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

Electro-optical coupling of a circular Airy beam in a uniaxial crystal

Open Access Open Access

Abstract

This paper investigates theoretically and numerically on the electro-optical coupling (EOC) for a circular Airy beam (CAB) propagating along the optical axis of a uniaxial crystal after deducing the wave coupling equations of EOC. For a circularly polarized incident CAB, EOC can be used to generate vortex beam by coupling the incident left-handed component into the right-handed vortex component with a vortex topological charge of 2. What’s more, EOC plays important role in enhancing or suppressing the abrupt autofocusing, the most important property of CABs, for both left-handed and right-handed components. Near the focal plane, EOC can result in electrically controllable optical “needle” and “cage”, which shall be interesting in micromanipulation. In addition, EOC can influence or even forbid the exchange between spin angular momentum (SAM) and orbit angular momentum (OAM). For a linearly polarized incident CAB, its two Cartesian field components of the beam cannot only couple their powers to each other, but also lead to the changes of the intensity pattern and polarization distributions. The polarization state becomes spatially inhomogeneous, and possesses vortex phase with a topological charge of 2 during propagation. EOC presents a new way to control an Airy beam fast and efficiently.

© 2017 Optical Society of America

1. Introduction

A circular Airy beam (CAB), i.e., a circularly symmetric beam with an Airy radial profile, has stirred wide interest in the past few years because of the abruptly autofocusing property [1–5]. The abruptly autofocusing property allows CABs to focus their intensity by an increase of orders of magnitude immediately before a target, which is applicable to biomedical treatment [6] and nonlinear optical processes [7]. CABs can also be used to trap and guide micro-particles [8,9] and to form special beams such as “bottle” beams [10], optical “needles” [11], optical “cages” [11], optical dark “channels” [11], and optical “bullets” [7]. Therefore, controlling the properties of CABs is a powerful way to develop the new applications of CABs such as modulating and switching. 1D and 2D Airy beams can be generated from a broad Gaussian beam through a Fourier Transform (FT) for a imposed cubic phase, which can be realized by several methods such as using spatial light modulator (SLM) and binary phase mask [12,13]. Fortunately, a CAB can also be generated in Fourier space by using a highly accurate closed-form expression for the FT of a CAB [10]. By modifying the Fourier spectrum of CABs, one can enhance the abruptly autofocusing property [14] and produce some interesting beams, such as the optical “needle” and optical “bottle” [10,15]. Recent research has shown that the propagation trajectory of CABs can be controlled and manipulated by potentials [16–18]. But there is still lack of a way in controlling the propagation properties of Airy beam. As we know, linear electro-optic (LEO) effect is extensively used to control the properties of light, such as polarization, phase and intensity. Therefore, it is expected that LEO effect can play its role in controlling a CAB.

On the other hand, most wave coupling theories describing the LEO effect are based on the plane-wave model [19,20]. This treatment is valid only when the crystal length is much shorter than the confocal parameter of the light beam such that the beam width keeps approximately constant within the crystal. This condition is not fulfilled in the Airy beam, where the beam is often micron sized or even smaller. In this case, the transverse distribution of the optical field should be taken into account. In fact, LEO effect for a Gaussian or Gaussian-like beam considering transverse distribution in a uniaxial crystal [21–25] and an optical superlattice [26] have been numerically or experimentally studied. In addition, a simpler approach that addresses paraxial propagation in a uniaxial crystal by FT was recently developed by Ciattoni et al [27–29]. With the inspiration of [27], Airy beams [30], Airy vortex beams [31], Airy Gaussian beams [32], Airy Gaussian vortex beams [33], and CABs [34] propagating through uniaxial crystals have been investigated. However, to our best knowledge, the LEO effect for CABs has never been reported. In the remainder of this paper, we will focus on the EOC of a CAB in uniaxial crystals. We present the evolution of a CAB with LEO effect considering the transverse distribution by means of 2D FT. We find that the abrupt autofocusing of the CAB can be enhanced or weakened by EOC. The exchange between SAM and OAM, the polarization state and intensity pattern can also be controlled electrically. Since LEO effect usually has very short response time (tens of nanoseconds), we believe that our work presents an efficient and fast way to control the propagation characteristics of a CAB, which has potential application in micromanipulation.

