Open Access
10 July 2019 M2 factor of a vector Schell-model beam
Author Affiliations +
Abstract
Extending existing scalar Schell-model source work, we derive the M2 factor for a general electromagnetic or vector Schell-model source to assess beam quality. In particular, we compute the M2 factors for two vector Schell-model sources found in the literature. We then describe how to synthesize vector Schell-model beams in terms of specified, desired M2 and present Monte Carlo simulation results to validate our analysis.

In the early 1990s, Siegman proposed the M2 factor as a metric to assess laser beam quality.1 Siegman defined the M2 factor as the ratio of a test beam’s space-beamwidth product to that of an ideal Gaussian TEM00 beam. In so doing, he showed that the test beam’s spot size Wx(z) (in the x direction as a function of range z) obeyed the quadratic propagation formula:

Eq. (1)

Wx2(z)=Wx2(z0)+Mx4λ2π2Wx2(z0)(zz0)2,
where λ is the wavelength, z0 is the z location of the waist plane, and Wx(z0) is the test beam’s spot size in the waist plane, i.e., Wx(z0) is the test beam’s waist. Expressed in this form, the physical interpretation of M2 becomes clear: M21 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 M2 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.58 The latter, which are germane to this work, focus exclusively on scalar partially coherent beams. There have been several papers that derived the M2 factors for specific electromagnetic (EM) or vector partially coherent beams, e.g., EM Gaussian Schell-model (EGSM) beams.912 However, none to date have derived an expression for the M2 factor of a general EM Schell-model beam.

This analysis is useful considering the wide-spread use of M2 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.1326 For directed energy, in particular, the use of M2 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 M2.27 Most importantly, having a relation for the M2 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 M2 analysis presented in Refs. 5 and 6 to vector Schell-model beams. Starting with Siegman’s M2 definition, we derive an expression for the M2 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 M2 factor of a general vector Schell-model beam. We present examples, where we calculate M2 for two EM Schell-model sources using our M2 relation and describe how to synthesize vector Schell-model beams in terms of specified, desired M2.

In Sec. 2, we present Monte Carlo simulations, where we generate two examples of EM Schell-model sources in terms of M2 to validate our analysis. We then study the convergence of the stochastic vector field realizations to the specified, desired M2. Last, we conclude this paper with a summary of the work presented herein.

1.

Theory

In this section, we first review the scalar M2 theory found in Refs. 5 and 6. Next, we calculate the M2 factors for two example vector Schell-model sources. We then describe how to generate vector Schell-model beams in terms of specified, desired M2.

1.1.

Siegman’s M2

As defined by Siegman,1 the M2 factor in the x direction Mx2 is as follows:

Eq. (2)

Mx2=4πσxσfx,
with a similar definition for My2. The normalized beam widths σx and σfx are as follows:

Eq. (3)

σx2=(xx)2Tr[W_(ρ,ρ)]d2ρTr[W_(ρ,ρ)]d2ρ=(xx)2Sx(ρ)d2ρSx(ρ)d2ρ+Sy(ρ)d2ρ+(xx)2Sy(ρ)d2ρSx(ρ)d2ρ+Sy(ρ)d2ρσfx2=(fxfx)2Tr[W˜_(f,f)]d2fTr[W˜_(f,f)]d2f=(fxfx)2S˜x(f)d2fS˜x(f)d2f+S˜y(f)d2f+(fxfx)2S˜y(f)d2fS˜x(f)d2f+S˜y(f)d2f,
where W_ is the cross-spectral density (CSD) matrix17,28,29 of the beam at its waist location, Tr is the trace, and Sα(ρ)=Wαα(ρ,ρ) are the spectral densities (SDs)17,28,29 of the beam’s α=x,y polarization components.

Continuing to define the symbols in Eq. (3), W˜_ is the Fourier transform of W_, i.e.:

Eq. (4)

W˜_(f1,f2)=W_(ρ1,ρ2)exp[j2π(ρ1·f1ρ2·f2)]d2ρ1d2ρ2,
where S˜α is as follows:

Eq. (5)

S˜α(f)=Wαα(ρ1,ρ2)exp[j2π(ρ1ρ2)·f]d2ρ1d2ρ2,
where ρ=x^x+y^y and f=x^fx+y^fy. Note that the denominators of σx2 and σfx2 are equal due to Parseval’s theorem. As such, we define and use the symbol P=Px+Py hereafter to represent this normalization factor.

