Diseño y Fabricación de Hologramas Con Matlab
Diseño y Fabricación de Hologramas Con Matlab
Diseño y Fabricación de Hologramas Con Matlab
Published by
SPIE
P.O. Box 10
Bellingham, Washington 98227-0010 USA
Phone: + 1 360.676.3290
Fax: + 1 360.647.1445
Email: [email protected]
Web: http://spie.org
The content of this book reflects the work and thought of the authors. Every effort has
been made to publish reliable and accurate information herein, but the publisher is not
responsible for the validity of the information or for any outcomes resulting from
reliance thereon.
To my husband, Subbu:
Note: An empty day with clan is better handled with you! (2, 4, 6)
–Shanti
1 Introduction 1
1.1 Fundamentals of Diffractive Optics 1
1.1.1 Introduction 1
1.1.2 Refractive and diffractive optics 2
1.1.3 Scalar diffraction formulation 9
1.2 Software for Designing Diffractive Optics 12
1.3 Concluding Remarks 13
References 14
2 Design of Diffractive Optical Elements 19
2.1 Design of Simple Diffractive Optical Elements 19
2.1.1 Design and analysis of 1D gratings 19
2.1.2 Design of 1D gratings with MATLAB® 24
2.1.2.1 Design of 1D amplitude gratings with MATLAB 25
2.1.2.2 Design of 1D phase gratings with MATLAB 29
2.1.2.3 Design of 1D amplitude and phase sinusoidal
gratings with MATLAB 31
2.1.3 Design of 2D gratings 32
2.1.4 Binary circular gratings 37
2.1.5 Fresnel zone plates 37
2.2 Conclusions 41
2.3 Exercises 42
References 42
3 Design and Analysis of Advanced Diffractive Optical Elements 45
3.1 Design of Multilevel and Grayscale DOEs 45
3.1.1 Design procedure for multilevel and grayscale gratings 45
3.1.1.1 Design of a 4-level 1D phase grating with
MATLAB 46
ix
3.1.1.2
Design of an 8-level 1D phase grating with
MATLAB 48
3.1.1.3 Design of a blazed 1D phase grating with
MATLAB 49
3.1.1.4 Design of an 8-level axicon with MATLAB 50
3.1.1.5 Design of a blazed axicon with MATLAB 50
3.1.2 Design procedure for multilevel and grayscale FZPs 51
3.1.2.1 MATLAB simulation of a blazed FZP 52
3.1.2.2 MATLAB simulation of a four-level FZP 53
3.1.3 Design procedure for multilevel and grayscale spiral phase
plates 53
3.1.3.1 MATLAB simulation of a grayscale SPP 55
3.1.3.2 MATLAB simulation of a multilevel SPP 57
3.1.4 Design procedure for a gradient axilens 58
3.2 Design of DOEs using Algorithms 59
3.2.1 Design procedure of DOEs using the IFTA 60
3.3 Design Procedure of Beam-Shaping DOEs using a Simplified Mesh
Technique 62
3.3.1 Mesh generation method 63
3.3.2 Determination of the phase element 66
3.4 Conclusions 73
3.5 Exercises 73
References 73
4 Analysis of DOEs in the Fresnel Diffraction Regimes 77
4.1 Analysis of DOEs with the Fresnel Diffraction Formula 77
4.1.1 Fresnel diffraction pattern of a circular aperture 79
4.1.2 MATLAB simulations of the Fresnel diffraction pattern of
a binary phase axicon 79
4.1.3 MATLAB simulations of the Fresnel diffraction pattern
of an axilens 81
4.1.4 MATLAB simulations of the Fresnel diffraction pattern of
nonperiodic DOEs 82
4.2 Talbot Imaging 84
4.3 Conclusions 87
4.4 Exercises 88
References 88
5 Substrate Aberration Correction Techniques and Error Analysis 91
5.1 FZP in Finite Conjugate Mode 91
5.1.1 Design of FZPs in finite conjugate mode 92
5.2 Characterization of Substrate Aberrations 96
5.3 Aberration Correction Schemes 103
xiii
xvii
Happy Diffracting!
A. Vijayakumar and Shanti Bhattacharya
November 2016
Downloaded From: https://www.spiedigitallibrary.org/ebooks/ on 19 Apr 2019
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Acknowledgments
As with any project, this book began with one small step or rather one seemingly
small suggestion. Vijay (Vijayakumar), who was then my doctoral student in the
finishing stages of his work, proposed that we record the basics of what we had
done over the last few years in the form of a book. He felt that although a number
of books on diffractive optics exist, they did not cover the hands-on knowledge
required to actually start designing and fabricating such elements. Over the next
few days, we discussed his proposal, and in one of those coincidences that seem to
happen uncannily often, an email from SPIE Press popped into both of our email
boxes, encouraging researchers to publish. It seemed that the time was right to
create this book. The rest, as they say, is history.
The first person to thank, therefore, is my fellow author Vijay, for his
enthusiasm and hard work. This book would not exist if it were not for him.
Many of the examples and results presented were developed over the course of
Vijay’s Ph.D. work, and we both thank Prof. Enakshi Bhattacharya, Prof.
Harishankar Ramachandran, Dr. Deepa Venkitesh, and Dr. Balaji Srinivasan
of the Department of Electrical Engineering, IIT Madras, for their invaluable
suggestions during that period and Dr. Bijoy Krishna Das, from the same
department, for lending us his fiber-coupled sources with which to test our
elements.
All of the electron beam lithography (EBL) work presented in this book
was carried out at the Centre for NEMS and Nanophotonics, IIT Madras. We
thank the Ministry of Communications and Information Technology for
funding this Centre and our research work. The EBL work would not have
been possible without help from several staff members associated with the
EBL system, and we thank them and the students who have helped with the
running of the machine or preparation of the samples. In particular, we are
deeply indebted to Dr. Manu Jaiswal, Department of Physics, IIT Madras, for
his support and help when the RAITH electron beam lithography system was
being set up and to the ever-helpful, ever-present staff of the Centre, namely,
Mr. Rajendran, Mr. Prakash, Mr. Joseph, and Mr. Sridhar. We also thank
Mr. Karthick Raj, EBL operator (2013–2014), Centre for NEMS and
Nanophotonics, IIT Madras, for his help relating to the work described in
Section 8.3.4.
xix
The focused ion beam (FIB) work was carried out at two institutions. At IIT
Madras, the Quanta 3D FEG FIB system located in the Nano-Functional
Materials Technology Centre (NFMTC), IIT Madras, was used for all of the
initial FIB work. We are grateful to the NFMTC faculty in charge for giving us
access to the tool. The FIB work on fibers was carried out at the Max Planck
Institute (MPI) for Intelligent Systems, Stuttgart, Germany. We thank
Dr. Michael Hirscher and Prof. Joachim P. Spatz, MPI, for offering the use of
their FIB system to us any time we needed it. We are also grateful for their helpful
suggestions and discussions. Several of the results presented in Chapter 8 arose
from collaborative work between Dr. Shanti Bhattacharya and Dr. Pramitha
Vayalamkuzhi, Inspire Faculty, IIT Madras, and Ms. Ulrike Eigenthaler and
Dr. Kahraman Kesinbora, MPI. Figures 8.52, 8.54, 8.56, 8.57, 8.58, and 8.59
were taken with the FIB system made available to us at MPI. We gratefully
acknowledge their permission to use these results in this book. We thank
Mr. Jayavel, staff member, and Ms. Gayathri, M.S. doctoral student, Department
of Electrical Engineering, IIT Madras, for their suggestions and help in handling
and preparing fibers for diffractive optic fabrication with the FIB system.
Dr. Pramitha is specially thanked for her many contributions to Chapter 8.
We would be remiss if we did not acknowledge Prof. S. S. Bhattacharya
and Dr. Deepa Venkitesh for reviewing the initial chapters of this book. Their
suggestions greatly enhanced the clarity of those chapters. Vijay also thanks
his friend and former colleague, Mr. Vinoth, for his suggestions while writing
this book.
Special thanks goes to the SPIE Press editor, Ms. Dara Burrows, for her
useful suggestions and constant words of encouragement throughout the book
project.
Writing a book requires two kinds of support: that relating to the book
itself and that provided by the network of family and friends who surround us.
We hope that we have thanked everyone connected with the former and
apologize if, by oversight, we have left someone out.
We reserve the final part of this acknowledgment to the people who made
the task of writing easy just by being part of our lives. We truly appreciate the
constant encouragement and support extended to us by our families during
this book project.
