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

Behavior of tiled-aperture arrays fed by vector partially coherent sources

Open Access Open Access

Abstract

This paper explores the far-zone behavior of partially coherent arrays. We derive an expression for the far-zone spectral density valid for any array composed of circular elements and fed by fields with Schell-model cross-spectral density functions. This expression is written as the sum of convolution integrals, making it easy to physically interpret. We discuss this expression at length and present examples. Lastly, we validate our analysis by comparing Monte Carlo averages from wave-optics simulations with theory. We conclude this paper with a brief summary of the results and potential uses of our work.

1. INTRODUCTION

Research into partially coherent sources or beams is quite popular because of their ability to maintain directionality while also being resistant to speckle or scintillation [1,2]. In addition, partially coherent sources are highly tailorable in shape, polarization, and coherence properties [36]. These attributes make them potentially useful in directed-energy, remote-sensing, optical-communication, and medical applications.

Research into tiled-aperture array technology has largely occurred in parallel with the coherence work outlined above. This research is motivated by the significant, potential improvements in system size, weight, and power afforded by fiber laser beam combining.

In general, beam combining in tiled-aperture arrays is accomplished in one of two ways—incoherently or coherently. Incoherent systems are simpler in design and, therefore, more technologically mature. The US Navy has invested heavily in this technology and has deployed an incoherently combined array [see the Laser Weapon System (LaWS) program] [7,8].

Coherently combined arrays hold the promise of much higher peak intensities than incoherent systems [912]. Since this scheme requires polarization control, path-length matching, and phase locking of multiple fiber lasers in a master-oscillator-power-amplifier configuration, the design of coherently combined arrays is much more complex than incoherent systems. As such, coherent beam combining is still largely in the research/laboratory phase of development [9,1315].

Anticipating the possible benefits of combining coherence control with laser array technology, several authors have investigated partially coherent array sources and beams. For obvious reasons, a large majority of these studies are concerned with how these arrays perform in the presence of turbulence, both atmospheric and oceanic [1621].

In order to arrive at analytical expressions for the average intensity, degree of coherence, scintillation index, etc., these studies use Gaussian Schell models, or their many variants, as the mutual coherence [or cross-spectral density (CSD)] functions of the array element fields [16,1825]. Using Gaussian Schell models as the basis for the source statistics results in output statistics that are generally sums of Gaussian functions. These expressions can be hard to physically interpret and are typically explored computationally.

Here, we step back and examine the far-zone behavior of a tiled-aperture array fed by vector partially coherent sources. We present the general analytical far-zone spectral density with the assumptions that the array elements are circular and that the partially coherent sources feeding the array are wide-sense stationary with Schell-model CSD functions.

The spectral density obtained here is not a closed-form function; however, we express it as a convolution integral, which is easy to physically interpret. We discuss the physical nature of this expression in detail and further explore it with three examples. We then validate this far-zone spectral density expression by comparing and contrasting analytical predictions with Monte Carlo averages from wave-optics simulations. Lastly, we conclude this paper with a brief summary of contributions.

2. THEORY

Using Love’s equivalence theorem [26,27], image theory [28,29], and the physical optics approximation [30,31], the approximate far-zone electric field radiated from a planar, conformal tiled-aperture array takes the general form

E(r,ω)=θ^jk4πrejkr[Lx(r,ω)sinϕ+Ly(r,ω)cosϕ]+ϕ^jk4πrejkr[Lx(r,ω)cosθcosϕLy(r,ω)cosθsinϕ],
where θ^ is the unit vector in the polar angle θ direction, ϕ^ is the unit vector in the azimuth angle ϕ direction, r=x^x+y^y+z^z is the observation vector, k=2π/λ, λ is the wavelength, and ω is the radian frequency.

The far-zone vector potential components Lx and Ly are

Lx,y(r,ω)=2SEy,xsrc(ρ,ω)ejkr^·ρd2ρ,
where ρ=x^x+y^y is the source vector, r^=r/r is the unit vector in the direction of r, Esrc is the stochastic electric field feeding the array (assumed to be a sample function drawn from a wide-sense stationary random process) [32], and S represents the area of the array. Note that the exp(jωt) time convention is used throughout.

Because of our interest in directed-energy applications, we are primary concerned with the spatial distribution of power in the far-zone. This information is captured in the spectral density, namely,

S(r,ω)E(r,ω)·E*(r,ω)=k2(4πr)2(cos2θcos2ϕ+sin2ϕ)Lx(r,ω)Lx*(r,ω)+k2(4πr)2(cos2θsin2ϕ+cos2ϕ)Ly(r,ω)Ly*(r,ω)+k2(4πr)2sin2θsin(2ϕ)Re[Lx(r,ω)Ly*(r,ω)],
where · is the average over the ensemble.

