Open Access
18 March 2020 Picosecond optical vortex-induced chiral surface relief in an azo-polymer film
Keigo Masuda, Ryo Shinozaki, Ami Shiraishi, Mitsuki Ichijo, Keisaku Yamane, Katsuhiko Miyamoto, Takashige Omatsu
Author Affiliations +
Abstract

We report on the first demonstration of picosecond optical vortex-induced chiral surface relief in an azo-polymer film due to two-photon absorption isomerization. The chiral surface relief exhibits an extremely narrow defocusing tolerance without undesired outer rings due to the Airy pattern of highly focused light. Such chiral surface relief reflects a z-polarized electric field with an azimuthal helical phase caused by spin–orbital angular momentum coupling.

1.

Introduction

Azo polymers enable photoinduced reversible surface relief to be realized;1 therefore, they will play an important role in the realization of rewritable optical data storage24 and light-driven micro/nanorobots.57 The formation of surface relief in azo polymers generally occurs through several steps: thermodynamically stable trans-form azo polymers are mostly stable at room temperature.8,9 Irradiation of trans-form azo polymers with nonuniform light leads to nonuniform trans-cis photoisomerization of azo polymers via light absorption and subsequent reversion of the azo polymers to the trans-form by spontaneous thermal reaction and the cis-trans reverse photoisomerization.1013 This trans-cis/cis-trans photoisomerization cycle (typical cycle time of 1 to 2 ps) forces a spring-like motion of the azo polymers, which results in spatially anisotropic photofluidity.1417 The mass transport of azo polymers thus occurs from a bright region toward a dark region along the polarization direction of the irradiated light by the gradient force based on interaction between photo-oriented azo polymers and the gradient of an optical electric field, i.e., light-induced mass transport, which results in the establishment of surface relief. The structures of the surface relief are thus mostly determined by the spatial intensity profile and polarization of the irradiated light.1821 Optical vortices carry an annular intensity profile and orbital angular momentum (OAM), characterized by a quantum number , which associated with its helical wavefront.2226 To date, several researchers have reported that irradiation with a continuous-wave visible optical vortex can force the helical mass transport of irradiated azo polymers to form wavefront-sensitive surface relief, i.e., chiral surface relief, facilitated by the spin angular momentum (SAM) of circularly polarized light with a helical electric field.2731

An ultrafast laser-induced two-photon-absorption (TPA) process offers the trans-cis photoisomerization of azo polymers with high spatial resolution, which could yield further advanced technologies, such as high-density rewritable data storage.3234 However, most studies on surface relief formation in azo polymers are still based on single-photon-absorption (SPA).1,821,2711 Surface relief formation via TPA has not yet been well established. Furthermore, there have been no reports on optical vortex-induced surface relief formation in azo polymers via TPA isomerization. Optical vortex-induced TPA will provide not only the helical mass transport of azo polymers for high-density rewritable data storage with a freedom of OAM3537 but also entirely new nonlinear and photochemical phenomena, such as an OAM conservation law in TPA, and spin and OAM coupling in TPA.3840

Here we report on TPA-induced chiral surface relief formation by irradiation with a near-infrared picosecond optical vortex.

2.

Experiments

2.1.

Two-Photon-Absorption Induced Surface Relief

A 1-μm thick azo-polymer (poly-orange tom-1) thin film spin-coated on a slide glass was used in the present experiments, which exhibited strong absorption at 532 nm (no absorption at 1.06  μm), as shown in Fig. 1(a). Therefore, two-photon isomerization should deform the azo-polymer film. A 1.06-μm picosecond laser with a pulse width of 8 ps, a pulse repetition frequency of 35 MHz, and an average power of 40 mW was used as a laser source, and its output was converted into a right-handed circularly polarized optical vortex with a total angular momentum of J=2 by employing a spiral phase plate (OAM: =1) and a quarter wave plate (SAM: s=1). The generated vortex beam was focused to be an annular spot with a diameter of 3.6  μm onto the azo-polymer film surface from the front side, i.e., the atmosphere side, using a high numerical aperture (NA) objective lens (NA=0.9) [Fig. 1(b)]. The corresponding intensity of the focused vortex beam and the exposure time were then 1.4  GW/cm2 and 300 s (dose energy of 12 J), respectively.

