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

Skew line ray model of nonparaxial Gaussian beam

Open Access Open Access

Abstract

Many ray-optics models have been proposed to describe the propagation of paraxial Gaussian beam. However, those paraxial ray-optics models are inapplicable to the beams that violate the paraxial condition. In this paper, we present a skew line ray (SLR) based model to represent the propagation properties of nonparaxial Gaussian beam under the oblate spheroidal coordinates. The free-space evolution of complex wavefront of the light beam including amplitude and phase is derived via this model. Our analysis demonstrates that the SLR model is available for both nonparaxial and paraxial conditions, and can be used to precisely describe the propagation of complex wavefront. Furthermore, this model changes the transverse density of rays while propagating. The behavior influences the transverse intensity distribution and makes the optical rays become concentrated towards the center. We believe that this ray-optics model can be further developed to describe other kind of structured beams such as Laguerre-Gauss and Bessel-Gauss beams.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Gaussian beams are widely used in optical imaging [1, 2], optical communications [3, 4] and optical manipulations [5, 6]. The physical expression of Gaussian beams is of great significance for theoretically investigating the beams’ propagation or inter-actions between beam and other objects. Since nonparaxial form of Gaussian beams provides a universal way to describe beams’ propagation, it has been studied for years. Lots of approaches have been proposed to describe the propagations of nonparaxial Gaussian beams, such as the perturbation method [7], power-series expansion [8, 9], angular spectrum representation [10, 11] and integral solution of the wave equation [12, 13]. All of them involve complicated mathematical algorithms based on wave optics. Notably, geometrical optics based description of beam propagation provides a convenient and intuitive explanation for many aspects of light beams, such as “self-healing” [14, 15], “self-accelerating” [16, 17] as well as the Gouy and Berry phases (also, geometrical phases). Thus, the geometrical optical field is of continuous interest, adding to the momentum of the rapidly growing field of structured light beams. Meanwhile, a ray-optics model of nonparaxial Gaussian beam offers a rather convenient way to reveal the beam’s propagation behaviors.

Ray-optics models of paraxial Gaussian beam have been studied for years, including the complex ray representation of Gaussian beam [18, 19], the skew line ray (SLR) model of Gaussian beam [20, 21], the Poynting vector model of Laguerre-Gauss beam and Bessel-Gauss beam [22], and the Poincaré sphere ray-optics model of structured Gaussian beams [23]. Notably, the SLR model regards the skew lines (straight generatrixes) of one-sheet hyperboloid as the trajectories of light rays. However,those ray-optics models are based on the paraxial beam and fail when paraxial condition is broken. Paraxial beam is the solution of scalar Helmholtz equation on paraxial approximation. In other words, nonparaxial beam is the exact solutions of scalar Helmholtz equation. Note that the exact solutions of scalar Helmholtz equation have been given in oblate spheroidal coordinates [24–26]. On the paraxial approximation, the zero order of the solutions tends to be a fundamental Gaussian beam. Undoubtedly, this beam can be regarded as a reference of nonparaxial Gaussian beam. Considering that one-sheet hyperboloid is a component of the oblate spheroidal coordinates. In the spirit of SLR model, here we present a SLR model of nonparaxial Gaussian beam. This model is based on skew lines of one-sheet hyperboloid in the oblate spheroidal coordinates.

Since the skew lines are regarded as trajectories of light rays, the geometrical optical field of nonparaxial Gaussian beam, including amplitude and phase, is traveling along these rays. Our mathematical derivations demonstrate that the expression of geometrical optical field, derived by SLR model, is consistent with the result obtained by wave optics for both nonparaxial and paraxial conditions. The free-space evolutions of amplitude and phase of beam are given in terms of rays. Further study shows that the transverse distributions of ray density change with beam propagation in the nonparaxial case, leading to the corresponding changes of intensity distribution in the transverse plane. By contrast, on the paraxial condition, the change of the ray density in the transverse plane is so small that it can be ignored. Therefore, our SLR model enables precise description of the propagation behaviors of nonparaxial light fields without explicit diffraction calculations, and provides new insight into the physics of structured light beams.

The structure of this paper is arranged as follows. In the next section, we give a brief introduction of the nonparaxial Gaussian beam under oblate spheroidal coordinates. In Section 3, the geometrical optical fields of the nonparaxial Gaussian beam are deduced via SLR model. In Section 4, the evolution of SLR model on the paraxial approximation is demonstrated. An explanation of the broken self-similar property on nonparaxial regime is given in terms of rays in Section 5. The relationship between ellipsoid wavefront and rays is revealed in Section 6. The last section includes our conclusions.