Shanti thanks her friend Anu for their weekly runs that helped keep her
sane and her daughter Sumi for all the fun they share. As for her husband
Subbu, who once commented, “One day, I will get you to put in writing how
you cannot manage without me,” he may consider that day as having arrived!
Shanti would also like to use this opportunity to thank her Ph.D. advisor,
Prof. R. S. Sirohi for starting her out on this journey with light.
We hope that readers find this book useful and encourage them to contact
the authors with questions, comments, or suggestions at physics.vijay@gmail.
com or [email protected].
xxi
t Thickness/height of resist/substrate
T Transmittance value
u Object distance
v Image distance
w0 Waist of the Gaussian beam
zT Talbot distance
a Base angle of prism
b Deviation/diffraction angle
D Sampling period
Df Focal depth
l Wavelength
L Period of the grating
Lx Period of the grating along x direction
Ly Period of the grating along y direction
F Phase
F1D Phase of 1D binary phase grating
F2D Phase of 2D binary phase grating
FA Phase aberration
FAxilens Phase of axilens
FFZP Phase of Fresnel zone plate
FG Phase of grating
Fin Phase of input wave
Fm Phase of multifunctional diffractive optical element
Fout Phase of output wave
FR Phase of reference wave
FRing lens Phase of ring lens
FSPP Phase of spiral phase plate
c Complex amplitude of a wave
∇c Gradient of a wavefront
xxiii
Figure 1.1 Diffraction of light in slits with different widths: (a)–(c) decreasing slit widths
show increasing domination of diffraction.
Figure 1.2 Diffraction of a plane wavefront with a Gaussian intensity profile at a single-slit
aperture.
Figure 1.3 Scheme showing the generation of a Fresnel lens from a conventional plano-
convex lens: (a) conventional lens and (b) Fresnel lens. If t and d ≫ l, the lens is refractive. If
t and d are on the order of, or less than, the wavelength, the element is diffractive.
Figure 1.4 (a) Refractive prism and (b) generation of a Fresnel prism from it. (c) Blazed
diffraction grating with dimensions on the order of the wavelength.
that will fill the aperture to obtain a desired intensity distribution at an output
plane. The first DOEs were modified versions of refractive elements with
feature sizes on the order of the incident wavelength. To understand the above
statement, let us consider the conversion of a prism into a diffraction grating.
In the refractive regime, a prism can be used to disperse light or change the
direction of an incident monochromatic light. A similar element in the
diffractive regime is a grating.
The construction of a diffraction grating from a prism is shown in Fig. 1.4.
The base angle of the prism and its refractive index are given by a and ng,
respectively. The refractive bulk prism is converted into a thin element using
Fresnel’s technique, as shown in Figs. 1.4(a) and (b). Figure 1.4(c) shows a
diffraction grating, which is similar to Fig. 1.4(b), except that the feature sizes
are closer to the wavelength of light with a period L and thickness t. A Fresnel
prism is a refraction-dominated system, while a diffraction grating is a
diffraction-dominated one. A major difference arising because of this is the
fact that the former will bend the incident beam into one direction, while the
latter will generate multiple orders. The shape of the diffraction grating will
determine the number of orders and will be discussed in more detail in later
chapters. In order to force most of the diffracted light into a single diffraction
order, the diffraction grating must be blazed (i.e., have a triangular shape)
with a height or thickness t given by
l
t¼ , (1.1)
ðng 1Þ
1 l
a ¼ tan : (1.2)
Lðng 1Þ
Equation (1.2) shows that the geometrical profile of the prism is related to the
period of the diffraction grating.
The deviation angle b of the prism with a base angle of a and refractive
index ng can be calculated using trigonometry as
In order to become familiar with the terms and the language of diffractive
optics, we briefly introduce the amplitude grating here. Many of the concepts
will directly hold true for a phase grating as well. The second chapter provides
a much more detailed look at gratings and methods by which to design and
simulate their behavior.
Imagine a structure comprising a number of reflective slits surrounded by
opaque regions, as shown in Fig. 1.5. The slits are periodically spaced with a
distance L. This structure defines a basic diffraction grating. Light is incident
at an angle of bi with respect to the grating normal, which is indicated as
vertical dashed lines in the figure. The question is what determines the angle(s)
br of the beam after incidence on this surface? Since we have chosen reflective
slits, the light will travel back into the region of incidence, but we will not call
this reflection as the resulting intensity is due to the superposition of many
beams. For a more detailed description, we refer readers to a number of books
that discuss Huygens’ principle (every point in the slit acts a secondary source)
and scalar diffraction.1–5 Suffice to say that the multitudes of beams from each
slit interact with each other and result in an intensity pattern in the far field.
This pattern is not uniform, and the goal is to determine the locations of the
intensity peaks.
To arrive at the pattern, we look at two rays AB and A0 B0 that both
originated from the wavefront AA0 . In other words, at the plane AA0 , both
rays started with the same phase. For the diffracted wave shown in the figure,
to represent an actual wave, BB0 should be a wavefront. That is, the path
n1 A0 B0 n2 AB ¼ ml0 , (1.4)
where, n1 and n2 are the refractive indices seen by the incident and reflected
rays, respectively. Since both are in the same medium, n1 ¼ n2. With this
information, and comparing the triangles AA0 B0 and BB0 A, the equation can
be rewritten as
where, l ¼ l0/n1, and m is the order number. The implication of the order
parameter m is that Eq. (1.5) is satisfied for different values of br. Therefore,
the equation could be more accurately be written as
where, bmr represents the mth diffraction order. Thus, a picture of what is
happening after diffraction from the grating slowly emerges. Unlike specular
reflection, where the reflected light travels in one direction only, or scattered
light disperses into a solid angle from a surface, several ‘diffraction’ orders
exist. One could think of this as reflection occurring in a finite number of
preferred directions. If the grating had been a transmissive one, then
refraction would occur in more than one direction. The condition m ¼ 0
represents the classical optics case. For example, in the above grating, the
condition m ¼ 0 results in bi ¼ br, which is the law of reflection. For a
transmission grating, m ¼ 0 would reduce the equation to Snell’s law. While
the equation allows us to predict the possible directions of travel, it gives us no
information about how much light travels in each order. This means that we
cannot predict the efficiency of the diffractive structure. Obviously, efficiency
is important, and later chapters will include further equations that can be used
during the design stage to maximize it. The convention used to name the
various orders of a grating is indicated in Fig. 1.6. The figure can be used for
either a transmission or reflection grating. Angles are always measured from
the grating normal. The sign of the angle depends on the direction of rotation
of the ray from the normal (indicated by signs in the figure). On the other
hand, the sign of the order parameter m depends on the direction from the 0th
order. For example, for the reflected m ¼ þ1 order in the figure, angle b1 is
positive but would have been negative if the ray happened to lie on the other
side of the normal.
For simplicity, normal incidence is considered, and for the 1st diffraction
order, Eq. (1.6) can be simplified. The diffraction angle b1 of the 1st
diffraction order of the diffraction grating with a period of L is given by
l 1
b1 ¼ sin : (1.7)
L
From Eqs. (1.3) and (1.8), for cases where the base angle of prism a and the
refractive index ng are small, or where the period of the diffraction grating L is
large, Eqs. (1.3) and (1.8) reduce to a simpler equation:
This is an interesting result. When the period of the diffraction grating is large,
it behaves more like a refractive element; however, its behavior is different
when L approaches l. To quantitatively understand this, a few typical cases
are considered with ng ¼ 1.1, 1.5, and 1.9. The deviation angles b of a prism
and diffraction angles b1 of a grating were calculated using Eqs. (1.3) and
(1.4), respectively, and plotted against base angles a of a prism, as shown in
Fig. 1.7.
For smaller values of a, there is good overlap between b and b1.
Therefore, for DOEs with small diffraction angles, it is possible to derive the
profile blueprint from a ROE with an equivalent function. Examples of some
other elements that can be achieved in a similar way are the axicon,9 circular
grating,10 ring lens,11 and diffractive ring lens.12 We must however, always
keep in mind that in any optical element, both refraction and diffraction
coexist. The dominating effect, dictated by the feature sizes, decides whether
the element is diffractive or refractive.
