Supercontinuum optical vortex pulse generation ... - OSA Publishing

0 downloads 0 Views 609KB Size Report
of the polarizer to transfer polarizations was elucidated. ..... should be emphasized that not a radial half-wave plate for far-infrared radiation (10.6 µm) [33].
Supercontinuum optical vortex pulse generation without spatial or topological-charge dispersion Yu Tokizane, Kazuhiko Oka and Ryuji Morita Department of Applied Physics, Hokkaido University, Kita-13, Nishi-8, Kita-ku, Sapporo 060-8628, Japan, [email protected] [email protected]

Abstract: A new achromatic method to generate the optical vortex was proposed and supercontinuum optical vortex generation ranging ∼500 to ∼800 nm was experimentally demonstrated without spatial nor topological-charge dispersions. In addition, polarization evolution in our system using Jones vectors and matrices was discussed and the condition of the polarizer to transfer polarizations was elucidated. This method is useful for the application to time-resolved nonlinear spectroscopy utilizing ultrabroadband optical vortex pulses in topological materials such as ring-shaped crystals or annular materials. © 2009 Optical Society of America OCIS codes: (050.4865) Optical vortices; (320.6629) Supercontinuum generation; (320.1590) Chirping; (230.5440) Polarization-selective devices

References and links 1. 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). 2. A. Ashkin, “Forces of a single-beam gradient laser trap on a dielectric sphere in the ray optics regime,” Biophys. J. 61, 569–582 (1992). 3. A. D. Mehta, M. Rief, J. A. Spudich, D. A. Smith, and R. M. Simmons, “Single-molecule biomechanics with optical methods,” Science 283, 1689–1695 (1999). 4. K. T. Gahagan and G. A. Swartzlander, Jr., “Optical vortex trapping of particles,” Opt. Lett. 21, 827–829 (1996). 5. H. He, M. E. J. Friese, N. R. Heckenberg, and H. Rubinsztein-Dunlop, “Direct observation of transfer of angular momentum to absorptive particles from a laser beam with a phase singularity”, Phys. Rev. Lett. 75, 826–829 (1995). 6. T. Kuga, Y. Torii, N. Shiokawa, T. Hirano, Y. Shimizu, and H. Sasada, “Novel optical trap of atoms with a doughnut beam,” Phys. Rev. Lett. 78, 4713–4716 (1997). 7. E. M. Wright, J. Arlt, K. Dholakia, K. T. Gahagan, and G. A. Swartzlander, Jr., “Toroidal optical dipole traps for atomic Bose-Einstein condensates using Laguerre-Gaussian beams,” Phys. Rev. A 63, 013608-1–7 (2001). 8. J. Tempere, J. T. Devreese, E. R. I. Abraham, K. T. Gahagan, and G. A. Swartzlander, Jr., “Vortices in BoseEinstein condensates confined in a multiply connected Laguerre-Gaussian optical trap,” Phys. Rev. A 64, 023603-1–8 (2001). 9. N. B. Simpson, K. Dholakia, L. Allen, and M. J. Padgett, “Mechanical equivalence of spin and orbital angular momentum of light: an optical spanner,” Opt. Lett. 22, 52–54 (1997). 10. L. Paterson, M. P. MacDonald, J. Arlt, W. Sibbet, P. E. Bryant, and K. Dholakia, “Controlled rotation of optically trapped microscopic particles,” Science 292, 912–914 (2001). 11. V. Garc´es-Ch´avez, D. M. McGloin, M. J. Padgett, W. Dultz, H. Schmitzer, and K. Dholakia, “Observation of the transfer of the local angular momentum density of a multiringed light beam to an optically trapped particle,” Phys. Rev. Lett. 91, 093602-1–4 (2003). 12. S. W. Hell and J. Wichmann, “Breaking the diffraction resolution limit by stimulated emission: stimulatedemission-depletion fluorescence microscopy”, Opt. Lett. 19, 780–782 (1994).