2. Theory

Supposing that all other second-order nonlinear effects are so weak because of phase mismatch that only the LEO effect must be considered, we can derive from Maxwell’s equations that

2E(r)[E(r)]+k02εE(r)+μ0ω2PEO(r)=0,
where k0 = ω/c, ω is the angular frequency of light, μ0 and c are the magnetic susceptibility and the speed of light in vacuum, ε is the relative dielectric tensor, given by ε=diag(no2,no2,ne2) for uniaxial crystals, PEO=ε0χ(2)(ω,ω,0):EE0 is the nonlinear polarization corresponding to the LEO effect, E0 is the external electric field, and χ(2)(ω,ω,0) is the second-order susceptibility tensor of the LEO effect. To observe the EOC effect, the incident light propagates along the optical axis (z-axis) for the phase matching. To obtain the exact solution of Eq. (1), we expand the optical fields E and PEO into the two-dimensional Fourier integrals
E(r)=E(r,z)=d2k2exp(ikr)E˜(k,z),
and
P(r)=P(r,z)=d2k2exp(ikr)P˜(k,z).
Substituting Eqs. (2) and (3) into Eq. (1) and inverting the obtained Fourier integrals, we obtain [21, 27]
(2z2+k02no2ky2)E˜x+kxkyE˜yikxzE˜z+μ0ω2P˜xEO=0,
(2z2+k02no2kx2)E˜y+kxkyE˜xikyzE˜z+μ0ω2P˜yEO=0,
(k02ne2k2)E˜zikxzE˜xikyzE˜y+μ0ω2P˜zEO=0.
Here, E˜(k,z)=(2π)2d2rexp(ikr)E(r,z), and P˜iEO=ε0jkχijk(2)(ω,ω,0)E˜j(ω)E˜k(0) ( denotes the convolution operator). The solution expressions of Eq. (4) depend on the specific expression of P˜iEO because P˜iEO may involve one, two or all of E˜i(i = x, y, z). For the uniaxial electro-optical SBN crystal, no = 2.3117, ne = 2.2987, and the non-vanishing electro-optic coefficients γ13 = γ23 = 67, γ33 = 1340, and γ42 = γ51 = 42 (in pm/V) at 632.8 nm [35]. To take advantage of the largest electro-optic coefficient γ33, we apply the external electric field along the z-axis. In this case, we have P˜xEO=ε0n02ne2γ13E0E˜x, P˜yEO=ε0n02ne2γ13E0E˜y and P˜zEO=ε0ne4γ33E0E˜z. We can see that P˜iEO only involves the optical field component E˜i. Therefore, the wave coupling equations shown in Eq. (4) can be simplified as

2z2E˜xikxzE˜z+kxkyE˜y+[k02n02(1ne2γ13E0)ky2]E˜x=0,
2z2E˜yikyzE˜z+kxkyE˜x+[k02n02(1ne2γ13E0)kx2]E˜y=0,
[k02ne2(1ne2γ33E0)k2]E˜zikxzE˜xikyzE˜y=0.

Because the initial boundary of the CAB is circularly symmetric, i.e., E(r,0)=E(r,0), the Fourier transform can be expressed as a zeroth-order Hankel transform,

E˜(k)=12π0drrJ0(kr)E(r,0),
where J0 stands for the zeroth order Bessel function of the first kind. The complex amplitude E(r) of the CAB at the input plane can be expressed as
E(r,0)=CAi(r0rw)exp(ar0rw),
where C is a constant, while Ai, r0, r, w and a stand for the Airy function, initial radius of the main ring, radial distance, radially scaled coefficient, and decay parameter, respectively. By relying on a suitable plane wave angular spectrum representation of the beam, the closed-form approximation of the CAB spatial FT can be expressed as [10]
E˜(k)=Cw2(r0w+k2w2)exp(ak2w2)3kr0+k3w33kr0+3k3w3J0(kr0+k3w33)
where k is the radial spatial frequency.

On the other hand, Ciattoni et al have developed a simple approach to deal with paraxial propagation along the optical axis of a uniaxial crystal [27–29]. They directly derived the plane wave angular spectrum of the superimposed ordinary and extraordinary monochromatic plane waves that are present inside a uniaxial crystal. The approach allows us to solve the boundary-value problem, that is, to obtain the field of beam propagating in the uniaxial crystal (z>0) when its distribution is defined on the initial plane (z = 0). The approach is still valid here because the external electric field applied along the optical axis does not change the symmetry of SBN crystal. For convenience to find the solution of Eq. (5), we express r and k by r=r(e^xcosϕ+e^ysinϕ) and k=k(e^xcosθ+e^ysinθ). Following the approach presented in [21,27–29], it is very easy to obtain the solution of Eq. (5)

E(r,ϕ,z)=exp(ik0n01ne2γ13E0z)[Α(0)(r,z)+R(ϕ)Α(2)(r,z)],
with
A(n)(r,z)=π0dk[exp(izk22k0n01ne2γ13E0)+exp(izn0k21ne2γ13E02k0ne2(1ne2γ33E0))]kJn(kr)E˜(k),
and
R(ϕ)=[cos2ϕsin2ϕsin2ϕcos2ϕ]
Here, Jn in Eq. (10) stands for the n-th order Bessel function of the first kind. From Eqs. (9)-(11), we can see that E(r,ϕ,z) is controlled by the external electric field E0 and that A(n) plays the key role here. Unlike the symmetry of the first term A(0) of Eq. (9), the second term R(ϕ)Α(2) is azimuth angle (ϕ)-dependent, which means the distribution of the optical electric field is asymmetric during the propagation.