Several researchers have studied the transition between refractive and
diffractive elements quite extensively.13,14 As discussed, this can be done by
Figure 1.7 Plot of deviation angle b of a prism (solid line) and diffraction angle b1 of
a grating (dashed line) versus variations in the base angle a of the prism for ng ¼ 1.1, 1.5,
and 1.9.
changing the feature sizes of an element and studying its behavior as the size
changes. In particular, the dispersive nature will vary depending on the feature
size. Again, let us take the examples of a lens and a prism. In the latter case,
we will compare a 1D (diffractive) grating and a (refractive) prism, as both
result in an off-axis deflection of the incident beam.14 In both cases, the
refractive index and, hence, the wavelength plays a role in the amount of
deflection. It is shown in Ref. 14 that the nature of dispersion is quite different
for the refractive and diffractive cases, with the latter experiencing much
greater dispersion for the same angle of deflection. Even more interesting is
the negative sign of the grating dispersion compared to that of the prism. In
other words, when light bends, different wavelengths bend by different
amounts, and the direction of bending is determined by the base optical
behavior of the element. The opposite signs of the dispersion of refraction and
diffraction have been used from very early on to provide some amount of
achromatization.15 Recent publications show that this concept is still being
manipulated for achromatization.16,17 In the refractive case, dispersion caused
by the material dominates, whereas, the structure of the element controls the
diffractive dispersion. Given these two very different causes, diffraction and
refraction cannot be balanced in a single element. Researchers are now
studying harmonic DOEs that lie somewhere between these two distinct
effects.18 In conclusion, it is clear that the structure of a DOE can be deduced
from the structure of a ROE that performs a similar optical operation.
interface or an optical element whose dimensions are close to that of the incident
wavelength. Diffraction theory allows one to calculate how wavefronts change
and how they travel after interaction with such an element. A complete analysis
of a diffractive system can be carried out using the vector diffraction
equations.19,20 Vector diffraction formulation can be used to understand the
behavior of DOEs with features both superwavelength as well as subwavelength.
This formulation can determine not only the intensity and phase profiles at
different planes, but also the state of polarization at these planes. However, the
formulation is quite difficult to implement and also to simulate. For most of the
analysis, which does not involve DOEs with features smaller than or equal to
the wavelength of light, and does not need to explain the polarization state of the
diffracted field, the simpler scalar diffraction formulation is sufficient. The focus
of this text book is only on superwavelength DOEs; therefore, the discussion is
limited only to the scalar diffraction formulation. Few models have been
developed that can explain the polarization state of a diffracted field using only
scalar diffraction formulation.21 The scalar diffraction formulation is described
in numerous text books.4,5 With the assumption that the features of the DOE are
larger than the wavelength of the source, and for spherical wavefronts, the scalar
diffraction formula based on Huygens–Fresnel theory is given by
ðð
z expðjkrÞ
Eðu,vÞ ¼ Aðx,yÞ dxdy, (1.10)
lj P r2
where (x, y) is the diffraction plane, and (u, v) is the observation plane. The
radius r can be given by
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r ¼ z2 þ ðu xÞ2 þ ðv yÞ2 : (1.11)
kðu2 þ v2 Þ
z≫ : (1.15)
2
At these distances, the radius of the spherical wavefront is large, so a section
of the spherical wavefront can be approximated to be a plane wavefront.
A summary of the approximations of the scalar diffraction formulation is shown
in Fig. 1.8. In this book, only the Fresnel and Fraunhofer approximations are
used for analysis of DOEs. The scalar diffraction integrals can be thought of as
Figure 1.8 Depiction of the validity of different approximations of the scalar diffraction
formula.
scalar diffraction propagators. In other words, they are a means to find the
diffraction amplitude and phase at any plane, given these values at an earlier
plane. These integral equations are used in their analog continuous form
only when analytic solutions for the diffraction problem being studied exist.
For all other cases, different techniques can be used to solve them to
determine the diffraction field at some plane. While the similarity to a
Fourier transform (FT) is clear when studying Fraunhofer diffraction, even
diffraction at closer planes can use the FT concept, as is obvious from
Eq. (1.13). (Further details are provided in Chapter 4.) Therefore, one
common method to solve scalar diffraction integral equations uses
discretized Fourier transforms (DFTs). One particularly efficient and fast
algorithm that implements a DFT is known as the fast Fourier transform
(FFT). This algorithm carries out the FT operation for N discrete samples in
O(N log N) steps, rather than the O(N2) steps of a standard FT operation.
Algorithms that carry out a DFT with O(N2) steps can also be used. These
algorithms have other benefits; for example: the size of the matrices (arrays)
used does not need to be powers of 2; there is more freedom in choosing
matrix sizes (hence, more freedom in fixing resolution at the diffraction
plane); they can tackle problems that cannot be handled by the FFT
algorithm, etc. While these benefits may seem attractive, they come at the
price of longer computation times. In addition to techniques that use the FT
as a basis for a diffraction solution, other beam propagation techniques such
as wavelets,22 finite element methods,23 etc.,24–26 can be used.
resolution of the element can be less than 10 nm, which is at least three orders
of magnitude higher than that of conventional spatial light modulators; in
addition, DOEs are lighter. Clearly, one can obtain a resolution better than
that of refractive elements while maintaining a compact optics configuration.
Besides the above advantages, some beam profiles, such as vortex beams,
chiral beams, etc.,49–52 cannot be generated using ROEs. Therefore, it is
possible to revolutionize the field of optical instrumentation by replacing
ROEs with equivalent lightweight DOEs and DOE- based optical instru-
ments. Recent research reports the fabrication of DOEs on the tip of optical
fiber, where the patterned fiber can be attached directly to a fiber laser to
achieve high-power beam shaping.53 These results could be useful for a wide
variety of biomedical applications such as laser-based surgery,54 endoscopy,55
laparoscopy,56 etc.
References
1. A. Authier, Early Days of X-Ray Crystallography,” Oxford University
Press, Oxford (2013).
2. J. Z. Buchwald and I. B. Cohen, Eds., Isaac Newton’s Natural Philosophy,
MIT Press, Cambridge, Massachusetts (2001).
3. S. Singh, Fundamentals of Optical Engineering, Discovery Publishing
House Pvt. Ltd, New Delhi (2009).
4. E. Hecht, Optics, Fourth Edition, Addison Wesley, San Francisco (2002).
5. J. W. Goodman, Introduction to Fourier Optics, Second Edition,
McGraw-Hill Companies Inc., New York (1996).
6. H. P. Herzig, Micro-optics: Elements, Systems and Applications, Taylor
and Francis Ltd., London (1997).
7. Y. G. Soskind, Field Guide to Diffractive Optics,” SPIE Press, Bellingham,
Washington (2011) [doi: 10.1117/3.895041].
8. K. Miyamoto, “The phase Fresnel lens,” J. Opt. Soc. Am. 51, 17–20
(1961).
9. J. H. McLeod, “The axicon: a new type of optical element,” J. Opt. Soc.
Am. 44, 592–592 (1954).
10. J. Dyson, “Circular and spiral diffraction gratings,” Proc. R. Soc. London
Ser. A 248, 93–106 (1958).
11. J. B. Goodell, “Eccentric lenses for producing ring images,” Appl. Opt. 8,
2566–2566 (1969).
12. M. R. Descour, D. I. Simon, and W.-H. Yei, “Ring-toric lens for focus-
error sensing in optical data storage,” Appl. Opt. 38, 1388–1392 (1999).
13. I. Snigireva, A. Snigirev, C. Rau, T. Weitkamp, V. Aristov, M. Grigoriev,
S. Kuznetsov, L. Shabelnikov, V. Yunkin, M. Hoffmann, and E. Voges,
“Holographic x-ray optical elements: Transition between refraction and
diffraction,” Nucl. Instrum. Methods Phys. Res. Sect. A 467-468, 982–985
(2001).
19
Figure 2.1 Phase profile of a 1D binary diffraction grating with period L and FF ¼ 0.5.
two parameters: the physical height or thickness t of the ridge and the
refractive index of the material of which the grating is made. In general, the
material is either glass or resist, which means that the refractive index is fixed,
and the design parameter used is only the thickness. To start, we look only at
gratings with a FF of 0.5, as this value provides the maximum efficiency.
(It is left to the readers to prove this using the MATLAB codes given in this
chapter.) The phase profile of the 1D binary phase grating with a FF of 0.5 is
given by
(
F 0 ≤ x ≤ L2
F1D ðxÞ ¼ F1D ðxÞ ¼ F1D ðx þ LÞ: (2.1)
0 L2 ≤ x ≤ L
The period of the grating controls the diffraction angle and, hence, the
spacing between the diffraction spots at the plane of interest. On the other
hand, the FF and the phase height F control the distribution of intensities in
different orders. The basic grating equation relating the wavelength of light l,
the period of the grating L, and the diffraction angle can be derived from
Fig. 2.2.