2. Gaussian Beam in oblate spheroidal coordinates

The oblate spheroidal coordinates are orthogonally built on confocal oblate spheroids and one-sheet hyperboloids in three-dimensions. The z-axis is defined as the rotation axis, and the distance between foci is set to be 2D. Therefore, the scalar Helmholtz can be separated as the product of the spheroidal radial functions and spheroidal angular functions [24, 26]. The set of coordinates (ξ,η,θ) describes a unique point in Cartesian coordinates (x,y,z) with.

x=D(1+ξ2)(1η2)cosθ,x=D(1+ξ2)(1η2)sinθ,z=Dξη,
where < ξ<+, 0η1, and 0θ2π. Note that all points with a nonzero constant η form a one-sheet hyperboloid, and all points with nonzero constant ξ form an oblate spheroid.

A family of exact solutions of the scalar Helmholtz equation in the oblate spheroidal coordinates have been derived by Landesman and Barrett [24]. The zero-order solution is given by

ψ(ξ,η,θ)=exp[kD(1η)]kDξ2+η2exp(ikDξ)exp(iarctanξη),
where k=2π/λ is the wave number, λ is the wavelength in vacuum. On paraxial approximation, Eq. (2) corresponds to a fundamental Gaussian beam. The relationship between parameter D, beam waist radius ω0 and half divergence angle ε is determined by [24, 26]

D=ω0sinε,ω0=1k2(1+1+k2D2).

As mentioned in [24], this solution is a nonparaxial Gaussian beam. The wave front is an oblate ellipsoid (ξ = nonzero constant). When ξ =0, the beam has a plane wavefront at z=0.

3. SLR model of nonparaxial Gaussian beam

3.1 The skew lines of a hyperboloid

According to Eq. (1), one-sheet hyperboloid with η=ηQ can be determined in Cartesian coordinates by

x2+y2D2(1ηQ2)z2D2ηQ2=1.

As shown in Fig. 1, considering that an arbitrary point Q(xQ,yQ,0) is located in waist plane, the parameter ηQ of the hyperboloid, containing point Q, is derived by Eq. (4) as

 figure: Fig. 1

Fig. 1 Two skew lines of point Q(xQ,yQ,0). |QS1|=| QS2|.

Download Full Size | PDF

ηQ=1xQ2+yQ2D2.

While there are two skew lines cross the same point Q, we define the QS1 (orange line) as the negative skew line lQ of point Q, and QS2 (blue line) as the positive skew line lQ+. These two skew lines are given as

lQ+:x=xQyQτDηQ,y=yQ+xQτDηQ,z=τ, (6-1)
lQ:x=xQ+yQτDηQ,y=yQxQτDηQ,z=τ, (6-2)
τ is a parameter. We assume that the beam propagates in a linear, homogeneous and isotropic medium, so that rays are straight lines. The skew lines can be regarded as the trajectories of rays of Gaussian beam. Both the amplitudes and phases of optical fields travel along those rays.

3.2 Amplitude propagation along the SLR

Since two skew-line rays are associated with point S(x, y, z), these two rays launch from different points in the waist plane. The geometrical optical field at point S is the superposition of the fields carried by these rays. As shown in Fig. 2, assuming that rays lQ1+ and lQ2 pass through the observation point S, the equation of lQ1+ and lQ2 can be derived by Eqs. (6-1) and (6-2). In the following text, the subscript “GO” denotes the optical field derived by geometrical optics. Then the geometrical optical field Ψgo(S) at point S is the sum of fields of the rays associated with point S [15, 27]. It can be expressed as

ψGO(S)=ψGO+(S)+ψGO(S),
where ψGO+(S) and ψGO(S) are the geometrical optical fields carried by rays lQ1+ and lQ2, respectively. Here, we choose the waist plane of z=0 as the initial plane, so the ψGO+(S) and ψGO(S) are given by [27]
ψGO+(S)=12AQ1JQ1JSexp[i(φQ1+φl)], (8-1)
ψGO(S)=12AQ2JQ2JSexp[i(φQ2+φl)]. (8-2)
Where AQ1(AQ2) and φQ1(φQ2) are the amplitude and phase at initial point Q1 (Q2) of ray, respectively. According to the symmetry of fundamental Gaussian beam, AQ1=AQ2, and φQ1=φQ2=0. Phase φl accumulates with the ray’s traveling length l. Variable J denotes the determinant of Jacobian matrix and is determined by J=det[(x,y,z)/(xQ,yQ,τ)] [19, 27]. Note that variables x, y and ηQ are the functions of xQ, yQ, thus J can be derived from Eqs. (5), (6-1) and (6-2) as
J=1+(τDηQn2)2,n=1,2.
The subscript index in Eq. (9), of ηQ denotes the specific point.

 figure: Fig. 2

Fig. 2 Sketch of the rays arriving at point S outside the waist plane.

