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

Method for generating spatiotemporal coherency vortices and spatiotemporal dislocation curves

Open Access Open Access

Abstract

A simple method for designing spatiotemporal coherency vortices (STCVs) and spatiotemporal dislocation curves (STDCs) is introduced by means of coherent-mode representation and Fourier transforms. A partially coherent pulsed beam is represented by an incoherent superposition of a Gaussian and a Hermite-Gaussian pulsed beam with different waist positions. It well demonstrates that there exist STCVs and STDCs in the space-time plane. The detailed numerical calculations are performed to address the dependence of waist distance of two modes, reference position, beam order, distribution of original spectrum, topological charge and mode weights ratio on the STCVs and STDCs. The physical interpretation behind numerical results is shown. A possible scheme for experimental synthesis of the STCVs is proposed. The obtained results may have potential applications in the fields of light-matter interaction, spatiotemporal spin-orbit angular momentum coupling and STCV-based optical trapping and optical manipulation.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Optical fields carrying orbital angular momentum (OAM) have acquired and sustained, throughout the last three decades, an unprecedented attention from the scientific community, after Allen et al. introduced the Laguerre-Gaussian(LGpl) modes [1] with vortices of nonzero azimuthal index l. A classic optical vortex is pertinent to a beam-like field possessing a spiral phase in the transverse plane to the direction of propagation, say x-y plane, while the OAM vector is longitudinal, i.e., parallel to that direction. The beams carrying longitudinal OAM have already found numerous applications in classical and quantum optics, such as optical communications [2], optical tweezers [3], optical detection [4], quantum information processing and imaging [5,6], etc.

Recently, optical fields carrying transverse OAM, i.e., those in which the OAM vector is orthogonal to the direction of propagation has gained a great deal of attention. Optical vortices with transverse OAM were shown to exist in the x-t plane (or, equivalently, the x-z plane) and are referred to as spatiotemporal optical vortices (STOVs) [7]. While propagating, the STOVs undergo dispersion and diffraction effects simultaneously. The STOV is not specific to electromagnetic radiation: a number of natural aero- and hydro-dynamic phenomena such as tropical cyclones, moving tornadoes and human left ventricular blood flow are the examples of the mechanical analogues of the STOV. Bliokh and Nori pointed out that the STOV states could describe either photon or electron vortex states [8,9]. In 2016, Jhajj et al. presented the first experimental evidence of the STOVs and revealed that they constitute the fundamental element of the nonlinear collapse and are responsible for the subsequent propagation of short optical pulses in air [10]. More recently, the STOVs were generated experimentally in free space [11,12]. Bliokh then carried out an accurate theoretical analysis of such STOV states and their angular momentum properties [13]. At the same time, Huang et al. theoretically and experimentally demonstrated generation, propagation and diffraction of the STOVs by the use of 4f pulse shaper system [14,15]. Wan et al. transferred the STOV tube into a vortex ring, hence introducing toroidal vortices, through optical conformal mapping [16], and Chen et al. reported the spatiotemporal Bessel vortices with transverse OAM having topological charges beyond 100 [17]. More recently, Cao et al. showed the propagation of the STOVs through few-mode fiber and found the possible application in telecommunications [18]. However, all of the aforementioned investigations are confined to fully coherent pulsed sources.

On the other hand, it has also been realized that coherence of light plays the quintessential role for structured light propagation [1921], especially in long-distance communications [22], optical trapping [23] and information encryption [24]. It is known that on interaction of deterministic light with optically fluctuating media, a classic optical vortex breaks down giving a rise to a multitude of coherency vortices with different indices [25]. It is also possible to synthesize light sources carrying a single or a finite set of coherence vortices with specified indices [26]. Theoretically, the coherency vortices are described as phase singularities of the second-order correlation functions [2634]. Recently a unified approach for complete characterization of the OAM in stationary light beams via the coherence-OAM matrices whose elements represent self and joint radial correlations among the pairs of the OAM indices was achieved [35,36].

Meanwhile, optical vortices with transverse OAM embedded in partially coherent pulsed beams, are called spatiotemporal coherency vortices (STCVs) [37]. With the rapid development of ultrashort pulse technology, partially spectral or temporal coherence states in the random optical pulses have drawn widespread attention because of their groundbreaking applications in pulse shaping, ghost imaging, laser micromachining, medical diagnosis and treatment [3850]. Till now, the STCVs have not been investigated thoroughly, with a few exceptions: in 2020, Hyde found out that the twisted space-frequency and space-time partially coherent pulsed sources could carry transverse OAM due to the statistical twists between their spatial and temporal dimensions [51]; in 2021, Mirando et al. demonstrated that a STCV can be generated from a pulsed source with partial temporal coherence and complete spatial coherence [52]; more recently, we investigated the coherence control of STCVs and found that source coherence makes it possible to control the dynamics of phase distributions and the positions of STCVs upon propagation [37]. However, these investigations of the STCVs relating to coherence still have some limitations. For instance, the currently known STCVs are limited to the transverse OAM with low topological charge values, l = 1 or 2. Moreover, their theoretical description remains far from trivial. Further, it is currently of considerable interest how to realize a STCV with an arbitrary value of topological charge in a simple manner.

In this paper we have explored a method for generating STCVs with the help of coherent-mode representation, which can readily deal with the high-order STCVs. We also have discovered the existence of spatiotemporal dislocation curve (STDC) in the x-t plane, in the partially coherent pulsed field. Usually, for stationary fully coherent light field, there are two kinds of pure dislocations. One is the screw dislocation, often called the optical vortex, which is a spiral phase ramp around a dark spot where the phase of the field is undefined and thus its amplitude is equal to zero; the other is the edge dislocation located along a line in the transverse x-y plane across which a π phase shift appears [53,54]. However, for partially coherent light field, the screw dislocation is often called the coherency vortex, which takes place at the position where the modulus of cross-spectral density of the field is equal to zero [55]. The edge dislocation also located along a line in the transverse x-y plane across which a π phase shift appears [56]. Analogous to the above definition, for partially coherent pulsed field, STDC take places in transverse x-t plane across which a π phase shift arise. Physically, along the STDC line, the modulus of mutual coherence function of partially coherent pulsed beam is equal to zero. In Section 2 we show the basic approach to modeling and present an analytical formula for describing the STCVs. The detailed numerical results and corresponding physical explanations are given in Section 3. Finally, in Section 4 we summarized our results and draw the main conclusion.

2. Theoretical approach

Consider a partially coherent pulsed beam, whose cross-spectral density can be expressed as an incoherent superposition of a number of coherent modes by means of eigenfunctions ϕmnk(r, ω, z; z0) [57], i.e.,