Fig. 1

(a) Absorption spectrum of the azo-polymer film used in the present experiment. (b) Experimental-setup. (c) Experimental spatial profile of the strongly focused optical vortex. The focused optical vortex exhibits multiple rings due to Airy pattern.

JNP_14_1_016012_f001.png

The optical vortex caused the upward mass transport of azo polymers along the beam propagation axis (z axis). The azo polymers were subsequently directed inward toward the dark core of the optical vortex and revolved along the clockwise direction to establish single-armed chiral surface relief, which reflects the helical wavefront of the optical vortex field within 120 s (4.8 J) [Fig. 2(a)]. The chiral surface relief formed had a diameter of 3.5  μm and a height of 300  nm. The formation of chiral surface relief on the azo polymers was difficult for an optical vortex with a dose energy of <3  J, whereas the SPA-induced helical mass transport of azo polymers can occur with only a dose energy of 2  mJ, as previously reported.2830 The TPA-induced mass transport of azo polymers41 is enabled by a longitudinal gradient of light intensity of the focused spot only within a small displacement of focused position from the film along z axis, i.e., defocusing depth Δz, as defined in Fig. 1(b). In fact, the formation of chiral surface relief occurred only at the defocusing depth Δz of 2  μm, which indicates that this is TPA-induced photoisomerization [Fig. 2(b)]. The strongly focused optical vortex, in general, exhibits multiple rings owing to the Airy pattern [Fig. 1(c)]. However, the created surface relief is shaped to be a single-armed ring without outer rings that also manifests strongly TPA photoisomerization with significant optical intensity threshold.

Fig. 2

(a) Temporal evolution of surface relief formation in an azo-polymer film by irradiation with 8 ps optical vortex pulses with J=2. (b) Atomic force microscopy (AFM) images of surface relief produced with various defocusing depths. Chiral relief can be formed only at a defocusing depth Δz of 2  μm. The lower panels show line profiles of surface relief along the red lines in the AFM images.

JNP_14_1_016012_f002.png

The optical vortex pulse was compressed to 2  ps without any change of the peak intensity (average power 10  mW). The optical vortex then typically induced a ring-shaped (not well twisted) surface relief, which mostly reflects the spatial intensity profile of the focused optical vortex field [Fig. 3(a)]. A ring-shaped surface relief (although slightly asymmetric) was produced even with an exposure time of 1200 s.

Fig. 3

(a) Surface relief formation in azo-polymer film by irradiation with a 2-ps optical vortex with J=2. The surface relief formed was mostly ring-shaped (not well twisted). (b) Surface relief with two petals was formed by a 2-ps optical vortex. The intensity of the focused optical vortex was measured to be 4.2  GW/cm2.

JNP_14_1_016012_f003.png

It is further noteworthy that the optical vortex with an even higher peak intensity induced photon-bleaching of the azo polymer and ablation of the film itself, which resulted in the formation of an abnormal ring-shaped surface relief with two petals [Fig. 3(b)].

These results further indicate that the helical mass transport, i.e., chiral surface relief formation, requires an optical vortex with a pulse duration of at least several photoisomerization cycles. The results of experiments performed by irradiation with optical vortices with various pulse durations are summarized in Fig. 4.

Fig. 4

TPA-induced surface relief formed by optical vortices with various pulse durations (2, 4, 6, and 8 ps).

JNP_14_1_016012_f004.png

An optical vortex with J=0 was also generated by inversion of the quarter wave plate, which produced only a nonhelical single damp-shaped surface relief, regardless of the pulse duration and exposure time (Fig. 5).

Fig. 5

TPA-induced surface relief formed by optical vortices (J=0) with various pulse durations.

JNP_14_1_016012_f005.png

2.2.