Download Full Size | PDF

Since there are two skew lines from a same initial point (Fig. 1), both of them play the same role in wave propagation. The weight of field’ amplitude at initial point should be divided equally to two rays. A factor of 1/2 should be multiplied at initial amplitude in Eqs. (8-1) and (8-2).

3.3 Phase propagation along the SLR

Assuming geometrical optical field propagates a tiny length dl along a SLR of QS in Fig. 3, the phase changes dφl. From Eq. (2), we can obtain that

dφldl=darg(U)dl=d[kDξSarctan(ξS/ηS)]dl,
where ξS and ηS are the oblate spheroidal coordinates of point S. Note that the optical path length l (Eikonal) is equal to the distance between points Q and S. From Eqs. (6-1) and (6-2), l=τ/ηQ. According to Eq. (1), z=DξSηS=τ and ηQ=ηS (points S and Q are on the same hyperboloid), then
ξS=lD.
Combining with Eq. (10), we get
φl=klarctanlDηQ.
Equation (12) shows that the phase shift along a ray can be divided into two parts. One is the Berry phase shift of φb=kl. The other is the Gouy phase shift of φg=arctan[l/(DηQ)] which is related with rotation angle of l in terms of the initial point. A geometrical explanation of Gouy phase shift has been given by Landesman [20], as shown in Fig. 3. The edge ray (QS) from point Q has arrived at point S on a wave front ξS. The azimuth angle from point Q to S is α. However, the central ray (OC) arrives at point C at the same time. If the edge ray passes through plane C, it should travel an additional path Δl0. The Δl0 will lead to the azimuth angle difference Δα between S’ and S. It can be find that
α+Δα=arctanlDηQ.
From Eqs. (7) to (13), the geometrical optical field of point S(x, y, z) should be
ψGO(S)=n=121/2AQn1+[z/(DηQn2)]2exp[i(klarctanlDηQn)].
Here, AQn is the amplitude at initial points, and AQ1=AQ2. According to Eq. (2), we get the initial amplitude AQn=exp[kD(1ηQn)]/(kDηQn). If we make the substitution of z=DξSηS, ηQn=ηS and Eq. (11) into Eq. (14), it can be rewritten as
ψGO(S)=exp[kD(1ηS)]kDξS2+ηS2exp[i(kDξSarctanξSηS)].
This equation has the same form as Eq. (2).

 figure: Fig. 3

Fig. 3 Wave front of SLR. The increment Δα corresponds to the segment Δl0.

Download Full Size | PDF

An intriguing phenomenon is that the parameter ηQ in Eq. (5) must have a real root if the equation of skew line is to be determined. That means the initial points of skew line are restricted in a circle area with radius equal to D. In the subsequent of this paper, we call this circle area the “foci disk”. This feature indicates that fields of nonparaxial Gaussian beam in other transverse planes (z0) are traced by rays from the confined foci disk.

4. Paraxial form of SLR model

As shown in Fig. 4, points S(x,y,z) and C1(0,0,z) are in the same transverse plane. Supposing that a ray launching from point Q to S, the optical path length l between points Q and S is |QS|, and can be separated into two parts as

l=|QS|=|QS1|+|S1S|,
where |QS1| is the distance between points Q and S1, |S1S| is the distance between points S1 and S. Point S1 is located on spheroidal wavefront ξS1 (vertex at point C1, blue curve); and point S is located on wavefront ξS2 (vertex at point C2, green curve). If the parameters k, D and η of the nonparaxial Gaussian beam satisfy the following condition [28]
kD>>1,η1,
then the beam propagates on the paraxial approximation. The condition of kD1 means that the size of beam waist is large enough compared with its wavelength. The condition of η1 makes those one-sheet hyperboloids tend to be cylinders which are very close to z-axis.

 figure: Fig. 4

Fig. 4 Sketch of the ray tracing along the skew line.