In tiled-aperture arrays, λd, where d is the diameter (or width) of an array element. A vast majority of the radiated power is, therefore, contained within a small angle centered on the z axis. Applying what is equivalently the paraxial approximation (namely, cosθ1; sinθθ, such that sin2θθ20; rz; and r^=r/rx^x/z+y^y/z+z^) to Eq. (3) yields

S(ρ,z,ω)k2(4πz)2[Lx(ρ,z,ω)Lx*(ρ,z,ω)+Ly(ρ,z,ω)Ly*(ρ,z,ω)],
where Lx and Ly become
Lx,y(ρ,z,ω)2SEy,xsrc(ρ,ω)exp(jkzρ·ρ)d2ρ.
In the next section, we assume a Schell-model form for the CSD function of Esrc. This permits Eq. (4) to be simplified into a physically meaningful expression.

A. Esrc Cross-Spectral Density

To make further progress, an analytical model for the random field Esrc must be chosen. Let a single realization of Esrc be

Esrc(ρ,ω)=m=1Ncirc(|ρρm|d/2)×[x^Cx,m(ω)tx,m(ρ,ω)+y^Cy,m(ω)ty,m(ρ,ω)],
where d is the diameter of an array element, α=x,y, Cα,m and tα,m are the complex amplitude and random complex transmittance function of the α component of the field in the mth array element, respectively, circ(x) is the circle function defined by Goodman [33], and ρm is a vector that points from the origin to the center of the mth array element.

The random complex transmittances, which make the source field in Eq. (6) stochastic, can arise physically from platform vibrations (or jitter) or from thermal gradients in the transmitter optics. They could also be applied intentionally, for example, to reduce turbulence-induced scintillation.

The expression for Esrc given in Eq. (6) is very general. Note that each array element has a unique polarization state and spatial degree of coherence. The latter can be controlled to make an element mutually, spatially coherent (or incoherent) with any of the other array elements.

Taking the vector autocorrelation of Eq. (6) to form the CSD dyad (only the diagonal elements are required), where the ω dependence is assumed and, henceforth, suppressed, yields

Eαsrc(ρ1)Eαsrc*(ρ2)=m=1Nn=1NCα,mCα,n*Pm(ρ1)Pn(ρ2)tα,m(ρ1)tα,n*(ρ2)=m=1Nn=1NCα,mCα,n*Pm(ρ1)Pn(ρ2)μαα,mn(ρ1ρ2),
where Pm(ρ)=circ[|ρρm|/(d/2)], and μαα,mn is the spatial cross-correlation function between the mth and nth array element complex transmittances. Note that the CSD in Eq. (7) takes a Schell-model form with μαα,mn(ρ)=μαα,nm*(ρ), μαα,mm(0)=1, and |μαα,mn(0)|1 [1,34].

B. Spectral Density S(ρ,z)

Substituting Eq. (7) into Eq. (4), making the variable substitutions s=ρ1 and t=ρ1ρ2, and simplifying produces

S(ρ,z)=1(λz)2α=x,ym=1Nn=1NCα,mCα,n*μαα,mn(t)×[Pm(s)Pn(st)d2s]exp(jkzt·ρ)d2t=1(λz)2α=x,ym=1Nn=1NCα,mCα,n*×μαα,mn(t)Λmn(t)exp(j2πt·ρλz)d2t,
where Λmn(t), the quantity in brackets, is the autocorrelation of the mth and nth element functions.

Equation (8) is the inverse Fourier transform of the product of two functions; thus, by using the convolution theorem, Eq. (8) becomes

S(ρ,z)=1(λz)2α=x,ym=1Nn=1NCα,mCα,n*×μ˜αα,mn(f)Λ˜mn(fρλz)d2f,
where f=x^fx+y^fy is the spatial frequency vector, and the overscore “” denotes the Fourier transform.

Evaluating Λ˜mn and simplifying results in

S(ρ,z)=α=x,ym=1Nn=1NCα,mCα,n*[π(d/2)2λz]2jinc2(2πd2ρ)×exp[j2π(xρmρn)·ρ]μ˜αα,mn(ρ)|ρ=ρλz,
where represents two-dimensional convolution, jinc(x)=2J1(x)/x, and J1(x) is a first-order Bessel function of the first kind. Lastly, Eq. (10) can be written as the sum of incoherent and coherent terms, such that
S(ρ,z)=Sincoh(ρ,z)+Scoh(ρ,z),Sincoh(ρ,z)=α=x,ym=1N|Cα,m|2[π(d/2)2λz]2×jinc2(2πd2ρ)μ˜αα,mm(ρ)|ρ=ρλz,Scoh(ρ,z)=α=x,ym=1NnmNCα,mCα,n*[π(d/2)2λz]2jinc2(2πd2ρ)×exp[j2π(ρmρn)·ρ]μ˜αα,mn(ρ)|ρ=ρλz.
Equation (11) is the main analytical result of this paper. We explore a few examples in the next section to demonstrate the physical intuitiveness of this expression.

C. Examples

In this section, we present three examples labeled Scenario 1, Scenario 2, and Scenario 3, respectively. The goal of this exercise is to reveal the physical nature of Eq. (11).

