You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/338520535

Understanding diffraction grating behavior, part II: parametric diffraction


efficiency of sinusoidal reflection (holographic) gratings

Article in Optical Engineering · January 2020


DOI: 10.1117/1.OE.59.1.017103

CITATIONS READS

9 1,679

2 authors, including:

James Harvey
University of Central Florida
205 PUBLICATIONS 3,319 CITATIONS

SEE PROFILE

All content following this page was uploaded by James Harvey on 17 January 2020.

The user has requested enhancement of the downloaded file.


Understanding diffraction grating
behavior, part II: parametric
diffraction efficiency of sinusoidal
reflection (holographic) gratings

James E. Harvey
Richard N. Pfisterer

James E. Harvey, Richard N. Pfisterer, “Understanding diffraction grating behavior, part II:
parametric diffraction efficiency of sinusoidal reflection (holographic) gratings,” Opt. Eng. 59(1),
017103 (2020), doi: 10.1117/1.OE.59.1.017103.

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Optical Engineering 59(1), 017103 (January 2020)

Understanding diffraction grating behavior, part II:


parametric diffraction efficiency of sinusoidal
reflection (holographic) gratings
James E. Harvey* and Richard N. Pfisterer
Photon Engineering, LLC, Tucson, Arizona, United States

Abstract. With the widespread availability of electromagnetic (vector) analysis codes for describing the diffrac-
tion of electromagnetic waves by periodic grating structures, the insight and understanding of nonparaxial
parametric diffraction grating behavior afforded by approximate methods (i.e., scalar diffraction theory) is being
ignored in the education of most optical engineers today. We show that the linear systems formulation of non-
paraxial scalar diffraction theory enables the development of a scalar parametric diffraction grating model [for
transverse electric (TE) polarization] for sinusoidal reflection gratings with arbitrary groove depths and arbitrary
nonparaxial incident and diffracted angles. This scalar parametric analysis is remarkably accurate as it includes
the ability to redistribute the energy from evanescent orders into the propagating ones, thus allowing the cal-
culation of nonparaxial diffraction efficiencies to be predicted with an accuracy usually thought to require rigorous
electromagnetic theory. These scalar parametric predictions of diffraction efficiency are compared to paraxial
scalar and rigorous electromagnetic (vector) predictions for a variety of nonparaxial diffraction grating configu-
rations, thus providing quantitative limits of applicability of nonparaxial scalar diffraction theory to sinusoidal
reflection gratings as a function of the grating period-to-wavelength ratio (λ∕d ). © 2020 Society of Photo-Optical
Instrumentation Engineers (SPIE) [DOI: 10.1117/1.OE.59.1.017103]
Keywords: sinusoidal phase gratings; scalar parametric nonparaxial diffraction grating efficiency model; generalization of the classical
paraxial expression for diffraction efficiency of sinusoidal reflection gratings.
Paper 191193 received Aug. 26, 2019; accepted for publication Dec. 6, 2019; published online Jan. 9, 2020.

1 Introduction for sinusoidal reflection (phase) gratings with arbitrary


This paper is intended to be an extension of Ref. 1 in which groove depths and arbitrary nonparaxial incident and dif-
(1) elementary diffraction grating behavior was reviewed, fracted angles.7 This scalar parametric analysis is remarkably
(2) the importance of maintaining consistency in the sign accurate as it includes the ability to redistribute the energy
convention for the planar diffraction grating equation was from evanescent orders into the propagating ones, thus
emphasized, (3) the advantages of discussing “conical” dif- allowing the calculation of nonparaxial diffraction efficien-
fraction grating behavior in terms of the direction cosines of cies to be predicted with an accuracy usually thought to
the incident and diffracted angles were demonstrated, and require rigorous electromagnetic (vector) theory. These sca-
(4) the general grating equation for obliquely incident lar parametric predictions of diffraction efficiency compare
beams and arbitrarily oriented gratings was presented and well with rigorous predictions for a variety of nonparaxial
discussed. “Paraxial” grating behavior for coarse gratings diffraction grating configurations. Special attention is
(d ≫ λ) was then derived and displayed graphically for devoted to quantitatively discussing the limits of applicabil-
five elementary grating types: sinusoidal amplitude gra- ity of the presented results as a function of the grating period-
tings, square-wave amplitude gratings, sinusoidal phase to-wavelength ratio (λ∕d).
gratings, square-wave phase gratings, and classical blazed Since the late 1960s, holographic gratings, fabricated by
gratings. Paraxial diffraction efficiencies were calculated, the exposure of photoresist by a stationary sinusoidal inter-
tabulated, and compared for these five elementary grating ference fringe field, have become commonplace.8,9 The pho-
types. A linear systems formulation of “nonparaxial scalar toresist substrate is chemically developed after exposure to
diffraction theory”2–4 was then briefly reviewed and used produce a master holographic grating with sinusoidal groove
to predict the nonparaxial behavior [for transverse electric profiles. These master holographic gratings are routinely
(TE) polarization] of both the sinusoidal and the square- coated and replicated, as are master ruled gratings.
wave amplitude transmission gratings when the þ1 dif- Prior to the widespread use of holographic gratings for a
fracted order is maintained in the Littrow condition. This variety of different applications, the diffraction characteris-
nonparaxial behavior included the well-known Rayleigh tics of sinusoidal phase gratings were of interest primarily
(Wood’s) anomaly effects.5 Scalar theory is widely thought because other groove profiles (lamellar and blazed gratings)
to be unable to predict these Rayleigh anomalies.6 can be Fourier analyzed into a superposition of sinusoidal
In this paper, we show how this linear systems formu- profiles.10 Likewise, arbitrary (random) scattering surfaces
lation of nonparaxial scalar diffraction theory allows the are routinely modeled as a superposition of sinusoidal
development of a scalar parametric diffraction grating model surfaces of different amplitudes, periods, orientations, and
phases.11–14 Therein lies one of the motivations for writing

*Address all correspondence to James E. Harvey, E-mail: jimh@photonengr


.com 0091-3286/2020/$28.00 © 2020 SPIE

Optical Engineering 017103-1 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Harvey and Pfisterer: Understanding diffraction grating behavior, part II. . .