Last, x and fx are the normalized beam “first moments” at the beam’s waist location and in the spatial frequency domain, respectively, such that

Eq. (6)

x=xSx(ρ)d2ρPx+Py+xSy(ρ)d2ρPx+Pyfx=fxS˜x(f)d2fPx+Py+fxS˜y(f)d2fPx+Py.

Following the scalar partially coherent source analysis presented in Refs. 5 and 6, we choose the z axis—the mean propagation direction of the random beam—such that x=fx=0. This permits us to write Eq. (3) as follows:

Eq. (7)

σx2=PxPx+Pyσx,x2+PyPx+Pyσx,y2σfx2=PxPx+Pyσfx,x2+PyPx+Pyσfx,y2,
where σx,α2 and σfx,α2 are as follows:

Eq. (8)

σx,α2=1Pαx2Sα(ρ)d2ρσfx,α2=1Pαfx2S˜α(f)d2f.
Substituting Eq. (7) into Eq. (2) and simplifying produce as follows:

Eq. (9)

Mx4=Px(Px+Py)2(Px+σx,y2σx,x2Py)[4πσx,xσfx,x]2+Py(Px+Py)2(Py+σx,x2σx,y2Px)[4πσx,yσfx,y]2.
The bracketed quantities in Eq. (9) are the vector component beam quality factors, and therefore,

Eq. (10)

Mx4=1Px+Py(PxPx+Pyσx,x2+PyPx+Pyσx,y2)(Pxσx,x2Mx,x4+Pyσx,y2Mx,y4),
where Mx,α2=4πσx,ασfx,α are the beam quality factors for the α=x,y vector components.

1.2.

Prior Mx2 Scalar Analysis

Through extensive analysis and clever mathematics, Refs. 5 and 6 showed that Mx2 for a scalar beam of any state of coherence is as follows:

Eq. (11)

Mx4=16π2σx2σfx2J2,
where σx2 is given in Eq. (8) but computed in the source plane, σfx2 is as follows:

Eq. (12)

σfx2=14π2P2W(ρ1,ρ2)x1x2|ρ,ρd2ρ
and J is

Eq. (13)

J=1PIm[fxW˜(f1,f2)f1x|f,fd2f].
The |ρ,ρ and |f,f denote that ρ1=ρ2=ρ and f1=f2=f after computing the partial derivatives, respectively.

For Schell-model sources,17,28 Ref. 6 showed that Eq. (11) further reduces to

Eq. (14)

Mx4=(Mxc)44σx22a(ρd)xd2|ρd=0,
where (Mxc)2 is the beam quality factor of the corresponding coherent source, ρd=x^xd+y^yd=x^(x1x2)+y^(y1y2), and a(ρd) is related to the source’s spectral autocorrelation function by

Eq. (15)

μ(ρd)=a(ρd)exp[jψ(ρd)].
Both a and ψ are real functions. Since μ(0)=1 and μ(ρd)=μ*(ρd), a(0)=1, ψ(0)=0, a(ρd)=a(ρd), and ψ(ρd)=ψ(ρd). If μ is real, then a=μ, and μ can be substituted directly into Eq. (14).

1.3.

New Mx2 Vector Analysis

Substituting in Eq. (14) for the vector component beam quality factors in Eq. (10) and simplifying produce as follows:

Eq. (16)

Mx4=1Px+Py[PxPx+Pyσx,x2+PyPx+Pyσx,y2]×{Pxσx,x2[(Mx,xc)44σx,x22axx(ρd)xd2|ρd=0]+Pxσx,y2[(Mx,yc)44σx,y22ayy(ρd)xd2|ρd=0]}.
Because Mx2 in Eq. (16) is expressed in terms of vector component quantities (i.e., Mx,αc, σx,α2, Pα, and aαα), 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 σx2 via Eq. (7), expanding, and then simplifying Eq. (16) yields as follows:

