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

Theory and generation of heterogeneous 2D arrays of optical vortices by using 2D fork-shaped gratings: topological charge and power sharing management

Open Access Open Access

Abstract

In this work, by providing comprehensive theoretical foundations, we revisit and improve a simple and efficient method that has been used for generation of 2D orthogonal arrays of optical vortices with components having different topological charges (TCs). This method has been implemented by the diffraction of a plane wave from 2D gratings where the gratings’ profiles are determined by iterative computational process. Here, based on the theoretical predictions, specifications of the diffraction gratings can be easily adjusted in a way to generate experimentally a heterogeneous vortex array with the desired power shares among different elements of the array. We use the diffraction of a Gaussian beam from a class of pure phase 2D orthogonal periodic structures having sinusoidal or binary profiles possessing a phase singularity, calling pure phase 2D fork-shaped gratings (FSGs). The transmittance of each of the introduced gratings is obtained by multiplying the transmittance of two pure phase 1D FSGs along x and y directions, having topological defect numbers lx and ly and phase variation amplitudes γx and γy, respectively. By solving the Fresnel integral, we show that the diffraction of a Gaussian beam from a pure phase 2D FSG leads to generation of a 2D array of vortex beams having different TCs and power shares. The power distribution among the generated optical vortices over the different diffraction orders can be adjusted by γx and γy, and it strongly depends on the profile of the grating. Meanwhile the TCs of the generated vortices depend on lx and ly and the corresponding diffraction orders, namely lm,n = −(mlx + nly) presents the TC of (m, n)th diffraction order. We recorded the intensity patterns of the experimentally generated vortex arrays which are fully consistent with the theoretically predicted results. Furthermore, the TCs of the experimentally generated vortices are measured individually by the diffraction of each of them through a pure amplitude quadratic curved-line (parabolic-line) grating. The absolute values and signs of the measured TCs are consistent with the theoretical prediction. The generated configuration of vortices with adjustable TC and power sharing features might find many applications such as non-homogeneous mixing of a solution consisting trapped particles.

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

1. Introduction

In recent three decades, singular optics has grown as an interesting and highly applied field which has been well extended to other fields of physics such as acoustic, electron and neutron physics, etc. A variety of different methods have been introduced for both generation and detection of optical vortices [1]. In some of methods used for the generation of vortex beams, a Gaussian laser mode passes through some optical elements such as spiral phase plates [2,3], forked gratings [49], spiral zone plates [1012], and Q-plates [13,14] in which the vortices, based on the effect of the element on the impinging beam, form at different propagation distances. A considerable number of methods have been also proposed to characterize optical vortices specifically to determine their TCs, including the diffraction from different structured apertures [1519] and gratings [2025], interferometry [26], and moiré deflectometry [27].

On the other hand, generation of optical vortex arrays (vortex lattices) has recently found interesting applications in the manipulation of microoptomechanical pumps, microlithography, orbital angular momentum (OAM) multiplexing, and quantum processing [28]. One of the methods used for the generation of optical vortex arrays is based on the near-field diffraction of a vortex beam from optical gratings [2933]. Moreover, a detailed study on the near- and far-field diffraction from 2D orthogonal multiplicatively separable and nonseparable periodic structures has been presented in Refs. [34,35]. It is shown that the spatial spectrum or far-field diffraction pattern of these structures is a 2D orthogonal lattice of impulses.

In this work we introduce a family of pure phase 2D orthogonal multiplicatively separable periodic structures having sinusoidal or binary phase profiles including a phase singularity. Indeed, the proposed structure is multiplication of two pure phase 1D fork-shaped gratings (FSGs) defined along $x$ and $y$ directions, having TCs $l_x$ and $l_y$ and phase variation amplitudes $\gamma _x$ and $\gamma _y$, respectively. Diffraction of a Gaussian beam by these structures leads to generation of a 2D array of vortex beams with different TCs and power sharings. We show that the power share of any diffraction order equals the absolute square of the corresponding Fourier coefficient which is a two variable function of $\gamma _x$ and $\gamma _y$. Therefore, sharing of the incident power among different diffraction orders can be managed by changing the values of $\gamma _x$ and $\gamma _y$. It is also shown that the TC of $(m,n)$th diffraction order is $l_{m,n}=-\left ( ml_x+nl_y \right )$. Then, the value of $l$ can be manipulated by adjusting the values of $l_x$ and $l_y$.

It is worth noting that in a very close work [36] generation of a 2D orthogonal array of optical vortices using a pure phase 2D FSG having a modified blazed profile has been proposed. This modified blazed profile is defined by an anti-blazing function that is obtained by a numerical method, Gerchberg-Saxton algorithm [37]. Using different anti-blazing functions leads to different power distributions in the diffraction orders. In another work the produced array of optical vortices have been used for free space information transfer by the same research group [38]. It is also worth mentioning that the optical vortex arrays in both Refs. [36,38] were generated under plane wave illumination of the modified blazed based pure phase 2D FSGs. While in the current work we present a full mathematical insight to the power sharing and TC management by explicit solving the Fresnel integral and using the 2D Fourier analyses of the transmittance. This kind of facing to the issue enables us to continuously calculate the diffraction behaviour in terms of $\gamma _x$, $\gamma _y$, $l_x$, and $l_y$. The main advantage of our approach is that we can determine the mentioned parameters of the grating to obtain a desired power sharing optimally. For instance, by choosing a binary profile for the grating and $\gamma _x+\gamma _y=\pi$ all of the second (even) orders completely vanish.

2. Pure phase 1D FSGs with sinusoidal and binary profiles

By considering $(x,y)$ as Cartesian coordinates and $(r,\theta )$ as the corresponding polar coordinates, generalized transmission function of a 1D fork-shaped structure can be written as [39,40]:

$$T_x(x,y) = \sum_{m ={-} \infty }^{ + \infty } {{t_m}\exp \left[ {im\left( {\frac{{2\pi }}{{{\Lambda _x}}}x - {l_x}\theta } \right)} \right]},$$
where $\Lambda _x$ is the period of the grating at far distances from the dislocation point located at the center of coordinate system, and $l_x$ is the topological defect number or defect number of the grating. The values of $t_m$ coefficients can be specified considering the grating’s profile. The transmission functions of pure phase 1D FSGs with sinusoidal and binary profiles have the following transmittances:
$$T_x(x,y) = \exp \left[ {i{\gamma_x}\cos \left( {\frac{{2\pi }}{{{\Lambda _x}}}x - {l_x}\theta } \right)} \right],$$
$$T_x(x,y) = \exp \Bigg\{ {i{\gamma_x}\text{sign}\left[ {\cos \left( {\frac{{2\pi }}{{{\Lambda_x}}}x - {l_x}\theta } \right)} \right]} \Bigg\},$$
respectively, where $\gamma _x$ is the amplitude of the phase variation and “sign” indicates the signum function. The first column of Fig. 1 illustrates the phase profiles of these gratings having $l_x=\,2$ and $\Lambda _x=\,0.1$ mm with sinusoidal (first row) and binary (second row) profiles. The Fourier expansion of these transmission functions can be expressed in the unified form of Eq. (1) in which $t_m={t_m}(\gamma _x )$ is a function of $\gamma _x$. For the sinusoidal profile ${t_m}(\gamma _x )=i^mJ_m(\gamma _x )$ and for the binary profile [41]
$${t_m}(\gamma_x )=\begin{cases}\cos\left(\gamma_x \right),\qquad\qquad\ m=0 , \\ i\sin \left( {{\gamma _x}} \right)\text{sinc}\left( {\frac{m}{2}} \right) ,\quad m\neq 0 ,\end{cases}$$
in which $\text {sinc}(x)=\frac {{\sin (\pi x)}}{{\pi x}}$ if $x\neq 0$ and $\text {sinc}(x)=1$ if $x= 0$. As $\text {sinc}( m/2)=0$ for even values of $m$ and $\text {sinc}( m/2)=\frac {2}{m\pi } i^{(m-1)}$ for odd values of $m$ (see [41]), Eq. (3) can be rewritten as
$${t_m}(\gamma_x )=\begin{cases}\cos\left(\gamma_x \right),\quad\quad m=0 , \\ \frac{2i^m}{m\pi}\sin\left(\gamma_x \right) , \quad m\neq 0 \quad \text{and it is odd }, \\ 0, \quad\qquad\qquad m\neq 0 \quad \text{and it is even }. \end{cases}$$

For a pure phase periodic structure the absolute square of the corresponding Fourier coefficient, ${ \left | {{t_m}\left ( {{\gamma _x}} \right )} \right |}^2$, equals the power share of the $m$ the diffraction order from the power of the incident/transmitted light beam, see Appendix A. Using this fact the power shares of the diffraction orders of a pure phase 1D FSG with a sinusoidal profile are illustrated in terms of $\gamma _x$ in the first row of Fig. 2. Considering ${ \left | {{t_m}\left ( {{\gamma _x}} \right )} \right |}^2=J_m^2\left ( {{\gamma _x}} \right )$ we can find some values of $\gamma _x$ for which power sharing gets interesting properties, see Fig. 3 of Ref. [42]. For instance, considering $\gamma _x=2.4048$, as the first root of the function $J_0(\gamma _x)$, zero (central) diffraction order vanishes as is shown in the second row of Fig. 2. Considering $\gamma _x=1.4347$, as the first intersection of the functions $J_0(\gamma _x)$ and $J_1(\gamma _x)$, the power shares of the zero and first diffraction orders are equal as is seen in the third row of Fig. 2. Furthermore, by setting $\gamma _x=2.6298$, as the first intersection of the functions $J_1(\gamma _x)$ and $J_2(\gamma _x)$, the power shares of the first and second diffraction orders are equal as is shown in the fourth row of Fig. 2. By setting $\gamma _x=1.8412$ the function $J_1(\gamma _x)$ reaches to its maximum, then the power share of the first diffraction order is maximized, see the fifth row of Fig. 2. Moreover, the power sharing among diffracted beams is calculated in the first row of Visualization 1 for a pure phase 1D FSG with a sinusoidal profile by changing $\gamma _x$.

 figure: Fig. 1.