Using Fourier analysis,2 a binary phase 1D grating of period L can be
considered as a sum of many sinusoidal gratings of different amplitudes,
phases, and periods, namely, L, L/2, L/3, and so on. However, for a binary
phase grating with a FF of 0.5, the even diffraction orders vanish,10 and only
Figure 2.3 Sinusoidal grating components with periods (L, L/3, L/5, . . . ) of the binary 1D
grating with period L.
2 j ðFþmp Þ F mp
Dm ¼ e 2 sin sin : (2.3)
jmp 2 2
From Eq. (2.3) it can be noted that Dm ¼ 0 for all even values of m. The
diffracted beams’ intensity values Im relative to the incident beam can be
obtained from the square of the absolute values of the Fourier coefficients.
The relative intensity for diffraction orders m ¼ 0, 1, and 3 are given by
Eqs. (2.4), (2.5), and (2.6), respectively:
2
F
I 0 ¼ jD0 j ¼ cos
2
: (2.4)
2
2
2 F
I 1 ¼ jD1 j ¼2
sin , (2.5)
p 2
2
2 F
I 3 ¼ jD3 j ¼
2
sin : (2.6)
3p 2
Figure 2.4 Plot of the normalized relative intensity values for 0th, 1st and 3rd diffraction
orders with a variation in wavelength (or phase) for a binary grating with a resist thickness of
633 nm.
wavelength for which the intensity of the 0th diffraction order is same as that
of the 3rd diffraction order. The maximum efficiency for the 3rd diffraction
order is 4.5%.
This technique can be used to engineer 1D gratings to obtain the essential
intensity profiles at the application plane. In many cases, the grating is used in
conjunction with a focusing lens having a focal length f as shown in Fig. 2.5.
The spacing between the 0th diffraction order and the mth diffraction order can
be estimated using trigonometry as f tan b.
In general, the diffraction pattern in the far field is estimated by
calculation of the Fourier transform of the aperture function.2,3 If A(x, y) is
the aperture function, then the amplitude profile of the diffraction pattern in
the Fourier domain (spatial frequency domain) E(u, v) is given by
Figure 2.5 Optics configuration for focusing the diffraction orders generated by a binary
phase grating (BPG) with a convex lens.
ð þ`
þ` ð
ejkf ej 2f ðu þv Þ
k 2 2
2p
Eðu,vÞ ¼ Aðx,yÞ exp j ðxu þ yvÞ dxdy, (2.7)
jlf lf
` `
where f is the focal length of the lens used to cancel the quadratic phase factor
of the Fresnel diffraction integral.2,3 The spatial frequencies along the x and y
directions are fx ¼ x/lf and fy ¼ y/lf, respectively.
Figure 2.6 Algorithm for design and analysis of a DOE. The algorithm uses analytic
expressions and is applicable to any software.
direction, instead of writing the full code and verifying later. In this book, the
design and analysis is performed using MATLAB. For SCILAB users, the
MATLAB code can be easily converted into SCILAB code.12 Schemes for
converting MATLAB code to C, Cþþ, Mathcad, and Python codes are
available in numerous websites.
1D gratings are designed as matrices of size N M, where N and M are
the number of pixels along x and y axes, respectively. In most of the examples
in this book, N ¼ M. For simulation of diffraction patterns, the matrix size
has been chosen as 500 500 to minimize the computation time. However,
the choice of the value of N and M depends on various factors, such as the
overall size of the element, the resolution with which it needs to be described,
and even the resolution of the tool that will be used to actually fabricate the
element. The fundamentals of sampling are discussed in various text books13
and research papers.14,15 The values stored in each pixel represent a sampled
version of the structure to be designed. The amplitude and phase values are
assigned assuming the size of each pixel to be 1 mm. The far-field diffraction
pattern can be calculated by performing a Fourier transform of the aperture
profile generated in MATLAB.15,16 In particular, the Fourier transform for a
2D matrix is calculated using the ‘fft2’ function in MATLAB.
Table 2.1 MATLAB code for design and analysis of 1D binary amplitude grating.
%%1D Amplitude grating%%
clear %Clear all memory
% Defining Grating Parameters
N¼500; %Define Matrix size
A¼zeros(1,N); %Define a row Matrix by assigning 0 to all pixels
P¼100; %Define the period of the grating
FF¼0.5; %Define fill factor
% Constructing the Grating
for q¼1:N;
if rem(q,P),P*FF; %Use remainder function ‘rem’ to construct
%the grating
A(1,q)¼1;
end
end
A¼repmat(A,N,1); %replicate the row to create a 2D grating
% Alternative code for constructing the grating
O¼ones(N,FF*P);
Z¼zeros(N, P-FF*P);
unit¼[O Z];
A¼repmat(unit,1,N/P); %replicate to create a 1D grating
%Observing the grating output in the far field
E¼fftshift(fft2(A)); %fftshift is used to re-order the terms
%in their natural order
IN¼(abs(E)/(N*N)).*(abs(E)/(N*N)); % Calculating intensity
figure (1)
colormap(gray); %colormap(gray) is used to display grayscale
%image
imagesc(A);% imagesc is used to display a high constrast image
figure (2)
colormap(gray);
imagesc(IN);
folder’ and then choose the folder that contains the functions that need to be
permanently available, similar to inbuilt MATLAB commands. Each
function must be saved as an individual file with the name that will be used
to call it. We have created a set of functions that use either normalized
intensity or the intensity values as directly calculated. These functions are
provided in the Appendix.
After executing the program, the grating profile is visible in the first figure
window [Fig. 2.7(a)], while the diffraction spots will be seen in the second one
[Fig. 2.7(b)]. In the latter case, one should “zoom in” until the spots are clear.
Now, by selecting the ‘data cursor’ icon in the figure and clicking on the
diffraction spots, the relative intensity directed in that particular order will be
displayed. It can be noted that the efficiency of the 1st order is only 10.14%,
which is only 1/4 of the maximum intensity in the case of 1D grating profile
arising due to blocking of 50% of the light input at the grating plane. The
MATLAB code of Table 2.1 can be implemented for different FFs by
assigning different values to the parameter FF. Figure 2.1 can be generated
using the command ‘meshz(A)’.
Figure 2.7 Output of the code shown in Table 2.1 (a) Profile of the 1D amplitude grating
and (b) diffraction patterns at the Fourier plane. The latter image was obtained by zooming in
on the generated output figure.
Figure 2.8 (a) Output figure for diffraction by an amplitude grating with a period twice that
of the sampling period and (b) output figure for diffraction by a circular grating (Table 2.8)
when the period is 12 mm sampled at a rate of 10 mm/pixel.
as shown in Fig. 2.8(a). Since the central peak is located at pixel number 251,
only one of the orders is visible on one side, while the other order is out of
range (by one pixel). The aliasing effects are clearly visible when the DOE
period is smaller than the sampling period.21,22 For instance, let us consider
the case where the sampling period is 10 mm but the period is 12 mm for a
circular grating (generating by running the code of Table 2.8). Instead of a
circle, the pattern in Fig. 2.8(b) is obtained.
The second requirement is zero padding of the diffraction plane. Zero
padding is the process of adding zeros around a matrix such that the
information-carrying part actually occupies a smaller (central) area of the
matrix. The use of this technique always involves some trade-offs, as must be
obvious from the fact that useful resolution in the DOE plane is being
sacrificed. For instance, in the case of Fresnel diffraction, which we will
introduce in Chapter 4, zero padding is used to provide more information at
the observation plane.23 While this can be thought of as an improvement in
resolution in the far-field pattern, one has to be careful with that statement, as
zero padding cannot improve resolution beyond what is provided at the DOE
plane.24,25 The only way to do that would be to increase the spatial size rather
than decrease it. The higher level of detail at the observation plane with zero
padding is therefore obtained at the cost of resolution at the DOE plane itself.
Let us once again consider the simulation of a 1D amplitude grating. This
time, let us zero pad the grating structure and repeat the calculation. Images
of the diffraction patterns when the diffraction grating is zero padded by 125
pixels and 200 pixels are shown in Figs. 2.9(a) and (b), respectively.
