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

Propagation of multi-cosine-Laguerre-Gaussian correlated Schell-model beams in free space and atmospheric turbulence

Open Access Open Access

Abstract

We introduce a class of random stationary, scalar source named as multi-cosine-Laguerre-Gaussian-correlated Schell-model (McLGCSM) source whose spectral degree of coherence (SDOC) is a combination of the Laguerre-Gaussian correlated Schell-model (LGCSM) and multi-cosine-Gaussian correlated Schell-model (McGCSM) sources. The analytical expressions for the spectral density function and the propagation factor of a McLGCSM beam propagating in turbulent atmosphere are derived. The statistical properties, such as the spectral intensity and the propagation factor, of a McLGCSM beam are illustrated numerically. It is shown that a McLGCSM beam exhibits a robust ring-shaped beam array with adjustable number and positions in the far field by directly modulating the spatial structure of its SDOC in the source plane. Moreover, we provide a detailed insight into the theoretical origin and characteristics of such a ring-shaped beam array. It is demonstrated that these peculiar shaping properties are the concentrated manifestation of the individual merits respectively associated with the Laguerre- and multi-cosine-related factors of the whole SDOC. Our results provide a novel scheme to generate robust and controllable ring-shaped beam arrays over large distances, and will widen the potentials for manipulation of multiple particles, free-space optical communications and imaging in the atmosphere.

© 2017 Optical Society of America

1. Introduction

As is well known, for a partially coherent beam, the properties of the spectral degree of coherence (SDOC) or spatial correlation function significantly affect its propagation properties, such as the intensity distribution profiles and the propagation factors [1]. Due to some constraints in the choice of the mathematical form of the SDOCs for partially coherent optical fields, earlier work mainly concentrated on the partially coherent beams with the conventional SDOCs being Gaussian-correlated Schell-model functions, but studies of partially coherent beams with non-conventional SDOCs were quite few in number and confined to only the Bessel-correlated Schell-model beams [2–4]. In order to fulfil the physical limits of the SDOCs, Gori and Santarsiero generally established the sufficient condition for devising a genuine SDOC of a scalar or electromagnetic non-conventional partially coherent beam based on the theory of reproducing kernel Hilbert spaces [5]. Since then plenty of partially coherent beams with non-conventional SDOCs have been explored, including circular or rectangular Laguerre-Gaussian correlated Schell-model (LGCSM) beam [4, 6–16], cos-Gaussian correlated Schell-model (cGCSM) beam [17–24], multi-cos-Gaussian correlated Schell-model (McGCSM) beam [25–31], and so on [32–40]. Theoretical analyses have demonstrated that these novel beams exhibit extraordinary propagation characteristics. For example, a LGCSM beam with circular symmetry displays a single ring-shaped beam profile in the far field or a controllable optical cage near the focal plane and a McGCSM beam displays robust optical lattice patterns in the far field, although both beams are Gaussian beam profiles in the source plane. Experiments have been performed that have verified these theoretical predications [7–10, 15, 19, 23, 32, 33, 35, 38–40]. Very recently, we introduced a novel kind of partially coherent beam whose complicated SDOC contains both non-conventional cosine- and Hermite-type components [41]. Such beams have been found to exhibit cascade self-splitting properties for two successive times during propagation in free space. Additionally, the propagation of partially coherent beams with conventional or non-conventional SDOCs in turbulent atmosphere has also been studied in detail. It is generally believed that modulating the initial coherence properties is perhaps one of the most efficient way for reducing the negative influence of the turbulence [4, 11, 14, 15, 18, 20–22, 24, 26, 27, 33, 36, 42–45].

On the other hand, ring-shaped beams (also known as dark hollow beams (DHBs)) and optical cages have attracted considerable attentions recently due to their potential applications in optical trapping [46–49], optical imaging [50, 51], and optical cloaking [52, 53]. Partially coherent DHBs with conventional or non-conventional SDOCs have also been introduced and some of them have been experimentally realized. As mentioned above, a ring-shaped beam profile or an optical cage can be arrived in the far field (on near the focal plane) by employing a circular LGCSM beam [8, 46–53]. However, note that most of the existing technologies are limited to generate only one single ring-shaped beam or optical cage, which is not sufficient for meeting the engineering requirement. For instance, it was found that for the optical imaging in a stimulated emission depletion (STED) microscopy, a regular single spot STED setup has only a quarter of scanning speed compared to the parallelized STED setup in which four ring-shaped spots are combined to reduce the recording time [54]. Finding an efficient way to generate robust multiple ring-shaped beams or optical cages is therefore highly desirable. Very recently, using a full polarization-controlled method the optical cage array with tunable number and positions was first obtained by focusing a so-called Dammann vector beam [55]. In this paper, we propose a new kind of non-conventional partially coherent beam named as multi-cosine-Laguerre-Gaussian correlated Schell-model (McLGCSM) beam whose non-conventional SDOC is a Gaussian function modulated by both Laguerre- and multi-cosine-related factors. By adjusting the initial coherence states such beams are able to naturally create a robust ring-shaped beam array with prescribed number and tunable positions in the far field during propagation.

The paper is organized as follows: in section 2, we introduce the theoretical model for a McLGCSM beam and determine its genuine SDOC function. In section 3, the analytical expression for the cross-spectral density (CSD) function of a McLGCSM beam propagating in atmospheric turbulence is derived based on the extended Huygens-Fresnel integral. In section 4, we numerically study the evolution properties of the spectral density of a McLGCSM beam and compare the results with a McGCSM beam and LGCSM beam. While the propagation factor of the McLGCSM beam is discussed in section 5, some extraordinary and useful results are summarized in section 6.

2. Model for multi-cosine-Laguerre-Gaussian-correlated Schell-model (McLGCSM) source

Let W(0)(r1,r2,ω) denote the CSD at two spatial positions r1=(x1,y1) and r2=(x2,y2) which are two arbitrary transverse position vectors. For brevity, the dependence of W(0) on frequency ω is not explicitly shown in the following equations. The CSD must have a non-negative definite kernel, meaning that, for any choice of f (r′), the following inequality must be established [5]:

d2r1d2r2W(0)(r1,r2)f(r1)f(r2)0.
To satisfy the constraint given by inequality (1) a sufficiency condition is that the CSD has to be expressed as a superposition integral of the form [5]
W(0)(r1,r2)=d2vp(v)H0*(r1,v)H0(r2,v),
where H0 is an arbitrary kernel and p(v) is a non-negative function.

If the kernel H0 is chosen as a Fourier-like structure, i.e.,

H0(r,v)=τ(r)exp(irv),
where τ(r) is a generally complex profile function, then Eq. (2) becomes
W(0)(r1,r2)=τ*(r1)τ(r2)d2vp(v)exp[iv(r1r2)]=τ*(r1)τ(r2)μ(r1r2)
where μ(r1r2) represents the two-dimensional Fourier transform of p(v) and characterizes the SDOC.