   
this paper: if one does not thoroughly understand the non- pffiffiffiffiffiffi r a
paraxial diffraction behavior of a single sinusoidal reflection tA ðx1 ; y1 Þ ¼
EQ-TARGET;temp:intralink-;e001;326;752 E0 Gaus 1 exp i sinð2πx1 ∕dÞ ; (1)
b 2
grating, then there is little chance of understanding the wide-
angle scatter behavior of random rough surfaces. where
In Sec. 2 of this paper, we derive the well-known classical
expression of the (monochromatic) diffraction efficiency of     2  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r r
a coarse (d ≫ λ) sinusoidal reflection grating with arbitrary Gaus 1 ¼ exp −π 1
EQ-TARGET;temp:intralink-;e002;326;705 and r1 ¼ x21 þ y21 ;
b b
groove depth as predicted by classical scalar diffraction
theory for paraxial incident and diffracted angles.15 This (2)
detailed derivation was not included in Refs. 1, 7, or even
15. We also look at the well-known approximation of that where E0 is the peak irradiance of the incident Gaussian
expression for “shallow” (smooth) sinusoidal reflection gra- beam. We have ignored a factor representing the average
tings and quantitatively evaluate how shallow is shallow phase delay through the grating. The parameter a represents
enough for that approximation to be valid. In addition we the “peak-to-peak excursion of the sinusoidal phase variation
illustrate the diffracted intensity profile as a function of of the wavefront emerging from the grating.”15
both groove depth and diffraction angle as predicted by the The irradiance distribution produced by a Gaussian beam
paraxial model for a sinusoidal reflection grating of period in an arbitrary plane perpendicular to the propagation direc-
d ¼ 20λ operating at normal incidence and discuss its tion is given by the squared modulus of the complex ampli-
limitations. tude distribution in that plane
In Sec. 3, we review the linear systems formulation of  
r
nonparaxial scalar diffraction theory,2–4 which suggests a Eðr; zÞ ¼ jtA ðx; yÞj2 ¼ E0 Gaus2
EQ-TARGET;temp:intralink-;e003;326;547: (3)
simple generalization of the classical paraxial parametric b
expression for the diffraction efficiency of sinusoidal reflec-
tion gratings. This nonparaxial scalar diffraction theory The total radiant power in the Gaussian beam is obtained by
enables the formulation of a parametric diffraction grating integrating the irradiance over that perpendicular plane
model for sinusoidal reflection (phase) gratings with arbi- Z Z  
trary groove depths and arbitrary nonparaxial incident and 2π ∞ r
PT ¼ E0 Gaus2
rdr dφ: (4)
diffracted angles.
EQ-TARGET;temp:intralink-;e004;326;478

φ¼0 r¼0 b
In Sec. 4, we generalize the classical paraxial parametric
expression for diffraction efficiency to include arbitrary As described in Ref. 18 (p. 421),
groove depth and nonparaxial incident and diffraction
  pffiffiffi 
angles. This exercise will provide insight into the nature r 2r
of the proper obliquity factor not only for nonparaxial dif- Gaus2 ¼ Gaus :
EQ-TARGET;temp:intralink-;e005;326;421 (5)
b b
fraction grating phenomena but also for a general wide-angle
surface scatter theory.13,14,16,17 Likewise, as shown in Ref. 18 (p. 74), the volume under
In Sec. 5, we compare the diffraction efficiencies pre- Gauss(r∕d) is equal to d2 .
dicted by the classical paraxial expression, the generalized The relationship between the total radiant power of the
scalar nonparaxial parametric expression, and rigorous laser beam and the peak irradiance in a given plane
electromagnetic (vector) diffraction theory for a variety of perpendicular to its propagation is thus given as
nonparaxial diffraction grating configurations. A summary,
statement of conclusions, and an extensive set of references b2
then completes the paper. PT ¼ E : (6)
2 0
EQ-TARGET;temp:intralink-;e006;326;320

2 Derivation of the Classical Expression for Making use of the Bessel function identity15 provided by
the Paraxial Efficiency of Sinusoidal Eq. (7)
Phase Gratings
  X∞  
We will first derive the well-known classical paraxial expres- a a
exp i sinð2πx1 ∕dÞ ¼ Jm expði2πmx1 ∕dÞ;
sion of the (monochromatic) diffraction efficiency of para-
EQ-TARGET;temp:intralink-;e007;326;255

2 m¼−∞
2
xial (d ≫ λ) sinusoidal reflection gratings with arbitrary
groove depth by using classical scalar diffraction theory. (7)
We will assume that the grating is illuminated by a small where Jm is a Bessel function of the first kind, order m, and
two-dimensional radially symmetric Gaussian laser beam the fact that the exponential function Fourier transforms into
of width b (radius at which the field drops to e−π of its peak a shifted delta function,18 we can apply the convolution theo-
value) that underfills the oversized physical grating. rem and write the Fourier transform of Eq. (1) as
X ∞   
2.1 For Arbitrary Groove Depth, Normal Incidence, a
and Paraxial Diffraction Angles F ftA ðx1 ; y1 Þg ¼ Jm δðξ − m∕d; ηÞ
EQ-TARGET;temp:intralink-;e008;326;152

m¼−∞
2
Using the symbolic notation for special functions popular- pffiffiffiffiffiffi
ized by Goodman15 and Gaskill,18 the complex amplitude   E0 b2 GausðbρÞ; (8)
transmittance (reflectance) function of a perfectly reflecting pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sinusoidal phase grating illuminated as described above can where ρ ¼ ξ2 þ η2 , and the spatial frequencies ξ and η are
be expressed as the reciprocal variables to x and y, respectively.

Optical Engineering 017103-2 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Harvey and Pfisterer: Understanding diffraction grating behavior, part II. . .