Suppose that the initial beam is linearly polarized, that is, E(r,0)=E(r,0)t^(α) with t^(α)=cosαe^x+sinαe^y, which is a unit vector that forms an angle α with the x-axis. Inserting E(r,0) into Eq. (6) and the corresponding spectrum E˜(k)=E˜(k)t^(α)into Eq. (10) and Eq. (9), we can get the two Cartesian components of the beam in the polar coordinate system [29],

Ex(r,ϕ,z)=exp(ik0n01ne2γ13E0z)[A(0)(r,z)cosα+A(2)(r,z)cos(2ϕα)],Ey(r,ϕ,z)=exp(ik0n01ne2γ13E0z)[A(0)(r,z)sinα+A(2)(r,z)sin(2ϕα)].
From Eq. (12), we can see that Ex and Ey are coupled through A(0) and A(2), and both of them can be controlled by the external electric field. Additionally, Ex and Ey are asymmetric due to their dependences of the azimuth angle ϕ. Interestingly, Eq. (12) reveals that the beam can be regarded as the superposition of two beams with different polarizations: one is linear polarization while the other is vortex polarization with a topological charge of 2 [34].

To obtain a more suitable solution representation for a circularly polarized incident field, we introduce two complex unit vectors, e^+=2/2(e^x+ie^y), and e^=2/2(e^xie^y), corresponding to left-handed and right-handed circularly polarized light waves, respectively. Without loss of generality, we assume that the incident field is a left-handed circularly polarized beam without a vortex, namely, E(r,0)=E(r,0)e^+, so the solution of Eq. (5) can be expressed as [21,28]:

E+(r,ϕ,z)=exp(ik0n01ne2γ13E0z)A(0)(r,z),E(r,ϕ,z)=exp[i(k0n01ne2γ13E0z+2φ)]A(2)(r,z).
Here, A(0) (r, z) and A(2)(r, z) are the same as those in Eq. (12). From Eq. (13), one can see that the intensity distributions of both the left-handed and the right-handed component are symmetric, meanwhile, the right-handed component is a vortex beam of topological charge 2 [28]. As we know, a left- and right-handed circularly polarized photons possess their SAMs of ±, respectively. The OAM depends on the helical phase structure exp(ilϕ) of light, which carries an eigenvalue of l per photon [36]. From Eq. (13), we can see that the right-handed component is with an OAM of 2 per photon. That is to say, the exchange of the two angular momentums happens during propagation, and its efficiency can be controlled by electro-optic effect. To investigate the efficiency of the angular momentum exchange, we evaluate the total power carried by the two fields [28]
W±(z)=12W+(0)±4π0dkkcos(zΔ2k0n01ne2γ13E0k2)|E˜+(k)|2,
where Δ=no2(1ne2γ13E0)/ne2(1ne2γ33E0)1. We can see that Δ plays the key role in the power evolution of the two field, however, it can be adjusted by LEO effect. Especially, we have W+(z)=W+(0) and W(z)=0 when Δ=0, which means that the power exchange between the two fields is forbidden. Correspondingly, the fluxes of SAM Φs and OAM Φo carried by the light field turn out to be [28]
Φs(z)=noε0c2ω[W+(z)W(z)],Φo(z)=2noε0c2ωW(z).
Unfortunately, it is almost impossible to present an analytical description of the power dynamics for a CAB beam as that for a Gaussian beam because the FT of CAB is much complicated than that of Gaussian beams.

3. Numerical Calculations

Taking advantage of Eqs. (12) and (13), the propagation characteristics of the beam can be analysed. However, the most important factor in Eqs. (12) and (13) A(n), presented in Eq. (10), does not allow an analytic treatment. We use an accurate numerical method to perform the integral related to oscillating function E˜(k) and obtain the numerical solution. Similarly, the numerical integrals are also needed for calculating SAM Φs and OAM Φo using Eqs. (14) and (15). Here, we adopt adaptive Simpson quadrature to compute the integrals and use MATLAB software to perform the simulations.