1. Scenario 1

First, assume that Cα,m=Cα and μαα,mn=μαα, then Eq. (11) simplifies to

Sincoh(ρ,z)=α=x,y|Cα|2N[π(d/2)2λz]2jinc2(2πd2ρ)μ˜αα(ρ)|ρ=ρλz,Scoh(ρ,z)=α=x,y|Cα|2m=1NnmN[π(d/2)2λz]2jinc2(2πd2ρ)×exp[j2π(ρmρn)·ρ]μ˜αα(ρ)|ρ=ρλz,S(ρ,z)=α=x,y|Cα|2{N2[π(d/2)2λz]2jinc2(2πd2ρ)×|1Nm=1Nexp(j2πρm·ρ)|2}μ˜αα(ρ)|ρ=ρλz.
This scenario models an array fed by a common partially coherent source.

The expression in braces is equivalent to the far-zone array pattern when the elements are fed by a common spatially coherent source (e.g., a laser). We refer to this expression throughout the remainder of this paper; therefore, we assign it the symbol L for convenience.

L has two scale sizes. The first is due to the jinc(x) function (also known as the element pattern), whose width is proportional to λz/d; the second is due to the summation term (better known as the array factor), whose width is inversely proportional to the array’s full diameter, i.e., λz/D.

These dimensions, when compared to the widths of μαα (represented hereafter as δαα), result in three general behaviors for S(ρ,z):

  • 1. If δαα>D, μ˜αα are fast functions (narrow functions) compared to the array factor. The light from each array element adds coherently, and S(ρ,z) is approximately equal to L.
  • 2. If d<δαα<D, μ˜αα are fast functions compared to the element pattern and slow functions (wide functions) compared to the array factor. The light from different array elements partially interferes, and S(ρ,z) transforms continuously from L (when δααD) to the element pattern (when δααd).
  • 3. If δαα<d, μ˜αα are slow functions compared to the element pattern. The light from each array element adds incoherently, and S(ρ,z) asymptotically approaches μ˜αα as δαα0.

2. Scenario 2

For the next example, let Cα,m=Cα and μαα,mn=μααδmn, where δmn is the Kronecker delta function. Substituting these into Eq. (11) and simplifying produces

Sincoh(ρ,z)=S(ρ,z)=α=x,y|Cα|2{N[π(d/2)2λz]2jinc2(2πd2ρ)}μ˜αα(ρ)|ρ=ρλz.
This scenario models an array fed by N mutually incoherent sources.

The quantity in braces is N times the element pattern. As previously discussed, the width of the element pattern is proportional to λz/d. Thus, how the δαα compare to d determines the behavior of S(ρ,z):

  • 1. If δαα>d, then μ˜αα are fast functions compared to the element pattern, and S(ρ,z) is approximately equal to N times the element pattern.
  • 2. If δαα<d, then μ˜αα are slow functions compared to the element pattern, and S(ρ,z) transforms continuously from N times the element pattern (when δααd) to N times μ˜αα as δαα0.

3. Scenario 3

For the last example, assume that μαα,mn are slow functions (μ˜αα,mn are fast functions) compared to the other expressions in Eq. (11). Under these conditions, Eq. (11) simplifies to

Sincoh(ρ,z)=α=x,ym=1N|Cα,m|2[π(d/2)2λz]2jinc2(kzd2ρ),Scoh(ρ,z)=α=x,ym=1NnmNCα,mCα,n*[π(d/2)2λz]2×jinc2(kzd2ρ)exp[jkz(ρmρn)·ρ].
This scenario models an array fed by N mutually coherent sources with different polarization states.

Scenario 3A To explore this scenario further, assume that N=2, and Cx,1=exp(jπ/4)/2, Cx,2=exp(jπ/4)/2, Cy,1=exp(jπ/4)/2, and Cy,2=exp(jπ/4)/2 (i.e., right-hand circular polarization in element 1 and left-hand circular polarization in element 2). Substituting these assumptions into Eq. (14) and simplifying yields

Sincoh(ρ,z)=S(ρ,z)=2[π(d/2)2λz]2jinc2(kzd2ρ),
which is the incoherent and physically expected result, considering that the element polarization states are orthogonal.

Scenario 3B For a slightly more interesting example, let N=2, and Cx,1=1, Cx,2=1/2, Cy,1=0, and Cy,2=1/2 (i.e., horizontal linear polarization in element 1 and 45° linear polarization in element 2). Substituting these into Eq. (14) and simplifying yields

Sincoh(ρ,z)=2[π(d/2)2λz]2jinc2(kzd2ρ),Scoh(ρ,z)=Sincoh(ρ,z)12cos[kz(ρ1ρ2)·ρ],
which is the incoherent result (always present) plus a coherent contribution due to the interference of the x components of the fields in elements 1 and 2.

3. SIMULATIONS

Here, we present simulation results to validate Scenarios 1–3 discussed above—in particular, Eqs. (12), (13), (15), and (16) and their physical interpretations. Before presenting the results, we discuss the simulation setup.