$$W({{\boldsymbol r}_1},{{\boldsymbol r}_2},{\omega _1},{\omega _2},{z_1},{z_2}) = \sum\limits_{m,n,k}^{} {{\lambda _{mnk}}\phi _{mnk}^\ast ({{\boldsymbol r}_1},{\omega _1},{z_1};{z_0}){\phi _{mnk}}({{\boldsymbol r}_2},{\omega _2},{z_2};{z_0})}, $$
where mode weighs λmnk are necessarily real and non-negative, and the subscripts m, n, k may represent more than one index of summation. The eigenfunctions will be set as Hermite-Gaussian modes, defined as
$$\begin{aligned} {\phi _{mnk}}({\boldsymbol r},\omega ,z;{z_0}) &= {\phi _k}(\omega ){\left( {\frac{2}{\pi }} \right)^{1/2}}\frac{{\exp \{{i[{{{\omega (z\textrm{ - }{z_0})} / c} - ({m + n + 1} )\theta ({z\textrm{ - }{z_0},\omega } )} ]} \}}}{{w(z\textrm{ - }{z_0},\omega )\sqrt {{2^{m + n}}m!n!} }}\\ & \times {H_m}\left[ {\frac{{x\sqrt 2 }}{{w(z\textrm{ - }{z_0},\omega )}}} \right]{H_n}\left[ {\frac{{y\sqrt 2 }}{{w(z\textrm{ - }{z_0},\omega )}}} \right]\exp \left[ {i\frac{{\omega ({x^2} + {y^2})}}{{2cR(z\textrm{ - }{z_0},\omega )}} - \frac{{{x^2} + {y^2}}}{{{w^2}(z\textrm{ - }{z_0},\omega )}}} \right] \end{aligned}, $$
$$w({z,\omega } )= {w_0}\sqrt {1 + {{\left( {\frac{{2cz}}{{{w_0}{w_c}\omega }}} \right)}^2}}, $$
$$R({z,\omega } )= z\left[ {1 + {{\left( {\frac{{{w_0}{w_c}\omega }}{{2cz}}} \right)}^2}} \right], $$
$$\tan \theta ({z,\omega } )= \frac{{{w^2}(z,\omega )\omega }}{{2cR(z,\omega )}} = \frac{{2cz}}{{w_c^2\omega }}, $$
$$\frac{1}{{w_c^2}} = \frac{1}{{w_0^2}} + \frac{1}{{\sigma _0^2}}, $$
$${\phi _k}(\omega ) = \frac{1}{{\sqrt {{2^k}k!} }}{\left( {\frac{{2d}}{\pi }} \right)^{{1 / 4}}}{H_k}\left[ {(\omega - {\omega_0})\sqrt {2d} } \right]\exp [{ - d{{(\omega - {\omega_0})}^2}} ], $$
$$d = {({{a^2} + 2ab} )^{{1 / 2}}}, $$
$$a = {1 / {2\Omega _0^2}},b = {1 / {2\Omega _c^2}}, $$
$$\Omega _0^2 = {1 / {T_0^2}} + {2 / {T_c^2}},{\Omega _c} = {{{{T_c}{\Omega _0}} / T}_0}, $$
where Hj indicates the Hermite polynomial of order j, r = (x, y) is the 2D transverse position vector, z is the propagation distance, z0 denotes the position of the waist plane of the mode, c is the speed of light in vacuum. Further, w0 defines the width of the mode at the waist, while w(z, ω) gives the width of the mode at distance z and frequency ω; R(z, ω) represents the radius of curvature of the equiphase surfaces, and θ(z, ω) is the a longitudinal phase. The modes in Eqs. (2)-(10) represent pulsed beams paraxially propagating into the half-space z > 0, along the z-axis. In addition, ϕk(ω) is the spectrum distribution of the pulsed beam; ω0 is the pulse’s carrier frequency, while σ0 denotes the beam’s spatial coherence width; Ω0 and Ωc are the spectral width and the spectral coherence width of the pulsed beam, respectively. Also, T0 and Tc characterize the average pulse duration and its coherence time.

In time-space domain, the mutual coherence function of a partially coherent pulsed beam has form

$$\Gamma ({{\boldsymbol r}_1},{{\boldsymbol r}_2},{t_1},{t_2},{z_1},{z_2}) = \sum\limits_{m,n,k}^{} {{\lambda _{mnk}}\varphi _{mnk}^\ast ({{\boldsymbol r}_1},{t_1},{z_1};{z_0}){\varphi _{mnk}}({{\boldsymbol r}_2},{t_2},{z_2};{z_0})}$$
where φmnk(r, t, z; z0) is the Fourier transform of ϕmnk(r, ω, z; z0), expressed by FT[·], i.e.,
$${\varphi _{mnk}}({\boldsymbol r},t,z;{z_0}) = FT[{\phi _{nnk}}({\boldsymbol r},\omega ,z;{z_0})] = \int\limits_{ - \infty }^\infty {{\phi _{nnk}}({\boldsymbol r},\omega ,z;{z_0})} \exp [ - i\omega t]d\omega. $$

Let us consider a STOV phase with topological charge q[11],

$${\left[ {\frac{{{t_{}}}}{{{t_s}}} + i\textrm{sgn}(q) \cdot \frac{{x_{}^{}}}{{x_s^{}}}} \right]^{|q |}}$$
where, t, x are time instant and space coordinate, respectively; sgn (•) is the signum function; and τs and xs are the temporal and the spatial scale widths of the STOV, respectively. On imbedding the STOV phase into the eigenfunction φmnk(r, t, z; z0) at the z plane, we obtain the new eigenfunction φ'mnk(r, t, z; z0):
$${\varphi ^{\prime}_{mnk}}({\boldsymbol r},t,z;{z_0}) = {\varphi _{mnk}}({\boldsymbol r},t,z;{z_0}){\left[ {\frac{{{t_{}}}}{{{t_s}}} + i\textrm{sgn}({q_{mnk}}) \cdot \frac{{x_{}^{}}}{{x_s^{}}}} \right]^{|{{q_{mnk}}} |}}. $$

Thus, the mutual coherence function of a partially coherent pulsed beam with the single STCV phase of topological index qmnk becomes

$$\scalebox{0.86}{$\begin{aligned} &\Gamma ^{\prime}({{\boldsymbol r}_1},{{\boldsymbol r}_2},{t_1},{t_2},{z_1},{z_2}) = \sum\limits_{m,n,k}^{} {{\lambda _{mnk}}\varphi ^{\prime\ast}_{mnk} ({{\boldsymbol r}_1},{t_1},{z_1};{z_0}){{\varphi ^{\prime}}_{mnk}}({{\boldsymbol r}_2},{t_2},{z_2};{z_0})} \\ &= \sum\limits_{m,n,k}^{} {{\lambda _{mnk}}\varphi _{mnk}^\ast ({{\boldsymbol r}_1},{t_1},{z_1};{z_0}){\varphi _{mnk}}({{\boldsymbol r}_2},{t_2},{z_2};{z_0})} {\left[ {\frac{{{t_1}}}{{{t_s}}}\textrm{ - }i\textrm{sgn}({q_{mnk}}) \cdot \frac{{x_1^{}}}{{x_s^{}}}} \right]^{|{{q_{mnk}}} |}}{\left[ {\frac{{{t_2}}}{{{t_s}}} + i\textrm{sgn}({q_{mnk}}) \cdot \frac{{x_2^{}}}{{x_s^{}}}} \right]^{|{{q_{mnk}}} |}}. \end{aligned}$}$$