Single-Photon-Absorption Induced Surface Relief

We also investigated SPA-induced surface relief formation by employing a 532-nm optical vortex with a pulse width of 8 ps, a pulse repetition frequency of 35 MHz, and an average power of 3  μW. The optical vortex was focused to be an annular spot with a diameter of 3.6  μm at the focal plane; the corresponding intensity was estimated to be 105  kW/cm2 and the exposure time was 300 s (dose energy: 0.9 mJ). The irradiated azo polymers also exhibited surface relief that was directed in a clockwise direction [Fig. 6(a)]. However, the resultant chiral surface relief had outer rings due to the Airy pattern of the strongly focused optical vortex. The height of the surface relief was typically measured to be 100 nm. Chiral surface relief was further formed at any defocusing depth Δz within the range of 2 to +2  μm [Fig. 6(b)]. These results indicate that TPA-induced photoisomerization enables the formation of chiral surface relief in azo polymers with high longitudinal and transverse spatial resolution [Fig. 6(c)].

Fig. 6

(a) Temporal evolution of surface relief formation in an azo-polymer film by irradiation with 532 nm 8 ps optical vortex pulses with J=2. (b) AFM images of surface relief formed with various defocusing depths. Chiral relief is formed at any defocusing depth Δz within the range of 2 to +2  μm. (c) SPA or TPA induced chiral surface relief. TPA produces chiral relief without any undesired outer rings due to the Airy pattern. (d) Surface relief formed by 532 nm optical vortices with pulse durations of 2 and 8 ps.

JNP_14_1_016012_f006.png

It should be noted that a 2-ps optical vortex did not induce efficient helical mass transport of the azo polymer [Fig. 6(d)]. Therefore, the helical mass transport of the azo polymer, even via SPA, requires an optical vortex with a pulse duration of at least several photoisomerization cycles.

3.

Discussion

The experiments on TPA and SPA suggest that TPA contributes primarily to photoisomerization rather than mass transport of azo polymers. Here we discuss the surface relief formation based on classical optical radiation force based on time-averaged Poynting vector.

The electromagnetic fields of the circularly polarized optical vortex strongly focused by a lens with a high NA (=sinθmax)>0.8 can be analyzed using the theoretical formula as the Debye vector integral:42

Eq. (1)

E(r,φ,z)=[ExEyEz]=ikf2π0θmax02π(2fsinθw)||exp(f2sin2θw2)sinθcosθexp(iϕ)×[(cos2ϕcosθ+sin2ϕ)±icosϕsinϕ(cosθ1)cosϕsinϕ(cosθ1)±i(cos2ϕ+sin2ϕcosθ)sinθexp(±iϕ)]eik[zcosθ+rsinθcos(ϕφ)]dϕdθ,
where Ex, Ey, and Ez are the x-, y-, and z-polarized electric fields in the focused plane; f is the focal distance; w is the beam waist; k is the wave vector; and r, φ, θ, and ϕ are shown in Fig. 7(a). The circularly polarized optical vortex with J=2 produces an annular amplitude profile and a 4π azimuthal phase shift due to constructive spin–orbital coupling effects in an isotropic and homogeneous material, thereby yielding an annular focused spot, which is defined as the sum of the transverse and z-polarized electric fields4346 [Figs. 7(b)7(d)]. The resulting scattering force on the azo-polymer film, given as29

Eq. (2)

Fscattering=iωε0χE×B=ωε02[χrIm(E×B*)+χiRe(E×B*)]
twists and directs the azo polymer toward the dark core with the assistance of the gradient force (Fgradient=dI/dr), so as to form a ring-shaped helical surface relief [Fig. 7(h)].

Fig. 7

(a) The coordinate system for strongly focused optical vortex. Theoretical spatial intensity profiles and phases of (b), (e) transverse (in the focused plane) and (c), (f) z-polarized electric fields of strongly focused circularly polarized optical vortices with J=2 and J=0. (d), (g) Spatial intensity profiles of strongly focused circularly polarized optical vortices with J=2 and J=0. Optical scattering and gradient forces created by strongly focused optical vortices with (h) J=2 and (i) J=0.