Fig. 1. Phase profiles of $T_x(x,y)$ with $l_x=\,2$ (first column), $T_y(x,y)$ with $l_y\,=\,3$ (second column), $T(x,y)$ with $l_x=\,2 , l_y\,=\,3$ (third column), and $T(x,y)$ with $l_x=l_y=0$ (fourth column) having $\gamma _x=\gamma _y=\frac {\gamma }{2}$ and $\Lambda _x=\Lambda _y=\,0.1$ mm. The first and second rows are for the sinusoidal and binary transmittances, respectively.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. First row: power shares of the zero to second diffraction orders of a pure phase 1D FSG with a sinusoidal profile in terms of $\gamma _x$. Second to fifth rows: power shares of the zero to third diffraction orders for different values of $\gamma _x$. See the first row of Visualization 1.

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. First row: power shares of the zero, first, and third diffraction orders of a pure phase 1D FSG with a binary profile in terms of $\gamma _x$. Second to third rows: power shares of the zero to third diffraction orders for different values of $\gamma _x$. See the second row of Visualization 1.

Download Full Size | PDF

By using Eq. (4), the power shares of the diffraction orders for a pure phase 1D FSG with a binary profile are calculated in terms of $\gamma _x$ and the results are depicted in the first row of Fig. 3. As is expected we only have the odd diffraction orders, except the zero order. Furthermore, we have $\left | {{t_m}\left ( {\pi - {\gamma _x}} \right )} \right | = \left | {{t_m}\left ( {{\gamma _x}} \right )} \right |$ that leads to symmetry of the plots around $\gamma _x=\frac {\pi }{2}$. For $\gamma _x=\frac {\pi }{2}$ the zero (central) diffraction order vanishes and the first diffraction order is maximum as is shown in the second row of Fig. 3. By setting $\left | t_0\left ( {{\gamma _x}} \right )\right |=\left | t_{\pm 1}\left ( {{\gamma _x}} \right )\right |$ in Eq. (4) we get ${\gamma _x} = {\tan ^{ - 1}}(\pi /2) \approx 1.004$ and ${\gamma _x} =\pi - {\tan ^{ - 1}}(\pi /2) \approx 2.138$, the intersections of blue solid and dashed red curves in the first row of Fig. 3. For these two values of $\gamma _x$ the power shares of the zero (central) and first diffraction orders are equal, see third row of Fig. 3. Moreover, for a pure phase 1D FSG with a binary profile, the power sharing among diffracted beams by changing $\gamma _x$ is illustrated in the second row of Visualization 1.

Similarly $T_y(x,y)$ can be defined by replacing $x$ by $y$ in the Eqs. (1)–(4). The second column of Fig. 1 shows the phase distributions of $T_y(x,y)$ with $l_y\,=\,3$ and $\Lambda _x=\,0.1$ mm having sinusoidal (first row) and binary (second row) profiles.

3. Pure phase 2D FSGs: power sharing among diffraction orders

Let us now define a 2D pure phase FSG with the following transmittance:

$$T(x, y) = T_{x}(x, y)T_{y}(x, y).$$

Substituting $T_x(x,y)$ and $T_y(x,y)$ from Eq. (2) in Eq. (5) we get

$$T(x,y) = \exp \left[ {i{\gamma _x}\cos \left( {\frac{{2\pi }}{{{\Lambda _x}}}x - {l_x}\theta } \right)}+{i{\gamma _y}\cos \left( {\frac{{2\pi }}{{{\Lambda _y}}}y- {l_y}\theta } \right)} \right],$$
$${T}(x,y) = \exp \Bigg\{{i{\gamma _x}\text{sign}\left[ {\cos \left( {\frac{{2\pi }}{{{\Lambda _x}}}x - {l_x}\theta } \right)} \right]}+ {i{\gamma _y}\text{sign}\left[ {\cos \left( {\frac{{2\pi }}{{{\Lambda _y}}}y- {l_y}\theta } \right)} \right]} \Bigg\},$$
as the transmission function of the pure phase 2D FSG having sinusoidal and binary profiles, respectively. By using Eq. (6), phase distributions for $T(x,y)$ with $l_x=\,2 , l_y\,=\,3$ and $l_x=l_y=0$ are illustrated in the third and fourth columns of Fig. 1, respectively. It should be mentioned that, based on Eq. (6), the phase amplitude of $T(x,y)$, say $\gamma$, equals sum of the phase amplitudes of $T_x(x,y)$ and $T_y(x,y)$, namely $\gamma =\gamma _x+\gamma _y$. Using Eq. (1), the corresponding 2D Fourier series can be written in the following unified form:
$$T(x,y)= \sum_{m ={-} \infty }^{ + \infty } {\sum_{n ={-} \infty }^{ + \infty } {{t_{m,n}(\gamma_x,\gamma_y ) }{e^{im\left( {\frac{{2\pi }}{\Lambda_x }x - l_x\theta } \right) + in\left( {\frac{{2\pi }}{\Lambda_y }y - l_y\theta } \right)}}} },$$
where ${t_{m,n}(\gamma _x, \gamma _y) }= {t_m}(\gamma _x ){t_n}(\gamma _y )$. Since Eq. (7) represents 2D Fourier expansion of a pure phase 2D FSG, the power share of $(m,n)$th diffraction order from the transmitted/incident light beam equals ${\left | {{t_{m,n}}(\gamma _x, \gamma _y)} \right |^2}$, see Appendix A. As is apparent, ${\left | {{t_{m,n}}(\gamma _x, \gamma _y)} \right |^2}$ is a two-variable function. However, considering the total phase amplitude $\gamma$ as a constant, power contribution of $(m,n)$th diffraction order will be a one-variable function of a $\gamma _x$.

For the sinusoidal profile we have ${t_{m,n}(\gamma _x, \gamma _y) }= {i^{m + n}}{J_m}(\gamma _x ){J_n}(\gamma _y )$ which is generally a two-variable function. Nevertheless, by considering $\gamma =\gamma _x+\gamma _y$ as a constant we get $t_{m,n}(\gamma _x) = i^{m + n}J_m(\gamma _x )J_n(\gamma -\gamma _x )$ which is a one-variable function. Using this identity, the power contributions of the zero to second diffraction orders of a sinusoidal pure phase 2D FSG are plotted in Figs. 4 and 5 in terms of $\gamma _x$ for the $\gamma =\pi$ (first column) and $\gamma =\pi / 2$ (second column). In the first column of Fig. 4, for the zero (central) diffraction order we have ${t_{0,0}}({\gamma _x}) = {J_0}({\gamma _x}){J_0}\left ( {\pi - {\gamma _x}} \right )$ then $\gamma _0 \approx 2.4048$ (the first root of the function $J_0(\gamma _x)$), and $\pi -\gamma _0 \approx 0.7368$ can be considered as the roots of ${t_{0,0}}({\gamma _x})$. It is also seen that setting $\gamma _x=\gamma _0$ the values of $\left | {{t_{0, \pm 1}}({\gamma _x})} \right |$ go to zero and the values of $\left | {{t_{ \pm 1,0}}({\gamma _x})} \right |$ reach to the corresponding maxima. Similarly setting $\gamma _x=\pi -\gamma _0$ the values of $\left | {{t_{0, \pm 1}}({\gamma _x})} \right |$ reach to their maxima and the values of $\left | {{t_{ \pm 1,0}}({\gamma _x})} \right |$ go to zero. In Visualization 2 the power sharing among diffracted beams is illustrated for $\gamma =\pi$ by changing $\gamma _x$. Looking at the second column of Fig. 4, we see that the zero (central) diffraction order is dominant and has the most power share comparing the higher orders. Moreover, the power sharing among diffracted beams for $\gamma =\pi /2$ by changing $\gamma _x$ is illustrated in Visualization 3. Figure 5 shows the power contributions of the second diffraction orders. As is seen, for $\gamma =\pi / 2$ (second column) all the second order power shares are less than $7\%$.

 figure: Fig. 4.

Fig. 4. The power share of the zero and first diffraction orders of a pure phase 2D FSG with a sinusoidal profile having $\gamma =\pi$ (first column) and $\gamma = \pi / 2$ (second column) in terms of $\gamma _x$. See Visualization 2 and Visualization 3.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. The power share of the second diffraction orders in terms of $\gamma _x$ for a a pure phase 2D FSG with a sinusoidal profile having $\gamma =\pi$ (first column) and $\gamma =\pi / 2$ (second column).