Download Full Size | PDF

From the conditions of (17), we have divergence angle ε0. Since ε=λ/πω0 [29], the parameter D tends to be

limε0D=limε0ω0sinε=πω02λ=zr,
where zr is the Rayleigh distance. On this approximation, we have |QS1||OC1|. Length |S1S| can be replaced by the distance Δd between the transverse plane of S and spheroidal surface ξS1. For the condition of η1, the spheroidal surface ξS1 can be regard as a spherical surface at vicinity of the vertex C1. The radius of this spherical surface is R(z) at point C1(0, 0, z). Combining Eqs. (1) with (18), we can derive
R(z)=z(1+D2z2)z(1+zr2z2),
then the distance Δd=(x2+y2)/(2R) [29]. Therefore, the optical path length l in Eq. (16) tends to be
lz+x2+y22R(z).
Meanwhile, the skew lines in Eqs. (6-1) and (6-2) have the paraxial form of
lQ+:x=xQyQτzr,y=yQ+xQτzr,z=τ, (21-1)
lQ:x=xQ+yQτzr,y=yQxQτzr,z=τ. (21-2)
Equations (21-1) and (21-2) are consistence with the skew line equation in [2,21]. Then the determinant J of Jacobian is
J=1+(zzr)2.
Since Δα→0 and tanα=|S0Q|/|OQ|, according to Eqs. (13), (21-1) and (21-2), the Gouy phase is given by φg=arctan(z/zr). In this case, the total phase shift φl along a skew line is
φl=k[z+x2+y22R(z)]arctanzzr.
We choose the A0=exp[(xQ2+yQ2)/ω02] for the initial amplitude, the paraxial geometrical optical field at point S(x, y, z) is
ψGO(S)=11+(z/zr)2exp(xQ2+yQ2ω02)×exp{ik[z+x2+y22R(z)]}exp(iarctanzzr).
For paraxial Gaussian beam, ω(z)=ω01+(z/zr)2, there is a relation of ω02/ω2(z)=(xQ2+yQ2)/(x2+y2) according to Eqs. (21-1) and (21-2). Then Eq. (24) can be rewritten as
ψGO(S)=ω0ω(z)exp[x2+y2ω2(z)]×exp{ik[z+x2+y22R(z)]}exp(iarctanzzr).
Obviously, Eq. (25) is the field of the well-known paraxial Gaussian beam [29].

In paraxial condition, the restriction of Eq. (5) vanishes for the radius of foci disk is much larger than beam waist. Although the paraxial SLR model can be enforced to trace the propagation of rays in nonparaxial regime, the initial points of rays might be out of the foci disk. Besides, we note that the optical path length calculated by Eq. (20) is an approximate treatment in paraxial condition. In practical ray tracing methodology, the optical path length is always rigorously equal to the distance between the initial and terminal points of a ray. In order to distinguish these two models, we define that the ray trajectories traced by Eqs. (6-1) and (6-2) is nonparaxial SLR model and by Eqs. (21-1) and (21-2) are paraxial SLR model, respectively.

5. Ray-based explanation on broken self-similar property

In previous works on nonparaxial Gaussian beam, the beam profiles are quite different from those of the conventional transverse pattern of paraxial beam [12, 13]. This phenomenon indicates that self-similar property of beam is broken in nonparaxial region. However, none explanation is given. Combining with ray tracing methodology, here we give an explanation in terms of rays. In the following simulations, we choose λ=1.064 μm and take kD=100 (ω0,12.38 μm) for a paraxial beam. In this case, the half divergence angle ε8° which ensures the parameter η1. By contrast, we choose kD=1 (ω0,20.15 μm) for a nonparaxial beam where ε65.5°. The optical fields are observed in the plane of z=5zr.

Trajectories of rays of the nonparaxial SLR model are presented in Fig. 5. Those rays are traced by Eqs. (6-1) and (6-2) and the initial points of rays are arranged in equidistant concentric circles. The coordinates are nondimensionalized by 1/ω0 for x-/y-coordinates and by 1/zr for z-coordinate. In the case of kD=100, the outer envelope is one-sheet hyperboloids as shown in Fig. 5(a). Trajectories of rays of paraxial SLR model for both kD=100 and kD=1 (not shown) are consistent with those in Fig. 5(a). For kD=1, as shown in Fig. 5(b), the outer envelope is similar to that in Fig. 5(a) because of the nondimensionalization of coordinates. However, the distribution of rays’ tips is more concentrated in central zone than that in Fig. 5(a).

 figure: Fig. 5

Fig. 5 Rays’ trajectories of nonparaxial SLR model with different parameter. (a) kD=100; (b) kD=1.

Download Full Size | PDF

For detailed discussions, we set the rays with uniform distribution in the initial plane (z=0), and trace the rays’ distribution in an observation plane as shown in Fig. 6. Gray images denote the ray’s distribution pattern. The gray value at each pixel represents the ray’s density of that area. Both x and y axes are nondimensionalized by  1/ω0. Figure 6(a) shows the distribution of rays in the initial plane (z=0). The rays are uniformly distributed within the circle radius of ω0. Figures 6(b)-6(e) are distributions of rays in the observation plane (z=5zr). For the beam with kD=100 in Figs. 6(b) and 6(c), ray distributions are traced by paraxial and nonparaxial SLR model, respectively. In this case, the gray values are still uniformly distributed in a circle with radius approaches to 5ω0. When kD=100, the rays’ distribution after propagation is similar to the initial distribution of rays. This property ensures the self-similar feature with beam propagation. Hence, the intensity distributions are modulated by rays’ distributions. In this condition, the intensity patterns traced by the two models should be similar.

 figure: Fig. 6

Fig. 6 Ray’s distribution patterns in an observation plane (z=5zr). (a) Initial plane z=0; (b) and (d) are traced by paraxial SLR model for kD=100 and kD=1, respectively. (c) and (e) are traced by nonparaxial SLR model for kD=100 and kD=1, respectively.