A. Setup

The simulated array had seven circular elements arranged in a regular hexagon for Scenarios 1 and 2, and two circular elements arranged side-by-side on the x axis for Scenario 3. Each circular element was d=10cm in diameter; they were spaced (center-to-center) D=10.5cm apart. The computational grid consisted of N=1024 points per side with Δ=5mm grid spacing.

For Scenarios 1 and 2, Cα,m=Cα=1/2 and μαα,mn were Gaussian, namely,

μαα,mn(ρ)=Γαα,mnexp(ρ22δαα2),
where Γαα,mn=1 and δmn (the Kronecker delta function) for Scenarios 1 and 2, respectively. Here, the δαα (1/2 times the 1/e radius) were 31.0, 7.75, and 1.94 cm to simulate the array under the three conditions described in Scenario 1. For Scenario 3, the Cα,m were the same as those discussed in Scenarios 3A and 3B, and the μαα,mn were equal to Eq. (17) with Γαα,mn=1 and δαα=31cm.

Complex transmittance screens tα,m were generated using the Monte Carlo spectral method [3538], viz.,

tα,m[i,j]=klrα,m[k,l]μ˜αα,mm[k,l]2×exp(j2πNik)exp(j2πNjl)1NΔ,
where i, j were discrete spatial indices, k, l were discrete spatial frequency indices, and rα,m was an N×N matrix of zero-mean, unit-variance, circular complex Gaussian random numbers.

For all three scenarios, tx was assumed to be statistically independent of ty. Recall that only the diagonal elements of the source field CSD dyad were required, i.e., the statistical relationship between the x and y components of the source field was irrelevant. For Scenarios 1 and 3, all of the element screens were perfectly correlated (tα,m=tα), and, for Scenario 2, they were statistically independent.

Lastly, the simulated far-zone spectral densities were computed and then averaged over 2000 propagated instances of the array source field [see Eq. (6)]. The simulated propagation distance and wavelength were z=10km and λ=632.8nm, respectively. All propagations were computed using fast Fourier transforms [39,40].

B. Results

This section presents results corresponding to Scenarios 1–3 discussed in Section C. In general, the Monte Carlo averages for the simulated far-zone spectral densities compare well to the theory. This comparison validates Eq. (11).

1. Scenario 1

Figure 1 shows the Scenario 1 simulation results. The figure is organized as follows: the left column subfigures [(a), (c), and (e)] show the theoretical spectral densities, and the right column subfigures [(b), (d), and (f)] show the simulated spectral densities. Proceeding down the rows are the results for δxx=δyy=31cm [Figs. 1(a) and 1(b)], δxx=δyy=7.75cm [Figs. 1(c) and 1(d)], and δxx=δyy=1.94cm [Figs. 1(e) and 1(f)], respectively.

 figure: Fig. 1.

Fig. 1. Scenario 1 results: (a) Sthy and (b) Ssim with δxx=δyy=31cm; (c) Sthy and (d) Ssim with δxx=δyy=7.75cm (see Visualization 1); (e) Sthy and (f) Ssim with δxx=δyy=1.94cm; (g) y=0 slices (a)–(f); and (h) x=0 slices (a)–(f).

Download Full Size | PDF

Lastly, Figs. 1(g) and 1(h) show the y=0 and x=0 slices through Figs. 1(a)1(f), respectively. The spectral density results in Figs. 1(g) and 1(h) are plotted in decibels so that all results are clearly visible on a single plot. Note that the peak spectral densities in Figs. 1(a) and 1(b) are roughly 30 times larger than those in Figs. 1(e) and 1(f).

We also included a 30 s movie (Visualization 1) showing the formation of the far-zone spectral density result in Fig. 1(d). The movie shows the first 300 array and observation field realizations, as well as the running average of the “instantaneous” spectral densities.

Overall, the agreement between the simulated and theoretical spectral densities is very good. As predicted from Eq. (12), the far-zone spectral density has three semi-distinct behaviors, which are shown here.

In row 1 of Fig. 1, the 1/e point of μ (hereafter referred to as the coherence radius) is greater than the full diameter of the array—the array is essentially fed by a uniform coherent source—resulting in a pattern similar to L.

In row 2, the coherence radius lies between the full array and element diameters d. The light emitted from each element is generally spatially coherent; however, the coherence radius is such that the interference of light emitted from different elements is only partially developed. The result is a blurred pattern that has approximately the same overall size as L.

Lastly, in row 3, the coherence radius is smaller than d. The light emitted from each element is spatially incoherent, and the far-zone spectral density resembles μ˜ (Gaussian in this case). Note that all information about the array (e.g., element shapes, diameters, spacings) is lost.

2. Scenario 2

Figure 2 shows the results for Scenario 2. The layout of the figure is exactly the same as Fig. 1. Like in the Scenario 1 results, we included a 30 s movie (Visualization 2) showing the formation of Fig. 2(d).

 figure: Fig. 2.