#112742 - $15.00 USDReceived 12 Jun 2009; revised 29 Jul 2009; accepted 30 Jul 2009; published 3 Aug 2009

(C) 2009 OSA

17 August 2009 / Vol. 17, No. 17 / OPTICS EXPRESS 14517

13. K. I. Willig, S. O. Rizzoli, V. Westphal, S. W. Hell, and R. Jahn, “STED microscopy reveals that synaptotagmin remains clustered after synaptic vesicle exocytosis”, Nature 440, 935–939 (2006). 14. Y. Iketaki, T. Watanabe, N. Bokor, T. Omatsu, T. Hiraga, K. Yamamoto, and M. Fujii, “Measurement of contrast transfer function in super-resolution microscopy using Two-color Fluorescence Dip Spectroscopyh, Appl. Spectroscopy 61 6–10 (2007). 15. A. Mair, A. Vaziri, G. Weihs, and A. Zeilinger, “Entanglement of the orbital angular momentum states of photons,” Nature 412, 313–316 (2001). 16. A. Vaziri, G. Weihs, and A. Zeilinger, “Experimental two-photon, three-dimensional entanglement for quantum communication,” Phys. Rev. Lett. 89, 240401-1–4 (2002). 17. G. Molina-Terriza, J. P. Torres, and L. Torner, “Management of the angular momentum of light: preparation of photons in multidimensional vector states of angular momentum,” Phys. Rev. Lett. 88, 013601-1–4 (2001). 18. J. Hamazaki, Y. Mineta, K. Oka, and R. Morita, “Direct observation of Gouy phase shift in a propagating optical vortex”, Opt. Express 14 8382–8392 (2006). 19. A. I. Yakimenko, Y. A.Zaliznyak, and Y. Kivshar, “Stable vortex soliton in nonlocal self-focusing nonlinear media”, Phys. Rev. E 71 065603(R)-1–4 (2005). 20. 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 133901-1–4 (2006). 21. S. Tanda, T. Tsuneta, Y. Okajima, K. Inagaki, K. Yamaya, and N. Hatakenaka, “A M¨obius strip of single crystals”, Nature 417, 397–398 (2002). 22. K. Shimatake, Y. Toda, and S. Tanda, “Quenching of phase coherence in quasi-one-dimensional ring crystals”, Phys. Rev. B 73, 153403-1–4 (2006). 23. Y. Aharonov and D. Bohm, “Significance of electromagnetic potentials in the quantum theory”, Phys. Rev. 115, 485–491 (1959). 24. M. W. Beijersbergen, R. P. C. Coeerwinkel, M. Kristensen, and J. P. Woerdman, “Helical-wavefront laser beams produced with a spiral phaseplate”, Opt. Commun. 112, 321–327 (1994). 25. V. Y. Bazhenov, M. V. Vasnetsov, and M. S. Soskin, “Laser beams with screw dislocations in their wavefronts”, JETP Lett. 52, 1037–1039 (1990). 26. K. J. Moh, X.-C. Yuan, D. Y. Tang, W. C. Cheong, L. S. Zhang, D. K. Y. Low, X. Peng, H. B. Niu, and Z. Y. Lin, “Generation of femtosecond optical vortices using a single refractive optical element”, Appl. Phys. Lett. 88, 091103-1–3 (2006). 27. K. Bezuhanov, A. Dreischuh, G. G. Paulus, M. G. Sch¨atzel, and H. Walter, “Vortices in femtosecond laser fields”, Opt. Lett. 29, 1942–1944 (2004). 28. K. Bezuhanov, A. Dreischuh, G. G. Paulus, M. G. Sch¨atzel, H. Walter, D. Neshev, W. Kr´olikowski, and Y. Kivshar, “Spatial phase dislocation in femtosecond laser pulses”, J. Opt. Soc. Am. B 23, 26–35 (2006). 29. I. Zeylikovich, H. I. Sztul, V. Kartazaev, T. Le, and R. R. Alfano, “Ultrashort Laguerre-Gaussian pulses with angular and group velocity dispersion compensation”, Opt. Lett. 32, 2025–2027 (2007). 30. I. G. Mariyenko, J. Strohaber, and C. J. G. J. Uiterwaal, “Creation of optical vortices in femtosecond pulses”, Opt. Express 13, 7599–7608 (2005). 31. A. Schwarz and W. Rudolph, “Dispersion-compensating beam shaper for femtosecond optical vortex beams”, Opt. Lett. 33, 2970–2972 (2008). 32. J. Leach and M. J. Padgett, “Observation of chromatic effects near a white-light vortex”, New J. Phys. 5, 154-1–7 (2003). 33. A. Niv, G. Biener, V. Kleiner, and E. Hasman, “Polychromatic vectorial vortex formed by geometric phase elements”, Opt. Lett. 32, 847–849 (2002). 34. A. J. Wright, J. M. Girkin, G. M. Gibson, J. Leach, and M. J. Padgett, “Transfer of orbital angular momentum from a super-continuum, white-light beam”, Opt. Express 16, 9495–9500 (2008). 35. G. A. Swartzlander, Jr. “Achromatic optical vortex lens”, Opt. Lett. 31, 2042–2044 (2006). 36. R. C. Jones, “A new calculus for the treatment of optical systems”, J. Opt. Soc. Am. 31, 488–493 (1942). 37. K. Yamane, Z. Zhang, K. Oka, R. Morita, M. Yamashita, and A. Suguro, “Optical pulse compression to 3.4 fs in the monocycle region by feedback phase compensation”, Opt. Lett. 28, 2258–2260 (2003). 38. M. Yamashita, K. Yamane, and R. Morita, “Quasi-automatic phase-control technique for chirp compensation of pulses with over-one-octave bandwidth—generation of few- to mono-cycle optical pulses”, IEEE J. Sel. Top. Quantum Electron. 12, 213–222 (2006).