Eq. (17)

Mx4=σx2[PxPx+Py(Mx,xc)4σx,x2+PyPx+Py(Mx,yc)4σx,y2]4σx2(PxPx+Py2axx(ρd)xd2|ρd=0+PyPx+Py2ayy(ρd)xd2|ρd=0).
Last, substituting in (Mx,αc)2=4πσx,ασfx,αc, recalling Eq. (7), and simplifying, the desired result is obtained as follows:

Eq. (18)

Mx4=(4πσx)2[PxPx+Py(σfx,xc)2+PyPx+Py(σfx,yc)2]4σx2[PxPx+Py2axx(ρd)xd2|ρd=0+PyPx+Py2ayy(ρd)xd2|ρd=0]=(Mxc)44σx2[PxPx+Py2axx(ρd)xd2|ρd=0+PyPx+Py2ayy(ρd)xd2|ρd=0].

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 Mx2 depends on the coherence of the source only through the second derivatives of the vector correlation functions evaluated at ρd=0. Rather intuitively, it differs in the fact that the “coherence contributions,” σx and σfxc, 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 Mx,α2. Unfortunately, this simple, physical form for Mx2 is spoiled by cross terms, which are evident in Eq. (16).

1.4.

Examples

Here, we calculate the Mx2 for two example, vector Schell-model sources. We then simulate them in terms of Mx2 in Sec. 2.

1.4.1.

EM Gaussian Schell-model source

The elements of the CSD matrix for an EGSM source are as follows:

Eq. (19)

Wαβ(ρ1,ρ2)=Aαexp(ρ124σα2)Aβexp(ρ224σβ2)Bαβexp(|ρ1ρ2|22δαβ2),
where α,β=x,y, Aα is the amplitude of the α field component, and σα is the RMS width of the α component SD. Also in Eq. (19), Bαβ is the complex correlation coefficient between the α and β field components and δαβ is the RMS width of the cross-correlation function μαβ, i.e., the cross-correlation function of the α and β field components.17,29 The EGSM source parameters—Aα, σα, δαβ, and Bαβ—must satisfy the realizability conditions derived in Refs. 17 and 30. Since only the diagonal elements of the CSD matrix are required to compute Mx2, the Bαα=1 criterion is the most relevant here.

To compute Mx2, we use Eq. (16). The normalized beam widths of the x and y components of the EGSM source, computed using Eq. (8), are σx,α=σα. The powers in the x and y components of the source are computed using the denominators in Eq. (3) and are Pα=2πσα2Aα2. The corresponding coherent source to the EGSM source defined in Eq. (19) is as follows:

Eq. (20)

Wαβc(ρ1,ρ2)=AαAβBαβexp(ρ124σα2)exp(ρ224σβ2).
This CSD function describes a vector source composed of horizontally and vertically polarized, spatially coherent, correlated, Gaussian beams. Since the x and y vector components of the source are Gaussian beams, the component beam quality factors are (Mx,αc)2=1. Last, the component spectral correlation functions μαα are real; therefore, aαα=μαα and the second derivatives in Eq. (16) evaluate to

Eq. (21)

2aαα(ρd)xd2|ρd=0=2xd2exp(ρd22δαα2)|ρd=0=1δαα2.

Substituting the quantities discussed in the previous paragraph into Eq. (16) and simplifying produce as follows:

Eq. (22)

Mx2=Ax2σx4+Ay2σy4(Ax2σx2+Ay2σy2)2[Ax2(1+4σx2δxx2)+Ay2(1+4σy2δyy2)]1/2.
This expression is equal to the Mx2 for an EGSM source derived in Refs. 910.11.12 using different methods. Note that letting either Ax=0 or Ay=0 yields the Mx2 for a scalar GSM source first derived in Ref. 6. We generate an EGSM source with a specified Mx2 in Sec. 2.

1.4.2.