Download Full Size | PDF

For the binary profile, considering Eq. (4), when $m$ and/or $n$ gets non-zero even values ${t_{m,n}(\gamma _x, \gamma _y) }=0$. It means the $(m,n)$ the diffraction order vanishes if $m$ and/or $n$ is a non-zero even number. Here again, considering $\gamma =\gamma _x+\gamma _y$ as a constant, we have ${t_{m,n}(\gamma _x) }= {t_m}(\gamma _x ){t_n}(\gamma -\gamma _x )$ where ${t_m}(\gamma _x )$ and ${t_n}(\gamma -\gamma _x )$ are obtained from Eq. (4). For the special case $\gamma =\pi$, using Eq. (4), we get

$$\left|{t_{m,n}(\gamma_x) }\right|= \left| {t_m(\gamma_x )t_n(\gamma_x )} \right|.$$

Therefore, in this case we have $\left | {{t_{0,0}}({\gamma _x})} \right | = {\cos ^2}({\gamma _x})$ which goes to zero for ${\gamma _x}=\pi /2$ and reaches to its maximum for ${\gamma _x}=0,\, \pi$, see the first column, first row of Fig. 6. In this case we also have $\left | {{t_{ \pm 1, \pm 1}}({\gamma _x})} \right | = 4{\sin ^2}({\gamma _x})/{\pi ^2}$ which go to zero for ${\gamma _x}=0,\, \pi$ and reach to their maxima for ${\gamma _x}=\pi /2$. Moreover, we get $\left | {{t_{0, \pm 1}}({\gamma _x})} \right | = \left | {{t_{ \pm 1,0}}({\gamma _x})} \right | = \sin (2{\gamma _x})/\pi$ which goes to zero for ${\gamma _x}=0,\, \pi /2, \, \pi$ and reaches to its maximum for ${\gamma _x}=\pi /4, \, 3\pi /4$. For ${\gamma _x} = {\tan ^{ - 1}}(\pi /2)$ and ${\gamma _x} = \pi - {\tan ^{ - 1}}(\pi /2)$ at which $\left |t_0(\gamma _x)\right |= \left |t_{\pm 1}(\gamma _x)\right |$ we get $\left | {{t_{ \pm 1, \pm 1}}({\gamma _x})} \right | = \left | {{t_{0, \pm 1}}({\gamma _x})} \right |= \left | {{t_{\pm 1,0}}({\gamma _x})} \right |= \left | {{t_{0, 0}}({\gamma _x})} \right |$ by considering Eq. (8). It means that for these values of ${\gamma _x}$ zero and all the first diffraction orders have the same power shares. In Visualization 4 the power sharing among diffracted beams is illustrated for $\gamma =\pi$ by changing $\gamma _x$. Moreover, By comparing second columns of Figs. 4 and 6 we see that for $\gamma =\pi /2$, the power shares of the first diffraction orders for the binary profile are considerably more than sinusoidal profile. The power sharing among diffracted beams by changing $\gamma _x$ is illustrated for $\gamma =\pi /2$ in Visualization 5. It should be noted that power contributions of the second diffraction orders vanish for binary profile and therefore we do not illustrated them.

 figure: Fig. 6.

Fig. 6. The power share of the zero and first diffraction orders of a pure phase 2D FSG with a binary profile having $\gamma =\pi$ (first column) and $\gamma =\pi / 2$ (second column) in terms of $\gamma _x$. See also Visualization 4 and Visualization 5.

Download Full Size | PDF

As a different approach we consider the total phase amplitude $\gamma$ as a variable parameter in which it is equally shared between the corresponding phase amplitudes along $x$ and $y$ directions, namely $\gamma _x=\gamma _y=\frac {\gamma }{2}$. Then the he Fourier coefficient ${t_{m,n}(\gamma _x, \gamma _y) }$ gets the following form:

$$t_{m,n}(\gamma)=t_m\left( \frac{\gamma}{2}\right)t_n\left( \frac{\gamma}{2}\right),$$
which is a one-variable function of $\gamma$ and also we have $t_{m,n}(\gamma )=t_{n,m}(\gamma )$. For the sinusoidal profile Eq. (9) leads to
$${t_{m,n}}(\gamma ) = {i^{m + n}}{J_m}(\gamma /2){J_n}(\gamma /2).$$

Using this identity, the different diffraction orders’ power shares are calculated and the results are depicted in the first column of Fig. 7. For $\gamma _x=\gamma _y=\gamma /2=1.4347$ at which $\left |t_0(\gamma /2)\right |= \left |t_{\pm 1}(\gamma /2)\right |$ we get $\left | {{t_{ \pm 1, \pm 1}}({\gamma })} \right | = \left | {{t_{0, \pm 1}}({\gamma })} \right |= \left | {{t_{\pm 1,0}}({\gamma })} \right |= \left | {{t_{0, 0}}({\gamma })} \right |$ by considering Eq. (9). Consequently, for $\gamma =2\times 1.4347$ all zero and the first diffraction orders have the same power shares. Furthermore, for $\gamma _x=\gamma _y=\gamma /2=1.0825$ the values of $\left | {{t_{0, \pm 1}}({\gamma })} \right |= \left | {{t_{\pm 1,0}}({\gamma })} \right |$ reach to their maximum. In Visualization 6 the power sharing among diffracted beams is animated for a sinusoidal profile and $\gamma _x=\gamma _y=\gamma /2$ by changing $\gamma$.

 figure: Fig. 7.

Fig. 7. First column: the power share of the zero (first row), first (second row), and second (third row) diffraction orders in terms of $\gamma$ for a 2D FSG with a sinusoidal profile. Second column: the power share of the zero (first row), first (second row), and third (third row) diffraction orders in terms of $\gamma$ for a 2D FSG with a binary profile. See Visualization 6 and Visualization 7.

Download Full Size | PDF

For the binary profile, by using Eqs. (4) and (9), the diffraction orders’ power shares are calculated and the results are depicted in the second column of Fig. 7. Comparing Eqs. (8) and (9) we see that $\left |{t_{m,n}}(\gamma )\right |$ and $\left |{t_{m,n}}(\gamma _x )\right |$ have the same functionality except a coefficient $\frac {1}{2}$ in the arguments of Eq. (9). Therefore power share plots in the second column (first and second rows) of Fig. 7 is the half of the plots of the first column of Fig. 6 which are horizontally stretched twice. Looking at the third row of Fig. 7, it is seen that the power shares of the second diffraction orders of the sinusoidal profile are about the power shares of the third diffraction orders of the binary profile. The power sharing among diffracted beams is animated for a binary profile and $\gamma _x=\gamma _y=\gamma /2$ by changing $\gamma$ in Visualization 7.

4. Diffraction of a Gaussian beam from pure phase 2D FSGs

Now suppose that the defined structures by Eq. (6) are illuminated by a Gaussian beam, fixing the waist position in the aperture plane $z = 0$, we can write:

$$\psi (x,y,0) = \exp \left( { - \frac{{{x^2+y^2}}}{{w_0^2}}} \right)T(x,y),$$
where $\psi (x,y,0)$ indicates the complex amplitude of the light immediately after the structure and $w_0$ is the beam waist radius. The complex amplitude of the diffracted light can be calculated using the Fresnel integral as follows:
$$\psi (x,y,z) = h\int\limits_{ - \infty }^{ + \infty } {\int\limits_{ - \infty }^{ + \infty } {\psi (x',y',0){e^{i\alpha \left[ {x{'^2} + y{'^2} - 2\left( {xx' + yy'} \right)} \right]}}dx'dy'} } ,$$
where $\alpha =\frac {\pi }{\lambda z}$ and
$$h =h(x,y,z)= \frac{{\exp \left[ {ikz + i\alpha \left( {{x^2} + {y^2}} \right)} \right]}}{{i\lambda z}},$$
in which $k=\frac {2\pi }{\lambda }$ is the wavenumber. Substituting Eq. (7) in Eq. (11) and substituting the result in Eq. (12) we get
$$\psi (x,y,z) = h\sum_{m ={-} \infty }^{ + \infty } {\sum_{n ={-} \infty }^{ + \infty } {{t_{m,n}}\int\limits_{ - \infty }^{ + \infty } {\int\limits_{ - \infty }^{ + \infty } {{e^{ - \frac{{x{'^2} + y{'^2}}}{{{w^2}}} - 2i\alpha \left( {x'{X_m} + y'{Y_n}} \right) + il_{m,n}\theta '}}} dx'dy'} } } ,$$
where $t_{m,n}=t_{m,n}(\gamma _x, \gamma _y)$, $l_{m,n}=-\left ( ml_x+nl_y \right )$, $w$ is defined as follows:
$$\frac{1}{w^2}= \frac{1}{{w_0^2}} - i\frac{\pi }{{\lambda z}} = \frac{1}{{w_0^2}}\left( {1 - i\frac{{z_0}}{z}} \right),$$
and
$${X_m} = x - \frac{{m\lambda z}}{\Lambda_x },$$
$${Y_n} = y - \frac{{n\lambda z}}{\Lambda_y } .$$