1.

Introduction

A Laguerre-Gaussian (LG) mode is one of the modes of paraxial solutions to wave equation and has a helical wavefront. The light beam with helical wavefront is called an optical vortex. The optical vortex or LG mode has unique properties. The beam has a phase singularity on its center, which shows a dark part on its center of intensity profile, and carries an orbital angular momen-

#112742 - $15.00 USDReceived 12 Jun 2009; revised 29 Jul 2009; accepted 30 Jul 2009; published 3 Aug 2009

(C) 2009 OSA

17 August 2009 / Vol. 17, No. 17 / OPTICS EXPRESS 14518

tum of light well defined by the topological charge [1]. These characteristics recently attracted much attention because of their increasing applications in many fields, such as the optical trapping [2, 3, 4, 5] (especially, trapping for Bose-Einstein condensates [6, 7, 8]), microstructure rotation in laser tweezers and spanners [9, 10, 11], super-resolution microscopy [12, 13, 14], quantum information using multidimensional entangled states [15, 16, 17], geometrical (or Berry’s) phase observation [18] and nonlinear spatial-vortex propagation [19, 20]. However, they have not been fully utilized so far for optical spectroscopy, in particular, ultrafast spectroscopy for topological materials such as a ring-shaped crystal or annular-shaped materials [21, 22], in which closed-loop coherence or Aharonov-Bohm effect [23] is to be investigated by them. It has been due to the fact that generating methods of ultrashort optical vortex pulse is not well established from the view point of spatial or topological charge-chirp free techniques. Thus, for making good use of ultrashort optical vortex pulses, it is a key technique for managing spatialand/or topological charge-chirps as well as a conventional frequency chirp. There are two typical methods to generate the optical vortex from the Hermite-Gaussian beams. One uses spiral phase plates [24]; the other employs holograms [25] generated by spatial phase modulators. These two methods are not suitable for generating ultrashort [26, 27, 28, 29, 30, 31] or ultrabroadband vortex pulses [32, 33, 34]. Usually spiral plates are designed for a certain wavelength [26], hence, they induce topological-charge dispersion for ultrashort or ultrabroadband pulses, leading mixture of eigenstates of a vortex. An achromatic lens consisted of two materials was demonstrated as a modified spiral plate, the bandwidth was, however, limited to ∼100 nm [35]. Computer generated holograms composing of diffraction and phase singularity patterns, which are often used for separation of fundamental and transferred vortex beam, bring spatial dispersion without beam position coincidence [30, 32, 34]. Although this spatial chirp can be eliminated by 4 f [27, 28, 29], 2 f -2 f [30] or prism configuration [31, 34], it is comparably complicated or rather bandwidth-limited. In the present paper, we propose a new achromatic method without spatial- nor topological charge-dispersions for generation of ultrabroadband optical vortex pulses. In addition, we experimentally demonstrate our new achromatic method to generate a supercontinuum optical vortex pulse with an almost octave-spanning bandwidth. Moreover, we discuss polarization evolution in our system using Jones vectors and matrices. 2.

