• Photonics Research
  • Vol. 13, Issue 5, 1408 (2025)
Fangze Deng1, Ke Ma1, Yumeng Ma1, Xiang Hou1..., Zhihua Han1, Yuchao Li1, Keke Cheng1, Yansheng Shao1, Chenglong Wang1, Meng Liu1, Huiyun Zhang1,2,* and Yuping Zhang1,3,*|Show fewer author(s)
Author Affiliations
  • 1Qingdao Key Laboratory of Terahertz Technology, College of Electronic and Information Engineering, Shandong University of Science and Technology, Qingdao 266590, China
  • 2e-mail: sdust_thz@126.com
  • 3e-mail: sdust_thz@163.com
  • show less
    DOI: 10.1364/PRJ.555214 Cite this Article Set citation alerts
    Fangze Deng, Ke Ma, Yumeng Ma, Xiang Hou, Zhihua Han, Yuchao Li, Keke Cheng, Yansheng Shao, Chenglong Wang, Meng Liu, Huiyun Zhang, Yuping Zhang, "Dual-channel tunable multipolarization adapted terahertz spatiotemporal vortices generating device," Photonics Res. 13, 1408 (2025) Copy Citation Text show less

    Abstract

    Spatiotemporal optical vortices (STOVs) exhibit characteristics of transverse orbital angular momentum (OAM) that is perpendicular to the direction of pulse propagation, indicating significant potential for diverse applications. In this study, we employ vanadium dioxide and photonic crystal plates to design tunable transreflective dual-channel terahertz (THz) spatiotemporal vortex generation devices that possess multipolarization adaptability. In the reflection channel, we achieve active tunability of the topological dark lines by utilizing circularly polarized light, based on the topological dark phenomenon, and observe variations in the number of singularities across the parameter space from different observational perspectives. In the transmission channel, we generate independent vortex singularities using linearly polarized light. This multifunctional terahertz device offers a novel approach for the generation and active tuning of spatiotemporal vortices.

    1. INTRODUCTION

    Vortices are prevalent in nature in various forms, ranging from quantum vortices to atmospheric vortices, and their existence is not constrained by spatial and temporal scales. In optics, metamaterial and vortex beams with longitudinal orbital angular momentum (OAM) aligned with the wave vector have been extensively studied and applied [115]. This optical vortex manifests as a helical phase wavefront featuring a phase singularity at its center and zero intensity. In contrast to longitudinal OAM vortex beams, spatiotemporal optical vortices (STOVs) possess a transverse OAM characteristic that is perpendicular to the wave vector’s direction. This novel class of optical vortex exhibits unique properties across numerous optical phenomena, thereby garnering significant attention [1624]. In recent years, researchers have achieved remarkable advancements in STOV generation through the utilization of spatial light modulators, subwavelength nanogratings, and photonic crystal plates [2533]. These developments offer a variety of promising strategies for generating spatiotemporal vortices; however, existing methods either depend on complex time-frequency pulse-shaping devices or face limitations due to challenges in nanostructure fabrication. Consequently, the generation of STOVs within a compact and easily manufacturable platform presents a significant challenge. Specifically, balancing the controllability of the spatiotemporal vortex generating device with the need to maximize the versatility of STOV generation and active tuning characteristics remains a critical issue.

    In this paper, we present a tunable transreflective dual-channel and multipolarization terahertz spatiotemporal optical vortices (T-STOV) generation scheme. The reflection and transmission channels can be interconverted by adjusting the temperature of vanadium dioxide (VO2). In the reflection channel, we employ a circularly polarized beam incident structure to observe, for the first time, the existence of topological dark lines in parallel tilt. We document the annihilation process of four vortex singularities into two within the wave vector space plane at a single frequency. This observation indicates that our proposed structure can form either a single singularity or two topological singularities with opposite charges in different observation directions, thereby breaking the constraint that topological dark lines are limited to the vertical plane. Consequently, the degree of freedom and flexibility in generating spatiotemporal vortices are significantly enhanced. Moreover, by leveraging the tunable properties of the VO2 phase transition, we successfully observe and plot the trend and direction of topological dark line movement as a function of temperature, further expanding the tunability of our spatiotemporal vortex generation devices. In the transmission channel, we observe independent and topologically robust vortex singularities in the frequency-momentum plane using linearly polarized beams. The phenomenon of singularity migration is analyzed in detail. Through this analysis, we explore a novel direction for achieving tunable frequency and wave vector properties of spatiotemporal vortices. The active tuning scheme for spatiotemporal vortex generation proposed herein greatly enhances the degree of freedom and flexibility required for such generators, broadens the adaptability and application range of the generation device, and provides a novel approach for designing actively regulated spatiotemporal vortex generators.

    2. THEORETICAL ANALYSIS AND STRUCTURAL DESIGN

    The photonic crystal unit structure we designed is depicted in Fig. 1(a). The top layer consists of photonic crystals made of Si3N4, with a thickness denoted as t1, featuring a transverse U-shaped hole at its center. The periodicity of the structure extends in the x and y directions, with a period of a. The middle layer is composed of SiO2, with a thickness of t2, while the bottom layer is made of VO2, as illustrated in Fig. 1(d). The angle of inclination for the hole in the photonic crystal plate is denoted as θ.

    Structure of the spatiotemporal vortex generating device. (a), (b), (c), and (d) Schematic structure of the photonic crystal plate, with the following structural parameters: a=210 μm; b1=130 μm; b2=150 μm; w=60 μm; t1=52.5 μm; and t2=37.5 μm. Additionally, the aperture tilt angle of the photonic crystal plate is denoted as θ.

    Figure 1.Structure of the spatiotemporal vortex generating device. (a), (b), (c), and (d) Schematic structure of the photonic crystal plate, with the following structural parameters: a=210  μm; b1=130  μm; b2=150  μm; w=60  μm; t1=52.5  μm; and t2=37.5  μm. Additionally, the aperture tilt angle of the photonic crystal plate is denoted as θ.

