|
In the early 1990s, Siegman proposed the factor as a metric to assess laser beam quality.1 Siegman defined the factor as the ratio of a test beam’s space-beamwidth product to that of an ideal Gaussian beam. In so doing, he showed that the test beam’s spot size (in the direction as a function of range ) obeyed the quadratic propagation formula: where is the wavelength, is the location of the waist plane, and is the test beam’s spot size in the waist plane, i.e., is the test beam’s waist. Expressed in this form, the physical interpretation of becomes clear: is the test beam’s far-zone (FZ) beam spread relative to an ideal Gaussian beam with a waist equal to that of the test beam.Siegman’s quickly gained acceptance and became the laser industry’s standard for assessing beam quality. It was soon extended to include hard-aperture,2,3 vortex,4 and stochastic, or partially coherent beams.5–8 The latter, which are germane to this work, focus exclusively on scalar partially coherent beams. There have been several papers that derived the factors for specific electromagnetic (EM) or vector partially coherent beams, e.g., EM Gaussian Schell-model (EGSM) beams.9–12 However, none to date have derived an expression for the factor of a general EM Schell-model beam. This analysis is useful considering the wide-spread use of to assess beam quality and a large number of vector Schell-model sources that have been developed for applications, such as free-space/underwater optical communications, directed energy, optical trapping/tweezers, etc.13–26 For directed energy, in particular, the use of to describe beam quality is pervasive, and there has been recent work in modeling dynamic, stochastic, noncommon-path phase errors in high-energy-laser systems in terms of .27 Most importantly, having a relation for the factor of a general EM Schell-model source allows one to consider beam quality in the design of new vector Schell-model beams. For the reasons stated above, here, we extend the scalar partially coherent source analysis presented in Refs. 5 and 6 to vector Schell-model beams. Starting with Siegman’s definition, we derive an expression for the factor of a vector partially coherent source in terms of the vector component beam quality factors. We then apply the results in Refs. 5 and 6 to derive a simple and physical relation for the factor of a general vector Schell-model beam. We present examples, where we calculate for two EM Schell-model sources using our relation and describe how to synthesize vector Schell-model beams in terms of specified, desired . In Sec. 2, we present Monte Carlo simulations, where we generate two examples of EM Schell-model sources in terms of to validate our analysis. We then study the convergence of the stochastic vector field realizations to the specified, desired . Last, we conclude this paper with a summary of the work presented herein. 1.TheoryIn this section, we first review the scalar theory found in Refs. 5 and 6. Next, we calculate the factors for two example vector Schell-model sources. We then describe how to generate vector Schell-model beams in terms of specified, desired . 1.1.Siegman’sAs defined by Siegman,1 the factor in the direction is as follows: with a similar definition for . The normalized beam widths and are as follows: where is the cross-spectral density (CSD) matrix17,28,29 of the beam at its waist location, is the trace, and are the spectral densities (SDs)17,28,29 of the beam’s polarization components.Continuing to define the symbols in Eq. (3), is the Fourier transform of , i.e.: where is as follows: where and . Note that the denominators of and are equal due to Parseval’s theorem. As such, we define and use the symbol hereafter to represent this normalization factor.Last, and are the normalized beam “first moments” at the beam’s waist location and in the spatial frequency domain, respectively, such that Following the scalar partially coherent source analysis presented in Refs. 5 and 6, we choose the axis—the mean propagation direction of the random beam—such that . This permits us to write Eq. (3) as follows: where and are as follows: Substituting Eq. (7) into Eq. (2) and simplifying produce as follows: The bracketed quantities in Eq. (9) are the vector component beam quality factors, and therefore, where are the beam quality factors for the vector components.1.2.Prior Scalar AnalysisThrough extensive analysis and clever mathematics, Refs. 5 and 6 showed that for a scalar beam of any state of coherence is as follows: where is given in Eq. (8) but computed in the source plane, is as follows: and is The and denote that and after computing the partial derivatives, respectively.For Schell-model sources,17,28 Ref. 6 showed that Eq. (11) further reduces to where is the beam quality factor of the corresponding coherent source, , and is related to the source’s spectral autocorrelation function by Both and