Principle

We explain the principle of our new achromatic method in this section. Figure 1 shows a conceptual scheme of the experiment by our system. It consists of a polarizer (P), an achromatic quarter-wave plate (AQWP1), an axially symmetric polarizer (ASP), another achromatic quarter-wave plate2 (AQWP2), and an analyzer (A). It converts the incident Gaussian beam into the optical vortex or the LG beam with a certain topological charge, which is controllable with the design of the axially symmetric polarizer (ASP). In our case, ASP is designed for generation of an optical vortex with a topological charge ` = ±2, having a bandwidth of ∼500–∼850 nm. The AQWPs’ bandwidths range from ∼400 to ∼800 nm. Our proposal is similar to ref. [33] in terms of utilizing polarization singularity for generation of an optical vortex. However, it should be emphasized that not a radial half-wave plate for far-infrared radiation (10.6 µ m) [33] but ASP and AQWP’s are employed here, enabling us to obtain a supercontinuum optical vortex pulse with an almost octave-spanning bandwidth in the visible and near-infrared region. In our proposed system, the polarization transformation is a key technique. Here, we explain the polarization evolution of the beam in our system, using Jones vectors and matrices [36]. An incident optical · beam ¸ with a spatial Gaussian profile is set to be linearly-polarized with the 1 Jones vector of , by the horizontal linear polarizer P whose Jones matrix is expressed 0

#112742 - $15.00 USDReceived 12 Jun 2009; revised 29 Jul 2009; accepted 30 Jul 2009; published 3 Aug 2009

(C) 2009 OSA

17 August 2009 / Vol. 17, No. 17 / OPTICS EXPRESS 14519

P AQWP1 ASP AQWP2 A

Fig. 1. Schematic drawing of ultrabroadband optical vortex generation system without spatial or topological-charge dispersions. Polarization distribution of the beam are shown after passing through optical components.

Table 1. Top row: Optical components used in our technique, second row: Jones matrix of each component, and bottom row: Jones vector after passing through the component in the experimental setup. component

P ·

matrix

1 0 ·

vector

·

AQWP1

0 0

1 0

¸

· √1 2

¸

1 i ·

√1 2

i 1

1 i

ASP ¸

·

cos2 φ 1 2 sin 2φ ·

¸ √1 eiφ 2

1 2

AQWP2

sin 2φ sin2 φ

cos φ sin φ

¸

· √1 2

¸

· 1 2

1 −i

A

−i 1

1 −iei2φ

¸

·

0 0

·

¸ − 2i ei2φ

¸

0 1

0 1

¸

¸ . It is then converted into left (counterclockwise) circularly-polarized beam giv· ¸ · ¸ 1 1 i 1 1 √ √ ing the vector 2 , by AQWP1 whose matrix is 2 . This beam is entered into i i 1 the axially symmetric polarizer ASP. It transmits ¸ only the radially-polarized component and · cos2 φ 21 sin 2φ , where φ is the azimuthal angle in the beam its matrix is yielded by 1 sin2 φ 2 sin 2φ cross section. This polarizer is made of a photonic crystal and purchased from Photonic Lattice Inc. After passing · ¸ through ASP, the beam is turned to be radially polarized with the vector cos φ 1 i φ of √2 e . The polarization is locally linear. However, the polarization directions desin φ pends on the azimuthal coordinates φ . It should be noted that not only their amplitude but their phase depends on φ in the form of exp(iφ ). It is essential for generating an optical vortex. The obtained radially polarized beam is guided into the second achromatic quarter-wave plate AQWP2, whose fast axis is perpendicular to that of AQWP1. The Jones matrix of AQWP2 is by

1 0 0 0

#112742 - $15.00 USDReceived 12 Jun 2009; revised 29 Jul 2009; accepted 30 Jul 2009; published 3 Aug 2009

(C) 2009 OSA

17 August 2009 / Vol. 17, No. 17 / OPTICS EXPRESS 14520

Interference system Delay stage

CCD

L1

L2

AQWP1 AQWP2 ASP P A

Filter Q

Ti:Sapphire laser + regenerative amplifer Sapphire crystal

Achromatic vortex generation system

Fig. 2. Experimental setup for ultrabroadband optical vortex generation without spatial or topological-charge dispersions.

