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

Generating laser transverse modes analogous to quantum Green’s functions of two-dimensional harmonic oscillators

Open Access Open Access

Abstract

We theoretically analyzed the relationship between quantum Green’s functions of two-dimensional harmonic oscillators and radial-order Laguerre–Gaussian laser modes of spherical resonators. By using a nearly hemispherical resonator and a tight focusing in the end-pumped solid-state laser, we successfully generated various laser transverse modes analogous to quantum Green’s functions. We further experimentally and numerically verified that the transverse order associated with quantum Green’s functions is noticeably raised with increasing the pump power induced by the thermal effect. More importantly, the high lasing efficiency and the salient structure enable the present laser source to be used in exploring the light–matter interaction.

© 2017 Chinese Laser Press

1. INTRODUCTION

The study of optical pattern formation has been one of the most active fields of research for decades, since there are some interesting similarities in behavior between optical and hydrodynamics systems [16]. Due to the isomorphism between the paraxial wave equation and the Schrödinger equation [79], the transverse modes of laser spherical cavities can be described with the eigenfunctions of two-dimensional (2D) quantum harmonic oscillators that can be analytically expressed as Hermite–Gaussian (HG) functions with Cartesian symmetry and Laguerre–Gaussian (LG) functions with circular symmetry. The HGn,m modes [10,11] and LGp,l modes [1214] have been directly generated in end-pumped solid-state lasers with the selective pumping [15], where n and m are the order indices in the x and y directions, respectively; p and l are the radial and azimuthal indices, respectively. The Ince–Gaussian modes, another form of eigenfunctions to the paraxial wave equation, have been recently introduced [16] and also experimentally observed in stable resonators [1719]. There are some intriguing features in the spatial structures of high-order transverse modes, such as low divergence of Bessel beams [20], orbital angular momentum of helical beams [21], free acceleration of Airy beams [22], and evolution of a light pulse in a nonlinear laser cavity [23]. Moreover, several interesting issues for generating optical vortices have been explored, such as the formation of vortex lattices in transverse-mode-locked processing [24,25], the phenomenon of Berezinskii–Kosterlitz–Thouless phase transition [2628], and twisted speckle patterns [29]. These characteristics enable high-order transverse structures to be exploited in numerous applications, including optical tweezers and microscopy [30], particle trapping [31], and quantum information [32].

Recently, the connection between the pattern formation of quantum Green’s functions and classical periodic orbits for the 2D harmonic oscillator has been numerically analyzed by plotting the classical trajectories starting from the initial position (x˜s,y˜s) [33]. As seen in Fig. 1, two cases for (x˜s,y˜s)=(0,0) and (x˜s,y˜s)=(0,2.6) display the correspondence between quantum Green’s functions and classical periodic-orbit bundles. Three different total numbers of classical trajectories of 6, 12, and 1000 are plotted in Fig. 1 to reveal the formation of periodic-orbit bundles. It certainly provides important insights into mesoscopic physics to manifest the connection between the spatial distributions of quantum Green’s functions and the classical periodic orbits. According to Hamilton’s opticomechanical theory [34,35], modern laser resonators have been widely used to analogously explore the formation of quantum coherent waves in the mesoscopic regime [3641]. Although a numerical analysis of quantum Green’s functions has been reported, so far there is no experimental observation to demonstrate that the lasing transverse patterns can be related to quantum Green’s functions for the 2D harmonic oscillator.

 figure: Fig. 1.

Fig. 1. Calculated results to display the correspondence between quantum Green’s functions and classical periodic-orbit bundles for two cases: (a) (x˜s,y˜s)=(0,0) and (b) (x˜s,y˜s)=(0,2.6).

Download Full Size | PDF

In this work, we analyze the transverse pattern formation in the spherical cavity with tight excitation by using the inhomogeneous Schrödinger equation for the 2D quantum harmonic oscillator. It is verified that the point-excited resonant modes, corresponding to the quantum Green’s function, are exactly the radial-order LGp,0 modes for the source at the origin. We further develop an end-pumped solid-state laser with a nearly hemispherical resonator to imitate the point-like source and to generate the resonant modes. It is confirmed that the tightly excited resonant modes can be efficiently obtained from low to very high orders just by increasing the pump power due to the thermal effect. It is believed that the tightly excited resonant modes can be used not only to develop novel applications in optical manipulation but also to explore the spatial patterns of quantum Green’s functions in an analogous way.