Owing to the replicating property of the delta function and the diffracted orders are approximately rotationally
under convolution,18 this can be written as an infinite array symmetric and centered. Hence we can go back to radial
of scaled and shifted Gaussian functions: coordinates
     
pffiffiffiffiffiffi X ∞
Eo b4 X

EQ-TARGET;temp:intralink-;e009;63;719

a 2 a 2 br2
F ftA ðx1 ; y1 Þg ¼ E0 b 2 Jm Gaus½bðξ − m∕d; ηÞ: E2 ðx2 ; y2 Þ ≈ 2 2 Jm Gaus : (14)
2 λ f m¼−∞ λf
EQ-TARGET;temp:intralink-;e014;326;719

m¼−∞ 2
(9)
Again from Eq. (5), this can be written as
If the grating is placed immediately behind an aberration-   pffiffiffi 
Eo b4 X

2 a 2br2
free positive lens of focal length f with the beam waist E2 ðx2 ; y2 Þ ≈ 2 2 Jm Gaus : (15)
λ f m¼−∞ λf
EQ-TARGET;temp:intralink-;e015;326;661

located at the lens as shown in Fig. 1, the Fraunhofer diffrac- 2


tion pattern appearing in the focal plane is given as15
 2 The radiant power in the m’th diffracted order is obtained
  by integrating the irradiance distribution of the m’th order
1  
over all space in the x2 − y2 plane:
E2 ðx2 ; y2 Þ ¼ 2 2 F ftA ðx1 ; y1 Þgj ξ ¼ x ∕λf  : (10)
λ f 
η ¼ y2 ∕λf 
EQ-TARGET;temp:intralink-;e010;63;611

2
 Z Z ∞ pffiffiffi 
Eo b4 2 a 2π 2br2
Pm ¼ 2 2 J m Gaus rdr dϕ:
λ f λf
EQ-TARGET;temp:intralink-;e016;326;579

From Eq. (10) the irradiance distribution on the focal 2 ϕ¼0 r¼0
plane can be written as (16)

E2 ðx2 ; y2 Þ
EQ-TARGET;temp:intralink-;e011;63;529 But as shown in Ref. 18, this integral is just equal to
 2 λ2 f 2 ∕2b2 , and hence
 ∞   
Eo b 4  X a   
¼ 2 2 Jm Gaus½bðξ − m∕d; ηÞj  ; b2
λ f  m¼−∞ 2 ξ ¼ x2 ∕λf  Pm ¼ Eo Jm 2 a
: (17)
η ¼ y2 ∕λf
EQ-TARGET;temp:intralink-;e017;326;497

2 2
(11)
Dividing Eq. (17) by Eq. (6), the diffraction efficiency of
or, evaluating this function at spatial frequencies ξ ¼ x2 ∕λf the m’th diffracted order of a sinusoidal phase grating oper-
and η ¼ y2 ∕λf, we obtain ating with normal incidence and paraxial diffracted angles
(θi ¼ 0, d ≫ λ) is given as
 ∞     2
Eo b4  X a b mλf   
E2 ðx2 ;y2 Þ¼ 2 2  Jm Gaus x2 − ;y2  : Pm 2 a
λ f m¼−∞ λf ηm ¼ ¼ RJ m
EQ-TARGET;temp:intralink-;e012;63;415

2 d ;
EQ-TARGET;temp:intralink-;e018;326;407 (18)
PT 2
(12)
where R is the reflectance of the reflection grating material
Since the size of the incident beam is large compared to the and a represents the peak-to-peak excursion of the sinusoidal
grating period (b ≫ d), there is a negligible overlap between phase variation of the wavefront emerging from the grating.
the individual Gaussian functions (diffracted orders). Thus, The phase variation induced upon the reflected wavefront
there are no cross terms in the squared modulus of the above by the surface variation of the grating is given by 2π∕λ times
summation, hence the optical path difference (OPD) of the reflected wavefront
     from one reflected by the mean surface shown in Fig. 2.7 The
E b4 X ∞
EQ-TARGET;temp:intralink-;e013;63;299

a b mλf
E2 ðx2 ;y2 Þ ¼ 2o 2 J2m Gaus2 x2 − ;y2 : total groove depth (peak to peak) of the sinusoidal surface is
λ f m¼−∞ 2 λf d equal to h. The surface height variation and the correspond-
ing reflected wavefront are illustrated as a function of x when
(13)
the grating grooves are aligned with the y axis:
For the current situation of normal incidence and paraxial  
2π nh 2π
diffracted orders (θi ¼ 0, d ≫ λ), the shift parameter in φðxÞ ¼ OPDðxÞ; OPDðxÞ ¼ sin x ; (19)
λ
EQ-TARGET;temp:intralink-;e019;326;241

Eq. (13) can be neglected (i.e., diffraction angles are small) 2 d

Fig. 1 Geometry for producing a Fraunhofer diffraction pattern of an Fig. 2 Illustration of a normally incident plane wavefront reflected
aperture (or grating) in the back focal plane of a lens. from a sinusoidal reflection grating.

Optical Engineering 017103-3 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Harvey and Pfisterer: Understanding diffraction grating behavior, part II. . .

where n is the refractive index of the media adjacent to the shallow grating approximation expressed in Eq. (22) is accu-
grating surface. If this media is air (n ¼ 1), the relative phase rate to within 1% for groove depths h < 0.03λ, or accurate to
variation is thus given as within 5% for h < 0.07λ but has an error >10% for h > 0.1λ.
  It also illustrates that the maximum diffraction efficiency that
hπ 2π can be achieved in the first order with a paraxial sinusoidal
φðxÞ ¼ sin x : (20)
λ phase grating is equal to η1 ≡ 0.3386.
EQ-TARGET;temp:intralink-;e020;63;719

The parameter a in Eq. (18) that determines the diffraction


efficiency of the various diffracted orders is the peak-to-peak 2.3 Illustration and Discussion of the Paraxial
excursion of the sinusoidal phase variation of the wavefront Diffracted Intensity Distribution and Its Limitations
emerging from the grating and, for paraxial incident and dif- Since the Fraunhofer diffraction integral implicitly contains
fracting angles, is thus equal to twice the peak amplitude of the paraxial approximation, by recalling our definitions of
the relative phase variation: radiometric quantities, we can write the diffracted intensity
distribution (radiant power per unit solid angle) emanating
a ¼ 4πh∕λ: (21)
from a sinusoidal phase grating of period d ≫ λ when illu-
EQ-TARGET;temp:intralink-;e021;63;609

minated by a narrow Gaussian beam as


With Eqs. (18) and (21) we have derived the parametric
description for diffraction efficiency of paraxial (d ≫ λ) X
∞  
b
sinusoidal reflection (phase) gratings operating at normal Iðθx ; θy Þ ¼ I o J m ða∕2ÞGaus
2 2 ðθ − mλ∕d; θy Þ ;
λ x
EQ-TARGET;temp:intralink-;e023;326;582