It should be mentioned that $\left ( {\frac {{m\lambda z}}{\Lambda _x },\frac {{n\lambda z}}{\Lambda _y }} \right )$ show the coordinates of $(m,n)$th diffraction order of the grating at the observation plane, see Eq. (45) of Ref. [34]. Therefore Eq. (13) can be considered as a transformation of coordinates to the center of $(m,n)$th diffraction order. Now let us define $\rho _{m,n}$ and $\varphi _{m,n}$ so that ${X_m} = {\rho _{m,n}}\cos \left ( {{\varphi _{m,n}}} \right )$ and ${Y_n} = {\rho _{m,n}}\sin \left ( {{\varphi _{m,n}}} \right )$, then we have $\rho _{m,n}=\sqrt {X_m^2+Y_n^2}$ and ${\varphi _{m,n}} = {\tan ^{ - 1}}\left ( {\frac {{{Y_n}}}{{{X_m}}}} \right )$. In the polar coordinates, Eq. (14) gets the following form:

$$\psi (r,\theta ,z) = h\sum_{m,n ={-} \infty }^{ + \infty } {t_{m,n}\int\limits_0^{2\pi } {\int\limits_0^\infty {{e^{ - {{\left( {\frac{{r'}}{w}} \right)}^2} - 2i\alpha {\rho _{m,n}}r'\cos \left({\theta ' }-{\varphi _{m,n}} \right) + il_{m,n}\theta '}}r'dr'd\theta '} } } ,$$
where $(r,\theta )$ and $(r',\theta ')$ are the polar coordinates at the grating and observation planes, respectively, see Fig. 1 of the Ref. [43]. By using Jacobi-Anger expansion we have
$$e^{ - 2i\alpha {\rho _{m,n}}r'\cos \left({\theta ' }-{\varphi _{m,n}} \right)}= \sum_{s ={-} \infty }^{ + \infty } {i^{-\left| s \right|}}J_{\left| s \right|}\left( {2\alpha {\rho _{m,n}}r'} \right) {e^{ - is\left({\theta ' }- {\varphi _{m,n}} \right)}},$$
where we used $(-i)^s J_s(x)=i^{-\left | s \right |}J_{\left | s \right |}(x)$ which is resulted from $J_{-s}(x)=(-1)^sJ_s(x)$ for any integer $s$. Substituting Eq. (18) in Eq. (17) we get
$$\psi (r,\theta ,z) = h\sum_{m,n,s ={-} \infty }^\infty {t_{m,n}{i^{ - \left| s \right|}}{e^{is{\varphi _{m,n}}}}\int\limits_0^{2\pi } {{e^{i(l_{m,n} - s)\theta }}d\theta \int\limits_0^\infty {{e^{ - {{\left( {\frac{{r'}}{w}} \right)}^2}}}{J_{\left| s \right|}}\left( {2\alpha {\rho _{m,n}}r'} \right)r'dr'} } }.$$

Considering $\int\limits_0^{2\pi } {{e^{i(l - s)\theta }}d\theta = 2\pi {\delta _{l,s}}}$ in which $\delta _{l,s}$ indicates the Kronecker delta, Eq. (19) leads to

$$\psi (r,\theta ,z) = 2\pi h\sum_{m,n ={-} \infty }^{ + \infty } {t_{m,n}}i^{-\left| l_{m,n} \right|}{e^{ i l_{m,n} {\varphi _{m,n}}}}\int\limits_0^\infty {{J_{\left| l_{m,n} \right|}\left( {2\alpha {\rho _{m,n}}r'} \right){e^{ - {{\left( {\frac{{r'}}{w}} \right)}^2}}}r'dr'} } .$$

By using the following reference integral [44]:

$$\int\limits_0^\infty {r{e^{ - a{r^2}}}{J_v}(br)} = \frac{{\sqrt \pi b}}{{8{a^{\frac{3}{2}}}}}\exp \left( { - \frac{{{b^2}}}{{8a}}} \right)\left[ {{I_{\frac{{\nu - 1}}{2}}}\left( {\frac{{{b^2}}}{{8a}}} \right) - {I_{\frac{{\nu + 1}}{2}}}\left( {\frac{{{b^2}}}{{8a}}} \right)} \right], \quad {\mathop{\rm Re}\nolimits} \, a > 0, \, {\mathop{\rm Re}\nolimits} \,\nu >{-}2,$$
in which $I_{\frac {{\nu \pm 1}}{2}}$ denotes modified Bessel functions of order ${\frac {{\nu \pm 1}}{2}}$, Eq. (20) leads to
$$\psi (r,\theta ,z) =h{w^2}\sum_{m,n ={-} \infty }^\infty {{c_{m,n}}{e^{ i l_{m,n}{\varphi _{m,n}}}}{\mathcal{R}_{m,n}}{e^{ - \mathcal{R}_{m,n}^2}}\left[ {{I_{\frac{{\left| l_{m,n} \right| - 1}}{2}}}\left( {\mathcal{R}_{m,n}^2} \right) - {I_{\frac{{\left| l_{m,n} \right| + 1}}{2}}}\left( {\mathcal{R}_{m,n}^2} \right)} \right]} ,$$
where ${\mathcal {R}_{m,n}}= \frac {\pi }{{\sqrt 2 }}\frac {{w{\rho _{m,n}}}}{{\lambda z}}$ is a dimensionless parameter, $c_{m,n}=\frac {{{{(2\pi )}^{\frac {3}{2}}}}}{4}{i^{- \left | l_{m,n} \right |}}t_{m,n}(\gamma _x, \gamma _y)$, and rewriting Eq. (13) in polar coordinates we have
$$h=h(r,z) = \frac{{\exp \left[ {i\left( {kz + \alpha {r^2}} \right)} \right]}}{{i\lambda z}}.$$

Equation (22) is the main result and specifies the complex amplitude of diffraction of a Gaussian beam from a 2D FSG at any arbitrary propagation distance $z$. Equation (22) implies that $(m,n)$th diffraction order is a vortex beam which its TC equals $l_{m,n}=-\left ( ml_x+nl_y \right )$ and its power share equals ${\left | {{t_{m,n}}(\gamma _x, \gamma _y)} \right |^2}$.

For better demonstration of the results, let us consider some examples. As the first example, we consider the diffraction of a Gaussian beam with $\lambda =532$ nm and $w_0\,=\,2.5$ mm from four typical pure phase 2D FSGs having $l_x=l_y=1$ and $\Lambda _x=\Lambda _y=0.1$ mm with sinusoidal and binary profiles. The resulted diffraction patterns at distance $z=2$ m from the gratings are illustrated in Fig. 8 in which $\gamma _x=\gamma _y=\frac {\pi }{2}$ in the first column, and $\gamma _x=\gamma _y=\frac {\pi }{4}$ in the second column. Considering $l_x=l_y$ and according to $l_{m,n}=-\left ( ml_x+nl_y \right )$, the intensity patterns of $(m,n)$th diffraction orders over the second and fourth Cartesian quadrants ($n=-m$) are not doughnut-shaped and their TCs are zero. Considering the first row of Fig. 8, from the power sharing perspective, the power shares of the $(\pm 1, \pm 1 )$th orders are dominant (negligible) in the first (second ) column. This feature was predicted in the first row of Fig. 4, by comparing red dashed plots in the first and second columns. Furthermore, in the second row, the considerable power sharing difference between the first and second columns of Fig. 8 can be also explained by comparing the first and second columns of Fig. 6. The diffraction patterns of Fig. 8 are animated in Visualization 8, Visualization 9, Visualization 10 and Visualization 11 under propagation from $z=0$ to $z=2$ m.

 figure: Fig. 8.

Fig. 8. Diffraction patterns at distance $z=2$ m from four typical pure phase 2D FSGs having $l_x=l_y=1$, $\Lambda _x=\Lambda _y$ = 0.1 mm, $\gamma _x=\gamma _y=\frac {\pi }{2}$ (first column), $\gamma _x=\gamma _y=\frac {\pi }{4}$ (second column), having sinusoidal (first row), and binary (second row) profiles illuminated by a Gaussian beam with $\lambda$ = 532 nm and $w_0$ = 2.5 mm. See Visualization 8, Visualization 9, Visualization 10 and Visualization 11.

Download Full Size | PDF

To explicitly demonstrate the TCs of the diffraction orders, and verify the identity $l_{m,n}=-\left ( ml_x+nl_y \right )$ for TC of the $(m,n)$th diffraction order, we depict the corresponding phase profiles of the intensity patterns presented in the first row of Fig. 8 in Fig. 9 using Eq. (22). For the sake of simplicity in representation we set $\Lambda =500$ nm instead of $\Lambda =532$ nm so that the coordinates of the center of $(m,n)$th diffraction order at distance $z=2$ m is obtained $(10m,10n)$ in millimetres. For instant, the phase profile of the diffraction order $(1,1)$ around the point $(x=10\, \text {mm}, y=10\,\text {mm})$ includes two singularities with TC=-1 so that the total TC equals −2. As another example, looking carefully at the $(-2,-1)$th diffraction order, it includes three singular points around the point $(x=-20\, \text {mm}, y=-10\,\text {mm})$ with TC=+1 so that the total TC equals +3.

 figure: Fig. 9.

Fig. 9. The corresponding phase profiles of the diffraction patterns illustrated in the first row of Fig. 8. The central areas of diffraction orders are covered by the patterns.

