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

Generation of arbitrary axisymmetrically polarized pulses by using the combination of 4-f spatial light modulator and common-path optical system

Open Access Open Access

Abstract

We proposed and constructed a system to realize broadband generation of arbitrary axisymmetrically polarized (AP) pulses with spatial complex amplitude modulation. This system employs the combination of a spatial light modulator in the 4-f configuration (4-f SLM), and a space variant wave plate as a common path interferometer. The 4-f SLM and the common path interferometer offer compensation for spatial dispersion with respect to wavelength and stability to perturbation, respectively. We experimentally demonstrated the various AP pulses generation by applying modulations of fundamental and higher-order Laguerre-Gauss modes, whose radial indices were, respectively, p = 0 and 1, with high purity, which showed that we were able to generate arbitral AP pulses with spatial complex amplitude modulation. This technique is expected to be applied in both classical and quantum communications with higher-order modes.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

An axisymmetrically polarized (AP) mode, a kind of the cylindrically polarized modes or the vector vortex modes, has the symmetry of C in both of the polarization and the intensity distributions [1]. The high symmetry of the AP modes has attracted attention for the various research topics such as super-resolution microscopy [2], particle acceleration [3–6], laser processing [7–14], and optical tractor [15].

Recently, some applications using ultrashort or broadband AP pulses have been reported such as quantum communication [16, 17], generation of extreme ultraviolet AP beams [18], and nonlinear spectroscopy [19–22]. In order to fully explore the potential of the AP modes, it is important to develop a generation system of AP pulses with an ultrashort duration or a broadband spectrum, while there are fewer researches on the generation of AP pulses in comparison with the generation of AP continuous wave beams. We have reported on the generation of AP broadband pulses by using the combination of a spatial light modulator (SLM) in the 4-f configuration and a coherent combining system, and showed that the arbitrary AP states can be generated by superposing the left and right circularly polarized (LCP and RCP, respectively) optical vortex (OV) broadband pulses [23]. Thanks to the 4-f configuration SLM (4-f SLM) [24, 25], this method enables us to modulate the complex amplitude distribution in the radial axis of AP broadband pulse beams. However, a separate-path interferometer in the coherent combining system, being sensitive to disturbance (e.g. vibration and air turbulence), causes a practical issue.

One way to solve this issue is the use of common-path interferometers, which are robust over disturbance compared with our previous method. The generation of arbitrary AP beams using common-path interferometers has been demonstrated by utilizing a space variant wave plate (SVWP) or a q-plate [26], SLMs [27–29] or anisotropic crystal beam displacers [30]. In all cases, the generation methods need a SLM [27–29] including a digital micromirror device (DMD) [30] for complex amplitude modulation, preventing from generation of a point vortex [26] or a hypergeometric-Gaussian state [31]. However, the configurations of such devices do not compensate for the spatial chromatic dispersion caused by grating hologram patterns displayed on the DMD or the SLM. These optical systems are, therefore, unsuited for broadband pulses.

In the present paper, we generate arbitrary AP pulses with spatial complex amplitude modulation by the system composed of a 4-f SLM and a SVWP. The benefit of using the SVWP is robust against perturbations, but the complex amplitude modulation in the radial direction cannot be done with the SVWP. The use of 4-f SLM, which is suitable only for linear polarized states, can perform the complex amplitude modulation with compensation of chromatic dispersions. In contrast, the combination of the 4-f SLM and SVWP gives the complex amplitude modulation in all the directions without chromatic dispersions for not only linear polarization states but AP states. Moreover, we evaluate the properties of the AP pulses qualitatively and quantitatively through interference fringe pattern and extended Stokes parameters (ESPs) [23, 32–34], respectively. The present paper is organized as follows. We propose the system capable for the generation of arbitrary AP broadband pulses with spatial complex amplitude modulation. After that, we show the experimental results of the generation of p = 0 and p = 1 Laguerre-Gaussian (LG) AP pulses. Here, p denotes the radial index of LG modes [35]. The AP state and the spatial phase distribution are characterized by using the Stokes polarimetry and capturing the interference fringe pattern, respectively. We discuss the experimental results and comment prospects of this research. Finally, we summarize the conclusion.

2. Experiments

In this section, after proposing a generation system of arbitrary AP pulses possessing a broadband spectrum with spatial complex amplitude modulation, we show an experimental setup for evaluation of generated AP pulses. Finally, we describe experimental results of the generation of p = 0 and p = 1 LG AP pulses.

2.1. Proposal of system capable for generation of arbitrary AP pulses with a broadband spectrum

Our schematic generation system is shown in Fig. 1(a). This generation system has four principle components: a 4-f SLM system, a relay lens system, a polarization converter system and a SVWP.

 figure: Fig. 1

Fig. 1 (a) System setup capable for generation of arbitrary AP broadband pulses. SLM1,2, a liquid crystal on silicon spatial light modulators; L1-L4, convex lenses (f1,2,3 = 500 mm, f4 = 1000 mm); I1,2, irises; AHWP1,2, achromatic half-wave plates; AQWP1, an achromatic quarter-wave plate; SVWP, a space variant wave plate (Photonic Lattice SWP-808). |s, l〉 (= |s〉 |l〉) represents the state of spin and orbit angular momentum of light in the ħ units [23, 32]. (b) an AP state of final generated pulses plotted on the l =1 normalized extended Poincaré sphere (EPS). The final generated state is manipulated by the angles of AHWP1 (θH1) and AHWP2 (θH2) in the polarization converter. (c) (ϕl=1, θl=1)-plane of the l =1 normalized EPS. Here, ϕl=1 and θl=1 are, respectively, azimuthal angle and polar angle. The rotations of AHWP1 and AHWP2 make the moves of the final generated AP state parallel to the θl=1 and ϕl=1 axes, respectively.

Download Full Size | PDF

The 4-f SLM system, which is composed of two SLMs (SLM1,2), two lenses (L1,2) and an iris (I1) (I1 selects the first diffraction ray), can modulate the complex amplitude distribution of the input horizontally-polarized Gaussian pulses without the spatial chromatic dispersion [24,25]. Normally, a 4-f SLM system is used to convert Gaussian broadband pulses into OV broadband pulses by applying both amplitude and phase modulations. In our system, we apply just amplitude modulation to the input Gaussian broadband pulses by giving the depth modulated grating pattern [36–39] on SLM2.

The pulse after SLM2 is no longer an eigenmode of the paraxial Helmholtz wave equation. The relay lens system (L3 and L4) provides the amplitude-modulated image on SLM2 to SVWP. Here, I2 in the relay lens system selects only the first diffraction ray from SLM2. The complex amplitude on the input facet of SVWP can be described by fSLM21st(rf4/f3), where fSLM21st(r) is the complex amplitude of the first diffraction beam in the beam cross section on SLM2,(f)3,4 are, respectively, the focal lengths of L3 and L4.