incidence. m¼−∞
The conservation of energy is easily shown for a perfectly (23)
conducting paraxial (d ≫ λ) reflection grating at normal
incidence because the sum over m from −∞ to ∞ of the where a is the peak-to-peak phase excursion and b is the
squared Bessel function in Eq. (18) is equal to unity. radius of the Gaussian beam (radius at which the field drops
Thus, we get the same expression for the diffracted effi- to e−π of its peak value).7
ciency of the m’th diffracted order from a sinusoidal phase Figure 4 illustrates this diffracted intensity profile plotted
grating as Goodman reported for normal incidence and para- as a function of groove depth along the x axis and diffraction
xial diffracted angles in the case of a perfectly conducting angle along the y axis. The maximum value of J21 ða∕2Þ is
(R ¼ 1) finite square grating illuminated by a unit amplitude 0.3386 and occurs for a ¼ 3.68, corresponding to a groove
normally incident plane wave.15 depth of h ¼ 0.293λ. The diffraction efficiency of the first
few orders for this value of a are tabulated in Table 1.
Note that the energy falls off rapidly, with 99.88% of the
2.2 Approximation for Diffraction Efficiency of diffracted radiant power contained in diffracted orders
Shallow Sinusoidal Phase Gratings jmj ≤ 3. This paraxial model is accurate only for very coarse
In addition to being a paraxial (d ≫ λ) grating, if the sinus- gratings (d ≫ λ). This paraxial model, thus, leads to the
oidal reflection grating is also shallow (i.e., the groove depth common misconception that it is impossible to get more than
is much less than a wavelength of the incident light), then the 33.86% of the incident energy into the first diffracted order
diffraction efficiency of the first orders of the sinusoidal of a sinusoidal phase grating.
reflection grating can be approximated as A simple thought experiment can prove this misconcep-
tion to be false. Suppose that we have a perfectly conducting
η1 ≡ J21 ða∕2Þ ≈ a2 ∕16:
EQ-TARGET;temp:intralink-;e022;63;331 (22) grating with a normally incident beam. If you decrease the
grating period, the diffracted angles increase and the higher
In Sec. 5.3 of Part I of this two-part paper,1 we provided orders eventually go evanescent. When only the 0 and 1
the expression in Eq. (18) without a detailed derivation and orders remain, changing the incident angle will cause the
graphically compared the diffraction efficiency of the first −1 order to go evanescent. Then one can vary the groove
diffracted orders as predicted by Eqs. (18) and (22). That depth to squelch the energy in the zero order. For a perfectly
curve is duplicated here as Fig. 3, and it establishes that the conducting sinusoidal reflectance grating, we can thus get
100% of the incident energy in the þ1 diffracted order.
Since there are now only two propagating orders, and there
is no energy in the 0’th order, “all” of the energy has to be
contained in the þ1 order. The maximum diffraction effi-
ciency of the first order for a perfectly conducting sinusoidal
reflection grating is therefore equal to unity, not 0.3386 as
predicted by the paraxial model.
If we use this paraxial model to describe the behavior of
fine (small spatial period) gratings whose diffracted orders
do not fall within the paraxial region, we will obtain incorrect
results that lead to a variety of other misconceptions. For
example: (1) the small angle approximation to the grating
equation inherent in Eq. (23) leads to the erroneous prediction
Fig. 3 Comparison of diffracted efficiency of a sinusoidal phase of equally spaced diffracted orders (in diffraction angle);1
grating as predicted by Eq. (18) and the common approximation for (2) when a diffracted order of finite width is located near
shallow (smooth) gratings expressed in Eq. (22). 90 deg, the predicted diffracted intensity (or irradiance)

Optical Engineering 017103-4 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Harvey and Pfisterer: Understanding diffraction grating behavior, part II. . .

Fig. 4 Diffracted intensity profile as a function of groove depth and diffraction angle as predicted for
a sinusoidal reflection grating of period d ¼ 20λ operating at normal incidence.

Table 1 Diffraction efficiencies. spatial variables are normalized by the wavelength of


light:2

Order # Efficiencies EQ-TARGET;temp:intralink-;e024;326;463 x^ ¼ x∕λ; y^ ¼ y∕λ; z^ ¼ z∕λ; etc: (24)


0 1.003 × 10−1
The reciprocal variables in Fourier transform space are
1 3.386 × 10−1 then precisely the direction cosines α and β of the propaga-
tion vectors of the plane wave components:
2 9.970 × 10−2
α ¼ x^ ∕^r; β ¼ y^ ∕^r; and γ ¼ z^ ∕^r (25)
3 1.093 × 10−2
EQ-TARGET;temp:intralink-;e025;326;399

4 6.320 × 10−4 in the angular spectrum of plane waves discussed by


Ratcliff,20 Goodman,15 and Gaskill.18
It is then shown that wide-angle scalar diffraction phe-
nomena are shift-invariant with respect to the changes in
exhibits a discontinuity that is an unacceptable nonphysical incident angle “only in direction cosine space.” Furthermore,
phenomenon;2 (3) the predicted angular width of the dif- it is the diffracted “radiance” (not intensity or irradiance) that
fracted orders does not indicate the broadening that occurs is shift-invariant in direction cosine space. This realization
with increasing diffraction angle;19 and finally (4) the diffrac- greatly extends the range of parameters over which simple
tion efficiencies of the propagating orders are not adjusted to Fourier techniques can be used to make accurate calculations
account for the Wood’s anomaly effect of redistributing the concerning wide-angle diffraction phenomena. Nonparaxial
energy associated with evanescent orders among the remain- diffraction grating behavior and surface scattering effects are
ing propagating orders.5 These inaccurate predictions were two diffraction phenomena that are not limited to the paraxial
discussed in Refs. 2 and 4 and are illustrated graphically region and benefit greatly from this new development.
in Ref. 2. By formulating scalar diffraction theory in terms of the
direction cosines of the propagation vectors of the angular
3 Review of the Linear Systems Formulation of spectrum of plane waves represented by the kernel of the
Nonparaxial Scalar Diffraction Theory Fourier transform integral, and incorporating sound radio-
A nonparaxial scalar diffraction theory has been developed metric principles, we obtained the following expression for
by using a scaled coordinate system in which all of the diffracted radiance2

L 0 ðα; β − βo Þ ¼ Kγ o Aλ s jF fU o ð^x; y^ ; 0Þ expði2πβo y^ Þgj2


2
for α2 þ β2 ≤ 1;
(26)
L 0 ðα; β − βo Þ ¼ 0 for α2 þ β2 > 1;
EQ-TARGET;temp:intralink-;e026;63;145

where K is a renormalization factor to be defined in Eq. (28) For large incident and diffraction angles, a portion of the
below, γ o is the cosine of the incident angle, and As is the differ- diffracted radiance distribution function will fall outside of
ential source area. the unit circle in direction cosine space. Those evanescent

Optical Engineering 017103-5 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Harvey and Pfisterer: Understanding diffraction grating behavior, part II. . .