Depending on the choice of p(v), a family of sources with different SDOCs are determined. In this paper we use the following expression for the non-negative function p(v)

p(vx,vx)=N4lx=MxMxly=MyMy[pL(vx+lxβxδ,vy+lyβyδ)+pL(vx+lxβxδ,vylyβyδ)+pL(vxlxβxδ,vy+lyβyδ)+pL(vxlxβxδ,vylyβyδ)],
with
pL(vx,vy)=(vx2+vy2δ2)nexp[δ2(vx2+vy2)2],
where βx, βy and δ are real constants, n is a positive integer, Mx = (Px−1)/2 and My = (Py−1)/2 with Px and Py being positive integers, and N=1/max[μ(r1r2)] to make the maximum values of μ(r1r2) being 1. In fact, pL (vx, vy) is just the function associated with a LGCSM beam defined in [6–8].

Let us assume τ(r)=exp(r24w02) to be a Gaussian with w0 being the beam width, after substituting Eqs. (5) and (6) into Eq. (4), and making use of the shift theorem of the Fourier transformation [56]

f(zz0)exp(ixz)dz=exp(iz0x)f(y)exp(ixy)dy,
we obtain the following CSD
W(0)(r1,r2)=exp(rs22w02rd28w02)μ(rd),
where
μ(rd)=NLn(rd2δ2)exp(rd22δ2)lx=MxMxly=MyMycos(lxβxxdδ)cos(lyβyydδ),
with rs=(r1+r2)/2, rd=r1r2 being the “sum” and “difference” coordinates in the source plane, respectively, and Ln() denoting the Laguerre polynomial of mode order n. Such a random source with CSD function determined by Eqs. (8) and (9) is termed as a McLGCSM source and represents a more general source model. It can reduce to a circular LGCSM source (for βx = βy = 0), a rectangular McGCSM source (for n = 0) and a well-known Gaussian-Schell model (GSM) source (for βx = βy = n = 0). As was proved, a circular LGCSM beam displays a ring-shaped beam profile [6–8] and a rectangular McGCSM beam exhibits a self-splitting beam spot array [25–27] in the far field due to their non-conventional correlation functions.

Figure 1 comparatively shows the density plot of the square of the modulus of the SDOC for a McGCSM source, LGCSM source, and McLGCSM source. It is interesting to note that the SDOC of the McLGCSM source significantly differs from that of the counterpart LGCSM source or McGCSM source. Although all of these beams have the same Gaussian intensity distributions in the source plane, the variations in the source structured coherence state lead to extraordinary propagation properties of the McLGCSM beam as shown later.

 figure: Fig. 1

Fig. 1 Density plots of the square of the modulus of the SDOC for (a) a McGCSM source with n = 0 and βx = βy = 8, (b) a LGCSM source with n = 5 and βx = βy = 0, and (c) a McLGCSM source with n = 5 and βx = βy = 8. The other parameters are Px = Py = 3.

Download Full Size | PDF

In fact, the McLGCSM source with the SDOC given by Eqs. (8) and (9) can be considered as a superposition of PxPy LGCSM sources [7, 8] respectively with tilting factors exp[ ± i(lxβxx ± lyβyy)/δ]. Hence, it is straight that the required condition for such a source representing a beam is 1δ2+14w02<<2π2λ2, that is, just the same as that of a LGCSM source [6–8] or a cGCSM source [17, 20, 21].

3. Propagation of a McLGCSM beam in atmospheric turbulence

Based on the extended Huygens–Fresnel principle, paraxial propagation of the CSD of a McLGCSM beam propagating in turbulent atmosphere can be written as [11, 14, 20–22, 26, 33, 36, 42–45]

W(r1,r2,z)=1λ2z2exp[ik2z(r22r12)]d2r1d2r2W(0)(r1,r2)×exp[ik2z(r12r22)+ikz(r1r1r2r2)]exp[Φ(r1,r1)+Φ*(r2,r2)],
where r1 = (x1, y1) and r2 = (x2, y2) are two arbitrary transverse position vectors in the output plane, z is the propagation distance, and k = 2π/λ is the wave number with λ being the wavelength. The asterisk (*) denotes the complex conjugation and Φ(ri,ri)(i=1,2) represents the random part of the complex phase of a spherical wave due to the atmospheric turbulence. The last term in the integrand of Eq. (10) can be expressed as [11, 26, 33, 42–45]
exp[Φ(r1,r1)+Φ*(r2,r2)]=exp[Tα(rd2+rdrd+rd2)],
where rs = (r1 + r2)/2 and rd = r1r2 are the “sum” and “difference” coordinates in the receiving plane, respectively, and the symbol Tα reflects the intensity of the turbulence. In this paper, the Kolmogorov spectrum and a quadratic approximation of the 5/3 power law for the Rytov’s phase structure function are employed, which are accepted to be valid not only for weak fluctuations but also for strong ones. Then Tα=ρ02 where ρ0=(0.545Cn2k2z)3/5 denotes the spatial coherence radius of a spherical wave propagating in a turbulent atmosphere with Cn2 being the refraction index structure constant and describing the strength of the atmospheric turbulence [11, 26, 33, 42–45]. Therefore, Eq. (10) can be rewritten in the following form
W(r1,r2,z)=Nλ2z2exp(ikzrrdTαrd2)lx=1Mxly=1Myd2rsd2rdLn(rd22δ2)×cos(Qlxxdδ)cos(Qlyydδ)exp[rs22w02(1δ2+14w02+2Tα)rd22]×exp[ikzrsrd+ikz(rdrs+rrd)Tαrdrd].
When Τα→0, Eq. (12) represents the propagation of the beam within free space.

Completing integration over rs, we obtain

W(r1,r2,z)=2πw02Nλ2z2exp[2izRzw02rrd(2zR2z2w02+Tα)rd2]×lx=MxMxly=MyMy[Wl,++(r1,r2)+Wl,+(r1,r2)+Wl,+(r1,r2)+Wl,(r1,r2)],
where
Wl,uv(r1,r2)=14d2rdLn(rd22δ2)exp[Ωrd28w02+ikz(blx,xuxd+bly,yvyd)],
Ω=1+4δw2+8w02Tα+16zR2z2,
bq±=q+i(Tαzk2zRz)qd±zw02zRδwlqβq,(q=x,yu,v=±),
with δw=δ/w0 and zR=πw02/λ being the Rayleigh distance.

Using the following formulae [56]

Ln(x)=p=0n(np)(1)pp!xp
(x2+y2)p=m=0p(pm)x2(pm)y2m,
znexp[(zβ)2]dz=(2i)nπHn(iβ),
and performing tedious mathematical operations, Eq. (14) turns to be
Wl,uv(r1,r2)=2πw02Ωexp[2k2w02(blx,xu2+bly,yv2)Ωz2]p=0n(np)1p!(1Ωδw2)p×m=0p(pm)H2(pm)(2Ωkw0zblx,xu)H2m(2Ωkw0zbly,yv),
where Hn denotes the Hermite polynomial of mode order n. Applying the following formulae [56–58]
m=0n(nm)H2m(x)H2n2m(y)=(4)nn!Ln(x2+y2),
Ln(bx)=(1b)nm=0n(b1b)m(nm)Lm(x),
Equation (20) can be further simplified to
Wl,uv(r1,r2)=2πw02Ωexp[2k2w02(blx,xu2+bly,yv2)Ωz2]m=0n(nm)(4Ωδw2)mLm[2k2w02(blx,xu2+bly,yv2)Ωz2]=2πw02Ω(Ωδw2Ωδw24)nexp[2k2w02(blx,xu2+bly,yv2)Ωz2]Lm[8k2w02(blx,xu2+bly,yv2)Ωz2(4Ωδw2)].
Equations (13) and (23) represent the CSD of a McLGCSM beam in the output plane. Letting ρ1 = ρ2 = ρ, we obtain the following explicit expressions for the average intensity of a McLGCSM beam propagating in turbulent atmosphere
I(r,z)=W(r,r,z)=πNw02z2zR2lx=1Mxly=1My[Il,++(r,z)+Il,+(r,z)+Il,+(r,z)+Il,(r,z)],
where
Il,±±(r,z)=(Ωδw2Ωδw24)nexp(8zR2Ωz2Rlx,ly)Ln[32zR2δw2z2(4Ωδw2)Rlx,ly]
with Rlx,ly=(xw±zlxβx2zRδw)2+(yw±zlyβy2zRδw)2 and (xw, yw) = (x, y)/w0 being the scaled transversal coordinates.