Download Full Size | PDF

As the second example, the diffraction of a Gaussian beam with $\lambda =532$ nm and $w_0=2.5$ mm from six pure phase 2D FSGs having $l_x=2, l_y=3$ and $\Lambda _x=\Lambda _y=0.1$ mm and different values of $\gamma _x$ and $\gamma _y$ with sinusoidal and binary profiles. The resulted diffraction patterns at a distance $z=2$ m from the gratings are depicted in Fig. 10 in which the first and second rows are allocated to the sinusoidal and binary profiles, respectively. The values $\gamma _x$ and $\gamma _y$ have been chosen based on the extrema and intersections of the plots of the first-order power shares in Figs. 4, 6, and 7. As is seen in the first column of Fig. 4, for a grating with a sinusoidal profile and $\gamma = \pi$, for $\gamma _x=\gamma _0 \approx 2.4048$ and $\gamma _x=\pi -\gamma _0 \approx 0.7368$ the power shares of $(\pm 1,0)$/$(0,\pm 1)$ and $(0,\pm 1)$/$(\pm 1,0)$ diffraction orders reach to maximum/zero, respectively, and the $(0,0)$ diffraction order vanishes for both values of $\gamma _x$. This feature is verified and demonstrated at the first row, the first and second columns, of Fig. 10. Moreover, as was mentioned in the explanations of the first column of Fig. 7, for $\gamma _x=\gamma _y=1.4347$ power is shared equally among zero and all the first diffraction orders. Thus, a $3 \times 3$ array of diffracted beams is generated with different TCs and having total power share of $81\%$ from the incident beam’s power, see the first row third column of Fig. 10. This feature might find application in free-space optical communications. For the binary profile, returning to the first column of Fig. 6, we see that for $\gamma _x=\gamma _y=\pi /2$, the $(0,0)$, $(0,\pm 1)$ and $(\pm 1,0)$ diffraction orders vanish. This result is reflected in the second row first column of Fig. 10. As an interesting result, we see that the total power share of four illustrated vortices is more than $65\%$ of the incident beam’s power. This feature can be used for generation of high power optical vortices. In the second row second column of Fig. 10 the values of $\gamma _x$ and $\gamma _y$ are chosen so that the power shares of $(0,\pm 1)$ and $(\pm 1,0)$ diffraction orders gets their maximum values. In the second row third column we choose the values of $\gamma _x$ and $\gamma _y$ so that all the zero and all the first diffraction orders have the same power shares. Therefore an isolated $3 \times 3$ array of diffracted beams is generated with total power share of $74\%$ from the incident beam’s power. As it was proved in the previous section, all of the second diffraction orders vanish for a pure phase FSG with a binary profile. In Visualization 12, Visualization 13, Visualization 14, Visualization 15, Visualization 16 and Visualization 17 the diffraction patterns of Fig. 10 are animated under propagation from $z=0$ to $z=2$ m.

 figure: Fig. 10.

Fig. 10. Diffraction patterns at distance z = 2 m from three different pure phase 2D FSGs with $l_x$ = 2, $l_y$ = 3, $\Lambda _x=\Lambda _y$ = 0.1 mm, and different values of $\gamma _x$ and $\gamma _y$ having sinusoidal (first row), and binary (second row) profiles illuminated by a Gaussian beam with $\lambda$ = 532 nm and $w_0$ = 2.5 mm. See Visualization 12, Visualization 13, Visualization 14, Visualization 15, Visualization 16 and Visualization 17.

Download Full Size | PDF

5. Experiments

We examine some of the presented theoretical results experimentally by impinging a Gaussian beam over some pure phase 2D FSGs generated by using a spatial light modulator (SLM). We used a conventional SLM extracted from a video projector (LCD projector KM3, model no. X50) for which the total phase amplitude of the prepared gratings is limited to $\gamma =\frac {\pi }{2}$. The second harmonic of an Nd:YAG diode-pumped laser beam having a Gaussian profile and collimated wavefront with a wavelength $\lambda =532$ nm impinges the SLM where the desired phase profile is imposed on it. At distance $z=300$ cm from the SLM, we recorded the diffraction pattern by a camera (Nikon D100). Six different 2D FSGs having sinusoidal and binary profiles with $l_x= l_y=5$, and different values of $\gamma _x$ and $\gamma _y$ are imposed on the SLM. The experimentally recorded and the corresponding calculated diffraction patterns are illustrated in Fig. 11. The same patterns for $l_x=3\; {\textrm {and}} \; l_y=-2$ are shown in Fig. 12. Because of $\gamma =\frac {\pi }{2}$ limitation of the used SLM we have $\gamma _x+\gamma _y=\frac {\pi }{2}$ for all of the illustrated cases.

 figure: Fig. 11.

Fig. 11. Experimentally recorded (first and third rows) and calculated (second and fourth rows) diffraction patterns of a Gaussian beam with $\lambda =532$ nm and $w_0=5$ mm from six different pure phase 2D FSGs with $l_x= l_y=5$, $\Lambda _x=\Lambda _y=0.1$ mm, and different values of $\gamma _x$ and $\gamma _y$ having sinusoidal (first and second rows), and binary (third and fourth rows) profiles at distance $z=300$ cm from the gratings.

Download Full Size | PDF

 figure: Fig. 12.

Fig. 12. The same patterns of Fig. 11 for $l_x$ = 3 and $l_y$ = −2.

Download Full Size | PDF

In the next stage, we directly measure the TCs of optical vortices generated over the diffraction orders of a pure phase 2D FSG to verify the identity $l_{m,n}=-\left ( ml_x+nl_y \right )$ experimentally. In this regard, we use the diffraction from a quadratic curved-line (parabolic-line) grating introduced in Ref. [24] as a simple and efficient tool for characterization of optical vortices. In this method, we let each of the generated vortex beams individually illuminates a quadratic curved-line grating (also known parabolic-line grating), see Fig. 13. The first diffraction order of the quadratic curved-line grating is only a set of elongated intensity spots along a straight line. The number of intensity spots over the first-order diffraction pattern and its orientation respectively determine the absolute value and sign of the TC of the impinging vortex beam. The absolute value of the TC is obtained as $\left | l \right | =N-1$ where $N$ denotes the number of elongated intensity spots. Alternatively, we can account the number of intensity nulls between elongated intensity spots as the absolute value of TC. If the orientation of the first-order diffraction pattern is placed along the first and third (second and fourth) Cartesian quadrants the sign of TC is negative (positive). Figure 14 illustrates the first-order diffraction patterns of the optical vortices generated in the second columns of Fig. 11 and Fig. 12 passing through a quadratic curved-line grating.

 figure: Fig. 13.

Fig. 13. Experimental setup for generating a 2D array of optical vortices by a 2D FSG imposed on an SLM and measuring of their TCs using a quadratic curved-line grating (also known parabolic-line grating, PLG). S.F. stands for spatial filter.

Download Full Size | PDF

 figure: Fig. 14.

Fig. 14. First-order diffraction patterns of the optical vortices generated in the second columns of Fig. 9 (first to third columns) and Fig. 10 (fourth to sixth columns) passing through a quadratic curved-line grating.

Download Full Size | PDF

6. Conclusion

A new class of pure phase 2D FGSs was introduced having topological defect numbers $l_x$ and $l_y$ and phase variation amplitudes $\gamma _x$ and $\gamma _y$ with sinusoidal and binary profiles. Diffraction of a Gaussian beam from this class of grating leads to generation of 2D arrays of optical vortices having different power and TC distributions. It was shown that the TC of $(m,n)$ diffraction order is $l_{m,n}=-\left ( ml_x+nl_y \right )$ and the power distribution among diffraction orders can be managed by adjusting the values of $\gamma _x$ and $\gamma _y$. For instance, setting $\gamma _x=\gamma _y=1.435$ for the sinusoidal profile and $\gamma _x=\gamma _y= {\tan ^{ - 1}}(\pi /2) \approx 1.004$ for the binary profile lead to generation of a $3 \times 3$ array of optical beams having equal powers and different TCs. This feature has application in OAM based free-space optical communications. Furthermore, by setting $\gamma _x=\gamma _y=\pi /2$ for the binary profile we obtain a $2 \times 2$ array of optical vortices with considerable power shares and different TCs. This feature might find application in the field of optical tweezers. Finally, some of the presented theoretical results experimentally verified by impinging a Gaussian beam over some pure phase 2D FSGs generated by using a spatial light modulator (SLM). Moreover, TCs of the generated optical vortices have been measured by using a parabolic-line grating to verify the identity $l_{m,n}=-\left ( ml_x+nl_y \right )$ experimentally.