This means that for a 500 500 pixel matrix, the grating occupies an area
corresponding to 250 250 pixels and 100 100 pixels, respectively, for these
Figure 2.9 Images of the diffraction patterns when the diffraction grating is zero padded by
(a) 125 pixels and (b) 200 pixels.
two cases. The result is that at the observation plane, an increase in resolution
of the diffraction patterns is observed, resulting in a better match between
simulation and experimental results. However, for analysis of efficiency, the
case without zero padding is convenient, as it is easy to extract the
information about the efficiency by just clicking on the spot. In Section 2.1.4
a similar analysis is shown for a circular grating.
Table 2.2 MATLAB code for design and analysis of a 1D binary phase grating.
%%1D Phase grating%%
clear; %Clear all memory
% Defining Grating Parameters
N¼500; %Define Matrix size
A¼ones(1,N); %Define a Matrix by assigning 1 to all pixels
P¼100; %Define the period of the grating
FF¼0.5; %Define fill factor
% Constructing the Grating
for q¼1:N; %Scan pixel by pixel
if rem(q,P),P*FF; %Use remainder function ‘rem’ to construct
%the grating
A(1,q)¼exp(1i*pi);
end
end
A¼repmat(A,N,1); %replicate the row to create a 2D grating
%Observing the grating output in the far-field
E¼fftshift(fft2(A)); %fftshift is used to re-order the terms
in their natural order
IN¼(abs(E)/(N*N)).*(abs(E)/(N*N)); % Calculating intensity
figure(1)
colormap(gray);
imagesc(angle(A))
figure(2)
colormap(gray);
imagesc(IN);
Figure 2.10 Output of the code shown in Table 2.2. (a) Profile of the 1D phase grating and
(b) diffraction patterns at the Fourier plane. The latter image was obtained by zooming in on
the generated output figure.
diffraction orders will display the same value: 0.29. One can repeat the
exercise for the case to equalize the intensity of the 0th diffraction order with
that of 3rd diffraction orders. Once again, the exercise may be repeated by
varying the FF of the grating.
Figure 2.11 Output of the code shown in Table 2.3 (a) Profile of the 1D sinusoidal
amplitude grating and (b) diffraction patterns at the Fourier plane. The latter image was
obtained by zooming in on the generated output figure.
to 0th, 1st, 2nd, and 3rd diffraction orders with efficiency values of 1%,
33.9%, 10%, and 1%, respectively.
The periodicity along the x and y directions gives rise to diffraction spots
along those directions. The relative intensity in various diffraction orders can
be calculated using Eqs. (2.4)–(2.6). The mx diffraction orders generated by
x periodicity of the 2D binary phase grating is divided further into my
diffraction orders due to periodicity along the y direction. The relative
intensity in (mx, my) diffraction orders can be calculated by multiplication of
the respective relative intensities in different orders along the x and y
directions. For instance, the relative intensity in (mx ¼ 1, my ¼ 1) and
(mx ¼ 1, my ¼ 3) can be calculated using Eqs. (2.9) and (2.10), respectively:
2 2
2 F 2 F
I 1,1 ¼ jD1 j jD1 j ¼
2 2
sin sin , (2.9)
p 2 p 2
2
2
2 F 2 F
I 1,3 ¼ jD1 j jD3 j ¼
2 2
sin sin : (2.10)
p 2 3p 2
Hence, the maximum possible efficiency in the (mx ¼ 1, my ¼ 1) and
(mx ¼ 1, my ¼ 3) diffraction orders are only 16% and 1.8%, respectively.
The image of a 2D grating seen in many text books13 and literature27–32 is
shown in Fig. 2.12. The grating has periodicity in both the x and y directions.
In the former case (amplitude elements), the white and black portions
represent areas with transmittance ‘1’ and ‘0’, respectively. For the latter case
(phase elements), the entire area has transmittance ‘1’, but the white regions
all have a constant phase difference with respect to the black regions, which
can be considered to have phase ‘0’.
Figure 2.14 Fundamental building block of the 2D grating shown in Fig. 2.12.
grating to be evaluated with Eqs. (2.9) and (2.10) is not the grating shown in
Fig. 2.12.
Let us consider the fundamental building unit (Fig. 2.14) of the 2D grating
shown in Fig. 2.12. It can be noted that 75% of the unit has a phase value of 0,
while only 25% of the unit has a phase value of p. In that sense, the actual FF
of this grating is 0.25. The Fourier coefficients for F ¼ p yield the values
corresponding to Eqs. (2.4)–(2.6) only when the FF is 0.5. This can easily be
tested by designing a 1D phase grating such that the FF is not 0.5. The FF can
be modified in the MATLAB code shown in Table 2.2. Another way to
understand this effect by recalling the fact that the 1st diffraction order has
maximum efficiency only when the 0th diffraction order is completely
cancelled. This can only happen when the light emanating from the 0 phase
regions and the p phase regions are equal. Hence, the areas of the two binary
phase regions must be always equal to each other in order to extract the
maximum efficiency values. Therefore, the grating that more accurately
describes the 2D version of the 1D grating with a FF of 0.5 is known as a
checkerboard grating and is shown in Fig. 2.15.
The checkerboard grating27,28 can be generated by many methods;
MATLAB even has a function named ‘checkerboard’. MATLAB codes
for designing a checkerboard phase grating are presented Tables 2.6 and 2.7.
Once again, we present two versions. One using for and if loops, which is easy
to understand, and a second version, which is much more compact and
Table 2.6 MATLAB code for design of a checkerboard grating (version 1).
% Version 1
%%Checkerboard phase grating%%
clear; %Clear all memory
% Defining Grating Parameters
N¼500; %Define Matrix size
A1¼zeros(N,N); %Define Matrices by assigning 0 to all pixels
A2¼ zeros (N,N);
A¼ zeros (N,N);
Px¼100; %Define the periods of the gratings
Py¼100;
FFx¼0.5; %Define fill factors
FFy¼0.5;
% Constructing the grating
tic
for p¼1:N;
for q¼1:N;
if rem(q,Px),Px*FFx;
A1(p,q)¼1;
end
if rem(p,Py),Py*FFy;
A2(p,q)¼1;
end
end
end
A¼exp(1i*pi*xor(A1,A2));%% XOR operation between A1 and A2
toc
elegant. Using the command ‘tic-toc’, one can determine how long the
‘Constructing the grating’ module runs in both cases. The version without the
for and if loops is already an order of magnitude faster for this file size. The
module ‘Observing the grating output in the far field’ is not shown, as it is
identical to the corresponding module in Table 2.2.
Table 2.7 MATLAB code for design of a checkerboard grating (version 2).
% Version 2: Checker pattern created without loops
%%Checkerboard phase grating%%
clear; %Clear all memory
% Defining Grating Parameters
N¼500; %Define Matrix size
Px¼100; %Define the periods of the gratings
Py¼100;
FFx¼0.5; %Define fill factors
FFy¼0.5;
%Define one unit of the grating
tic
O¼ones(FFy*Py, FFx*Px)*exp(1i*pi);
Z¼ones(FFy*Py, Px-FFx*Px);
O1¼ones(FFy*Py, Px-FFx*Px)*exp(1i*pi);
Z1¼ones(FFy*Py, FFx*Px);
unit¼[Z O; O1 Z1];
s¼size(unit)
%Constructing the entire grating
A¼repmat(unit,N/s(1),N/s(2)); %replicate to create a 2D
grating
toc
When either version of the codes from Table 2.6 and Table 2.7 is run,
along with the part that generates the output, the diffraction orders will
appear as seen in Fig. 2.16. The efficiency values displayed in the figure
window are in complete agreement with the values calculated using Eqs. (2.9)
and (2.10).
Hence, from the above discussions, care must be taken when designing
elements in MATLAB in order to obtain the correct simulation results.
Figure 2.18 (a) Image of the circular grating and (b) the ring diffraction patterns with and
without zero padding.
phase with respect to the light diffracted by the zones with even zone numbers.
Such an FZP has only 10% efficiency. This disadvantage can be overcome by
making a phase element, with the addition of phase p to either the even or odd
zones using a refractive material of calculated thickness and refractive index
(such as photo resist or electron beam resist). Such binary phase FZPs have a
relatively higher efficiency of 40%.