JNP_14_1_016012_f007.png

In contrast, an optical vortex with J=0 creates a theoretically focused spot with a central bright spot due to destructive spin–orbital coupling effects [Figs. 7(e)7(g)], thereby yielding a damp-shaped relief via TPA induced photoisomerization [Fig. 7(i)]. These analyses support the observed experimental surface relief. Though it is beyond the scope of this paper, further modeling of optical radiation force, including TPA effects, should be investigated as a future work.

4.

Conclusion

We have successfully demonstrated the formation of TPA-induced chiral surface relief in an azo-polymer film for the first time by employing a picosecond optical vortex source. The TPA-induced photoisomerization allows the formation of chiral surface relief with high longitudinal and transverse spatial resolution, i.e., without undesired outer rings due to Airy patterns and only within an extremely narrow defocusing tolerance. The formation of such surface relief also requires an optical vortex with a pulse duration of at least several trans-cis/cis-trans photoisomerization cycles. An optical vortex with a pulse duration of <6  ps impacted the formation of chiral surface relief with the use of any focusing region and exposure time. The SAM-OAM coupling effects create a strong z-polarized electric field with an azimuthal phase, which contributes significantly to the helical mass transport of the azo polymer.

Such TPA-induced helical mass transport is expected to lead to the realization of high-density optical data storage with the freedom of OAM and offer entirely new physical insights for the interplay between structured light fields and materials.

Acknowledgments

The authors acknowledge the financial support in the form of Kakenhi Grants-in-Aid (Nos. JP16H06507, JP17K19070, and JP18H03884) from the Japan Society for the Promotion of Science (JSPS) and a CREST grant (No. JPMJCR1903) from the Japan Science and Technology Agency (JST).

References

1. 

N. K. Viswanathan et al., “Surface relief structures on azo polymer films,” J. Mater. Chem., 9 (9), 1941 –1955 (1999). https://doi.org/10.1039/a902424g JMACEP 0959-9428 Google Scholar

2. 

G. Xu et al., “Application of all-optical poling in reversible optical storage in azopolymer films,” Opt. Commun., 159 (1–3), 88 –92 (1999). https://doi.org/10.1016/S0030-4018(98)00607-5 OPCOB8 0030-4018 Google Scholar

3. 

W. Zhang et al., “High-efficiency holographic volume index gratings in DR1-dye-doped poly(methyl methacrylate),” Opt. Lett., 27 (13), 1105 –1107 (2002). https://doi.org/10.1364/OL.27.001105 OPLEDP 0146-9592 Google Scholar

4. 

D. Gindre et al., “Rewritable optical data storage in azobenzene copolymers,” Opt. Express, 14 (21), 9896 –9901 (2006). https://doi.org/10.1364/OE.14.009896 OPEXFF 1094-4087 Google Scholar

5. 

Y. Yu, M. Nakano and T. Ikeda, “Directed bending of a polymer film by light,” Nature, 425 145 –145 (2003). https://doi.org/10.1038/425145a Google Scholar

6. 

T. Muraoka et al., “Light-driven open-close motion of chiral molecular scissors,” J. Am. Chem. Soc., 125 (19), 5612 –5613 (2003). https://doi.org/10.1021/ja034994f JACSAT 0002-7863 Google Scholar

7. 

J. J. Wie, M. R. Shankar and T. J. White, “Photomotility of polymers,” Nat. Commun., 7 13260 (2016). https://doi.org/10.1038/ncomms13260 NCAOBW 2041-1723 Google Scholar

8. 

A. Natansohn, P. Rochon, “Reversible optical storage in azo polymers,” Progress in Pacific Polymer Science, 3 295 –305 Springer, Berlin, Heidelberg (1994). Google Scholar

9. 

M. S. Ho et al., “Azo polymers for reversible optical storage. 8. The effect of polarity of the azobenzene groups,” Can. J. Chem., 73 (11), 1773 –1778 (1995). https://doi.org/10.1139/v95-218 CJCHAG 0008-4042 Google Scholar

10. 