2. RELATIONSHIP BETWEEN QUANTUM GREEN’S FUNCTIONS AND RADIAL-ORDER LG MODES

Using the 2D isotropic harmonic oscillator to model the transverse part of the spherical cavity, the transverse eigenmodes can be given by [42]

ψn,m(x˜,y˜)=(2n+mn!m!π)1/2e(x˜2+y˜2)/2Hn(x˜)Hm(y˜),
where x˜=2x/wc and y˜=2y/wc are the dimensionless variables for the xy space, Hn(·) is the Hermite polynomials of order n, and wc is the beam radius of the fundamental mode. The corresponding transverse eigenfrequencies fn,m can be normalized as fn,m/fT=vn,m=(n+m+1), where fT is the frequency spacing of the transverse modes. The characteristic equation for ψn,m(x˜,y˜) and vn,m is given by
[12(d2dx˜2+d2dy˜2x˜2y˜2)+vn,m]ψn,m(x˜,y˜)=0.
The inhomogeneous wave equation for the point function can be written as
[12(d2dx˜2+d2dy˜2x˜2y˜2)+v]Gv(r;rs)=δ(rrs),
where r=(x˜,y˜), rs=(x˜s,y˜s), (x˜s,y˜s) is the location of the point source, Gv(r;rs) is the resonant function, and v is the normalized resonant frequency. With the eigenmode expansion to solve Eq. (3), the resonant function Gv(r;rs) can be derived as
Gv(r;rs)=m=0n=0ψn,m(x˜s,y˜s)ψn,m(x˜,y˜)vvn,m.
Note that the resonant function Gv(r;rs) in Eq. (4) is explicitly the quantum Green’s function for the 2D harmonic oscillator. For the resonant condition, the zeros of the denominator in Eq. (4) determine the predominantly contributed eigenmodes. For v=N+1, the function Gv(r;rs) in Eq. (4) can be expressed as a superposition of the degenerate eigenmodes with N=n+m:
GN(r;rs)=n=0Nψn,Nn(x˜s,y˜s)ψn,Nn(x˜,y˜).
Due to the symmetrical property, the location of the point source is set to be (x˜s,y˜s)=(0,ds) for further analysis. The resonant function for general (x˜s,y˜s) can be obtained by using a simple rotational transformation with respect to the y axis. Recall that the formula for the Hermite polynomials is given by [43]
Hn(x)=k=0[n/2](1)kn!k!(n2k)!(2x)n2k,
where [u] denotes the largest integer u. Substituting (x˜s,y˜s)=(0,ds) into Eq. (5) and using Eq. (6), after some algebra, the resonant function in Eq. (5) can be derived as
GN(x˜,y˜;ds)=1π12Nn=0[N/2]AN,n(ds)H2n(x˜)HN2n(y˜)e(x˜2+y˜2)/2,
where
AN,n(ds)=k=0[N/2]n(1)n+k(2ds)N2n2keds2/2n!k!(N2n2k)!.
Note that H2n(0)=(1)n(2n)!/n! and H2n+1(0)=0 have been used in the derivation. It is worthwhile to discuss the case that the point source is at the origin, i.e., ds=0. For odd N, Eq. (8) can be used to show that AN,n(0)=0. As a consequence, GN(x˜,y˜;0) cannot exist for odd N. In contrast, substituting N=2p with p=0,1,2 into Eq. (8) for even N, it can be obtained that A2p,n(0)=(1)p/[n!(pn)!]. Substituting this result into Eq. (7), the resonant function G2p(x˜,y˜;0) can be simplified as
G2p(x˜,y˜;0)=1π(1)p22pn=0pH2n(x˜)H2p2n(y˜)n!(pn)!e(x˜2+y˜2)/2.
From the formula [43]
Lp(x2+y2)=(1)p22pn=0pH2n(x)H2p2n(y)n!(pn)!,
the resonant function G2p(x˜,y˜;0) can be verified to be exactly the radial-order LGp,0 mode. Recently, the high-radial-order LG modes have been used for focal volume reduction [44]. Furthermore, the LGp,0 mode with p can be related to the zero-order Bessel beam J0(kr) by the asymptotic formula [43]
G2p(x˜,y˜;0)=1πer˜2/2Lp(r˜2)p1πJ0(2p+1/2r˜),
where r˜=x˜2+y˜2. To be brief, the present analysis reveals that the point excitation can be used to generate the laser modes associated with the quantum Green’s functions. When the point source is at the origin, the resonant modes are exactly the high-order LGp,0 modes that can be asymptotic to the zero-order Bessel beam in the limit p.