A polarization converter system is located between L4 and SVWP. The system is composed of a sequence of achromatic wave plates in the order of an achromatic half-wave plate (AHWP1), an achromatic quarter-wave plate (AQWP1) and another achromatic half-wave plate (AHWP2). The Jones vector after the combination of AHWP1 and AQWP1 with a horizontally-polarized light input can be expressed by

exp(iπ/4)2(1ii1)(cos2θH1sin2θH1)=exp(+iπ/4)sin2θH1|s=+1+exp(iπ/4)cos2θH1|s=1,
where θH1 is the fast axis angle of AHWP1, and |s=(1is)T/2 represents the state of spin angular momentum of light in the ħ units [23, 32]. The LCP and RCP components correspond to |s = +1〉 and |s = −1〉 states, respectively. The rotation angle of AHWP1 θH1 determines the ellipticity of the output polarization state because θH1 adjusts the absolute value of superposition coefficient of the LCP and RCP states. The Jones matrix of AHWP2 can be written by
(cos2θH2sin2θH2sin2θH2cos2θH2)=RθH2(1001)RθH2=|s=+1eiθH2s=1|+|s=1e+iθH2|s=+1,
where θH2 is the fast axis angle of AHWP2, and
Rθ=(cosθsinθsinθcosθ)
is a rotation matrix. AHWP2 reverses the polarization state on its fast axis. Consequently, the rotation angle θH2 of AHWP2 determines the long axis direction of the output polarization state from the polarization converter system. Thereby, the polarization converter system enables us to convert the horizontally-polarized state into any uniform polarization states by rotating AHWP1 and AHWP2. As we describe in detail later, the uniform polarization state from the polarization converter system is the base of the polarization state of the final generated AP pulse state.

In our proposing system, SVWP is a q = 1/2 half-wave plate (or q-plate [40]), which transforms the uniformly-distributed polarization state into an AP state, i.e., l = 1 cylindrically polarized (CP) state. Here, q and l describe the fast-axis distribution and the rotational symmetry of CP state [40, 41], respectively. In terms of the spin-to-orbit angular momentum conversion, the space variant wave plate can be regarded as a converter that flips spin angular momentum of light s and gives orbital angular momentum of light l [Fig. 1(a)] [32, 40, 42–44] because the Jones matrix of a SVWP is described by

(cos2qϕsin2qϕsin2qϕcos2qϕ)=|s=+1ϕ|l=2qs=1|+|s=1ϕ|l=+2qs=+1|,
where 〈φ|l〉 = ei represents the state of orbit angular momentum of light in the ħ units [23,32] and |ϕ〉 is a position base state on the ϕ axis. We take account of this property when we analyze spatial phase distribution of the final generated pulse.

The complex amplitude state f (r, ϕ) of the final generated pulse in the beam cross section can be described as

f(r,ϕ)=fSLM21st(rf4/f3)(sin2θH1ei(2θH2π/4)|s=+1ϕ|l=1+cos2θH1ei(2θH2π/4)|s=1ϕ|l=+1).
From Eq. (5), the complex amplitude of the final generated beam is modulated on its radial axis by SLM2. We can interpret Eq. (5) as the superposition of |s = +1, l = −1〉 (= |s = +1〉 |l = −1〉) and |s = −1, l = +1〉 (= |s = −1〉 |l = +1〉) states. In the paper that we previously reported [23], we generated AP pulse states by using a separate-path interferometer based on Mach-Zehnder interferometer. Separate-path interferometers are unstable on account of sensitive to disturbance (e.g. vibration and air turbulence). In contrast to that, the system that we propose in this paper, being regarded as a common-path interferometer, has robustness to such disturbance.

The normalized ESPs S˜l=1E=(S˜1,l=1ES˜2,l=1ES˜3,l=1E)T [23, 32–34], which is an extension of conventional Stokes parameters into spatial dependent polarization states, describe the AP state of the final generated pulse as following:

S˜l=1E=(sin4θH2sin4θH1cos4θH2sin4θH1cos4θH1).
By rotating AHWP1 and AHWP2, arbitrary AP states can be generated as shown in Fig. 1(b). Here, Fig. 1(b) is the l = 1 normalized extended Poincaré sphere (EPS) constructed by the space of l = 1 normalized ESPs. The l = 1 normalized EPS has two degree of freedom: azimuthal angle ϕl = 1 and polar angle θl = 1 [34]. By using these parameters, the normalized ESPs are expressed by S˜l=1E(ϕl=1,θl=1)=(sinθl=1cosϕl=1sinθl=1sinϕl=1cosθl=1)T, thus
ϕl=1=π2+4θH2,
θl=1=π4θH1.
As depicted in Fig. 1(c), the rotations of AHWP1 and AHWP2, respectively, move the final generated AP state parallel to the θl=1 and ϕl=1 axes, which corresponds to the move on a meridian and a latitude line of l = 1 EPS [Fig. 1(b)]. AHWP1 and AHWP2 changes the ellipticity and the orientation of the final generated AP state. Because of spherical coordinate system, (ϕl=1, θl=1)-plane [Fig. 1(c)] has periodical boundary conditions such as
S˜l=1E(ϕl=1,θl=1)=S˜l=1E(ϕl=1+2π,θl=1)=S˜l=1E(ϕl=1,θl=1+2π)=S˜l=1E(ϕl=1+π,2πθl=1).

Thereby, this generation system can generate arbitrary AP broadband pulses with spatial complex amplitude modulation.

2.2. Experimental setup

Figure 2(a) shows the schematic experimental setup (See Appendix for the complete experimental setup). A horizontally-polarized femtosecond pulse (central wavelength, 800 nm; bandwidth, 60 nm) from a Ti:Sa oscillator passed through a bandpass filter (BPF; central wavelength, 800 nm; bandwidth, 3 nm) and its pulse duration was evaluated to be ∼ 4 ps. Although the generation system we proposed supports broadband pulses, we used here a narrowband pulse as an input in order to perform fundamental demonstration experiments. The pulse was divided into two by a broadband non-polarizing beam splitters (BS1); they were guided into the main and the reference branches, respectively. The reference pulse was necessary for evaluating the spatial phase distribution of the finally generated pulse but unnecessary just for generation of AP pulses. The pulse in the main branch passed through the generation system and finally the generated pulse was evaluated in terms of two aspects: polarization and phase distributions. The former and the latter were, respectively, evaluated by using a rotating retarder-type imaging polarimeter [23, 45] and by capturing a spatial interference with the reference pulse beam. A rotating retarder-type imaging polarimeter was composed of an achromatic quarter-wave plate (AQWP2), a polarizer (POL1) and a charge coupled device camera (CCD). When we measured polarization distributions of the final generated pulses, we blocked the reference pulse beam. For the evaluation of phase distributions, we acquired the interference patterns of |s = 1〉 (|s = −1〉) component of the final generated pulse with the reference pulse beam, setting the angle of AQWP2 θQ2 = 135 deg (45 deg) [Fig. 2(b)]. When θQ2 = 135 deg (45 deg), the polarization analyzer composed of AQWP2 and POL1 outputs the |s = 1〉 (|s = −1〉) component as a vertically polarized state, thus the interference pattern with the vertically polarized reference pulses was observed. The crossing angle in the interference was 0.7 deg. The reference pulse was enlarged enough in the beam cross section by a beam expander, composed of L7 and L8, to cover the whole beam pattern of the final generated pulses. Since a pinhole (P2) was inserted into the beam expander for the sake of beam cleaning, the intensity distribution of the reference pulse was almost Gaussian shape. Its beam size was estimated to be 3 mm. An achromatic half-wave plate (AHWP3) and a polarizer (POL2) purified vertically-polarized reference pulses and adjusted their intensity in order to maximize visibility of the interference patterns.