From Eq. (15), one can predict the average intensity and the coherence state of a partially coherent pulsed beam according to the second-order coherence theory [58]. In our calculation, for simplicity, let us consider partially coherent pulsed beams consisting of only two modes, i.e., with a mutual coherence function given by the expression [28]

$$\scalebox{0.86}{$\begin{aligned} &\Gamma ^{\prime}({{\boldsymbol r}_1},{{\boldsymbol r}_2},{t_1},{t_2},{z_1},{z_2})\\ &= {\lambda _1}{\{{FT[{\phi_{000}}({{\boldsymbol r}_1},{\omega_1},{z_1};0)]} \}^{\ast }}FT[{\phi _{000}}({{\boldsymbol r}_2},{\omega _2},{z_2};0)]{\left[ {\frac{{{t_1}}}{{{t_s}}}\textrm{ - }i\textrm{sgn}({q_{000}}) \cdot \frac{{x_1^{}}}{{x_s^{}}}} \right]^{|{{q_{000}}} |}}{\left[ {\frac{{{t_2}}}{{{t_s}}} + i\textrm{sgn}({q_{000}}) \cdot \frac{{x_2^{}}}{{x_s^{}}}} \right]^{|{{q_{000}}} |}}\\ &+ {\lambda _2}{\{{FT[{\phi_{mnk}}({{\boldsymbol r}_1},{\omega_1},{z_1};{z_0})]} \}^{\ast }}FT[{\phi _{mnk}}({{\boldsymbol r}_2},{\omega _2},{z_2};{z_0})]{\left[ {\frac{{{t_1}}}{{{t_s}}}\textrm{ - }i\textrm{sgn}({q_{mnk}}) \cdot \frac{{x_1^{}}}{{x_s^{}}}} \right]^{|{{q_{mnk}}} |}}{\left[ {\frac{{{t_2}}}{{{t_s}}} + i\textrm{sgn}({q_{mnk}}) \cdot \frac{{x_2^{}}}{{x_s^{}}}} \right]^{|{{q_{mnk}}} |}}, \end{aligned}$}$$
where the ϕmnk(r, ω, z; z0) are defined in Eq. (2). Equation (16) represents an incoherent superposition of a Gaussian pulsed beam (mode 1) with waist plane at z = 0 and a Hermite-Gaussian pulsed beam (mode 2) of order mnk with waist plane at z = z0. Here we choose different waist planes in order to avoid a real-valued mutual coherence function in the waist plane. The phase singularities of Γ(r1, r2, t1, t2, z1, z2) would then generate in the space-time domain. Without loss of generality, we confine our attention to one-dimensional case, the two-dimensional extension being straightforward because of separability in x-y plane. In that case, the expressions of the first and second mode can be written as
$${\phi _{000}}({\boldsymbol r},\omega ,z;0) = {\left( {\frac{{2d}}{\pi }} \right)^{{1 / 4}}}\exp [{ - d{{(\omega - {\omega_0})}^2}} ]{\left( {\frac{2}{\pi }} \right)^{1/2}}\frac{1}{{{w_0}}}\exp \left[ { - \frac{{{x^2}}}{{w_0^2}}} \right], $$
and
$$\begin{aligned} {\phi _{m0k}}({\boldsymbol r},\omega ,z;{z_0}) &= {\left( {\frac{{2d}}{\pi }} \right)^{{1 / 4}}}\frac{1}{{\sqrt {{2^k}k!} }}{H_k}\left[ {(\omega - {\omega_0})\sqrt {2d} } \right]\exp [{ - d{{(\omega - {\omega_0})}^2}} ]\\ & \times {\left( {\frac{2}{\pi }} \right)^{1/2}}\frac{{\exp \{{i[{{{\omega (z\textrm{ - }{z_0})} / c} - ({m + 1} )\theta ({z\textrm{ - }{z_0},\omega } )} ]} \}}}{{w(z\textrm{ - }{z_0},\omega )\sqrt {{2^m}m!} }}\\ & \times {H_m}\left[ {\frac{{x\sqrt 2 }}{{w(z\textrm{ - }{z_0},\omega )}}} \right]\exp \left[ {i\frac{{\omega {x^2}}}{{2cR(z\textrm{ - }{z_0},\omega )}} - \frac{{{x^2}}}{{{w^2}(z\textrm{ - }{z_0},\omega )}}} \right] \end{aligned}, $$
respectively. Using Eq. (16), the STCVs of partially coherent pulsed beams can be investigated by means of the zero level curves in the real and imaginary parts of mutual coherence functions Γ(x1, x2, t1, t2, z1, z2). Resembling the definition of coherence vortices in the stationary beam [28], the position of the STCVs is hence determined by relations
$${\textrm{Re}} [{\Gamma^{\prime}({x_1},{x_2},{t_1},{t_2},{z_1},{z_2})} ]= 0,$$
$${\mathop{\rm Im}\nolimits} [{\Gamma^{\prime}({x_1},{x_2},{t_1},{t_2},{z_1},{z_2})} ]= 0,$$
where Re[·] and Im[·] indicate the real and imaginary parts of a complex number, respectively. The topological charge and its sign are determined by the vorticity of phase contours around the singularity [59]. The spatiotemporal intensity I′(x2, t2, z2) can be obtained from Γ(x1, x2, t1, t2, z1, z2) when x1 = x2, t1 = t2, z1 = z2.

3. Generation of the STCVs and STDC

By use of Eqs. (16)-(20), the detailed numerical calculation related to the STCVs and spatiotemporal dislocation of partially coherent pulsed beam are shown for fixed x1, t1 and a variety of values of x2, t2, the calculation being done in a single z-plane, i.e., z1 = z2. The calculation parameters are z1 = z2 = 0, z0=-0.1 m, λ=800 nm, c = 3 × 108 m/s, w0 = σ=2 mm, T0 = Tc = 5fs, xs = 1 mm, τs = 1fs, x1 = 1 mm, t1 = 1fs, k = 0, q000 = qmnk = q = 1, λ1 + λ2 = 1, λ1 = 0.5, λ2 = 0.5 and m = 1 unless other values are specified.

Figure 1 shows the phase distribution of Γ, curves of Re Γ′ = 0 (solid line) and Im Γ′ = 0 (dashed line), modulus distribution of Γ and spatiotemporal intensity I′ of a partially coherent pulsed beam with topological charge q = + 1 for different waist positions of two modes. One sees from Fig. 1(a) that when z0=-0.1 m, the STCV takes place in the spatiotemporal center (x2 = 0, t2 = 0), while the STDC also appears in the x2-t2 plane. The positions of STCV and STDC appear in the zeros of modulus of Γ rather than that of the spatiotemporal intensity (see Fig. 1(i) and Fig. 1(m)). From Eqs. (16) and (17), one can also see that the spatiotemporal intensity of a partially coherent pulsed beam with Eq. (16) is nonzero throughout the time-space plane because the mode ϕ000(r, ω, z; z0) of Eq. (17) has no zeros. These properties are similar to the case of coherent vortices in stationary partially beams [28]. When the distances between waist positions of the two modes become larger, the position of the STCV is always in the spatiotemporal center, but the STDCs gradually reduce, weakening and vanishing for negative t2 (see Figs. 1(f)-(h)). Further numerical calculation indicates that the STDCs disappear when z0<-3 m. This occurs because with increasing distance the coupling between two modes becomes weak. Nevertheless, the STCV always exists in the spatiotemporal center because the spatiotemporal vortex phase exists in modes 1 and mode 2, as shown in Eq. (16).

 figure: Fig. 1.

Fig. 1. (a)-(d) Phase distribution of Γ, (e)-(h) level curves of Re Γ′ = 0 (solid line) and Im Γ′ = 0 (dashed line), (i)-(l) modulus distribution of Γ and (m)-(p) spatiotemporal intensity I′ of a partially coherent pulsed beam with topological charge q = + 1 for different waist positions of two modes.

Download Full Size | PDF

Figure 2 gives the phase and modulus distributions of Γ of a partially coherent pulsed beam with topological charge q = + 1 for different reference positions x1, t1. As can be seen from Fig. 2, reference positions significantly affect the position of the STDC, but do not substantially influence the STCV itself. When the reference positions are gradually separated from the spatiotemporal center, the STDC gets closer to the STCV. The STDC and the STCV coincide when x1 = 10 mm, t1 = 10fs. At the same time, in the vicinity of the STDC, the modulus of Γ gets larger with increasing reference positions (see Figs. 2(d)-(f)).

 figure: Fig. 2.

Fig. 2. (a)-(c) Phase distribution and (d)-(f) Modulus distribution of Γ of a partially coherent pulsed beam with topological charge q = + 1 for different reference positions x1, t1.

