• Photonics Research
  • Vol. 9, Issue 4, 439 (2021)
Xiao-Bo Hu1, Benjamin Perez-Garcia2、5, Valeria Rodríguez-Fajardo3, Raul I. Hernandez-Aranda2, Andrew Forbes3, and Carmelo Rosales-Guzmán1、4、*
Author Affiliations
  • 1Wang Da-Heng Collaborative Innovation Center, Heilongjiang Provincial Key Laboratory of Quantum Manipulation and Control, Harbin University of Science and Technology, Harbin 150080, China
  • 2Photonics and Mathematical Optics Group, Tecnologico de Monterrey, Monterrey 64849, Mexico
  • 3School of Physics, University of the Witwatersrand, Johannesburg 2050, South Africa
  • 4Centro de Investigaciones en Óptica, A.C., Loma del Bosque 115, Colonia Lomas del campestre, 37150 León, Gto., Mexico
  • 5e-mail: b.pegar@tec.mx
  • show less
    DOI: 10.1364/PRJ.416342 Cite this Article Set citation alerts
    Xiao-Bo Hu, Benjamin Perez-Garcia, Valeria Rodríguez-Fajardo, Raul I. Hernandez-Aranda, Andrew Forbes, Carmelo Rosales-Guzmán. Free-space local nonseparability dynamics of vector modes[J]. Photonics Research, 2021, 9(4): 439 Copy Citation Text show less

    Abstract

    One of the most prominent features of quantum entanglement is its invariability under local unitary transformations, which implies that the degree of entanglement or nonseparability remains constant during free-space propagation, true for both quantum and classically entangled modes. Here we demonstrate an exception to this rule using a carefully engineered vectorial light field, and we study its nonseparability dynamics upon free-space propagation. We show that the local nonseparability between the spatial and polarization degrees of freedom dramatically decays to zero while preserving the purity of the state and hence the global nonseparability. We show this by numerical simulations and corroborate it experimentally. Our results evince novel properties of classically entangled modes and point to the need for new measures of nonseparability for such vectorial fields, while paving the way for novel applications for customized structured light.

    1. INTRODUCTION

    It is well known that entanglement is invariant to local unitary transformations. An example of such is the well-known fact that the degree of entanglement remains unmodified upon free-space propagation. Crucially, nonseparability, the fundamental aspect of entanglement, is not exclusive to quantum systems, and this principle applies to both nonlocal and local entanglement at the single- and multiphoton levels. The former is observed between photons that simultaneously exist in physically separated locations, and the latter between the internal degrees of freedom of photons. Classical states of light such as complex vectorial light fields [1] also exhibit local entanglement, attributed to the nonseparability between their spatial and polarization degrees of freedom (DoFs). Here it is worth clarifying what is meant by such a statement, and to this end we follow the terminology of several seminal works [25]. The definition of entanglement is simply the nonseparability of sums of product states that exist in different vector spaces. While there are many options for the DoFs to use with optical fields, e.g., spatial, temporal, frequency, mode index, and even more exotic choices such as path and trajectory [6], here we are interested in the vector spaces made of the polarization and spatial modes, defined by our two DoFs, the former two dimensional and the latter infinite dimensional. The resulting Hilbert space describing the field is the tensor product of the two, i.e., an infinite number of two-dimensional spaces. We will consider the dynamics of one pure state vector in such a Hilbert space. We adopt conventional notation in the literature and call this state a “vector beam” if the polarization is nonhomogeneously distributed in space and a “scalar beam” if homogeneously distributed. Our interest here is in the former: vector beams nonseparable in the polarization and spatial mode DoFs, controversially called “classically entangled” [5,79]. Despite the controversy, it is becoming clear that some quantum systems can be tested and probed with classically entangled light [1012]. For example, classical entanglement has been exploited in quantum error correction [13], quantum state tomography [14], optical communications [1517], and optical metrology [1820]. Additionally, the tightly focusing properties of complex vector modes have been exploited in the field of optical tweezers [2127], micromachining [28], as well as in super-resolution microscopy [2931].