    To fully satisfy our design objectives, we conducted comprehensive screening and optimization of the structural parameters of the metasurface units employed. First, we started with a square structure, and in order to avoid having in-plane C2 symmetry, we created a groove on one side of the square structure. The groove structure has a width of w and a length of b/2. We explored the effects of different values for the groove width and length on the reflection topological dark line and the transmission topological singularity. After investigation, we optimized the parameters to be a=210  μm, b1=130  μm, b2=150  μm, w=60  μm, t1=52.5  μm, and t2=37.5  μm. As for the tilt angle θ, we treated it as a variable and conducted a detailed exploration. We observed its effect on the properties of the topological dark line in the reflection channel and its influence on the frequency and wave vector properties of the topological singularity in the transmission channel.

    VO2 is a unique phase change material; when it transitions from an insulating state to a metallic state, its conductivity increases dramatically from σ=20  S/m to σ=200,000  S/m. This conversion can be achieved by varying the temperature, allowing VO2 to revert from the metallic state back to the insulating state. In the terahertz frequency range, the relative dielectric constant of VO2 can be characterized using the Drude–Lorentz model [34]: εVO2(ω)=εωp2(σVO2)ω2+iγω,where ε=12 is the relative dielectric constant at the limit frequency, γ=5.75×1013  rad/s is the damped frequency, and the plasma frequency can be approximately described by Eq. (2) [35], where σ0=3×105  S/m and ωp(σ0)=1.4×1015  rad/s: ωp2(σVO2)=σσ0ωp2(σ0).

    We employ temporal coupled-mode theory (TCMT) to analyze the T-STOV phenomena of the system in transmission and reflection modes. First, when the temperature of VO2 exceeds 68°C, vanadium dioxide enters a fully metallic state, and the system operates in reflective mode, as illustrated in Fig. 2(a). Our proposed structure exhibits longitudinal mirror symmetry; however, it lacks in-plane C2 rotational symmetry. This characteristic enables the conversion of incident circularly polarized light into its orthogonal counterpart, leading to the formation of topological dark singularities. We have provided relevant explanations in Appendix A.

    Reflection working mode. (a) When the state of VO2 is in a fully metallic state, the system’s reflection channel is activated, allowing the structure to operate in reflection mode. At this stage, the system is responsive to left- and right-circularly polarized light, enabling the generation of spatiotemporal vortices with distinct orbital angular momentum (OAM) directions depending on the observation angle. (b) The diagram includes four ports labeled 1, 2, 3, and 4 for the photonic crystal plate structure. (c) Complete polarization conversion at the resonant frequency follows a trajectory from one pole to the other on the Poincaré sphere, as indicated by the black arrow in the figure. (d) A 2D resonance band diagram of the photonic crystal plate structure is presented, showcasing vanadium dioxide in various states.

    Figure 2.Reflection working mode. (a) When the state of VO2 is in a fully metallic state, the system’s reflection channel is activated, allowing the structure to operate in reflection mode. At this stage, the system is responsive to left- and right-circularly polarized light, enabling the generation of spatiotemporal vortices with distinct orbital angular momentum (OAM) directions depending on the observation angle. (b) The diagram includes four ports labeled 1, 2, 3, and 4 for the photonic crystal plate structure. (c) Complete polarization conversion at the resonant frequency follows a trajectory from one pole to the other on the Poincaré sphere, as indicated by the black arrow in the figure. (d) A 2D resonance band diagram of the photonic crystal plate structure is presented, showcasing vanadium dioxide in various states.

    At the resonant frequency, the additional phase shift of π in the resonant reflection within the resonant scattering channel causes the photonic crystal plate to function as a half-wave plate. This configuration converts incident left circularly polarized light into right circularly polarized light while simultaneously converting incident right circularly polarized light into left circularly polarized light. This process corresponds to the transition of one pole to the other on the Poincaré sphere, as illustrated in Fig. 2(c).

    As the temperature increases, VO2 undergoes a phase transition from an insulator to a metal, resulting in a significant increase in its conductivity. This phase transition does not occur abruptly at a single temperature point; rather, it progresses gradually within a specific temperature range. The characteristics of this diffuse phase transition enable VO2 and its applications to exhibit smoother temperature control and adjustment capabilities in practical use [36]. This phenomenon provides substantial support for the active tuning of the properties of space-time vortices. The relationship between temperature and conductivity is detailed in Appendix B. Since the conductivity, dielectric constant, and loss of VO2 vary with temperature, the time-reversal symmetry condition and the resonant frequency of the system are altered as the state of VO2 changes.

    We selected states of VO2 corresponding to conductivities of 20,000, 30,000, 50,000, 70,000, 100,000, and 200,000 S/m and plotted the 2D energy band distributions of the photonic crystal system for these different states, as illustrated in Fig. 2(d). The figure demonstrates that the conductivity of VO2 increases with rising temperature, leading to an observed increase in the resonance frequency. Notably, when the state of VO2 approaches a complete transition to the metallic state, the 2D energy bands nearly coincide.

    When VO2 is at room temperature, the system operates in transmission mode, allowing it to adapt to the incidence of a linearly polarized beam and generate isolated singularities in the frequency-momentum plane. This process results in the formation of spatiotemporal vortices, as illustrated in Fig. 3(a). We also employed the double-resonance four-port time-coupled mode theory (TCMT) model, as illustrated in Fig. 3(b). The relevant aspects of the time-coupled mode theory concerning the system’s transmission modes and the conditions for singularity generation are further elaborated in Appendix A.

    Transmission working mode. (a) When the temperature of VO2 is below 68°C, the system operates in transmission mode, with the transmission channel open. This configuration adapts to the generation of spatiotemporal vortices caused by the incidence of linearly polarized light along the y direction. (b) The four ports labeled 1, 2, 3, and 4 are designated for the photonic crystal plate structure.