Download Full Size | PDF

Figure 3 shows the phase and the modulus distributions of Γ of a partially coherent pulsed beam with topological charge q = + 1 for different beam order values of mode 2, where the reference position is x1 = 10 mm, t1 = 10fs. From Figs. 3(a)-(f), there is no STDC when m = 0. There are 1, 2, 3, 4, 5 STDCs when beam order m = 1, 2, 3, 4, 5, respectively i.e., the number of the STDCs coincides with m. Also, the modulus of Γ distribution is almost symmetric about line x2 = 0. These results can be explained as follows: in Eq. (18) the second-mode eigenfunction ϕm0 k(r, ω, z; z0) includes Hermite polynomial Hm(·), which holds m solutions for equation Hm(x) = 0 corresponding to m STDCs. At the same time, Fig. 3 implies that the spatiotemporal dislocations result from the mode 2.

 figure: Fig. 3.

Fig. 3. (a)-(f) Phase distribution and (g)-(l) modulus distribution of Γ of a partially coherent pulsed beam with topological charge q = + 1 for different beam order of the mode 2.

Download Full Size | PDF

Figure 4 presents the phase and modulus distributions of Γ of a partially coherent pulsed beam with topological charge q = + 1 for different values of parameter k. It is shown that k parameter has quite significant influence on the STDCs, which are cut off and become k + 1 segments according to the different k values. This is also a result of spatiotemporal coupling. This can be explained by use of Eq. (7). From Eq. (7), the original spectrum of the pulsed beams is $S(\omega ) = {|{{\phi_k}(\omega )} |^2}$ and the temporal intensity is $I(t) = {|{FT[{{\phi_k}(\omega )} ]} |^2}$.

 figure: Fig. 4.

Fig. 4. (a)-(c) Phase distribution and (d)-(f) Modulus distribution of Γ of a partially coherent pulsed beam with topological charge q = + 1 for different k parameter.

Download Full Size | PDF

Figure 5 shows the normalized spectrum of the source and normalized temporal intensity corresponding to a partially coherent pulsed beam for different values of k. S0(ω) and I0(t) indicate the maxima of S(ω) and I(t), respectively. As can be seen from Fig. 5, different values of k lead to different original spectrum distributions (Fig. 5(a)-(c)), and different temporal intensity distributions (Fig. 5(d)-(f)). From Fig. 5(d)-(f), there is a zero intensity value at t = 0 for k = 1, while there are two zero intensity values at t=±2.7fs for k = 2, and three zero values at t = 0, t=±4.7fs for k = 3. The positions where the temporal intensities are zero correspond to the positions where the STDCs are cut off in the Fig. 4(a)-(c), which indicates a spatiotemporal coupling.

 figure: Fig. 5.

Fig. 5. (a)-(c) Normalized spectrum of the source and (d)-(f) normalized temporal intensity corresponding to partially coherent pulsed beams with different k.

Download Full Size | PDF

Figure 6 shows the phase and modulus distributions of Γ of a partially coherent pulsed beam for different topological charge q. As can be seen in Fig. 6, the phase of STCV rotates along a closed counterclockwise path and increases by 2π for q=+1. While the phase of STCV increases by 4π and 6π for q=+2 and q=+3, respectively. However, the phase of STCV rotates along a closed clockwise path and increases 2π, 4π, 6π for q = -1, q = -2, q = -3, respectively. The topological charge q has no noticeable influence on the STDCs. From Eqs. (6)(g)-(i), the size of vortex core of STCV increases with increasing |q|.

 figure: Fig. 6.

Fig. 6. (a)-(f) Phase distribution and (g)-(i) modulus distribution of Γ of a partially coherent pulsed beam for different topological charge.

Download Full Size | PDF

Figure 7 shows the phase distribution of Γ, level curves of Re Γ′ = 0 (solid line) and Im Γ′ = 0 (dashed line) of a partially coherent pulsed beam with topological charge q = + 1 for different mode weights. Figure 7(a)-(f) and Fig. 6(a) suggest that the STDC gets farther from the spatiotemporal center with increasing mode weight ratio λ1/λ2, and disappears when the mode weight ratio λ1/λ2 becomes large enough. These results imply that the STDC depends on the effect of mode 2, which degrades with increase of λ1/λ2. However, if we increase the effect of mode 2, i.e., decrease λ1/λ2, shown in Fig. 7(g)-(l), the STDCs always exist and are near the STCV. The two coincide in x2 coordinate when the mode weight λ1/λ2 is small enough. In addition, the value of λ1/λ2 is closely related with the coherence state of the partially coherent pulsed beam. Namely, the pulsed beam becomes fully coherent when either λ1/λ2 = ∞ or 0, otherwise, it acquires a partially coherent state. In general, the mode weight ratio λ1/λ2 implies the influence of coherence state on the STDC and the STCV, including the spatial and temporal coherence counterparts. These qualitative results have a good agreement with our previous research [37].

 figure: Fig. 7.

Fig. 7. (a)-(c), (g)-(i) Phase distribution of Γ, (d)-(f), (j)-(l), level curves of Re Γ′ = 0 (solid line) and Im Γ′ = 0 (dashed line) of a partially coherent pulsed beam with topological charge q = + 1 for different mode weights ratio.

Download Full Size | PDF

The above results are confined to the same topological charge q for different modes. Figure 8 shows the phase distribution of Γ of a partially coherent pulsed beam for different topological charges q000 and q100 corresponding to mode 1 and mode 2, respectively. The calculation parameters are (a) q000 = 0, q100 = 1; (b) q000 = 1, q100 = 2; (c) q000 = 2, q100 = 3; (d) q000 = 3, q100 = 4; (e) q000 = 0, q100 = 2; (f) q000 = 1, q100 = 3; (g) q000 = 2, q100 = 4; (h) q000 = 3, q100 = 5. Other parameters are the same as Fig. 1. As can be seen form Fig. 8(a), there are no the STCV and the STDC compared with Fig. 1(a). For larger q value, there is one STCV in the spatiotemporal center and the phase of STCV increases by 2π, 4π and 6π for Fig. 8(b), Fig. 8(c) and Fig. 8(d), respectively. From Fig. 8(e), there are two STCVs near the spatiotemporal center and the STDC appears at the line x2 = 0. For larger q value, i.e., Fig. 8(f), one more STCV with topological charge +1 emerges at the spatiotemporal center. While, with the increase of q, there are one STCV with topological charge +2 and one STCV with topological charge +3 at the spatiotemporal center for Fig. 8(g) and Fig. 8(h), respectively. These results can be explained as follows. For example, when q000 = 2 and q100 = 4, Eq. (16) can be re-expressed as

$$\scalebox{0.9}{$\begin{aligned} \Gamma ^{\prime}({x_1},{x_2},{t_1},{t_2},{z_1},{z_2}) &= {\left[ {\frac{{{t_1}}}{{{t_s}}}\textrm{ - }i \cdot \frac{{x_1^{}}}{{x_s^{}}}} \right]^2}{\left[ {\frac{{{t_2}}}{{{t_s}}} + i \cdot \frac{{x_2^{}}}{{x_s^{}}}} \right]^2}({{\lambda_1}{{\{{FT[{\phi_{000}}({x_1},{\omega_1},{z_1};0)]} \}}^{\ast }}FT[{\phi_{000}}({x_2},{\omega_2},{z_2};0)]} \\ &\left. { + {\lambda_2}{{\{{FT[{\phi_{100}}({x_1},{\omega_1},{z_1};{z_0})]} \}}^{\ast }}FT[{\phi_{100}}({x_2},{\omega_2},{z_2};{z_0})]{{\left[ {\frac{{{t_1}}}{{{t_s}}}\textrm{ - }i \cdot \frac{{x_1^{}}}{{x_s^{}}}} \right]}^2}{{\left[ {\frac{{{t_2}}}{{{t_s}}} + i \cdot \frac{{x_2^{}}}{{x_s^{}}}} \right]}^2}} \right). \end{aligned}$}$$

 figure: Fig. 8.