 figure: Fig. 2

Fig. 2 (a) Outline drawing of whole experimental setup. Ti:Sa Osc., a Ti:Sapphire oscillator (central wavelength, 800 nm; bandwidth, 60 nm); BPF, a bandpass filter (central wavelength, 800 nm; bandwidth, 3 nm); BS1,2, broadband non-polarizing beam splitters; L5-8, convex lenses (f5,8 = 200 mm, f6,7 = 100 mm); P1,2, pinholes; AQWP2,3, achromatic quarter-wave plates; POL1,2, polarizers; a delay stage is put in the reference pulse branch (see the Appendix). (b) Method for measurement of interference pattern of the |s = ±1〉 component of the final generated pulses. When θQ2 = 45 [deg], the polarization analyzer outputs the |s = −1〉 component as a vertically polarized state. By combining vertically polarized reference pulses, we observe an interference fringe pattern of the |s = −1〉 component of the final generated pulses with the reference pulse beam. When θQ2 = 135 [deg], we observe that of the |s = 1〉 component of the final generated pulses.

Download Full Size | PDF

A pinhole (P1) purified the spatial rotational symmetry of the final generated pulse because SVWP that we used was a 12-segmented half-wave plate and deteriorates somewhat its spatial rotational symmetry into C12. If we use a smoothly fast-axis distributed SVWP, P1 is essentially not necessary.

In the experiment, we applied two patterns of amplitude modulations on SLM2; the one [Fig. 3(a)] and the other [Fig. 3(b)] were to generate p = 0 and p = 1 LG AP pulses, respectively. Here, the patterns were calculated by using the Davis’s method [37, 39]. In particular, the p = 1 LG mode has phase difference of π between the inner ring and the outer ring, which reflected the half-cycle gap of grating pattern between these rings [Fig. 3(b)]. As we mentioned before, we omitted the spiral phase term ei of LG modes from these fringe patterns. If we apply the spiral phase ei to the pattern on SLM2, the final generated pulse is not an eigenmode of LG AP mode any longer; in other words, the vector vortex state which is the superposition of |s = 1, l = m − 1〉 and |s = 1, l = m + 1〉 OVs is generated in that case.

 figure: Fig. 3

Fig. 3 Grating patterns on SLM2 in order to generate (a) p = 0 and (b) p = 1 LG AP pulses.

Download Full Size | PDF

2.3. Experimental results

2.3.1. Generation of radially polarized pulses with spatial complex amplitude modulation

First, we generated the p = 0 and p = 1 LG radially polarized (RP) pulses in order to see how precisely the amplitude and phase modulation operated. The polarization and intensity distribution for p = 0 and p = 1 LG RP pulses are shown in Figs. 4(a) and 4(e), respectively.

 figure: Fig. 4

Fig. 4 Generation results of (a–d) p = 0 and (e–h) p = 1 LG RP pulses. (a,e) Intensity and polarization distributions of (a) p = 0 and (e) p = 1 LG RP pulses. (b,f) Intensity on the radial axis r of (b) p = 0 and (f) p = 1 LG RP pulses. Red and cyan curves represent experimental and fitting ones, respectively. (c,d) Interference fringe patterns of (c) |s = 1〉 and (d) |s = 1〉 components in p = 0 LG RP pulses with the reference pulse beam. (g,h) Interference fringe patterns of (g) |s = −1〉 and (h) |s = 1〉 components in p = 1 LG RP pulses with the reference pulse beam. Note that the images in (e), (g) and (h) are displayed on a 75% scale of the images in (a), (c) and (d).

Download Full Size | PDF

As shown in Table 1, the polarization distributions were quantitatively evaluated by using the normalized l = 1 first ESP (strictly speaking, all the values of ESPs in this article were normalized by the energy of the temporally- and spatially-perfect-polarized states [23]) and the degree of polarization for the spatial distribution (DOP-SD) [23, 33]. Since the normalized l = 1 first ESP S1,l=1E gives the energy ratio of the RP state to the azimuthally polarized state as (1+S˜1,l=1E)/(2Pl=1space):(1S˜1,l=1E)/(2Pl=1space) [23], over 99.5% power of both the generated p = 0 and p = 1 pulses were RP states. Moreover, the DOP-SD Pl=1space represents the energy purity in the symmetry of C|l−1| (when l = 1, C) polarization distribution [23]. The energy proportions of the states having C polarization distributions were ∼99% in both the generated pulses. These facts indicate that we generated RP pulses with high purity in terms of polarization distribution.

Tables Icon

Table 1. The values of the first ESP S˜1,l=1 and the DOP-SD Pl=1space of generated p = 0 LG RP pulses [Fig. 4(a)] and p = 1 LG RP pulses [Fig. 4(e)].

Figures 4(b) and 4(f) show the intensity distributions (red curves) on the radial axis r of p = 0 and p = 1 LG RP pulses, respectively. The intensity on the radial axis I(r) were obtained by using the following equation:

I(r)=12π02πI(r,ϕ)dϕ,
where I(r, ϕ) is intensity as a function of r and ϕ, and the center position (r = 0) was selected to be the position where DOP-SD Pl=1space was maximized. The cyan curves in Figs. 4(b) and 4(f) are fitting curves of the intensity of p = 0 and p = 1 LG modes Il=1,pLG(r), which are described [35] by
IlLG(r)=|A|22p!π(p+|l|)!(2r2w2)|l|{Lp|l|(2r2w2)}21w2exp(2r2w2),
where Lp|l|(ξ) represents the generalized Laguerre polynomials, and we performed fitting with adjustable parameters of an amplitude A and a beam size w. The intensity distribution of p = 0 LG RP pulses [Fig. 4(b)] well overlapped with the fitting curve. In contrast to that, the experimental curve of p = 1 LG RP pulses somewhat deviated from the fitting curve [Fig. 4(f)]. However, we evaluated the energy purity of p = 1 LG RP mode was high enough (as high as ≳ 93% by using mode analysis via phase reconstruction method [46]), more details of which are the subject of a future paper.