Download Full Size | PDF

For kD=1, rays’ distributions in Figs. 6(d) and 6(e) are traced by paraxial and nonparaxial SLR models, respectively. Although kD=1 breaks the paraxial condition, Fig. 6(d) has little difference with Fig. 6(b) which comes from the paraxial SLR model. Figure 6(e) has a bright flare at the center and gradually dim toward outside. There is no obvious edge of the distribution. Compared with the initial rays’ distribution in Fig. 6(a), the distribution is obviously distorted after beam’s propagation. The self-similar feature of beam is therefore broken. The intensity pattern traced by nonparaxial SLR model should be more concentrated than that of paraxial SLR model. Since self-similar property requires that the transverse distribution of rays should also propagate in a self-similar way [23]. The change of distribution of rays’ tips in Fig. 6(e) indicates that self-similar property is broken in nonparaxial case.

To compare intensity patterns traced by two models, we ought to choose the same distribution of amplitude in the initial plane. Note that Eq. (2) has amplitude singularity in z=0 plane which can cause intensity distortion in other planes after ray tracing. Although the modified amplitude has been proposed in [25], it is a superposition of an outgoing wave and an incoming wave, and it has infinite energy. For practical purposes, we use the Gaussian distribution A0=exp(r2/ω02) as initial amplitude in the waist plane, where r is the distance from the initial point of ray to the origin.

In the observation plane, the geometrical optical fields have been traced with nonparaxial [Eqs. (6-1) and (6-2)] and paraxial [Eqs. (21-1) and (21-2)] SLR models in Fig. 7. The intensity distributions I1 and I2 are obtained via ray tracing, while I0,1 and I0,2 that are calculated by Eq. (2) and Eq. (25), respectively. When kD=100 in Fig. 7(a), I1 is very closed to I2. Both of them are fitted with I0,1 and I0,2. If we take kD=1 [Fig. 5(b)], I1 is narrower than I0,1 due to the difference of the initial amplitudes between Eq. (2) and Gaussian beam. The curve I2 is fitted with I0,2. The curve I1 is obviously more concentrated in central zone than curve I2, and the peak of I1 is also higher than the peak of I2. These appearances are consistent with discussions in paragraph above. While the paraxial condition is broken, the self-similar feature is vanished. The paraxial SLR model has a great deviation than nonparaxial SLR model, and makes the intensity distribution broaden and weaken.

 figure: Fig. 7

Fig. 7 Comparison of the intensity distribution in the plane of z=5zr between nonparaxial and paraxial SLR models. (a) kD=100; (b) kD=1.

Download Full Size | PDF

6. Ellipsoid wavefront as the envelope of equidistance rays

Considering that the wavefront can be regarded as an envelope of rays’ tips with equal length. We plot the rays’ trajectories for kD=100 and kD=1, where the lengths of rays equal to λ. The trajectories of rays in each case are separately traced by paraxial and nonparaxial SLR models, as shown in Fig. 8. The transverse coordinate is nondimensionalized by ωλ=ω01+(λ/zr)2. When kD=100, the envelope of rays’ tips overlays on the ellipsoid with parameter ξ=λ/D (the black dotted line). As shown in the Figs. 8(a) and 8(b), the ellipsoid can be regarded as a flat for its curvature radius is large enough compared with λ. The wavefront tends to be a plane wave. For the case of kD=1, the envelope of tips, far from xoy plane, is still an ellipsoid as shown in Figs. 8(c) and 8(d). The curvature radius of the ellipsoid becomes large and the dotted line is obviously curved. However, in Fig. 8(c), the envelope near xoy plane appears to be distorted away from the origin. This would lead to a serious distortion on the wavefront. Since the paraxial SLR model is enforced in nonparaxial condition, the restriction that ray’s initial point must be located in foci disk of Eq. (5) disappears. The initial points of rays far away from z-axis are out of the foci disk, which leads to the distorted envelope.

 figure: Fig. 8

Fig. 8 Rays’ trajectories with length equal to λ. (a) and (c) are traced by paraxial SLR model for kD=100 and kD=1. (b) and (d) are traced by nonparaxial SLR model for kD=100 and kD=1, respectively.

