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

Self-mode-locked Laguerre-Gaussian beam with staged topological charge by thermal-optical field coupling

Open Access Open Access

Abstract

A light beam with a helical phase is associated with an optical vortex and carries optical orbital angular momentum. Mode-locked optical vortex pulses impart orbital angular momentum to photons in short pulses and have attractive applications. However, due to the conflict between mature mode-locking and the generation of optical vortices, directly generated mode-locked optical vortex short pulses seem to be unavailable, thus constraining the development and applications of optical vortex short pulses. Laguerre-Gaussian (LG) modes are eigenfunctions for a laser cavity. Besides carrying optical orbital angular momentum, LG beams also have self-healing and quasi-nondiffracting properties. Here, we report the realization of a self-mode-locked LG lasers with tunable orbital angular momentum. By coupling between the thermal and optical fields, the orbital angular momentum was found to be staged. These results verify the possibility of direct mode-locking of optical vortices, and may open the way for several applications of short pulses. Moreover, mode-locked pulses with high-repetition rates also have particularly attractive applications such as optical frequency comb spectroscopy, high capacity optical networks, spectroscopy of metallic nanoparticles, arbitrary waveform generation, etc..

© 2016 Optical Society of America

Introduction

The vector nature of light supports both spin and orbital phase singularities. The former is associated with light polarization and has the value of ± per photon, depending on the chirality of the polarization [1–4 ], where ±is Planck’s constant; the latter is associated with a vortex that is determined by the phase distribution [4–6 ].When the wavefront of a light beam becomes helical and is described byexp(ilφ), where φ is the azimuthal coordinate and l is the helicity, the phase at the center where an optical vortex exists becomes undefined and the amplitude of the electric vectors vanishes. Using the angular momentum operator along the z-direction, Lz=iφ to calculate the eigenvalues, the photons in a light beam with a helical phase have an orbital angular momentum of l per photon, where l also denotes the topological charge. Studies of optical vortices, including their generation [7–9 ], structure [10], collapse [11], propagation dynamics [12–14 ], knots [15,16 ], entanglement [17,18 ], interaction with matter [19–21 ], etc., have helped to shed light on their nature and have also led to the consideration of intriguing applications such as twisting and controlling particles [22], quantum information communications [23,24 ], microscopy and lithography that surpass the diffraction limit [25,26 ] etc.

Ultrashort optical pulses with pulse length in the range of picoseconds or femtoseconds have become essential tools in the study of ultrafast processes and in the generation of high-intensity physical fields. Ultrashort optical vortex pulses would impart a new physical parameter to optical vortex fields and combine the properties of ultrashort and optical vortex beams. This prospect in turn suggests promising applications in the study and control of ultrafast physical and chemical processes, ultrafast nonlinear spectroscopy, nanomachining, terahertz vortex generation [27–30 ], etc. However, ultrashort pulses can only be achieved by mode-locking that results from the wave interference of phase-locked longitudinal laser modes which possess a stationary phase relationship in the longitudinal direction. Due to the interaction between multiple-transverse modes and different optical frequencies as determined by the transverse distribution and the different optical phases, the existence of a fundamental transverse (TEM00) mode is the primary condition necessary for mode-locking in general. Up to now, mode-locked optical vortex pulses have only been obtained by phase modulation of mode-locked light beams with spatial light modulators [27,29–32 ]. Moreover, the resolution of the modulator determines the purity and the other properties of the optical vortices, and may induce break up or loss of optical vortices during propagation or light-matter interaction [33–36 ]. No experimental evidence shows that an optical vortex can be mode-locked. The current situation constrains further investigation and applications of ultrashort optical vortices, such as carrier-vortex interaction dynamics, compact lasers for nanomachining, study and control of ultrafast processes, etc. Directly generated mode-locked optical vortex pulses of high purity are urgentlydesired. Laguerre-Gaussian (LGp,l,s) modes are transverse laser modes with topological charge l [37], where integers p, l and s denote the radial, azimuthal and longitudinal mode index, respectively. Besides carrying optical orbital angular momentum, LGp,l,s beams also have self-healing and quasi-nondiffracting properties which have some advantages compared to Bessel beams, which are another type of beam that carrys optical orbital angular momentum and that have, in some respects, well-known self-healing quasi-nondiffracting properties [38]. Recently, we also found that the thermal-optical field coupling that occurs during laser generation can change the transverse LGp,l,s modes in a way that corresponds to topological charge by mode-matching [39,40 ]. In this work, based on the development of mode-locking theory [41], we report the production of a self-mode-locked LGp,l,s pulse with staged topological charge using a simple configuration by that employs the coupling between thermal and optical fields during the lasing process.

