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

Theoretical generation of arbitrarily homogeneously 3D spin-orientated optical needles and chains

Open Access Open Access

Abstract

The generation of homogeneously 3D spin-orientated optical needles and optical chains is investigated by focusing the annular collimated beam composed of radially polarized component and azimuthally polarized vortex components with aplanatic focusing systems. Using the Richards-Wolf vector diffraction theory, analytic expressions of focal electric fields and target spin-orientation are given. The results show that the spin-orientation is tunable by changing two parameters of the incident beam while keeping the mean size of the focal spot under 0.5λ. Spin-orientation homogeneity purity is introduced to evaluate spin-orientation homogeneity quantitatively and the results present that spin-orientation homogeneity purity is always beyond 0.996 when spin-orientation varies.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Over the past few years, super-resolved optical fields have always been attracting intensive research interest for their novel properties and potential applications in many fields including particle acceleration [1,2], optical data storage [3,4], optical coherence tomography [5,6], fluorescent imaging [7,8] and so on. Optical needles, also called non-diffraction beams, and optical chains are typically two-dimensional (2D) and three-dimensional (3D) super-resolved optical fields, respectively. These special optical fields are achieved and optimized by focusing modulated incident beams with different focusing systems by various researchers [9–21]. To extend the applications of the super-resolved optical fields, some research is focusing on controlling their spin angular momentum (SAM) [22,23]. The gradient force derived from intensity distribution can only trap particles [24] but the SAM can make the particles rotated. A circularly polarized beam only has longitudinal SAM while transverse SAM can be generated under certain conditions, including plasmonic, evanescent wave and tightly focused beams [22, 25–27]. Among the aforementioned several methods, the simplest way to obtain super-resolved optical fields with controllable SAM is to focus modulated incident beams with a high numerical aperture (NA) focusing system.

In these regards, there are still some obstacles remaining in the tight focusing realm. First, it is hard to realize homogeneous spin-orientation distribution in a 3D region. In the literature [23], only in a plane (such as xz plane) can the spin-orientation distribution be homogeneous. Second, there is no standard to evaluate spin-orientation homogeneity quantitatively. It is important to develop a standard because the spin-orientation cannot be totally homogeneous for an optical field except a few special occasions such as a circularly polarized plane wave. Third, the obtained 3D super-resolved optical field is usually a single focal spot because it is difficult to make multiple spots in different positions have the same spin-orientation. To solve these problems, we choose a thin annular beam as the incident beam to obtain an optical needle (non-diffraction beam) so that the longitudinal distribution and the transverse distribution of the focal field can be separated. In this way, the spin-orientation will be homogeneous in a 3D region as long as it is homogeneous in a plane which is perpendicular to the propagating direction. With the help of a 4π focusing system, a needle can be tailored to a chain [28]. Besides, we propose a conception called “spin-orientated homogeneity purity” (SOHP) to evaluate spin-orientation homogeneity quantitatively. SOHP of the obtained super-resolved optical fields in this paper are all beyond 0.996.

2. Theory and configuration

Figure 1 shows schematic diagrams of tight focusing systems. Figure 1(a) is a parabolic mirror, which can be utilized to generate optical needles [15]. Figure 1(b) is a 4 π system composed of two confocal aplanatic lenses, which can be used to generate optical chains [28]. A parabolic mirror’s NA can reach 1 (i.e. θ0 = 90°), which means it can generate a smaller focal spot than an aplanatic lens [15].

 figure: Fig. 1

Fig. 1 Schematic diagrams of tight focusing systems. (a) a parabolic mirror; (b) a 4π system. The incident beams are collimated modulated thin annular beams. These systems are azimuthally symmetric.