Download Full Size | PDF

The phase patterns traced by two SLR models with kD=1 are shown in Figs. 9(a) and 9(b). In Fig. 9(a), the shape of contour of isophase equal to π is consistent with the shape of rays’ envelope in Fig. 8(a). In Fig. 9(b), the contour is an ellipsoid. The phase patterns calculated directly by Eqs. (25) and (2) are also shown in Figs. 9(c) and 9(d), respectively. Since Eq. (25) is not suit for kD=1 condition, the pattern in Fig. 9(c) is obviously different from that in Figs. 9(b) or 9(d). The pattern in Fig. 9(b) is consistent with that of in Fig. 9(d), proving that nonparaxial SLR model can trace the evolution of phase precisely. We noticed that the phase pattern in Fig. 9(c) varies rapidly along x-axis with periodicity near z=0 plane. This is mainly induced by the approximation of Eq. (20). This treatment is the paraxial approximation of optical path length. That leads to an increased optical path length.

 figure: Fig. 9

Fig. 9 Axial patterns of phase. kD=1. (a) and (b) traced by paraxial and nonparaxial SLR model. (c) and (d) are calculated by Eq. (25) and Eq. (2) respectively.

Download Full Size | PDF

7. Concluding remarks

In this paper, we have presented a SLR model for geometrical description of nonparaxial Gaussian beams. This model enables us to precisely describe the propagation of nonparaxial Gaussian beam including the amplitude and phase. The optical field deduced by nonparaxial SLR model is consistent with that obtained from the wave equation. The explanations of both Berry phase and Gouy phase are given in term of ray optics. We demonstrate the rays’ trajectories of nonparaxial model in both paraxial and nonparaxial regimes. The simulation results show that our SLR model is suitable for both paraxial and nonparaxial cases. Moreover, on the paraxial approximation, the equations of skew line are consistent with the previous works.

Notably, this model depicts a ray by the initial point and the direction vector of ray. That makes it easy to combine with the ray-tracing methodology. Such a geometrical description provides a useful way to study beam’s behaviors after reflection/refraction for nonparaxial beam, or through the ABCD matrix on paraxial condition. Furthermore, apart from the Gaussian beam we focus on here, the SLR model would be further exploited for studying other kind of nonparaxial light beams, such as Laguerre-Gauss beam and Bessel-Gauss beam. Generally, positive and negative skew lines play the same role in presenting beam’ propagation. Considering that the SLR model can work as a ray-optics model for force calculation in optical trapping, it is necessary to choose both of positive and negative skew lines at the same times to counteract the artificial force torque.

Funding

The Key Project of Natural Science Foundation of the Anhui Higher Education Institutions (KJ2016A361); the Grants for Scientific Research of BSKY from Anhui Medical University (XJ201518); Scientific Research Foundation of the Institute for Translational Medicine of Anhui Province(2017zhyx25).

References and links

1. L. Novotny, E. J. Sanchez, and X. S. Xie, “Near-field optical imaging using metal tips illuminated by higher-order Hermite—Gaussian beams,” Ultramicroscopy 71(1-4), 21–29 (1998). [CrossRef]  

2. A. Tycho, T. M. Jørgensen, H. T. Yura, and P. E. Andersen, “Derivation of a Monte Carlo method for modeling heterodyne detection in optical coherence tomography systems,” Appl. Opt. 41(31), 6676–6691 (2002). [CrossRef]   [PubMed]  

3. J. C. Ricklin and F. M. Davidson, “Atmospheric turbulence effects on a partially coherent Gaussian beam: implications for free-space laser communication,” J. Opt. Soc. Am. A 19(9), 1794–1802 (2002). [CrossRef]   [PubMed]  

4. M. Toyoshima, T. Jono, K. Nakagawa, and A. Yamamoto, “Optimum divergence angle of a Gaussian beam wave in the presence of random jitter in free-space laser communication systems,” J. Opt. Soc. Am. A 19(3), 567–571 (2002). [CrossRef]   [PubMed]  

5. A. Ashkin, “Forces of a Single-Beam Gradient Laser Trap on a Dielectric Sphere in the Ray Optics Regime,” Biophys. J. 61(2), 569–582 (1992). [CrossRef]   [PubMed]  

6. D. J. Stevenson, F. Gunn-Moore, and K. Dholakia, “Light forces the pace: optical manipulation for biophotonics,” J. Biomed. Opt. 15(4), 041503 (2010). [CrossRef]   [PubMed]  