2. Theoretical and experimental section

The time (t) dependent wave function of the LGp,l,s mode in cylindrical coordinates (ρ, φ, z) is given as Eq. (1) [8,42 ]:

ψ(ρ,φ,z,t)=eiωteilφϕp,l,s=eiωteilφ2p!π(p+|l|)!1w(z)(2ρw(z))|l|Lp|l|(2ρ2w(z)2)×exp(ρ2w(z)2)exp{ikz[1+ρ22(z2+zR2)]}×exp[i(2p+|l|+1)θG(z)]
where ω is the resonant angular frequency in the laser cavity, w(z)=w01+(z/zR)2, w0 is the beam radius at the waist, zR=πw02/λ is the Rayleigh range, λ is the light wavelength, Lp|l| are the Laguerre polynomials, k is the wave number, and θG(z)=tan1(z/zR) is the Gouy phase. For a Fabry-Perot (F-P) cavity with length L, the resonant frequency is expressed as ω=sΔfL+(2p+|l|)ΔfT, where ΔfL=2πΔf is the longitudinal mode range, Δf is the resonant frequency range in the laser cavity, andΔfT is the transverse mode range. In an empty symmetric resonator, the ratio between the transverse and longitudinal mode ranges is a constant [42]. For LGp,l,s modes with multiple-longitudinal modes and a single transverse mode, the indices p and l are constants and but the the s index is variable, and indicates that the angular frequency range Δω=ΔfL and the wave function of the LGp,l,s mode can be described as:
ψi(ρ,φ,z,t)=seisΔωteiω0teilφϕp,l,s
where ω0 is the fundamental resonant angular frequency. From Eq. (2), it is found that the interference of the LGp,l,s modes with the multiple-longitudinal modes and the single transverse mode can generate ideal mode-locking. It should be noted that the etalon effect formed by the surfaces of the crystals used in the cavity can separate the resonant frequency range and generate high-order harmonic mode-locking [43,44 ]. For instance, when the ratio (R) between the optical length of the crystal (d) and the external F-P cavity formed by the end face of the crystal and the output coupler (L) is equal to 12, the frequency range is doubled, and second-order harmonic mode-locking can be accomplished with the half-pulse period of the empty cavity.

In the experiment, it has been demonstrated that a pump beam with a doughnut shape can generate LGp,l,s laser modes with tunable topological charges, since the intensity distribution of the pump spot can be easily mode-matched with the LGp,l,s laser modes [39,40 ]. In this experiment, a commercial fiber-coupled laser diode (LD) with a central wavelength of 808 nm and doughnut intensity distribution was used as the pump source for the exciting the laser [40,45 ]. The configuration of the mode-locked laser was shown in Fig. 1 . By means of an optical fiber, the pump power was delivered to the focal system. The core size of the fiber was 400 μm in radius, and the numerical aperture was 0.22. The focal system focused the power from the fiber onto the crystal with a beam compression ratio of 2:1. The pump light on the crystal (pump mode) had a beam waist of 200 μm in radius. The gain medium is a neodymium doped (Lu0.5Y0.5)2SiO5 (Nd:LYSO) crystal with doping concentration of 0.5 at%. This crystal was cut along the crystalline b-axis with dimensions of 3 mm × 3 mm × 7 mm, which meaned that the pump and oscillating light propagate along the b-axis. The optical length of the crystal was 12.7 mm. The surfaces of the crystal perpendicular to the b-axis were polished. The surfaces were high-transmission (HT) and high-reflection(HR) coated at 808 nm and 1077 nm, respectively. In order to reduce the heat generated in the lasing process, the crystal was wrapped with indium foil and mounted on a water-cooled copper block with a cooling temperature of 20°C. A concave mirror with a radius of curvature of 50 mm was used as the input mirror, and was HT coated at the pump wavelength and HR coated at 1078 nm. The crystal was placed at a position as close as possible to the input mirror. A plane mirror with a reflection of 95% at 1078 nm was used as the output coupler. The laser cavity consisted of the front mirror and the output coupler, and had a length of about 33 mm, which means that the length of the external F-P cavity between the end surface of the crystal and the output coupler was about 26 mm, corresponding to R=12 . The pulse performance was recorded with a photoelectric detector and oscilloscope, both of which had a bandwidth of 25 GHz and a rise time of 16 ps.The laser patterns were recorded by a charge-coupled device (CCD). The mode-locked pulse trains were detected with an InGaAs photodetector (New focus, 1414 model) with a rise time of 14 ps, and recorded by a digital oscilloscope (Tektronix, MSO 72504DX) with a bandwidth of 25 GHz and a rise time of 16 ps. The laser spectrum was monitored by a spectrometer (Thorlab, OSA205) with a resolution of 0.03 nm.