Fig. 5 The profile of a broad Gaussian radiance distribution in direction cosine space, shifted due to a
64-deg incident angle is illustrated. The renormalized propagating portion of the radiance distribution is
also illustrated.

orders are imaginary and do not diffract any radiant energy a function of the incident and/or the diffracted angles in addi-
away from the grating surface. Parseval’s theorem from tion to the wavelength and the groove depth as expressed in
Fourier transform theory then requires that a renormalization Eq. (21). This general situation is depicted in Fig. 6.
factor, K, be applied to the diffracted radiance distribution The relative phase variation clearly depends upon both
remaining inside the unit circle in direction cosine space, the incident and the diffracted angles and is now given as
as illustrated in Fig. 5.2
The normalization factor, K, is given by the ratio of the 2π
ϕð^x; y^ Þ ¼ OPD; OPD ≈ ðγ i þ γ m Þhð^x; y^ Þ∕2; (28)
integral of the radiance distribution function over infinite λ
EQ-TARGET;temp:intralink-;e028;326;547

limits to the integral of the radiance-distribution function


over the unit circle in direction cosine space2 or
R∞ R∞
α¼−∞ β¼−∞ Lðα; β − βo Þdα dβ ϕð^x; y^ Þ ¼ πðγ i þ γ m Þhð^x; y^ Þ∕λ: (29)
Kðγ i Þ ¼ R R pffiffiffiffiffiffiffi

EQ-TARGET;temp:intralink-;e029;326;492

pffiffiffiffiffiffiffiffi Lðα; β − β o Þdα dβ


EQ-TARGET;temp:intralink-;e027;63;492

1−α 2
1
α¼−1 2 β¼− 1−α The peak-to-peak excursion of the sinusoidal phase varia-
tion is again equal to twice the peak amplitude of the relative
≡ Renormalization factor: (27)
phase variation
This renormalization factor K differs from unity only if a ¼ 2πðγ i þ γ m Þh∕λ: (30)
the diffracted radiance distribution function extends beyond
EQ-TARGET;temp:intralink-;e030;326;428

the unit circle in direction cosine space (i.e., only if evanes- Recall that Eq. (18) was an expression for diffraction effi-
cent waves are produced). The well-known Wood’s anoma- ciency for a paraxial application where radiant intensity and
lies that occur in diffraction-grating efficiency measurements radiance are essentially equivalent. However, the primary
are entirely consistent with this predicted renormalization in premise of the linear systems formulation of nonparaxial sca-
the presence of evanescent waves.2,4 lar diffraction theory described in Ref. 2 is that diffracted
radiance (not irradiance or intensity) is the fundamental
4 Generalizing the Classical Scalar Parametric radiometric quantity that exhibits shift invariance with
Expression for Diffraction Efficiency respect to changes in the incident and diffracted angles only
We will now show that the classical paraxial parametric when formulated in direction cosine space. Hence, for non-
expression for diffraction efficiency given in Eq. (18) can paraxial diffraction angles, we must use the following pro-
be generalized to accurately model nonparaxial situations. cedure for calculating diffraction grating efficiencies with
We will do this by modifying the parameter a (the peak- our scalar-based nonparaxial parametric expression for dif-
to-peak phase excursion of the diffracted wavefront) to be fraction efficiencies of sinusoidal phase gratings:

Fig. 6 Illustration of the wavefront diffracted at an arbitrary angle for an arbitrary incident angle.

Optical Engineering 017103-6 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Harvey and Pfisterer: Understanding diffraction grating behavior, part II. . .

• Start with the paraxial Eq. (18). λ ¼ 0.6328 μm;


EQ-TARGET;temp:intralink-;e033;326;752 d ¼ 9.492 μm; and h ¼ 0.293λ:
• Divide by γ m ¼ cos θm to convert diffracted intensity (33)
to diffracted radiance.
• Multiply by the renormalization factor K. The values for the actual diffraction angles and the cosine
• Substitute R with the polarization reflectance Q if other
of the diffracted angles are tabulated in Table 2, along
with the paraxial predictions of diffraction efficiency from
than TE polarization effects are expected. Eq. (18) with a given by Eq. (21), the scalar parametric pre-
The generalized parametric expression for diffraction effi- dictions from Eq. (31) with a given by Eq. (30), and rigorous
ciency of the m’th diffracted order for a sinusoidal reflection electromagnetic (vector) predictions by the Waterman inte-
grating with arbitrary groove depth and arbitrary incident and gral method for diffracted orders through the fifth order.
diffracted angles is thus now given as Percent errors of the paraxial and the scalar parametric meth-
ods from the rigorous calculations are also tabulated. The
   
P a diffraction efficiency calculations were made to four significant
ηm ¼ m ¼ K QJ 2m ∕γ m ;
EQ-TARGET;temp:intralink-;e031;63;625 (31) figures and tabulated in conventional scientific notation, as
PT 2
shown in Table 2. Percent errors are expressed to 100th of
where the peak-to-peak phase excursion a is given by a percent.
Eq. (30) and γ i ¼ cos θi and γ m ¼ cos θm . For this diffrac- The agreement among the three different predictions is
tion grating application, the renormalization factor previ- excellent (<1% error) for θm < 8 deg. The paraxial model
ously given by Eq. (27) is now obviously given as exhibits errors >11% for θm > 15 deg and >25% as the dif-
fraction angle approaches 20 deg. These diffraction efficien-
P∞
ηm 1 cies are compared graphically in Fig. 7 for normal incidence
K ¼ m¼−∞P ¼P : (32) and near-paraxial diffracted angles. The scalar parametric
ηm Prop:Orders ηm
EQ-TARGET;temp:intralink-;e032;63;536

Prop:Orders predictions exhibit <1% error until θm ≈19 deg.