Fig. 8. Phase distribution of Γ of a partially coherent pulsed beam for different topological charges q000 and q100 corresponding to mode 1 and mode 2, respectively. (a) q000 = 0, q100 = 1; (b) q000 = 1, q100 = 2; (c) q000 = 2, q100 = 3; (d) q000 = 3, q100 = 4; (e) q000 = 0, q100 = 2; (f) q000 = 1, q100 = 3; (g) q000 = 2, q100 = 4; (h) q000 = 3, q100 = 5.

Download Full Size | PDF

As can be seen from Eq. (21) that there one STCV outside of the parentheses at the spatiotemporal center with topological charge +2. Inside the parentheses, there are the superposition of mode 1 and mode 2. The mode 2 has another STCV with topological charge +2. Equation (2) can be also regarded as a modulation of mode 1 and mode 2 on the STCV outside of the parentheses.

Last, an illustrating schematic of the experimental setup for generating STCVs are presented in Fig. 9. A pulsed beam, may be dispersed along the horizontal direction by a grating, and collimated by a cylindrical lens. A spatial light modulator can be employed, loaded with the desired holograms to shape the dispersed pulsed beam. The produced beam can then be converged by the second cylindrical lens and a grating, which then produces a STCV. Further experimental details belong to Ref. [12]. The key operational difference of our method is in the design of the holograms. Considering that STCVs are produced via incoherent superposition of two (or more) modes, here we first introduce the random numbers into both modes and sum them. By refreshing the random numbers, we realize the incoherent superposition of the modes. Considering that the commercial SLMs are normally phase-type and our approach involves combining the STCVs with two or more modes [complex functions, see Eqs. (15) or (16)], we prefer the protocol for creating holograms described by Frobes et al [6062]. It can encode the complex amplitude function into the phase-type holograms [refer to Refs. [6062] for more details].

 figure: Fig. 9.

Fig. 9. Schematic diagram of experimental generation of the STCVs. The distance between all adjacent optical devices is the focal length of the cylindrical lens in the horizontal direction.

Download Full Size | PDF

4. Conclusion

With the help of the coherent-mode representation and Fourier transforms, we have introduced a straightforward, both theoretically and practically, method for generating STCVs and STDCs. The method involves a partially coherent pulsed beam represented by the incoherent superposition of a Gaussian pulsed beam and a Hermite-Gaussian pulsed beam with different waist positions. The detailed numerical calculations are shown to address the dependence of waist distance z0 of two modes, reference position (x1, t1), beam order m, distribution of original spectrum related to the k parameter, topological charge q and mode weights ratio λ1/λ2 on the STCVs and spatiotemporal dislocation. It is revealed that, for the same topological charge q corresponding to the mode 1 and mode 2, there exist the STCVs and the STDCs in the space-time plane. The phase distribution and the core size of STCVs vary with increasing topological charge |q|. The spatiotemporal dislocation lines vanish when |z0| or λ1/λ2 are large enough. There are m STDCs corresponding to the beam order m of the Hermite-Gaussian mode. The original spectrum distribution plays a crucial role on the STDC which is cut into k + 1 segments according to different k values related to the original spectrum. When the reference position is far from the spatiotemporal center or λ1/λ2 decreases within ratio 2, the STDCs always exist and move towards the STCV, and the two finally coincide when the reference position is far enough or λ1/λ2 is small enough. On the other hand, for the different topological charge q values, corresponding to the mode 1 and mode 2, the STCVs and the STDCs simultaneously exists in the space-time plane only for a larger topological charge difference (q100-q000 ≥ 2). In addition, a simple scheme of the experimental setup for generating the STCVs and STDCs is proposed. The results obtained in this paper may have potential applications in the fields of light-matter interaction, spatiotemporal spin-orbit angular momentum coupling and STCV-based optical trapping and optical manipulation.

Funding

National Natural Science Foundation of China (12174171, 12004220); Central Plains Talents Program of Henan (ZYYCYU202012144); Natural Science Foundation of Henan Province (232102210169); Excellent Overseas Visiting Scholar Program of Henan (026); Agence Nationale de la Recherche (ANR-19-QUAN-0001-04); University of Miami via the Cooper Fellowship.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are available from the corresponding authors upon reasonable request.

References

1. L. Allen, M. W. Beijersbergen, R. J. C. Spreeuw, et al., “Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes,” Phys. Rev. A 45(11), 8185–8189 (1992). [CrossRef]  

2. J. Wang, J.-Y. Yang, I. M. Fazal, et al., “Terabit free-space data transmission employing orbital angular momentum multiplexing,” Nat. Photonics 6(7), 488–496 (2012). [CrossRef]  

3. M. Padgett and R. Bowman, “Tweezers with a twist,” Nat. Photonics 5(6), 343–348 (2011). [CrossRef]  

4. M. P. J. Lavery, F. C. Speirits, S. M. Barnett, et al., “Detection of a spinning object using light's orbital angular momentum,” Science 341(6145), 537–540 (2013). [CrossRef]  

5. A. Vaziri, J. W. Pan, T. Jennewein, et al., “Concentration of higher dimensional entanglement: qutrits of photon orbital angular momentum,” Phys. Rev. Lett. 91(22), 227902 (2003). [CrossRef]  

6. L. Chen, J. Lei, and J. Romero, “Quantum digital spiral imaging,” Light: Sci. Appl. 3(3), e153 (2014). [CrossRef]  

7. A. Sukhorukov and V. Yangirova, “Spatio-temporal vortices: properties, generation and recording,” Proc. SPIE 5949, 594906 (2005). [CrossRef]  

8. K. Y. Bliokh and F. Nori, “Spatiotemporal vortex beams and angular momentum,” Phys. Rev. A 86(3), 033824 (2012). [CrossRef]  

9. K. Y. Bliokh and F. Nori, “Transverse and longitudinal angular momenta of light,” Phys. Rep. 592, 1–38 (2015). [CrossRef]  

10. N. Jhajj, I. Larkin, E. W. Rosenthal, et al., “Spatiotemporal optical vortices,” Phys. Rev. X 6(3), 031037 (2016). [CrossRef]  

11. S. W. Hancock, S. Zahedpour, A. Goffin, et al., “Free-space propagation of spatiotemporal optical vortices,” Optica 6(12), 1547–1553 (2019). [CrossRef]  

12. A. Chong, C. Wan, J. Chen, et al., “Generation of spatiotemporal optical vortices with controllable transverse orbital angular momentum,” Nat. Photonics 14(6), 350–354 (2020). [CrossRef]  

13. K. Y. Bliokh, “Spatiotemporal vortex pulses: angular momenta and spin-orbit interaction,” Phys. Rev. Lett. 126(24), 243601 (2021). [CrossRef]  

14. S. Huang, P. Wang, X. Shen, et al., “Properties of the generation and propagation of spatiotemporal optical vortices,” Opt. Express 29(17), 26995–27003 (2021). [CrossRef]  

15. S. Huang, P. Wang, X. Shen, et al., “Diffraction properties of light with transverse orbital angular momentum,” Optica 9(5), 469–472 (2022). [CrossRef]  