    Figure 3.Transmission working mode. (a) When the temperature of VO2 is below 68°C, the system operates in transmission mode, with the transmission channel open. This configuration adapts to the generation of spatiotemporal vortices caused by the incidence of linearly polarized light along the y direction. (b) The four ports labeled 1, 2, 3, and 4 are designated for the photonic crystal plate structure.

    3. RESULTS AND DISCUSSION

    In the fully metallic state of VO2, characterized by a conductivity of 200,000 S/m, a left circularly polarized light is incident on a photonic crystal structure with a frequency of 1.108 THz. The copolarized reflectivity and copolarized reflection phase are plotted in the wave vector plane, as illustrated in Fig. 4. We observe the presence of four singularities in the wave vector plane, symmetric about the kx=0 plane, indicated as red labeled dots in the figure. The existence of these singularities markedly contrasts with the conventional single nodal and topological dark lines.

    Plots of the copolarized reflection phase and copolarized reflectivity in the plane of the wave vector at different θ. (a) and (c) Copolarized reflection phase diagram and copolarized reflectivity diagram, respectively, for θ=8° and an incident frequency of 1.108 THz. In these plots, the red points denote the intersections of the topologically dark lines within the wave vector plane in the 3D frequency-momentum space. (b) and (d) Copolarized reflection phase and copolarized reflectivity plots at the same frequency, with the angle θ increased to 15°. As θ increases, the red singularity shifts outward; however, the vortex phase pattern remains unchanged.

    Figure 4.Plots of the copolarized reflection phase and copolarized reflectivity in the plane of the wave vector at different θ. (a) and (c) Copolarized reflection phase diagram and copolarized reflectivity diagram, respectively, for θ=8° and an incident frequency of 1.108 THz. In these plots, the red points denote the intersections of the topologically dark lines within the wave vector plane in the 3D frequency-momentum space. (b) and (d) Copolarized reflection phase and copolarized reflectivity plots at the same frequency, with the angle θ increased to 15°. As θ increases, the red singularity shifts outward; however, the vortex phase pattern remains unchanged.

    Figures 4(a) and 4(c) present the copolarized reflection phase and copolarized reflectivity maps in the vector space at an inclination angle of θ=8° for the photonic crystal dug hole, where four vortex singularities with two opposite topological charges are observable. These vortex singularities represent the intersections of two topological dark lines with the wave vector plane in 3D frequency-momentum space. Figures 4(b) and 4(d) illustrate the copolarized reflection phase and reflectivity maps in the vector space for θ=15°. It is evident that variations in the inclination angle θ result in differing positions of the four singularities within the wave vector plane. This phenomenon corresponds to the increased curvature of the dark line with the inclination angle, which arises from the C2 symmetry of the structure and the degree of broken z-mirror symmetry that intensifies with an increase in θ.

    The inclination angle θ significantly influences the copolarized reflectivity and conversion efficiency of the structure; thus, we selected a moderate angle of θ=15°. The bottom-left inset of Fig. 5 illustrates the topological dark lines present in the 3D frequency-momentum space of the system at 1.1045 to 1.108 THz. We generated copolarized reflectivity and reflection phase maps in the wave vector plane at various frequencies, connecting the vortex singularities in the wave vector plane with black curves. This resulted in two tilted curves in the 3D frequency-momentum space that are symmetric about the kx=0 plane. In the wave vector space across different frequencies, we identified four independent singularities with two opposite topological charges. By fixing the value of kxa/(2π) at 0.028 and plotting the copolarized reflection phase cuts in 3D frequency-momentum space, we observed an isolated vortex singularity. Additionally, when we fixed the values of kya/(2π) at 0.02 and 0.02, we detected a pair of vortex singularities with opposite topological charges, respectively. This observation is attributed to the fact that each topological dark line does not align routinely parallel to either the frequency-transverse wave vector plane or the frequency-longitudinal wave vector plane but rather exists in a tilt-symmetric manner. Notably, these two pairs of topological singularities are distinctly positioned in the figure, and their locations in the frequency-transverse wave vector plane do not coincide. This phenomenon aligns with the position and state of the topological dark lines depicted in the lower-left inset.

    Topological dark line diagram in frequency-momentum space and at frequency-transverse wave vector plane section and frequency-longitudinal wave vector plane section. The illustration at the lower-left corner of the diagram is the schematic illustration where the two parallel oblique topological dark lines are located in the section of the wave vector plane in the 3D frequency-momentum space, as shown in the polarization reflection amplitude diagram and phase diagram. Four phase vortices appear in the diagram with higher frequency. The phase vortices (±1) with opposite topological charges in the diagram all annihilate at the frequency of 1.1045 THz. The section diagrams of the topological dark line and the frequence-x wave vector plane when kya/(2π)=0.02 and kya/(2π)=−0.02, and the section diagrams of the topological dark line and the frequence-y wave vector plane when kxa/(2π)=−0.02 are, respectively, drawn.

    Figure 5.Topological dark line diagram in frequency-momentum space and at frequency-transverse wave vector plane section and frequency-longitudinal wave vector plane section. The illustration at the lower-left corner of the diagram is the schematic illustration where the two parallel oblique topological dark lines are located in the section of the wave vector plane in the 3D frequency-momentum space, as shown in the polarization reflection amplitude diagram and phase diagram. Four phase vortices appear in the diagram with higher frequency. The phase vortices (±1) with opposite topological charges in the diagram all annihilate at the frequency of 1.1045 THz. The section diagrams of the topological dark line and the frequence-x wave vector plane when kya/(2π)=0.02 and kya/(2π)=0.02, and the section diagrams of the topological dark line and the frequence-y wave vector plane when kxa/(2π)=0.02 are, respectively, drawn.