are real functions. Since and , , , , and . If is real, then , and can be substituted directly into Eq. (14).1.3.New Vector AnalysisSubstituting in Eq. (14) for the vector component beam quality factors in Eq. (10) and simplifying produce as follows: Because in Eq. (16) is expressed in terms of vector component quantities (i.e., , , , and ), it is easy to use, in practice, to calculate the beam quality factor. It is not, however, the most simplified or physical form. Noting that the bracketed quantity on the first line of Eq. (16) is equal to via Eq. (7), expanding, and then simplifying Eq. (16) yields as follows: Last, substituting in , recalling Eq. (7), and simplifying, the desired result is obtained as follows:Equation (18)—in particular, the last line of Eq. (18)—is the main analytical result of this paper and generalizes the scalar result presented in Ref. 6. Its form is very similar to the scalar result in that depends on the coherence of the source only through the second derivatives of the vector correlation functions evaluated at . Rather intuitively, it differs in the fact that the “coherence contributions,” and , are weighted by the fraction of power in the associated vector component. This being the case, one is tempted to try to manipulate Eq. (18) into the sum of weighted, component, beam quality factors . Unfortunately, this simple, physical form for is spoiled by cross terms, which are evident in Eq. (16). 1.4.ExamplesHere, we calculate the for two example, vector Schell-model sources. We then simulate them in terms of in Sec. 2. 1.4.1.EM Gaussian Schell-model sourceThe elements of the CSD matrix for an EGSM source are as follows: where , is the amplitude of the field component, and is the RMS width of the component SD. Also in Eq. (19), is the complex correlation coefficient between the and field components and is the RMS width of the cross-correlation function , i.e., the cross-correlation function of the and field components.17,29 The EGSM source parameters—, , , and —must satisfy the realizability conditions derived in Refs. 17 and 30. Since only the diagonal elements of the CSD matrix are required to compute , the criterion is the most relevant here.To compute , we use Eq. (16). The normalized beam widths of the and components of the EGSM source, computed using Eq. (8), are . The powers in the and components of the source are computed using the denominators in Eq. (3) and are . The corresponding coherent source to the EGSM source defined in Eq. (19) is as follows: This CSD function describes a vector source composed of horizontally and vertically polarized, spatially coherent, correlated, Gaussian beams. Since the and vector components of the source are Gaussian beams, the component beam quality factors are . Last, the component spectral correlation functions are real; therefore, and the second derivatives in Eq. (16) evaluate toSubstituting the quantities discussed in the previous paragraph into Eq. (16) and simplifying produce as follows: This expression is equal to the for an EGSM source derived in Refs. 910.11.–12 using different methods. Note that letting either or yields the for a scalar GSM source first derived in Ref. 6. We generate an EGSM source with a specified in Sec. 2.1.4.2.Vector optical coherence latticesAnother popular and relatively new example of a vector partially coherent source is the so-called vector optical coherence lattice (VOCL).14 The CSD matrix elements for a VOCL take the form: where , is a first-order Bessel function of the first kind, and is a vector that points from the origin to the ’th node of the “coherence lattice.” The other symbols have been defined previously and generally have the same physical interpretation as the EGSM source. One should immediately recognize that the spectral cross-correlation function in Eq. (23) is equivalent to the FZ pattern (spatial Fourier transform) of an array of circular transmitters.To find the beam quality factor, we again use Eq. (16). The , , and are the same as in the EGSM source example. The second derivative of is much more difficult. Here, for analytical convenience, we assume that the lattice nodes are symmetric about the origin. This assumption means that is real, and thus, . After much calculus and algebra (details in Sec. 4 Appendix A), the second derivatives in Eq. (16) evaluate to Substituting , , ; Eq. (24) into Eq. (16); and simplifying produce as follows: We generate and analyze a VOCL, in terms of , in Sec. 2.1.5.Generating Stochastic Vector Fields in Terms ofAn instance of a Schell-model source can be generated by filtering a two-dimensional (2-D) array of circular complex Gaussian random numbers. For computational efficiency, it is best to perform the filtering in the spectral domain using the convolution theorem. This process has been described in the literature many times.31–37 Here, we present the necessary equations to