Download Full Size | PDF

In general, a collimated incident beam can be express as

Ei(θ,φ)=m=+fm(θ)exp(imφ)eρ+m=+gm(θ)exp(imφ)eφ,
where fm, gm are complex amplitude factors of radially and azimuthally polarized components when the topological charge is m, respectively. When the incident beam is thin, fm, gm can be regarded as constants. Such annular light can be generated by an axicon and a lens [15]. Considering a single aplanatic focusing system such as Fig. 1(a), according to the Richards-Wolf vector diffraction theory [29], the electric field near the focus can be expressed as
E(r,φ,z)=Aθ0Δθ2θ0+Δθ2P(θ)sinθeikzcosθm=imeimϕ×[icosθfm(eiϕJm1+eiϕJm+1)ex+gm(eiϕJm1+eiϕJm+1)excosθfm(eiϕJm1+eiϕJm+1)ey+igm(eiϕJm1eiϕJm+1)ey2sinθfmJmez]dθ,
where A is a constant related to wavelength λ, k = 2π/λ is wave number, θ0 represents the angular position of the central ray of the incident beam and Δθ is the angular thickness of the incident beam as shown in Fig. 1(a). P(θ) is apodization factor, obtained from energy conservation. For a parabolic mirror, P(θ) = 2/(1 − cos θ); for an aplanatic lens, P(θ)=cos(θ). Jn = Jn(kr sin θ), where Jn(·) denotes the nth-order Bessel function of the first kind. e⃗x, e⃗y and e⃗z are Cartesian unit vectors. It is easy to verify that Eq. (2) fulfills the equations ∇⃗ · E⃗ = 0 and (∇⃗2 + k2)E⃗ = 0.

With some approximations when Δθ << θ0 [28], Eq. (2) can be expressed as

E(r,φ,z)=AP(θ0)2sin(kzsinθ0sinΔθ2)exp(ikzcosθ0cosΔθ2)kzm=imeimϕ×[icosθ0fm(eiϕJm1+eiϕJm+1)ex+gm(eiϕJm1+eiϕJm+1)excosθ0fm(eiϕJm1+eiϕJm+1)ey+igm(eiϕJm1eiϕJm+1)ey2sinθ0fmJmez],
where Jn = Jn(kr sin θ0). Equation (3) shows that the longitudinal distribution and the transverse distribution of the focused field are separated and a non-diffraction beam is obtained. To solve the equation |E(r,ϕ,z)|2=12|E(r,ϕ,0)|2, it is easy to gain that the longitudinal full width at half maximum (LFWHM) or depth of focus is about 0.8743λsinθ0Δθ [28].

When r = 0, only J0 ≠ 0. Jn of non-zero orders represent hollow fields and will make resolution worse. To achieve a super-resolved optical field, only J0 is desired. Hence, m should only be 0 or ±1. So only f−1, f0, f1, g−1 and g1 can be retained. A larger sin θ0 (θ0 is near 90°) will cause a smaller focal spot [15, 18, 28], so a large sin θ0 will be adopted in this paper. In the extreme case θ0 = 90°, f1 and f−1 will not generate J0 component in Eq. (3), so we choose f1, f−1 = 0. Now there are still six degrees of freedom (because f0, g−1 and g1 are complex numbers), but the direction of SAM only has two degrees of freedom. We fix the phase and add an amplitude restriction and then f0, g−1 and g1 can be expressed as

f0=cosα/(2sinθ0)g1=isinαcosβg1=sinαsinβ,
where α ∈ [0, π], β ∈ [0, 2π]. The amplitude weighting factors of three components of the incident beam can be controlled by α and β. So f0, g−1 and g1 only have two degrees of freedom, which can be used to control spin-orientation. The factor (2 sin θ0)−1 is for simplifying the calculation. Such a beam can be realized by a vectorial optical field generator, and α and β can be adjusted by loading the proper phase patterns onto spatial light modulators [30]. Equation (4) represents the input beams which can generate optical needles. Substituting Eq. (4) into Eq. (3), the electric field can be expressed as
E(r,φ,z)=AP(θ0)2sin(kzsinθ0sinΔθ2)exp(ikzcosθ0cosΔθ2)kz×[[J0sinαeiβ+iJ1cosαcosϕcotθ0+J2sinα(isinβei2ϕcosβei2ϕ)]ex[iJ0sinαeiβ+iJ1cosαsinϕcotθ0+J2sinα(sinβei2ϕicosβei2ϕ)]ey[J0cosα]ez].

According to the reference [31], the SAM density can be given explicitly as

sIm(E*×E),
where E⃗* is the complex conjugation of E⃗, and Im(·) denotes the imaginary part. To describe the direction of the SAM density s⃗, we define a unit vector γ⃗, which is expressed as
γ=sss(s0).