Obviously, Eqs. (24) and (25) will go back to the average intensity of a LGCSM beam for βx = βy = 0 [6–8] or of a McGCSM beam for n = 0 [25–27]. Moreover, we point out that if βx ≠ 0 or βy ≠ 0, the four terms in the summation of Eq. (24) represent the average intensity patterns of LGCSM beams respectively centered at z2zRδw(±lxβx,±lyβy), implying that the McLGCSM beam can behave self-splitting behavior when the parameters βx and βy are sufficiently large. Therefore, the results clearly reveal that for the proposed McLGCSM beam on propagation, modulating the multi-cosine factor of the total correlation function can induce the self-splitting behavior, and adjusting the Laguerre factor of the total correlation function can give rise to the formation of the ring-shaped beam profile.

The SDOC of a McLGCSM beam in the output plane is obtained as

μ(r1,r2,z)=W(r1,r2,z)W(r1,r1,z)W(r2,r2,z).
Applying Eqs. (13), (23)-(26), the propagation properties of the CSD and SDOC of a McLGCSM beam in free space or Kolmogorov turbulence can be numerically demonstrated in a convenient way.

4. Numerical examples for the spectral density of a McLGCSM beam in free space and atmospheric turbulence

In this section, we will numerically study the evolution of the spectral density of a McLGCSM beam on propagation in free space and turbulent atmosphere. Because we only emphasize on the evolution of the beam pattern configurations but do not care about the size of the beam patterns, in the following graphical discussions arbitrary units for transversal coordinates in different z-planes are used, although they have the same transversal dimensions in the same z-plane. In addition, the beam parameters are set as λ = 632.8nm, and βx = βy = β.

Figure 2 shows the normalized intensity distributions of a McLGCSM, McGCSM, and LGCSM beams in free space at several propagation distances with w0 = 5mm, Px = Py = 3, and δw = 1. For one typical case of β = 9, n = 3 as shown in Fig. 2(a), the McLGCSM beam exhibits a ring-shaped beam array during propagation in free space, i.e., the initial single Gaussian beam spot is first self-split apart into a beam spot array at short propagation distance z1 (z1 = 0.1km); then as the travelling distance increases the beam spot array further turns into a ring-shaped beam array at farther propagation distance z2 (z2 = 0.2km). Although the size of the patterns increases with increasing propagation distance, the configuration of the ring-shaped beam array is very robust and remains unchanged till z = 40km. Figures 2(b) and 2(c) show that a McGCSM beam is split into a beam spot array and a LGCSM beam is shaped into a single ring-shaped beam profile in the far field, which agrees well with the results that have been reported in [6–8, 25–27]. Note that the propagation behaviors of the proposed McLGCSM beam significantly differs from that of the McGCSM beam or LGCSM beam; or more specifically, it can be regarded as a superposition of the self-splitting of the McGCSM beam and self-shaping of the LGCSM beam. As a result, the number of the ring-shaped beams obtained by a McLGCSM beam is equal to PxPy, which is just same as the number of the self-splitting beam spots produced by a McGCSM beam [25–27]. Moreover, since previous studies have reported that by focusing a LGCSM beam the ring-shaped beam pattern created in free space can evolve into an optical cage near the focal plane [8], it is reasonable to infer that a tunable optical cage array can also be generated by focusing the proposed McLGCSM beam. To the best of our knowledge, such a phenomenon of a single beam spot self-splitting and self-shaping into a ring-shaped beam array in the far field has not been reported elsewhere.

 figure: Fig. 2

Fig. 2 Intensity distribution patterns of a McLGCSM beam (a), McGCSM beam (b), and LGCSM beam (c) at several propagation distances in free space with w0 = 5mm, Px = Py = 3, and δw = 1. The x and y axes taken as the transverse coordinates are in arbitrary units.

Download Full Size | PDF

Figure 3 shows the influence of the beam parameters β and δw on the formation of the ring-shaped beam array of a McLGCSM beam on propagation in free space with w0 = 5mm, Px = Py = 3, and n = 4. One finds from Fig. 3(b) that only for sufficiently large values of β and relatively small values of δw, the McLGCSM beam exhibits a ring-shaped beam array in the far field. Comparing Fig. 3(b) with Fig. 3(c) reveals that, when δw is sufficiently large, only self-splitting behavior can be observed during propagation, indicating that the self-shaping of a McLGCSM beam originates from the Laguerre-related factor of the entire SDOC. Furthermore, contrasting Fig. 3(a) with 3(b) shows that, when β is appropriately small, only self-shaping phenomenon can be observed during propagation, demonstrating that the self-splitting of a McLGCSM beam is induced by the multi-cosine-related factor of the entire SDOC. Obviously, these results are in good agreement with the analytical conclusions aforesaid in the previous section.

 figure: Fig. 3

Fig. 3 Intensity distribution patterns of a McLGCSM beam at several propagation distances in free space for different values of β and δw with w0 = 5mm, Px = Py = 3, and n = 4.

Download Full Size | PDF

Generally, the control of the number, position and null field of the ring-shaped beam profiles can provide the ring-shaped beam array more possibilities for practical applications. As a result, numerical calculations are performed to study the relationship between the properties of the ring-shaped beam array and the beam parameters of the McLGCSM source. In particular, we find that the size of the null field region of each ring-shaped beam pattern can be gigantic and the velocity evolving into a ring-shaped beam array can be enhanced with increasing the order n of Laguerre-polynomial. Performing calculations also indicate that the separation distances among these ring-shaped beam patterns increase in proportion with the beam parameter β. As already mentioned, the number of the ring-shaped beam patterns in the array is determined by the product value of PxPy. However, note that increasing the coherence length δ can disrupt the evolution process and the self-shaping phenomenon fades away when δ is greater than the critical value of approximately 2 as shown in Fig. 3(c). In short, the results reveal that our method can give rise to a controllable ring-shaped beam array by modulating the source structured coherence state of a McLGCSM beam on propagation in free space.