The path difference equation for the design of a 1D FZP is given in by
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
f 2 þ x2n f ¼ nl: (2.11)
Solving Eq. (2.11) and rearranging the terms, we obtain the width of the
grating lines xn as
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n2 l2
xn ¼ þ nf l: (2.12)
4
The first term inside the square root is negligible compared to the second
term. Hence, the approximate value of the widths of the grating lines is
given by
pffiffiffiffiffiffiffiffi
xn ≅ nf l: (2.13)
The 1D FZP has a line focus similar to that of a cylindrical lens.29 The
MATLAB code for simulation and analysis of a 1D phase FZP is shown in
Table 2.9. The image of the 1D FZP and its diffraction pattern are shown in
Figs. 2.20(a) and (b), respectively. The same MATLAB code with some
modifications described in the previous cases can be used for the design of the
amplitude version of the grating. A 2D FZP can be constructed by designing
two FZPs orthogonal to each other and combining them using the XOR
operation discussed in Section 2.1.3.
A circular FZP can be designed using a similar procedure. In that case, the
radius rn of each zone will be defined instead of the distance away from the
axis in one direction:
pffiffiffiffiffiffiffiffi
rn ≅ nf l: (2.14)
The MATLAB code for a circular FZP can be designed by replacing the
x or y coordinate in the nested-for-loop section using radial coordinates.
The circular FZP design part using nested-for-loops is shown in Table 2.10.
The M value is modified from 50 to 32 in order to fit the circular FZP to the
matrix size.
Figure 2.20 (a) Image of the 1D FZP and (b) its diffraction pattern.
The image of the generated circular FZP is shown in Fig. 2.21. In the case
of a circular FZP, the first diffraction order (either þ1st or –1st) is the center
spot, while the other diffraction orders (3rd, 5th) with larger diffraction
angles appear as circles around the central diffraction order.
The MATLAB code can be modified to design FZPs with other
geometries such as elliptical FZPs, 2D FZPs, etc., or even elements with
sinusoidal phase variations.
Table 2.10 MATLAB code for design of a circular binary phase FZP.
%%Circular FZP%%
% Constructing the FZP
for n¼1:M; %Calculate the width of the grating lines
r1(n)¼sqrt(n*f*lambda);
end
for n¼1:2:M;
for p¼1:N;
for q¼1:N;
r(p,q)¼sqrt((p-N/2)*(p-N/2)þ(q-N/2)*(q-N/2));
if r(p,q) . r1(n) && r(p,q) , r1(nþ1);
A(p,q)¼exp(1i*pi);
end
end
end
end
Figure 2.21 Image of the circular FZP with a focal length of 3 mm generated using the
MATLAB code in Table 2.10.
2.2 Conclusions
In this chapter, the designs of basic DOEs such as different types of gratings
and FZPs have been presented. In each case, the fundamental concepts are
introduced in such a way that a novice programmer can begin making his or
her own DOE designs almost immediately. The MATLAB codes are
discussed in detail so that each line of code makes sense. After completing
the chapter, the reader is encouraged to design other complex periodic
structures as well.
2.3 Exercises
E.2.1 Design a 1D phase grating to generate diffraction spots such that the
relative intensity of 1st diffraction order is only 50% that of the 0th diffraction
order. Verify the result by estimating the efficiency values.
E.2.3 Design a triangular grating array with a period of 100 mm along both
the x and y directions and calculate its far-field diffraction pattern using
MATLAB.
E.2.4 Design a 2D phase FZP with a focal length of 3000 mm along the
x direction and 6000 mm along the y direction.
E.2.5 Design an elliptical FZP with a 1.5 ratio of maximum to minimum focal
length values.41,42
E.2.6 Design an amplitude axicon array in which the diameter of each element
is 100 mm and the period is 10 mm.43,44
References
1. G. G. Lendaris and G. L. Stanley, “Diffraction-pattern sampling for
automatic pattern recognition,” Proc. IEEE 58, 198–216 (1970).
2. J. W. Goodman, Introduction to Fourier Optics, Second Edition,
McGraw-Hill Companies Inc., New York (1996).
3. B. C. Kress and P. Meyrueis, Applied Digital Optics, John Wiley & Sons,
Chichester, UK (2009).
4. B. Zhang and D. Zhao, “Focusing properties of Fresnel zone plates with
spiral phase,” Opt. Express 18, 12818–12823 (2010).
5. L. Rayleigh, “XII. On the manufacture and theory of diffraction-
gratings,” Philos. Mag. 47, 81–93 (1874).
6. J. B. Keller, “Geometrical theory of diffraction,” J. Opt. Soc. Am. 52,
116–130 (1962).
7. H. Kogelnik, “Coupled wave theory for thick hologram gratings,” Bell
Systems Technical Journal 48, 2909–2947 (1969).
8. J. M. Cowley and A. F. Moodie, “Fourier images IV: The phase grating,”
Proc. Phys. Soc. 76, 378–384 (1960).
9. T. A. Shankoff, “Phase holograms in dichromated gelatin,” Appl. Opt. 7,
2101–2105 (1968).
10. K. Knop, “Rigorous diffraction theory for transmission phase gratings
with deep rectangular grooves,” J. Opt. Soc. Am. 68, 1206–1210 (1978).
45
Hence, from Eq. (3.1), the efficiency of phase gratings with g ¼ 2, 4, 8, 16, 32,
etc., levels are 41%, 81%, 95%, 99%, and .99%, respectively.
However, the height of each phase level is crucial in achieving the
calculated efficiency values. As was done in the binary case, designing the
multilevel structure for maximum efficiency will be done by cancelling the 0th
diffraction order. Hence, it is necessary to carefully choose the parameters
relating to the areas of different phase values in the element.
Figure 3.1 (a) Phase values of the 4-level 1D phase grating and (b) their corresponding
resist thickness values for l ¼ 633 nm.
Table 3.1 MATLAB code for design and analysis of a 4-level phase grating.
%% 4-level 1D grating
clear %Clear all memory
%% Defining grating parameters
N¼500; % Matrix size
P¼100; % Grating period
A1¼ones(P,N); % Size of fundamental building block of grating
g¼4; % Number of phase levels
delphase¼ 2*pi/g; %Phase step size
% Constructing one n-level section of the phase grating
sub ¼round(P/g)-1;
for count ¼ 1:g;
A1((count -1)*subþ1:count*sub,:)¼exp(1i*(count-1)*delphase);
end
%Constructing the full grating
A2¼repmat(A1,N/P,1);
%Observation of the diffraction pattern
%See appendix for function definition
Norm_outputP(A2, N)
The phase values 2p (or 0) and p/2 have opposite orientations with respect
to the phase values p and 3p/2, respectively. Hence, light emanating at these
phase levels cancel each other at the 0th diffraction order. Let us verify the
efficiency values obtained from Eq. (3.1) with MATLAB simulation. The
MATLAB code for design of a 4-level phase grating is shown in Table 3.1.
The profile of the 1D 4-level phase grating and the diffraction spots can be
displayed using the standard MATLAB codes presented in Chapter 2. The
image of the phase variation of the 1D 4-level phase grating and the zoomed-
in figure of the diffraction spots are shown in Figs. 3.2(a) and (b), respectively.
By clicking on the diffraction spot using the data cursor of the MATLAB
window, the efficiency value can be read as 81%, matching the value
calculated using Eq. (3.1). Varying the period of the grating such that the
areas of the four levels are not equal will give a lower efficiency value
depending upon the discrepancy. In multilevel structures it might be useful to
Figure 3.2 (a) Phase profile of a 4-level 1D phase grating generated by the output function.
(b) Diffraction spots in the zoomed image.
Figure 3.3 Phase profile of a 4-level phase grating plotted across the y direction.
Figure 3.4 Phase values and resist thickness values of an 8-level 1D phase grating.
Figure 3.5 (a) Plot of an 8-level 1D phase grating and (b) the zoomed-in diffraction pattern.
Figure 3.7 (a) Phase profile of the 8-level axicon and (b) its diffraction pattern.
is given in Table 3.4. The phase profile of the blazed grating is shown in
Fig. 3.8. The diffraction pattern is a ring pattern, and the efficiency was
measured to be 98%, which is close to the expected 100%. The phase profile in
the figure may seem incorrect, but, as discussed earlier, in MATLAB, phase
values above p will be converted to a negative phase, resulting in a phase
profile range [–p, p].
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
f 2 þ r2 f ¼ nl, (3.2)
where f is the focal length of the FZP, and l is the wavelength. The phase
profile is given by
Figure 3.9 Phase profile of a FZP (a) before (dashed line) and (b) after discretizing into
modulo-2p zones (solid line).
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2p
FFZP ðrÞ ¼ f þr f
2 2
: (3.3)
l 2p
Figure 3.10 Image of the phase profile of a blazed FZP generated using MATLAB.