It is not easy to calculate an analytic result of γ at an arbitrary point. And the result is too complex to give a distinct physical picture. But it is easy to calculate γ⃗ at the origin r = 0, z = 0. According to Eqs. (5)(7), γ⃗(0, 0, 0) is

γ(0,0,0)=[cosαcosβcos2α+sin2αcos22βexcosαsinβcos2α+sin2αcos22βeysinαcos2βcos2α+sin2αcos22βez](sinα0).

When sin α = 0, s⃗ is 0⃗, which will not be discussed in this paper.

γ⃗(0, 0, 0) can be regarded as a “theory value” or “target value”. It is controllable by adjusting α and β. To evaluate the spin-orientation homogeneity quantitatively, we define SOHP as

ηS(a)=SsadSS|s||a|dS(|a|0)ηV(a)=VsadVV|s||a|dV(|a|0),
where a⃗ is a constant vector, ηS and ηV denote SOHP with respect to a⃗ in a 2D region and a 3D region, respectively. When s⃗ is always parallel to a⃗ in the integral domain, η = 1. When s⃗ is always inversely parallel to a⃗, η = −1. When s⃗ is always perpendicular to a⃗, η = 0. When s⃗ is parallel to a⃗ in half space of the integral region and antiparallel to a⃗ in the rest half space and |s⃗| is constant, η = 0. Only when η is close to 1, the homogeneity of spin-orientation is satisfactory. These basic properties indicate that this definition is logical. In the following, a⃗ = γ⃗(0, 0, 0).

3. Numerical simulation and discussion

3.1. Optical needle

The angular position θ0 and the angular width Δθ of the incident beam can influence LFWHM and the transverse full width at half maximum (TFWHM) [28]. Equation (3) shows that Δθ has little influence on electric field in a transverse profile. So it will be hardly discussed in the following. In this subsection, θ0 = 90° and Δθ = 0.01.

Figure 2 shows an axial profile at r = 0 and three lateral profiles at z = 0 of an optical needle by two methods with α=π4 and β=π4. The solid blue lines are generated from Eqs. (2) and (4) by numerical integration; the dashed red lines are produced by the approximation Eq. (5). These two methods lead to consistent results, which implies that the approximation is accurate. Figures 3(a)–(c) are coherent with Figs. 2(b)–(d), respectively. They display that the longitudinal distributions are similar while the transverse distributions are different when φ is different. The former is from the diffraction of the thin annular incident beam, which is very similar to Fraunhofer diffraction at the slit [32]. The latter is due to the interference of two azimuthally polarized components with different topological charges, which will lead to TFWHM as a function of φ. Equation (5) can be re-expressed as E⃗(r, φ, z) ≈ F(z) G⃗(r, φ). So |F(z)|2 and |G⃗(r, φ)|2 represent longitudinal and lateral distribution of electric energy density, respectively. With the major premise θ0 = 90°, |G⃗(r, φ)|2 will be independent of φ only when sin2 α sin 2β = 0 (i.e. g1g−1 = 0).

 figure: Fig. 2

Fig. 2 Distribution of the normalized electric energy density along z-axis and r-axis with θ0=90°=π2, Δθ = 0.01, α=π4 and β=π4. Figure 2(a) is the longitudinal profile and Figs. 2(b)–(d) are the transverse profiles with φ = 0, π2, π4, respectively. The solid blue lines are generated from Eqs. (2) and (4); the dashed red lines are produced by Eq. (5). This figure shows that numerical integration and the approximation by Eq. (5) lead to consistent results for optical needles.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Distribution of the normalized electric energy density in rz plane with α=π4 and β=π4. Figures 3(a)–(c) are the profiles with φ = 0, π2, π4, respectively. These images correspond to Eq. (5).

Download Full Size | PDF

Solving the equation |E(r,ϕ,z)|2=12|E(0,ϕ,z)|2, TFWHMs as a function of φ are shown in Figs. 4(a1) and (a2) with β=π4, 0, , respectively. Figures 4(b1) and (b2) are relevant transverse profiles to Figs. 4(a1) and (a2). Figures 4(a2) and (b2) show TFWHM=0.366λ is a constant, owing to g1 = 0. On the contrary, when β=π4 or β=34π (i.e.|g1|/|g−1| = 1), the interference between two incident azimuthally polarized components reaches the maximum, which results in the maximal deviation of TFWHM when φ varies. In spite of the deviation, TFWHM at φ = 0 or π2 is a constant when β changes. Besides, Fig. 4(a1) indicates that TFWHM at φ = 0 or π2 is a mean value. Therefore, it is suitable to use TFWHM at φ = 0 or π2 as a mean TFWHM.

 figure: Fig. 4

Fig. 4 (a1) and (a2) display TFWHM as a function of φ with β=π4, 0, respectively. (b1) and (b2) are transverse profiles of the normalized electric energy at z = 0 with β=π4 and 0, respectively. These images are generated from Eq. (5) with α=π4.