implement the technique. Any type of vector Schell-model source can be formed by the following optical field realization:31,33 where is the complex amplitude and is the stochastic complex transmittance screen for the component of the field, respectively. is formed from correlated, circular complex Gaussian random numbers. Here, we have assumed a Gaussian shape for the source, considering that Sec. 1.4 examples also have Gaussian shapes. In general, the source can take any shape.Hereafter, we specialize the mathematics to generate the sources discussed in Sec. 1.4. The procedure for synthesizing any other type of vector Schell-model source is the same as presented here, only the mathematical details change. Taking the cross-correlation of Eq. (26) with and comparing the resulting expression to Eqs. (19) and (23) yields the following equalities: , , and Because potentially , and , in general, must be generated from correlated Gaussian random numbers. To see this and reveal the conditions on the values of and , we note that an instance of is generated by31,32 where , are the numbers of grid points in the , directions, , are the lengths of the grid in the , directions in meters, and is the grid spacing. In Eq. (28), is an grid of zero-mean, unit-variance circular complex Gaussian random numbers and is the spatial power spectrum of , i.e., the Fourier transform of the autocorrelation of . For the examples discussed in Sec. 1.4: where is the circle function defined in Ref. 38. We note that Eq. (28) is in the form of a discrete, inverse Fourier transform; thus, can be generated quickly and efficiently using the fast Fourier transform algorithm.Using Eqs. (28) and (29), a with a specified is produced. Physically, this leads to an EGSM source or VOCL with the correct, diagonal, CSD matrix elements. Recall that depends only on these elements, and therefore, one could ignore any correlation between the and vector components (i.e., produce instances of and from independent Gaussian random numbers) and still produce a beam with the desired . Doing this step results in a beam that is randomly, partially, or fully linearly polarized. To generate a vector Schell-model source with a general polarization state, one must control the off-diagonal elements of the CSD matrix as well. This requires examination of the cross-correlation of with , namely, The moment , where is the correlation coefficient between the and random numbers and is the discrete Dirac delta function. Substituting this into Eq. (30) and simplifying produce as follows: The must equal the cross-power spectra, or the Fourier transforms of the expressions in Eq. (27), viz., This step produces the following relation for EGSM sources: From here, it is quite clear that where . These conditions were first derived in Ref. 31.The VOCL conditions on and are derived from Because of this expression, it is not likely that conditions similar to Eq. (34) can be derived for a VOCL. Two rather trivial, but physically important conditions can be derived by letting :
These same conditions apply to many other vector Schell-model sources, e.g., EM multi-Gaussian Schell-model17,21 and EM Bessel-Gaussian Schell-model sources.22,32 The last step is to express the EGSM source and VOCL conditions derived above in terms of . The EGSM source and VOCL component beam quality factors are as follows: Referring back to Eqs. (22) and (25), substituting Eq. (36) into those expressions and inverting produces as follows:We assume that and are given. This makes sense as we would expect the on-axis intensity and size of the source to be known. In addition, since generally (or equivalently, ), one of those must be given. Without loss of generality, we assume is known. Simplifying Eq. (37) further produces as follows: Last, substituting Eq. (36) into Eq. (38) and simplifying yields for an EGSM source, and for a VOCL. We note that the locations of the coherence lattice nodes must be known to find .In summary, to produce an EGSM source or VOCL with a desired :
Since the generated fields are stochastic, this procedure will produce an EGSM source or VOCL with a desired, “on-average” . 2.SimulationHere, we present simulation results to validate the analysis in the previous section. We also examine the convergence of the stochastic vector fields to the desired . Before presenting the results, we discuss the details of the setup. 2.1.SetupFor these Monte Carlo simulations, we used computational grids with points per side. The grid spacings were chosen such that resulting in 3 and 0.97 mm for the EGSM source and VOCL simulations, respectively. These spacings easily satisfied the Nyquist sampling criterion derived for Gaussian signals in Ref. 39. The EGSM source and VOCL parameters are given in Table 1. For the VOCL, the coherence lattice was rectangular with 5 rows and 4 columns of nodes spaced apart. Table 1EGSM source and VOCL parameters.