    Recently, there has been an increasing interest in the engineering of vector light beams whose polarization state varies upon free-space propagation. Most of these have focused on the generation of pure vector beams that oscillate from one vector state to another while keeping a constant degree of nonseparability [3237]. A more interesting case, which encloses the previous cases, reported the generation of light beams whose degree of nonseparability oscillates between scalar and vector modes, offering a tool for the on-demand delivery of specific states to desired positions [38]. Such oscillating modes are generated from the superposition of two counterpropagating vector beams, whose implementation can be cumbersome. While these studies have only considered cylindrical vector vortex modes, the use of new symmetries (spatial shapes), such as parabolic or elliptical, could potentially allow us to unveil properties of vector modes that up to now have remained hidden.

    Here we demonstrate a new class of vector beam whose degree of nonseparability features interesting dynamics as it propagates, evolving from a nonhomogeneously polarized vector beam to a quasi-homogeneously polarized beam. Such modes are generated from a nonseparable superposition of orthogonal parabolic beams, which are natural solutions to the Helmholtz equation in parabolic cylindrical coordinates [3944], and orthogonal polarization states. The entanglement dynamics are quantified through a modified measurement of concurrence C, commonly used to measure the degree of nonseparability in vector modes [4549]. It is noteworthy that, while it is tempting to view these beams as separable in their two degrees of freedom, vindicated by the local measure of C, their global concurrence remains unchanged.

    2. THEORETICAL BACKGROUND

    Schematic representation of a classically entangled light mode featuring a separation of both degrees of freedom upon free space propagation. Right and left circular polarizations are represented by orange and green ellipses, while linear polarization is represented by gray lines.

    Figure 1.Schematic representation of a classically entangled light mode featuring a separation of both degrees of freedom upon free space propagation. Right and left circular polarizations are represented by orange and green ellipses, while linear polarization is represented by gray lines.

    The engineered vector beams are constructed from a superposition of traveling parabolic-Gaussian (TPG±) beams. Mathematically, TPG beams are given by superposition of the even and odd parabolic-Gaussian (PG) beams as [39]TPG±(r;a)=PGe(r;a)±iPGo(r;a).

    Here the functions PGe,o(·) are the even and odd PG modes of the parabolic cylindrical coordinates r=(η,ξ,z), given by [44]PGe(r;a)=exp(ikt22kzμ)GB(r)|Γ1|2π2Pe(2ktμξ;a)×Pe(2ktμη;a),PGo(r;a)=exp(ikt22kzμ)GB(r)2|Γ3|2π2Po(2ktμξ;a)×Po(2ktμη;a),where Pe(·) and Po(·) are the even and odd solutions of the parabolic cylindrical differential equation [d2/dx2+(x2/4a)]P(x;a)=0, and a(,) represents the continuous order of the beam. Importantly, the solutions Pe(·) and Po(·) can be written in terms of Kummer confluent hypergeometric functions [50], which are available in many numerical libraries. Furthermore, Γ1=Γ[(1/4)+(1/2)ia] and Γ3=Γ[(3/4)+(1/2)ia], with Γ(·) the gamma function; and kt is the transverse component of the wave vector k, whose magnitude is related to the longitudinal component kz as k2=kt2+kz2. Additionally, GB(r) is the fundamental Gaussian beam given by GB(r)=exp(ikz)μexp(r2μω02),where μ=μ(z)=1+iz/zr with zr=kω02/2 being the usual Rayleigh range of a Gaussian beam. The parabolic coordinates r=(η,ξ,z) are related to the Cartesian coordinates as x=(η2ξ2)/2 and y=ηξ, where η[0,) and ξ(,). For values γ=ktωo1, the PG beam propagates in a nondiffracting way within the range [zmax,zmax], where zmax=ω0k/kt.