Figure 4(c) [(d)] shows the interference pattern of |s = −1〉 (RCP) [|s = 1〉 (LCP)] component of p = 0 LG RP pulses with the reference pulses. Figure 4(g) [(h)] shows the interference pattern of |s = −1〉 [|s = 1〉] component of p = 1 LG RP pulses with the reference pulses. All the patterns had two-pronged fork patterns, which indicated that the absolute values of the dominant topological charge of these components were 1 [47, 48]. By considering the direction of the folk patterns, both the p = 0 and 1 LG RP pulses possessed |s = −1, l = +1〉 and |s = 1, l = −1〉 components. We surely generated the true LG RP pulses. Moreover, in-phase and out-of-phase positions of the inner and the outer rings of p = 1 LG RP pulses [Figs. 4(g) and 4(h)] flipped. The flip reflected the π phase jump between inner and outer rings of p = 1 LG modes. Thus, the spatial phase distribution as well as intensity distribution was successfully modulated by SLM2.

2.3.2. Generation of arbitral axisymmetrically polarized pulses with spatial complex amplitude modulation

In the next place, we demonstrated the generation of various AP pulses with p = 0 or p = 1 LG modes in order to show the arbitral AP pulses can be generated.

By using the polarization analyzer, we observed the transition of AP states when AHWP1 was rotated with the angle of AHWP2 fixed to θH2 = −22.5 deg. Figures 5(a) and 5(c) are, respectively, the transition of polarization distribution for p = 0 and p = 1 LG AP pulses with a 10 deg increase in θH1. We observed generation of elliptic AP states by rotating AHWP1, and hence the move of the AP state on the meridian of the l = 1 EPS [Fig. 1(b)] or the move of that parallel to θl=1 axis of (ϕl=1, θl=1)-plane [Fig. 1(c)] was realized. In contrast, Figs. 5(b) and 5(d) show the transition of AP states when AHWP2 was rotated with the angle of AHWP1 fixed to θH1 = +22.5 deg. The results show that the rotation of AHWP2 made the move of the AP state on the latitude line of the l = 1 EPS.

 figure: Fig. 5

Fig. 5 Transition of polarization distribution when (a,c) AHWP1 was rotated under the condition that θH2 = −22.5 deg, or when (b,d) AHWP2 was rotated under the condition that θH1 = 22.5 deg. The grating pattern on SLM2 was (a,b) the pattern in Fig. 3(a) (p = 0 LG AP mode) or (c,d) the pattern in Fig. 3(b) (p = 1 LG AP mode). These polarization distributions are colored under the following rule: black, linear polarization; red, left-handed elliptical polarization; blue, right-handed elliptical polarization. The images in (c) and (d) are displayed on a 75% scale of the images in (a) and (b).

Download Full Size | PDF

Polarization distributions in Fig. 5 are visually apparent but give only qualitative information, hence we fully quantitatively evaluated these polarization distribution by using the normalized l = 1 ESPs and the l = 1 DOP-SD [23, 33]. Figures 6(a)6(d), respectively, show the quantitatively-evaluated values of the polarization distributions depicted in Figs. 5(a)5(d). No significant differences of ESPs and DOP-SD were found between the modulations, indicating that the manipulation of AP states by the polarization converter works well with independent of the spatial beam patterns. Basically, the values of ESPs followed the curves of Eq. (6). However, while S˜2,l=1E in Figs. 6(a) and 6(c), and S˜3,l=1E in Figs. 6(b) and 6(d) should be constant functions with respect to θH1 and θH2 respectively, they slightly deviated from the constant functions. This was ascribed to retardation errors of waveplates. We plotted the experimental data in Fig. 6 on the (ϕl=1, θl=1)-plane (Fig. 7), which showed that the shifts of the AP state parallel to θl=1 and ϕl=1 axes on the (ϕl=1, θl=1)-plane were achieved by rotating AHWP1 and AHWP2, respectively. Thus, this generation system can generate arbitrary AP pulse states. The values of DOP-SD Pl=1space were over 0.988 in all the evaluated results, which meant that all the generated AP pulses had as high as over 98.8% purity of C polarization distribution.

 figure: Fig. 6

Fig. 6 (a–d) The value of normalized l = 1 ESPs S˜i,l=1E(i=13) and l = 1 DOP-SD corresponding to the polarization distributions displayed in Fig. 5(a–d). Solid curves represent ideal ones described by Eq. (6).

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Experimental results for (a) p = 0 and (b) p = 1 LG AP states plotted on the (ϕl=1, θl=1)-planes. Dashed lines are ideal trajectories [Eqs. (7)(8)]

Download Full Size | PDF

3. Discussions and prospects

The experimental demonstration in Sec. 2.3.1 yielded high purity (≳ 93%) of p = 0 and p = 1 LG RP pulse states. Some researches have been shown that the p = 0 LG mode with high purity can be easily generated from a point vortex state by using a low-spatial-frequency-pass filter such as a pinhole. In comparison, the spatial quality of our generated beam was superior to that in an earlier research [49]. This improvement in our method was done by as follows: first, we pre-shaped the spatial beam pattern to the target pattern (except the phase profiles), which is imaged on the SVWP, by using a 4-f SLM so that the higher-order mode components are reduced. Second, by use of the low-spatial-frequency-pass filter, the residual higher-order modes (p > 0 LG modes) were fully eliminated. Thereby, in the present study, we essentially modulated the spatial complex amplitude through SLM and generated p = 0 and p = 1 LG RP pulse states. In addition to that, the improvement of the purity can be achieved by means of a feedback mechanism [50]. Some researches [51, 52] have reported on generations of single mode LG RP CW beams with high spatial quality. Their method utilized a Fabry-Perot interferometer as a mode cleaner after a SVWP. They, however, cannot be applied to broadband pulses. Unlike systems using Fabry-Perot interferometers, our system is suitable for broadband pulses, being able to generate superposition states of AP LG modes with different p indices.

If we replace the q = 1/2 half-wave plate with a q = Q half-wave plate, the final output pulse is l = 2q (= 2Q) CP pulse. Moreover, when we make complex amplitude modulation of l = 2q LG mode, the spatial mode of the final generated pulse will be l = 2q LG CP mode, which is an eigenmode of the paraxial Helmholtz wave equation. The system we proposed has thereby scalability of any order of LG CP modes.