Here, we investigate numerically the EOC of a circularly and linearly polarized CAB in a SBN crystal. The beam propagates along the optical axis (z-axis) of the crystal. The beam parameters are selected as follows: r0 = 0.2 mm, w = 8 µm, a = 0.05, wavelengthλ = 632.8 nm. We assume that Im is the maximum intensity of the incident beam at the initial plane. To see the change of the intensity during the propagation, all intensities are normalized by Im in our calculations. First, we focus on the situation where the input CAB is left-handed circularly polarized. Since the intensity distributions of E+ and E- are cylindrically symmetric, the intensity distributions can easily be shown by displaying the dependence of the intensities of E+ and E- on the radial distance r. Figure 1 shows the intensity distributions of E+ and E- at the propagation distance z = 10mm with different external electric field. From Fig. 1, we can see that E- is much weaker than E+ with absence of an external electric field. When E0 = 5 kV/mm, the maximum intensity of the E-’ main lobe is above 10 times larger than that without external electric field, and become comparable with that of the E+’ main lobe. If E0 increases further to10 kV/mm, the first lobe of E+ weakens, while the second one becomes the main lobe, which is stronger than the main lobe of E-. For E0 = 15 kV/mm, the second lobe of E- becomes dominant, whereas all lobes of E+ turns to diminish. That is to say, the exchange between the two components and the intensity pattern can be controlled by the external electric field.

 figure: Fig. 1

Fig. 1 Intensity distributions (I/Im) of E+ (dark solid line) and E- (red dash line) with different external electric fields: (a) E0 = 0 kV/mm; (b) E0 = 5 kV/mm; (c) E0 = 10 kV/mm; (d) E0 = 15 kV/mm.

Download Full Size | PDF

To observe the influence of EOC on the beam’ propagation dynamics, Figs. 2 and 3 provide the detailed plots of the evolutions of E+ and E- with 0 kV/mm and 5kV/mm external electric field, respectively. Figures 2 and 3 obviously shows that the radii of both E+ and E- become increasingly smaller with the propagation distance, whereas the intensities rapidly increases at the focal point, which means the abruptly autofocusing happens in both E+ and E-. From Fig. 2(a), we can see that the width of beam E+ becomes very small from the focal point (about z = 15mm) while the depth of focal spot lasts about 10 mm long along the propagation axis. Since the beam E+ is solid near the focal point, it becomes a “needle-like” beam [8, 10, 11, 15, 37]. Similarly, we can see from Fig. 2(b) that the beam E- becomes narrow but still hollow near the focal point, and forms a “cage” or “bottle” beam [11, 38]. This can be interpreted that the vortex phase of E- does not allow the beam to focus into a spot. Moreover, the intensities of optical “needle” and “cage” can be controlled electrically, which is useful in micromanipulation. In the absence of external electric field, one can see from Figs. 2(a) and 2(b) that abruptly autofocusing is strong for E+, while E- is weak due to the weak medium anisotropy [28]. When the external electric field is applied, we can see that from Fig. 3(a) and 3(b) that the abruptly autofocusing of E+ is weakened while that of E- is enhanced because E+ couple its power to E- through EOC during propagation.

 figure: Fig. 2

Fig. 2 Propagation dynamics of E+ (a) and E- (b) with 0 kV/mm external electric field.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Propagation dynamics of E+ (a) and E- (b) with 5 kV/mm external electric field..

Download Full Size | PDF

Figure 4 shows the intensity patterns of the beam at the propagation distance z = 10 mm when the external electric field is E0 = 5 kV/mm. As comparison, we plot the distribution of the beam at the initial plane (z = 0) as shown in Fig. 4(a). Due to the autofocusing, the radii of E+ and E- at z = 10 mm are smaller than those at initial plane. The maximum intensity of E- is stronger than that of E+, which agrees with the result in Fig. 1(b). To visualize the topological charge carried by the left-handed component E-, Fig. 4(b) presents the interference pattern between E- (E(r,φ,z)/max(|E(r,φ,z)|)) and the plane wave (exp[i(k0n01ne2γ13E0z)]). One can see that the left-handed component carries a vortex phase with a topological charge of 2.

 figure: Fig. 4

Fig. 4 Intensity patterns (I/Im) of the beam when the external electric field E0 = 5kV/mm at z = 10 mm: (a) initial distribution of the beam; (b) interference pattern of |E(r,φ,z)/max(|E(r,φ,z)|)+exp[i(k0n01ne2γ13E0z)]|2; (c) intensity distributions of E+; (d) intensity distributions of E-.