Vector optical coherence lattices

Another 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:

Eq. (23)

Wαβ(ρ1,ρ2)=Aαexp(ρ124σα2)Aβexp(ρ224σβ2)×Bαβjinc(|ρ1ρ2|2δαβ)1Nn=1Nexp[jvn·(ρ1ρ2)],
where jinc(x)=2J1(x)/x, J1 is a first-order Bessel function of the first kind, and vn=x^vnx+y^vny is a vector that points from the origin to the n’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 σx,α=σα, Pα=2πσα2Aα2, and (Mx,αc)2=1 are the same as in the EGSM source example. The second derivative of aαα is much more difficult. Here, for analytical convenience, we assume that the lattice nodes are symmetric about the origin. This assumption means that μαα is real, and thus, aαα=μαα. After much calculus and algebra (details in Sec. 4 Appendix A), the second derivatives in Eq. (16) evaluate to

Eq. (24)

2aαα(ρd)xd2|ρd=0=2xd2[jinc(ρd2δαα)1Nn=1Nexp(jvn·ρd)]|ρd=0=18δαα21Nn=1Nvnx2.

Substituting σx,α=σα, Pα=2πσα2Aα2, (Mx,αc)2=1; Eq. (24) into Eq. (16); and simplifying produce as follows:

Eq. (25)

Mx2=Ax2σx4+Ay2σy4(Ax2σx2+Ay2σy2)2×{Ax2[1+12σx2δxx2(1+8Nn=1Nδxx2vnx2)]+Ay2[1+12σy2δyy2(1+8Nn=1Nδyy2vnx2)]}1/2.
We generate and analyze a VOCL, in terms of Mx2, in Sec. 2.

1.5.

Generating Stochastic Vector Fields in Terms of Mx2

An 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.3137 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

Eq. (26)

Eα(ρ)=Cαexp(ρ24σα2)Tα(ρ),
where Cα is the complex amplitude and Tα is the stochastic complex transmittance screen for the α=x,y component of the field, respectively. Tα 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 Eβ and comparing the resulting expression to Eqs. (19) and (23) yields the following equalities: |Cα|=Aα, arg(Cα)arg(Cβ)=arg(Bαβ), and

Eq. (27)

Tα(ρ1)Tβ*(ρ2)={|Bαβ|exp(|ρ1ρ2|22δαβ2)EGSMsource|Bαβ|jinc(|ρ1ρ2|2δαβ)1Nn=1Nexp[jvn·(ρ1ρ2)]VOCL.

Because potentially |Bxy|>0, Tx and Ty, in general, must be generated from correlated Gaussian random numbers. To see this and reveal the conditions on the values of |Bxy| and δxy, we note that an instance of Tα is generated by31,32

Eq. (28)

Tα[i,j]=mnrα[m,n]ΦTα[m,n]2LxLyexp(j2πNxmi)exp(j2πNynj),
where Nx, Ny are the numbers of grid points in the x, y directions, Lx=NxΔ, Ly=NyΔ are the lengths of the grid in the x, y directions in meters, and Δ is the grid spacing. In Eq. (28), rα is an Ny×Nx grid of zero-mean, unit-variance circular complex Gaussian random numbers and ΦTα is the spatial power spectrum of Tα, i.e., the Fourier transform of the autocorrelation of Tα. For the examples discussed in Sec. 1.4:

Eq. (29)

ΦTα(f)={2πδαα2  exp(2π2δαα2f2)EGSMsource8πδαα21Nn=1Ncirc(8πδαα|f+vn2π|)VOCL,
where circ(x) is the circle function defined in Ref. 38. We note that Eq. (28) is in the form of a discrete, inverse Fourier transform; thus, Tα can be generated quickly and efficiently using the fast Fourier transform algorithm.