·

¸ described by . The polarizations of the passed beam depend on the azimuthal · ¸ 1 . Finally, only the vertical component is extracted angle φ , giving the Jones vector 12 −iei2φ · ¸ 0 0 by the vertical linear analyzer A with the matrix of . Consequently, the Jones vector 0 1 · ¸ 0 . Its polarization is linearly vertical. However, unlike of the beam comes to be − 2i ei2φ 1 usual uniform linear polarization, its phase depends on φ as indicated by the factor exp(i2φ ), expressing an `=2 optical vortex. Thus we obtain the optical vortex beam. For `=−2 optical vortex generation with another sign, AQWP’s should be interchanged. In addition, other designs of ASP supply other `-value vortices. The polarizations represented by Jones vectors at points after passing through the components are summarized in Table 1. It should be emphasized that Jones matrices of components used in our system are independent of wavelength λ as well as the radial coordinate ρ in the beam cross section, resulting in capability of spatial and topological-charge free optical vortex generation even with an ultrabroadband pulse. √1 2

3.

1 −i

−i 1

Experimental setup and results

In this section, we describe experimental demonstration of ultrabroadband optical vortex pulse generation without spatial nor topological-charge dispersions. Figure 2 shows our experimental setup. The light source that we used was a Ti:sapphire laser regenerative amplifier with a center wavelength of 795 nm and a repetition rate of 1 kHz. The pulse from the regenerative amplifier was focused into a 3.5 mm-thick sapphire crystal by a plano-convex lens L1 with a focal length of 150 mm for supercontinuum or white light continuum generation. The generated white light pulse, spectrally ranging from ∼450 to ∼900 nm with a full-width at one-thousandth maximum as depicted in Fig. 3, was roughly collimated by a plano-convex lens L2 with a focal length of 70 mm. The supercontinuum pulse with a Gaussain spatial profile was guided into our achromatic vortex generation system. After passing through the achromatic vortex generation system, a supercontinuum vortex

#112742 - $15.00 USDReceived 12 Jun 2009; revised 29 Jul 2009; accepted 30 Jul 2009; published 3 Aug 2009

(C) 2009 OSA

17 August 2009 / Vol. 17, No. 17 / OPTICS EXPRESS 14521

Intensity (arb. units)

0.005

0.001

300

400

500 600 700 800 Wavelength (nm)

900

1000

Fig. 3. Spectral intensity of a generated supercontinuum, normalized by the maximum intensity at 799 nm. The spectrum ranges from ∼450 to ∼900 nm and its full-width at onethousandth maximum is ∼450 nm.

Fig. 4. Spectrally-resolved vortex pulses with (a) a center wavelength λ0 =800 nm and a bandwidth ∆ λ =11 nm, (b) λ0 =680 nm and ∆ λ =11 nm, and (c) λ0 =500 nm and ∆ λ =65 nm from a generated supercontinuum. Line profiles show horizontal (x-direction) and vertical (y-direction) intensity along the lines including the beam center. (d) superposition of intensity profiles and line profiles of (a)-(c).

pulse was obtained. It was spectrally filtered by bandpass filters, and its beam profile was monitored at the point Q by a charge-coupled device (CCD) or directed into the Mach-Zehnder-type interference system to investigate topological charge. By changing bandpass filters, the supercontinuum vortex pulse was spectrally resolved. First, we investigate position of the dark part of the spectral-resolved beams to examine the spatial dispersion. Beam intensity profiles and line profiles at the point Q are shown in Fig. 4 for the spectrally-resolved vortex pulses with (a) a center wavelength λ0 =800 nm and a bandwidth ∆ λ =11 nm, (b) λ0 =680 nm and ∆ λ =11 nm, and (c) λ0 =500 nm and ∆ λ =65 nm from a generated supercontinuum. Line profiles show horizontal and vertical intensities along the lines including the beam center. They all have dark spots on the center, reflecting a character of optical vortices. Figure 4(d) shows the superposition of their beam intensity profiles. Their dark spots well coincide at the center position with one another, giving no spatial dispersion, at least in the wavelength range of 500-800 nm. Second, we examine interference patterns between spectrally-resolved vortex pulses. The vortex pulse propagating along the one arm in our Mach-Zehnder-type interferometer is reflected three times, while the vortex pulse is reflected six times along the other arm (threereflection retroreflector RR was used for the delay line). Hence, interference patterns between vortex pulses with the same |`| value but different signs were observed, since a reflection changes the sign of topological charge. Observed interference patterns for (a) λ0 =800 nm and