16. C. Wan, Q. Cao, J. Chen, et al., “Toroidal vortices of light,” Nat. Photonics 16(7), 519–522 (2022). [CrossRef]  

17. W. Chen, W. Zhang, Y. Liu, et al., “Time diffraction-free transverse orbital angular momentum beams,” Nat. Commun. 13(1), 4021 (2022). [CrossRef]  

18. Q. Cao, Z. Chen, C. Zhang, et al., “Propagation of transverse photonic orbital angular momentum through few-mode fiber,” Adv. Photonics 5(03), 036002 (2023). [CrossRef]  

19. H. Rubinsztein-Dunlop, A. Forbes, M. V. Berry, et al., “Roadmap on structured light,” J. Opt. 19(1), 013001 (2017). [CrossRef]  

20. J. Wang and Y. Liang, “Generation and detection of structured light: a review,” Front. Phys. 9, 688284 (2021). [CrossRef]  

21. J. Chen, C. Wan, and Q. Zhan, “Engineering photonic angular momentum with structured light: a review,” Adv. Photonics 3(06), 064001 (2021). [CrossRef]  

22. X. Liu, Y. Shen, L. Liu, et al., “Experimental demonstration of vortex phase-induced reduction in scintillation of a partially coherent beam,” Opt. Lett. 38(24), 5323–5326 (2013). [CrossRef]  

23. Y. Yang, Y. Ren, M. Chen, et al., “Optical trapping with structured light: a review,” Adv. Photonics 3(03), 034001 (2021). [CrossRef]  

24. X. Liu, X. Peng, L. Liu, et al., “Self-reconstruction of the degree of coherence of a partially coherent vortex beam obstructed by an opaque obstacle,” Appl. Phys. Lett. 110(18), 181104 (2017). [CrossRef]  

25. G. Gbur and R. K. Tyson, “Vortex beam propagation through atmospheric turbulence and topological charge conservation,” J. Opt. Soc. Am. A 25(1), 225–230 (2008). [CrossRef]  

26. G. V. Bogatyryova, C. V. Fel’de, P. V. Polyanskii, et al., “Partially coherent vortex beams with a separable phase,” Opt. Lett. 28(11), 878–880 (2003). [CrossRef]  

27. F. Gori, M. Santarsiero, R. Borghi, et al., “Partially coherent sources with helicoidal modes,” J. Mod. Opt. 45(3), 539–554 (1998). [CrossRef]  

28. G. Gbur and T. D. Visser, “Coherence vortices in partially coherent beams,” Opt. Commun. 222(1-6), 117–125 (2003). [CrossRef]  

29. D. M. Palacios, I. D. Maleev, A. S. Marathay, et al., “Spatial correlation singularity of a vortex field,” Phys. Rev. Lett. 92(14), 143905 (2004). [CrossRef]  

30. R. K. Singh, A. M. Sharma, and P. Senthilkumaran, “Vortex array embedded in a partially coherent beam,” Opt. Lett. 40(12), 2751–2754 (2015). [CrossRef]  

31. Y. Zhang, O. Korotkova, Y. Cai, et al., “Correlation-induced orbital angular momentum changes,” Phys. Rev. A 102(6), 063513 (2020). [CrossRef]  

32. J. Wang, S. Yang, M. Guo, et al., “Change in phase singularities of a partially coherent Gaussian vortex beam propagating in a GRIN fiber,” Opt. Express 28(4), 4661–4673 (2020). [CrossRef]  

33. X. Liu, J. Zeng, and Y. Cai, “Review on vortex beams with low spatial coherence,” Adv. Phys.: X 4(1), 1626766 (2019). [CrossRef]  

34. J. Zeng, R. Lin, X. Liu, et al., “Review on partially coherent vortex beams,” Front. Optoelectron. 12(3), 229–248 (2019). [CrossRef]  

35. O. Korotkova and G. Gbur, “Unified matrix representation for spin and orbital angular momentum in partially coherent beams,” Phys. Rev. A 103(2), 023529 (2021). [CrossRef]  

36. Z. Yang, H. Wang, Y. Chen, et al., “Measurement of the coherence-orbital angular momentum matrix of a partially coherent beam,” Opt. Lett. 47(17), 4467–4470 (2022). [CrossRef]  

37. C. Ding, D. Horoshko, O. Korotkova, et al., “Source coherence-induced control of spatiotemporal coherency vortices,” Opt. Express 30(11), 19871–19888 (2022). [CrossRef]  

38. R. Dutta, J. Turunen, and A. T. Friberg, “Michelson's interferometer and the temporal coherence of pulse trains,” Opt. Lett. 40(2), 166–169 (2015). [CrossRef]  

39. A. Abbas, C. Xu, and L. Wang, “Spatiotemporal ghost imaging and interference,” Phys. Rev. A 101(4), 043805 (2020). [CrossRef]  

40. A. Abbas and L. Wang, “Hanbury Brown and Twiss effect in spatiotemporal domain,” Opt. Express 28(21), 32077–32086 (2020). [CrossRef]  

41. H. Pesonen, P. Li, T. Setälä, et al., “Temporal coherence and polarization modulation of pulse trains by resonance gratings,” J. Opt. Soc. Am. A 37(1), 27–38 (2020). [CrossRef]  

42. H. Wang, X. L. Ji, Y. Deng, et al., “Theory of the quasi-steady-state self-focusing of partially coherent light pulses in nonlinear media,” Opt. Lett. 45(3), 710–713 (2020). [CrossRef]  

43. A. Le Marec, O. Guilbaud, O. Larroche, et al., “Evidence of partial temporal coherence effects in the linear autocorrelation of extreme ultraviolet laser pulses,” Opt. Lett. 41(14), 3387–3390 (2016). [CrossRef]  

44. C. Bourassin-Bouchet and M. E. Couprie, “Partially coherent ultrafast spectrography,” Nat. Commun. 6(1), 6465 (2015). [CrossRef]  

45. H. Lajunen and T. Saastamoinen, “Non-uniformly correlated partially coherent pulses,” Opt. Express 21(1), 190–195 (2013). [CrossRef]  

46. M. Koivurova, L. Ahad, G. Geloni, et al., “Interferometry and coherence of nonstationary light,” Opt. Lett. 44(3), 522–525 (2019). [CrossRef]  

47. M. Koivurova, C. Ding, J. Turunen, et al., “Partially coherent isodiffracting pulsed beams,” Phys. Rev. A 97(2), 023825 (2018). [CrossRef]  

48. C. Ding, O. Korotkova, Y. Zhang, et al., “Cosine-Gaussian correlated Schell-model pulsed beams,” Opt. Express 22(1), 931–942 (2014). [CrossRef]  

49. C. Ding, M. Koivurova, J. Turunen, et al., “Temporal self-splitting of optical pulses,” Phys. Rev. A 97(5), 053838 (2018). [CrossRef]  

50. C. Ding, O. Korotkova, D. Zhao, et al., “Propagation of temporal coherence gratings in dispersive medium with a chirper,” Opt. Express 28(5), 7463–7474 (2020). [CrossRef]  

51. M. W. Hyde, “Twisted space-frequency and space-time partially coherent beams,” Sci. Rep. 10(1), 12443 (2020). [CrossRef]  

52. A. Mirando, Y. Zang, Q. W. Zhan, et al., “Generation of spatiotemporal optical vortices with partial temporal coherence,” Opt. Express 29(19), 30426–30435 (2021). [CrossRef]  

53. G. Indebetouw, “Optical Vortices and Their Propagation,” J. Mod. Opt. 40(1), 73–87 (1993). [CrossRef]  

54. H. Yan and B. Lü, “Dynamic evolution of an edge dislocation through aligned and misaligned paraxial optical ABCD systems,” J. Opt. Soc. Am. A 26(4), 985–992 (2009). [CrossRef]  