Considering the transverse distribution of the pump source F(r;rs) to deal with a realistically end-pumped configuration in solid-state lasers, the inhomogeneous wave equation for deriving the resonant modes Ψv(r;rs) is given by

[12(d2dx˜2+d2dy˜2x˜2y˜2)+v]Ψv(r;rs)=F(r;rs).
With the eigenmode expansion, the resonant function can be derived as
Ψv(r;rs)=m=0n=0bn,mψn,m(x˜,y˜)vvn,m,
where the coefficient bn,m solved from the orthonormal property of eigenmodes can be expressed as
bn,m=y˜say˜s+ax˜sax˜s+aψn,m(x˜,y˜)F(r;rs)dx˜dy˜,
for the pump-to-mode size ratio a=wp/wc and the pumping profile with a Gaussian distribution
F(r;rs)=(12a)2exp[(x˜x˜s)2+(y˜y˜s)2a2].
Since the coefficient bn,m can be separated into x and y components, the relationship between the pump-to-mode size ratio and the coefficient |bn,m|2 in the y direction can be shown in Fig. 2. The numerical calculations for the location of the excitation source at the origin (i.e., y˜s=0), y˜s=1, and y˜s=2.6 are depicted in Figs. 2(a)2(c), respectively. It can be seem that the coefficient |bn,m|2 is approximately a constant for the pump-to-mode size ratio within the value of 0.3. Here the critical criterion of the pump-to-mode size ratio is validated that the laser transverse modes with a tight pumping can be analogous to quantum Green’s functions of 2D harmonic oscillators with a point source. In the following, we experimentally demonstrate that the tight excitation can be genuinely realized by using an end-pumped solid-state laser with a nearly hemispherical cavity.

 figure: Fig. 2.

Fig. 2. Numerical calculations for the relationship between the pump-to-mode size ratio and the coefficient |bn,m|2 in the y direction for the location of the excitation source at (a) y˜s=0, (b) y˜s=1, and (c) y˜s=2.6.

Download Full Size | PDF

3. EXPERIMENTAL RESULTS AND DISCUSSION

The experimental setup was a concave-flat laser cavity with selectively longitudinal pumping, as shown in Fig. 3. The gain medium was an a-cut 2.0 at. % Nd:YVO4 crystal with a length of 2 mm and an aperture of 10mm×10mm. The large aperture of the gain medium was very critical for generating extremely high-order transverse modes. The Nd:YVO4 crystal was coated on both end surfaces to be antireflective at lasing wavelength (R<0.2% at 1064 nm). The laser crystal was wrapped with indium foil and mounted in a water-cooled copper holder to ensure stable laser output. The front mirror was a 20 mm radius-of-curvature concave mirror with antireflection coating at pumping wavelength (808 nm) on the entrance face and with high-reflectance coating at lasing wavelength (>99.8%) and high-transmittance coating (T>95%) at pumping wavelength on the second surface. The output coupler was a flat mirror with transmission as low as 0.8% at lasing wavelength. The pump source was a 3.0 W 808 nm fiber-coupled laser diode with a core diameter of 100 μm and a numerical aperture of 0.16. A lens with a 25 mm focal length was used to focus the pump beam into the laser crystal. The pump radius wp was approximately 25 μm.

 figure: Fig. 3.

Fig. 3. Experimental setup for a solid-state laser selectively end-pumped by a laser diode in a nearly hemispherical cavity.

Download Full Size | PDF

To achieve the comparable effect of tight pumping, the pump radius wp should be considerably smaller than the cavity mode radius wc in the gain medium. Since the pump radius wp is limited by the brightness of the pump source, enlarging wc is the critical criterion. Here a nearly hemispherical resonator was exploited to obtain a large cavity mode radius for satisfying the criterion of wpwc. For a 20 mm radius-of-curvature concave mirror, it was experimentally found that the tight excitation could be effectively realized when the optical cavity length L was designed to be nearby 19.95 mm. Under the circumstance L=19.95mm, the beam waist at the output coupler can be calculated by using wo=λL(RL)/π [45]. With R=20mm and λ=1064nm, the beam waist wo is approximately 18 μm. With wo=18μm, the mode radius in the gain medium can be calculated by using wc=wo1+(Δz/zR)2, where zR=πwo2/λ and Δz is the distance between the laser crystal and the output coupler. In the experiment, Δz is approximately 7.0 mm with which wc can be found to be approximately 126 μm. Consequently, the ratio wp/wc can be found to be approximately 0.2 smaller than the value of critical criterion 0.3 verified in Fig. 2. In other words, the pump radius wp25μm in the experiment is comparably tight enough that the pump radius wp should be smaller than 38 μm for the cavity mode radius wc126μm to mimic the quantum Green’s function with a point source. For the present hemispherical configuration, the ratio fT/fL is nearly equal to 1/2, where fT and fL are the transverse and longitudinal mode spacings, respectively.