Download Full Size | PDF

The parameters α and β are used to control the spin-orientation. But before the spin-orientation is discussed, the size of the focal spot should be given. Figure 5 shows the mean TFWHM as a function of α and β. It implies that the mean TFWHM is independent of β and will increase when sin α increases. This is because the radially polarized component of the incident beam only causes J0 while the azimuthally polarized components cause J0 and J2. Hence, the mean TFWHM reaches the maximum when the radially polarized component vanishes (i.e. α=π2), and drops to the minimum when the azimuthally polarized components vanish (i.e. α = 0, π). As shown in Fig. 5, the mean TFWHM varies from 0.36λ to 0.37λ when α and β vary. To make the calculation simpler, we will choose 0.366λ as the diameter of the focal spot in the following in this subsection.

 figure: Fig. 5

Fig. 5 Mean TFWHM as a function of α and β. This figure is generated from Eq. (5).

Download Full Size | PDF

Figure 6 shows distribution of SAM density in two extreme cases. For Figs. 6(a1) and (b1), the parameters α=π4, β=π4 cause γ(0,0,0)=22ex22ey, which means the spin-orientation at the origin is purely transverse. On the contrary, the parameters α=π2, β = 0 cause γ⃗(0, 0, 0) = e⃗z in Figs. 6(a2) and (b2), which leads to purely longitudinal spin-orientation at the origin. Figures 6(a1) and (a2) both display almost perfect spin-orientated homogeneity in the focal region. Though the 3D pictures are visual, the results are rough and usually depend on the angle of the observation. Their 2D projections Figs. 6(b1) and (b2) can help to imagine their 3D spin-orientation.

 figure: Fig. 6

Fig. 6 Distribution of SAM density and related 2D projections. For Figs. 6(a1) and (b1), α=β=π4. For Figs. 6(a2) and (b2), α=π2, β = 0. These images are generated from Eqs. (5) and (6).

Download Full Size | PDF

To evaluate spin-orientated homogeneity quantitatively, Fig. 7 shows that SOHP is always beyond 0.996 when α and β vary. The integral domain is r ≤ 0.183λ. It should be noted that in Fig. 7 α ∈ [0.01, π − 0.01] instead of α ∈ [0, π]. This is because the target value will be meaningless when sin α = 0. When α=π2, the radially polarized component will vanish, leading to the focal field with zero longitudinal component. Hence, purely longitudinal spin-orientation will be obtained in the focal region and SOHP will equal to 1 as shown in Fig. 7.

 figure: Fig. 7

Fig. 7 SOHP as a function of α and β. This figure is produced by Eqs. (5)(9). The integral domain is r ≤ 0.183λ.

Download Full Size | PDF

3.2. Optical chain

When a 4π confocal system replaces a single focusing system, an optical chain will appear. The angular position θ0 can influence the aspect ratio of every spot of the chain [28]. In order to obtain a chain with spherical spot, the aspect ratio should equal 1. According to the mathematical method in [28], θ0 = 55.66° = 0.9715 is used in this section. Besides, Δθ = 0.01 is still used.