#112742 - $15.00 USDReceived 12 Jun 2009; revised 29 Jul 2009; accepted 30 Jul 2009; published 3 Aug 2009

(C) 2009 OSA

17 August 2009 / Vol. 17, No. 17 / OPTICS EXPRESS 14522

(a)

(b)

(c)

Fig. 5. Observed interference patterns for (a) λ0 =800 nm and ∆ λ =11 nm, (b) λ0 =680 nm and ∆ λ =11 nm, and (c) λ0 =500 nm and ∆ λ =65 nm, which are spectrally-resolved from a generated supercontinuum. They all yield clear four-pronged fork patterns, indicating that spectrally-resolved vortex pulses before entering the interferometer definitely possess the same topological charge of `=2 as designed.

∆ λ =11 nm, (b) λ0 =680 nm and ∆ λ =11 nm, and (c) λ0 =500 nm and ∆ λ =65 nm, which are spectrally-resolved from a generated supecontinuum, are depicted in Fig. 5(a), (b) and (c), respectively. They all yield clear four-pronged fork patterns, indicating that spectrally-resolved vortex pulses before entering the interferometer definitely possess the same topological charge of `=2 as designed. These results show generation of topological charge-free optical vortex pulse with a broadband width, at least in the wavelength range of 500-800 nm. The conversion efficiency from spatially-Gaussian to vortex with an ultrabroadband was evaluated to be ∼18%, which was comparable with the maximum efficiency in principle described in the discussion section. Somewhat of discrepancy between the experimental and theoretical conversion efficiencies is attributed to the imperfectness of polarization transformation of ASP and AQWP’s for ultrabroadband pulses as well as their transmittivities (∼90 % for ASP and ∼99 % for AQWP’s). Although not fully investigated, the maximum input fluence of the supercontinuum pulse available to our achromatic vortex generation system is limited to several µ J/cm2 mainly by the damage threshold of ASP. As mentioned above, our new achromatic method enables us to generate supercontinuum optical vortices without spatial-chirp nor topological-charge dispersions. In this experiment, although active compensation of frequency chirp was not carried out, a Fourier-transform limited pulse can be obtained by using a 4 f -system with a spatial phase modulator [37, 38], even for a pulse with an over-octave-spanning bandwidth. 4.

Discussion

We discuss our system from the view point of circular polarization decomposition. Arbi˜ trary polarization · ¸ E is decomposed into a linear combination · ¸ of left circular polarization 1 1 E˜ L = √12 and right circular polarization E˜ R = √12 , as i −i ˜ = (E˜ L , E) ˜ E ˜ L + (E˜ R , E) ˜ E˜ R . E

(1)

Here the bracket (a, b) represents the inner product as (a, b) = a† b, where a† is the adjoint vector of a (transposed complex conjugate vector of a). For the set of P and AQWP1, its joint · ¸ 1 0 ˜ R = √1 E˜ L . Thus, F1 converts arbitrary Jones matrix F1 is F1 = √12 , and F1 E˜ L = F1 E 2 i 0 √ ˜ polarizations to the left circular polarization EL with amplitude reduction of factor · ¸ 1/ · 2. Sim¸ 0T 0 0 ilarly, for the set of AQWP2 and A, its Jones matrix F2 is F2 = √12 = −i ˜ † , −i 1 ER

#112742 - $15.00 USDReceived 12 Jun 2009; revised 29 Jul 2009; accepted 30 Jul 2009; published 3 Aug 2009

(C) 2009 OSA

17 August 2009 / Vol. 17, No. 17 / OPTICS EXPRESS 14523

˜ rad can be decomposed into the superposition of constant amFig. 6. Radial polarization E ˜ L and right polarization E ˜ R . Only the right polarization plitude left circular polarization E component has azimuthal angle ϕ -dependent phase, while the left circular polarization component has the uniform phase. In this case, a linear polarized optical vortex with topological charge ` = 2 can be generated.