In the end-pumping scheme, the location of the excitation source (Δx,Δy) can be precisely controlled by the manual translation stages, where Δx and Δy represent the distances away from the center of the optical axis in the x and y directions, respectively. The position (Δx,Δy) is related to the theoretical parameter (x˜s,y˜s) by x˜s=2Δx/wc and y˜s=2Δy/wc. To compare with the resonant function GN(x˜,y˜;ds) in Eq. (7), the displacement Δx was fixed to be zero and Δy was adjustable to correspond to the theoretical parameter ds. Figure 4 shows experimental results for the output power and the far-field patterns of lasing modes obtained by varying the pump power Pin for the source at the origin, i.e., ds=0. The far-field patterns were measured by projecting the output beam on a paper screen at a distance of 50 cm away from the laser cavity and using a digital camera to capture the scattered light. It can be seen that the threshold pump power Pth is as low as 50 mW and the output power is up to 0.36 W at a pump power of 2.4 W. It is worthwhile to mention that the order of the lasing mode gradually increases with increasing the pump power. It is clear that all the lasing modes are in good agreement with the spatial patterns |GN(x˜,y˜;0)|, as shown in the bottom of Fig. 4. Based on experimental observation, the relationship between the effective order index N and the pump strength can be empirically expressed as Pin/Pth2N+1. Since GN(x˜,y˜;0) is exactly the LGp,0 mode with N=2p, the maximum radial order for the LGp,0 mode can be found to be up to p=15 in the present experiment. As discussed in Eq. (11), the high-radial-order LGp,0 mode can behave as the zero-order Bessel beam.

 figure: Fig. 4.

Fig. 4. Experimental results for the output power and the lasing modes obtained by varying the pump power Pin for the source at (x˜s,y˜s)=(0,0). Bottom: theoretical patterns |GN(x˜,y˜;0)| for comparison.

Download Full Size | PDF

Quite recently, Dong and collaborators have developed several passively Q-switched Nd:YVO4 microchip lasers to generate various high-order transverse modes by means of adjusting the longitudinally focal position [46], the incident angle [47,48], or the polarization angle [49] of the pump beam. Since the cavity mode sizes in microchip lasers strongly depend on the pump intensity, the transverse orders of lasing modes are asymmetrically expanded within the pump area with increasing the pump power for considering the asymmetrical distribution of the inversion population inside the gain medium [4749]. For the present hemispherical configuration with the critical condition of stable cavities, there exists a critical pump power at which the thermal effect will easily cause the laser cavity to be unstable. The thermal effect leads to a lensing behavior in the gain medium due to the beam profile of a fiber-coupled laser diode approximately described as a Gaussian distribution. Therefore, the thermal effect can be estimated by the overlap integral between the intensity distribution of eigenmodes and the thermal lensing profile:

cn,m=y˜say˜s+ax˜sax˜s+a|ψn,m(x˜,y˜)|2e(x˜x˜s)2+(y˜y˜s)2a2dx˜dy˜.
Figure 5 indicates the calculation result for the coefficient cn,m as a function of the transverse order m of eigenmodes with various pump positions y˜s in the y direction. It is clearly identified that the influence of the thermal effect is relatively more significant on the fundamental mode for the pump source at the origin represented by the red solid curve in Fig. 5. On the other hand, the fundamental mode with the maximum value of the overlap integral is corresponding to the minimum threshold power. For the near-threshold power, the fair thermal lensing causes the fundamental mode with the lowest threshold breaking into oscillation at first. Considering the pump power far above the threshold value, the thermal effect of the fundamental mode rapidly increases, which gives rise to an unstable fundamental mode and is replaced by higher-order modes. With the excitation source away from the origin, the maximum value of the overlap integral will be gradually shifted to higher-order transverse modes represented by the blue dashed curve for y˜s=1 and the green dashed–dotted curve for y˜s=2.6, as shown in Fig. 5. Here the transverse order of lasing modes can be symmetrically enlarged with increasing the pump intensity for the pump size much smaller than the cavity mode size by considering the thermal effect of different transverse order modes.

 figure: Fig. 5.

Fig. 5. Calculation result for the coefficient cn,m as a function of the transverse order m of eigenmodes with various pump positions y˜s in the y direction.