As we mentioned in the Sec. 2.2, when the hologram grating pattern with ei is applied on SLM2, the finally generated pulses are the superposition of |s = 1, l = m − 1〉 and |s = −1, l = m + 1〉 OVs. Thus, they are no longer AP state pulses (since AP state is the superposition of |s = 1, l = −1〉 and |s = −1, l = 1〉 OVs), but vector vortex pulses. Since it is difficult to distinguish AP states and vector vortex state solely through polarization distributions, we observed the interference patterns of |s = 1〉 and |s = −1〉 OV of the final generated pulses with the reference pulses (Fig. 4(c,d) and (g,h)), respectively. Thereby, we conclude that we generated AP pulses. In a precise sense, the evaluation of mode decomposition is needed, but it is beyond the area of this paper and we will report on it in another article.

The system we proposed in Sec. 2.1, which supports the generation of broadband AP pulses, can generate ultrashort pulses when we just apply a frequency chirp compensation to the linear polarized broadband pulses before SLM1. Moreover, when a femtosecond polarization pulse shaper [53] is used in place of the polarization converter, we can realize the transformation of AP state in a single pulse, which may enable us to precisely manipulate light-matter interactions in nonlinear spectroscopy.

Recently, space division multiplexing transmission experiments by using LG modes with the higher radial index p in the free space have been demonstrated [54, 55]. In addition to that, the radial index of LG modes is proved to be a quantum number [56–58] and the fundamental experiments of Hong-Ou-Mandel interference was performed [59]. Our technique of generation of LG AP pulses can be applied in the fields of classical and quantum communication.

4. Conclusion

We have proposed a system to realize broadband generation of arbitral AP pulses with spatial complex amplitude modulation. In this system, a 4-f SLM and a space variant wave plate, respectively, compensates spatial chromatic dispersion and gives stability to perturbation. On the basis of our proposal, we constructed the demonstration system and generated various AP pulses by applying p = 0 and 1 LG mode modulations. The polarization purity evaluated by ESPs was high in each modulation (DOP-SD Pl=1space>0.988), and there was no significant difference between the modulations. We confirmed that the complex amplitude modulation worked well through intensity patterns and interference fringes with not so broadband pulses. The evaluation for broadband pulses is a future study. Our proposed system, which is easily extended to generation of ultrashort LG CP pulses, can be utilized in classical and quantum communication with higher-order spatial modes.

Appendix

In this section, we describe the detail of experimental setup. Figure 8 shows the actual experimental setup. Single mode fiber (SMF) performed a function as a mode cleaner. A pair of an achromatic half-wave plate (AHWP0) and a polarized beam splitter (PBS) modulated pulse energy. The 4-f SLM system was realized with the folded path configuration as depicted in Fig. 9. The grating patterns for SLM1 and SLM2 in Fig. 1(a) were shown on the dimidiate areas of a single spatial light modulator (SLM). The sequence of relay lenses in the reference pulse beam path (Fig. 8, the dashed line) formed an image on CCD and stabilized the pointing of the reference pulse beam. P2 cleaned the spatial beam profile of reference pulses.

 figure: Fig. 8

Fig. 8 Whole experimental setup. Ti:Sa Osc., a Ti:Sapphire oscillator (central wavelength, 800 nm; bandwidth, 60 nm); OL1,2, objective lenses; SMF, single mode fiber (∼10 m); M1-15, mirrors; L1-L14, convex lenses (L1-3,9,10,12,14, f = 500 mm; L4,11, f = 1000 mm; L5,7,13, f = 200 mm; L6,8, f = 100 mm); PBS, a polarizing beam splitter; BS1,2, broadband non-polarizing beam splitters; SLM, a liquid crystal on silicon spatial light modulator; MZ, a zero-incidence mirror; I1,2, irises; GL1,2, Glan-Laser Calcite polarizers; AHWP0-2, achromatic half-wave plates; AQWP1,2, achromatic quarter-wave plates; SVWP, a space variant wave plate (Photonic Lattice SWP-808); Delay, a delay stage; POL2, a polarizer; BPF, a bandpass filter (central wavelength, 800 nm; bandwidth, 3 nm).

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 A three-dimensional configuration of optical components in a 4-f SLM system.

Download Full Size | PDF

Funding

Japan Society for the Promotion of Science (JSPS) (JP26286056, JP15J00038, JP16H06506); Japan Science and Technology agency (JST) (CREST).

References and links

1. Q. Zhan, “Cylindrical vector beams: from mathematical concepts to applications,” Adv. Opt. Photon. 1, 1–57 (2009). [CrossRef]  

2. L. Novotny, M. R. Beversluis, K. S. Youngworth, and T. G. Brown, “Longitudinal field modes probed by single molecules,” Phys. Rev. Lett. 86, 5251–5254 (2001). [CrossRef]   [PubMed]  

3. M. O. Scully and M. S. Zubairy, “Simple laser accelerator: Optics and particle dynamics,” Phys. Rev. A 44, 2656–2663 (1991). [CrossRef]   [PubMed]  

4. Y. I. Salamin and C. H. Keitel, “Electron acceleration by a tightly focused laser beam,” Phys. Rev. Lett. 88, 095005 (2002). [CrossRef]   [PubMed]  

5. Y. I. Salamin, Z. Harman, and C. H. Keitel, “Direct high-power laser acceleration of ions for medical applications,” Phys. Rev. Lett. 100, 155004 (2008). [CrossRef]   [PubMed]  

6. V. Marceau, C. Varin, T. Brabec, and M. Piché, “Femtosecond 240-keV electron pulses from direct laser acceleration in a low-density gas,” Phys. Rev. Lett. 111, 224801 (2013). [CrossRef]   [PubMed]  

7. M. Meier, V. Romano, and T. Feurer, “Material processing with pulsed radially and azimuthally polarized laser radiation,” Appl. Phys. A 86, 329–334 (2007). [CrossRef]  

8. M. Kraus, M. A. Ahmed, A. Michalowski, A. Voss, R. Weber, and T. Graf, “Microdrilling in steel using ultrashort pulsed laser beams with radial and azimuthal polarization,” Opt. Express 18, 22305–22313 (2010). [CrossRef]   [PubMed]  

9. T. Omatsu, K. Chujo, K. Miyamoto, M. Okida, K. Nakamura, N. Aoki, and R. Morita, “Metal microneedle fabrication using twisted light with spin,” Opt. Express 18, 17967–17973 (2010). [CrossRef]   [PubMed]  

10. C. Hnatovsky, V. Shvedov, W. Krolikowski, and A. Rode, “Revealing local field structure of focused ultrashort pulses,” Phys. Rev. Lett. 106, 123901 (2011). [CrossRef]   [PubMed]  

11. O. Allegre, W. Perrie, S. Edwardson, G. Dearden, and K. Watkins, “Laser microprocessing of steel with radially and azimuthally polarized femtosecond vortex pulses,” J. Opt. 14, 085601 (2012). [CrossRef]  