Fig. 2. Scenario 2 results: (a) Sthy and (b) Ssim with δxx=δyy=31cm; (c) Sthy and (d) Ssim with δxx=δyy=7.75cm (see Visualization 2); (e) Sthy and (f) Ssim with δxx=δyy=1.94cm; (g) y=0 slices (a)–(f); and (h) x=0 slices (a)–(f).

Download Full Size | PDF

Again, the agreement between the theoretical and simulated spectral densities is very good with some minor differences in the diffraction rings in Figs. 2(b) and 2(d). Being equivalent to an array fed by seven mutually incoherent sources [recall the physical interpretation of Eq. (13)], the far-zone spectral density has two semi-distinct behaviors determined by the size of the coherence radius in comparison to d.

In rows 1 and 2 of Fig. 2, the coherence radius is greater than d. The light emitted from each array element is spatially coherent but mutually incoherent with respect to all other elements. The far-zone spectral density is approximately the sum of seven Airy patterns. In row 2, the diffraction ring has nearly disappeared because the coherence radius is approximately d; however, the peak spectral density is roughly the same as in row 1. This is quite different from the results in rows 1 and 2 of Fig. 1.

Lastly, in row 3 of Fig. 2, the coherence radius is less than d. The physical situation here is equivalent to the corresponding situation in Scenario 1, and the results are generally the same (note the Gaussian shape and peak spectral density values).

3. Scenario 3

Figure 3 shows the Scenario 3 results. Figures 3(a)3(d) and 3(e)3(h) pertain to Scenarios 3A [see Eq. (15)] and 3B [see Eq. (16)], respectively. Figures 3(a) and 3(e) show the theoretical spectral densities, Figs. 3(b) and 3(f) show the simulated spectral densities, Figs. 3(c) and 3(g) show the y=0 slices through Sthy and Ssim in decibels, and Figs. 3(d) and 3(h) show the x=0 slices through Sthy and Ssim in decibels.

 figure: Fig. 3.

Fig. 3. Scenario 3 results: Scenario 3A—(a) Sthy and (b) Ssim with δxx=δyy=31cm (see Visualization 3); (c) y=0 slices (a) and (b); and (d) x=0 slices (a) and (b); Scenario 3B—(e) Sthy and (f) Ssim with δxx=δyy=31cm (see Visualization 4); (g) y=0 slices (e) and (f); (h) x=0 slices (e) and (f).

Download Full Size | PDF

Like in the results of Scenarios 1 and 2, the agreement between simulation and theory is excellent. We included two 30 s movies—Visualization 3 and Visualization 4—showing the formations of Figs. 3(b) and 3(f), respectively.

Recall that Scenario 3A concerned a mutually coherent, two-element array with right-hand circular polarization transmitted from element 1 and left-hand circular polarization transmitted from element 2. Intuitively, the far-zone spectral density should be equivalent to the spatially incoherent result, and, indeed, the incoherent pattern is evident in the first two rows of Fig. 3.

Scenario 3B again concerned a mutually coherent, two-element array; however, this time horizontal linear polarization was transmitted from element 1, and 45° linear polarization was transmitted from element 2. In this case, we should observe an interference pattern in the far-zone spectral density because of the interference of the x components of the element fields. This interference pattern is clearly visible in the last two rows of Fig. 3.

4. CONCLUSION

Here, we examined the far-zone behavior of a tiled-aperture array fed by vector partially coherent sources. As opposed to previous studies on partially coherent array beams, which used the popular Gaussian Schell-model family of sources, we derived a general form for the far-zone spectral density, assuming only that the array elements were circular and fed by optical fields with Schell-model CSD functions.

The far-zone spectral density was expressed as the sum of spatially coherent and incoherent terms. Both were convolution integrals, where the associated (coherent or incoherent) far-field array pattern was convolved with the Fourier transforms of the spatial cross-correlation functions. We explored the physical nature of this result with three examples. Lastly, we validated the far-zone spectral density expression by comparing and contrasting the Monte Carlo averages from wave-optics simulations with our analytical predictions. The simulation results were in excellent agreement with theory.

This work will find use in the prediction, simulation, and design of partially coherent beam arrays. These systems are now being developed for use in directed-energy, remote-sensing, optical-communications, and medical applications.

Acknowledgment

The views expressed in this paper are those of the authors and do not reflect the official policy or position of the US Air Force, the Department of Defense, or the US government.