 figure: Fig. 1

Fig. 1 Experimental configuration for generation of mode-locked optical vortex pulses. The pump source is a LD with doughnut-shaped output intensity. The output laser pulse performance was recorded by the oscilloscope and the pattern simultaneously observed with a CCD.

Download Full Size | PDF

Based on the detailed self-mode-locked model developed by Xie et al. [46], thermal lens aberration induces considerable phase distortion and diffraction losses in the lasing process. The diffraction loss is negligible only in the center of the pumped region and rapidly increases with increase in the oscillating transverse laser modes associated with the mode size. However, an increase of the Kerr effect would decrease the mode size and thus generate modulation of the diffraction losses. With the ABCD matrix formation of the present cavity, the oscillating laser mode size changes in the crystal due to the Kerr lens effect, and is calculated to be about 0.1 μm at the absorbed pump power of 2.5 W and output power of about 200 mW in the LG0,0,s modes. This effect induced a round trip diffraction loss modulation decrease of 6.0 × 10−4, enough for the generation of self-mode-locking [47]. Using the same method, and considering the distribution of the oscillating LGp,l,s modes shown in Eq. (1), the LG0,1,s laser mode size change due to the Kerr lens effect in the crystal was calculated to be about 0.21 μm at the absorbed pump power of 4.27 W and output power of about 594 mW. The modulation of the round trip diffraction loss is about 1.6 × 10−3. For the LG0,2,s modes, the laser mode size change in the crystal was calculated to be about 0.48 μm, and the decrease of the round trip diffraction loss modulation was about 4.0 × 10−3 at the absorbed pump power of 6.1 W and output power of about 1082 mW. It can also be assumed that with an increase of pump power, the diffraction loss modulation will also be increased due to the increase of the Kerr lens effect. The pulse width of the LGp,l,s modes will decrease with the increase in pump power and increase with p and l.

3. Results and discussions

Continuous-wave mode-locking appeared when the pump power was raised above the threshold. Figure 2(a) shown the pulse trains observed on the 1.25 ns scope with the pump power above the threshold. From these figures, we also observe that the mode-locking had a period of 121 ps, corresponding to a frequency of 7.8 GHz. This frequency indicated that the observed laser excitation was due to second-harmonic mode-locking with R=12 in the present experimental conditions. The pulse width was measured to be 37 ps. In that situation, the observed pattern, as shown in Fig. 3(a) (top) indicats that the achieved mode-locking was with the transverse TEM00 mode, corresponding to LGp,l,s modes with p = l = 0. The phase distribution of the laser beam can be analyzed with a cylindrical lens [48]. Upon examination with a cylindrical lens, the corresponding pattern is shown in Fig. 3(a) (middle), and it showns that there is only one strip and that the mode-locked beam obtained has no topological charge.

 figure: Fig. 2

Fig. 2 Performance of the experimental self-mode-locked laser. a-c. Pulse trains observed in the time range of 1.25 ns (250ps/div) with a .LG0,0,s, b. LG0,1,s and c. LG0,2,s modes. d. Pulse trains observed in the time range of 150 ns. e.RF spectrum with frequency of 7.8 GHz. f. Laser spectrum of the mode-locked LG0,2,s pulses at wavelength of 1078.2 nm.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Detailed characterization of experimental transverse pattern of mode-locked LGp,l,s beams with topological charges of l. The top images are mode-locked LGp,l,s beams detected by the CCD,the middle images are the phase distributions of LGp,l,s beams separated by a cylindrical lens, the bottom images are the intensity distributions of the LGp,l,s beams fitted by Eq. (1) (a) LG0, 0,s mode without topological charge; (b) LG0, 1,s mode with topological charge of 1; and (c) LG0, 2,s mode with topological charge of 2.