Using Eqs. (28) and (29), a Tα with a specified δαα is produced. Physically, this leads to an EGSM source or VOCL with the correct, diagonal, CSD matrix elements. Recall that Mx2 depends only on these elements, and therefore, one could ignore any correlation between the x and y vector components (i.e., produce instances of Tx and Ty from independent Gaussian random numbers) and still produce a beam with the desired Mx2. 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 Tx with Ty, namely,

Eq. (30)

Tx[i1,j1]Ty*[i2,j2]=m1n1m2n2rx[m1,n1]ry*[m2,n2]2LxLyΦTx[m1,n1]ΦTy[m2,n2]×exp(j2πNxm1i1)exp(j2πNyn1j1)exp(j2πNxm2i2)exp(j2πNyn2j2).
The moment rx[m1,n1]ry*[m2,n2]=2Γδ[m1m2]δ[n1n2], where Γ is the correlation coefficient between the rx and ry random numbers and δ[n] is the discrete Dirac delta function. Substituting this into Eq. (30) and simplifying produce as follows:

Eq. (31)

Tx[i1,j1]Ty*[i2,j2]=mnΓΦTx[m,n]ΦTy[m,n]×1LxLyexp[j2πNxm(i1i2)]exp[j2πNyn(j1j2)].
The ΓΦTxΦTy must equal the cross-power spectra, or the Fourier transforms of the expressions in Eq. (27), viz.,

Eq. (32)

ΦTxTy(f)={|Bxy|2πδxy2exp(2π2δxy2f2)EGSMsource|Bxy|8πδxy21Nn=1Ncirc(8πδxy|f+vn2π|)VOCL.
This step produces the following relation for EGSM sources:

Eq. (33)

Γδxxδyyexp[2π2(δxx2+δyy22)f2]=|Bxy|δxy2exp(2π2δxy2f2).
From here, it is quite clear that

Eq. (34)

δxy=δxx2+δyy22|Bxy|=Γ2δxxδyyδxx2+δyy2,
where 0Γ1. These conditions were first derived in Ref. 31.

The VOCL conditions on |Bxy| and δxy are derived from

Eq. (35)

Γα=x,y[δαα2n=1Ncirc(8πδαα|fvn2π|)]1/2=|Bxy|δxy2n=1Ncirc(8πδxy|fvn2π|).
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 Γ=|Bxy|:

  • 1. If Bxy=0, δxy is physically meaningless and its value is irrelevant, Tx and Ty are statistically independent, and δαα can be chosen freely. This produces a VOCL with a diagonal CSD matrix and the beam is randomly, partially, or fully linearly polarized.

  • 2. If |Bxy|0, then δxx=δyy=δxy, and Tx and Ty are correlated to some degree. This produces a VOCL with a full CSD matrix and the beam is, most generally, elliptically partially polarized.

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 Mx2. The EGSM source and VOCL component beam quality factors are as follows:

Eq. (36)

Mx,α4={1+4σα2δαα2EGSMsource1+12σα2δαα2(1+8Nn=1Nδαα2vnx2)VOCL.
Referring back to Eqs. (22) and (25), substituting Eq. (36) into those expressions and inverting produces as follows:

Eq. (37)

Mx4(Ax2σx2+Ay2σy2)2Ax2σx4+Ay2σy4=Ax2Mx,x4+Ay2Mx,y4.

We assume that Aα and σα are given. This makes sense as we would expect the on-axis intensity Aα2 and size of the source to be known. In addition, since generally Mx,x2Mx,y2 (or equivalently, δxxδyy), one of those must be given. Without loss of generality, we assume Mx,y2 is known. Simplifying Eq. (37) further produces as follows:

Eq. (38)

Mx,x4=1Ax2[Mx4(Ax2σx2+Ay2σy2)2Ax2σx4+Ay2σy4Ay2Mx,y4].
Last, substituting Eq. (36) into Eq. (38) and simplifying yields

Eq. (39)

δxx2=4Ax2σx2(Ax2σx4+Ay2σy4){(Ax2σx2+Ay2σy2)2Mx4(Ax2σx4+Ay2σy4)[Ax2+Ay2(1+4σy2δyy2)]}1
for an EGSM source, and

Eq. (40)

δxx2=12Ax2σx2(Ax2σx4+Ay2σy4){(Ax2σx2+Ay2σy2)2Mx4(Ax2σx4+Ay2σy4)[Ax2(1+4σx2Nn=1Nvnx2)+Ay2(1+4σy2Nn=1Nvnx2+12σy2δyy2)]}1
for a VOCL. We note that the locations of the coherence lattice nodes vn must be known to find δxx.

