P_{cr}}$ at the location of the self-focusing length $z = 1.30{z_R}$.

Equations (11)

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

2 i k E z + 2 E + 4 n 2 k 2 3 n 0 [ 2 ( E E ) E + ( E E ) E ] = 0 ,
E ( r , φ , z = 0 ) = A ( r ) ( cos δ e h + sin δ e v ) ,
2i k E h z + 2 E h + 4 n 2 k 2 3 n 0 [ 3 | E h | 2 E h + 2 | E v | 2 E h + E v 2 E h ] = 0 2i k E v z + 2 E v + 4 n 2 k 2 3 n 0 [ 3 | E v | 2 E v + 2 | E h | 2 E v + E h 2 E v ] = 0
E ( r , z ) = A z ( r w z ) exp ( r 2 w z 2 + i r 2 φ z ) ( cos δ e h + sin δ e v )
d w z d z = 2 k φ z A z ,
d φ z d z = 1 2 k w z 4 2 φ z 2 k + 2 n 2 k A z 2 n 0 w z 2 ,
d A z d z = 2 k φ z A z .
d 2 w z d z 2 = 1 k 2 w z 3 + 4 n 2 A z 2 n 0 w z ,
( d w z d z ) 2 + ( 4 w z 2 4 w 0 2 ) + 16 n 2 k A z 2 n 0 ln ( w z w 0 ) = 0.
w z 2 = w 0 2 + z 2 k 2 w 0 2 + 4 n 2 A z 2 z 2 n 0 ( 1 ln A z A 0 ) .
z f = z R [ 1 + ln ( A z / A 0 ) ] P in / P cr 0.25 .
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.