Download Full Size | PDF

With an increase of the pump power, the higher-order LG0,1,s and LG0,2,s modes with topological charge of 1 and 2, respectively, appeared step by step, as shown in Fig. 4 . The mode-locking trains of the LG0,1,s and LG0,2,s modes were presented in Figs. 2(b) and 2(c) with pulse widths of 35 and 28 ps, respectively. Compared with the mode-locked pulse width of the LG0,0,s modes (37 ps), we also find that under high pump power, the nonlinear Kerr-effect indeed participates in the mode-locking process, and increases the modulation depth, generating shorter pulses in agreement with the analysis of the modulation by thermally induced loss. The observed patterns were shown in Figs. 3(b) and 3(c) (top), respectively. With the same cylindrical lens, the corresponding focused patterns of the LG0,1,s and LG0,2,s modes are shown in Figs. 3(b) and 3(c) (middle). Two and three clearly obvious strips appear, which verify that the achieved mode-locking consisted of optical vortex pulses with topological charge of 1 and 2, respectively. In order to describe the beams in more detailed, the intensity distributions of the LGp,l,s beams are shown in Fig. 3(bottom). As can be seen at the bottom of the figure, the experimental data curve can be fitted well by Eq. (1) [8,49 ]. Taking the LG0,1,s mode as representative, the radio-frequency (RF) spectrum with 100 KHz resolution was measured to confirm the stability of the self-mode-locked Nd:LYSO laser and is shown in Fig. 2(d). According to the figure, the signal-to-noise ratio is greater than 45 dB above the background level. The FWHM of the single pulse trace in the LG0,1,s mode shown on the oscilloscope was about 35 ps. Assuming a sech2 pulse profile, the pulse width of the self-mode-locked laser is estimated to be about 23 ps, which is shown in Fig. 2(e). The optical spectrum of the self-mode-locked pulses in the LG0,2,s mode is also displayed in Fig. 2(f) with a full width at half maximum (FWHM) of 0.11 nm at the wavelength of 1078.2 nm. Consequently, the time-bandwidth product of the mode-locked pulse is 0.59, which is a bit larger than the transform-limited sech2 pulse, indicating that the pulses are frequency-chirped. The other two experiments on LG0,0,s and LG0,2,s also had similar results. The maximum output power was 1.08 W under a the pump power of 6.07 W with conversion efficiency of 23%.

 figure: Fig. 4

Fig. 4 Laser output power and staged topological charge of LGp,1,s modes with the increase of absorbed pump power.

Download Full Size | PDF

For the staged topological charge shown in Fig. 4, mode-matching should be considered in this simple experimental configuration. In fact, mode-matching is a dynamic balancing process that is determined by the overlap of the oscillating and pump modes, and which is influenced by the thermally induced diffraction loss in the cavity [45]. The mode-matching dynamics can be qualitatively described as the follows: first, at the threshold, mode-matching requires that the oscillating mode with the smallest size determine the LG0,0,s mode, since the size of the LGp,l,s modes is expressed as wp,l,s=w02p+|l|+1 [50,51 ]; second, in order to convert the pump energy efficiently, the oscillating mode size must be comparable with that of the pump mode; third, the thermally induced diffraction loss generated by the thermal focal lens produced by the pump power dictates a smaller oscillating mode than the pump mode; and finally, the pump light intensity is distributed in a doughnut shape which removes the low-order laser modes when the higher-order modes are matched well with the pump beam. The optimized mode-size in the presence of the thermal focal lens is calculated to be about 0.9 times the pump mode size. Besides, the thermal focal lens also decreases the oscillating mode size in the present experimental configuration, a consequence of the thermal-optical coupling effect. By measuring the thermal focal length under different pump power values, the optimized oscillating modes were calculated based on the mode-matching theory [39,45 ]. They were associated with the present laser performance, which showed that with an increase of pump power, the optimized oscillating mode size increased and approached 180 μm. The optimized oscillating mode size at the end of the topological charge stages is located in the middle of the adjacent mode sizes. It should be noted that the LG1,0,s mode with no topological charge requires higher pump power than the LG0,2,s mode [52]. Associated with the mode-matching between the doughnut pump and the LGp,l,s modes [49,52 ], the topological charge was staged by the thermal-optical coupling process.