The image of the phase profile of the FZP is shown in Fig. 3.10. The code can
be modified for an elliptical FZP, as presented in Chapter 2, by stretching
either the x or the y coordinate with respect to the other.
Figure 3.11 Phase profile of (a) a blazed FZP (dashed line) and its (b) 4-level
approximation (solid line).
L and represents the number of spiral cycles the beam makes within a distance
of the wavelength of light l. The charge L equals 1, 2, 3, . . . p, depending on
whether the beam’s phase varies by 2p, 4p, 6p, . . . 2pp along the distance l.
The diameter of the donut beam increases with increasing topological charge
value.20,21 Recently, SPPs with fractional topological charges have attracted
attention due to their generated asymmetric donut patterns, which are useful
for optical trapping applications.22–24
When a plane wave is incident on the SPP, it generates a helical beam with
a number of azimuthal rotations matching L. The fabrication of a grayscale
SPP is a challenge; hence, in most cases, only multilevel SPPs have been
designed and fabricated.20,21 The design of such elements is similar to that of
gratings except that the multiple levels are made along the azimuthal
direction. The phase profile of a SPP with topological charge L is given by
The phase step size of a multilevel SPP with g phase levels is given by
2p
DFSPP ðuÞ ¼ L: (3.5)
g
Figure 3.13 Image of the donut patterns generated for (a) L ¼ 1, (b) L ¼ 2, (c) L ¼ 3, and
(d) L ¼ 4.
Figure 3.14 Diffraction patterns of spiral phase plates with topological charges L ¼ 0.25 to
2 in steps of 0.25.
The images of the asymmetric donut beams generated for L ¼ 0.25, 0.5, 0.75,
1, 1.25, 1.5, 1.75 and 2 are shown in Fig. 3.14.
Figure 3.15 Diffraction patterns of spiral phase plates with topological charge L ¼ 1 and
phase levels (a) g ¼ 2, (b) g ¼ 3, (c) g ¼ 4, and (d) g ¼ 5.
where f0 is the focal length, Δf is the focal depth, and R is the radius of the
element. The phase profile of an axilens can be then given by
Figure 3.16 Optics configuration for focusing a plane wavefront using an axilens.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2p
FAxilens ðrÞ ¼ f ðrÞ þ r f ðrÞ
2 2
: (3.8)
l 2p
The MATLAB code for design of a blazed axilens is given in Table 3.9.
The element was designed with f0 ¼ 5 mm, Df ¼ 1 mm, and radius R ¼ 1 mm.
The image and phase profile of an axilens are shown in Figs. 3.17(a) and (b),
respectively.
Figure 3.17 (a) Image and (b) plot of the phase profile of an axilens.
Figure 3.19 Images of the logo generated from (a) a grayscale DOE and (b) a binary DOE.
iterations before and after binarization of the phase profile are shown in
Figs. 3.19(a) and (b), respectively. The loop termination can be specified by
the number of iterations or by the average error between the ideal and iterated
images.
In Fig. 3.19(b), twin images of the logo are formed as the DOE has
centrosymmetric support.33 The twin image is flipped along both the
horizontal and vertical direction, as expected. The lateral shift is not seen
because the object is aligned with the optical axis. If an object that is not
collinear with the optical axis is used, the twin image will occur laterally
shifted along the horizontal and vertical directions. The twin image can be
removed by approximating the grayscale DOE with a 4-level DOE or with a
higher number of phase levels. The problem of twin images arises in many
different fields. Research is still ongoing in this area, and several ways of
removing one of the twin images have been reported.34–37 Higher diffraction
orders can be seen by zero padding the image and calculating the phase using
the algorithm. However, the efficiency of these higher orders is poor (only
4.5% distributed over the entire image).38
Although the IFTA and similar algorithms are very powerful (they can be
used to create DOES that generate almost any intensity pattern), their main
problem is that there is no specific phase relationship from one pixel to
another. This leads to scattering and, hence, an overall decrease in efficiency
and contrast of the final image.
Alternate techniques that avoid these problems can be used at least for
generation of specific intensity profiles. One such technique is described in the
next section.
are used to connect the input beam with the desired output beam. The
fundamental postulate of geometric optics is the Eikonal equation:40
ð∇SÞ2 ¼ n2 , (3.9)
where n is the refractive index. The surfaces S(r) ¼ constant are called the
geometrical wavefronts. Geometrical light rays are perpendicular to the
wavefronts and indicate the direction of flow of energy. Initially, the areas
over the incident and output beams are divided into a mesh, i.e., a number of
zones of equal power. The input and output mesh nodes are connected using
Eikonal equations. The phase distribution required to produce the desired
output is obtained by solving these equations.41
The basic idea of this technique is that the energy in each zone in the input
plane is redirected to a similarly located zone in the output plane. Since the
equations are solved simultaneously for the entire DOE, the obtained phase is
continuous.
w20
PTOT ¼ p I , (3.10)
2 0
where w0 is the beam waist of the Gaussian beam, and I0 is the on-axis
intensity. Therefore, the power in a circle of radius r is
r2
Pr ¼ 1 exp 2 2 PTOT : (3.11)
w0
The beam is then split into a number of rings of equal power. If there are
n1 (¼ I/2) rings, the power in each ring will be Pr/n1. Normalizing the power
such that PTOT ¼ 1, the radius of the nth ring is given by
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi
1 1
rn ¼ w0 ln , (3.12)
2 1 nPr ∕n1
where, n varies from 1 to n1. Figure 3.20 shows a Gaussian beam with annular
rings of equal power. [This is the output curve obtained as ‘figure(2)’ in
the MATLAB code given in Table 3.11.] These rings will eventually form part
of the mesh and, hence, are plotted only from the starting node to the final
node, which is the reason for the apparent gap near 0 on the y axis.
Figure 3.20 Creating the input mesh. Step One: annular rings (reprinted from Ref. 41).
Table 3.11 Creating the input mesh for the Eikonal technique when the incident light is
Gaussian.
%program creates a mesh of zones of equal energy for a Gaussian
%beam
%mesh is rotated away from axes to avoid zeros
clear all
close all
% all units in mm
global I J R xh yh sigma
sigma¼0.8; % Gaussian spot size
R¼1; % radius of beam
%mesh co-ordinates
I¼20; %related to rows and therefore also to y coords
J¼20; %related to columns and therefore also to x coords
N¼(I*J); %total number of zones
TP¼pi*sigma*sigma/2; %total power
P¼(1-exp(-2*R*R/(sigma*sigma)))*TP; %total power in radius R
rings¼I/2; %no of rings
Pring¼P/rings; %Power per ring
%calculating radii of rings
n¼1:rings;
r¼sigma*sqrt(0.5*log(1./(1-n.*Pring)));
%%add very small value for first ring (close to zero)
rad(1)¼1e-10;
i¼1:n(end);
rad(iþ1)¼r(i);
r¼rad;
% number of zones in a ring-decided by angles chosen
angle_step¼pi/I;
angle_shift¼pi/100;
(Continued )
In order to obtain areas of equal power, the annular rings are further
divided into zones. These zones can easily be made to have equal power by
drawing radial lines from the origin. In these few steps, the Gaussian circle has
now been divided into a number of zones of unequal area but equal power.
Table 3.11 gives the code that carries out this operation for a Gaussian beam.
Figure 3.21 shows a Gaussian beam that has been divided using this technique.
The black points in Fig. 3.21 indicate the corners or nodes of each zone.
In order to program the Eikonal technique correctly, certain factors must be
taken into account. Firstly, no zeros should occur in the x and y coordinates,
Figure 3.21 A Gaussian beam divided into 20 20 nodes (reprinted from Ref. 41).
Figure 3.22 Output mesh consisting of 20 20 nodes. All zones have equal area
(reprinted from Ref. 41).
coordinates over the DOE. The corresponding coordinates at the image plane
are a function of ū and are given by x(ū). If the distance between the DOE and
image plane is d, then the gradient of the wavefront will be given by
u xðuÞ
∇c ¼ : (3.14)
d
Table 3.13 Using the Eikonal method to generate the DOE phase.