Appendix A

The transmittance of a 2D orthogonal periodic structure can be expressed in terms of a 2D Fourier series as follows [45]:

$$T(x,y) = \sum_{m,n ={-} \infty }^{ + \infty } {{t_{m,n}}\exp \left[ {2\pi i\left( {\frac{m}{{{\Lambda _x}}}x + \frac{n}{{{\Lambda _y}}}y} \right)} \right]},$$
where $\Lambda _x$ and $\Lambda _y$ are the periods in the $x$ and $y$ directions, respectively, and $t_{m,n}$ indicate the $(m,n)$th Fourier coefficients. In the diffraction of a light beam from this structure the power contribution of the $(m,n)$th diffraction order from the power of incident beam is
$$\frac{p_{m,n}}{p_i}=\left| t_{m,n} \right|^2 ,$$
where $p_{m,n}$ and $p_i$ are power of the $(m,n)$th diffraction order and power of the incident beam, respectively. Furthermore $\left | t_{m,n} \right |^2$ denotes the absolute square of $(m,n)$th Fourier coefficient. Calculating the summation of Eq. (25) over all $(m,n)$s we get
$$\frac{1}{{{p_i}}}\sum_{m ={-} \infty }^{ + \infty } {\sum_{n ={-} \infty }^{ + \infty } {{p_{m,n}}} } = \sum_{m ={-} \infty }^{ + \infty } {\sum_{n ={-} \infty }^{ + \infty } {{{\left| {{t_{m,n}}} \right|}^2}} } .$$

Using energy conservation, summation of all the diffraction orders’ powers equals the transmitted power,

$${p_t} = \sum_{m ={-} \infty }^{ + \infty } {\sum_{n ={-} \infty }^{ + \infty } {{p_{m,n}}} },$$
where $p_t$ indicates the transmitted power. Substituting Eq. (27) in Eq. (26) we get
$$\frac{p_t}{p_i}= \sum_{m ={-} \infty }^{ + \infty } {\sum_{n ={-} \infty }^{ + \infty } {{{\left| {{t_{m,n}}} \right|}^2}} } .$$
Dividing through both sides of Eq. (25) by both sides of Eq. (28) we obtain
$$\frac{p_{m,n}}{p_t}=\frac{{{{\left| {{t_{m,n}}} \right|}^2}}}{{\sum\limits_{m' ={-} \infty }^{ + \infty } {\sum\limits_{n' ={-} \infty }^{ + \infty } {{{\left| {{t_{m',n'}}} \right|}^2}} } }},$$
as the power contribution of the $(m,n)$th diffraction order from the transmitted power. It should be mentioned that Eq. (28) is completely consistent with the mathematical Parseval’s theorem in two dimensions [46]
$$\sum_{m ={-} \infty }^{ + \infty } {\sum_{n ={-} \infty }^{ + \infty } {{{\left| {{t_{m,n}}} \right|}^2}} } = \frac{1}{{{\Lambda _x}{\Lambda _y}}}\int\limits_{ - {\Lambda _x}/2}^{{\Lambda _x}/2} {\int\limits_{-{\Lambda _y}/2}^{{\Lambda _y}/2} {{{\left| {T(x,y)} \right|}^2}dydx} }.$$

For a pure phase transmittance we have ${\left | {T(x,y)} \right |}^2=1$, then Eq. (30) reduces to

$$\sum_{m ={-} \infty }^{ + \infty } {\sum_{n ={-} \infty }^{ + \infty } {{{\left| {{t_{m,n}}} \right|}^2}} } =1.$$

Substituting Eq. (31) in Eq. (29) one can deduce that the power contribution of the $(m,n)$th diffraction order from transmitted/incident power is equal with $\left | t_{m,n} \right |^2$ for a pure phase periodic structure.

Funding

Iran National Science Foundation (99002408); Institute for Advanced Studies in Basic Sciences (G2022IASBS12632).

Acknowledgment

The author S. Rasouli would like to acknowledge the Abdus Salam International Center for Theoretical Physics (ICTP), Trieste, Italy, for the Senior Associate Fellowship.

Disclosures

The authors declare no conflicts of interest.

Data availability

No data were generated or analyzed in the presented research.

References

1. G. J. Gbur, Singular Optics (CRC Press, 2016).

2. M. Beijersbergen, R. Coerwinkel, M. Kristensen, and J. Woerdman, “Helical-wavefront laser beams produced with a spiral phaseplate,” Opt. Commun. 112(5-6), 321–327 (1994). [CrossRef]  

3. V. V. Kotlyar, A. A. Almazov, S. N. Khonina, V. A. Soifer, H. Elfstrom, and J. Turunen, “Generation of phase singularity through diffracting a plane or gaussian beam by a spiral phase plate,” J. Opt. Soc. Am. A 22(5), 849–861 (2005). [CrossRef]  

4. V. Y. Bazhenov, M. Vasnetsov, and M. Soskin, “Laser beams with screw dislocations in their wavefronts,” Jetp Lett 52, 429–431 (1990).

5. N. Heckenberg, R. McDuff, C. Smith, H. Rubinsztein-Dunlop, and M. Wegener, “Laser beams with phase singularities,” Opt. Quantum Electron. 24(9), S951–S962 (1992). [CrossRef]  

6. G. Brand, “Phase singularities in beams,” Am. J. Phys. 67(1), 55–60 (1999). [CrossRef]  

7. H. He, N. Heckenberg, and H. Rubinsztein-Dunlop, “Optical particle trapping with higher-order doughnut beams produced using high efficiency computer generated holograms,” J. Mod. Opt. 42(1), 217–223 (1995). [CrossRef]  

8. S. Topuzoski, “Generation of optical vortices with curved fork-shaped holograms,” Opt. Quantum Electron. 48(2), 138 (2016). [CrossRef]  

9. P. Soleimani, H. Khoshsima, and M. Yeganeh, “Optical vortex beam controlling based on fork grating stored in a dye-doped liquid crystal cell,” Sci. Rep. 12(1), 21271 (2022). [CrossRef]  

10. N. Heckenberg, R. McDuff, C. Smith, and A. White, “Generation of optical phase singularities by computer-generated holograms,” Opt. Lett. 17(3), 221–223 (1992). [CrossRef]  

11. Y. Liang, E. Wang, Y. Hua, C. Xie, and T. Ye, “Single-focus spiral zone plates,” Opt. Lett. 42(13), 2663–2666 (2017). [CrossRef]  

12. Z. Ji, H. Zang, C. Fan, J. Wang, C. Zheng, L. Wei, C. Wang, and L. Cao, “Fractal spiral zone plates,” J. Opt. Soc. Am. A 35(5), 726–731 (2018). [CrossRef]  

13. L. Marrucci, C. Manzo, and D. Paparo, “Optical spin-to-orbital angular momentum conversion in inhomogeneous anisotropic media,” Phys. Rev. Lett. 96(16), 163905 (2006). [CrossRef]  

14. E. Karimi, B. Piccirillo, E. Nagali, L. Marrucci, and E. Santamato, “Efficient generation and sorting of orbital angular momentum eigenmodes of light by thermally tuned q-plates,” Appl. Phys. Lett. 94(23), 231124 (2009). [CrossRef]  

15. C.-S. Guo, L.-L. Lu, and H.-T. Wang, “Characterizing topological charge of optical vortices by using an annular aperture,” Opt. Lett. 34(23), 3686–3688 (2009). [CrossRef]  

16. J. Hickmann, E. Fonseca, W. Soares, and S. Chávez-Cerda, “Unveiling a truncated optical lattice associated with a triangular aperture using light’s orbital angular momentum,” Phys. Rev. Lett. 105(5), 053904 (2010). [CrossRef]  

17. L. E. de Araujo and M. E. Anderson, “Measuring vortex charge with a triangular aperture,” Opt. Lett. 36(6), 787–789 (2011). [CrossRef]  

18. D. Hebri, S. Rasouli, and A. M. Dezfouli, “Theory of diffraction of vortex beams from structured apertures and generation of elegant elliptical vortex hermite–gaussian beams,” J. Opt. Soc. Am. A 36(5), 839–852 (2019). [CrossRef]  

19. S. Hosseini-Saber, E. A. Akhlaghi, and A. Saber, “Diffractometry-based vortex beams fractional topological charge measurement,” Opt. Lett. 45(13), 3478–3481 (2020). [CrossRef]  

20. I. Moreno, J. A. Davis, B. M. L. Pascoguin, M. J. Mitry, and D. M. Cottrell, “Vortex sensing diffraction gratings,” Opt. Lett. 34(19), 2927–2929 (2009). [CrossRef]  

21. K. Dai, C. Gao, L. Zhong, Q. Na, and Q. Wang, “Measuring oam states of light beams with gradually-changing-period gratings,” Opt. Lett. 40(4), 562–565 (2015). [CrossRef]  

22. S. Zheng and J. Wang, “Measuring orbital angular momentum (oam) states of vortex beams with annular gratings,” Sci. Rep. 7(1), 1–9 (2017). [CrossRef]  

23. D. Hebri, S. Rasouli, and M. Yeganeh, “Intensity-based measuring of the topological charge alteration by the diffraction of vortex beams from amplitude sinusoidal radial gratings,” J. Opt. Soc. Am. B 35(4), 724–730 (2018). [CrossRef]  

24. P. Amiri, A. M. Dezfouli, and S. Rasouli, “Efficient characterization of optical vortices via diffraction from parabolic-line linear gratings,” J. Opt. Soc. Am. B 37(9), 2668–2677 (2020). [CrossRef]  

25. S. Rasouli, S. Fathollazade, and P. Amiri, “Simple, efficient and reliable characterization of laguerre-gaussian beams with non-zero radial indices in diffraction from an amplitude parabolic-line linear grating,” Opt. Express 29(19), 29661–29675 (2021). [CrossRef]  

26. J. Leach, J. Courtial, K. Skeldon, S. M. Barnett, S. Franke-Arnold, and M. J. Padgett, “Interferometric methods to measure orbital and spin, or the total angular momentum of a single photon,” Phys. Rev. Lett. 92(1), 013601 (2004). [CrossRef]  