We generated 20,000 EGSM source and VOCL field instances. From these, we computed the near-zone (NZ) (i.e., source plane) and FZ Stokes parameters, , and . The Stokes parameters in terms of the CSD matrix elements are as follows:17,29 For and , the , , , and [see Eq. (3)] were computed using trapezoidal numerical integration.To verify that, we produced a source with the correct NZ and FZ polarization properties, we compared the simulated Stokes parameters with the corresponding theoretical quantities. The theoretical expressions for the EGSM source and VOCL NZ Stokes parameters were computed using Eqs. (19) and (23) and the values in Table 1, respectively. We computed the theoretical FZ Stokes parameters using those same equations and the Fourier transform in Eq. (4) resulting in The integral in the VOCL can be evaluated in closed form using Mellin transform techniques.40 The resulting expression is an infinite series of hypergeometric functions.41 Computing the result using this analytical relation is very slow; therefore, we evaluated the integral directly using numerical quadrature.2.2.ResultsFigures 1 and 2 show the EGSM source and VOCL results, respectively. The figures are organized as follows: the Stokes parameters are displayed in the first four rows—, , , and , respectively. The first two columns show the NZ results, with column 1 (the left column) showing the theoretical (Thy) results and column 2 (the right column) showing the simulated (Sim) results. Columns 3 and 4 show the FZ results, with an identical left-right arrangement of Thy (column 3) and Sim (column 4) results. Note that we have added row and column headings to Figs. 1 and 2 to aid the reader. Last, the fifth rows [Figs. 1(q) and 2(q)] show the convergence of the simulated and to the corresponding theoretical values versus Monte Carlo trial number. Overall, the agreement between simulation and theory is excellent. We note that the visually conspicuous differences between the and Thy and Sim VOCL results in Fig. 2 are in fact quantitatively small (see the associated color bars above the subfigures). These discrepancies arise because of our choice of color scale and Thy . Running more Monte Carlo trials would reduce these errors; however, and Sim will never be identically zero for that would require an infinite number of trials. The simulated and converge to the desired values in approximately 1,000 trials. This finding is consistent with the scalar Schell-model beam results in Ref. 27. Figures 1 and 2 validate the theoretical analysis presented in Sec. 1. 3.ConclusionHere, we derived the factor for a general vector Schell-model beam. Starting with Siegman’s definition, we found an expression for the factor of a vector partially coherent beam in terms of its vector component beam quality factors. Then, applying the prior scalar analysis, we derived a physical expression for the factor of a general vector Schell-model source. As an example, we computed the beam quality factors for two EM Schell-model beams found in the literature using our new relation and described how to synthesize vector Schell-model beams in terms of specified, desired . To validate our analysis, we performed Monte Carlo simulations, where we generated two partially coherent sources. The simulated results were found to be in excellent agreement with the corresponding theory, and convergence to the specified, desired occurred within approximately 1000 vector field realizations. Although not performed here because of equipment availability, experimental synthesis and subsequent measurement of the factor for vector Schell-model sources can be performed using optical setups described in Refs. 1516.17.–18, 32, and 4243.