The field near the confocal spot can be expressed as

Etotal(r,ϕ,z)=EL(r,ϕ,z)+ER(r,ϕ,z),
where E⃗L and E⃗R denote the beam from the left and the right, respectively. The left input beam is the same as the input beam to generate optical needles. As for the right one, its azimuthally components are the same as the left while its radially polarized component should have an opposite polarization. Then E⃗R can be expressed as
ER(r,ϕ,z)=Aθ0Δθ2θ0+Δθ2dθP(θ)sinθeikzcosθ×[[J0sinαeiβiJ1cosαcosϕcosθ/sinθ0+J2sinα(isinβei2ϕcosβei2ϕ)]ex[iJ0sinαeiβiJ1cosαsinϕcosθ/sinθ0+J2sinα(sinβei2ϕicosβei2ϕ)]ey[J0cosαsinθ/sinθ0]ez]AP(θ0)2sin(kzsinθ0sinΔθ2)exp(ikzcosθ0cosΔθ2)kz[[J0sinαeiβiJ1cosαcosϕcotθ0+J2sinα(isinβei2ϕcosβei2ϕ)]ex[iJ0sinαeiβiJ1cosαsinϕcotθ0+J2sinα(sinβei2ϕicosβei2ϕ)]ey[J0cosα]ez].
It is easy to verify that Eq. (11) also respects the equations ∇⃗ · E⃗R = 0 and (∇⃗2 + k2)E⃗R = 0 before the approximation.

Figure 8 shows the profiles of the chain in xz plane (a) and yz plane (b). The figure displays 17 spherical spots, in which the maximum value of the rightmost/leftmost spot is 0.9886. The number of bright spots is inversely proportionate to Δθ [28]. When Δθ = 0.1, there are only 3∼5 bright spots. Mean TFWHM and LFWHM of the chain is about λ4cosθ0=0.043λ. The focal volume is estimated to be 0.046λ3, which is far samller than the result 0.222λ3 in [22]. To make the calculation simpler, we will choose 0.444λ as the diameter of the focal spot in this section.

 figure: Fig. 8

Fig. 8 Distribution of the normalized electric energy density in xz plane and yz plane with α=π4 and β=π4. These images are generated from Eqs. (5), (10) and (11).

Download Full Size | PDF

The existence of J1 leads to the fact that the longitudinal distribution and the transverse distribution of the confocal electric field cannot be separated completely. Therefore, SOHP in a 3D region should be considered.

Figure 9 shows SOHPs for the spots of the chain. The integral domains of Figs. 9(a)–(c) are r2 + z2 ≤ 0.2222λ2, r2 + (zz0)2 ≤ 0.2222λ2 and r2 + (z − 8z0)2 ≤ 0.2222λ2, where z0=λ2cosθ0. They represent the first, the second and the ninth spot with respect to the origin. Like the occasion of the needles, α varies from 0.01 to π-0.01 and SOHPs are always beyond 0.996.

 figure: Fig. 9

Fig. 9 SOHP as a function of α and β. This figure is produced by Eqs. (5)(11). The integral domains of (a)–(c) are the first, the second and the ninth spot with respect to the origin in the chain. The related mathematical expressions are r2 + z2 ≤ 0.2222λ2, r2 + (zz0)2 ≤ 0.2222λ2 and r2 + (z − 8z0)2 ≤ 0.2222λ2, respectively.

Download Full Size | PDF

4. Conclusion

In conclusion, we theoretically investigate the generation of homogeneously 3D spin-orientated optical needles and optical chains. The spin-orientation are continuously tunable by adjusting two parameters of the incident beam. Meanwhile, SOHP can keep beyond 0.996. This research can be also used to generate magnetization needles and magnetization chains with controllable orientation, because the magnetization field induced by the inverse Faraday effect is proportional to SAM density of the electric field [28].

References

1. J. Fontana and R. Pantell, “A high-energy, laser accelerator for electrons using the inverse cherenkov effect,” J. Appl. Phys. 54, 4285–4288 (1983). [CrossRef]  

2. R. D. Romea and W. D. Kimura, “Modeling of inverse čerenkov laser acceleration with axicon laser-beam focusing,” Phys. Rev. D 42, 1807–1818 (1990). [CrossRef]  

3. Y. Zhang and J. Bai, “Improving the recording ability of a near-field optical storage system by higher-order radially polarized beams,” Opt. Express 17, 3698–3706 (2009). [CrossRef]   [PubMed]  

4. H. Wang, G. Yuan, W. Tan, L. Shi, and C. T. Chong, “Spot size and depth of focus in optical data storage system,” Opt. Eng. 46, 065201 (2007). [CrossRef]  

5. R. Leitgeb, M. Villiger, A. Bachmann, L. Steinmann, and T. Lasser, “Extended focus depth for fourier domain optical coherence microscopy,” Opt. Lett. 31, 2450–2452 (2006). [CrossRef]  