55. G. Gbur, T. D. Visser, and E. Wolf, “‘Hidden’ singularities in partially coherent wavefields,” J. Opt. A: Pure Appl. Opt. 6(5), S239–S242 (2004). [CrossRef]  

56. P. Gao, L. Bai, Z. Wu, et al., “Evolution of linear edge dislocation in atmospheric turbulence and free space,” J. Mod. Opt. 66(1), 17–25 (2019). [CrossRef]  

57. H. Lajunen, P. Vahimaa, and J. Tervo, “Theory of spatially and spectrally partially coherent pulses,” J. Opt. Soc. Am. A 22(8), 1536–1545 (2005). [CrossRef]  

58. L. Sereda, M. Bertolotti, and A. Ferrari, “Coherence properties of nonstationary light wave fields,” J. Opt. Soc. Am. A 15(3), 695–705 (1998). [CrossRef]  

59. I. Freund and N. Shvartsman, “Wave-field phase singularities:the sign principle,” Phys. Rev. A 50(6), 5164–5172 (1994). [CrossRef]  

60. C. Rosales-Guzmán and A. Forbes, How to Shape Light with Spatial Light Modulators (SPIE Press, 2017, Chap.6, pp. 37–46).

61. X. Liu, Y. E. Monfared, R. Pan, et al., “Experimental realization of scalar and vector perfect Laguerre–Gaussian beams,” Appl. Phys. Lett. 119(2), 021105 (2021). [CrossRef]  

62. X. Wang, J. Tang, Y. Wang, et al., “Complex and phase screen methods for studying arbitrary genuine Schell-model partially coherent pulses in nonlinear media,” Opt. Express 30(14), 24222–24231 (2022). [CrossRef]  

Data availability

Data underlying the results presented in this paper are available from the corresponding authors upon reasonable request.

Cited By

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

Alert me when this article is cited.


Figures (9)

Fig. 1.
Fig. 1. (a)-(d) Phase distribution of Γ, (e)-(h) level curves of Re Γ′ = 0 (solid line) and Im Γ′ = 0 (dashed line), (i)-(l) modulus distribution of Γ and (m)-(p) spatiotemporal intensity I′ of a partially coherent pulsed beam with topological charge q = + 1 for different waist positions of two modes.
Fig. 2.
Fig. 2. (a)-(c) Phase distribution and (d)-(f) Modulus distribution of Γ of a partially coherent pulsed beam with topological charge q = + 1 for different reference positions x1, t1.
Fig. 3.
Fig. 3. (a)-(f) Phase distribution and (g)-(l) modulus distribution of Γ of a partially coherent pulsed beam with topological charge q = + 1 for different beam order of the mode 2.
Fig. 4.
Fig. 4. (a)-(c) Phase distribution and (d)-(f) Modulus distribution of Γ of a partially coherent pulsed beam with topological charge q = + 1 for different k parameter.
Fig. 5.
Fig. 5. (a)-(c) Normalized spectrum of the source and (d)-(f) normalized temporal intensity corresponding to partially coherent pulsed beams with different k.
Fig. 6.
Fig. 6. (a)-(f) Phase distribution and (g)-(i) modulus distribution of Γ of a partially coherent pulsed beam for different topological charge.
Fig. 7.
Fig. 7. (a)-(c), (g)-(i) Phase distribution of Γ, (d)-(f), (j)-(l), level curves of Re Γ′ = 0 (solid line) and Im Γ′ = 0 (dashed line) of a partially coherent pulsed beam with topological charge q = + 1 for different mode weights ratio.
Fig. 8.
Fig. 8. Phase distribution of Γ of a partially coherent pulsed beam for different topological charges q000 and q100 corresponding to mode 1 and mode 2, respectively. (a) q000 = 0, q100 = 1; (b) q000 = 1, q100 = 2; (c) q000 = 2, q100 = 3; (d) q000 = 3, q100 = 4; (e) q000 = 0, q100 = 2; (f) q000 = 1, q100 = 3; (g) q000 = 2, q100 = 4; (h) q000 = 3, q100 = 5.
Fig. 9.
Fig. 9. Schematic diagram of experimental generation of the STCVs. The distance between all adjacent optical devices is the focal length of the cylindrical lens in the horizontal direction.

Equations (21)