Download Full Size | PDF

Even though the far-field pattern of the lasing mode agrees well with the spatial distribution of a single mode GN(x˜,y˜;0), it is important to explore whether there are other different mode components in the individual lasing mode. To analyze the mode components, a cylindrical-lens mode converter outside the laser resonator was used to transform the lasing mode from the LG basis to the HG basis [50]. The focal length of the cylindrical lenses was f=25mm and the distance was precisely adjusted to be 2f for the operation of the π/2 converter. Figure 6 shows the transformed patterns for the lasing modes in Figs. 4(c) and 4(d). It can be seen that the transformed pattern is not a pure single HGp,p mode. In other words, the lasing mode is not a pure LGp,0 mode but includes some low-order modes. The mode components of the lasing mode are deduced by using the numerical fitting to reconstruct the transformed patterns, as shown in the right-hand side of Fig. 6. For the lasing mode in Fig. 4(c), the mode components are fitted as G2p(x˜,y˜;0)+[(1/4)k=02G2k(x˜,y˜;0)] with p=4. For the lasing mode in Fig. 4(d), the mode components are fitted as G2p(x˜,y˜;0)+[(1/8)k=02G2k(x˜,y˜;0)] with p=7. The low-order LG components in the main LGp,0 mode come from the frequency locking with different longitudinal orders. Since the mode-spacing ratio fT/fL is nearly equal to 1/2, different transverse modes can be frequency locked with the help of different longitudinal orders. On the other hand, the transformed patterns reveal that the lasing modes are mainly dominated by the standing waves.

 figure: Fig. 6.

Fig. 6. Transformed patterns for the lasing modes in Figs. 4(c) and 4(d). Right side: numerically reconstructed patterns.

Download Full Size | PDF

Figure 7 shows experimental results for the output power and the lasing modes obtained by varying the pump power Pin for the excitation source at (x˜s,y˜s)=(0,1), i.e., ds=1. The output power can reach 0.33 W at a pump power of 2.4 W. The overall efficiency is slightly lower than the result obtained for the source at (x˜s,y˜s)=(0,0). It can be clearly seen that all the experimental lasing modes agree very well with the spatial patterns |GN(x˜,y˜;1)| shown in the bottom of Fig. 7 from low to high orders. Figure 8 shows experimental results obtained at (x˜s,y˜s)=(0,2.6). Once again, all the experimental lasing modes are consistent with the theoretical distributions |GN(x˜,y˜;2.6)| shown in the bottom of Fig. 8.

 figure: Fig. 7.

Fig. 7. Experimental results for the output power and the lasing modes obtained by varying the pump power Pin for the source at (x˜s,y˜s)=(0,1). Bottom: theoretical patterns |GN(x˜,y˜;1)| for comparison.

Download Full Size | PDF

 figure: Fig. 8.

Fig. 8. Experimental results for the output power and the lasing modes obtained by varying the pump power Pin for the source at (x˜s,y˜s)=(0,2.6). Bottom: theoretical patterns |GN(x˜,y˜;2.6)| for comparison.

Download Full Size | PDF

4. CONCLUSIONS

In summary, we have theoretically explored the pattern formation of quantum Green’s functions with the point excitation. The point-excited resonant mode has been verified to be exactly the radial-order LGp,0 mode that can be asymptotic to the zero-order Bessel beam in the limit p. In the experiment, an end-pumped solid-state laser with a nearly hemispherical resonator was employed to generate the tightly excited resonant modes from low to very high orders in an efficient way. It is believed that the present finding not only creates an important innovation to generate the structured beams for laser applications but also provides a remarkable method to visualize the quantum–classical correspondence.

Funding

Ministry of Science and Technology, Taiwan (MOST) (106-2628-M-009-001); Japan Society for the Promotion of Science (JSPS) (JP 15H03571, JP 15K13373, JP 16H06507).

REFERENCES

1. K. Staliunas and V. J. Sanchez-Morcillo, Transverse Patterns in Nonlinear Optical Resonators (Springer, 2003).

2. 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, 5090–5113 (1991). [CrossRef]  

3. D. Dangoisse, D. Hennequin, C. Lepers, E. Louvergneaux, and P. Glorieux, “Two-dimensional optical lattices in a CO2 laser,” Phys. Rev. A 46, 5955–5958 (1992). [CrossRef]  

4. E. Louvergneaux, D. Hennequin, D. Dangoisse, and P. Glorieux, “Transverse mode competition in a CO2 laser,” Phys. Rev. A 53, 4435–4438 (1996). [CrossRef]  

5. K. Staliunas, G. Slekys, and C. O. Weiss, “Nonlinear pattern formation in active optical systems: shocks, domains of tilted waves, and cross-roll patterns,” Phys. Rev. Lett. 79, 2658–2661 (1997). [CrossRef]  

6. V. B. Taranenko, K. Staliunas, and C. O. Weiss, “Pattern formation and localized structures in degenerate optical parametric mixing,” Phys. Rev. Lett. 81, 2236–2239 (1998). [CrossRef]  