7. M. Lax, W. H. Louisell, and W. B. McKnight, “From Maxwell to paraxial wave optics,” Phys. Rev. A 11(4), 1365–1370 (1975). [CrossRef]  

8. T. Takenaka, M. Yokota, and O. Fukumitsu, “Propagation of light beams beyond the paraxial approximation,” J. Opt. Soc. Am. A 2(6), 826–829 (1985). [CrossRef]  

9. H. Laabs, “Propagation of Hermite-Gaussian-beams beyond the paraxial approximation,” Opt. Commun. 147(1-3), 1–4 (1998). [CrossRef]  

10. G. P. Agrawal and D. N. Pattanayak, “Gaussian beam propagation beyond the paraxial approximation,” J. Opt. Soc. Am. 69(4), 575–578 (1979). [CrossRef]  

11. C. G. Chen, P. T. Konkola, J. Ferrera, R. K. Heilmann, and M. L. Schattenburg, “Analyses of vector Gaussian beam propagation and the validity of paraxial and spherical approximations,” J. Opt. Soc. Am. A 19(2), 404–412 (2002). [CrossRef]   [PubMed]  

12. K. Duan, B. Wang, and B. Lü, “Propagation of Hermite-Gaussian and Laguerre-Gaussian beams beyond the paraxial approximation,” J. Opt. Soc. Am. A 22(9), 1976–1980 (2005). [CrossRef]   [PubMed]  

13. B. Z. Wang, Z. G. Zhao, B. D. Lü, and K. L. Duan, “Vectorial Hermite–Laguerre–Gaussian beams beyond the paraxial approximation,” Chin. Phys. 16(1), 143–147 (2007). [CrossRef]  

14. J. Broky, G. A. Siviloglou, A. Dogariu, and D. N. Christodoulides, “Self-healing properties of optical Airy beams,” Opt. Express 16(17), 12880–12891 (2008). [CrossRef]   [PubMed]  

15. Y. Kaganovsky and E. Heyman, “Wave analysis of Airy beams,” Opt. Express 18(8), 8440–8452 (2010). [CrossRef]   [PubMed]  

16. M. A. Alonso and M. A. Bandres, “Generation of nonparaxial accelerating fields through mirrors. I: two dimensions,” Opt. Express 22(6), 7124–7132 (2014). [CrossRef]   [PubMed]  

17. M. A. Alonso and M. A. Bandres, “Generation of nonparaxial accelerating fields through mirrors. II: Three dimensions,” Opt. Express 22(12), 14738–14749 (2014). [CrossRef]   [PubMed]  

18. J. Arnaud, “Representation of Gaussian beams by complex rays,” Appl. Opt. 24(4), 538–543 (1985). [CrossRef]   [PubMed]  

19. Y. A. Kravtsov, G. W. Frobes, and A. A. Asatryan, “Theory and applications of complex rays,” Prog. Opt. 39, 1–67 (1999). [CrossRef]  

20. B. T. Landesman, “Geometrical representation of the fundamental mode of a Gaussian beam in oblate spheroidal coordinates,” J. Opt. Soc. Am. A 6(1), 5–17 (1989). [CrossRef]  

21. B. Shi, Z. Meng, L. Wang, and T. Liu, “Monte Carlo modeling of human tooth optical coherence tomography imaging,” J. Opt. 15(7), 075304 (2013). [CrossRef]  

22. M. V. Berry and K. T. McDonald, “Exact and geometrical optics energy trajectories in twisted beams,” J. Opt. A, Pure Appl. Opt. 10(3), 035005 (2008). [CrossRef]  

23. M. A. Alonso and M. R. Dennis, “Ray-optical Poincaré sphere for structured Gaussian beams,” Optica 4(4), 476–486 (2017). [CrossRef]  

24. B. T. Landesman and H. H. Barrett, “Gaussian amplitude functions that are exact solutions to the scalar Helmholtz equation,” J. Opt. Soc. Am. A 5(10), 1610–1619 (1988). [CrossRef]  

25. Z. Ulanowski and I. K. Ludlow, “Scalar field of nonparaxial Gaussian beams,” Opt. Lett. 25(24), 1792–1794 (2000). [CrossRef]   [PubMed]  

26. G. Rodríguez-Morales and S. Chávez-Cerda, “Exact nonparaxial beams of the scalar Helmholtz equation,” Opt. Lett. 29(5), 430–432 (2004). [CrossRef]   [PubMed]  

27. D. A. McNamara, C. W. I. Pistorius, and J. A. G. Malherbe, Introduction to the Uniform Geometrical Theory of Diffraction (ARTECH HOUSE, 1990), Chap. 2.