    The traveling parabolic-Gaussian vector (TPGV) beams are generated as a nonseparable weighted superposition of the polarization and spatial degrees of freedom. Here the polarization degree of freedom is encoded in the circular polarization basis while the spatial degree of freedom is precisely encoded in the TPG±(·) modes. Mathematically, such superposition can be written as TPGV(r;a)=12[TPG+(r;a)e^R+TPG(r;a)exp(iϕ)e^L],where the unitary vectors e^R and e^L represent the right and left circular states of polarization, respectively. Finally, the term exp(iϕ) (ϕ[π/4,π/4]) is a phase difference between both constituting modes.

    Phase (back panels) and polarization distribution overlapped with the intensity profiles (front panels) of the scalar modes TPG+(r;a)e^R (left) and TPG−(r;a)e^L (middle), which are combined to generate the TPGV mode (right) at the (a) near and (b) far field. In the case of the TPGV mode, we depict the phase of the complex Stokes field S=S1+iS2.

    Figure 2.Phase (back panels) and polarization distribution overlapped with the intensity profiles (front panels) of the scalar modes TPG+(r;a)e^R (left) and TPG(r;a)e^L (middle), which are combined to generate the TPGV mode (right) at the (a) near and (b) far field. In the case of the TPGV mode, we depict the phase of the complex Stokes field S=S1+iS2.

    3. EXPERIMENTAL GENERATION OF TPGV MODES

    (a) To experimentally generate the TPGV mode, we used a novel approach based on a digital micromirror device (DMD). A laser beam expanded and collimated (by lenses L1 and L2) is diagonally polarized with a half-wave plate (HWP1) and split by a polarizing beam splitter (PBS) according to its polarization components. The two beams are redirected to the DMD, impinging at slightly different angles but overlapped at the center of a binary multiplexed hologram where the TPG+ and TPG− scalar modes are encoded, each with a unique linear grating. After the DMD, the first diffraction order of each beam overlaps along a common propagation axis where the TPGV is generated. A spatial filter (SF) placed at the focusing plane of a telescope formed by lenses L3 and L4 removes all higher diffraction orders. A quarter-wave plate (QWP1) transforms the mode from the linear to the circular polarization basis. The nonseparability dynamics are quantified through Stokes polarimetry, for which a set of four intensities are recorded with a charge-coupled device (CCD) camera. (b) Experimental Stokes parameters S1, S2, and S3 of the TPGV mode TPGV(r,3). (c) Intensity profile overlapped with the reconstructed polarization distribution.

    Figure 3.(a) To experimentally generate the TPGV mode, we used a novel approach based on a digital micromirror device (DMD). A laser beam expanded and collimated (by lenses L1 and L2) is diagonally polarized with a half-wave plate (HWP1) and split by a polarizing beam splitter (PBS) according to its polarization components. The two beams are redirected to the DMD, impinging at slightly different angles but overlapped at the center of a binary multiplexed hologram where the TPG+ and TPG scalar modes are encoded, each with a unique linear grating. After the DMD, the first diffraction order of each beam overlaps along a common propagation axis where the TPGV is generated. A spatial filter (SF) placed at the focusing plane of a telescope formed by lenses L3 and L4 removes all higher diffraction orders. A quarter-wave plate (QWP1) transforms the mode from the linear to the circular polarization basis. The nonseparability dynamics are quantified through Stokes polarimetry, for which a set of four intensities are recorded with a charge-coupled device (CCD) camera. (b) Experimental Stokes parameters S1, S2, and S3 of the TPGV mode TPGV(r,3). (c) Intensity profile overlapped with the reconstructed polarization distribution.