Now we study the propagation properties of a McLGCSM beam in turbulent atmosphere. Figure 4 shows the average intensity distributions of a McLGCSM, McGCSM, and LGCSM beams in turbulence at several propagation distances with the identical beam parameters used in Fig. 2 and the structure constant Cn2=5×1014m2/3. It is interesting to find that at short propagation distance, the influence of turbulence are negligible and the propagation properties of these beams in turbulence are similar to those in free space (see Fig. 2), while at long propagation distance, the influence of turbulence is greatly enhanced and all of these beams exhibit combining properties. Particularly, one finds from Fig. 4(a) that for a McLGCSM beam in Kolmogorov turbulence, the initial single Gaussian beam spot still successively evolves into a beam spot array and a ring-shaped beam array at short propagation distance; however, due to the accumulative effect of turbulence with increasing the propagation distance, the ring-shaped beam array is being gradually destroyed and eventually merge to a single Gaussian-like beam spot at sufficiently large propagation distance. Of course, it should be pointed out that the size of the finally fused single Gaussian-like beam spot in the far field is much larger than the initial one in the source plane although the configurations of these beam spots are quite similar with each other.

 figure: Fig. 4

Fig. 4 Average intensity distribution patterns of a McLGCSM beam (a), McGCSM beam (b), and LGCSM beam (c) at several propagation distances in atmospheric turbulence with w0 = 5mm, Px = Py = 3, δw = 1, and Cn2 = 5 × 10−14m−2/3.

Download Full Size | PDF

To further illustrate the influence of the atmospheric turbulence on the average intensity characteristics of a McLGCSM beam, we calculate in Fig. 5 the average intensity distributions of a McLGCSM beam in turbulence at several propagation distances for different values of the structure constantCn2 and beam parameter δw with w0 = 5mm, β = 10, and n = 5. When the structure constant Cn2 increases as shown in Figs. 5(a) and 5(b), the ring-shaped beam array becomes more vulnerable and quickly evolves into a Gaussian-like configuration after travelling a relatively short distance. Evidently, this is caused by the fact that the propagation properties are determined by the free-space diffraction and the atmospheric disturbance together. For a weak or moderate turbulence (Cn2=8×1015m2/3), the free-space diffraction plays a dominant role and the ring-shaped beam patterns can be obviously observed at propagation distances z = 1km and z = 8km. However, for a stronger turbulence (Cn2=5×1014m2/3), the atmospheric disturbance is strengthened and the ring-shaped beam profiles degrade seriously at propagation distance z = 8km. Furthermore, as illustrated in Figs. 5(a) and 5(c), we demonstrate that the ring-shaped beam array can retain for a longer propagation distance as the coherence length δ is reduced. Therefore, it implies that producing a McLGCSM source with smaller coherence length δ will have a stronger capability of resisting the destructive effect of atmospheric turbulence on the ring-shaped beam array formation.

 figure: Fig. 5

Fig. 5 Average intensity distribution patterns of a McLGCSM beam at several propagation distances in atmospheric turbulence for different values of Cn2 and δw with w0 = 5mm, β = 10, and n = 5.

Download Full Size | PDF

An analysis of the source structured coherence states suggests that the generation of the ring-shaped beam array derived from a McLGCSM beam can be attributed to its intrinsic coherence properties. It is shown in Fig. 1 that the SDOC distribution of a McLGCSM beam in the source plane presents a more complex structure compared with that of a LGCSM beam or a McGCSM beam. In fact, the model we proposed here is clearly more general: the non-conventional Laguerre- and multi-cosine-type components have been integrated together into one entire SDOC. By combining both non-conventional SDOC components to incorporate their individual excellent properties, the synthesis effect leads to the formation of the ring-shaped beam array for a McLGCSM beam in the far field. Hence, the principle for achieving a ring-shaped beam array here is based on modulating the spatial structure of non-conventional SDOC of the incident beam, which is totally different from that reported in [55] where multiple dark hollow patterns in the transverse plane are obtained with a full polarization-controlled method. In other words, careful design and proper tuning of the spatial coherence of a McLGCSM beam in the source plane can give rise to various beam patterns in the far field such as arriving a desired ring-shaped beam array. Since recently performed experiment indicates that the random Schell-model source can radiate any desired far-zone average intensity pattern with the help of the nematic, phase-only spatial light modulators [38], we believe that this technology can be directly applied here to generate this novel model source as well.

5. Propagation factor of a McLGCSM beam

The propagation factor (also named M2 factor) is an important characteristic parameter of a beam and can be regarded as a beam quality factor in many practical applications. In this section, we are going to derive the analytical expressions for the propagation factors of such beams within free space or in a turbulent atmosphere from the second-order moments of the Wigner distribution function (WDF) for a McLGCSM beam.

According to the method given in [42, 44, 45], the intensity moments of the order n1 + n2 + m1 + m2 of the WDF for partially coherent beams are defined as

xn1yn2θxm1θym2z=1Pd2ρd2θxn1yn2θxm1θym2h(ρ,θ,z)=1Pd2ρd2θG(ρ,θ,z)h(ρ,θ,0),
where P=h(ρ,θ,z)d2ρd2θ is the total power of the beam, and
h(ρ,θ,0)=(k2π)2W(0)(ρ,ρd,0)exp(ikρdθ)d2ρd,
h(ρ,θ,z)=(k2π)2W(ρ,ρd,z)exp(ikρdθ)d2ρd,
are respectively the WDFs in the source plane z = 0 and the output z-plane, and
G(ρ,θ,z)=(i)n1+n2+m1+m2Bn1+n2kn1+n2+m1+m2d2ρdd2ρdδ(n1)(xdDxd)δ(n2)(ydDyd)×δ(m1)(xd)δ(m2)(yd)exp[ikBρ(Aρdρd)+ikθρdH(ρd,ρd,z)],
with δ and δ(n) being the Dirac delta function and its nth derivative, respectively. For the propagation within free space or in a turbulent atmosphere it is clear that A = D = 1 and B = z.

Substituting Eqs. (13) and (20) into Eqs. (27)-(29) and after tedious integration, we obtain the following expressions for the second-order moments of the WDF of a McLGCSM beam in a turbulent atmosphere

x2z=2z2k2Tα+ρx20+2zρxθx0+z2θx20,
θx2z=6k2Tα+θx20,
xθxz=3zk2Tα+ρxθx0+zθx20,
where
Q(r,θ)0=1Pd2rd2θQ(r,θ)h(r,θ,0).
Completing some mathematical operations we finally obtain
x2+y2z=(4k2Tα+Mn,β2w02)z2+2w02,
xθx+yθyz=(6k2Tα+Mn,β2w02)z,
θx2+θy2z=12k2Tα+Mn,β2w02,
Mn,β=1k2[1+4w02δ2(n+1)+2w02δ2QM,β],
QM,β=βx2Px1lx=MxMxlx2+βy2Py1ly=MyMyly2.
Obviously, for βx = βy = 0, the expressions obtained above can reduce to those for a LGCSM beam [11, 14], while for n = 0, the expressions can reduce to those for a McGCSM beam [26].

The propagation factor of a partially coherent beam in atmospheric turbulence is defined in terms of the second-order moments as follows [42, 45]

M2(z)=k(ρ2zθ2zρθz2)1/2.
Inserting Eqs. (35)-(39) into Eq. (40), we obtain the following expression for the propagation factor of a McLGCSM beam in atmospheric turbulent
Mn,Q2(z)=[Mn,β+2Tαk2(12w02+6Tαk2z2+Mn,βw02z2)]1/2.
Under the condition of Tα → 0, Eq. (41) represents the propagation factor of a McLGCSM beam in free space and reduces to Mn,Q2(z)=Mn,β. Note that the propagation factor of the beam in free space is independent of the propagation distance as expected. Based on Eq. (41), we obtain Mn,Q2(z)>Mn,02(z)[M0,Q2(z)]>M0,02(z), indicating that the propagation factor of a McLGCSM beam is the largest compared with that of a LGCSM beam (with β = 0), McGCSM beam (with n = 0), and GSM beam (with β = n = 0).