Download Full Size | PDF

Further, we obtain the evolutions of SAM and OAM using Eqs. (14) and (15) as shown in Fig. 5, where Φs gradually decreases while Φo symmetrically increases. This can be interpreted that the left-handed circularly polarized component with initial zero OAM and SAM of + per photon is gradually converted into the right-handed component with OAM of 2 and SAM of - per photon. However, the total angular momentum Φs + Φo keeps constant because the electro-optic effect here does not break the rotational invariance around the optical axis [21]. Interestingly, Fig. 5 also indicates that the efficiency of angular momentum exchange can be controlled by external electric field. Especially, when E0 = −1.687kV/mm, EOC forbids the exchange between SAM and OAM, which is similar to the case of Gaussian beam [21].

 figure: Fig. 5

Fig. 5 Plots of Φs(z)/Φs(0) (solid line) and Φo(z)/Φs(0) (dash line) of a CAB vs propagation distance with different external electric field.

Download Full Size | PDF

In the following, we study the other situation where the input CAB is linearly polarized along the x-axis of the crystal. Figures 6 and 7 respectively show the intensity distributions of Ex and Ey with different external electric fields at the propagation distance z = 10 mm. Similar to the circularly polarized CAB, the autofocusing also occurs and make the radius of the beam smaller than that at initial plane. From Fig. 6, we can see that the intensity pattern of Ex is no longer concentric rings when the external electric field is applied; instead, it varies with external electric field. From Fig. 7, we can see that the intensity of Ey is low without an electric field but becomes higher with the external electric field because of the EOC. In addition, we can see that the intensity patterns of Ex and Ey change with the external electric field.

 figure: Fig. 6

Fig. 6 Intensity patterns of Ix/Im with different external electric fields at z = 10 mm: (a) E0 = 0 kV/mm; (b) E0 = 5 kV/mm; (c) E0 = 10 kV/mm; (d) E0 = 15 kV/mm.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Intensity patterns of Iy/Im with different external electric fields at z = 10 mm: (a) E0 = 0 kV/mm; (b) E0 = 5 kV/mm; (c) E0 = 10 kV/mm; (d) E0 = 15 kV/mm.

Download Full Size | PDF

Figure 8 shows the polarization distribution of beam with different external electric field at the propagation distance z = 10mm. As expected, the polarization state becomes spatially inhomogeneous, and possesses vortex with a topological charge of 2. This result agrees with Eq. (12), which shows obviously that the beam contains one component of vortex polarization [34]. More importantly, the polarization distribution of beam can be controlled by external electric field.

 figure: Fig. 8

Fig. 8 Polarization distribution with different external electric fields at z = 10 mm: (a) E0 = 0 kV/mm; (b) E0 = 5 kV/mm. The size of the ellipse denotes the beam intensity.

Download Full Size | PDF

4. Conclusion

In summary, as far as we know, this paper reports firstly the theoretical investigation on the EOC of a circular Airy beam in a uniaxial crystal. Basing on the deduction of the wave coupling equations considering EOC, we simulate numerically the propagations of CABs for a left-handed circularly and a linearly polarized incident CABs in a SBN crystal. Our results show EOC causes a left-handed circularly polarized beam to couple toward a right-handed circularly polarized beam with a vortex topological charge of 2, which opens a novel way to generate vortex light beam. It is worth noting that the EOC can be used to enhance or weaken, thereby manipulate the abrupt autofocusing, the most important property for CAB during their propagation. Interestingly, EOC also plays a role in forming and controlling an optical “needle” and “cage” for both left-handed and right-handed components near the focal point, which can find its application in micromanipulation. In addition, EOC can influence or even forbid the exchange between SAM and OAM. For the case of linearly polarized CAB, EOC can be applied to control the powers exchange of the two Cartesian field components, as well as the intensity pattern and polarization distribution of the CAB. Excitingly, the polarization state can become spatially inhomogeneous, and possess vortex with a topological charge of 2 during propagation. Anyway, our work finds a very useful method to control an Airy beam.

Funding

National Natural Science Foundation of China (NSFC) (11404220, 61275101, 61605128); Shenzhen Science and Technology Development Fund (JCYJ20160308093947132).

References and links

1. N. K. Efremidis and D. N. Christodoulides, “Abruptly autofocusing waves,” Opt. Lett. 35(23), 4045–4047 (2010). [CrossRef]   [PubMed]  

2. D. G. Papazoglou, N. K. Efremidis, D. N. Christodoulides, and S. Tzortzakis, “Observation of abruptly autofocusing waves,” Opt. Lett. 36(10), 1842–1844 (2011). [CrossRef]   [PubMed]  

3. S. Liu, M. Wang, P. Li, P. Zhang, and J. Zhao, “Abrupt polarization transition of vector autofocusing Airy beams,” Opt. Lett. 38(14), 2416–2418 (2013). [CrossRef]   [PubMed]  