    To quantify the degree of nonseparability we relied on a well-known measure from quantum mechanics, the concurrence C, which assigns a value in the range [0,1] to the degree of entanglement [4549]. The concurrence C is measured by integrating the Stokes parameters Si,i=0,  1,2,  3 over the entire transverse plane through the relation [46,47]C=1(S1S0)2(S2S0)2(S3S0)2,where Si=R2SidA. The Stokes parameters Si are computed from a set of four intensity measurements as [54] S0=I0,S1=2IHS0,S2=2IDS0,S3=2IRS0,where I0 is the total intensity of the mode and IH, ID, and IR the intensity of the horizontal, diagonal, and right-handed polarization components, respectively. As illustrated in Fig. 3(a), such intensity measurements were acquired by a CCD camera through the combination of a linear polarizer (P) and a quarter-wave plate (QWP2) [55]. Specifically, the intensities of the horizontal (IH) and diagonal (ID) polarization components were obtained by passing the beam through a linear polarizer at 0° and 45°, respectively, whereas intensity corresponding to the RCP component (IR) was obtained by passing the beam simultaneously through a QWP at 45° and a linear polarizer at 90°. As an example, Fig. 3(b) shows the experimental Stokes parameters S0, S1, S2, and S3 for the specific mode TPGV(r,3) at z=0. Such parameters were used to reconstruct the transverse polarization distribution on a 20×20 grid as shown in Fig. 3(c). Here, for the sake of clarity, we also display the transverse intensity profile. For this specific example, we have S1/S0=0.12, S2/S0=0.09, and S3/S0=0.02, which upon substitution in Eq. (5) yield the value C=0.98, which, as expected, corresponds to a maximally entangled mode.

    4. ANALYSIS OF THE NONSEPARABILITY DYNAMICS

    (a) Experimental and (b) simulated evolution of intensity and polarization distribution of TPGV beams as a function of propagation distance. (c) Experimental and (d) simulated representation of the transverse polarization distribution on a Poincaré sphere, each in correspondence with the planes shown in (a) and (b).

    Figure 4.(a) Experimental and (b) simulated evolution of intensity and polarization distribution of TPGV beams as a function of propagation distance. (c) Experimental and (d) simulated representation of the transverse polarization distribution on a Poincaré sphere, each in correspondence with the planes shown in (a) and (b).

    With the idea of better visualizing the evolution of polarization upon propagation, we mapped the different states of polarization acquired at each plane onto the well-known Poincaré sphere (PS), in which the different polarization states are associated to unique points on the surface of the sphere [54] as shown in Figs. 4(c) and 4(d) for experiment and numerical simulation, respectively. Here the coordinate axes are given in terms of the Stokes parameters S1, S2, and S3. For z=0, all the states of polarization in the transverse plane are linear and therefore mapped to points along the equator. For z>0, such linear polarization states gradually evolve from linear into elliptical and finally to circular. In the PS, this is seen as points along a spiral trajectory connecting the north and south poles, in the intermediate planes, and in the far field as points on the north and south poles. Notice that even though the polarization structure evolves from completely mixed to completely unmixed, the amount of right- and left-elliptically polarized photons remains in equilibrium.

    (a) Concurrence as a function of propagation and transverse coordinates. The solid and dashed lines show two examples of C as a function of z for the specific sections of the beam, shown in (b) and (c), which correspond to z=0 and z=4zmax, respectively. The surface corresponds to numerical simulations and the data points to experimental measurements at the propagation planes z/zmax=0, 0.2, 0.7, 1.3, 2.2, and 3.6. An analogous behavior has been reported in the context of temporal coherence [58], where the generation of beams whose local degree of temporal coherence varies as a function of the time difference is discussed.

    Figure 5.(a) Concurrence as a function of propagation and transverse coordinates. The solid and dashed lines show two examples of C as a function of z for the specific sections of the beam, shown in (b) and (c), which correspond to z=0 and z=4zmax, respectively. The surface corresponds to numerical simulations and the data points to experimental measurements at the propagation planes z/zmax=0,  0.2,  0.7,  1.3,2.2, and 3.6. An analogous behavior has been reported in the context of temporal coherence [58], where the generation of beams whose local degree of temporal coherence varies as a function of the time difference is discussed.