12. K. Toyoda, F. Takahashi, S. Takizawa, Y. Tokizane, K. Miyamoto, R. Morita, and T. Omatsu, “Transfer of light helicity to nanostructures,” Phys. Rev. Lett. 110, 143603 (2013). [CrossRef]   [PubMed]  

13. J. J. Nivas, S. He, A. Rubano, A. Vecchione, D. Paparo, L. Marrucci, R. Bruzzese, and S. Amoruso, “Direct femtosecond laser surface structuring with optical vortex beams generated by a q-plate,” Sci. Rep. 5, 17929 (2015). [CrossRef]   [PubMed]  

14. L. Rapp, R. Meyer, L. Furfaro, C. Billet, R. Giust, and F. Courvoisier, “High speed cleaving of crystals with ultrafast bessel beams,” Opt. Express 25, 9312–9317 (2017). [CrossRef]   [PubMed]  

15. V. Shvedov, A. R. Davoyan, C. Hnatovsky, N. Engheta, and W. Krolikowski, “A long-range polarization-controlled optical tractor beam,” Nature Photon. 8, 846–850 (2014). [CrossRef]  

16. C. Gabriel, A. Aiello, W. Zhong, T. G. Euser, N. Y. Joly, P. Banzer, M. Förtsch, D. Elser, U. L. Andersen, C. Marquardt, P. S. J. Russell, and G. Leuchs, “Entangling different degrees of freedom by quadrature squeezing cylindrically polarized modes,” Phys. Rev. Lett. 106, 060502 (2011). [CrossRef]   [PubMed]  

17. C. Gabriel, A. Aiello, S. Berg-Johansen, C. Marquardt, and G. Leuchs, “Tools for detecting entanglement between different degrees of freedom in quadrature squeezed cylindrically polarized modes,” Eur. Phys. J. D 66, 1–5 (2012). [CrossRef]  

18. C. Hernández-García, A. Turpin, J. S. Román, A. Picón, R. Drevinskas, A. Cerkauskaite, P. G. Kazansky, C. G. Durfee, I. nigo, and J. Sola, “Extreme ultraviolet vector beams driven by infrared lasers,” Optica 4, 520–526 (2017). [CrossRef]  

19. Y. Tokizane, K. Shimatake, Y. Toda, K. Oka, M. Tsubota, S. Tanda, and R. Morita, “Global evaluation of closed-loop electron dynamics in quasi-one-dimensional conductors using polarization vortices,” Opt. Express 17, 24198–24207 (2009). [CrossRef]  

20. F. K. Fatemi, “Cylindrical vector beams for rapid polarization-dependent measurements in atomic systems,” Opt. Express 19, 25143–25150 (2011). [CrossRef]  

21. H. Kim, M. Akbarimoosavi, and T. Feurer, “Probing ultrafast phenomena with radially polarized light,” Appl. Opt. 55, 4389–4394 (2016). [CrossRef]   [PubMed]  

22. K. Shigematsu, M. Suzuki, K. Yamane, R. Morita, and Y. Toda, “Snap-shot optical polarization spectroscopy using radially polarized pulses,” Appl. Phys. Express 9, 122401 (2016). [CrossRef]  

23. M. Suzuki, K. Yamane, K. Oka, Y. Toda, and R. Morita, “Full quantitative analysis of arbitrary cylindrically polarized pulses by using extended Stokes parameters,” Sci. Rep. 5, 17797 (2015). [CrossRef]   [PubMed]  

24. I. G. Mariyenko, J. Strohaber, and C. J. G. J. Uiterwaal, “Creation of optical vortices in femtosecond pulses,” Opt. Express 13, 7599–7608 (2005). [CrossRef]   [PubMed]  

25. 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). [CrossRef]   [PubMed]  

26. M. Sakamoto, K. Oka, and R. Morita, “Diffraction characteristics of optical and polarization vortices generated by an axially symmetric polarizer,” Proc. SPIE 8274, 827414 (2012). [CrossRef]  

27. C. Maurer, A. Jesacher, S. Fürhapter, S. Bernet, and M. Ritsch-Marte, “Tailoring of arbitrary optical vector beams,” New J. Phys. 9, 78 (2007). [CrossRef]  

28. M. Sakamoto, T. Sasaki, K. Noda, T. M. Tien, N. Kawatsuki, and H. Ono, “Three-dimensional vector recording in polarization sensitive liquid crystal composites by using axisymmetrically polarized beam,” Opt. Lett. 41, 642–645 (2016). [CrossRef]   [PubMed]  

29. V. Chille, S. Berg-Johansen, M. Semmler, P. Banzer, A. Aiello, G. Leuchs, and C. Marquardt, “Experimental generation of amplitude squeezed vector beams,” Opt. Express 24, 12385–12394 (2016). [CrossRef]   [PubMed]  

30. K. J. Mitchell, S. Turtaev, M. J. Padgett, T. Čižmár, and D. B. Phillips, “High-speed spatial control of the intensity, phase and polarisation of vector beams using a digital micro-mirror device,” Opt. Express 24, 29269–29282 (2016). [CrossRef]   [PubMed]  

31. E. Karimi, G. Zito, B. Piccirillo, L. Marrucci, and E. Santamato, “Hypergeometric-Gaussian modes,” Opt. Lett. 32, 3053–3055 (2007). [CrossRef]   [PubMed]  

32. M. Suzuki, K. Yamane, K. Oka, Y. Toda, and R. Morita, “Nonlinear coupling between axisymmetrically-polarized ultrashort optical pulses in a uniaxial crystal,” Opt. Express 22, 16903–16915 (2014). [CrossRef]   [PubMed]  

33. M. Suzuki, K. Yamane, K. Oka, Y. Toda, and R. Morita, “Extended Stokes parameters for cylindrically polarized beams,” Opt. Rev. 22, 179–183 (2015). [CrossRef]  

34. M. Suzuki, “A comprehensive study on cylindrical symmetry in optical physics : Full-quantitative characterization of cylindrically polarized optical pulses,” Ph.D. thesis, Hokkaido Univ. (2016).