27. M. Yeganeh, S. Rasouli, M. Dashti, S. Slussarenko, E. Santamato, and E. Karimi, “Reconstructing the poynting vector skew angle and wavefront of optical vortex beams via two-channel moiré deflectometery,” Opt. Lett. 38(6), 887–889 (2013). [CrossRef]  

28. X. Wang, Z. Nie, Y. Liang, J. Wang, T. Li, and B. Jia, “Recent advances on optical vortex generation,” Nanophotonics 7(9), 1533–1556 (2018). [CrossRef]  

29. B. Knyazev, O. Kameshkov, N. Vinokurov, V. Cherkassky, Y. Choporova, and V. Pavelyev, “Quasi-talbot effect with vortex beams and formation of vortex beamlet arrays,” Opt. Express 26(11), 14174–14185 (2018). [CrossRef]  

30. Z.-Y. Rong, Y.-J. Han, L. Zhang, and X.-Y. Chen, “Generation of a vortex and helix with square arrays with high-efficiency by the use of a 2d binary phase mask,” OSA Continuum 2(12), 3482–3489 (2019). [CrossRef]  

31. D. A. Ikonnikov, S. A. Myslivets, M. N. Volochaev, V. G. Arkhipkin, and A. M. Vyunishev, “Two-dimensional talbot effect of the optical vortices and their spatial evolution,” Sci. Rep. 10(1), 20315 (2020). [CrossRef]  

32. S. Schwarz, C. Kapahi, R. Xu, A. R. Cameron, D. Sarenac, J.-P. W. MacLean, K. B. Kuntz, D. G. Cory, T. Jennewein, K. J. Resch, and D. A. Pushin, “Talbot effect of orbital angular momentum lattices with single photons,” Phys. Rev. A 101(4), 043815 (2020). [CrossRef]  

33. D. Hebri and S. Rasouli, “Theoretical study on the diffraction-based generation of a 2d orthogonal lattice of optical beams: physical bases and application for a vortex beam multiplication,” J. Opt. Soc. Am. A 39(9), 1694–1711 (2022). [CrossRef]  

34. S. Rasouli, D. Hebri, and A. M. Khazaei, “Investigation of various behaviors of near-and far-field diffractions from multiplicatively separable structures in the x and y directions, and a detailed study of the near-field diffraction patterns of 2d multiplicatively separable periodic structures using the contrast variation method,” J. Opt. 19(9), 095601 (2017). [CrossRef]  

35. D. Hebri and S. Rasouli, “Diffraction from two-dimensional orthogonal nonseparable periodic structures: Talbot distance dependence on the number theoretic properties of the structures,” J. Opt. Soc. Am. A 36(2), 253–263 (2019). [CrossRef]  

36. G. Gibson II, J. Courtial, M. Vasnetsov, S. Barnett, S. Franke-Arnold, and M. Padgett, “Increasing the data density of free-space optical communications using orbital angular momentum,” in Free-Space Laser Communications IV, vol. 5550 (SPIE, 2004), pp. 367–373.

37. R. Gerhberg and W. Saxton, “’a practical algorithm for the determination of phase from image and diffraction plane picture,” Optik (Stuttgart) 35, 237–246 (1972).

38. G. Gibson, J. Courtial, M. J. Padgett, M. Vasnetsov, V. Pas’ko, S. M. Barnett, and S. Franke-Arnold, “Free-space information transfer using light beams carrying orbital angular momentum,” Opt. Express 12(22), 5448–5456 (2004). [CrossRef]  

39. S. Rasouli and M. Yeganeh, “Formulation of the moiré patterns formed by superimposing of gratings consisting topological defects: moiré technique as a tool in singular optics detections,” J. Opt. 17(10), 105604 (2015). [CrossRef]  

40. L. Janicijevic and S. Topuzoski, “Fresnel and fraunhofer diffraction of a gaussian laser beam by fork-shaped gratings,” J. Opt. Soc. Am. A 25(11), 2659–2669 (2008). [CrossRef]  

41. S. Rasouli, A. M. Khazaei, and D. Hebri, “Radial carpet beams: a class of nondiffracting, accelerating, and self-healing beams,” Phys. Rev. A 97(3), 033844 (2018). [CrossRef]  

42. M. Zarei, D. Hebri, and S. Rasouli, “1 d spatially chirped periodic structures: managing their spatial spectrum and investigating their near-field diffraction,” J. Opt. Soc. Am. A 39(12), 2354–2375 (2022). [CrossRef]  

43. S. Rasouli, A. M. Khazaei, and D. Hebri, “Talbot carpet at the transverse plane produced in the diffraction of plane wave from amplitude radial gratings,” J. Opt. Soc. Am. A 35(1), 55–64 (2018). [CrossRef]  

44. D. Zwillinger and A. Jeffrey, Table of Integrals, Series, and Products (Elsevier, 2007).

45. I. Amidror, The Theory of the Moiré Phenomenon: Volume I: Periodic Layers (Springer, 2009).

46. R. Bracewell, Fourier Analysis and Imaging (Springer Science & Business Media, 2004).

Supplementary Material (17)

NameDescription
Visualization 1       Evolution of the power shares of the zero to third diffraction orders for pure phase 1D FSGs with sinusoidal (first row) and binary (second row) profiles by changing ?_x.
Visualization 2       Evolution of the power shares of the zero to second diffraction orders for a pure phase 2D FSG with ?=p and sinusoidal profile by changing ?_x.
Visualization 3       Evolution of the power shares of the zero to second diffraction orders for a pure phase 2D FSG with ?=p/2 and sinusoidal profile by changing ?_x.
Visualization 4       Evolution of the power shares of the zero to second diffraction orders for a pure phase 2D FSG with ?=p and binary profile by changing ?_x.
Visualization 5       Evolution of the power shares of the zero to second diffraction orders for a pure phase 2D FSG with ?=p/2 and binary profile by changing ?_x.
Visualization 6       Evolution of the power shares of the zero to second diffraction orders for a pure phase 2D FSG with ?_x=?_y=?/2 and sinusoidal profile by changing.
Visualization 7       Evolution of the power shares of the zero to second diffraction orders for a pure phase 2D FSG with ?_x=?_y=?/2 and binary profile by changing ?.
Visualization 8       Diffraction of a Gaussian beam with ?=532 nm and w_0=2.5 mm from a pure phase 2D FSG having a sinusoidal profile with l_x=l_y=1, ?_x=?_y=0.1 mm, and ?_x=?_y=p/2 under propagation from z=0 to z=2 m.
Visualization 9       Diffraction of a Gaussian beam with ?=532 nm and w_0=2.5 mm from a pure phase 2D FSG having a sinusoidal profile with l_x=l_y=1, ?_x=?_y=0.1 mm, and ?_x=?_y=p/4 under propagation from z=0 to z=2 m.
Visualization 10       Diffraction of a Gaussian beam with ?=532 nm and w_0=2.5 mm from a pure phase 2D FSG having a binary profile with l_x=l_y=1, ?_x=?_y=0.1 mm, and ?_x=?_y=p/2 under propagation from z=0 to z=2 m.
Visualization 11       Diffraction of a Gaussian beam with ?=532 nm and w_0=2.5 mm from a pure phase 2D FSG having a binary profile with l_x=l_y=1, ?_x=?_y=0.1 mm, and ?_x=?_y=p/4 under propagation from z=0 to z=2 m.
Visualization 12       Diffraction of a Gaussian beam with ?=532 nm and w_0=2.5 mm from a pure phase 2D FSG having a sinusoidal profile with l_x=2,l_y=3, ?_x=?_y=0.1 mm, and ?_x=2405, ?_y=0.737 under propagation from z=0 to z=2 m.
Visualization 13       Diffraction of a Gaussian beam with ?=532 nm and w_0=2.5 mm from a pure phase 2D FSG having a sinusoidal profile with l_x=2,l_y=3, ?_x=?_y=0.1 mm, and ?_x=0.737 , ?_y=2405 under propagation from z=0 to z=2 m.
Visualization 14       Diffraction of a Gaussian beam with ?=532 nm and w_0=2.5 mm from a pure phase 2D FSG having a sinusoidal profile with l_x=2,l_y=3, ?_x=?_y=0.1 mm, and ?_x=?_y=1.435 under propagation from z=0 to z=2 m.
Visualization 15       Diffraction of a Gaussian beam with ?=532 nm and w_0=2.5 mm from a pure phase 2D FSG having a binary profile with l_x=2,l_y=3, ?_x=?_y=0.1 mm, and ?_x=?_y=p/2 under propagation from z=0 to z=2 m.
Visualization 16       Diffraction of a Gaussian beam with ?=532 nm and w_0=2.5 mm from a pure phase 2D FSG having a binary profile with l_x=2,l_y=3, ?_x=?_y=0.1 mm, and ?_x=?_y=p/4 under propagation from z=0 to z=2 m.
Visualization 17       Diffraction of a Gaussian beam with ?=532 nm and w_0=2.5 mm from a pure phase 2D FSG having a binary profile with l_x=2,l_y=3, ?_x=?_y=0.1 mm, and ?_x=?_y=1.004 under propagation from z=0 to z=2 m.