–44. The analysis presented in this paper will be useful in the design of vector Schell-model sources for applications from optical communications and directed energy to atomic optics and optical tweezers. 4.Appendix A: Derivation of Eq. (24)Starting with we apply the derivative multiplication rule twice to yield as follows: The first derivatives in Eq. (45) are odd functions of ; therefore, evaluating them at is trivially zero. Simplifying Eq. (45) produces as follows:We now evaluate the second derivatives in Eq. (46) separately. We begin with the second derivative of the coherence lattice term. Bringing the derivatives inside the summation and evaluating the second derivative of the exponential yields as follows: To evaluate the second derivative of the jinc function and argument, we first let the argument . The associated second derivative becomes We now use the Bessel function identity: to simplify Eq. (48) to Equation (50) can be further simplified by using the Bessel function identity: yielding In going from line 1 to 2 in Eq. (52), we used the Bessel function identity , and in going from line 2 to 3, we used Eq. (49).Continuing, we apply the derivative multiplication and chain rules producing The derivative of can be found by applying Eqs. (49) and (51), such thatWe now evaluate the first and second derivatives of . The first derivative of is found by applying the chain rule, namely, The second derivative of is found by applying the multiplication and chain rules: Substituting Eqs. (55) and (56) into Eq. (54) and simplifying produce as follows:Last, evaluating Eq. (57) at yields as follows: We obtain the second line of Eq. (24)—the desired result—by substituting Eqs. (58) and (47) into Eq. (46), i.e.: AcknowledgmentsThe 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. The authors have no conflicts of interest. ReferencesA. E. Siegman,
“New developments in laser resonators,”
Proc. SPIE, 1224 2
–14
(1990). https://doi.org/10.1117/12.18425 PSISDG 0277-786X Google Scholar
R. Martínez-Herrero and P. M. Mejías,
“Second-order spatial characterization of hard-edge diffracted beams,”
Opt. Lett., 18 1669
–1671
(1993). https://doi.org/10.1364/OL.18.001669 OPLEDP 0146-9592 Google Scholar
R. Martínez-Herrero, P. M. Mejías and M. Arias,
“Parametric characterization of coherent, lowest-order Gaussian beams propagating through hard-edged apertures,”
Opt. Lett., 20 124
–126
(1995). https://doi.org/10.1364/OL.20.000124 OPLEDP 0146-9592 Google Scholar
S. Ramee and R. Simon,
“Effect of holes and vortices on beam quality,”
J. Opt. Soc. Am. A, 17 84
–94
(2000). https://doi.org/10.1364/JOSAA.17.000084 JOAOD6 0740-3232 Google Scholar
F. Gori, M. Santarsiero and A. Sona,
“The change of width for a partially coherent beam on paraxial propagation,”
Opt. Commun., 82
(3), 197
–203
(1991). https://doi.org/10.1016/0030-4018(91)90444-I OPCOB8 0030-4018 Google Scholar
M. Santarsiero et al.,
“Spreading properties of beams radiated by partially coherent Schell-model sources,”
J. Opt. Soc. Am. A, 16 106
–112
(1999). https://doi.org/10.1364/JOSAA.16.000106 JOAOD6 0740-3232 Google Scholar
B. Zhang, X. Chu and Q. Li,
“Generalized beam-propagation factor of partially coherent beams propagating through hard-edged apertures,”
J. Opt. Soc. Am. A, 19 1370
–1375
(2002). https://doi.org/10.1364/JOSAA.19.001370 JOAOD6 0740-3232 Google Scholar
X. Chu, B. Zhang and Q. Wen,
“Generalized factor of a partially coherent beam propagating through a circular hard-edged aperture,”
Appl. Opt., 42 4280
–4284
(2003). https://doi.org/10.1364/AO.42.004280 APOPAI 0003-6935 Google Scholar
S. Zhu, Y. Cai and O. Korotkova,
“Propagation factor of a stochastic electromagnetic Gaussian Schell-model beam,”
Opt. Express, 18 12587
–12598