4. Y. Jiang, K. Huang, and X. Lu, “Propagation dynamics of abruptly autofocusing Airy beams with optical vortices,” Opt. Express 20(17), 18579–18584 (2012). [CrossRef]   [PubMed]  

5. I. Chremmos, P. Zhang, J. Prakash, N. K. Efremidis, D. N. Christodoulides, and Z. Chen, “Fourier-space generation of abruptly autofocusing beams and optical bottle beams,” Opt. Lett. 36(18), 3675–3677 (2011). [CrossRef]   [PubMed]  

6. R. Dasgupta, S. Ahlawat, R. S. Verma, and P. K. Gupta, “Optical orientation and rotation of trapped red blood cells with Laguerre-Gaussian mode,” Opt. Express 19(8), 7680–7688 (2011). [CrossRef]   [PubMed]  

7. P. Panagiotopoulos, D. G. Papazoglou, A. Couairon, and S. Tzortzakis, “Sharply autofocused ring-Airy beams transforming into non-linear intense light bullets,” Nat. Commun. 4, 2622 (2013). [CrossRef]   [PubMed]  

8. P. Zhang, J. Prakash, Z. Zhang, M. S. Mills, N. K. Efremidis, D. N. Christodoulides, and Z. Chen, “Trapping and guiding microparticles with morphing autofocusing Airy beams,” Opt. Lett. 36(15), 2883–2885 (2011). [CrossRef]   [PubMed]  

9. P. Li, S. Liu, T. Peng, G. Xie, X. Gan, and J. Zhao, “Spiral autofocusing Airy beams carrying power-exponent-phase vortices,” Opt. Express 22(7), 7598–7606 (2014). [CrossRef]   [PubMed]  

10. I. Chremmos, P. Zhang, J. Prakash, N. K. Efremidis, D. N. Christodoulides, and Z. Chen, “Fourier-space generation of abruptly autofocusing beams and optical bottle beams,” Opt. Lett. 36(18), 3675–3677 (2011). [CrossRef]   [PubMed]  

11. W. G. Zhu and W. L. She, “Tightly focusing vector circular airy beam through a hard aperture,” Opt. Commun. 334, 303–307 (2015). [CrossRef]  

12. G. A. Siviloglou, J. Broky, A. Dogariu, and D. N. Christodoulides, “Observation of Accelerating Airy Aeams,” Phys. Rev. Lett. 99(21), 213901 (2007). [CrossRef]   [PubMed]  

13. Y. Hu, G. A. Siviloglou, P. Zhang, N. K. Efremidis, D. N. Christodoulides, and Z. Chen, “Self-accelerating Airy Beams: Generation, Control, and Applications,” in Nonlinear Photonics and Novel Optical Phenomena, Z. Chen, and R. Morandotti, eds. (Springer, 2012), pp. 1–46.

14. Y. Jiang, X. Zhu, W. Yu, H. Shao, W. Zheng, and X. Lu, “Propagation characteristics of the modified circular Airy beam,” Opt. Express 23(23), 29834–29841 (2015). [CrossRef]   [PubMed]  

15. G. L. Zheng, X. Q. Deng, S. X. Xu, and Q. Y. Wu, “The propagation characteristics of circular Airy beam with low-pass filtering modification,” Proc. SPIE 10256, 102561L (2017). [CrossRef]  

16. C. Y. Hwang, K. Y. Kim, and B. Lee, “Dynamic control of circular Airy beams with linear optical potentials,” IEEE Photonics J. 4(1), 174–180 (2012). [CrossRef]  

17. I. D. Chremmos, Z. Chen, D. N. Christodoulides, and N. K. Efremidis, “Abruptly autofocusing and autodefocusing optical beams with arbitrary caustics,” Phys. Rev. A 85(2), 023828 (2012). [CrossRef]  

18. H. Zhong, Y. Zhang, M. R. Belić, C. Li, F. Wen, Z. Zhang, and Y. Zhang, “Controllable circular Airy beams via dynamic linear potential,” Opt. Express 24(7), 7495–7506 (2016). [CrossRef]   [PubMed]  

19. W. L. She and W. K. Lee, “Wave coupling theory of linear electrooptic effect,” Opt. Commun. 195(1-4), 303–311 (2001). [CrossRef]  

20. G. Zheng, H. Wang, and W. She, “Wave coupling theory of Quasi-Phase-Matched linear electro-optic effect,” Opt. Express 14(12), 5535–5540 (2006). [CrossRef]   [PubMed]  

21. L. Chen and W. She, “Electro-optically forbidden or enhanced spin-to-orbital angular momentum conversion in a focused light beam,” Opt. Lett. 33(7), 696–698 (2008). [CrossRef]   [PubMed]  