In summary, to produce an EGSM source or VOCL with a desired Mx2:

  • 1. Specify Ax, σx, Ay, σy, Mx2, Mx,y2 and for a VOCL, vn. Recall that |Cα|=Aα. The values of Mx,y2, σy, and vn (for a VOCL) determine δyy [see Eq. (36)].

  • 2. Use Eq. (39) for an EGSM source or Eq. (40) for a VOCL to find δxx. For both sources, this effectively sets δxy [see Eq. (34) and VOCL conditions immediately following Eq. (35)].

  • 3. Specify arg(Cx), arg(Cy), and |Bxy|. Recall that arg(Cx)arg(Cy)=arg(Bxy). For both sources, the value of |Bxy| determines Γ [see Eq. (34) and for a VOCL, Γ=|Bxy|].

  • 4. Use a multivariate Gaussian random number generator to produce correlated rx and ry. The means and covariance matrix are as follows:

    Eq. (41)

    rxr=rxi=ryr=ryi=0Σ=[(rxr)2rxrrxirxrryrrxrryirxirxr(rxi)2rxiryrrxiryiryrrxrryrrxi(ryr)2ryrryiryirxrryirxiryiryr(ryi)2]=[10Γ0010ΓΓ0100Γ01],rx=rxr+jrxiry=ryr+jryi
    where the superscripts “r” and “i” stand for real and imaginary parts, respectively.

  • 5. Use Eqs. (28) and (29), the values of δxx, δyy, and vn (for a VOCL), and the rx and ry from step 4 to generate instances of Tx and Ty.

  • 6. Use Eq. (26), the values of Cx, Cy, σx, σy, and the Tx and Ty generated in step 5 to create an EGSM or VOCL field instance.

Since the generated fields are stochastic, this procedure will produce an EGSM source or VOCL with a desired, “on-average” Mx2.

2.

Simulation

Here, 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 Mx2. Before presenting the results, we discuss the details of the setup.

2.1.

Setup

For these Monte Carlo simulations, we used computational grids with Ny=Nx=512 points per side. The grid spacings were chosen such that Δ=min{σx,σy}/10 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 500  m1 apart.

Table 1

EGSM source and VOCL parameters.

EGSMVOCL
Ax1.31
Ay11.7
σx5 cm1 cm
σy3 cm0.97 cm
Bxy0.15exp(jπ/6)0
Mx21011.4
My21014.2028
Mx,x21112.5
Mx,y25.144610.9880
δxx0.9129 cm0.1286 cm
δyy1.1889 cm0.4707 cm
δxy1.0599 cmN/A

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, Mx2, and My2. The Stokes parameters in terms of the CSD matrix elements are as follows:17,29

Eq. (42)

S0(ρ)=Wxx(ρ,ρ)+Wyy(ρ,ρ)S1(ρ)=Wxx(ρ,ρ)Wyy(ρ,ρ)S2(ρ)=Wxy(ρ,ρ)+Wyx(ρ,ρ)S3(ρ)=j[Wyx(ρ,ρ)Wxy(ρ,ρ)].
For Mx2 and My2, the σx, σfx, σy, and σfy [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

Eq. (43)

W˜αβ(f,f)={AαAβBαβ16π2σα2σβ2δαβ22σα2+2σβ2+δαβ2exp[4π2δαβ2(σα2+σβ2)2σα2+2σβ2+δαβ2f2]EGSMsourceAαAβBαβ32π2σα2σβ2δαβ2σα2+σβ20exp[δαβ22(σα2+σβ2)u2]×J1(u)1Nn=1NJ0(8πδαβ|f+vn2π|u)duVOCL.
The integral in the VOCL W˜αβ 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.

Results

Figures 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—S0, S1, S2, and S3, 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 Mx2 and My2 to the corresponding theoretical values versus Monte Carlo trial number.

Fig. 1

EGSM source results—(a) S0NZ Thy, (b) S0NZ Sim, (c) S0FZ Thy, (d) S0FZ Sim, (e) S1NZ Thy, (f) S1NZ Sim, (g) S1FZ Thy, (h) S1FZ Sim, (i) S2NZ Thy, (j) S2NZ Sim, (k) S2FZ Thy, (l) S2FZ Sim, (m) S3NZ Thy (n) S3NZ Sim, (o) S3FZ Thy, (p) S3FZ Sim, and (q) convergence of simulated Mx2 and My2 to the corresponding theoretical values versus trial number.