4. Conclusion

In conclusion, we have demonstrated the possibility of achieving self-mode-locking of short optical vortex pulses with tunable topological charge. By theoretical analysis, we found that mode-locked lasers with LGp,l,s modes can be achieved. With a simple two-mirror configuration, mode-locked pulses with a topological charge of 0, 1 and 2 were realized at a repetition rate of 7.8 GHz. Due to thermal-optical field coupling, the topological charge was selected by mode-matching, which determines its step-by-step appearance. Associated with the particle-like properties of phase singularity and mode-locking that result from wave interference, this work should provide a platform for the further study of wave-particle dualism in light beams. In optics, this work verifies the existence of the direct mode-locking of optical vortices and may open the way for further investigation and application of the short optical vortex, self-healing and quasi-nondiffracting pulses such as in ultrafast physical and chemical processes, carrier and vortex dynamics, optical communications [44],etc. We also propose that by optimizing the gain medium and by using suitable dispersion compensation, it is possible to realize and study the direct mode-locking of optical vortices with femtosecond pulses and optical solitons.

Acknowledgments

The authors wish to thank Prof. R.I.Boughton, Department of Physics and Astronomy of Bowling Green State University, for useful discussions on physics and linguistic advice. This work is supported by the National Natural Science Foundation of China (Nos. 51422205 and 51272131), the Natural Science Foundation for Distinguished Young Scholars of Shandong Province (JQ201415) and Taishan Scholar Foundation of Shandong Province, China.

References and links

1. M. V. Berry, “The electric and magnetic polarization singularities of paraxial waves,” J. Opt. A, Pure Appl. Opt. 6(5), 475–481 (2004). [CrossRef]  

2. J. F. Nye, “Polarization Effects in the diffraction of electromagnetic waves: the role of disclinations,” Proc. R. Soc. Lond. A Math. Phys. Sci. 387(1792), 105–132 (1983). [CrossRef]  

3. R. E. Beth, “Mechanical detection and measurement of the angular momentum of light,” Phys. Rev. 50(2), 115–125 (1936). [CrossRef]  

4. P. Němec, E. Rozkotová, N. Tesařová, F. Trojánek, E. De Ranieri, K. Olejník, J. Zemen, V. Novák, M. Cukr, P. Malý, and T. Jungwirth, “Experimental observation of the optical spin transfer torque,” Nat. Phys. 8(5), 411–415 (2012). [CrossRef]  

5. G. Molina-Terriza, J. P. Torres, and L. Torner, “Twisted photons,” Nat. Phys. 3(5), 305–310 (2007). [CrossRef]  

6. M. V. Berry, “The adiabatic phase and Pancharatnam’s phase for polarized light,” J. Mod. Opt. 34(11), 1401–1407 (1987). [CrossRef]  

7. J. Scheuer and M. Orenstein, “Optical vortices crystals: Spontaneous generation in nonlinear semiconductor microcavities,” Science 285(5425), 230–233 (1999). [CrossRef]   [PubMed]  

8. D. Naidoo, T. Godin, M. Fromager, E. Cagniot, N. Passilly, A. Forbes, and K. Aït-Ameur, “Transverse mode selection in a monolithic microchip laser,” Opt. Commun. 284(23), 5475–5479 (2011). [CrossRef]  

9. M. Brambilla, F. Battipede, L. A. Lugiato, V. Penna, F. Prati, C. Tamm, and C. O. Weiss, “Transverse laser patterns. I. Phase singularity crystals,” Phys. Rev. A 43(9), 5090–5113 (1991). [CrossRef]   [PubMed]  

10. L. T. Vuong, T. D. Grow, A. Ishaaya, A. L. Gaeta, G. W. ’t Hooft, E. R. Eliel, and G. Fibich, “Collapse of optical vortices,” Phys. Rev. Lett. 96(13), 133901 (2006). [CrossRef]   [PubMed]  