35. 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]   [PubMed]  

36. J. P. Kirk and A. L. Jones, “Phase-only complex-valued spatial filter,” J. Opt. Soc. Am. 61, 1023–1028 (1971). [CrossRef]  

37. J. A. Davis, D. M. Cottrell, J. Campos, M. J. Yzuel, and I. Moreno, “Encoding amplitude information onto phase-only filters,” Appl. Opt. 38, 5004–5013 (1999). [CrossRef]  

38. V. Arrizón, U. Ruiz, R. Carrada, and L. A. González, “Pixelated phase computer holograms for the accurate encoding of scalar complex fields,” J. Opt. Soc. Am. A 24, 3500–3507 (2007). [CrossRef]  

39. T. Ando, Y. Ohtake, N. Matsumoto, T. Inoue, and N. Fukuchi, “Mode purities of Laguerre–Gaussian beams generated via complex-amplitude modulation using phase-only spatial light modulators,” Opt. Lett. 34, 34–36 (2009). [CrossRef]  

40. L. Marrucci, C. Manzo, and D. Paparo, “Optical spin-to-orbital angular momentum conversion in inhomogeneous anisotropic media,” Phys. Rev. Lett. 96, 163905 (2006). [CrossRef]   [PubMed]  

41. M. Suzuki, K. Yamane, K. Oka, Y. Toda, and R. Morita, “Analysis of the Pancharatnam-Berry phase of vector vortex states using the Hamiltonian based on the Maxwell-Schrödinger equation,” Phys. Rev. A 94, 043851 (2016). [CrossRef]  

42. L. Allen, J. Courtial, and M. J. Padgett, “Matrix formulation for the propagation of light beams with orbital and spin angular momenta,” Phys. Rev. E 60, 7497–7503 (1999). [CrossRef]  

43. E. Brasselet, Y. Izdebskaya, V. Shvedov, A. S. Desyatnikov, W. Krolikowski, and Y. S. Kivshar, “Dynamics of optical spin-orbit coupling in uniaxial crystals,” Opt. Lett. 34, 1021–1023 (2009). [CrossRef]   [PubMed]  

44. R. C. Devlin, A. Ambrosio, D. Wintz, S. L. Oscurato, A. Y. Zhu, M. Khorasaninejad, J. Oh, P. Maddalena, and F. Capasso, “Spin-to-orbital angular momentum conversion in dielectric metasurfaces,” Opt. Express 25, 377–393 (2017). [CrossRef]   [PubMed]  

45. E. Collett, Polarized light in fiber optics (SPIE, 2003).

46. K. Yamane, Z. Yang, Y. Toda, and R. Morita, “Frequency-resolved measurement of the orbital angular momentum spectrum of femtosecond ultra-broadband optical-vortex pulses based on field reconstruction,” New J. Phys. 16, 053020 (2014). [CrossRef]  

47. N. Baranova, B. Y. Zel’Dovich, A. Mamaev, N. Pilipetskii, and V. Shkukov, “Dislocations of the wavefront of a speckle-inhomogeneous field (theory and experiment),” JETP Lett. 33, 206 (1981).

48. N. R. Heckenberg, R. McDuff, C. P. Smith, and A. G. White, “Generation of optical phase singularities by computer-generated holograms,” Opt. Lett. 17, 221–223 (1992). [CrossRef]   [PubMed]  

49. J. Kalwe, M. Neugebauer, C. Ominde, G. Leuchs, G. Rurimo, and P. Banzer, “Exploiting cellophane birefringence to generate radially and azimuthally polarised vector beams,” European J. Phys. 36, 025011 (2015). [CrossRef]  

50. 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 compensationM,” Opt. Lett. 28, 2258–2260 (2003). [CrossRef]   [PubMed]  

51. R. Dorn, S. Quabis, and G. Leuchs, “Sharper focus for a radially polarized light beam,” Phys. Rev. Lett. 91, 233901 (2003). [CrossRef]   [PubMed]  

52. S. Quabis, R. Dorn, and G. Leuchs, “Generation of a radially polarized doughnut mode of high quality,” Applied Physics B 81, 597–600 (2005). [CrossRef]  

53. T. Brixner and G. Gerber, “Femtosecond polarization pulse shaping,” Opt. Lett. 26, 557–559 (2001). [CrossRef]  

54. G. Xie, Y. Ren, Y. Yan, H. Huang, N. Ahmed, L. Li, Z. Zhao, C. Bao, M. Tur, S. Ashrafi, and A. E. Willner, “Experimental demonstration of a 200-Gbit/s free-space optical link by multiplexing Laguerre–Gaussian beams with different radial indices,” Opt. Lett. 41, 3447–3450 (2016). [CrossRef]   [PubMed]  

55. L. Li, G. Xie, Y. Yan, Y. Ren, P. Liao, Z. Zhao, N. Ahmed, Z. Wang, C. Bao, A. J. Willner, S. Ashrafi, M. Tur, and A. E. Willner, “Power loss mitigation of orbital-angular-momentum-multiplexed free-space optical links using nonzero radial index Laguerre–Gaussian beams,” J. Opt. Soc. Am. B 34, 1–6 (2017). [CrossRef]  

56. E. Karimi and E. Santamato, “Radial coherent and intelligent states of paraxial wave equation,” Opt. Lett. 37, 2484–2486 (2012). [CrossRef]   [PubMed]  

57. W. N. Plick, R. Lapkiewicz, S. Ramelow, and A. Zeilinger, “The forgotten quantum number: A short note on the radial modes of Laguerre-Gauss beams,” https://arXiv:1306.6517 (2013).

58. E. Karimi, R. W. Boyd, P. de la Hoz, and H. de Guise, J. Řeháček, Z. Hradil, A. Aiello, G. Leuchs, and L. L. Sánchez-Soto, “Radial quantum number of Laguerre-Gauss modes,” Phys. Rev. A 89, 063813 (2014). [CrossRef]  

59. E. Karimi, D. Giovannini, E. Bolduc, N. Bent, F. M. Miatto, M. J. Padgett, and R. W. Boyd, “Exploring the quantum nature of the radial degree of freedom of a photon via Hong-Ou-Mandel interference,” Phys. Rev. A 89, 013829 (2014). [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 (9)