6. L. Liu, C. Liu, W. C. Howe, C. Sheppard, and N. Chen, “Binary-phase spatial filter for real-time swept-source optical coherence microscopy,” Opt. Lett. 32, 2375–2377 (2007). [CrossRef]  

7. F. O. Fahrbach, V. Gurchenkov, K. Alessandri, P. Nassoy, and A. Rohrbach, “Light-sheet microscopy in thick media using scanned bessel beams and two-photon fluorescence excitation,” Opt. Express 21, 13824–13839 (2013). [CrossRef]   [PubMed]  

8. G. Thériault, Y. De Koninck, and N. McCarthy, “Extended depth of field microscopy for rapid volumetric two-photon imaging,” Opt. Express 21, 10095–10104 (2013). [CrossRef]   [PubMed]  

9. J. Fu, Y. Wang, and P. Chen, “Generation of sub-half-wavelength non-diffracting beams with tunable depth of focus by focusing azimuthally polarized beams,” Appl. Phys. B 123, 214 (2017). [CrossRef]  

10. H. Wang, L. Shi, B. Lukyanchuk, C. Sheppard, and C. T. Chong, “Creation of a needle of longitudinally polarized light in vacuum using binary optics,” Nat. Photonics 2, 501–505 (2008). [CrossRef]  

11. C.-C. Sun and C.-K. Liu, “Ultrasmall focusing spot with a long depth of focus based on polarization and phase modulation,” Opt. Lett. 28, 99–101 (2003). [CrossRef]   [PubMed]  

12. X. Hao, C. Kuang, T. Wang, and X. Liu, “Phase encoding for sharper focus of the azimuthally polarized beam,” Opt. Lett. 35, 3928–3930 (2010). [CrossRef]   [PubMed]  

13. J. Lin, K. Yin, Y. Li, and J. Tan, “Achievement of longitudinally polarized focusing with long focal depth by amplitude modulation,” Opt. Lett. 36, 1185–1187 (2011). [CrossRef]   [PubMed]  

14. P. Suresh, C. Mariyal, K. Rajesh, T. Pillai, and Z. Jaroszewicz, “Generation of a strong uniform transversely polarized nondiffracting beam using a high-numerical-aperture lens axicon with a binary phase mask,” Appl. Opt. 52, 849–853 (2013). [CrossRef]   [PubMed]  

15. H. Dehez, A. April, and M. Piché, “Needles of longitudinally polarized light: guidelines for minimum spot size and tunable axial extent,” Opt. Express 20, 14891–14905 (2012). [CrossRef]   [PubMed]  

16. M. Zhu, Q. Cao, and H. Gao, “Creation of a 50,000 λ long needle-like field with 0.36 λ width,” J. Opt. Soc. Am. A 31, 500–504 (2014). [CrossRef]  

17. D. Panneton, G. St-Onge, M. Piché, and S. Thibault, “Needles of light produced with a spherical mirror,” Opt. Lett. 40, 419–422 (2015). [CrossRef]   [PubMed]  

18. L. Hang, Y. Wang, and P. Chen, “Needles of light produced with a quasi-parabolic mirror,” J. Opt. Soc. Am. A 35, 174–178 (2018). [CrossRef]  

19. Y. Zhao, Q. Zhan, Y. Zhang, and Y.-P. Li, “Creation of a three-dimensional optical chain for controllable particle delivery,” Opt. Lett. 30, 848–850 (2005). [CrossRef]   [PubMed]  

20. J. Wang, W. Chen, and Q. Zhan, “Creation of uniform three-dimensional optical chain through tight focusing of space-variant polarized beams,” J. Opt. 14, 055004 (2012). [CrossRef]  

21. J. Cao, Q. Chen, and H. Guo, “Creation of a controllable three dimensional optical chain by tem01 mode radially polarized laguerre–gaussian beam,” Optik-International J. for Light. Electron Opt. 124, 2033–2036 (2013). [CrossRef]  

22. J. Chen, C. Wan, L. Kong, and Q. Zhan, “Experimental generation of complex optical fields for diffraction limited optical focus with purely transverse spin angular momentum,” Opt. Express 25, 8966–8974 (2017). [CrossRef]   [PubMed]  