REFERENCES AND NOTES

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

2. G. Gbur, “Partially coherent beam propagation in atmospheric turbulence,” J. Opt. Soc. Am. A 31, 2038–2045 (2014). [CrossRef]  

3. E. Wolf, Introduction to the Theory of Coherence and Polarization of Light (Cambridge University, 2007).

4. O. Korotkova, Random Light Beams: Theory and Applications (CRC Press, 2014).

5. G. Gbur and T. Visser, “The structure of partially coherent fields,” Prog. Opt. 55, 285–341 (2010). [CrossRef]  

6. Y. Cai, Y. Chen, J. Yu, X. Liu, and L. Liu, “Generation of partially coherent beams,” Prog. Opt. 62, 157–223 (2017). [CrossRef]  

7. P. Sprangle, B. Hafizi, A. Ting, and R. Fischer, “High-power lasers for directed-energy applications,” Appl. Opt. 54, F201–F209 (2015). [CrossRef]  

8. W. Nelson, P. Sprangle, and C. C. Davis, “Atmospheric propagation and combining of high-power lasers,” Appl. Opt. 55, 1757–1764 (2016). [CrossRef]  

9. A. Brignon, ed., Coherent Laser Beam Combining (Wiley-VCH, 2013).

10. N. R. Van Zandt, S. J. Cusumano, R. J. Bartell, S. Basu, J. E. McCrae, and S. T. Fiorino, “Comparison of coherent and incoherent laser beam combination for tactical engagements,” Opt. Eng. 51, 104301 (2012). [CrossRef]  

11. B. Lü and H. Ma, “Beam propagation properties of radial laser arrays,” J. Opt. Soc. Am. A 17, 2005–2009 (2000). [CrossRef]  

12. P. Zhou, Z. Liu, X. Xu, and X. Chu, “Comparative study on the propagation performance of coherently combined and incoherently combined beams,” Opt. Commun. 282, 1640–1647 (2009). [CrossRef]  

13. T. M. Shay, “Theory of electronically phased coherent beam combination without a reference beam,” Opt. Express 14, 12188–12195 (2006). [CrossRef]  

14. M. A. Vorontsov, “Speckle effects in target-in-the-loop laser beam projection systems,” Adv. Opt. Technol. 2, 369–395 (2013). [CrossRef]  

15. M. Vorontsov, G. Filimonov, V. Ovchinnikov, E. Polnau, S. Lachinova, T. Weyrauch, and J. Mangano, “Comparative efficiency analysis of fiber-array and conventional beam director systems in volume turbulence,” Appl. Opt. 55, 4170–4185 (2016). [CrossRef]  

16. Y. Yuan, Y. Cai, H. T. Eyyuboğlu, Y. Baykal, and J. Chen, “Propagation factor of partially coherent flat-topped beam array in free space and turbulent atmosphere,” Opt. Lasers Eng. 50, 752–759 (2012). [CrossRef]  

17. Y. Cai, Y. Chen, H. Eyyuboğlu, and Y. Baykal, “Propagation of laser array beams in a turbulent atmosphere,” Appl. Phys. B 88, 467–475 (2007). [CrossRef]  

18. Y. Zhu, D. Zhao, and X. Du, “Propagation of stochastic Gaussian-Schell model array beams in turbulent atmosphere,” Opt. Express 16, 18437–18442 (2008). [CrossRef]  

19. P. Zhou, Y. Ma, X. Wang, H. Ma, X. Xu, and Z. Liu, “Average intensity of a partially coherent rectangular flat-topped laser array propagating in a turbulent atmosphere,” Appl. Opt. 48, 5251–5258 (2009). [CrossRef]  

20. F. D. Kashani and M. Yousefi, “Analyzing the propagation behavior of coherence and polarization degrees of a phase-locked partially coherent radial flat-topped array laser beam in underwater turbulence,” Appl. Opt. 55, 6311–6320 (2016). [CrossRef]  

21. M. Yousefi, F. D. Kashani, S. Golmohammady, and A. Mashal, “Scintillation and bit error rate analysis of a phase-locked partially coherent flat-topped array laser beam in oceanic turbulence,” J. Opt. Soc. Am. A 34, 2126–2137 (2017). [CrossRef]  

22. X. Du and D. Zhao, “Propagation characteristics of stochastic electromagnetic array beams,” Appl. Phys. B 93, 901–905 (2008). [CrossRef]  

23. X. Du and D. Zhao, “Statistical properties of correlated radial stochastic electromagnetic array beams on propagation,” Opt. Commun. 282, 1993–1997 (2009). [CrossRef]  

24. X. Li, X. Ji, and F. Yang, “Beam quality of radial Gaussian Schell-model array beams,” Opt. Laser Technol. 42, 604–609 (2010). [CrossRef]  

25. Y. Yuan, Y. Cai, C. Zhao, H. T. Eyyuboğlu, and Y. Baykal, “Propagation factors of laser array beams in turbulent atmosphere,” J. Mod. Opt. 57, 621–631 (2010). [CrossRef]  

26. C. A. Balanis, Advanced Engineering Electromagnetics, 2nd ed. (Wiley, 2012), pp. 329–334.

27. C. A. Balanis, Antenna Theory: Analysis and Design, 4th ed. (Wiley, 2016), pp. 639–643.

28. C. A. Balanis, Advanced Engineering Electromagnetics, 2nd ed. (Wiley, 2012), pp. 314–323.

29. C. A. Balanis, Antenna Theory: Analysis and Design, 4th ed. (Wiley, 2016), pp. 182–191.

30. C. A. Balanis, Advanced Engineering Electromagnetics, 2nd ed. (Wiley, 2012), pp. 338–346.

31. C. A. Balanis, Antenna Theory: Analysis and Design, 4th ed. (Wiley, 2016), pp. 890–891.

32. By making this statement, we have effectively assumed that the Fourier transform of a sample function drawn from a wide-sense stationary random process exists. In general, this is not the case [1,3,34]. In the context of this work, the mathematically proper way to handle this outcome is to define the cross-spectral density (CSD) function in the array plane and propagate that CSD function to the target plane [1,4,34]. With all of this in mind, the spectral density obtained in this manner is equal to our expression given in Eq. (11).