Data availability

No data were generated or analyzed in the presented research.

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 (14)

Fig. 1.
Fig. 1. Phase profiles of $T_x(x,y)$ with $l_x=\,2$ (first column), $T_y(x,y)$ with $l_y\,=\,3$ (second column), $T(x,y)$ with $l_x=\,2 , l_y\,=\,3$ (third column), and $T(x,y)$ with $l_x=l_y=0$ (fourth column) having $\gamma _x=\gamma _y=\frac {\gamma }{2}$ and $\Lambda _x=\Lambda _y=\,0.1$ mm. The first and second rows are for the sinusoidal and binary transmittances, respectively.
Fig. 2.
Fig. 2. First row: power shares of the zero to second diffraction orders of a pure phase 1D FSG with a sinusoidal profile in terms of $\gamma _x$. Second to fifth rows: power shares of the zero to third diffraction orders for different values of $\gamma _x$. See the first row of Visualization 1.
Fig. 3.
Fig. 3. First row: power shares of the zero, first, and third diffraction orders of a pure phase 1D FSG with a binary profile in terms of $\gamma _x$. Second to third rows: power shares of the zero to third diffraction orders for different values of $\gamma _x$. See the second row of Visualization 1.
Fig. 4.
Fig. 4. The power share of the zero and first diffraction orders of a pure phase 2D FSG with a sinusoidal profile having $\gamma =\pi$ (first column) and $\gamma = \pi / 2$ (second column) in terms of $\gamma _x$. See Visualization 2 and Visualization 3.
Fig. 5.
Fig. 5. The power share of the second diffraction orders in terms of $\gamma _x$ for a a pure phase 2D FSG with a sinusoidal profile having $\gamma =\pi$ (first column) and $\gamma =\pi / 2$ (second column).
Fig. 6.
Fig. 6. The power share of the zero and first diffraction orders of a pure phase 2D FSG with a binary profile having $\gamma =\pi$ (first column) and $\gamma =\pi / 2$ (second column) in terms of $\gamma _x$. See also Visualization 4 and Visualization 5.
Fig. 7.
Fig. 7. First column: the power share of the zero (first row), first (second row), and second (third row) diffraction orders in terms of $\gamma$ for a 2D FSG with a sinusoidal profile. Second column: the power share of the zero (first row), first (second row), and third (third row) diffraction orders in terms of $\gamma$ for a 2D FSG with a binary profile. See Visualization 6 and Visualization 7.
Fig. 8.
Fig. 8. Diffraction patterns at distance $z=2$ m from four typical pure phase 2D FSGs having $l_x=l_y=1$, $\Lambda _x=\Lambda _y$ = 0.1 mm, $\gamma _x=\gamma _y=\frac {\pi }{2}$ (first column), $\gamma _x=\gamma _y=\frac {\pi }{4}$ (second column), having sinusoidal (first row), and binary (second row) profiles illuminated by a Gaussian beam with $\lambda$ = 532 nm and $w_0$ = 2.5 mm. See Visualization 8, Visualization 9, Visualization 10 and Visualization 11.
Fig. 9.
Fig. 9. The corresponding phase profiles of the diffraction patterns illustrated in the first row of Fig. 8. The central areas of diffraction orders are covered by the patterns.
Fig. 10.
Fig. 10. Diffraction patterns at distance z = 2 m from three different pure phase 2D FSGs with $l_x$ = 2, $l_y$ = 3, $\Lambda _x=\Lambda _y$ = 0.1 mm, and different values of $\gamma _x$ and $\gamma _y$ having sinusoidal (first row), and binary (second row) profiles illuminated by a Gaussian beam with $\lambda$ = 532 nm and $w_0$ = 2.5 mm. See Visualization 12, Visualization 13, Visualization 14, Visualization 15, Visualization 16 and Visualization 17.
Fig. 11.
Fig. 11. Experimentally recorded (first and third rows) and calculated (second and fourth rows) diffraction patterns of a Gaussian beam with $\lambda =532$ nm and $w_0=5$ mm from six different pure phase 2D FSGs with $l_x= l_y=5$, $\Lambda _x=\Lambda _y=0.1$ mm, and different values of $\gamma _x$ and $\gamma _y$ having sinusoidal (first and second rows), and binary (third and fourth rows) profiles at distance $z=300$ cm from the gratings.
Fig. 12.
Fig. 12. The same patterns of Fig. 11 for $l_x$ = 3 and $l_y$ = −2.
Fig. 13.
Fig. 13. Experimental setup for generating a 2D array of optical vortices by a 2D FSG imposed on an SLM and measuring of their TCs using a quadratic curved-line grating (also known parabolic-line grating, PLG). S.F. stands for spatial filter.
Fig. 14.
Fig. 14. First-order diffraction patterns of the optical vortices generated in the second columns of Fig. 9 (first to third columns) and Fig. 10 (fourth to sixth columns) passing through a quadratic curved-line grating.

Equations (34)

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

T x ( x , y ) = m = + t m exp [ i m ( 2 π Λ x x l x θ ) ] ,
T x ( x , y ) = exp [ i γ x cos ( 2 π Λ x x l x θ ) ] ,
T x ( x , y ) = exp { i γ x sign [ cos ( 2 π Λ x x l x θ ) ] } ,
t m ( γ x ) = { cos ( γ x ) ,   m = 0 , i sin ( γ x ) sinc ( m 2 ) , m 0 ,
t m ( γ x ) = { cos ( γ x ) , m = 0 , 2 i m m π sin ( γ x ) , m 0 and it is odd  , 0 , m 0 and it is even  .
T ( x , y ) = T x ( x , y ) T y ( x , y ) .
T ( x , y ) = exp [ i γ x cos ( 2 π Λ x x l x θ ) + i γ y cos ( 2 π Λ y y l y θ ) ] ,
T ( x , y ) = exp { i γ x sign [ cos ( 2 π Λ x x l x θ ) ] + i γ y sign [ cos ( 2 π Λ y y l y θ ) ] } ,
T ( x , y ) = m = + n = + t m , n ( γ x , γ y ) e i m ( 2 π Λ x x l x θ ) + i n ( 2 π Λ y y l y θ ) ,
| t m , n ( γ x ) | = | t m ( γ x ) t n ( γ x ) | .
t m , n ( γ ) = t m ( γ 2 ) t n ( γ 2 ) ,
t m , n ( γ ) = i m + n J m ( γ / 2 ) J n ( γ / 2 ) .
ψ ( x , y , 0 ) = exp ( x 2 + y 2 w 0 2 ) T ( x , y ) ,
ψ ( x , y , z ) = h + + ψ ( x , y , 0 ) e i α [ x 2 + y 2 2 ( x x + y y ) ] d x d y ,
h = h ( x , y , z ) = exp [ i k z + i α ( x 2 + y 2 ) ] i λ z ,
ψ ( x , y , z ) = h m = + n = + t m , n + + e x 2 + y 2 w 2 2 i α ( x X m + y Y n ) + i l m , n θ d x d y ,
1 w 2 = 1 w 0 2 i π λ z = 1 w 0 2 ( 1 i z 0 z ) ,
X m = x m λ z Λ x ,
Y n = y n λ z Λ y .
ψ ( r , θ , z ) = h m , n = + t m , n 0 2 π 0 e ( r w ) 2 2 i α ρ m , n r cos ( θ φ m , n ) + i l m , n θ r d r d θ ,
e 2 i α ρ m , n r cos ( θ φ m , n ) = s = + i | s | J | s | ( 2 α ρ m , n r ) e i s ( θ φ m , n ) ,
ψ ( r , θ , z ) = h m , n , s = t m , n i | s | e i s φ m , n 0 2 π e i ( l m , n s ) θ d θ 0 e ( r w ) 2 J | s | ( 2 α ρ m , n r ) r d r .
ψ ( r , θ , z ) = 2 π h m , n = + t m , n i | l m , n | e i l m , n φ m , n 0 J | l m , n | ( 2 α ρ m , n r ) e ( r w ) 2 r d r .
0 r e a r 2 J v ( b r ) = π b 8 a 3 2 exp ( b 2 8 a ) [ I ν 1 2 ( b 2 8 a ) I ν + 1 2 ( b 2 8 a ) ] , Re a > 0 , Re ν > 2 ,
ψ ( r , θ , z ) = h w 2 m , n = c m , n e i l m , n φ m , n R m , n e R m , n 2 [ I | l m , n | 1 2 ( R m , n 2 ) I | l m , n | + 1 2 ( R m , n 2 ) ] ,
h = h ( r , z ) = exp [ i ( k z + α r 2 ) ] i λ z .
T ( x , y ) = m , n = + t m , n exp [ 2 π i ( m Λ x x + n Λ y y ) ] ,
p m , n p i = | t m , n | 2 ,
1 p i m = + n = + p m , n = m = + n = + | t m , n | 2 .
p t = m = + n = + p m , n ,
p t p i = m = + n = + | t m , n | 2 .
p m , n p t = | t m , n | 2 m = + n = + | t m , n | 2 ,
m = + n = + | t m , n | 2 = 1 Λ x Λ y Λ x / 2 Λ x / 2 Λ y / 2 Λ y / 2 | T ( x , y ) | 2 d y d x .
m = + n = + | t m , n | 2 = 1.
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.