23. W. Yan, Z. Nie, X. Liu, X. Zhang, Y. Wang, and Y. Song, “Arbitrarily spin-orientated and super-resolved focal spot,” Opt. Lett. 43, 3826–3829 (2018). [CrossRef]   [PubMed]  

24. Q. Zhan, “Trapping metallic rayleigh particles with radial polarization,” Opt. Express 12, 3377–3382 (2004). [CrossRef]   [PubMed]  

25. A. Aiello, P. Banzer, M. Neugebauer, and G. Leuchs, “From transverse angular momentum to photonic wheels,” Nat. Photonics 9, 789–795 (2015). [CrossRef]  

26. P. Banzer, M. Neugebauer, A. Aiello, C. Marquardt, N. Lindlein, T. Bauer, and G. Leuchs, “The photonic wheel: demonstration of a state of light with purely transverse angular momentum,” arXiv preprint arXiv:1210.1772 (2012).

27. L. Hang, J. Fu, X. Yu, Y. Wang, and P. Chen, “Generation of a sub-half-wavelength focal spot with purely transverse spin angular momentum,” Appl. Phys. B 123, 266 (2017). [CrossRef]  

28. L. Hang, K. Huang, J. Fu, Y. Wang, and P. Chen, “Tunable magnetization chains induced with annular parabolic mirrors,” IEEE Photonics J. 10, 1–7 (2018). [CrossRef]  

29. B. Richards and E. Wolf, “Electromagnetic diffraction in optical systems, ii. structure of the image field in an aplanatic system,” Proc. R. Soc. Lond. A 253, 358–379 (1959). [CrossRef]  

30. W. Han, Y. Yang, W. Cheng, and Q. Zhan, “Vectorial optical field generator for the creation of arbitrarily complex fields,” Opt. Express 21, 20692–20706 (2013). [CrossRef]   [PubMed]  

31. M. Neugebauer, T. Bauer, A. Aiello, and P. Banzer, “Measuring the transverse spin density of light,” Phys. Rev. Lett. 114, 063901 (2015). [CrossRef]   [PubMed]  

32. M. Born and E. Wolf, Principle of Optics (Cambridge University, 1999). 436–439.

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 Schematic diagrams of tight focusing systems. (a) a parabolic mirror; (b) a 4π system. The incident beams are collimated modulated thin annular beams. These systems are azimuthally symmetric.
Fig. 2
Fig. 2 Distribution of the normalized electric energy density along z-axis and r-axis with θ 0 = 90 ° = π 2, Δθ = 0.01, α = π 4 and β = π 4. Figure 2(a) is the longitudinal profile and Figs. 2(b)–(d) are the transverse profiles with φ = 0, π 2, π 4, respectively. The solid blue lines are generated from Eqs. (2) and (4); the dashed red lines are produced by Eq. (5). This figure shows that numerical integration and the approximation by Eq. (5) lead to consistent results for optical needles.
Fig. 3
Fig. 3 Distribution of the normalized electric energy density in rz plane with α = π 4 and β = π 4. Figures 3(a)–(c) are the profiles with φ = 0, π 2, π 4, respectively. These images correspond to Eq. (5).
Fig. 4
Fig. 4 (a1) and (a2) display TFWHM as a function of φ with β = π 4, 0, respectively. (b1) and (b2) are transverse profiles of the normalized electric energy at z = 0 with β = π 4 and 0, respectively. These images are generated from Eq. (5) with α = π 4.
Fig. 5
Fig. 5 Mean TFWHM as a function of α and β. This figure is generated from Eq. (5).
Fig. 6
Fig. 6 Distribution of SAM density and related 2D projections. For Figs. 6(a1) and (b1), α = β = π 4. For Figs. 6(a2) and (b2), α = π 2, β = 0. These images are generated from Eqs. (5) and (6).
Fig. 7
Fig. 7 SOHP as a function of α and β. This figure is produced by Eqs. (5)(9). The integral domain is r ≤ 0.183λ.
Fig. 8
Fig. 8 Distribution of the normalized electric energy density in xz plane and yz plane with α = π 4 and β = π 4. These images are generated from Eqs. (5), (10) and (11).
Fig. 9
Fig. 9 SOHP as a function of α and β. This figure is produced by Eqs. (5)(11). The integral domains of (a)–(c) are the first, the second and the ninth spot with respect to the origin in the chain. The related mathematical expressions are r2 + z2 ≤ 0.2222λ2, r2 + (zz0)2 ≤ 0.2222λ2 and r2 + (z − 8z0)2 ≤ 0.2222λ2, respectively.