For this situation of small diffraction angles and diminish-
ing diffraction efficiencies, the renormalization factor is
only slightly different than unity (K ¼ 0.999998568). Note
5 Comparison of Predicted Diffraction Efficiencies: that this value is actually less than unity as dividing the
Paraxial, Scalar Parametric, and Rigorous Bessel function expression in Eq. (18) by γ m increased the
We will now present detailed comparisons of predicted dif- diffraction efficiency of each order, thus making the sum of
fracted efficiencies by the classical paraxial expression from
Eq. (18), the parametric expression obtained from the appli-
cation of nonparaxial scalar diffraction theory from Eq. (31),
and by rigorous vector predictions using the Waterman inte-
gral method.

5.1 For Normal Incidence and Near-Paraxial


Diffraction Angles
For our first example demonstrating the scalar parametric
expression in Eq. (31), for the diffraction efficiency of a
perfectly conducting sinusoidal reflection grating, we will
operate at normal incidence and have chosen a wavelength
and period such that the fifth diffracted order approaches
20 deg (somewhat larger than what is usually considered Fig. 7 Comparison of the predicted diffracted efficiencies for normal
paraxial). incidence and near-paraxial diffracted angles.

Table 2 Comparison of predicted diffracted efficiencies.

Diffraction efficiencies Paraxial Parametric

m θm γm Paraxial Parametric Rigorous % Error % Error

0 0 1.000 9.995E-02 9.995E-02 1.007E-01 0.74 0.74

1 3.823 0.998 3.386E-01 3.393E-01 3.393E-01 0.21 0.00

2 7.662 0.991 9.984E-02 9.949E-02 9.918E-02 −0.67 −0.31

3 11.537 0.980 1.096E-02 1.062E-02 1.056E-02 −3.79 −0.57

4 15.446 0.964 6.344E-04 5.759E-04 5.711E-04 −11.08 −0.84

5 19.471 0.943 2.280E-05 1.839E-05 1.819E-05 −25.34 −1.10

Optical Engineering 017103-7 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Harvey and Pfisterer: Understanding diffraction grating behavior, part II. . .

Fig. 8 Comparison of the predicted paraxial, scalar-based paramet-


ric, and rigorous diffracted efficiencies for normal incidence and
moderate diffracted angles.

the efficiencies of all propagating orders slightly larger than Fig. 9 Diffraction grating efficiency of the first order of a perfectly con-
unity. By applying the renormalization factor in Eq. (32), the ducting sinusoidal reflection grating (h∕d ¼ 0.20) in Littrow condition
sum of the efficiencies of all propagating orders is precisely as predicted by the Beckmann theory, the paraxial scalar theory,
the nonparaxial scalar diffraction theory presented in this paper, and
unity as required. a rigorous integral vector theory.

5.2 For Normal Incidence and Moderate


(Nonparaxial) Diffraction Angles Littrow condition, θm ¼ θi , into the grating equation for
reflection gratings that we obtain
In Fig. 8, we have left the wavelength and the groove depth
fixed and reduced the grating period to 5.942 μm. The dif- θi ¼ sin−1 ð0.5λ∕dÞ: (34)
fracted angle of the fifth order is now greater than 32 deg and
EQ-TARGET;temp:intralink-;e034;326;467

the %error of the paraxial prediction exceeds 86%, whereas Thus, as λ∕d varies from 0.1 to 1.9, as shown in Fig. 9, the
the %error of the scalar parametric prediction remains <4%.7 incident angle (and θ1 ) varies from ∼3 deg to 65 deg.
Figures 7 and 8 demonstrate that the classical paraxial It is quite remarkable that our nonparaxial scalar parametric
expression of Eq. (18) is indeed very restrictive with regard diffraction grating efficiency model is so accurate over this
to diffraction angle, and the generalized scalar parametric wide a range of diffraction angles.
expression greatly relaxes those restrictions, at least up to It is of practical interest to continue to test the angular
moderate diffraction angles exceeding 30 deg.7 range of validity of this generalized scalar parametric model
of diffraction grating efficiencies for various diffracted
5.3 Diffraction Efficiency of the +1 Order Satisfying orders at increasingly larger diffracted angles.
the Littrow Condition for a Reflection Grating
Many practical diffraction grating applications are not per-
formed at normal incidence but instead involve maintaining 5.4 For Normal Incidence and Large Nonparaxial
a given diffracted order in the Littrow condition (θm ¼ θi for Diffraction Angles
reflection gratings). Also, substantially more insight into dif- Recall from Sec. 3 of Ref. 1 that λ∕d is the separation of the
fraction grating behavior, and our scalar parametric model, equally spaced diffracted orders when displayed in a direc-
can be demonstrated by calculating and displaying diffrac- tion cosine diagram. Figure 10 illustrates the direction cosine
tion grating efficiencies as a function of the dimensionless diagrams for three different values of Δα ¼ 0.2, 0.25, and
parameter λ∕d. 0.333 where the fifth, fourth, and third diffracted orders are
Figure 9, therefore, illustrates the diffraction efficiency of just going evanescent, i.e., reaching a diffraction angle of
a perfectly conducting (R ¼ Q ¼ 1) sinusoidal reflection 90 deg.
grating with groove depth-to-period ratio (h∕d) equal to Figure 11 graphically illustrates the predicted diffracted
0.20 when the þ1 order is maintained in the Littrow condi- efficiency (paraxial, scalar intensity parametric, and rigorous
tion. The above nonparaxial scalar diffraction theory pro- vector calculations) versus the quantity λ∕d for diffracted
vides remarkably good agreement with rigorous integral orders 3, 4 and 5 at normal incidence.
electromagnetic theory, not merely in the paraxial regime and We have again set the wavelength at 0.6328 μm and the
the smooth surface (shallow grating) regime, but over the groove depth at 0.293λ. Clearly, from the simple planar gra-
entire range of λ∕d, including at λ∕d ¼ 0.667 where there ting equation, the fifth, fourth, and third orders will reach
is a major Rayleigh anomaly when the −1 and þ2 diffracted 90 deg and go evanescent at λ∕d equals 0.2, 0.25, and
orders go evanescent simultaneously (see discussion of 0.33, respectively. The dashed black lines in Fig. 11 are the
Fig. 34 in Ref. 1). paraxial predictions from Eq. (18) for a perfectly conducting
Figure 9 was previously published in Ref. 4, and it should sinusoidal reflection grating with the parameter a given by
be noted that to maintain the þ1 order in the Littrow con- Eq. (21). The black dotted lines are the rigorous predictions
dition, each data point on these curves requires a different from the Waterman integral method. The lower solid curve
incident angle. It is readily shown by substituting the for each case is the intensity parametric prediction from

Optical Engineering 017103-8 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Harvey and Pfisterer: Understanding diffraction grating behavior, part II. . .