(2010). https://doi.org/10.1364/OE.18.012587 OPEXFF 1094-4087 Google Scholar
S. Zhu and Y. Cai,
“-factor of a stochastic electromagnetic beam in a Gaussian cavity,”
Opt. Express, 18 27567
–27581
(2010). https://doi.org/10.1364/OE.18.027567 OPEXFF 1094-4087 Google Scholar
S. Zhu and Y. Cai,
“-factor of a truncated electromagnetic Gaussian Schell-model beam,”
Appl. Phys. B, 103 971
–984
(2011). https://doi.org/10.1007/s00340-011-4417-3 Google Scholar
S. Zhu et al.,
“State of polarization and propagation factor of a stochastic electromagnetic beam in a gradient-index fiber,”
J. Opt. Soc. Am. A, 30 2306
–2313
(2013). https://doi.org/10.1364/JOSAA.30.002306 JOAOD6 0740-3232 Google Scholar
Y. Chen et al.,
“Vector Hermite-Gaussian correlated Schell-model beam,”
Opt. Express, 24 15232
–15250
(2016). https://doi.org/10.1364/OE.24.015232 OPEXFF 1094-4087 Google Scholar
C. Liang et al.,
“Vector optical coherence lattices generating controllable far-field beam profiles,”
Opt. Express, 25 9872
–9885
(2017). https://doi.org/10.1364/OE.25.009872 OPEXFF 1094-4087 Google Scholar
Y. Cai, Y. Chen and F. Wang,
“Generation and propagation of partially coherent beams with nonconventional correlation functions: a review,”
J. Opt. Soc. Am. A, 31 2083
–2096
(2014). https://doi.org/10.1364/JOSAA.31.002083 JOAOD6 0740-3232 Google Scholar
Y. Cai et al.,
“Generation of partially coherent beams,”
Prog. Opt., 62 157
–223
(2017). https://doi.org/10.1016/bs.po.2016.11.001 POPTAN 0079-6638 Google Scholar
O. Korotkova, Random Light Beams: Theory and Applications, CRC, Boca Raton, Florida
(2014). Google Scholar
Vectorial Optical Fields, World Scientific, Hackensack, New Jersey
(2014). Google Scholar
H. Mao et al.,
“Self-steering partially coherent vector beams,”
Opt. Express, 27 14353
–14368
(2019). https://doi.org/10.1364/OE.27.014353 OPEXFF 1094-4087 Google Scholar
W. Wang et al.,
“Recent progress on the unified theory of polarization and coherence for stochastic electromagnetic fields,”
Proc. SPIE, 10612 10612
(2018). https://doi.org/10.1117/12.2306368 PSISDG 0277-786X Google Scholar
Z. Mei, O. Korotkova and E. Shchepakina,
“Electromagnetic multi-Gaussian Schell-model beams,”
J. Opt., 15
(2), 025705
(2013). https://doi.org/10.1088/2040-8978/15/2/025705 Google Scholar
S. Avramov-Zamurovic et al.,
“Experimental study of electromagnetic Bessel-Gaussian Schell model beams propagating in a turbulent channel,”
Opt. Commun., 359 207
–215
(2016). https://doi.org/10.1016/j.optcom.2015.09.078 OPCOB8 0030-4018 Google Scholar
S. Avramov-Zamurovic et al.,
“Polarization-induced reduction in scintillation of optical beams propagating in simulated turbulent atmospheric channels,”
Waves Random Complex Media, 24
(4), 452
–462
(2014). https://doi.org/10.1080/17455030.2014.944242 Google Scholar
O. Korotkova,
“Scintillation index of a stochastic electromagnetic beam propagating in random media,”
Opt. Commun., 281
(9), 2342
–2348
(2008). https://doi.org/10.1016/j.optcom.2007.12.047 OPCOB8 0030-4018 Google Scholar
O. Korotkova,
“Changes in the intensity fluctuations of a class of random electromagnetic beams on propagation,”
J. Opt. A: Pure Appl. Opt., 8
(1), 30
–37
(2006). https://doi.org/10.1088/1464-4258/8/1/005 Google Scholar
X. Xiao and D. Voelz,
“Wave optics simulation of partially coherent and partially polarized beam propagation in turbulence,”
Proc. SPIE, 7464 74640T
(2009). https://doi.org/10.1117/12.829336 PSISDG 0277-786X Google Scholar
IV M. W. Hyde and M. F. Spencer,
“Modeling the effects of high-energy-laser beam quality using scalar Schell-model sources,”
Proc. SPIE, 10981 109810O
(2019). https://doi.org/10.1117/12.2518648 PSISDG 0277-786X Google Scholar
L. Mandel and E. Wolf, Optical Coherence and Quantum Optics, Cambridge University, New York