    The illustrations in the lower-left corner of Fig. 5 for frequencies above 1.1045 THz all contain points with zero copolarization reflectivity. Near 1.1045 THz, dark lines with opposite topological charges converge and subsequently annihilate. Additionally, Fig. 5 reveals two distinct directions of copolarized reflection phases, illustrated in the frequency-transverse wave vector tangent and frequency-longitudinal wave vector tangent of copolarization. These different orientations of the reflection phase cuts demonstrate variations in the number and position of frequency-momentum vortex singularities, indicative of changes in the number of vortex singularities within the space-time domain when observed from different directions. This observation first confirms the existence of two tilted, juxtaposed symmetric topological dark lines in the 3D frequency-momentum space for our proposed photonic crystal structure. Further, it supports the realization of differential changes in the number and nature of spatiotemporal vortex singularities across various directions in the spatiotemporal domain.

    VO2 functions as a temperature-sensitive material, with its complex dielectric constant and conductivity exhibiting temperature-dependent variations. To qualitatively assess the movement direction and the trend of change of topological dark lines as the temperature is decreased, we systematically altered the state of VO2, facilitating the transition of VO2 from a fully metallic state. Figures 6(a)–6(h) present the reflectance phase and reflectivity plots in a 3D frequency-wave vector spatial section at kxa/(2π)=0.02. As the conductivity of VO2 is sequentially varied to 200,000, 150,000, 100,000, and 70,000 S/m, the movement direction and trend of the vortex singularity in the phase cross-section are as illustrated in Figs. 6(a), 6(b), 6(e), and 6(f), while Figs. 6(c), 6(d), 6(g), and 6(h) depict the corresponding reflectivities. Notably, the range of singularity dispersion increases slightly as conductivity decreases. Our findings indicate that, when a reduction in temperature leads to a significant decrease in the conductivity of VO2, while still remaining above or equal to 70,000 S/m, the VO2 gradually transitions away from the fully metallic state. At this juncture, the vortex singularity in the cross-section progressively shifts upward along the path of maximum phase change rather than moving vertically.

    Graphical representation of the copolarized reflection phase and copolarized reflectivity obtained at different cuts by changing the vanadium dioxide state. (a), (b), (e), and (f) Copolarized reflection phase diagrams for a fixed observation cut at kxa/(2π)=−0.02, with VO2 conductivities of 200,000, 150,000, 100,000, and 70,000 S/m, respectively. (i) and (j) Copolarized reflection phase diagrams in the wave vector plane at an incident frequency of 1.105 THz, specifically for VO2 conductivities of 200,000 and 100,000 S/m. The red dots in these figures indicate the intersection points between the topological dark line and the wave vector plane. (c), (d), (g), and (h) Copolarized reflection amplitude spectra corresponding to (a), (b), (e), and (f), highlighting how variations in VO2 conductivity affect the maximum reflectivity and the position of singularities, with maximum reflectivity fluctuating within a range of 0.1. (k) and (l) Copolarized reflection amplitude related to (i) and (j).

    Figure 6.Graphical representation of the copolarized reflection phase and copolarized reflectivity obtained at different cuts by changing the vanadium dioxide state. (a), (b), (e), and (f) Copolarized reflection phase diagrams for a fixed observation cut at kxa/(2π)=0.02, with VO2 conductivities of 200,000, 150,000, 100,000, and 70,000 S/m, respectively. (i) and (j) Copolarized reflection phase diagrams in the wave vector plane at an incident frequency of 1.105 THz, specifically for VO2 conductivities of 200,000 and 100,000 S/m. The red dots in these figures indicate the intersection points between the topological dark line and the wave vector plane. (c), (d), (g), and (h) Copolarized reflection amplitude spectra corresponding to (a), (b), (e), and (f), highlighting how variations in VO2 conductivity affect the maximum reflectivity and the position of singularities, with maximum reflectivity fluctuating within a range of 0.1. (k) and (l) Copolarized reflection amplitude related to (i) and (j).

    However, this does not provide a clear indication of the movement direction of the two topological dark lines. To address this, we plotted the reflection phase diagrams in the wave vector plane at 1.105 THz, as illustrated in Figs. 6(i) and 6(j). In these figures, the red markers indicate the intersections of the topological dark lines with the wave vector plane. Figure 6(i) depicts the frequency of the incident circularly polarized light on VO2 at a conductivity of 200,000 S/m, with the copolarized reflection phase diagram fixed at 1.105 THz. At this frequency, both topological dark lines approach the annihilation frequency, resulting in the red dots in the figure being close to the plane defined by kya/(2π)=0. In contrast, Fig. 6(j) presents the reflection phase diagram at the same frequency but with the conductivity of VO2 set to 100,000 S/m. Although the four vortex singularities exhibit a similar outward shift, this shift is attributed to changes in the conductivity and complex permittivity of VO2, which fundamentally differs from the outward shift of the singularities caused by the increased curvature due to changes in the aperture angle θ of the photonic crystals, as shown in Fig. 4, as well as the frequency-induced shift observed in the inset of Fig. 5. This phenomenon is linked to the downward shift of the topological dark line, consistent with the trend observed in Fig. 2(d), which shows that the structure-guided resonance frequency decreases with decreasing conductivity. The trend of phase singularity position shifts in response to decreasing VO2 conductivity is also evident from the positions of the dark points in Figs. 6(c), 6(d), 6(g), and 6(h). Finally, Figs. 6(k) and 6(l) display the copolarized reflection amplitudes corresponding to Figs. 6(i) and 6(j), clearly illustrating the four singularities where the two topological dark lines intersect the wave vector plane.

    The observed phenomenon indicates that the topological dark line shifts in response to changes in the state of VO2. We posit that this shift is primarily attributed to the alteration in the substrate loss of VO2. Consequently, we conclude that, as the external temperature decreases and VO2 transitions gradually from a fully metallic state, the motion of the two topological dark lines in 3D frequency-momentum space trends downward and inward, as illustrated in Fig. 7.