Fig. 10 Direction cosine diagrams for normal incidence and (a) Δα ¼ 0.2, (b) Δα ¼ 0.25, and
(c) Δα ¼ 0.333. For these three cases, the fifth, fourth, and third diffracted orders are just going
evanescent.

Fig. 11 Comparison of predicted paraxial, rigorous, and intensity Fig. 12 Predicted paraxial, rigorous, parametric intensities and para-
parametric diffracted efficiencies as a function of λ∕d for the third, metric radiance diffracted efficiencies as a function of λ∕d for the third,
fourth, and fifth diffracted orders for normal incidence. fourth, and fifth orders for normal incidence.

Eq. (18) with R ¼ 1 and the parameter a given by Eq. (30).


Note that this prediction is a big improvement over the para-
xial prediction; however, it becomes increasingly in error at
the larger diffracted angles as it is substantially lower than
the rigorous curve at large diffraction angles for all three
cases.
Figure 12 shows that the parametric radiance predictions
given by Eq. (31) with the peak-to-peak phase excursion
given by Eq. (30) are a significant improvement over para-
metric intensity predictions but are slightly higher than the
rigorous curve for the large diffraction angles.
Recall that the main lesson from the nonparaxial scalar
diffraction theory from Ref. 2 was that radiance (not intensity
or irradiance) is the fundamental quantity predicted by scalar
diffraction theory when proper radiometric principles are
applied.
At this point, we felt compelled to apply an empirical Fig. 13 Comparison of the paraxial prediction, rigorous electromag-
modification to our scalar radiance parametric model of dif- netic (EM) prediction, parametric intensity prediction, and empirically
fraction efficiency in an attempt to obtain an even better fit to modified parametric radiance prediction of diffracted efficiencies as
the rigorous calculations.7 Figure 13 shows that raising the a function of λ∕d for the third, fourth, and fifth diffracted orders for
normal incidence.
γ m in Eq. (31) to a power of 0.7 causes the parametric pre-
dictions for the fifth order to be a near-perfect fit to the rig-
orous predictions almost all the way out to a diffraction angle modified radiance parametric predictions are also an excel-
of 90 deg. The fitting procedure used was just a primitive lent fit out to slightly smaller angles. This empirically modi-
manual adjusting of this exponent and a visual observation fied radiance parametric expression for the diffraction
of the fit obtained in Fig. 13. The fourth and third order efficiency is thus given as

Optical Engineering 017103-9 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Harvey and Pfisterer: Understanding diffraction grating behavior, part II. . .

nonparaxial diffraction grating behavior of sinusoidal phase


gratings that is not readily obtained from rigorous vector
diffraction grating codes.
The key to obtaining this scalar nonparaxial parametric
expression for accurately modeling the diffraction efficiency
of sinusoidal reflection gratings for arbitrary groove depth
and arbitrary nonparaxial diffracting angles is to modify the
quantity a in Eq. (18) that represents the peak-to-peak excur-
sion of the sinusoidal phase variation in the classical paraxial
expression given in Eq. (18). With the use of Fig. 6, we cal-
culated the peak-to-peak excursion of the sinusoidal phase
variation of each diffracted order for arbitrary groove depth
and arbitrary incident and diffracted angles. This expression
was given by Eq. (30) as a ¼ 2πðγ i þ γ m Þh∕λ, where
Fig. 14 Illustration of the angular range of validity (given an allowable
γ i ¼ cos θi and γ m ¼ cos θm .
5% error from the rigorous prediction) of the diffraction efficiency pre- It was crucial to recognize that the classical paraxial
dictions by the empirically modified scalar radiance parametric model expression of Eq. (18) was for diffracted intensity, and the
of the third, fourth, and fifth diffracted orders for normal incidence. fundamental quantity of nonparaxial scalar diffraction theory
is diffracted radiance.2 We, thus, divided Eq. (18) by γ m to
    convert intensity to radiance and multiplied by the renormal-
Pm a ization factor, K, from nonparaxial scalar diffraction theory to
ηm ¼ ¼ K QJ 2m ∕ðγ m Þ0.7 : (35)
redistribute the radiant energy from the evanescent diffracted
EQ-TARGET;temp:intralink-;e035;63;534

PT 2
orders to the propagating ones. Finally, we substituted the
Finally, Fig. 14 shows the angular range of validity (given polarization reflectance, Q,13 for the scalar reflectance, R,
an allowable 5% error from the rigorous prediction) of the in Eq. (18) to obtain the generalized parametric expression
empirically modified scalar radiance parametric model for for diffraction efficiency of the m’th diffracted order for a
diffraction efficiency of sinusoidal reflection gratings. Note sinusoidal reflection grating with arbitrary groove depth and
that these empirically modified scalar radiance parametric arbitrary incident and diffracted angles given by Eq. (31).
predictions of diffraction efficiency are accurate (to within This generalized scalar nonparaxial parametric model for
5%) out to diffraction angles of 56 deg, 65 deg, and 77 deg predicting diffraction grating efficiencies is thus capable of
for the third, fourth, and fifth diffracted orders (at normal predicting (1) Rayleigh (Wood’s) anomalies and (2) polariza-
incidence), respectfully. tion effects when other than TE polarized light is present.
Table 3 shows the comparisons of the %errors of the vari- These two effects are usually thought to require rigorous
ous scalar diffraction models from the rigorous EM predic- electromagnetic theory.
tions. The second column of Table 3 tabulates the angular In Sec. 5, we graphically compared the predictions of this
range over which the %error of the empirically modified sca- generalized scalar nonparaxial parametric model of diffrac-
lar radiance parametric model remains less than 5% for the tion efficiency for sinusoidal reflection gratings to rigorous
third, fourth and fifth diffracted orders. The third and fourth calculations using the well-known vector Waterman method10
columns of Table 3 tabulates the corresponding %errors of and determined the angular range of validity (for a 5% error
the classical scalar paraxial model and the scalar intensity tolerance) for a variety of grating configurations.
parametric model respectively. These comparisons demon- An empirical modification was made by raising the γ m in
strate a dramatic improvement in the angular range of Eq. (31) to an arbitrary power in order to achieve a better fit
validity of simple intuitive nonparaxial scalar parametric to the rigorous predictions for the third, fourth, and fifth dif-
diffraction grating efficiency predictions. fracted orders over the entire range of diffracted angles illus-
trated in Figs. 12 and 13. Table 3 summarizes the results of
raising γ m to the 0.7 power, as shown bypEq. 7
ffiffiffi (35). Note that
6 Summary, Discussion, and Conclusions 0.7 is very close in value to both 2/3 and 2∕2. This suggests
We have developed a scalar nonparaxial parametric model of that there might be a theoretical basis for this empirical
diffraction efficiency for sinusoidal phase gratings that is modification.
useful for making accurate engineering calculations and vali- Finally, it was mentioned in Sec. 1 of this paper that ran-
dated it by rigorous vector calculations. This intuitive para- dom scattering surfaces are routinely modeled as a superpo-
metric model provides insight and understanding concerning sition of sinusoidal surfaces of different amplitudes, periods,