(1995). Google Scholar
E. Wolf, Introduction to the Theory of Coherence and Polarization of Light, Cambridge University, New York
(2007). Google Scholar
F. Gori et al.,
“Realizability condition for electromagnetic Schell-model sources,”
J. Opt. Soc. Am. A, 25 1016
–1021
(2008). https://doi.org/10.1364/JOSAA.25.001016 JOAOD6 0740-3232 Google Scholar
S. Basu et al.,
“Computational approaches for generating electromagnetic Gaussian Schell-model sources,”
Opt. Express, 22 31691
–31707
(2014). https://doi.org/10.1364/OE.22.031691 OPEXFF 1094-4087 Google Scholar
M. W. Hyde et al.,
“Generation of vector partially coherent optical sources using phase-only spatial light modulators,”
Phys. Rev. Appl., 6 064030
(2016). https://doi.org/10.1103/PhysRevApplied.6.064030 PRAHB2 2331-7019 Google Scholar
D. Voelz, X. Xiao and O. Korotkova,
“Numerical modeling of Schell-model beams with arbitrary far-field patterns,”
Opt. Lett., 40 352
–355
(2015). https://doi.org/10.1364/OL.40.000352 OPLEDP 0146-9592 Google Scholar
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). https://doi.org/10.1364/JOSAA.28.000675 JOAOD6 0740-3232 Google Scholar
C. A. Mack,
“Generating random rough edges, surfaces, and volumes,”
Appl. Opt., 52 1472
–1480
(2013). https://doi.org/10.1364/AO.52.001472 APOPAI 0003-6935 Google Scholar
M. W. Hyde et al.,
“Generating partially coherent Schell-model sources using a modified phase screen approach,”
Opt. Eng., 54
(12), 120501
(2015). https://doi.org/10.1117/1.OE.54.12.120501 Google Scholar
IV M. W. Hyde et al.,
“Producing any desired far-field mean irradiance pattern using a partially-coherent Schell-model source,”
J. Opt., 17
(5), 055607
(2015). https://doi.org/10.1088/2040-8978/17/5/055607 Google Scholar
J. W. Goodman, Introduction to Fourier Optics, 3rd ed.Roberts & Company, Englewood, Colorado
(2005). Google Scholar
J. D. Schmidt, Numerical Simulation of Optical Wave Propagation with Examples in MATLAB, SPIE Press, Bellingham, Washington
(2010). Google Scholar
R. J. Sasiela, Electromagnetic Wave Propagation in Turbulence: Evaluation and Application of Mellin Transforms, 2nd ed.SPIE Press, Bellingham, Washington
(2007). Google Scholar
I. S. Gradshteyn and I. M. Ryzhik, Table of Integrals, Series, and Products, 7th ed.Academic Press, Burlington, Massachusetts
(2007). Google Scholar
C. Rosales-Guzmán, B. Ndagano and A. Forbes,
“A review of complex vector light fields and their applications,”
J. Opt., 20 123001
(2018). https://doi.org/10.1088/2040-8986/aaeb7d Google Scholar
C. Rosales-Guzmán, N. Bhebhe and A. Forbes,
“Simultaneous generation of multiple vector beams on a single SLM,”
Opt. Express, 25 25697
–25706
(2017). https://doi.org/10.1364/OE.25.025697 OPEXFF 1094-4087 Google Scholar
Z. Chen et al.,
“Complete shaping of optical vector beams,”
Opt. Express, 23 17701
–17710
(2015). https://doi.org/10.1364/OE.23.017701 OPEXFF 1094-4087 Google Scholar
BiographyMilo W. Hyde IV received his BS degree in computer engineering from Georgia Tech in 2001 and his MS degree and PhD in electrical engineering from AFIT in 2006 and 2010, respectively. Currently, he is an adjunct professor in the Department of Electrical and Computer Engineering at AFIT. He is a senior member of IEEE and SPIE. He is also a member of OSA and DEPS. Mark F. Spencer is the principal investigator for the Aero Effects and Beam Control Program at the Air Force Research Laboratory, Directed Energy Directorate. He is also an adjunct assistant professor of optical sciences and engineering within the Department of Engineering Physics at the Air Force Institute of Technology. In addition to being a senior member of SPIE, he is a regular member of the Optical Society and the Directed Energy Professional Society. |