    Schematic of the direction of movement of the topological dark line with temperature in 3D frequency-momentum space. The black curve in the figure is the topological dark line, and the red markers are the intersections of the topological dark line with the wave vector plane at a fixed frequency.

    Figure 7.Schematic of the direction of movement of the topological dark line with temperature in 3D frequency-momentum space. The black curve in the figure is the topological dark line, and the red markers are the intersections of the topological dark line with the wave vector plane at a fixed frequency.

    As the temperature continues to decrease, VO2 gradually transitions from a fully metallic state to an insulating state, at which point the conductivity decreases to 20 S/m, resulting in a shift in the system’s mode of operation to a transmission mode. Building on our previous discussion, we utilized linearly polarized light incident on the photonic crystal along the y direction and plotted the copolarized transmission phase in the frequency-wave vector plane. In the frequency-transverse wave vector plane with ky=0, we observe isolated vortex singularities, as illustrated in Figs. 8(a) and 8(e).

    Demonstration plots of singularity displacements due to different inclination angles θ. (a)–(d) Evolution of the transmission amplitude in the frequency-transverse wave vector plane at θ=15°, θ=22°, θ=28°, and θ=32°. (e)–(h) Phase evolution of the singularity in the frequency-transverse wave vector plane for angles θ=15°, θ=22°, θ=28°, and θ=32°, respectively. The black circle in the figure indicates the position of the phase singularity. As θ increases, the singularity’s location progressively approaches the Γ point. (i) Evolution of the topological singularity in frequency-momentum space as the inclination varies from 8° to 54°.

    Figure 8.Demonstration plots of singularity displacements due to different inclination angles θ. (a)–(d) Evolution of the transmission amplitude in the frequency-transverse wave vector plane at θ=15°, θ=22°, θ=28°, and θ=32°. (e)–(h) Phase evolution of the singularity in the frequency-transverse wave vector plane for angles θ=15°, θ=22°, θ=28°, and θ=32°, respectively. The black circle in the figure indicates the position of the phase singularity. As θ increases, the singularity’s location progressively approaches the Γ point. (i) Evolution of the topological singularity in frequency-momentum space as the inclination varies from 8° to 54°.

    To better reflect the real situation, we take into account the nonradiative losses of the material in our design. Simulations indicate that the singularity remains present. We further examine the evolution of this singularity in the frequency-transverse wave vector space for various inclination angles. In this analysis, we assume that the nonradiative loss remains constant with respect to θ. As θ increases gradually, the vortex singularity approaches the Γ point; however, the singularity and the vortex phase persist and continue to shift with the inclination angle θ.

    As the inclination angle θ increases, the C2 symmetry and z-mirror symmetry of the system are progressively broken. The location where the second condition of Eq. (A13) is satisfied approaches the Γ point. At other locations, while the first and third conditions of Eq. (A13) are met, the second condition remains unsatisfied. With an increasing inclination angle θ, the singularity appears at a specific position in the frequency-wave vector plane, as illustrated in Fig. 8(i). Notably, even with a significant increase in the inclination angle θ, the singularity position does not fully coincide with the Γ point. This deviation is attributed to nonradiative losses induced by the photonic crystal structure and the VO2 substrate within the system, which causes the topological singularity to shift in frequency-momentum space. This results in a change in the time-reversal symmetry condition of the system, causing topological singularities to deviate in frequency-momentum space. We verify this in Appendix C, which opens up new ideas and directions for the tunable properties of spatiotemporal vortices. Nevertheless, the singularity and the phase helix remain unaffected, further demonstrating that the singularity at this point is topologically robust within the 2D parameter space, as anticipated.

    4. CONCLUSION

    In this study, we investigate the generation and tunability of vortex singularities in the frequency-wave vector plane within the reflection and transmission channels, utilizing the phase-change tunable properties of VO2. Our proposed structure employs a circularly polarized beam to irradiate the photonic crystal structure based on topological dark point theory. In the 3D frequency-wave vector space of the reflection channel, we observe a pair of topological dark lines composed of topological dark points. Two phase vortices with opposite topological charges (±1) are identified in the frequency-transverse wave vector section, while a single independent phase vortex is noted in the frequency-longitudinal wave vector section. By tuning the conductivity and complex permittivity of VO2 through temperature adjustments, we successfully observe the moving trajectories of the topological dark lines and achieve tuning of the center frequency of the spatiotemporal vortex. When VO2 transitions to the transmissive mode due to temperature reduction, we generate independent phase vortices in the frequency-wave vector plane using line-polarized light incident on the photonic crystal structure. The position of the singularity approaches the Γ point as the inclination angle θ increases, and the singularity demonstrates topological robustness within the parameter space, allowing for conversion to STOV via Fourier optics. The photonic crystal structure presented in this study exhibits exceptional performance in generating spatiotemporal vortices and tuning functions, significantly enhancing the tunability and degrees of freedom in the generation of spatiotemporal vortices using a photonic crystal system. This advancement offers a novel pathway for the design of actively tunable spatiotemporal vortex generation devices.

    APPENDIX A: DERIVATION OF THE EXISTENCE CONDITIONS FOR TOPOLOGICAL SINGULARITIES USING THE TEMPORAL COUPLED-MODE THEORY

    We employ temporal coupled-mode theory (TCMT) to analyze the T-STOV phenomena of the system in both transmission and reflection modes [3739]. First, when the temperature of VO2 exceeds 68°C, vanadium dioxide enters a fully metallic state, and the system operates in reflection mode. Due to the asymmetry of the structural geometry, we include the resonances at +k and k in the formula, thus forming a two-resonance four-port TCMT model: dadt=(jω0γ)a+KT|s+,|s=S|s+=C|s++Da,in which a=(akak),|s+=(s1+s2+s3+s4+),|s=(s1s2s3s4).