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

$$W({{\boldsymbol r}_1},{{\boldsymbol r}_2},{\omega _1},{\omega _2},{z_1},{z_2}) = \sum\limits_{m,n,k}^{} {{\lambda _{mnk}}\phi _{mnk}^\ast ({{\boldsymbol r}_1},{\omega _1},{z_1};{z_0}){\phi _{mnk}}({{\boldsymbol r}_2},{\omega _2},{z_2};{z_0})}, $$
$$\begin{aligned} {\phi _{mnk}}({\boldsymbol r},\omega ,z;{z_0}) &= {\phi _k}(\omega ){\left( {\frac{2}{\pi }} \right)^{1/2}}\frac{{\exp \{{i[{{{\omega (z\textrm{ - }{z_0})} / c} - ({m + n + 1} )\theta ({z\textrm{ - }{z_0},\omega } )} ]} \}}}{{w(z\textrm{ - }{z_0},\omega )\sqrt {{2^{m + n}}m!n!} }}\\ & \times {H_m}\left[ {\frac{{x\sqrt 2 }}{{w(z\textrm{ - }{z_0},\omega )}}} \right]{H_n}\left[ {\frac{{y\sqrt 2 }}{{w(z\textrm{ - }{z_0},\omega )}}} \right]\exp \left[ {i\frac{{\omega ({x^2} + {y^2})}}{{2cR(z\textrm{ - }{z_0},\omega )}} - \frac{{{x^2} + {y^2}}}{{{w^2}(z\textrm{ - }{z_0},\omega )}}} \right] \end{aligned}, $$
$$w({z,\omega } )= {w_0}\sqrt {1 + {{\left( {\frac{{2cz}}{{{w_0}{w_c}\omega }}} \right)}^2}}, $$
$$R({z,\omega } )= z\left[ {1 + {{\left( {\frac{{{w_0}{w_c}\omega }}{{2cz}}} \right)}^2}} \right], $$
$$\tan \theta ({z,\omega } )= \frac{{{w^2}(z,\omega )\omega }}{{2cR(z,\omega )}} = \frac{{2cz}}{{w_c^2\omega }}, $$
$$\frac{1}{{w_c^2}} = \frac{1}{{w_0^2}} + \frac{1}{{\sigma _0^2}}, $$
$${\phi _k}(\omega ) = \frac{1}{{\sqrt {{2^k}k!} }}{\left( {\frac{{2d}}{\pi }} \right)^{{1 / 4}}}{H_k}\left[ {(\omega - {\omega_0})\sqrt {2d} } \right]\exp [{ - d{{(\omega - {\omega_0})}^2}} ], $$
$$d = {({{a^2} + 2ab} )^{{1 / 2}}}, $$
$$a = {1 / {2\Omega _0^2}},b = {1 / {2\Omega _c^2}}, $$
$$\Omega _0^2 = {1 / {T_0^2}} + {2 / {T_c^2}},{\Omega _c} = {{{{T_c}{\Omega _0}} / T}_0}, $$
$$\Gamma ({{\boldsymbol r}_1},{{\boldsymbol r}_2},{t_1},{t_2},{z_1},{z_2}) = \sum\limits_{m,n,k}^{} {{\lambda _{mnk}}\varphi _{mnk}^\ast ({{\boldsymbol r}_1},{t_1},{z_1};{z_0}){\varphi _{mnk}}({{\boldsymbol r}_2},{t_2},{z_2};{z_0})}$$
$${\varphi _{mnk}}({\boldsymbol r},t,z;{z_0}) = FT[{\phi _{nnk}}({\boldsymbol r},\omega ,z;{z_0})] = \int\limits_{ - \infty }^\infty {{\phi _{nnk}}({\boldsymbol r},\omega ,z;{z_0})} \exp [ - i\omega t]d\omega. $$
$${\left[ {\frac{{{t_{}}}}{{{t_s}}} + i\textrm{sgn}(q) \cdot \frac{{x_{}^{}}}{{x_s^{}}}} \right]^{|q |}}$$
$${\varphi ^{\prime}_{mnk}}({\boldsymbol r},t,z;{z_0}) = {\varphi _{mnk}}({\boldsymbol r},t,z;{z_0}){\left[ {\frac{{{t_{}}}}{{{t_s}}} + i\textrm{sgn}({q_{mnk}}) \cdot \frac{{x_{}^{}}}{{x_s^{}}}} \right]^{|{{q_{mnk}}} |}}. $$
$$\scalebox{0.86}{$\begin{aligned} &\Gamma ^{\prime}({{\boldsymbol r}_1},{{\boldsymbol r}_2},{t_1},{t_2},{z_1},{z_2}) = \sum\limits_{m,n,k}^{} {{\lambda _{mnk}}\varphi ^{\prime\ast}_{mnk} ({{\boldsymbol r}_1},{t_1},{z_1};{z_0}){{\varphi ^{\prime}}_{mnk}}({{\boldsymbol r}_2},{t_2},{z_2};{z_0})} \\ &= \sum\limits_{m,n,k}^{} {{\lambda _{mnk}}\varphi _{mnk}^\ast ({{\boldsymbol r}_1},{t_1},{z_1};{z_0}){\varphi _{mnk}}({{\boldsymbol r}_2},{t_2},{z_2};{z_0})} {\left[ {\frac{{{t_1}}}{{{t_s}}}\textrm{ - }i\textrm{sgn}({q_{mnk}}) \cdot \frac{{x_1^{}}}{{x_s^{}}}} \right]^{|{{q_{mnk}}} |}}{\left[ {\frac{{{t_2}}}{{{t_s}}} + i\textrm{sgn}({q_{mnk}}) \cdot \frac{{x_2^{}}}{{x_s^{}}}} \right]^{|{{q_{mnk}}} |}}. \end{aligned}$}$$
$$\scalebox{0.86}{$\begin{aligned} &\Gamma ^{\prime}({{\boldsymbol r}_1},{{\boldsymbol r}_2},{t_1},{t_2},{z_1},{z_2})\\ &= {\lambda _1}{\{{FT[{\phi_{000}}({{\boldsymbol r}_1},{\omega_1},{z_1};0)]} \}^{\ast }}FT[{\phi _{000}}({{\boldsymbol r}_2},{\omega _2},{z_2};0)]{\left[ {\frac{{{t_1}}}{{{t_s}}}\textrm{ - }i\textrm{sgn}({q_{000}}) \cdot \frac{{x_1^{}}}{{x_s^{}}}} \right]^{|{{q_{000}}} |}}{\left[ {\frac{{{t_2}}}{{{t_s}}} + i\textrm{sgn}({q_{000}}) \cdot \frac{{x_2^{}}}{{x_s^{}}}} \right]^{|{{q_{000}}} |}}\\ &+ {\lambda _2}{\{{FT[{\phi_{mnk}}({{\boldsymbol r}_1},{\omega_1},{z_1};{z_0})]} \}^{\ast }}FT[{\phi _{mnk}}({{\boldsymbol r}_2},{\omega _2},{z_2};{z_0})]{\left[ {\frac{{{t_1}}}{{{t_s}}}\textrm{ - }i\textrm{sgn}({q_{mnk}}) \cdot \frac{{x_1^{}}}{{x_s^{}}}} \right]^{|{{q_{mnk}}} |}}{\left[ {\frac{{{t_2}}}{{{t_s}}} + i\textrm{sgn}({q_{mnk}}) \cdot \frac{{x_2^{}}}{{x_s^{}}}} \right]^{|{{q_{mnk}}} |}}, \end{aligned}$}$$
$${\phi _{000}}({\boldsymbol r},\omega ,z;0) = {\left( {\frac{{2d}}{\pi }} \right)^{{1 / 4}}}\exp [{ - d{{(\omega - {\omega_0})}^2}} ]{\left( {\frac{2}{\pi }} \right)^{1/2}}\frac{1}{{{w_0}}}\exp \left[ { - \frac{{{x^2}}}{{w_0^2}}} \right], $$
$$\begin{aligned} {\phi _{m0k}}({\boldsymbol r},\omega ,z;{z_0}) &= {\left( {\frac{{2d}}{\pi }} \right)^{{1 / 4}}}\frac{1}{{\sqrt {{2^k}k!} }}{H_k}\left[ {(\omega - {\omega_0})\sqrt {2d} } \right]\exp [{ - d{{(\omega - {\omega_0})}^2}} ]\\ & \times {\left( {\frac{2}{\pi }} \right)^{1/2}}\frac{{\exp \{{i[{{{\omega (z\textrm{ - }{z_0})} / c} - ({m + 1} )\theta ({z\textrm{ - }{z_0},\omega } )} ]} \}}}{{w(z\textrm{ - }{z_0},\omega )\sqrt {{2^m}m!} }}\\ & \times {H_m}\left[ {\frac{{x\sqrt 2 }}{{w(z\textrm{ - }{z_0},\omega )}}} \right]\exp \left[ {i\frac{{\omega {x^2}}}{{2cR(z\textrm{ - }{z_0},\omega )}} - \frac{{{x^2}}}{{{w^2}(z\textrm{ - }{z_0},\omega )}}} \right] \end{aligned}, $$
$${\textrm{Re}} [{\Gamma^{\prime}({x_1},{x_2},{t_1},{t_2},{z_1},{z_2})} ]= 0,$$
$${\mathop{\rm Im}\nolimits} [{\Gamma^{\prime}({x_1},{x_2},{t_1},{t_2},{z_1},{z_2})} ]= 0,$$
$$\scalebox{0.9}{$\begin{aligned} \Gamma ^{\prime}({x_1},{x_2},{t_1},{t_2},{z_1},{z_2}) &= {\left[ {\frac{{{t_1}}}{{{t_s}}}\textrm{ - }i \cdot \frac{{x_1^{}}}{{x_s^{}}}} \right]^2}{\left[ {\frac{{{t_2}}}{{{t_s}}} + i \cdot \frac{{x_2^{}}}{{x_s^{}}}} \right]^2}({{\lambda_1}{{\{{FT[{\phi_{000}}({x_1},{\omega_1},{z_1};0)]} \}}^{\ast }}FT[{\phi_{000}}({x_2},{\omega_2},{z_2};0)]} \\ &\left. { + {\lambda_2}{{\{{FT[{\phi_{100}}({x_1},{\omega_1},{z_1};{z_0})]} \}}^{\ast }}FT[{\phi_{100}}({x_2},{\omega_2},{z_2};{z_0})]{{\left[ {\frac{{{t_1}}}{{{t_s}}}\textrm{ - }i \cdot \frac{{x_1^{}}}{{x_s^{}}}} \right]}^2}{{\left[ {\frac{{{t_2}}}{{{t_s}}} + i \cdot \frac{{x_2^{}}}{{x_s^{}}}} \right]}^2}} \right). \end{aligned}$}$$
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.