The relative propagation factor is defined as

Mr,n,Q2(z)=Mn,Q2(z)/Mn,β.
Equation (42) describes the resistance of a beam to turbulence in comparison with the free-space propagation; i.e., the smaller the relative propagation factor is, the less the beam is affected by turbulence. Therefore, the relative propagation factor can be used to quantitatively compare the effect of turbulence for different partially coherent beams upon propagation. We comparatively analyze the relative propagation factors of a McLGCSM beam, LGCSM beam, McGCSM beam, and GSM beam, and generally confirm that
Mr,0,02(z)>Mr,n,02(z)[Mr,0,Q2(z)]>Mr,n,Q2(z).
For example,

Mr,n,Q2(z)Mr,n,02(z)=12Ω(Mn,02Mn,β2)(2w02+Ωz2/k2)k2Mn,βMn,0[Mn,0Mr,n,Q2(z)+Mn,βMr,n,02(z)]<0.

The theoretical predications are further demonstrated in Fig. 6 where the variations of the relative propagation factors against the propagation distance for a McLGCSM beam, LGCSM beam, McGCSM beam, and GSM beam in turbulent atmosphere are illustrated. It is shown that for all of these beams, the relative propagation factors grows rapidly from 1 with the increase of propagation distance, indicating the quality of these beams being degraded by the atmospheric turbulence. Moreover, we note that the growth of the relative propagation factor of a McLGCSM beam is the slowest compared with other beams, revealing that the McLGCSM beam is less affected by the atmospheric turbulence from the aspect of the propagation factor. Therefore, the results clearly show that the proposed McLGCSM beam has advantage in atmospheric turbulence, which will be useful in free-space optical communications.

 figure: Fig. 6

Fig. 6 Relative propagation factors Mr,n,Q2(z) versus the propagation distance z in atmospheric turbulence for (G) a GSM beam with n = βx = βy = 0, (L) a LGCSM beam with n = 5 and βx = βy = 0, (Mc) a McGCSM beam with n = 0 and βx = βy = 8, and (McL) a McLGCSM beam with n = 5 and βx = βy = 8. The other parameter are w0 = 10mm, δw = 1, Px = Py = 3, and Cn2 = 5 × 10−15m−2/3.

Download Full Size | PDF

6. Conclusions

We have introduced one kind of partially coherent beam, named as the McLGCSM beam whose SDOC in the source plane is a Gaussian function modulated by both non-conventional Laguerre- and multi-cosine-type factors, and studied its propagation properties in free space and atmospheric turbulence. Such a beam includes a LGCSM beam, McGCSM beam, and GSM beam as its special cases. The analytical expressions for the CSD and the propagation factor of a McLGCSM beam have been derived and its statistical properties, such as the spectral intensity and the propagation factor, have been illustrated numerically. We have found that a McLGCSM beam can form a ring-shaped beam array intensity profile in the far field on free-space propagation. Meanwhile, on propagation in atmospheric turbulence the McLGCSM beam still possesses a ring-shaped beam array profile at short distance but gradually merge to a Gaussian-like pattern at sufficiently large distance. Notably, differing from a LGCSM beam where only one single ring-shaped beam profile is realizable, with such a beam we can simultaneously obtain a ring-shaped beam array with adjustable number and positions by directly tuning the spatial structure of its SDOC. The physical mechanism behind the occurrence of the ring-shaped beam array derived from a McLGCSM beam is based on its peculiar source coherence state which combines the individual merits associated with each non-conventional SDOC component of the whole SDOC. Furthermore, we have generally proved that the propagation factor of a McLGCSM beam is minimally influenced by the atmospheric turbulence compared with that of a LGCSM, McGCSM and GSM beams. The work uncovers that the far-field ring-shaped beam array generated in free space can also be formed in atmospheric turbulence through synthesizing the non-conventional SDOC components to modulate the source spatial coherence properties, and the advantages of optical manipulation with controllable ring-shaped beam array will greatly benefit applications in micromanipulation of multiple particles, free-space optical communications and imaging in the atmosphere.

Funding

High level introduction of talent research start-up fund of Guizhou Institute of Technology

References and links

1. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University, 1995), pp. 33–39.

2. F. Gori, G. Guattari, and C. Padovani, “Modal expansion for J0-correlated Schell-model sources,” Opt. Commun. 64(4), 311–316 (1987). [CrossRef]  

3. J. Li, X. M. Gao, and Y. R. Chen, “Tight focusing of J0-correlated Gaussian Schell-model beam through high numerical aperture,” Opt. Commun. 285(16), 3403–3411 (2012). [CrossRef]  

4. J. Cang, P. Xiu, and X. Liu, “Propagation of Laguerre-Gaussian and Bessel-Gaussian Schell-model beams through paraxial optical systems in turbulent atmosphere,” Opt. Laser Technol. 54, 35–41 (2013). [CrossRef]  

5. F. Gori and M. Santarsiero, “Devising genuine spatial correlation functions,” Opt. Lett. 32(24), 3531–3533 (2007). [CrossRef]   [PubMed]  

6. Z. Mei and O. Korotkova, “Random sources generating ring-shaped beams,” Opt. Lett. 38(2), 91–93 (2013). [CrossRef]   [PubMed]  

7. F. Wang, X. Liu, Y. Yuan, and Y. Cai, “Experimental generation of partially coherent beams with different complex degrees of coherence,” Opt. Lett. 38(11), 1814–1816 (2013). [CrossRef]   [PubMed]  

8. Y. Chen and Y. Cai, “Generation of a controllable optical cage by focusing a Laguerre-Gaussian correlated Schell-model beam,” Opt. Lett. 39(9), 2549–2552 (2014). [CrossRef]   [PubMed]  

9. Y. Chen, F. Wang, C. Zhao, and Y. Cai, “Experimental demonstration of a Laguerre-Gaussian correlated Schell-model vortex beam,” Opt. Express 22(5), 5826–5838 (2014). [CrossRef]   [PubMed]  

10. Y. Chen, L. Liu, F. Wang, C. Zhao, and Y. Cai, “Elliptical Laguerre-Gaussian correlated Schell-model beam,” Opt. Express 22(11), 13975–13987 (2014). [CrossRef]   [PubMed]  

11. R. Chen, L. Liu, S. Zhu, G. Wu, F. Wang, and Y. Cai, “Statistical properties of a Laguerre-Gaussian Schell-model beam in turbulent atmosphere,” Opt. Express 22(2), 1871–1883 (2014). [CrossRef]   [PubMed]  

12. L. N. Guo, Y. H. Chen, L. Liu, and Y. J. Cai, “Propagation of a Laguerre-Gaussian correlated Schell-model beam beyond the paraxial approximation,” Opt. Commun. 352, 127–134 (2015). [CrossRef]  

13. H. F. Xu, Z. Zhang, J. Qu, and W. Huang, “The tight focusing properties of Laguerre–Gaussian-correlated Schell-model beams,” J. Mod. Opt. 63(15), 1429–1437 (2016). [CrossRef]  