    Here a is the 2×1 complex amplitude vector describing the two resonance modes at +k and k. The variable ω0 represents the resonance frequency, while γ denotes the radiation decay rate. The states |s+ and |s correspond to the input and output waves, respectively. K is the coupling matrix that connects the incident wave to the resonance mode, and D is the coupling matrix linking the resonance mode to the outgoing wave; both K and D are 4×2 matrices. C is a 4×4 scattering matrix that characterizes the scattering process in the absence of resonance, and it is a unitary symmetric matrix constrained by energy conservation and reciprocity. S is another 4×4 scattering matrix that describes the scattering process when resonance is present: KT=(κ10κ300κ20κ4),D=(0d1d200d3d40),C=ejϕ(0r0jtr0jt00jt0rjt0r0).

    In this context, let r and t represent the reflection and transmission coefficients, respectively. These coefficients satisfy the relationship r2+t2=1. The scattering matrix S is defined when the resonant mode is excited by an incident wave of frequency ω: S=C+DKTj(ωω0)+γ.

    The matrices K, D, and C satisfy the conditions due to the constraints of energy conservation and time-reversal symmetry conditions DD=2γ,CD*=K,K=D(0100100000010010).

    By coupling Eqs. (A6) and (A4), we obtain the form of the scattering matrix S, S=ejϕ(0r0jtr0jt00jt0rjt0r0)+1j(ωω0)+γ(0d1d20d1d4d2d10d2d300d3d20d3d4d4d10d4d30),and γ1+γ3=γ2+γ4=γ.

    When the incident beam is circularly polarized, the asymmetry in the structure is essential for facilitating polarization transitions and generating topological singularities. This understanding aids in establishing the conditions necessary for the generation of topological dark singularities [29], {rresonant=i(ω0ω)γ0i(ω0ω)+γ0·r,rnon-resonant=r;|rresonant|=|rnon-resonant|=|r|=1;arg(rresonantrnon-resonant)={0  (ω)π  (ω=ω0)2π  (ω+).

    Furthermore, the alteration in the VO2 substrate loss establishes the foundation for the active movement of the singularity.

    Then we consider that when the temperature drops to room temperature and the system operates in the transmission channel, we are able to obtain from Eq. (A6) ejϕ(rd2*+jtd4*)=d1,ejϕ(jtd2*+rd4*)=d3.

    From Eq. (A7), we obtain the transmission parameter S14 of the form S14=jtd4(rd2*+jtd4*)j(ωω0)+γ2+γ4.

    From Eqs. (A7), (A8), (A10), and (A11), the relationship between the transmission parameters and the attenuation rate of each port can be obtained: St=t(ωω0)±4r2t2γ2γ4(γ1r2γ2t2γ4)2tj(ωω0)+γ2+γ4+jt(γ2γ4)γ1r2γ2t2γ4tj(ωω0)+γ2+γ4.

    When the complex transmittance St is equal to zero, it is necessary to satisfy Re(St)=0 and Im(St)=0. This leads to the condition under which a transmission singularity is observed in the frequency-wave vector plane during transmission [40]: {(ωω0)2=4r2γ2γ4t2(γ2γ4)2γ1=γ2t0.

    For Eq. (A13), Condition 3 is satisfied under arbitrary circumstances. In Condition 1, the direct scattering process, specifically the Fabry–Perot background, can be balanced with the radiation imbalance by frequency tuning the difference ωω0 to achieve a specific resonance at any wave vector. Condition 2 requires that the system exhibits the same radiative attenuation rate at Port 1 (resonating at k) and Port 2 (resonating at +k). If the structure retains in-plane C2 symmetry and z-mirror symmetry, this condition will be automatically fulfilled, leading to a continuous nodal line. However, if the in-plane C2 symmetry and z-mirror symmetry are disrupted, Condition 2 is valid only at the Γ point (k=0), while the condition γ1=γ2 persists, resulting in an isolated singularity in the frequency-wave vector plane.

    APPENDIX B. TEMPERATURE RELATIONSHIP WHEN THE SYSTEM IS ACTIVELY TUNED

    As temperature increases, VO2 undergoes a phase transition from an insulator to a metal, resulting in a significant increase in its conductivity. This phase transition does not occur abruptly at a single temperature point; rather, it progresses gradually within a specific temperature range. The characteristics of this diffuse phase transition enable VO2 and its applications to exhibit smoother temperature control and adjustment capabilities in practical use. Within the temperature range of 300–340 K, the conductivity of VO2 rises from approximately 10 S/m to 2.7×105  S/m, demonstrating the material’s strong capacity for precise temperature regulation in devices [36]. We have plotted the changes in vanadium dioxide’s conductivity as it transitions from an insulating state to a metallic state and vice versa with temperature, as illustrated in Fig. 9 [41]. Upon returning to room temperature, vanadium dioxide’s conductivity stabilizes at 10 S/m, indicating that it is in an insulating state at this point.

    Conductivity of vanadium dioxide with temperature fluctuations.

    Figure 9.Conductivity of vanadium dioxide with temperature fluctuations.