7. D. Dragoman and M. Dragoman, Quantum-Classical Analogies (Springer, 2004).

8. V. Bužek and T. Quang, “Generalized coherent state for bosonic realization of SU(2) Lie algebra,” J. Opt. Soc. Am. B 6, 2447–2449 (1989). [CrossRef]  

9. G. S. Agarwal and J. Banerji, “Entanglement by linear SU(2) transformations: generation and evolution of quantum vortex states,” J. Phys. A 39, 11503–11519 (2006). [CrossRef]  

10. Y. F. Chen, T. M. Huang, C. F. Kao, C. L. Wang, and S. C. Wang, “Generation of Hermite–Gaussian modes in fiber-coupled laser-diode end-pumped lasers,” IEEE J. Quantum Electron. 33, 1025–1031 (1997). [CrossRef]  

11. H. Laabs and B. Ozygus, “Excitation of Hermite Gaussian modes in end-pumped solid-state lasers via off-axis pumping,” Opt. Laser Technol. 28, 213–214 (1996). [CrossRef]  

12. Y. F. Chen and Y. P. Lan, “Dynamics of the Laguerre Gaussian TEM0,l* mode in a solid-state laser,” Phys. Rev. A 63, 063807 (2001). [CrossRef]  

13. C. J. Flood, G. Giuliani, and H. M. van Driel, “Preferential operation of an end-pumped Nd:YAG laser in high-order Laguerre–Gauss modes,” Opt. Lett. 15, 215–217 (1990). [CrossRef]  

14. S. Ngcobo, K. Aït-Ameur, N. Passilly, A. Hasnaoui, and A. Forbes, “Exciting higher-order radial Laguerre–Gaussian modes in a diode-pumped solid-state laser resonator,” Appl. Opt. 52, 2093–2101 (2013). [CrossRef]  

15. A. Hu, J. Lei, P. Chen, Y. Wang, and S. Li, “Numerical investigation on the generation of high-order Laguerre–Gaussian beams in end-pumped solid-state lasers by introducing loss control,” Appl. Opt. 53, 7845–7853 (2014). [CrossRef]  

16. M. A. Bandres and J. C. Gutiérrez-Vega, “Ince–Gaussian modes of the paraxial wave equation and stable resonators,” J. Opt. Soc. Am. A 21, 873–880 (2004). [CrossRef]  

17. U. T. Schwarz, M. A. Bandres, and J. C. Gutiérrez-Vega, “Observation of Ince–Gaussian modes in stable resonators,” Opt. Lett. 29, 1870–1872 (2004). [CrossRef]  

18. T. Ohtomo, K. Kamikariya, K. Otsuka, and S. Chu, “Single-frequency Ince–Gaussian mode operations of laser-diode-pumped microchip solid-state lasers,” Opt. Express 15, 10705–10717 (2007). [CrossRef]  

19. N. Barré, M. Romanelli, and M. Brunel, “Role of cavity degeneracy for high-order mode excitation in end-pumped solid-state lasers,” Opt. Lett. 39, 1022–1025 (2014). [CrossRef]  

20. J. Durnin, J. J. Miceli, and J. H. Eberly, “Diffraction-free beams,” Phys. Rev. Lett. 58, 1499–1501 (1987). [CrossRef]  

21. L. Allen, M. W. Beijersbergen, R. J. C. Spreeuw, and J. P. Woerdman, “Orbital angular momentum of light and the transformation of Laguerre–Gaussian laser modes,” Phys. Rev. A 45, 8185–8189 (1992). [CrossRef]  

22. G. A. Siviloglou, J. Broky, A. Dogariu, and D. N. Christodoulides, “Observation of accelerating Airy beams,” Phys. Rev. Lett. 99, 213901 (2007). [CrossRef]  

23. A. Yu Okulov and A. N. Oraevsky, “Space-temporal behavior of a light pulse propagating in a nonlinear nondispersive medium,” J. Opt. Soc. Am. B 3, 741–746 (1986). [CrossRef]  

24. Y. F. Chen and Y. P. Lan, “Transverse pattern formation of optical vortices in a microchip laser with a large Fresnel number,” Phys. Rev. A 65, 013802 (2001). [CrossRef]  

25. A. Yu Okulov, “Vortex–antivortex wavefunction of a degenerate quantum gas,” Laser Phys. 19, 1796–1803 (2009). [CrossRef]  

26. V. L. Berezinskii, “Destruction of long-range order in one-dimensional and two-dimensional systems having a continuous symmetry group II. Quantum systems,” Sov. Phys. J. Exp. Theor. Phys. 34, 610–616 (1972).