28. L. A. Vainshtein, “Open resonators with spherical mirrors,” Sov. Phys. JETP 18, 471–479 (1964).

29. M. Dietrich, Light Transmission Optics (Van Nostrand Reinhold, 1972), Chap. 6.

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 Two skew lines of point Q( x Q , y Q ,0 ). |Q S 1 |=| Q S 2 |.
Fig. 2
Fig. 2 Sketch of the rays arriving at point S outside the waist plane.
Fig. 3
Fig. 3 Wave front of SLR. The increment Δα corresponds to the segment Δ l 0 .
Fig. 4
Fig. 4 Sketch of the ray tracing along the skew line.
Fig. 5
Fig. 5 Rays’ trajectories of nonparaxial SLR model with different parameter. (a) kD=100; (b) kD=1.
Fig. 6
Fig. 6 Ray’s distribution patterns in an observation plane ( z=5 z r ). (a) Initial plane z=0; (b) and (d) are traced by paraxial SLR model for kD=100 and kD=1, respectively. (c) and (e) are traced by nonparaxial SLR model for kD=100 and kD=1, respectively.
Fig. 7
Fig. 7 Comparison of the intensity distribution in the plane of z=5 z r between nonparaxial and paraxial SLR models. (a) kD=100; (b) kD=1.
Fig. 8
Fig. 8 Rays’ trajectories with length equal to λ. (a) and (c) are traced by paraxial SLR model for kD=100 and kD=1. (b) and (d) are traced by nonparaxial SLR model for kD=100 and kD=1, respectively.
Fig. 9
Fig. 9 Axial patterns of phase. kD=1. (a) and (b) traced by paraxial and nonparaxial SLR model. (c) and (d) are calculated by Eq. (25) and Eq. (2) respectively.

Equations (28)

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

x=D ( 1+ ξ 2 )( 1 η 2 ) cosθ, x=D ( 1+ ξ 2 )( 1 η 2 ) sinθ, z=Dξη,
ψ( ξ,η,θ )= exp[ kD( 1η ) ] kD ξ 2 + η 2 exp( ikDξ )exp( iarctan ξ η ),
D= ω 0 sinε , ω 0 = 1 k 2( 1+ 1+ k 2 D 2 ) .
x 2 + y 2 D 2 ( 1 η Q 2 ) z 2 D 2 η Q 2 =1.
η Q = 1 x Q 2 + y Q 2 D 2 .
l Q + : x= x Q y Q τ D η Q , y= y Q + x Q τ D η Q , z=τ,
l Q : x= x Q + y Q τ D η Q , y= y Q x Q τ D η Q , z=τ,
ψ GO ( S )= ψ GO + ( S )+ ψ GO ( S ),
ψ GO + ( S )= 1 2 A Q 1 J Q 1 J S exp[ i( φ Q 1 + φ l ) ],
ψ GO ( S )= 1 2 A Q 2 J Q 2 J S exp[ i( φ Q 2 + φ l ) ].
J=1+ ( τ D η Q n 2 ) 2 , n=1, 2.
d φ l dl = darg( U ) dl = d [ kD ξ S arctan( ξ S / η S ) ] dl ,
ξ S = l D .
φ l =klarctan l D η Q .
α+Δα=arctan l D η Q .
ψ GO ( S )= n=1 2 1/2 A Q n 1+ [ z/ ( D η Q n 2 ) ] 2 exp[ i( klarctan l D η Q n ) ].
ψ GO ( S )= exp[ kD( 1 η S ) ] kD ξ S 2 + η S 2 exp[ i( kD ξ S arctan ξ S η S ) ].
l=| QS |=| Q S 1 |+| S 1 S |,
kD>>1, η1,
lim ε0 D= lim ε0 ω 0 sinε = π ω 0 2 λ = z r ,
R( z )=z( 1+ D 2 z 2 )z( 1+ z r 2 z 2 ),
lz+ x 2 + y 2 2R( z ) .
l Q + : x= x Q y Q τ z r , y= y Q + x Q τ z r , z=τ,
l Q : x= x Q + y Q τ z r , y= y Q x Q τ z r , z=τ.
J=1+ ( z z r ) 2 .
φ l =k[ z+ x 2 + y 2 2R( z ) ]arctan z z r .
ψ GO ( S )= 1 1+ ( z/ z r ) 2 exp( x Q 2 + y Q 2 ω 0 2 ) ×exp{ ik[ z+ x 2 + y 2 2R( z ) ] }exp( iarctan z z r ).
ψ GO ( S )= ω 0 ω( z ) exp[ x 2 + y 2 ω 2 ( z ) ] ×exp{ ik[ z+ x 2 + y 2 2R( z ) ] }exp( iarctan z z r ).
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.