· where 0T is the transposed vector of 0 =

0 0

¸ . Since ·

˜ = −i(E˜ R , E) ˜ F2 E

0 1

¸ ,

˜ as a horizontal polarization. F2 extracts the E˜ R -component of arbitrary polarization E, ˜ Radial polarization Erad just after passing through ASP can be decomposed into · ¸ 1 iφ cos φ 1˜ 1 i2φ ˜ ˜ Erad = √ e = E L + e ER . sin φ 2 2 2

(2)

(3)

This indicates that E˜ rad is the superposition of constant-amplitude left circular polarization E˜ L ˜ R , as shown in Fig. 6. It should be noted here that only the and right circular polarization E right circular polarization component has azimuthal angle φ -dependent phase 2φ while the left circular polarization component has the uniform phase. Hence, it enables us to generate a linear polarized optical vortex with topological · ¸ charge ` = 2. Regarding the conversion efficiency · ¸ 1 0 from the initial linear polarization to final linear polarized optical vortex − 2i ei2φ , 0 1 the amplitude is reduced by a factor 1/2. Thus, maximum conversion efficiency in intensity in our system is 25 %. Next, we discuss the necessary and sufficient conditions of the optical component at the position of ASP in our system to generate an ultrabroadband optical vortex pulse without spatial nor topological-charge dispersions. We put the Jones matrix of the component as · ¸ P11 P12 (4) P= P21 P22 ˜ in = and the input constant polarization to the achromatic vortex pulse generation system as E · ¸ Ein,1 ˜ out from the system is expressed by . Hence, the output polarization E Ein,2 · ¸ 1 0 , (5) E˜ out = F2 R(θ )PR(−θ )F1 E˜ in = ei2θ {P12 + P21 − i(P11 − P22 )}Ein,1 2 · ¸ cos θ − sin θ where the rotation matrix R(θ ) = is used and it is assumed that the optical cos θ sin θ component is rotated by angle θ . Therefore, necessary and sufficient conditions of the component for generation an ultrabroadband optical vortex pulse without spatial nor topologicalcharge dispersions are:

#112742 - $15.00 USDReceived 12 Jun 2009; revised 29 Jul 2009; accepted 30 Jul 2009; published 3 Aug 2009

(C) 2009 OSA

17 August 2009 / Vol. 17, No. 17 / OPTICS EXPRESS 14524

1) rotation angle θ is nφ /2 (n = ±1, ±2, ±3, · · · ) 2) “P12 + P21 is nonzero and wavelength-independent” or “P11 − P22 is nonzero and wavelength-independent”. The ASP in our system is one of the simplest examples (n = ±2 and P11 = 1, P12 = P21 = P22 = 0). However, it should be noted that, only in even n cases, the component is axially symmetric (n/2-fold axially symmetry). 5.

Conclusion

We proposed a new achromatic method to generate the optical vortex without spatial- nor topological charge-dispersions and demonstrated the dispersion free supercontinuum optical vortex in the simple setup. Our experiment showed that our method generate the supercontinuum optical vortex without spatial nor topological charge dispersion (` = ±2) with an almost octave-spanning bandwidth in the wavelength of 500 to 800 nm. In addition, we discuss polarization evolution in our system using Jones vectors and matrices, clarifying the condition of the polarizer to transfer polarizations. Our method is useful and powerful for the application to time-resolved nonlinear spectroscopy employing ultrabroadband optical vortex pulses in topological materials such as ring-shaped crystals and annular-shaped materials, in order to investigate closed loop coherence. Acknowledgments The authors would like to thank S. Tanda and Y. Toda for their useful discussion and encouragement. This work was partially supported by Grant-in-Aid for the 21st Century COE program on “Topological Science and Technology” from the Ministry of Education, Culture, Sports, Science and Technology of Japan, and Grant-in-Aid for Scientific Research (B), 2008-2010, No. 20360025 from Japan Society for the Promotion of Science (JSPS).

#112742 - $15.00 USDReceived 12 Jun 2009; revised 29 Jul 2009; accepted 30 Jul 2009; published 3 Aug 2009

(C) 2009 OSA

17 August 2009 / Vol. 17, No. 17 / OPTICS EXPRESS 14525