14. Y. Zhou, Y. Yuan, J. Qu, and W. Huang, “Propagation properties of Laguerre-Gaussian correlated Schell-model beam in non-Kolmogorov turbulence,” Opt. Express 24(10), 10682–10693 (2016). [CrossRef]   [PubMed]  

15. Y. H. Chen, J. Y. Yu, Y. S. Yuan, F. Wang, and Y. J. Cai, “Theoretical and experimental studies of a rectangular Laguerre-Gaussian-correlated Schell-model beam,” Appl. Phys. B 122(2), 31 (2016). [CrossRef]  

16. Y. L. Qiu, Z. X. Chen, and Y. J. He, “Propagation of a Laguerre-Gaussian correlated Schell-model beam in strongly nonlocal nonlinear media,” Opt. Commun. 389, 303–309 (2017). [CrossRef]  

17. Z. Mei and O. Korotkova, “Cosine-Gaussian Schell-model sources,” Opt. Lett. 38(14), 2578–2580 (2013). [CrossRef]   [PubMed]  

18. Z. Mei, E. Schchepakina, and O. Korotkova, “Propagation of cosine-Gaussian-correlated Schell-model beams in atmospheric turbulence,” Opt. Express 21(15), 17512–17519 (2013). [CrossRef]   [PubMed]  

19. C. Liang, F. Wang, X. Liu, Y. Cai, and O. Korotkova, “Experimental generation of cosine-Gaussian-correlated Schell-model beams with rectangular symmetry,” Opt. Lett. 39(4), 769–772 (2014). [CrossRef]   [PubMed]  

20. Z. Mei, “Light sources generating self-splitting beams and their propagation in non-Kolmogorov turbulence,” Opt. Express 22(11), 13029–13040 (2014). [CrossRef]   [PubMed]  

21. H. F. Xu, Z. Zhang, J. Qu, and W. Huang, “Propagation factors of cosine-Gaussian-correlated Schell-model beams in non-Kolmogorov turbulence,” Opt. Express 22(19), 22479–22489 (2014). [CrossRef]   [PubMed]  

22. C. L. Ding, L. M. Liao, H. X. Wang, Y. T. Zhang, and L. Z. Pan, “Effect of oceanic turbulence on the propagation of cosine-Gaussian-correlated Schell-model beams,” J. Opt. 17(3), 035615 (2015). [CrossRef]  

23. S. Zhu, Y. Chen, J. Wang, H. Wang, Z. Li, and Y. Cai, “Generation and propagation of a vector cosine-Gaussian correlated beam with radial polarization,” Opt. Express 23(26), 33099–33115 (2015). [CrossRef]   [PubMed]  

24. J. Wang, S. Zhu, H. Wang, Y. Cai, and Z. Li, “Second-order statistics of a radially polarized cosine-Gaussian correlated Schell-model beam in anisotropic turbulence,” Opt. Express 24(11), 11626–11639 (2016). [CrossRef]   [PubMed]  

25. Z. Mei, D. Zhao, O. Korotkova, and Y. Mao, “Gaussian Schell-model arrays,” Opt. Lett. 40(23), 5662–5665 (2015). [CrossRef]   [PubMed]  

26. Z. Z. Song, Z. J. Liu, K. Y. Zhou, Q. G. Sun, and S. T. Liu, “Propagation properties of Gaussian Schell model array beams in non-Kolmogorov turbulence,” J. Opt. 18(10), 105601 (2016). [CrossRef]  

27. Y. H. Mao, Z. R. Mei, and J. G. Gu, “Propagation of Gaussian Schell-model Array beams in free space and atmospheric turbulence,” Opt. Laser Technol. 86, 14–20 (2016). [CrossRef]  

28. Y. H. Mao and Z. R. Mei, “Random sources generating ring-shaped optical lattice,” Opt. Commun. 381, 222–226 (2016). [CrossRef]  

29. Z. R. Mei and O. Korotkova, “Electromagnetic Schell-model sources generating far fields with stable and flexible concentric rings profiles,” Opt. Express 24(5), 5572–5583 (2016). [CrossRef]  

30. M. M. Tang, D. M. Zhao, X. Z. Li, and H. H. Li, “Focusing properties of radially polarized multi-cosine Gaussian correlated Schell-model beams,” Opt. Commun. 396, 249–256 (2017). [CrossRef]  

31. X. Wang, Z. R. Liu, K. L. Huang, and J. B. Sun, “Spectral shifts generated by scattering of Gaussian Schell-model arrays beam from a deterministic medium,” Opt. Commun. 387, 230–234 (2017). [CrossRef]  

32. Y. H. Chen, J. X. Gu, F. Wang, and Y. J. Cai, “Self-splitting properties of a Hermite-Gaussian correlated Schell-model beam,” Phys. Rev. A 91(1), 013823 (2015). [CrossRef]  

33. J. Yu, Y. Chen, L. Liu, X. Liu, and Y. Cai, “Splitting and combining properties of an elegant Hermite-Gaussian correlated Schell-model beam in Kolmogorov and non-Kolmogorov turbulence,” Opt. Express 23(10), 13467–13481 (2015). [CrossRef]   [PubMed]  

34. L. Ma and S. A. Ponomarenko, “Free-space propagation of optical coherence lattices and periodicity reciprocity,” Opt. Express 23(2), 1848–1856 (2015). [CrossRef]   [PubMed]  

35. J. Li, F. Wang, and O. Korotkova, “Random sources for cusped beams,” Opt. Express 24(16), 17779–17791 (2016). [CrossRef]   [PubMed]  

36. F. Wang and O. Korotkova, “Circularly symmetric cusped random beams in free space and atmospheric turbulence,” Opt. Express 25(5), 5057–5067 (2017). [CrossRef]   [PubMed]  

37. Z. Mei and O. Korotkova, “Random sources for rotating spectral densities,” Opt. Lett. 42(2), 255–258 (2017). [CrossRef]   [PubMed]  

38. M. W. Hyde IV, S. Basu, D. G. Voelz, and X. F. Xiao, “Experimentally generating any desired partially coherent Schell-model source using phase-only control,” J. Appl. Phys. 118(9), 093102 (2015). [CrossRef]  

39. Y. H. Chen, S. A. Ponomarenko, and Y. J. Cai, “Experimental generation of optical coherence lattices,” Appl. Phys. Lett. 109(6), 061107 (2016). [CrossRef]  

40. C. Liang, C. Mi, F. Wang, C. Zhao, Y. Cai, and S. A. Ponomarenko, “Vector optical coherence lattices generating controllable far-field beam profiles,” Opt. Express 25(9), 9872–9885 (2017). [CrossRef]   [PubMed]  

41. J. Zhu, H. Q. Tang, Q. Su, and K. C. Zhu, “Cascade self-splitting of a Hermite-cos-Gaussian correlated Schell-model beam,” Europhys. Lett. 118(1), 14001 (2017). [CrossRef]  

42. Y. Dan and B. Zhang, “Second moments of partially coherent beams in atmospheric turbulence,” Opt. Lett. 34(5), 563–565 (2009). [CrossRef]   [PubMed]  

43. Y. L. Qiu, Z. X. Chen, and L. Liu, “Partially coherent dark hollow beams propagating through real ABCD optical systems in a turbulent atmosphere,” J. Mod. Opt. 57(8), 662–669 (2010). [CrossRef]  