    5. DISCUSSION

    There are several mechanisms by which the concurrence of an initially pure state with maximum nonseparability could decay, but perhaps the most prominent in the context of spatial modes is an evolution toward a mixed state (by interacting with an open system, for example) or by modal scattering and subsequent subspace concatenation (ignoring the full expansion in the vector space and considering only the initial subspace). Neither happens here. Instead, our carefully constructed example highlights the counterintuitive dynamics at play: the local concurrence changes dramatically with propagation, eventually decreasing to zero everywhere; yet the global concurrence remains unchanged with propagation, consistent with the fact that free space is a unitary channel, and thus the state vector remains pure, i.e., there is no evolution to a mixed state and no modal scattering. Colloquially, it appears as if the initial beam is split into two homogeneously polarized parts, and so appears separable, it yet remains fully nonseparable in our Hilbert space, as it remains one coherent mode. Our work suggests the need for new or adapted parameters to quantify the nonhomogeneous distribution of polarization states for vector beams. It is intriguing to wonder if our test case may be viewed as a Young’s experiment in reverse, with our initial beam mimicking interference to the final beam reminiscent of two slits. Further, this interesting behavior may benefit from recent insights into polarization coherence, where an exact equality linking concurrence, visibility, and duality has been obtained [4]. These and other open questions are exciting challenges for follow-up work.

    6. CONCLUSIONS

    In this work we demonstrated a novel kind of complex light field that upon free-space propagation evolves from maximally mixed and locally nonseparable to completely unmixed and locally separable. More precisely, we generated a vector beam with a nonhomogeneous polarization distribution that upon free-space propagation evolves into a homogeneously polarized mode. Such behavior is directly observed at various propagation distances through a reconstruction of the transverse polarization distribution performed via Stokes polarimetry. This is further evinced by mapping the entire polarization distribution at each plane onto the Poincaré sphere, which exhibits an evolution of the state of polarization from the equator to the poles. Such evolution happens along spirals connecting the north and south poles. A quantification of this free-pace propagation dynamics was performed through the concurrence C, which takes the value C=1 for vector states and C=0 for scalar beams. We noted that a measure of the global concurrence yields a constant value, while when measuring this in smaller sections of the beam, what we called the “local concurrence,” it clearly shows a decrease as a function of the propagation distance, reaching the value C=0 in the far field. In other words, by measuring the local concurrence we observe what appears to be a decay in nonseparability, even though the entire state is coherent and nonseparable. This evinces the need for an alternative definition of that takes into account the local variations of the nonseparability, but this lies beyond the scope of this research. Importantly, the nonseparability dynamics reported here cannot be observed with cylindrical vector beams, as they are an intrinsic property of parabolic vector beams. It is also worth mentioning that the beam’s dynamics can be adjusted through the parameter zmax, which depends on ω0 and kt. Finally, these novel states of light offer a new tool for a wide variety of applications in fields such as optical communications, optical metrology, and optical tweezers, to mention a few.

    Acknowledgment

    Acknowledgment. The authors acknowledge fruitful discussions with Prof. Andrea Aiello from the Max Planck Institute.

    References

    [1] C. Rosales-Guzmán, B. Ndagano, A. Forbes. A review of complex vector light fields and their applications. J. Opt., 20, 123001(2018).

    [2] F. Gori, M. Santarsiero, R. Borghi. Vector mode analysis of a young interferometer. Opt. Lett., 31, 858-860(2006).

    [3] J. Eberly. Shimony–wolf states and hidden coherences in classical light. Contemp. Phys., 56, 407-416(2015).

    [4] X.-F. Qian, A. Vamivakas, J. Eberly. Entanglement limits duality and vice versa. Optica, 5, 942-947(2018).

    [5] J. H. Eberly, X.-F. Qian, A. A. Qasimi, H. Ali, M. A. Alonso, R. Gutiérrez-Cuevas, B. J. Little, J. C. Howell, T. Malhotra, A. N. Vamivakas. Quantum and classical optics–emerging links. Phys. Scr., 91, 063003(2016).