H. Rau, “Photoisomerization of benzenes,” Photoreactive Organic Thin Films, 3 –28 1st ed.Academic Press, Cambridge, Massachusetts (2002). Google Scholar

11. 

Z. Sekkat, “Photo-orientation by photoisomerization,” Photoreactive Organic Thin Films, 64 –102 1st ed.Academic Press(2002). Google Scholar

12. 

X. Wang, “Trans–cis isomerization,” Azo Polymers, 19 –56 1st ed.Springer, Berlin, Heidelberg, Germany (2016). Google Scholar

13. 

T. Fujino, S. Y. Arzhantsev and T. Tahara, “Femtosecond time-resolved fluorescence study of photoisomerization of trans-azobenzene,” J. Phys. Chem. A, 105 (35), 8123 –8129 (2001). https://doi.org/10.1021/jp0110713 JPCAFH 1089-5639 Google Scholar

14. 

Z. Sekkat, “Optical tweezing by photomigration,” Appl. Opt., 55 (2), 259 –268 (2016). https://doi.org/10.1364/AO.55.000259 APOPAI 0003-6935 Google Scholar

15. 

H. Ishitobi et al., “The anisotropic nanomovement of azo-polymers,” Opt. Express, 15 (2), 652 –659 (2007). https://doi.org/10.1364/OE.15.000652 OPEXFF 1094-4087 Google Scholar

16. 

P. Karageorgiev et al., “From anisotropic photo-fluidity towards nanomanipulation in the optical near-field,” Nat. Mater., 4 (9), 699 –703 (2005). https://doi.org/10.1038/nmat1459 NMAACR 1476-1122 Google Scholar

17. 

P. Lefin, C. Fiorini and J. M. Nunzi, “Anisotropy of the photoinduced translation diffusion of azo-dyes,” Opt. Mater., 9 (1–4), 323 –328 (1998). https://doi.org/10.1016/S0925-3467(97)00100-6 PAOAE3 0963-9659 Google Scholar

18. 

J. Kumar et al., “Gradient force: the mechanism for surface relief grating formation in azobenzene functionalized polymers,” Appl. Phys. Lett., 72 (17), 2096 –2098 (1998). https://doi.org/10.1063/1.121287 APPLAB 0003-6951 Google Scholar

19. 

C. J. Barrett, A. L. Natansohn and P. L. Rochon, “Mechanism of optically inscribed high-efficiency diffraction gratings in azo polymer films,” J. Phys. Chem., 100 (21), 8836 –8842 (1996). https://doi.org/10.1021/jp953300p JPCHAX 0022-3654 Google Scholar

20. 

P. Rochon, E. Batalla and A. Natansohn, “Optically induced surface gratings on azoaromatic polymer films,” Appl. Phys. Lett., 66 (2), 136 –138 (1995). https://doi.org/10.1063/1.113541 APPLAB 0003-6951 Google Scholar

21. 

A. Kravchenko et al., “Optical interference lithography using azobenzene-functionalized polymers for micro-and nanopatterning of silicon,” Adv. Mater., 23 (36), 4174 –4177 (2011). https://doi.org/10.1002/adma.201101888 ADVMEW 0935-9648 Google Scholar

22. 

L. Allen et al., “Orbital angular momentum of light and the transformation of Laguerre–Gaussian laser modes,” Phys. Rev. A, 45 (11), 8185 –8189 (1992). https://doi.org/10.1103/PhysRevA.45.8185 Google Scholar

23. 

M. Padgett, J. Courtial and L. Allen, “Light’s orbital angular momentum,” Phys. Today, 57 (5), 35 –40 (2004). https://doi.org/10.1063/1.1768672 PHTOAD 0031-9228 Google Scholar

24. 

X. Wang et al., “Recent advances on optical vortex generation,” Nanophotonics, 7 (9), 1533 –1556 (2018). https://doi.org/10.1515/nanoph-2018-0072 Google Scholar

25. 