OE_58_7_074101_f001.png

Fig. 2

VOCL results—(a) S0NZ Thy, (b) S0NZ Sim, (c) S0FZ Thy, (d) S0FZ Sim, (e) S1NZ Thy, (f) S1NZ Sim, (g) S1FZ Thy, (h) S1FZ Sim, (i) S2NZ Thy, (j) S2NZ Sim, (k) S2FZ Thy, (l) S2FZ Sim, (m) S3NZ Thy (n) S3NZ Sim, (o) S3FZ Thy, (p) S3FZ Sim, and (q) convergence of simulated Mx2 and My2 to the corresponding theoretical values versus trial number.

OE_58_7_074101_f002.png

Overall, the agreement between simulation and theory is excellent. We note that the visually conspicuous differences between the S2 and S3 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 S2=S3=0. Running more Monte Carlo trials would reduce these errors; however, S2 and S3 Sim will never be identically zero for that would require an infinite number of trials.

The simulated Mx2 and My2 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.

Conclusion

Here, we derived the M2 factor for a general vector Schell-model beam. Starting with Siegman’s M2 definition, we found an expression for the M2 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 M2 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 M2 relation and described how to synthesize vector Schell-model beams in terms of specified, desired M2.

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 M2 occurred within approximately 1000 vector field realizations. Although not performed here because of equipment availability, experimental synthesis and subsequent measurement of the M2 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

Eq. (44)

2aαα(ρd)xd2|ρd=0=2xd2[jinc(ρd2δαα)1Nn=1Nexp(jvn·ρd)]|ρd=0,
we apply the derivative multiplication rule twice to yield as follows:

Eq. (45)

2aαα(ρd)xd2|ρd=0=xd[1Nn=1Nexp(jvn·ρd)]|ρd=0xd[jinc(ρd2δαα)]|ρd=0+xd[jinc(ρd2δαα)]|ρd=0xd[1Nn=1Nexp(jvn·ρd)]|ρd=0+1Nn=1Nexp(jvn·ρd)|ρd=02xd2[jinc(ρd2δαα)]|ρd=0+jinc(ρd2δαα)|ρd=02xd2[1Nn=1Nexp(jvn·ρd)]|ρd=0.
The first derivatives in Eq. (45) are odd functions of xd; therefore, evaluating them at ρd=0 is trivially zero. Simplifying Eq. (45) produces as follows:

Eq. (46)

2aαα(ρd)xd2|ρd=0=2xd2[jinc(ρd2δαα)]|ρd=0+2xd2[1Nn=1Nexp(jvn·ρd)]|ρd=0.

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:

Eq. (47)

2xd2[1Nn=1Nexp(jvn·ρd)]|ρd=0=1Nn=1N(jvnx)2exp(jvn·ρd)|ρd=0=1Nn=1Nvnx2.
To evaluate the second derivative of the jinc function and argument, we first let the argument ρd/(2δαα)=u. The associated second derivative becomes

Eq. (48)

2xd2[jinc(ρd2δαα)]=xd{uxdu[2J1(u)u]}.
We now use the Bessel function identity:

Eq. (49)

2nxJn(x)=Jn1(x)+Jn+1(x)
to simplify Eq. (48) to

Eq. (50)

2xd2[jinc(ρd2δαα)]=xd{uxd[J0(u)+J2(u)]}.
Equation (50) can be further simplified by using the Bessel function identity:

Eq. (51)

2Jn(x)=Jn1(x)Jn+1(x),
yielding

Eq. (52)

2xd2[jinc(ρd2δαα)]=xd{uxd[12J1(u)12J3(u)]}=xd{uxd[12J1(u)+12J3(u)]}=xd{uxd2J2(u)u}.
In going from line 1 to 2 in Eq. (52), we used the Bessel function identity Jn(x)=(1)nJn(x), and in going from line 2 to 3, we used Eq. (49).