    [6] Y. Shen, X. Yang, D. Naidoo, X. Fu, A. Forbes. Structured ray-wave vector vortex beams in multiple degrees of freedom from a laser. Optica, 7, 820-831(2020).

    [7] T. Konrad, A. Forbes. Quantum mechanics and classical light. Contemp. Phys., 60, 1-22(2019).

    [8] A. Forbes, A. Aiello, B. Ndagano. Classically entangled light. Progress in Optics, 99-153(2019).

    [9] R. J. C. Spreeuw. A classical analogy of entanglement. Found. Phys., 28, 361-374(1998).

    [10] X.-F. Qian, J. H. Eberly. Entanglement and classical polarization states. Opt. Lett., 36, 4110-4112(2011).

    [11] X.-F. Qian, A. N. Vamivakas, J. H. Eberly. Emerging connections: classical and quantum optics. Opt. Photon. News, 28, 34-41(2017).

    [12] K. H. Kagalwala, G. Di Giuseppe, A. F. Abouraddy, B. E. Saleh. Bell’s measure in classical optical coherence. Nat. Photonics, 7, 72-78(2013).

    [13] B. Ndagano, B. Perez-Garcia, F. S. Roux, M. McLaren, C. Rosales-Guzmán, Y. Zhang, O. Mouane, R. I. Hernandez-Aranda, T. Konrad, A. Forbes. Characterizing quantum channels with non-separable states of classical light. Nat. Phys., 13, 397-402(2017).

    [14] E. Toninelli, B. Ndagano, A. Vallés, B. Sephton, I. Nape, A. Ambrosio, F. Capasso, M. J. Padgett, A. Forbes. Concepts in quantum state tomography and classical implementation with intense light: a tutorial. Adv. Opt. Photon., 11, 67-134(2019).

    [15] B. Ndagano, I. Nape, M. A. Cox, C. Rosales-Guzmán, A. Forbes. Creation and detection of vector vortex modes for classical and quantum communication. J. Lightwave Technol., 36, 292-301(2018).

    [16] G. Milione, M. P. Lavery, H. Huang, Y. Ren, G. Xie, T. A. Nguyen, E. Karimi, L. Marrucci, D. A. Nolan, R. R. Alfano, A. E. Willner. 4 × 20  Gbit/s mode division multiplexing over free space using vector modes and a q-plate mode (de)multiplexer. Opt. Lett., 40, 1980-1983(2015).

    [17] Y. Zhao, J. Wang. High-base vector beam encoding/decoding for visible-light communications. Opt. Lett., 40, 4843-4846(2015).

    [18] X.-B. Hu, B. Zhao, Z.-H. Zhu, W. Gao, C. Rosales-Guzmán. In situ detection of a cooperative target’s longitudinal and angular speed using structured light. Opt. Lett., 44, 3070-3073(2019).

    [19] F. Töppel, A. Aiello, C. Marquardt, E. Giacobino, G. Leuchs. Classical entanglement in polarization metrology. New J. Phys., 16, 073019(2014).

    [20] S. Berg-Johansen, F. Töppel, B. Stiller, P. Banzer, M. Ornigotti, E. Giacobino, G. Leuchs, A. Aiello, C. Marquardt. Classically entangled optical beams for high-speed kinematic sensing. Optica, 2, 864-868(2015).

    [21] V. Shvedov, A. R. Davoyan, C. Hnatovsky, N. Engheta, W. Krolikowski. A long-range polarization-controlled optical tractor beam. Nat. Photonics, 8, 846-850(2014).

    [22] N. Bhebhe, C. Rosales-Guzmán, A. Forbes. Classical and quantum analysis of propagation invariant vector flat-top beams. Appl. Opt., 57, 5451-5458(2018).