11. J. E. Curtis and D. G. Grier, “Structure of optical vortices,” Phys. Rev. Lett. 90(13), 133901 (2003). [CrossRef]   [PubMed]  

12. M. P. J. Lavery, F. C. Speirits, S. M. Barnett, and M. J. Padgett, “Detection of a spinning object using light’s orbital angular momentum,” Science 341(6145), 537–540 (2013). [CrossRef]   [PubMed]  

13. D. Rozas, C. T. Law, and G. A. Swartzlander Jr, “Propagation dynamics of optical vortices,” J. Opt. Soc. Am. B 14(11), 3054–3065 (1997). [CrossRef]  

14. F. C. Speirits, M. P. Lavery, M. J. Padgett, and S. M. Barnett, “Optical angular momentum in a rotating frame,” Opt. Lett. 39(10), 2944–2946 (2014). [CrossRef]   [PubMed]  

15. J. Leach, M. R. Dennis, J. Courtial, and M. J. Padgett, “Laser beams: Knotted threads of darkness,” Nature 432(7014), 165 (2004). [CrossRef]   [PubMed]  

16. M. R. Dennis, R. P. King, B. Jack, K. O’Holleran, and M. J. Padgett, “Isolated optical vortex knots,” Nat. Phys. 6(2), 118–121 (2010). [CrossRef]  

17. J. Romero, J. Leach, B. Jack, M. R. Dennis, S. Franke-Arnold, S. M. Barnett, and M. J. Padgett, “Entangled optical vortex links,” Phys. Rev. Lett. 106(10), 100407 (2011). [CrossRef]   [PubMed]  

18. A. Mair, A. Vaziri, G. Weihs, and A. Zeilinger, “Entanglement of the orbital angular momentum states of photons,” Nature 412(6844), 313–316 (2001). [CrossRef]   [PubMed]  

19. L. Huang, X. Chen, H. Mühlenbernd, G. Li, B. Bai, Q. Tan, G. Jin, T. Zentgraf, and S. Zhang, “Dispersionless phase discontinuities for controlling light propagation,” Nano Lett. 12(11), 5750–5755 (2012). [CrossRef]   [PubMed]  

20. B. Y. Wei, W. Hu, Y. Ming, F. Xu, S. Rubin, J. G. Wang, V. Chigrinov, and Y. Q. Lu, “Generating switchable and reconfigurable optical vortices via photopatterning of liquid crystals,” Adv. Mater. 26(10), 1590–1595 (2014). [CrossRef]   [PubMed]  

21. W. Cheong, W. Lee, X. Yuan, L. Zhang, K. Dholakia, and H. Wang, “Direct electron-beam writing of continuous spiral phase plates in negative resist with high power efficiency for optical manipulation,” Appl. Phys. Lett. 85(23), 5784–5786 (2004). [CrossRef]  

22. X. L. Wang, J. Chen, Y. Li, J. Ding, C. S. Guo, and H. T. Wang, “Optical orbital angular momentum from the curl of polarization,” Phys. Rev. Lett. 105(25), 253602 (2010). [CrossRef]   [PubMed]  

23. J. Wang, J. Yang, I. M. Fazal, N. Ahmed, Y. Yan, H. Huang, Y. Ren, Y. Yue, S. Dolinar, M. Tur, and A. E. Willner, “Terabit free-space data transmission employing orbital angular momentum multiplexing,” Nat. Photonics 6(7), 488–496 (2012). [CrossRef]  

24. X. L. Wang, X. D. Cai, Z. E. Su, M. C. Chen, D. Wu, L. Li, N. L. Liu, C. Y. Lu, and J. W. Pan, “Quantum teleportation of multiple degrees of freedom of a single photon,” Nature 518(7540), 516–519 (2015). [CrossRef]   [PubMed]  

25. W. L. Barnes, A. Dereux, and T. W. Ebbesen, “Surface plasmon subwavelength optics,” Nature 424(6950), 824–830 (2003). [CrossRef]   [PubMed]  

26. L. Du, D. Tang, and X. Yuan, “Detection of microscope-excited surface plasmon polaritons with Rayleigh scattering from metal nanoparticles,” Appl. Phys. Lett. 103(18), 181101 (2013). [CrossRef]  