Table 3 Angular range of validity of various parametric models.

Diffracted order Angular range (deg) Paraxial I-parametric Modified L-parametric

Third order 56 164% error 31% error <5% error

Fourth order 65 626% error 45% error <5% error

Fifth order 77 4185% error 65% error <5% error

Optical Engineering 017103-10 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Harvey and Pfisterer: Understanding diffraction grating behavior, part II. . .

orientations, and phases.11–14 The results of this paper may 11. P. Beckman and A. Spizzichino, The Scattering of Electromagnetic
Waves from Rough Surfaces, Pergamon Press, New York (1963).
thus be of substantial interest to the surface scattering com- 12. J. M. Bennett and L. Mattsson, Introduction to Surface Roughness and
munity. In particular, due to an ongoing debate concerning Scattering, Optical Society of America, Washington, DC (1989).
13. J. C. Stover, Optical Scattering, Measurement and Analysis, McGraw-
the difference in the obliquity factor of the classical Hill, New York (1990).
Rayleigh–Rice and the GHSSmooth surface scatter theories,14 14. J. E. Harvey, Understanding Surface Scatter Phenomena: A Linear
the generalized expression for the parameter a (peak-to-peak Systems Formulation, SPIE Press, Bellingham, Washington (2019).
15. J. W. Goodman, Introduction to Fourier Optics, p. 69, McGraw-Hill,
excursion of the sinusoidal phase variation) of Eq. (30) New York (1968).
clearly shows that the obliquity factor in the GHSSmooth sur- 16. S. O. Rice, “Reflection of electromagnetic waves from slightly rough
face scatter theory is rooted in the fundamental diffraction surfaces,” Commun. Pure Appl. Math. 4, 351–378 (1951).
17. J. E. Harvey and R. N. Pfisterer, “Comparison of the GHSSmooth and the
behavior of sinusoidal phase gratings, and this gives cre- Rayleigh-rice surface scatter theories,” Proc. SPIE 9961, 996103
dence to the GHSSmooth obliquity factor. (2016).
18. J. D. Gaskill, Linear Systems, Fourier Transforms, and Optics, Wiley,
New York (1978).
References 19. J. E. Harvey and E. A. Nevis, “Angular grating anomalies: effects of
1. J. E. Harvey and R. N. Pfisterer, “Understanding diffraction grating finite beam size upon wide-angle diffraction phenomena,” Appl. Opt.
behavior: including conical diffraction and Rayleigh anomalies from 31, 6783–6788 (1992).
transmission gratings,” Opt. Eng. 58(8), 087105 (2019). 20. J. A. Ratcliff, “Some aspects of diffraction theory and their application
2. J. E. Harvey et al., “Diffracted radiance: a fundamental quantity in to the ionosphere,” Rep. Prog. Phys. 19, 188–267 (1956).
a non-paraxial scalar diffraction theory,” Appl. Opt. 38, 6469–6481
(1999). James E. Harvey is a retired associate professor from CREOL, the
3. J. E. Harvey et al., “Diffracted radiance: a fundamental quantity in non- College of Optics and Photonics at the University of Central Florida,
paraxial scalar diffraction theory: errata,” Appl. Opt. 39(34), 6374–6375 and currently is a senior optical engineer with Photon Engineering,
(2000).
4. J. E. Harvey, A. Krywonos, and D. Bogunovic, “Non-paraxial scalar LLC in Tucson, Arizona. He has a PhD in optical sciences from the
treatment of sinusoidal phase gratings,” J. Opt. Soc. Am. A 23, 858– University of Arizona and is credited with over 220 publications and
865 (2006). conference presentations in diverse areas of applied optics. He is
5. R. W. Wood, “On a remarkable case of uneven distribution of light in a member of OSA and a fellow and past board member of SPIE.
a diffraction grating spectrum,” Philos. Mag. 4, 396–402 (1902).
6. E. G. Loewen and E. Popov, Diffraction Gratings and Applications, Richard N. Pfisterer is a cofounder and president of Photon
p. 376, Marcel Dekker, New York (1997). Engineering, LLC. He received his bachelor’s and master’s degrees
7. J. E. Harvey and R. N. Pfisterer, “Parametric diffraction efficiency of in optical engineering from the Institute of Optics at the University of
non-paraxial sinusoidal reflection gratings,” Proc. SPIE 10375,
103750B (2017). Rochester in 1979 and 1980, respectively. Previously, he was the
8. D. Rudolph and G. Schmahl, “Verfaren zur Herstellung von head of optical design at TRW (now Northrop-Grumman) and senior
Rongtenlinsen und Beugungsgittern (Method for producing x-ray lenses optical engineer at Breault Research Organization. He is credited with
and diffraction gratings),” Umsch. Wiss. Tech. 67, 225 (1967). over 20 articles and conference presentations in the areas of optical
9. A. Labeyrie and J. Flamand, “Spectrographic performance of holo- design, stray light analysis, and phenomenology. He is a member of
graphically made diffraction gratings,” Opt. Commun. 1, 5–8 (1969). OSA and SPIE.
10. R. Petit, Electromagnetic Theory of Gratings, p. 98, Springer-Verlag,
Berlin (1980).

Optical Engineering 017103-11 January 2020 • Vol. 59(1)

Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 16 Jan 2020


Terms of Use:View
https://www.spiedigitallibrary.org/terms-of-use
publication stats

You might also like