    [23] N. Bhebhe, P. A. C. Williams, C. Rosales-Guzmán, V. Rodriguez-Fajardo, A. Forbes. A vector holographic optical trap. Sci. Rep., 8, 17387(2018).

    [24] Y. Kozawa, S. Sato. Optical trapping of micrometer-sized dielectric particles by cylindrical vector beams. Opt. Express, 18, 10828-10833(2010).

    [25] S. E. Skelton, M. Sergides, R. Saija, M. A. Iatì, O. M. Maragó, P. H. Jones. Trapping volume control in optical tweezers using cylindrical vector beams. Opt. Lett., 38, 28-30(2013).

    [26] M. G. Donato, S. Vasi, R. Sayed, P. H. Jones, F. Bonaccorso, A. C. Ferrari, P. G. Gucciardi, O. M. Maragó. Optical trapping of nanotubes with cylindrical vector beams. Opt. Lett., 37, 3381-3383(2012).

    [27] B. Roxworthy, K. Toussaint. Optical trapping with π-phase cylindrical vector beams. New J. Phys., 12, 073012(2010).

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

    [29] P. Török, P. Munro. The use of Gauss-Laguerre vector beams in STED microscopy. Opt. Express, 12, 3605-3617(2004).

    [30] X. Hao, C. Kuang, T. Wang, X. Liu. Effects of polarization on the de-excitation dark focal spot in STED microscopy. J. Opt., 12, 115707(2010).

    [31] S. Segawa, Y. Kozawa, S. Sato. Resolution enhancement of confocal microscopy by subtraction method with vector beams. Opt. Lett., 39, 3118-3121(2014).

    [32] I. Moreno, J. A. Davis, M. M. Sánchez-López, K. Badham, D. M. Cottrell. Nondiffracting Bessel beams with polarization state that varies with propagation distance. Opt. Lett., 40, 5451-5454(2015).

    [33] S. Fu, S. Zhang, C. Gao. Bessel beams with spatial oscillating polarization. Sci. Rep., 6, 30765(2016).

    [34] J. A. Davis, I. Moreno, K. Badham, M. M. Sánchez-López, D. M. Cottrell. Nondiffracting vector beams where the charge and the polarization state vary with propagation distance. Opt. Lett., 41, 2270-2273(2016).

    [35] P. Li, Y. Zhang, S. Liu, H. Cheng, L. Han, D. Wu, J. Zhao. Generation and self-healing of vector Bessel-Gauss beams with variant state of polarizations upon propagation. Opt. Express, 25, 5821-5831(2017).

    [36] P. Li, D. Wu, Y. Zhang, S. Liu, Y. Li, S. Qi, J. Zhao. Polarization oscillating beams constructed by copropagating optical frozen waves. Photon. Res., 6, 756-761(2018).

    [37] P. Li, Y. Zhang, S. Liu, L. Han, H. Cheng, F. Yu, J. Zhao. Quasi-Bessel beams with longitudinally varying polarization state generated by employing spectrum engineering. Opt. Lett., 41, 4811-4814(2016).

    [38] E. Otte, C. Rosales-Guzmán, B. Ndagano, C. Denz, A. Forbes. Entanglement beating in free space through spin-orbit coupling. Light Sci. Appl., 7, 18009(2018).

    [39] M. A. Bandres, J. C. Gutiérrez-Vega, S. Chávez-Cerda. Parabolic nondiffracting optical wave fields. Opt. Lett., 29, 44-46(2004).

    [40] C. López-Mariscal, M. A. Bandres, J. C. Gutiérrez-Vega, S. Chávez-Cerda. Observation of parabolic nondiffracting optical fields. Opt. Express, 13, 2364-2369(2005).

    [41] M. A. Bandres. Accelerating parabolic beams. Opt. Lett., 33, 1678-1680(2008).

    [42] B. M. Rodrguez-Lara, R. Jáuregui. Dynamical constants of structured photons with parabolic-cylindrical symmetry. Phys. Rev. A, 79, 055806(2009).