27. J. M. Kosterlitz and D. J. Thouless, “Ordering, metastability and phase transitions in two-dimensional systems,” J. Phys. C 6, 1181–1203 (1973). [CrossRef]  

28. Z. Hadzibabic, P. Krüger, M. Cheneau, B. Battelier, and J. Dalibard, “Berezinskii–Kosterlitz–Thouless crossover in a trapped atomic gas,” Nature 441, 1118–1121 (2006). [CrossRef]  

29. A. Yu Okulov, “Twisted speckle entities inside wavefront reversal mirrors,” Phys. Rev. A 80, 013837 (2009). [CrossRef]  

30. M. Woerdemann, C. Alpmann, M. Esseling, and C. Denz, “Advanced optical trapping by complex beam shaping,” Laser Photon. Rev. 7, 839–854 (2013). [CrossRef]  

31. V. G. Shvedov, A. V. Rode, Y. V. Izdebskaya, A. S. Desyatnikov, W. Krolikowski, and Y. S. Kivshar, “Giant optical manipulation,” Phys. Rev. Lett. 105, 118103 (2010). [CrossRef]  

32. W. N. Plick, M. Krenn, R. Fickler, S. Ramelow, and A. Zeilinger, “Quantum orbital angular momentum of elliptically symmetric light,” Phys. Rev. A 87, 033806 (2013). [CrossRef]  

33. Y. F. Chen, J. C. Tung, P. H. Tuan, Y. T. Yu, H. C. Liang, and K. F. Huang, “Characterizing classical periodic orbits from quantum Green’s functions in two-dimensional integrable systems: harmonic oscillators and quantum billiards,” Phys. Rev. E 95, 012217 (2017). [CrossRef]  

34. E. Schrödinger, Collected Papers on Wave Mechanics (AMS Chelsea, 1982).

35. P. Holland, The Quantum Theory of Motion (Cambridge University, 1993).

36. Y. F. Chen, Y. P. Lan, and K. F. Huang, “Observation of quantum-classical correspondence from high-order transverse patterns,” Phys. Rev. A 68, 043803 (2003). [CrossRef]  

37. Y. F. Chen, T. H. Lu, K. W. Su, and K. F. Huang, “Devil’s staircase in three-dimensional coherent waves localized on Lissajous parametric surfaces,” Phys. Rev. Lett. 96, 213902 (2006). [CrossRef]  

38. J. Courtois, A. Mohamed, and D. Romanini, “Degenerate astigmatic cavities,” Phys. Rev. A 88, 043844 (2013). [CrossRef]  

39. J. C. Tung, P. H. Tuan, H. C. Liang, K. F. Huang, and Y. F. Chen, “Fractal frequency spectrum in laser resonators and three-dimensional geometric topology of optical coherent waves,” Phys. Rev. A 94, 023811 (2016). [CrossRef]  

40. K. F. Huang, Y. F. Chen, H. C. Lai, and Y. P. Lan, “Observation of the wave function of a quantum billiard from the transverse patterns of vertical cavity surface emitting lasers,” Phys. Rev. Lett. 89, 224102 (2002). [CrossRef]  

41. Y. T. Yu, P. H. Tuan, P. Y. Chiang, H. C. Liang, K. F. Huang, and Y. F. Chen, “Wave pattern and weak localization of chaotic versus scarred modes in stadium-shaped surface-emitting lasers,” Phys. Rev. E 84, 056201 (2011). [CrossRef]  

42. A. Messiah, Quantum Mechanics (Wiley, 1966).

43. N. N. Lebedev, Special Functions & Their Applications (Dover, 1972).

44. A. Hasnaoui, A. Bencheikh, and K. Aït-Ameur, “Tailored TEMp0 beams for large size 3-D laser prototyping,” Opt. Lasers Eng. 49, 248–251 (2011). [CrossRef]  

45. N. Hodgson and H. Weber, Laser Resonators and Beam Propagation (Springer, 2005).

46. J. Dong, S. C. Bai, S. H. Liu, K. I. Ueda, and A. A. Kaminskii, “A high repetition rate passively Q-switched microchip laser for controllable transverse laser modes,” J. Opt. 18, 055205 (2016). [CrossRef]  

47. J. Dong, Y. He, S. C. Bai, K. I. Ueda, and A. A. Kaminskii, “A Cr4+:YAG passively Q-switched Nd:YVO4 microchip laser for controllable high-order Hermite–Gaussian modes,” Laser Phys. 26, 095004 (2016). [CrossRef]  