Y. Shen et al., “Optical vortices 30 years on: OAM manipulation from topological charge to multiple singularities,” Light Sci. Appl., 8 90 (2019). https://doi.org/10.1038/s41377-019-0194-2 Google Scholar

26. 

T. Omatsu et al., “A new twist for materials science: the formation of chiral structures using the angular momentum of light,” Adv. Opt. Mater., 7 (14), 1801672 (2019). https://doi.org/10.1002/adom.201801672 2195-1071 Google Scholar

27. 

A. Ambrosio et al., “Light-induced spiral mass transport in azo-polymer films under vortex-beam illumination,” Nat. Commun., 3 989 (2012). https://doi.org/10.1038/ncomms1996 NCAOBW 2041-1723 Google Scholar

28. 

M. Watabe et al., “Light induced conch-shaped relief in an azo-polymer film,” Sci. Rep., 4 4281 (2015). https://doi.org/10.1038/srep04281 SRCEC3 2045-2322 Google Scholar

29. 

D. Barada et al., “Constructive spin-orbital angular momentum coupling can twist materials to create spiral structures in optical vortex illumination,” Appl. Phys. Lett., 108 (5), 051108 (2016). https://doi.org/10.1063/1.4941023 APPLAB 0003-6951 Google Scholar

30. 

K. Masuda et al., “Azo-polymer film twisted to form a helical surface relief by illumination with a circularly polarized Gaussian beam,” Opt. Express, 25 (11), 12499 –12507 (2017). https://doi.org/10.1364/OE.25.012499 OPEXFF 1094-4087 Google Scholar

31. 

K. Masuda et al., “Nanoscale chiral surface relief of azo- polymers with nearfield OAM light,” Opt. Express, 26 (17), 22197 –22207 (2018). https://doi.org/10.1364/OE.26.022197 OPEXFF 1094-4087 Google Scholar

32. 

S. Maruo, O. Nakamura and S. Kawata, “Three-dimensional microfabrication with two-photon-absorbed photopolymerization,” Opt. Lett., 22 (2), 132 –134 (1997). https://doi.org/10.1364/OL.22.000132 OPLEDP 0146-9592 Google Scholar

33. 

H. B. Sun, S. Kawata, “Two-photon photopolymerization and 3D lithographic microfabrication,” NMR · 3D Analysis · Photopolymerization, 170 169 –273 Springer, Berlin, Heidelberg, Germany (2004). Google Scholar

34. 

Z. Sekkat, “Vectorial motion of matter induced by light fueled molecular machines,” OSA Continuum, 1 (2), 668 –681 (2018). https://doi.org/10.1364/OSAC.1.000668 Google Scholar

35. 

M. Gu et al., “Effect of saturable response to two-photon absorption on the readout signal level of three-dimensional bit optical data storage in a photochromic polymer,” Appl. Phys. Lett., 79 (2), 148 –150 (2001). https://doi.org/10.1063/1.1383999 APPLAB 0003-6951 Google Scholar

36. 

R. J. Voogd, M. Singh and J. J. M. Braat, “The use of orbital angular momentum of light beams for optical data storage,” Proc. SPIE, 5380 387 –392 (2004). https://doi.org/10.1117/12.556939 Google Scholar

37. 

M. Gu, X. Li and Y. Cao, “Optical storage arrays: a perspective for future big data storage,” Light Sci. Appl., 3 e177 (2014). https://doi.org/10.1038/lsa.2014.58 Google Scholar

38. 

K. Y. Bliokh et al., “Spin-orbit interactions of light,” Nat. Photonics, 9 (12), 796 –808 (2015). https://doi.org/10.1038/nphoton.2015.201 NPAHBY 1749-4885 Google Scholar

39. 

S. Franke-Arnold, “Optical angular momentum and atoms,” Philos. Trans. R. Soc. A, 375 20150435 (2017). https://doi.org/10.1098/rsta.2015.0435 PTRMAD 1364-503X Google Scholar

40. 