    [43] A. Ruelas, S. Lopez-Aguayo, J. C. Gutiérrez-Vega. Engineering parabolic beams with dynamic intensity profiles. J. Opt. Soc. Am. A, 30, 1476-1483(2013).

    [44] J. C. Gutiérrez-Vega, M. A. Bandres. Helmholtz–Gauss waves. J. Opt. Soc. Am. A, 22, 289-298(2005).

    [45] M. McLaren, T. Konrad, A. Forbes. Measuring the nonseparability of vector vortex beams. Phys. Rev. A, 92, 023833(2015).

    [46] A. Selyem, C. Rosales-Guzmán, S. Croke, A. Forbes, S. Franke-Arnold. Basis-independent tomography and nonseparability witnesses of pure complex vectorial light fields by Stokes projections. Phys. Rev. A, 100, 063842(2019).

    [47] A. Manthalkar, I. Nape, N. T. Bordbar, C. Rosales-Guzmán, S. Bhattacharya, A. Forbes, A. Dudley. All-digital stokes polarimetry with a digital micromirror device. Opt. Lett., 45, 2319-2322(2020).

    [48] B. Zhao, X.-B. Hu, V. Rodríguez-Fajardo, A. Forbes, W. Gao, Z.-H. Zhu, C. Rosales-Guzmán. Determining the non-separability of vector modes with digital micromirror devices. Appl. Phys. Lett., 116, 091101(2020).

    [49] B. Ndagano, H. Sroor, M. McLaren, C. Rosales-Guzmán, A. Forbes. Beam quality measure for vector beams. Opt. Lett., 41, 3407-3410(2016).

    [50] F. Olver, National Institute, D. Lozier, R. Boisvert, C. Clark. NIST Handbook of Mathematical Functions Hardback and CD-ROM(2010).

    [51] C. Rosales-Guzmán, N. Bhebhe, A. Forbes. Simultaneous generation of multiple vector beams on a single SLM. Opt. Express, 25, 25697-25706(2017).

    [52] C. Rosales-Guzmán, X.-B. Hu, A. Selyem, P. Moreno-Acosta, S. Franke-Arnold, R. Ramos-Garcia, A. Forbes. Polarisation-insensitive generation of complex vector modes from a digital micromirror device. Sci. Rep., 10, 10434(2020).

    [53] Y. Li, X.-B. Hu, B. Perez-Garcia, B. Zhao, W. Gao, Z.-H. Zhu, C. Rosales-Guzmán. Classically entangled Ince–Gaussian modes. Appl. Phys. Lett., 116, 221105(2020).

    [54] D. H. Goldstein. Polarized Light(2011).

    [55] B. Zhao, X.-B. Hu, V. Rodríguez-Fajardo, Z.-H. Zhu, W. Gao, A. Forbes, C. Rosales-Guzmán. Real-time Stokes polarimetry using a digital micromirror device. Opt. Express, 27, 31087-31093(2019).

    [56] J. W. Goodman. Introduction to Fourier Optics(2017).

    [57] S. Scholes, R. Kara, J. Pinnell, V. Rodríguez-Fajardo, A. Forbes. Structured light with digital micromirror devices: a guide to best practice. Opt. Eng., 59, 041202(2019).

    [58] S. Yang, S. A. Ponomarenko, Z. D. Chen. Coherent pseudo-mode decomposition of a new partially coherent source class. Opt. Lett., 40, 3081-3084(2015).

    [59] P. Réfrégier. Polarization degree of optical waves with non-Gaussian probability density functions: Kullback relative entropy-based approach. Opt. Lett., 30, 1090-1092(2005).

    Xiao-Bo Hu, Benjamin Perez-Garcia, Valeria Rodríguez-Fajardo, Raul I. Hernandez-Aranda, Andrew Forbes, Carmelo Rosales-Guzmán. Free-space local nonseparability dynamics of vector modes[J]. Photonics Research, 2021, 9(4): 439
    Download Citation