44. K. Zhu, S. Li, Y. Tang, Y. Yu, and H. Tang, “Study on the propagation parameters of Bessel-Gaussian beams carrying optical vortices through atmospheric turbulence,” J. Opt. Soc. Am. A 29(3), 251–257 (2012). [CrossRef]   [PubMed]  

45. X. Q. Li and X. L. Ji, “Propagation of higher-order intensity moments through an optical system in atmospheric turbulence,” Opt. Commun. 298, 1–7 (2013). [CrossRef]  

46. Y. Zhao, Q. Zhan, Y. Zhang, and Y. P. Li, “Creation of a three-dimensional optical chain for controllable particle delivery,” Opt. Lett. 30(8), 848–850 (2005). [CrossRef]   [PubMed]  

47. L. Isenhower, W. Williams, A. Dally, and M. Saffman, “Atom trapping in an interferometrically generated bottle beam trap,” Opt. Lett. 34(8), 1159–1161 (2009). [CrossRef]   [PubMed]  

48. P. Xu, X. He, J. Wang, and M. Zhan, “Trapping a single atom in a blue detuned optical bottle beam trap,” Opt. Lett. 35(13), 2164–2166 (2010). [CrossRef]   [PubMed]  

49. P. Zhang, T. Li, J. Zhu, X. Zhu, S. Yang, Y. Wang, X. Yin, and X. Zhang, “Generation of acoustic self-bending and bottle beams by phase engineering,” Nat. Commun. 5, 4316 (2014). [PubMed]  

50. C. Wan, K. Huang, T. Han, E. S. P. Leong, W. Ding, L. Zhang, T. Yeo, X. Yu, J. Teng, D. Lei, S. A. Maier, B. Luk’yanchuk, S. Zhang, and C. Qiu, “Three-dimensional visible-light capsule enclosing perfect supersized darkness via antiresolution,” Laser Photonics Rev. 8(5), 743–749 (2014). [CrossRef]  

51. H. Wang, J. Lin, D. Zhang, Y. Wang, M. Gu, H. P. Urbach, F. Gan, and S. Zhuang, “Creation of an anti-imaging system using binary optics,” Sci. Rep. 6(1), 33064 (2016). [CrossRef]   [PubMed]  

52. R. Schmidt, C. A. Wurm, S. Jakobs, J. Engelhardt, A. Egner, and S. W. Hell, “Spherical nanosized focal spot unravels the interior of cells,” Nat. Methods 5(6), 539–544 (2008). [CrossRef]   [PubMed]  

53. Y. Iketaki, H. Kumagai, K. Jahn, and N. Bokor, “Creation of a three-dimensional spherical fluorescence spot for super-resolution microscopy using a two-color annular hybrid wave plate,” Opt. Lett. 40(6), 1057–1060 (2015). [CrossRef]   [PubMed]  

54. P. Bingen, M. Reuss, J. Engelhardt, and S. W. Hell, “Parallelized STED fluorescence nanoscopy,” Opt. Express 19(24), 23716–23726 (2011). [CrossRef]   [PubMed]  

55. X. Weng, L. Du, P. Shi, and X. Yuan, “Tunable optical cage array generated by Dammann vector beam,” Opt. Express 25(8), 9039–9048 (2017). [CrossRef]   [PubMed]  

56. M. Abramowitz and I. Stegun, Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables (U. S. Department of Commerce, 1970).

57. E. Jafarov, S. Lievens, and J. V. Jeugt, “The Wigner distribution function for the one-dimensional parabose oscillator,” J. Phys. A 41(23), 235301 (2008). [CrossRef]  

58. E. D. Rainville, Special Functions (Macmillan, 1960)

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 (6)

Fig. 1
Fig. 1 Density plots of the square of the modulus of the SDOC for (a) a McGCSM source with n = 0 and βx = βy = 8, (b) a LGCSM source with n = 5 and βx = βy = 0, and (c) a McLGCSM source with n = 5 and βx = βy = 8. The other parameters are Px = Py = 3.
Fig. 2
Fig. 2 Intensity distribution patterns of a McLGCSM beam (a), McGCSM beam (b), and LGCSM beam (c) at several propagation distances in free space with w0 = 5mm, Px = Py = 3, and δw = 1. The x and y axes taken as the transverse coordinates are in arbitrary units.
Fig. 3
Fig. 3 Intensity distribution patterns of a McLGCSM beam at several propagation distances in free space for different values of β and δw with w0 = 5mm, Px = Py = 3, and n = 4.
Fig. 4
Fig. 4 Average intensity distribution patterns of a McLGCSM beam (a), McGCSM beam (b), and LGCSM beam (c) at several propagation distances in atmospheric turbulence with w0 = 5mm, Px = Py = 3, δw = 1, and C n 2 = 5 × 10−14m−2/3.
Fig. 5
Fig. 5 Average intensity distribution patterns of a McLGCSM beam at several propagation distances in atmospheric turbulence for different values of C n 2 and δw with w0 = 5mm, β = 10, and n = 5.
Fig. 6
Fig. 6 Relative propagation factors M r,n,Q 2 ( z ) versus the propagation distance z in atmospheric turbulence for (G) a GSM beam with n = βx = βy = 0, (L) a LGCSM beam with n = 5 and βx = βy = 0, (Mc) a McGCSM beam with n = 0 and βx = βy = 8, and (McL) a McLGCSM beam with n = 5 and βx = βy = 8. The other parameter are w0 = 10mm, δw = 1, Px = Py = 3, and C n 2 = 5 × 10−15m−2/3.

Equations (44)