Fig. 1
Fig. 1 (a) System setup capable for generation of arbitrary AP broadband pulses. SLM1,2, a liquid crystal on silicon spatial light modulators; L1-L4, convex lenses (f1 , 2 , 3 = 500 mm, f4 = 1000 mm); I1,2, irises; AHWP1,2, achromatic half-wave plates; AQWP1, an achromatic quarter-wave plate; SVWP, a space variant wave plate (Photonic Lattice SWP-808). |s, l〉 (= |s〉 |l〉) represents the state of spin and orbit angular momentum of light in the ħ units [23, 32]. (b) an AP state of final generated pulses plotted on the l =1 normalized extended Poincaré sphere (EPS). The final generated state is manipulated by the angles of AHWP1 (θH1) and AHWP2 (θH2) in the polarization converter. (c) (ϕl=1, θl=1)-plane of the l =1 normalized EPS. Here, ϕl=1 and θl=1 are, respectively, azimuthal angle and polar angle. The rotations of AHWP1 and AHWP2 make the moves of the final generated AP state parallel to the θl=1 and ϕl=1 axes, respectively.
Fig. 2
Fig. 2 (a) Outline drawing of whole experimental setup. Ti:Sa Osc., a Ti:Sapphire oscillator (central wavelength, 800 nm; bandwidth, 60 nm); BPF, a bandpass filter (central wavelength, 800 nm; bandwidth, 3 nm); BS1,2, broadband non-polarizing beam splitters; L5-8, convex lenses (f5 , 8 = 200 mm, f6 , 7 = 100 mm); P1,2, pinholes; AQWP2,3, achromatic quarter-wave plates; POL1,2, polarizers; a delay stage is put in the reference pulse branch (see the Appendix). (b) Method for measurement of interference pattern of the |s = ±1〉 component of the final generated pulses. When θQ2 = 45 [deg], the polarization analyzer outputs the |s = −1〉 component as a vertically polarized state. By combining vertically polarized reference pulses, we observe an interference fringe pattern of the |s = −1〉 component of the final generated pulses with the reference pulse beam. When θQ2 = 135 [deg], we observe that of the |s = 1〉 component of the final generated pulses.
Fig. 3
Fig. 3 Grating patterns on SLM2 in order to generate (a) p = 0 and (b) p = 1 LG AP pulses.
Fig. 4
Fig. 4 Generation results of (a–d) p = 0 and (e–h) p = 1 LG RP pulses. (a,e) Intensity and polarization distributions of (a) p = 0 and (e) p = 1 LG RP pulses. (b,f) Intensity on the radial axis r of (b) p = 0 and (f) p = 1 LG RP pulses. Red and cyan curves represent experimental and fitting ones, respectively. (c,d) Interference fringe patterns of (c) |s = 1〉 and (d) |s = 1〉 components in p = 0 LG RP pulses with the reference pulse beam. (g,h) Interference fringe patterns of (g) |s = −1〉 and (h) |s = 1〉 components in p = 1 LG RP pulses with the reference pulse beam. Note that the images in (e), (g) and (h) are displayed on a 75% scale of the images in (a), (c) and (d).
Fig. 5
Fig. 5 Transition of polarization distribution when (a,c) AHWP1 was rotated under the condition that θH2 = −22.5 deg, or when (b,d) AHWP2 was rotated under the condition that θH1 = 22.5 deg. The grating pattern on SLM2 was (a,b) the pattern in Fig. 3(a) (p = 0 LG AP mode) or (c,d) the pattern in Fig. 3(b) (p = 1 LG AP mode). These polarization distributions are colored under the following rule: black, linear polarization; red, left-handed elliptical polarization; blue, right-handed elliptical polarization. The images in (c) and (d) are displayed on a 75% scale of the images in (a) and (b).
Fig. 6
Fig. 6 (a–d) The value of normalized l = 1 ESPs S ˜ i , l = 1 E ( i = 1 3 ) and l = 1 DOP-SD corresponding to the polarization distributions displayed in Fig. 5(a–d). Solid curves represent ideal ones described by Eq. (6).
Fig. 7
Fig. 7 Experimental results for (a) p = 0 and (b) p = 1 LG AP states plotted on the (ϕl=1, θl=1)-planes. Dashed lines are ideal trajectories [Eqs. (7)(8)]
Fig. 8
Fig. 8 Whole experimental setup. Ti:Sa Osc., a Ti:Sapphire oscillator (central wavelength, 800 nm; bandwidth, 60 nm); OL1,2, objective lenses; SMF, single mode fiber (∼10 m); M1-15, mirrors; L1-L14, convex lenses (L1-3,9,10,12,14, f = 500 mm; L4,11, f = 1000 mm; L5,7,13, f = 200 mm; L6,8, f = 100 mm); PBS, a polarizing beam splitter; BS1,2, broadband non-polarizing beam splitters; SLM, a liquid crystal on silicon spatial light modulator; MZ, a zero-incidence mirror; I1,2, irises; GL1,2, Glan-Laser Calcite polarizers; AHWP0-2, achromatic half-wave plates; AQWP1,2, achromatic quarter-wave plates; SVWP, a space variant wave plate (Photonic Lattice SWP-808); Delay, a delay stage; POL2, a polarizer; BPF, a bandpass filter (central wavelength, 800 nm; bandwidth, 3 nm).
Fig. 9
Fig. 9 A three-dimensional configuration of optical components in a 4-f SLM system.

Tables (1)

Tables Icon

Table 1 The values of the first ESP S ˜ 1 , l = 1 and the DOP-SD P l = 1 space of generated p = 0 LG RP pulses [Fig. 4(a)] and p = 1 LG RP pulses [Fig. 4(e)].

Equations (11)

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

exp ( i π / 4 ) 2 ( 1 i i 1 ) ( cos 2 θ H 1 sin 2 θ H 1 ) = exp ( + i π / 4 ) sin 2 θ H 1 | s = + 1 + exp ( i π / 4 ) cos 2 θ H 1 | s = 1 ,
( cos 2 θ H 2 sin 2 θ H 2 sin 2 θ H 2 cos 2 θ H 2 ) = R θ H 2 ( 1 0 0 1 ) R θ H 2 = | s = + 1 e i θ H 2 s = 1 | + | s = 1 e + i θ H 2 | s = + 1 ,
R θ = ( cos θ sin θ sin θ cos θ )
( cos 2 q ϕ sin 2 q ϕ sin 2 q ϕ cos 2 q ϕ ) = | s = + 1 ϕ | l = 2 q s = 1 | + | s = 1 ϕ | l = + 2 q s = + 1 | ,
f ( r , ϕ ) = f SLM 2 1 st ( r f 4 / f 3 ) ( sin 2 θ H 1 e i ( 2 θ H 2 π / 4 ) | s = + 1 ϕ | l = 1 + cos 2 θ H 1 e i ( 2 θ H 2 π / 4 ) | s = 1 ϕ | l = + 1 ) .
S ˜ l = 1 E = ( sin 4 θ H 2 sin 4 θ H 1 cos 4 θ H 2 sin 4 θ H 1 cos 4 θ H 1 ) .
ϕ l = 1 = π 2 + 4 θ H 2 ,
θ l = 1 = π 4 θ H 1 .
S ˜ l = 1 E ( ϕ l = 1 , θ l = 1 ) = S ˜ l = 1 E ( ϕ l = 1 + 2 π , θ l = 1 ) = S ˜ l = 1 E ( ϕ l = 1 , θ l = 1 + 2 π ) = S ˜ l = 1 E ( ϕ l = 1 + π , 2 π θ l = 1 ) .
I ( r ) = 1 2 π 0 2 π I ( r , ϕ ) d ϕ ,
I l LG ( r ) = | A | 2 2 p ! π ( p + | l | ) ! ( 2 r 2 w 2 ) | l | { L p | l | ( 2 r 2 w 2 ) } 2 1 w 2 exp ( 2 r 2 w 2 ) ,
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.