33. J. W. Goodman, Introduction to Fourier Optics, 3rd ed. (Roberts & Company, 2005).

34. J. W. Goodman, Statistical Optics, 2nd ed. (Wiley, 2015).

35. S. Basu, M. W. Hyde, X. Xiao, D. G. Voelz, and O. Korotkova, “Computational approaches for generating electromagnetic Gaussian Schell-model sources,” Opt. Express 22, 31691–31707 (2014). [CrossRef]  

36. D. Voelz, X. Xiao, and O. Korotkova, “Numerical modeling of Schell-model beams with arbitrary far-field patterns,” Opt. Lett. 40, 352–355 (2015). [CrossRef]  

37. H. T. Yura and S. G. Hanson, “Digital simulation of an arbitrary stationary stochastic process by spectral representation,” J. Opt. Soc. Am. A 28, 675–685 (2011). [CrossRef]  

38. C. A. Mack, “Generating random rough edges, surfaces, and volumes,” Appl. Opt. 52, 1472–1480 (2013). [CrossRef]  

39. J. D. Schmidt, Numerical Simulation of Optical Wave Propagation with Examples in MATLAB (SPIE, 2010).

40. D. G. Voelz, Computational Fourier Optics: A MATLAB Tutorial (SPIE, 2011).

Supplementary Material (4)

NameDescription
Visualization 1       Figure 1(d)
Visualization 2       Figure 2(d)
Visualization 3       Figure 3(b)
Visualization 4       Figure 3(f)

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 (3)

Fig. 1.
Fig. 1. Scenario 1 results: (a)  S thy and (b)  S sim with δ x x = δ y y = 31 cm ; (c)  S thy and (d)  S sim with δ x x = δ y y = 7.75 cm (see Visualization 1); (e)  S thy and (f)  S sim with δ x x = δ y y = 1.94 cm ; (g)  y = 0 slices (a)–(f); and (h)  x = 0 slices (a)–(f).
Fig. 2.
Fig. 2. Scenario 2 results: (a)  S thy and (b)  S sim with δ x x = δ y y = 31 cm ; (c)  S thy and (d)  S sim with δ x x = δ y y = 7.75 cm (see Visualization 2); (e)  S thy and (f)  S sim with δ x x = δ y y = 1.94 cm ; (g)  y = 0 slices (a)–(f); and (h)  x = 0 slices (a)–(f).
Fig. 3.
Fig. 3. Scenario 3 results: Scenario 3A—(a)  S thy and (b)  S sim with δ x x = δ y y = 31 cm (see Visualization 3); (c)  y = 0 slices (a) and (b); and (d)  x = 0 slices (a) and (b); Scenario 3B—(e)  S thy and (f)  S sim with δ x x = δ y y = 31 cm (see Visualization 4); (g)  y = 0 slices (e) and (f); (h)  x = 0 slices (e) and (f).

Equations (18)