27. G. F. Quinteiro and J. Berakdar, “Electric currents induced by twisted light in quantum rings,” Opt. Express 17(22), 20465–20475 (2009). [CrossRef]   [PubMed]  

28. Y. S. Lee, Principles of Terahertz Science and Technology (Springer, Berlin, 2009).

29. M. Bock, J. Jahns, and R. Grunwald, “Few-cycle high-contrast vortex pulses,” Opt. Lett. 37(18), 3804–3806 (2012). [CrossRef]   [PubMed]  

30. C. Hnatovsky, V. G. Shvedov, W. Krolikowski, and A. V. Rode, “Materials processing with a tightly focused femtosecond laser vortex pulse,” Opt. Lett. 35(20), 3417–3419 (2010). [CrossRef]   [PubMed]  

31. K. Yamane, Y. Toda, and R. Morita, “Ultrashort optical-vortex pulse generation in few-cycle regime,” Opt. Express 20(17), 18986–18993 (2012). [CrossRef]   [PubMed]  

32. H. C. Liang, Y. J. Huang, Y. C. Lin, T. H. Lu, Y. F. Chen, and K. F. Huang, “Picosecond optical vortex converted from multigigahertz self-mode-locked high-order Hermite-Gaussian Nd:GdVO4 lasers,” Opt. Lett. 34(24), 3842–3844 (2009). [CrossRef]   [PubMed]  

33. M. Zürch, C. Kern, P. Hansinger, A. Dreischuh, and C. Spielmann, “Strong-field physics with singular light beams,” Nat. Phys. 8(10), 743–746 (2012). [CrossRef]  

34. L. Torner and D. V. Petrov, “Splitting of light beams with spiral phase dislocations into solitons in bulk quadratic nonlinear media,” J. Opt. Soc. Am. B 14(8), 2017–2023 (1997). [CrossRef]  

35. D. Rozas, C. T. Law, and G. A. Swartzlander Jr, “Propagation dynamics of optical vortices,” J. Opt. Soc. Am. B 14(11), 3054–3065 (1997). [CrossRef]  

36. G. Indebetouw, “Optical vortices and their propagation,” J. Mod. Opt. 40(1), 73–87 (1993). [CrossRef]  

37. L. Allen, M. J. Padgett, and M. Babiker, “The orbital angular momentum of light,” Prog. Opt. 39, 291–372 (1999). [CrossRef]  

38. J. Mendoza-Hernández, M. L. Arroyo-Carrasco, M. D. Iturbe-Castillo, and S. Chávez-Cerda, “Laguerre-Gauss beams versus Bessel beams showdown: peer comparison,” Opt. Lett. 40(16), 3739–3742 (2015). [CrossRef]   [PubMed]  

39. Y. Zhao, Z. Wang, H. Yu, S. Zhuang, H. Zhang, X. Xu, J. Xu, X. Xu, and J. Wang, “Direct generation of optical vortex pulses,” Appl. Phys. Lett. 101(3), 031113 (2012). [CrossRef]  

40. Y. Chen and Y. Lan, “Observation of transverse patterns in an isotropic microchip laser,” Phys. Rev. A 67(4), 043814 (2003). [CrossRef]  

41. U. Keller, Ultrafast Solid-state Lasers, Vol.VIII, G. Herziger, H. Weber, and R. Proprawe, eds (Springer, 2007).

42. Y. F. Chen, T. H. Lu, and K. F. Huang, “Hyperboloid structures formed by polarization singularities in coherent vector fields with longitudinal-transverse coupling,” Phys. Rev. Lett. 97(23), 233903 (2006). [CrossRef]   [PubMed]  

43. M. F. Becker, D. J. Kuizenga, and A. E. Siegman, “Harmonic mode locking of the Nd:YAG laser,” IEEE J. Quantum Electron. 8(8), 687–693 (1972). [CrossRef]  

44. Y. Chen, M. Chang, W. Zhuang, K. Su, K. Huang, and H. Liang, “Generation of sub-terahertz repetition rates from a monolithic self-mode-locked laser coupled with an external Fabry-Perot cavity,” Laser Photonics Rev. 9(1), 91–97 (2015). [CrossRef]  