Continuing, we apply the derivative multiplication and chain rules producing

Eq. (53)

2xd2[jinc(ρd2δαα)]=2J2(u)u2uxd2(uxd)2u[2J2(u)u].
The derivative of 2J2(u)/u can be found by applying Eqs. (49) and (51), such that

Eq. (54)

2xd2[jinc(ρd2δαα)]=2J2(u)u2uxd2(uxd)2[14J0(u)14J4(u)].

We now evaluate the first and second derivatives of u. The first derivative of u is found by applying the chain rule, namely,

Eq. (55)

uxd=12δααρdxd=12δααxd(xd2+yd2)1/2=12δααxdρd1.
The second derivative of u is found by applying the multiplication and chain rules:

Eq. (56)

2uxd2=xd(12δααxdρd1)=12δααρd112δααxd2ρd3.
Substituting Eqs. (55) and (56) into Eq. (54) and simplifying produce as follows:

Eq. (57)

2xd2[jinc(ρd2δαα)]=2ρd2(1xd2ρd2)J2(ρd2δαα)18δαα2xd2ρd2[J0(ρd2δαα)J4(ρd2δαα)].

Last, evaluating Eq. (57) at ρd=0 yields as follows:

Eq. (58)

2xd2[jinc(ρd2δαα)]|ρd=0=18δαα2.

We obtain the second line of Eq. (24)—the desired result—by substituting Eqs. (58) and (47) into Eq. (46), i.e.:

Eq. (59)

2aαα(ρd)xd2|ρd=0=18δαα21Nn=1Nvnx2.

Acknowledgments

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. The authors have no conflicts of interest.

References

1. 

A. 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

2. 

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

3. 

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

4. 

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

5. 

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

6. 

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

7. 

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

8. 

X. Chu, B. Zhang and Q. Wen, “Generalized M2 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

9. 

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

10. 

S. Zhu and Y. Cai, “M2-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

11. 

S. Zhu and Y. Cai, “M2-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

12. 

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

13. 

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

14. 

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

15. 

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

16. 

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

17. 

O. Korotkova, Random Light Beams: Theory and Applications, CRC, Boca Raton, Florida (2014). Google Scholar

18. 

Vectorial Optical Fields, World Scientific, Hackensack, New Jersey (2014). Google Scholar

19. 

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

20. 

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

21. 

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

22. 

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

23. 

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

24. 

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

25. 

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

26. 

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

27. 

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

28. 

L. Mandel and E. Wolf, Optical Coherence and Quantum Optics, Cambridge University, New York (1995). Google Scholar

29. 

E. Wolf, Introduction to the Theory of Coherence and Polarization of Light, Cambridge University, New York (2007). Google Scholar

30. 

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

31. 

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

32. 

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

33. 

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

34. 

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

35. 

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

36. 

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

37. 

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

38. 

J. W. Goodman, Introduction to Fourier Optics, 3rd ed.Roberts & Company, Englewood, Colorado (2005). Google Scholar

39. 

J. D. Schmidt, Numerical Simulation of Optical Wave Propagation with Examples in MATLAB, SPIE Press, Bellingham, Washington (2010). Google Scholar

40. 

R. J. Sasiela, Electromagnetic Wave Propagation in Turbulence: Evaluation and Application of Mellin Transforms, 2nd ed.SPIE Press, Bellingham, Washington (2007). Google Scholar

41. 

I. S. Gradshteyn and I. M. Ryzhik, Table of Integrals, Series, and Products, 7th ed.Academic Press, Burlington, Massachusetts (2007). Google Scholar

42. 

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

43. 

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

44. 

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

Biography

Milo 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.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Milo W. Hyde and Mark F. Spencer "M2 factor of a vector Schell-model beam," Optical Engineering 58(7), 074101 (10 July 2019). https://doi.org/10.1117/1.OE.58.7.074101
Received: 25 April 2019; Accepted: 7 June 2019; Published: 10 July 2019
Lens.org Logo
CITATIONS
Cited by 2 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Monte Carlo methods

Optical engineering

Fourier transforms

Stochastic processes

Bessel functions

Coherence (optics)

Directed energy weapons

Back to Top