22. W. Zhu and W. She, “Electro-optically generating and controlling right- and left-handed circularly polarized multiring modes of light beams,” Opt. Lett. 37(14), 2823–2825 (2012). [CrossRef]   [PubMed]  

23. W. Zhu and W. She, “Electrically controlling spin and orbital angular momentum of a focused light beam in a uniaxial crystal,” Opt. Express 20(23), 25876–25883 (2012). [CrossRef]   [PubMed]  

24. X. Lu, Z. Wu, W. Zhang, and L. Chen, “Polarization singularities and orbital angular momentum sidebands from rotational symmetry broken by the Pockels effect,” Sci. Rep. 4(1), 4865 (2014). [CrossRef]   [PubMed]  

25. I. Skab, Y. Vasylkiv, I. Smaga, and R. Vlokh, “Spin-to-orbital momentum conversion via electro-optic Pockels effect in crystals,” Phys. Rev. A 84(4), 043815 (2011). [CrossRef]  

26. H. Tang, L. Chen, and W. She, “The spatially varying polarization of a focused Gaussian beam in quasi-phase-matched superlattice under electro-optic effect,” Opt. Express 18(24), 25000–25007 (2010). [CrossRef]   [PubMed]  

27. A. Ciattoni, B. Crosignani, and P. Di Porto, “Vectorial theory of propagation in uniaxially anisotropic media,” J. Opt. Soc. Am. A 18(7), 1656–1661 (2001). [CrossRef]   [PubMed]  

28. A. Ciattoni, G. Cincotti, and C. Palma, “Circularly polarized beams and vortex generation in uniaxial media,” J. Opt. Soc. Am. A 20(1), 163–171 (2003). [CrossRef]   [PubMed]  

29. A. Ciattoni, G. Cincotti, and C. Palma, “Propagation of cylindrically symmetric fields in uniaxial crystals,” J. Opt. Soc. Am. A 19(4), 792–796 (2002). [CrossRef]   [PubMed]  

30. G. Zhou, R. Chen, and X. Chu, “Propagation of Airy beams in uniaxial crystals orthogonal to the optical axis,” Opt. Express 20(3), 2196–2205 (2012). [CrossRef]   [PubMed]  

31. D. M. Deng, C. D. Chen, X. Zhao, and H. G. Li, “Propagation of an Airy vortex beam in uniaxial crystals,” Appl. Phys. B 110(3), 433–436 (2013). [CrossRef]  

32. M.-L. Zhou, C.-D. Chen, B. Chen, X. Peng, Y.-L. Peng, and D.-M. Deng, “Propagation of an Airy-Gaussian beam in uniaxial crystals,” Chin. Phys. B 24(12), 124102 (2014). [CrossRef]  

33. W. Yu, R. Zhao, F. Deng, J. Huang, C. Chen, X. Yang, Y. Zhao, and D. Deng, “Propagation of Airy Gaussian vortex beams in uniaxial crystals,” Chin. Phys. B 25(4), 044201 (2016). [CrossRef]  

34. G. Zheng, X. Deng, S. Xu, and Q. Wu, “Propagation dynamics of a circular Airy beam in a uniaxial crystal,” Appl. Opt. 56(9), 2444–2448 (2017). [CrossRef]   [PubMed]  

35. A. Yariv, Optical Waves in Crystals (Wiley, 1983).

36. 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(11), 8185–8189 (1992). [CrossRef]   [PubMed]  

37. H. F. Wang, L. P. 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(8), 501–505 (2008). [CrossRef]  

38. P. Zhang, Z. Zhang, J. Prakash, S. Huang, D. Hernandez, M. Salazar, D. N. Christodoulides, and Z. Chen, “Trapping and transporting aerosols with a single optical bottle beam generated by moiré techniques,” Opt. Lett. 36(8), 1491–1493 (2011). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (8)