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

d 2 r 1 d 2 r 2 W ( 0 ) ( r 1 , r 2 )f( r 1 )f( r 2 )0.
W ( 0 ) ( r 1 , r 2 )= d 2 v p( v ) H 0 * ( r 1 ,v ) H 0 ( r 2 ,v ),
H 0 ( r ,v )=τ( r )exp( i r v ),
W ( 0 ) ( r 1 , r 2 )= τ * ( r 1 )τ( r 2 ) d 2 v p( v )exp[ iv( r 1 r 2 ) ]= τ * ( r 1 )τ( r 2 )μ( r 1 r 2 )
p( v x , v x )= N 4 l x = M x M x l y = M y M y [ p L ( v x + l x β x δ , v y + l y β y δ ) + p L ( v x + l x β x δ , v y l y β y δ ) + p L ( v x l x β x δ , v y + l y β y δ )+ p L ( v x l x β x δ , v y l y β y δ ) ],
p L ( v x , v y )= ( v x 2 + v y 2 δ 2 ) n exp[ δ 2 ( v x 2 + v y 2 ) 2 ],
f( z z 0 ) exp( ixz )dz=exp( i z 0 x ) f( y ) exp( ixy )dy,
W ( 0 ) ( r 1 , r 2 )=exp( r s 2 2 w 0 2 r d 2 8 w 0 2 )μ( r d ),
μ( r d )=N L n ( r d 2 δ 2 )exp( r d 2 2 δ 2 ) l x = M x M x l y = M y M y cos( l x β x x d δ )cos( l y β y y d δ ) ,
W( r 1 , r 2 ,z )= 1 λ 2 z 2 exp[ ik 2z ( r 2 2 r 1 2 ) ] d 2 r 1 d 2 r 2 W ( 0 ) ( r 1 , r 2 ) ×exp[ ik 2z ( r 1 2 r 2 2 )+ ik z ( r 1 r 1 r 2 r 2 ) ] exp[ Φ( r 1 , r 1 )+ Φ * ( r 2 , r 2 ) ] ,
exp[ Φ( r 1 , r 1 )+ Φ * ( r 2 , r 2 ) ] =exp[ T α ( r d 2 + r d r d + r d 2 ) ],
W( r 1 , r 2 ,z )= N λ 2 z 2 exp( ik z r r d T α r d 2 ) l x =1 M x l y =1 M y d 2 r s d 2 r d L n ( r d 2 2 δ 2 ) ×cos( Q l x x d δ )cos( Q l y y d δ )exp[ r s 2 2 w 0 2 ( 1 δ 2 + 1 4 w 0 2 +2 T α ) r d 2 2 ] ×exp[ ik z r s r d + ik z ( r d r s +r r d ) T α r d r d ].
W( r 1 , r 2 ,z )= 2π w 0 2 N λ 2 z 2 exp[ 2i z R z w 0 2 r r d ( 2 z R 2 z 2 w 0 2 + T α ) r d 2 ] × l x = M x M x l y = M y M y [ W l,++ ( r 1 , r 2 )+ W l,+ ( r 1 , r 2 )+ W l,+ ( r 1 , r 2 )+ W l, ( r 1 , r 2 ) ] ,
W l,uv ( r 1 , r 2 )= 1 4 d 2 r d L n ( r d 2 2 δ 2 )exp[ Ω r d 2 8 w 0 2 + ik z ( b l x ,xu x d + b l y ,yv y d ) ],
Ω=1+ 4 δ w 2 +8 w 0 2 T α + 16 z R 2 z 2 ,
b q± =q+i( T α z k 2 z R z ) q d ± z w 0 2 z R δ w l q β q ,( q=x,yu,v=± ),
L n ( x )= p=0 n ( n p ) ( 1 ) p p! x p
( x 2 + y 2 ) p = m=0 p ( p m ) x 2( pm ) y 2m ,
z n exp[ ( zβ ) 2 ]dz= ( 2i ) n π H n ( iβ ),
W l,uv ( r 1 , r 2 )= 2π w 0 2 Ω exp[ 2 k 2 w 0 2 ( b l x ,xu 2 + b l y ,yv 2 ) Ω z 2 ] p=0 n ( n p ) 1 p! ( 1 Ω δ w 2 ) p × m=0 p ( p m ) H 2( pm ) ( 2 Ω k w 0 z b l x ,xu ) H 2m ( 2 Ω k w 0 z b l y ,yv ),
m=0 n ( n m ) H 2m ( x ) H 2n2m ( y ) = ( 4 ) n n! L n ( x 2 + y 2 ),
L n ( bx )= ( 1b ) n m=0 n ( b 1b ) m ( n m ) L m ( x ) ,
W l,uv ( r 1 , r 2 )= 2π w 0 2 Ω exp[ 2 k 2 w 0 2 ( b l x ,xu 2 + b l y ,yv 2 ) Ω z 2 ] m=0 n ( n m ) ( 4 Ω δ w 2 ) m L m [ 2 k 2 w 0 2 ( b l x ,xu 2 + b l y ,yv 2 ) Ω z 2 ] = 2π w 0 2 Ω ( Ω δ w 2 Ω δ w 2 4 ) n exp[ 2 k 2 w 0 2 ( b l x ,xu 2 + b l y ,yv 2 ) Ω z 2 ] L m [ 8 k 2 w 0 2 ( b l x ,xu 2 + b l y ,yv 2 ) Ω z 2 ( 4Ω δ w 2 ) ].
I( r,z )=W( r,r,z )= πN w 0 2 z 2 z R 2 l x =1 M x l y =1 M y [ I l,++ ( r,z )+ I l,+ ( r,z )+ I l,+ ( r,z )+ I l, ( r,z ) ] ,
I l,±± ( r,z )= ( Ω δ w 2 Ω δ w 2 4 ) n exp( 8 z R 2 Ω z 2 R l x , l y ) L n [ 32 z R 2 δ w 2 z 2 ( 4Ω δ w 2 ) R l x , l y ]
μ( r 1 , r 2 ,z )= W( r 1 , r 2 ,z ) W( r 1 , r 1 ,z )W( r 2 , r 2 ,z ) .
x n 1 y n 2 θ x m 1 θ y m 2 z = 1 P d 2 ρ d 2 θ x n 1 y n 2 θ x m 1 θ y m 2 h( ρ,θ,z ) = 1 P d 2 ρ d 2 θ G( ρ , θ ,z )h( ρ , θ ,0 ),
h( ρ,θ,0 )= ( k 2π ) 2 W ( 0 ) ( ρ, ρ d ,0 )exp( ik ρ d θ ) d 2 ρ d ,
h( ρ,θ,z )= ( k 2π ) 2 W ( ρ, ρ d ,z )exp( ik ρ d θ ) d 2 ρ d ,
G( ρ , θ ,z )= ( i ) n 1 + n 2 + m 1 + m 2 B n 1 + n 2 k n 1 + n 2 + m 1 + m 2 d 2 ρ d d 2 ρ d δ ( n 1 ) ( x d D x d ) δ ( n 2 ) ( y d D y d ) × δ ( m 1 ) ( x d ) δ ( m 2 ) ( y d )exp[ ik B ρ ( A ρ d ρ d )+ik θ ρ d H( ρ d , ρ d ,z ) ],
x 2 z = 2 z 2 k 2 T α + ρ x 2 0 +2z ρ x θ x 0 + z 2 θ x 2 0 ,
θ x 2 z = 6 k 2 T α + θ x 2 0 ,
x θ x z = 3z k 2 T α + ρ x θ x 0 +z θ x 2 0 ,
Q( r,θ ) 0 = 1 P d 2 r d 2 θ Q( r,θ )h( r,θ,0 ).
x 2 + y 2 z =( 4 k 2 T α + M n,β 2 w 0 2 ) z 2 +2 w 0 2 ,
x θ x +y θ y z =( 6 k 2 T α + M n,β 2 w 0 2 )z,
θ x 2 + θ y 2 z = 12 k 2 T α + M n,β 2 w 0 2 ,
M n,β = 1 k 2 [ 1+ 4 w 0 2 δ 2 ( n+1 )+ 2 w 0 2 δ 2 Q M,β ],
Q M,β = β x 2 P x 1 l x = M x M x l x 2 + β y 2 P y 1 l y = M y M y l y 2 .
M 2 ( z )=k ( ρ 2 z θ 2 z ρθ z 2 ) 1/2 .
M n,Q 2 ( z )= [ M n,β + 2 T α k 2 ( 12 w 0 2 + 6 T α k 2 z 2 + M n,β w 0 2 z 2 ) ] 1/2 .
M r,n,Q 2 ( z )= M n,Q 2 ( z )/ M n,β .
M r,0,0 2 ( z )> M r,n,0 2 ( z )[ M r,0,Q 2 ( z ) ]> M r,n,Q 2 ( z ).
M r,n,Q 2 ( z ) M r,n,0 2 ( z )= 12Ω( M n,0 2 M n,β 2 )( 2 w 0 2 + Ω z 2 / k 2 ) k 2 M n,β M n,0 [ M n,0 M r,n,Q 2 ( z )+ M n,β M r,n,0 2 ( z ) ] <0.
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.