    APPENDIX C. EFFECT OF TIME-REVERSED SYMMETRY BREAKING OF A SYSTEM ON SINGULARITY POSITION IN FREQUENCY-MOMENTUM SPACE

    In the process of simulating the transmission channel, we observe that, even with an increase in the inclination angle θ to a certain degree, the singularity does not coincide with the Γ point. This phenomenon can be attributed to the consideration of material losses in the simulation, indicating that the system is no longer lossless. The inclination angle θ influences the C2 symmetry and z-mirror symmetry of the structure, allowing us to manipulate the frequency-wave vector position of the topological singularity by adjusting θ. In addition to θ, the frequency condition and the wave vector condition under which the system satisfies time-reversal symmetry are critical factors affecting the position of the topological singularity. When utilizing the tunable phase change material vanadium dioxide, its inherent losses must be considered. As these losses are factored into the simulation, the system fails to maintain time-reversal symmetry, thereby precluding the possibility of achieving the same decay rate from the two radiation channels at the Γ point. Consequently, the topological singularity deviates from the Γ point in the frequency-wave vector plane. To validate our hypothesis and minimize the potential for result variability, we employ an alternative system capable of generating topological singularities in the frequency-momentum space. We examine the impacts on the frequency and wave vector properties of the topological singularities by disrupting the system’s time-reversal symmetry through the application of a magnetic field. The resulting positions of the singularity in the frequency-wave vector plane are illustrated in Fig. 10.

    The time reversal symmetry of the system is broken after the addition of an external magnetic field, resulting in changes in the position of the topological singularity. (a) Phase of topological singularities in frequency-momentum plane without an external magnetic field. (a) Phase of the frequency-momentum plane topological singularities when an external magnetic field is added.

    Figure 10.The time reversal symmetry of the system is broken after the addition of an external magnetic field, resulting in changes in the position of the topological singularity. (a) Phase of topological singularities in frequency-momentum plane without an external magnetic field. (a) Phase of the frequency-momentum plane topological singularities when an external magnetic field is added.

    The results show that, when the time-reversal symmetry of the system is broken, the frequency and wave vector properties of the topological singularities will change, which not only validates our idea but also provides a new way to actively tune the properties of space-time vortices.

    APPENDIX D. DEVICE FABRICATION AND TOLERANCE ANALYSIS

    Our device can be fabricated through the following process. First, a VO2 layer is grown on a silica substrate using radiofrequency plasma-assisted oxide molecular beam epitaxy (MBE) technology. Then, using plasma-enhanced chemical vapor deposition, we sequentially grow SiO2 spacer layers (52.5μm) and Si3N4 layers (37.5μm). The metasurface microstructure sequence is then fabricated by spin-coating a layer of CSAR 62 positive electron-beam resist on the Si3N4 layer photon crystal patterns written into the resist using electron-beam lithography. After development, the patterned resist serves as an etching mask. The pattern is then etched into the Si3N4 layer using a CHF3 and O2 gas-based reactive ion etching (RIE) process. Finally, the remaining resist mask is removed by O2 plasma via RIE. The general process is shown in Fig. 11.

    The metasurface machining process of temporal and spatial vortices in this paper.

    Figure 11.The metasurface machining process of temporal and spatial vortices in this paper.

    To determine the tolerance range suitable for our proposed device, we conducted simulations to obtain the maximum tolerances necessary to meet the specified functional requirements. This investigation primarily focused on the length and width of the groove: the groove width tolerance is approximately ±10  μm, while the length tolerance is around ±5  μm. The stringent tolerance requirements for the groove length, compared to the width, arise from its significant impact on the coupling strength of the device with linearly polarized light in the transmission channel as well as its effect on the breaking of the C2 symmetry of the system. If the groove length is too short, it may result in the disappearance of a pair of topological dark lines in the reflection channel. Conversely, if the length is excessively long, it can complicate the topological singularity phase of the transmission channel, making it more challenging to generate the space-time vortex beam. The tolerance requirements proposed in this paper are not stringent given the current machining processes and exhibit a relatively broad applicability.

    References

    [1] L. Allen, M. W. Beijersbergen, R. J. C. Spreeuw. Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes. Phys. Rev. A, 45, 8185-8189(1992).

    [2] E. Maguid, I. Yulevich, D. Veksler. Photonic spin-controlled multifunctional shared aperture antenna array. Science, 352, 1202-1206(2016).

    [3] L. Paterson, M. P. MacDonald, J. Arlt. Controlled rotation of optically trapped microscopic particles. Science, 292, 912-914(2001).

    [4] K. I. Willig, S. O. Rizzoli, V. Westphal. STED microscopy reveals that synaptotagmin remains clustered after synaptic vesicle exocytosis. Nature, 440, 935-939(2006).

    [5] N. Bozinovic, Y. Yue, Y. Ren. Terabit-scale orbital angular momentum mode division multiplexing in fibers. Science, 340, 1545-1548(2013).

    [6] R. C. Devlin, A. Ambrosio, N. A. Rubin. Arbitrary spin-to-orbital angular momentum conversion of light. Science, 358, 896-901(2017).

    [7] Y. Shen, X. Wang, Z. Xie. Optical vortices 30 years on: OAM manipulation from topological charge to multiple singularities. Light Sci. Appl., 8, 90(2019).

    [8] K. R. Patel, T. Paramanathan. Exploring the binding of mitoxantrone to DNA using optical tweezers. Biophys. J., 122, 75A(2023).

    [9] X. Y. Fang, H. R. Ren, K. Y. Li. Nanophotonic manipulation of optical angular momentum for high-dimensional information optics. Adv. Opt. Photonics, 13, 772-833(2021).

    [10] N. Lal, S. Mishra, A. Rani. Polarization-orbital angular momentum duality assisted entanglement observation for indistinguishable photons. Quantum Inf. Process., 22, 90(2023).

    [11] H. Yao, Z. Sun, L. Liang. Hybrid metasurface using graphene/graphitic carbon nitride heterojunctions for ultrasensitive terahertz biosensors with tunable energy band structure. Photonics Res., 11, 858-868(2023).

    [12] H. Yao, M. Yang, X. Yan. Patterned graphene and terahertz metasurface-enabled multidimensional ultra-sensitive flexible biosensors and bio-assisted optical modulation amplification. Results Phys., 40, 105884(2022).

    [13] C. Huang, L. Liang, P. Chang. Terahertz liquid biosensor based on a graphene metasurface for ultrasensitive detection with a quasi-bound state in the continuum. Adv. Mater., 36, 2310493(2024).