48. J. Dong, Y. He, X. Zhou, and S. Bai, “Highly efficient, versatile, self-Q-switched, high-repetition-rate microchip laser generating Ince–Gaussian modes for optical trapping,” Quantum Electron. 46, 218–222 (2016). [CrossRef]  

49. H. S. He, M. M. Zhang, J. Dong, and K. I. Ueda, “Linearly polarized pumped passively Q-switched Nd:YVO4 microchip laser for Ince–Gaussian laser modes with controllable orientations,” J. Opt. 18, 055205 (2016). [CrossRef]  

50. M. W. Beijersbergen, L. Allen, H. E. L. O. van der Veen, and J. P. Woerdman, “Astigmatic laser mode converters and transfer of orbital angular momentum,” Opt. Commun. 96, 123–132 (1993). [CrossRef]  

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. Calculated results to display the correspondence between quantum Green’s functions and classical periodic-orbit bundles for two cases: (a) (x˜s,y˜s)=(0,0) and (b) (x˜s,y˜s)=(0,2.6).
Fig. 2.
Fig. 2. Numerical calculations for the relationship between the pump-to-mode size ratio and the coefficient |bn,m|2 in the y direction for the location of the excitation source at (a) y˜s=0, (b) y˜s=1, and (c) y˜s=2.6.
Fig. 3.
Fig. 3. Experimental setup for a solid-state laser selectively end-pumped by a laser diode in a nearly hemispherical cavity.
Fig. 4.
Fig. 4. Experimental results for the output power and the lasing modes obtained by varying the pump power Pin for the source at (x˜s,y˜s)=(0,0). Bottom: theoretical patterns |GN(x˜,y˜;0)| for comparison.
Fig. 5.
Fig. 5. Calculation result for the coefficient cn,m as a function of the transverse order m of eigenmodes with various pump positions y˜s in the y direction.
Fig. 6.
Fig. 6. Transformed patterns for the lasing modes in Figs. 4(c) and 4(d). Right side: numerically reconstructed patterns.
Fig. 7.
Fig. 7. Experimental results for the output power and the lasing modes obtained by varying the pump power Pin for the source at (x˜s,y˜s)=(0,1). Bottom: theoretical patterns |GN(x˜,y˜;1)| for comparison.
Fig. 8.
Fig. 8. Experimental results for the output power and the lasing modes obtained by varying the pump power Pin for the source at (x˜s,y˜s)=(0,2.6). Bottom: theoretical patterns |GN(x˜,y˜;2.6)| for comparison.

Equations (16)

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

ψn,m(x˜,y˜)=(2n+mn!m!π)1/2e(x˜2+y˜2)/2Hn(x˜)Hm(y˜),
[12(d2dx˜2+d2dy˜2x˜2y˜2)+vn,m]ψn,m(x˜,y˜)=0.
[12(d2dx˜2+d2dy˜2x˜2y˜2)+v]Gv(r;rs)=δ(rrs),
Gv(r;rs)=m=0n=0ψn,m(x˜s,y˜s)ψn,m(x˜,y˜)vvn,m.
GN(r;rs)=n=0Nψn,Nn(x˜s,y˜s)ψn,Nn(x˜,y˜).
Hn(x)=k=0[n/2](1)kn!k!(n2k)!(2x)n2k,
GN(x˜,y˜;ds)=1π12Nn=0[N/2]AN,n(ds)H2n(x˜)HN2n(y˜)e(x˜2+y˜2)/2,
AN,n(ds)=k=0[N/2]n(1)n+k(2ds)N2n2keds2/2n!k!(N2n2k)!.
G2p(x˜,y˜;0)=1π(1)p22pn=0pH2n(x˜)H2p2n(y˜)n!(pn)!e(x˜2+y˜2)/2.
Lp(x2+y2)=(1)p22pn=0pH2n(x)H2p2n(y)n!(pn)!,
G2p(x˜,y˜;0)=1πer˜2/2Lp(r˜2)p1πJ0(2p+1/2r˜),
[12(d2dx˜2+d2dy˜2x˜2y˜2)+v]Ψv(r;rs)=F(r;rs).
Ψv(r;rs)=m=0n=0bn,mψn,m(x˜,y˜)vvn,m,
bn,m=y˜say˜s+ax˜sax˜s+aψn,m(x˜,y˜)F(r;rs)dx˜dy˜,
F(r;rs)=(12a)2exp[(x˜x˜s)2+(y˜y˜s)2a2].
cn,m=y˜say˜s+ax˜sax˜s+a|ψn,m(x˜,y˜)|2e(x˜x˜s)2+(y˜y˜s)2a2dx˜dy˜.
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.