45. Y. Chen, T. Huang, C. Kao, L. Wang, and S. Wang, “Optimization in scaling fiber-coupled laser-diode end-pumped lasers to higher power: influence of thermal effect,” IEEE J. Quantum Electron. 33(8), 1424–1429 (1997). [CrossRef]  

46. G. Q. Xie, D. Y. Tang, L. M. Zhao, L. J. Qian, and K. Ueda, “High-power self-mode-locked Yb:Y(2)O(3) ceramic laser,” Opt. Lett. 32(18), 2741–2743 (2007). [CrossRef]   [PubMed]  

47. H. Liang, H. Chang, W. Huang, K. Su, Y. Chen, and Y. Chen, “Self-mode-locked Nd:GdVO4 laser with multi-GHz oscillations: manifestation of third-order nonlinearity,” Appl. Phys. B 97(2), 451–455 (2009). [CrossRef]  

48. S. Li, L. Kong, Z. Ren, Y. Li, C. Tu, and H. Wang, “Managing orbital angular momentum in second-harmonic generation,” Phys. Rev. A 88(3), 035801 (2013). [CrossRef]  

49. I. A. Litvin, S. Ngcobo, D. Naidoo, K. Ait-Ameur, and A. Forbes, “Doughnut laser beam as an incoherent superposition of two petal beams,” Opt. Lett. 39(3), 704–707 (2014). [CrossRef]   [PubMed]  

50. R. J. Freiberg and A. S. Halsted, “Properties of low order transverse modes in Argon ion lasers,” Appl. Opt. 8(2), 355–362 (1969). [CrossRef]   [PubMed]  

51. S. M. Baumann, D. M. Kalb, L. H. MacMillan, and E. J. Galvez, “Propagation dynamics of optical vortices due to Gouy phase,” Opt. Express 17(12), 9818–9827 (2009). [CrossRef]   [PubMed]  

52. Y. Ding, M. Xu, Y. Zhao, H. Yu, H. Zhang, Z. Wang, and J. Wang, “Thermally driven continuous-wave and pulsed optical vortex,” Opt. Lett. 39(8), 2366–2369 (2014). [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 (4)

Fig. 1
Fig. 1 Experimental configuration for generation of mode-locked optical vortex pulses. The pump source is a LD with doughnut-shaped output intensity. The output laser pulse performance was recorded by the oscilloscope and the pattern simultaneously observed with a CCD.
Fig. 2
Fig. 2 Performance of the experimental self-mode-locked laser. a-c. Pulse trains observed in the time range of 1.25 ns (250ps/div) with a .LG0,0,s , b. LG0,1,s and c. LG0,2,s modes. d. Pulse trains observed in the time range of 150 ns. e.RF spectrum with frequency of 7.8 GHz. f. Laser spectrum of the mode-locked LG0,2,s pulses at wavelength of 1078.2 nm.
Fig. 3
Fig. 3 Detailed characterization of experimental transverse pattern of mode-locked LGp,l,s beams with topological charges of l. The top images are mode-locked LGp,l,s beams detected by the CCD,the middle images are the phase distributions of LGp,l,s beams separated by a cylindrical lens, the bottom images are the intensity distributions of the LGp,l,s beams fitted by Eq. (1) (a) LG0, 0,s mode without topological charge; (b) LG0, 1,s mode with topological charge of 1; and (c) LG0, 2,s mode with topological charge of 2.
Fig. 4
Fig. 4 Laser output power and staged topological charge of LGp,1,s modes with the increase of absorbed pump power.

Equations (2)

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

ψ ( ρ , φ , z , t ) = e i ω t e i l φ ϕ p , l , s = e i ω t e i l φ 2 p ! π ( p + | l | ) ! 1 w ( z ) ( 2 ρ w ( z ) ) | l | L p | l | ( 2 ρ 2 w ( z ) 2 ) × exp ( ρ 2 w ( z ) 2 ) exp { i k z [ 1 + ρ 2 2 ( z 2 + z R 2 ) ] } × exp [ i ( 2 p + | l | + 1 ) θ G ( z ) ]
ψ i ( ρ , φ , z , t ) = s e i s Δ ω t e i ω 0 t e i l φ ϕ p , l , s
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.