    [14] L. Liang, Y. Zhang, C. Huang. Metamaterial flexible GaN/graphene heterostructure-enabled multidimensional terahertz sensor for femtogram-level detection of aspartic acid. IEEE Sensors J., 23, 16814-16822(2023).

    [15] H. Yao, Z. Sun, X. Yan. Ultrasensitive, light-induced reversible multidimensional biosensing using THz metasurfaces hybridized with patterned graphene and perovskite. Nanophotonics, 11, 1219-1230(2022).

    [16] K. Y. Bliokh, F. Nori. Spatiotemporal vortex beams and angular momentum. Phys. Rev. A, 86, 033824(2012).

    [17] M. A. Porras. Transverse orbital angular momentum of spatiotemporal optical vortices. Prog. Electromagn. Res, 177, 95-105(2023).

    [18] N. Jhajj, I. Larkin, E. W. Rosenthal. Spatiotemporal optical vortices. Phys. Rev. X, 6, 031037(2016).

    [19] S. W. Hancock, S. Zahedpour, A. Goffin. Freespace propagation of spatiotemporal optical vortices. Optica, 6, 1547-1553(2019).

    [20] S. W. Hancock, S. Zahedpour, H. M. Milchberg. Mode structure and orbital angular momentum of spatiotemporal optical vortex pulses. Phys. Rev. Lett., 127, 193901(2021).

    [21] Q. Cao, J. Chen, K. Lu. Nonspreading Bessel spatiotemporal optical vortices. Sci. Bull., 67, 133-140(2022).

    [22] M. Yessenov, L. A. Hall, K. L. Schepler. Spacetime wave packets. Adv. Opt. Photonics, 14, 455-570(2022).

    [23] N. Dror, B. A. Malomed. Symmetric and asymmetric solitons and vortices in linearly coupled two-dimensional waveguides with the cubicquintic nonlinearity. Phys. D, 240, 526-541(2011).

    [24] Y. Shen, Q. Zhan, L. G. Wright. Roadmap on spatiotemporal light fields. J. Opt., 25, 093001(2023).

    [25] A. Chong, C. Wan, J. Chen. Generation of spatiotemporal optical vortices with controllable transverse orbital angular momentum. Nat. Photonics, 14, 350-354(2020).

    [26] H. Wang, C. Guo, W. Jin. Engineering arbitrarily oriented spatiotemporal optical vortices using transmission nodal lines. Optica, 8, 966-971(2021).

    [27] H. Ge, S. Liu, X.-Y. Xu. Spatiotemporal acoustic vortex beams with transverse orbital angular momentum. Phys. Rev. Lett., 131, 014001(2023).

    [28] J. Huang, J. Zhang, T. Zhu. Spatiotemporal differentiators generating optical vortices with transverse orbital angular momentum and detecting sharp change of pulse envelope. Laser Photonics Rev., 16, 2100357(2022).

    [29] W. Liu, J. Wang, Y. Tang. Exploiting topological darkness in photonic crystal slabs for spatiotemporal vortex generation. Nano Lett., 24, 943-949(2024).

    [30] A. I. Kashapov, E. A. Bezus, D. A. Bykov. Plasmonic generation of spatiotemporal optical vortices. Photonics, 10, 109(2023).

    [31] H. Zhang, Y. Sun, J. Huang. Topologically crafted spatiotemporal vortices in acoustics. Nat. Commun., 14, 6238(2023).

    [32] S. Divitt, W. Zhu, C. Zhang. Ultrafast optical pulse shaping using dielectric metasurfaces. Science, 364, 890-894(2019).

    [33] A. M. Shaltout, K. G. Lagoudakis, J. van de Groep. Spatiotemporal light control with frequency gradient metasurfaces. Science, 365, 374-377(2019).

    [34] X. Dong, X. Luo, Y. Zhou. Switchable broadband and wide-angular terahertz asymmetric transmission based on a hybrid metal-VO2 metasurface. Opt. Express, 28, 30675(2020).

    [35] Q. Z. Wang, S. Y. Liu, G. J. Ren. Multi-parameter tunable terahertz absorber based on graphene and vanadium dioxide. Opt. Commun., 494, 127050(2021).

    [36] Q.-Y. Wen, H. W. Zhang, Q. H. Yang. Terahertz metamaterials with VO2 cut-wires for thermal tenability. Appl. Phys. Lett., 97, 021111(2010).

    [37] S. Fan, W. Suh, J. D. Joannopoulos. Temporal coupled-mode theory for the Fano resonance in the optical resonator. J. Opt. Soc. Am., 20, 569-572(2003).

    [38] W. Suh, Z. Wang, S. Fan. Temporal coupled-mode theory and the presence of non-orthogonal modes in lossless multimode cavities. IEEE J. Quantum Electron., 40, 1511-1518(2004).

    [39] S. Fan, J. D. Joannopoulos. Analysis of guided resonances in photonic crystal slabs. Phys. Rev. B, 65, 235112(2002).

    [40] P. Huo, W. Chen, Z. Zhang. Observation of spatiotemporal optical vortices enabled by symmetry-breaking slanted nanograting. Nat. Commun., 15, 3055(2024).

    [41] P. U. Jepsen, B. M. Fischer, A. Thoman. Metal-insulator phase transition in a VO2 thin film observed with terahertz spectroscopy. Phys. Rev. B, 74, 205103(2006).

    Fangze Deng, Ke Ma, Yumeng Ma, Xiang Hou, Zhihua Han, Yuchao Li, Keke Cheng, Yansheng Shao, Chenglong Wang, Meng Liu, Huiyun Zhang, Yuping Zhang, "Dual-channel tunable multipolarization adapted terahertz spatiotemporal vortices generating device," Photonics Res. 13, 1408 (2025)
    Download Citation