Fig. 1
Fig. 1 Intensity distributions (I/Im) of E+ (dark solid line) and E- (red dash line) with different external electric fields: (a) E0 = 0 kV/mm; (b) E0 = 5 kV/mm; (c) E0 = 10 kV/mm; (d) E0 = 15 kV/mm.
Fig. 2
Fig. 2 Propagation dynamics of E+ (a) and E- (b) with 0 kV/mm external electric field.
Fig. 3
Fig. 3 Propagation dynamics of E+ (a) and E- (b) with 5 kV/mm external electric field..
Fig. 4
Fig. 4 Intensity patterns (I/Im) of the beam when the external electric field E0 = 5kV/mm at z = 10 mm: (a) initial distribution of the beam; (b) interference pattern of | E (r,φ,z) / max(| E (r,φ,z) | )+exp[i( k 0 n 0 1 n e 2 γ 13 E 0 z)] | 2 ; (c) intensity distributions of E+; (d) intensity distributions of E-.
Fig. 5
Fig. 5 Plots of Φs(z)/Φs(0) (solid line) and Φo(z)/Φs(0) (dash line) of a CAB vs propagation distance with different external electric field.
Fig. 6
Fig. 6 Intensity patterns of Ix/Im with different external electric fields at z = 10 mm: (a) E0 = 0 kV/mm; (b) E0 = 5 kV/mm; (c) E0 = 10 kV/mm; (d) E0 = 15 kV/mm.
Fig. 7
Fig. 7 Intensity patterns of Iy/Im with different external electric fields at z = 10 mm: (a) E0 = 0 kV/mm; (b) E0 = 5 kV/mm; (c) E0 = 10 kV/mm; (d) E0 = 15 kV/mm.
Fig. 8
Fig. 8 Polarization distribution with different external electric fields at z = 10 mm: (a) E0 = 0 kV/mm; (b) E0 = 5 kV/mm. The size of the ellipse denotes the beam intensity.

Equations (19)

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

2 E(r)[E(r)]+ k 0 2 ε E(r)+ μ 0 ω 2 P EO (r)=0,
E(r)=E( r ,z)= d 2 k 2 exp(i k r ) E ˜ ( k ,z),
P(r)=P( r ,z)= d 2 k 2 exp(i k r ) P ˜ ( k ,z).
( 2 z 2 + k 0 2 n o 2 k y 2 ) E ˜ x + k x k y E ˜ y i k x z E ˜ z + μ 0 ω 2 P ˜ x EO =0,
( 2 z 2 + k 0 2 n o 2 k x 2 ) E ˜ y + k x k y E ˜ x i k y z E ˜ z + μ 0 ω 2 P ˜ y EO =0,
( k 0 2 n e 2 k 2 ) E ˜ z i k x z E ˜ x i k y z E ˜ y + μ 0 ω 2 P ˜ z EO =0.
2 z 2 E ˜ x i k x z E ˜ z + k x k y E ˜ y +[ k 0 2 n 0 2 (1 n e 2 γ 13 E 0 ) k y 2 ] E ˜ x =0,
2 z 2 E ˜ y i k y z E ˜ z + k x k y E ˜ x +[ k 0 2 n 0 2 (1 n e 2 γ 13 E 0 ) k x 2 ] E ˜ y =0,
[ k 0 2 n e 2 (1 n e 2 γ 33 E 0 ) k 2 ] E ˜ z i k x z E ˜ x i k y z E ˜ y =0.
E ˜ (k)= 1 2π 0 dr r J 0 (kr)E(r,0),
E( r,0 )=CAi( r 0 r w )exp( a r 0 r w ),
E ˜ ( k )=C w 2 ( r 0 w + k 2 w 2 )exp( a k 2 w 2 ) 3k r 0 + k 3 w 3 3k r 0 +3 k 3 w 3 J 0 ( k r 0 + k 3 w 3 3 )
E(r,ϕ,z)=exp(i k 0 n 0 1 n e 2 γ 13 E 0 z)[ Α (0) (r,z)+R(ϕ) Α (2) (r,z)],
A (n) (r,z)=π 0 dk[ exp( iz k 2 2 k 0 n 0 1 n e 2 γ 13 E 0 )+exp( iz n 0 k 2 1 n e 2 γ 13 E 0 2 k 0 n e 2 (1 n e 2 γ 33 E 0 ) ) ]k J n (kr) E ˜ (k) ,
R(ϕ)=[ cos2ϕ sin2ϕ sin2ϕ cos2ϕ ]
E x (r,ϕ,z)=exp(i k 0 n 0 1 n e 2 γ 13 E 0 z)[ A (0) (r,z)cosα+ A (2) (r,z)cos(2ϕα)], E y (r,ϕ,z)=exp(i k 0 n 0 1 n e 2 γ 13 E 0 z)[ A (0) (r,z)sinα+ A (2) (r,z)sin(2ϕα)].
E + (r,ϕ,z)=exp(i k 0 n 0 1 n e 2 γ 13 E 0 z) A (0) (r,z), E (r,ϕ,z)=exp[i( k 0 n 0 1 n e 2 γ 13 E 0 z+2φ)] A (2) (r,z).
W ± (z)= 1 2 W + (0)±4π 0 dkkcos( zΔ 2 k 0 n 0 1 n e 2 γ 13 E 0 k 2 ) | E ˜ + (k) | 2 ,
Φ s (z)= n o ε 0 c 2ω [ W + (z) W (z)], Φ o (z)=2 n o ε 0 c 2ω W (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.