Equations (11)

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

E i ( θ , φ ) = m = + f m ( θ ) exp ( i m φ ) e ρ + m = + g m ( θ ) exp ( i m φ ) e φ ,
E ( r , φ , z ) = A θ 0 Δ θ 2 θ 0 + Δ θ 2 P ( θ ) sin θ e i k z cos θ m = i m e i m ϕ × [ i cos θ f m ( e i ϕ J m 1 + e i ϕ J m + 1 ) e x + g m ( e i ϕ J m 1 + e i ϕ J m + 1 ) e x cos θ f m ( e i ϕ J m 1 + e i ϕ J m + 1 ) e y + i g m ( e i ϕ J m 1 e i ϕ J m + 1 ) e y 2 sin θ f m J m e z ] d θ ,
E ( r , φ , z ) = AP ( θ 0 ) 2 sin ( k z sin θ 0 sin Δ θ 2 ) exp ( i k z cos θ 0 cos Δ θ 2 ) k z m = i m e i m ϕ × [ i cos θ 0 f m ( e i ϕ J m 1 + e i ϕ J m + 1 ) e x + g m ( e i ϕ J m 1 + e i ϕ J m + 1 ) e x cos θ 0 f m ( e i ϕ J m 1 + e i ϕ J m + 1 ) e y + i g m ( e i ϕ J m 1 e i ϕ J m + 1 ) e y 2 sin θ 0 f m J m e z ] ,
f 0 = cos α / ( 2 sin θ 0 ) g 1 = i sin α cos β g 1 = sin α sin β ,
E ( r , φ , z ) = AP ( θ 0 ) 2 sin ( k z sin θ 0 sin Δ θ 2 ) exp ( i k z cos θ 0 cos Δ θ 2 ) k z × [ [ J 0 sin α e i β + i J 1 cos α cos ϕ cot θ 0 + J 2 sin α ( i sin β e i 2 ϕ cos β e i 2 ϕ ) ] e x [ i J 0 sin α e i β + i J 1 cos α sin ϕ cot θ 0 + J 2 sin α ( sin β e i 2 ϕ i cos β e i 2 ϕ ) ] e y [ J 0 cos α ] e z ] .
s Im ( E * × E ) ,
γ = s s s ( s 0 ) .
γ ( 0 , 0 , 0 ) = [ cos α cos β cos 2 α + sin 2 α cos 2 2 β e x cos α sin β cos 2 α + sin 2 α cos 2 2 β e y sin α cos 2 β cos 2 α + sin 2 α cos 2 2 β e z ] ( sin α 0 ) .
η S ( a ) = S s a d S S | s | | a | d S ( | a | 0 ) η V ( a ) = V s a d V V | s | | a | d V ( | a | 0 ) ,
E total ( r , ϕ , z ) = E L ( r , ϕ , z ) + E R ( r , ϕ , z ) ,
E R ( r , ϕ , z ) = A θ 0 Δ θ 2 θ 0 + Δ θ 2 d θ P ( θ ) sin θ e i k z cos θ × [ [ J 0 sin α e i β i J 1 cos α cos ϕ cos θ / sin θ 0 + J 2 sin α ( i sin β e i 2 ϕ cos β e i 2 ϕ ) ] e x [ i J 0 sin α e i β i J 1 cos α sin ϕ cos θ / sin θ 0 + J 2 sin α ( sin β e i 2 ϕ i cos β e i 2 ϕ ) ] e y [ J 0 cos α sin θ / sin θ 0 ] e z ] AP ( θ 0 ) 2 sin ( k z sin θ 0 sin Δ θ 2 ) exp ( i k z cos θ 0 cos Δ θ 2 ) k z [ [ J 0 sin α e i β i J 1 cos α cos ϕ cot θ 0 + J 2 sin α ( i sin β e i 2 ϕ cos β e i 2 ϕ ) ] e x [ i J 0 sin α e i β i J 1 cos α sin ϕ cot θ 0 + J 2 sin α ( sin β e i 2 ϕ i cos β e i 2 ϕ ) ] e y [ J 0 cos α ] e z ] .
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.