S. Ishii, N. Yokoshi and H. Ishihara, “Optical selection rule of monolayer transition metal dichalcogenide by an optical vortex,” J. Phys. Conf. Ser., 1220 012056 (2019). https://doi.org/10.1088/1742-6596/1220/1/012056 Google Scholar

41. 

H. Ishitobi et al., “Two-photon induced polymer nanomovement,” Opt. Express, 16 (18), 14106 –14114 (2008). https://doi.org/10.1364/OE.16.014106 OPEXFF 1094-4087 Google Scholar

42. 

Y. Zhao et al., “Spin-to-orbital angular momentum conversion in a strongly focused optical beam,” Phys. Rev. Lett., 99 (7), 15 –18 (2007). https://doi.org/10.1103/PhysRevLett.99.073901 PRLTAO 0031-9007 Google Scholar

43. 

B. Richards and E. Wolf, “Electromagnetic diffraction in optical systems, II. Structure of the image field in an aplanatic system,” Proc. R. Soc. London Ser. A, 253 (1274), 358 –379 (1959). https://doi.org/10.1098/rspa.1959.0200 Google Scholar

44. 

L. Cicchitelli, H. Hora and R. Postle, “Longitudinal field components for laser beams in vacuum,” Phys. Rev. A, 41 (7), 3727 –3732 (1990). https://doi.org/10.1103/PhysRevA.41.3727 Google Scholar

45. 

L. Novotny et al., “Longitudinal field modes probed by single molecules,” Phys. Rev. Lett., 86 (23), 5251 –5254 (2001). https://doi.org/10.1103/PhysRevLett.86.5251 PRLTAO 0031-9007 Google Scholar

46. 

E. Yew and C. Sheppard, “Second harmonic generation polarization microscopy with tightly focused linearly and radially polarized beams,” Opt. Commun., 275 (2), 453 –457 (2007). https://doi.org/10.1016/j.optcom.2007.03.065 OPCOB8 0030-4018 Google Scholar

Biography

Keigo Masuda received his BE and ME degrees from Chiba University in 2016 and 2018, respectively. He is currently studying materials processing and manipulation. He is also a member of OSA, JSAP, and SPIE.

Ryo Shinozaki received his BE degree from Chiba University in 2018. He is currently working on surface relief formation of azo polymers.

Ami Shiraishi is an undergraduate student at Chiba University. She currently joins a project of optical materials processing and manipulation.

Mitsuki Ichijo received her BE degree from Chiba University in 2018. She is currently working on a numerical simulation of optical radiation force based on structured light.

Keisaku Yamane received his PhD from Kyoto University in 2005. He is an associate professor at Hokkaido University. His current research interests include ultrafast structured light and ultrafast physics. He is a member of OSA and JSAP.

Katsuhiko Miyamoto received his PhD from Tohoku Gakuin University in 2004. He is an associate professor at Chiba University. His research interests include terahertz photonics and nonlinear optics. He is a member of OSA, JSAP, and LSJ.

Takashige Omatsu received his PhD from the University of Tokyo in 1992. He was appointed as a professor at Chiba University in 2007. He is currently working on structured light and optical materials processing. He is also serving as an editor-in-chief, OSA continuum, and a steering committee of CLEO-Pacific Rim. He was elected a fellow of OSA, JSAP, and SPIE.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Keigo Masuda, Ryo Shinozaki, Ami Shiraishi, Mitsuki Ichijo, Keisaku Yamane, Katsuhiko Miyamoto, and Takashige Omatsu "Picosecond optical vortex-induced chiral surface relief in an azo-polymer film," Journal of Nanophotonics 14(1), 016012 (18 March 2020). https://doi.org/10.1117/1.JNP.14.016012
Received: 22 November 2019; Accepted: 3 March 2020; Published: 18 March 2020
Lens.org Logo
CITATIONS
Cited by 12 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Optical vortices

Polymers

Picosecond phenomena

Absorption

Spiral phase plates

Atomic force microscopy

Scattering

Back to Top