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

E ( r , ω ) = θ ^ j k 4 π r e j k r [ L x ( r , ω ) sin ϕ + L y ( r , ω ) cos ϕ ] + ϕ ^ j k 4 π r e j k r [ L x ( r , ω ) cos θ cos ϕ L y ( r , ω ) cos θ sin ϕ ] ,
L x , y ( r , ω ) = 2 S E y , x src ( ρ , ω ) e j k r ^ · ρ d 2 ρ ,
S ( r , ω ) E ( r , ω ) · E * ( r , ω ) = k 2 ( 4 π r ) 2 ( cos 2 θ cos 2 ϕ + sin 2 ϕ ) L x ( r , ω ) L x * ( r , ω ) + k 2 ( 4 π r ) 2 ( cos 2 θ sin 2 ϕ + cos 2 ϕ ) L y ( r , ω ) L y * ( r , ω ) + k 2 ( 4 π r ) 2 sin 2 θ sin ( 2 ϕ ) Re [ L x ( r , ω ) L y * ( r , ω ) ] ,
S ( ρ , z , ω ) k 2 ( 4 π z ) 2 [ L x ( ρ , z , ω ) L x * ( ρ , z , ω ) + L y ( ρ , z , ω ) L y * ( ρ , z , ω ) ] ,
L x , y ( ρ , z , ω ) 2 S E y , x src ( ρ , ω ) exp ( j k z ρ · ρ ) d 2 ρ .
E src ( ρ , ω ) = m = 1 N circ ( | ρ ρ m | d / 2 ) × [ x ^ C x , m ( ω ) t x , m ( ρ , ω ) + y ^ C y , m ( ω ) t y , m ( ρ , ω ) ] ,
E α src ( ρ 1 ) E α src * ( ρ 2 ) = m = 1 N n = 1 N C α , m C α , n * P m ( ρ 1 ) P n ( ρ 2 ) t α , m ( ρ 1 ) t α , n * ( ρ 2 ) = m = 1 N n = 1 N C α , m C α , n * P m ( ρ 1 ) P n ( ρ 2 ) μ α α , m n ( ρ 1 ρ 2 ) ,
S ( ρ , z ) = 1 ( λ z ) 2 α = x , y m = 1 N n = 1 N C α , m C α , n * μ α α , m n ( t ) × [ P m ( s ) P n ( s t ) d 2 s ] exp ( j k z t · ρ ) d 2 t = 1 ( λ z ) 2 α = x , y m = 1 N n = 1 N C α , m C α , n * × μ α α , m n ( t ) Λ m n ( t ) exp ( j 2 π t · ρ λ z ) d 2 t ,
S ( ρ , z ) = 1 ( λ z ) 2 α = x , y m = 1 N n = 1 N C α , m C α , n * × μ ˜ α α , m n ( f ) Λ ˜ m n ( f ρ λ z ) d 2 f ,
S ( ρ , z ) = α = x , y m = 1 N n = 1 N C α , m C α , n * [ π ( d / 2 ) 2 λ z ] 2 jinc 2 ( 2 π d 2 ρ ) × exp [ j 2 π ( x ρ m ρ n ) · ρ ] μ ˜ α α , m n ( ρ ) | ρ = ρ λ z ,
S ( ρ , z ) = S incoh ( ρ , z ) + S coh ( ρ , z ) , S incoh ( ρ , z ) = α = x , y m = 1 N | C α , m | 2 [ π ( d / 2 ) 2 λ z ] 2 × jinc 2 ( 2 π d 2 ρ ) μ ˜ α α , m m ( ρ ) | ρ = ρ λ z , S coh ( ρ , z ) = α = x , y m = 1 N n m N C α , m C α , n * [ π ( d / 2 ) 2 λ z ] 2 jinc 2 ( 2 π d 2 ρ ) × exp [ j 2 π ( ρ m ρ n ) · ρ ] μ ˜ α α , m n ( ρ ) | ρ = ρ λ z .
S incoh ( ρ , z ) = α = x , y | C α | 2 N [ π ( d / 2 ) 2 λ z ] 2 jinc 2 ( 2 π d 2 ρ ) μ ˜ α α ( ρ ) | ρ = ρ λ z , S coh ( ρ , z ) = α = x , y | C α | 2 m = 1 N n m N [ π ( d / 2 ) 2 λ z ] 2 jinc 2 ( 2 π d 2 ρ ) × exp [ j 2 π ( ρ m ρ n ) · ρ ] μ ˜ α α ( ρ ) | ρ = ρ λ z , S ( ρ , z ) = α = x , y | C α | 2 { N 2 [ π ( d / 2 ) 2 λ z ] 2 jinc 2 ( 2 π d 2 ρ ) × | 1 N m = 1 N exp ( j 2 π ρ m · ρ ) | 2 } μ ˜ α α ( ρ ) | ρ = ρ λ z .
S incoh ( ρ , z ) = S ( ρ , z ) = α = x , y | C α | 2 { N [ π ( d / 2 ) 2 λ z ] 2 jinc 2 ( 2 π d 2 ρ ) } μ ˜ α α ( ρ ) | ρ = ρ λ z .
S incoh ( ρ , z ) = α = x , y m = 1 N | C α , m | 2 [ π ( d / 2 ) 2 λ z ] 2 jinc 2 ( k z d 2 ρ ) , S coh ( ρ , z ) = α = x , y m = 1 N n m N C α , m C α , n * [ π ( d / 2 ) 2 λ z ] 2 × jinc 2 ( k z d 2 ρ ) exp [ j k z ( ρ m ρ n ) · ρ ] .
S incoh ( ρ , z ) = S ( ρ , z ) = 2 [ π ( d / 2 ) 2 λ z ] 2 jinc 2 ( k z d 2 ρ ) ,
S incoh ( ρ , z ) = 2 [ π ( d / 2 ) 2 λ z ] 2 jinc 2 ( k z d 2 ρ ) , S coh ( ρ , z ) = S incoh ( ρ , z ) 1 2 cos [ k z ( ρ 1 ρ 2 ) · ρ ] ,
μ α α , m n ( ρ ) = Γ α α , m n exp ( ρ 2 2 δ α α 2 ) ,
t α , m [ i , j ] = k l r α , m [ k , l ] μ ˜ α α , m m [ k , l ] 2 × exp ( j 2 π N i k ) exp ( j 2 π N j l ) 1 N Δ ,
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.