%This program reads mesh data from the files GaussianMesh and
%square
%Phase over the hologram is calculated using the Eikonal technique
%Phase is calculated at a limited number of irregularly spaced
%data points
close all
clear all
load GaussianMesh
load square
bits¼512;
x¼xsq;
y¼ysq;
%inputs
f¼300; %distance between input and output planes in mm
lambda¼0.000633; %wavelength in mm
figure(1)
plot(xh,yh,’r.’)
title(‘Mesh over hologram plane’)
%Eikonal retrieval
%delta psi
del_psix¼(x-xh)./f;
del_psiy¼(y-yh)./f;
%polynomial calculation to obtain phase of hologram
D¼5; %polynomial of degree D
M¼(Dþ2)*(Dþ1)/2;
%calculates k and l values for each m value
count¼0;
for k¼0:D
for l¼0:D-k
count¼countþ1;
m¼k*((2*D-kþ3)/2)þlþ1;
indices(count,1)¼m;
indices(count,2)¼k;
indices(count,3)¼l;
end
end
(Continued )
The output of the code given in Table 3.13 is the phase of the hologram that
will carry out the conversion from a Gaussian beam to a square-shaped beam
with constant intensity. The final figure of this code is shown in Fig. 3.23.
The phase retrieved using the polynomial method only gives discrete
values at the specified mesh nodes. For the circular Gaussian input, the nodes
are not equidistant and the number of nodes is also quite small. The phase will
not provide enough data with which to write the DOE. Also, to simulate the
output, we will be using a Fourier transform technique and, hence, need
evenly spaced points. Therefore, phase values at many more equidistant
points are required. There are several ways this can be achieved. MATLAB
Figure 3.23 Phase that converts a circular Gaussian beam into a flat-top intensity beam
with a square shape (reprinted from Ref. 21).
Table 3.14 Creating phase information at equally spaced points and simulating the output
intensity.
% This program calculates the phase over a grid
%loads uniformly spaced x and y coordinates
%calculates coefficients using Matlab fit
clear all
close all
load phase_circle %contains regularly spaced grid points
choice ¼ input(‘Enter 1 to generate fit coefficients in matlab and
2 to upload coefficients from other programme: ‘);
if choice ¼¼ 1
ft ¼ fittype(‘poly55’);
dataFit ¼ fit([xh1,yh1],psi1, ft);
coeff1¼coeffvalues(dataFit);
coeffN¼coeffnames(ft);
R¼confint(dataFit);
c¼size(coeff1);
psi1¼0;
for count¼1:c(2)
coeffI¼coeffN(count);
coeffI¼coeffI{:};
m¼str2double(coeffI(2));
n¼str2double(coeffI(3));
psi1¼psi1þcoeff1(count).*(XH1.^m).*(YH1.^n);
end
elseif choice ¼¼2
%coefficients obtained with R¼1, sigma¼0.8 and I¼J¼20, S¼2 from
%a data handling software such as excel
%SQUARE
a0 ¼ 0.106394007
a1 ¼ -0.836150168
a2 ¼ 0.020604165
a3 ¼ 17.61722947
a4 ¼ 15.83427723
a5 ¼ -7.266911537
a6 ¼ -4.752015829
(Continued )
Figure 3.24 Simulated intensity at output plane. The phase used to create this was
generated using choice 2 in the code of Table 3.14 (reprinted from Ref. 41).
3.4 Conclusions
In this chapter, the design, simulation, and analysis of simple multilevel and
grayscale elements are discussed with MATLAB codes. We have introduced
some methods to design diffractive optics for the generation of any arbitrary
intensity profile at a plane of interest. We have also discussed the creation of
special beams, such as donut beams, and beam-shaping fundamentals.43 In
addition to designing the diffractive element, in every case the output is also
simulated.
The types of diffractive optics presented in this chapter are by no means
comprehensive. There are many other interesting and useful DOEs such as
Dammann gratings, harmonic gratings, and spot array generators, to name a
few. However, the techniques introduced here can be adapted to any type of
binary or multilevel DOE.
3.5 Exercises
E.3.1 Design a 16-level 1D phase grating using MATLAB and estimate the
efficiency in the 1st diffraction order.
E.3.2 Design a negative 4-level FZP and simulate its far-field diffraction
pattern.
E.3.3 Design a binary spiral phase plate with L ¼ 5 and simulate its diffraction
pattern.
E.3.4 Design a 3-level DOE from an object of your choice and simulate its
diffraction pattern.
E.3.5 Design a gradient ring lens with a focal length f ¼ 2 mm and a ring
diameter r0 ¼ 0.1 mm.
References
1. M. Breidne, S. Johansson, L.-E. Nilsson, and H. Åhlèn, “Blazed
holographic gratings,” Opt. Acta 26, 1427–1441 (1979).
2. J. W. Goodman, Introduction to Fourier Optics, Second Edition,
McGraw-Hill Companies Inc., New York (1996).
3. B. C. Kress and P. Meyrueis, Applied Digital Optics, John Wiley & Sons,
Chichester, UK (2009).
4. J. Jahns and S. J. Walker, “Two-dimensional array of diffractive microlenses
fabricated by thin film deposition,” Appl. Opt. 29, 931–936 (1990).
5. S. Beretta and M. Cairoli, “Design of multilevel thin-film phase gratings
with optimized diffraction efficiency,” Opt. Lett. 16, 684–686 (1991).
ð(
ð þ`
þ` )
ejkz ej 2zðx þy Þ
k 2 2
k 2
Eðu,vÞ ¼ Aðx,yÞ exp j ðx þ y2 Þ
jlz 2z
` `
2p
exp j ðxu þ yvÞ dxdy: (4.1)
lz
77
From this equation it can be noted that the parabolic phase factor
ðk∕2zÞðx2 þ y2 Þ is multiplied with every pixel of the DOE function A(x, y) and
Fourier transformed. The equation is different from a pure Fourier transform
because of the parabolic phase factor ej 2zðx þy Þ and constant phase ejkz outside
k 2 2
of the integral. If we consider a point object in A(x, y), the resulting phase in
the plane (u, v) from Eq. (4.1) is nothing but a FZP.2 Hence, it can be
understood that, for any arbitrary object, the field in the observation plane is
nothing but a collection of FZPs. Since the parabolic phase factor for large
values of z reduces to 1, the equation reduces to a Fourier transform
operation, which is the far-field approximation (large z) of the scalar
diffraction formula.
As explained in Chapter 1, the Fresnel approximation is valid only when
2
p
z ≫
3
ðx uÞ þ ðy vÞ
2 2
: (4.2)
4l max
Hence, this approximation is not suitable for studying the field extremely close
to the diffraction element. A deeper look at the approximation reveals that it
is valid only over a smaller area (in the transverse plane) for shorter distances,
while it is valid over a larger area at longer distances. In other words, the
Fresnel approximation can be valid for distances very close to the diffraction
plane, but this validity is limited to a transverse area proportional to the
distance. The far-field approximation, also known as the Fraunhofer
approximation, is valid only for very large distances given by
k u2 þ v2
z≫ , (4.3)
2
Table 4.1 MATLAB code for simulation of Fresnel diffraction of a circular aperture.
%Fresnel diffraction of a circular aperture%
clear; %Clear all memory
%Defining the parameters
N¼500;% Define the matrix size
lambda¼0.633*10^-6;%Define wavelength in meters
z¼0.02;%Propagation distance ¼ 20 mm
r¼10^-3;%Radius of aperture ¼ 1 mm
wsamp¼10*lambda;%sampling period or width
%Creating sampled space
x¼1:N;
y¼1:N;
[X,Y]¼meshgrid(x,y);%Sampling
Rsamp¼sqrt((X-N/2).^2þ(Y-N/2).^2).*wsamp;%Define sampled
%radius
%Constructing the aperture
A¼ones(N,N);%Define matrix by assigning zeros to all pixels
A(Rsamp.¼r)¼0;
% Calculating the Fresnel diffraction pattern
PPF¼exp(1i*pi/(lambda*z).*Rsamp.*Rsamp); %Calculate the
%parabolic phase factor
A1¼A.*PPF;%Multiply the circular aperture function with the
parabolic phase factor
E¼abs(fftshift(fft2(fftshift(A1)))); %Calculate Fourier
%Transform
%Observation of the diffraction pattern
colormap(gray)%Display greyscale image
imagesc(E)%Display scaled image
Figure 4.1 Fresnel diffraction patterns for circular aperture at distances (a) z ¼ 20 mm,
(b) z ¼ 50 mm (c), z ¼ 100 mm, and (d) z ¼ 50 m.
not increased at large distances, as the period of the binary axicon itself was
selected to obtain larger ring patterns. However, if the period of the binary
axicon were larger, i.e., generating a smaller ring pattern, then the sampling
period would need to be increased in order to see the diffraction pattern.
The codes given in Tables 4.1 and 4.2 can be modified to simulate the
Fresnel diffraction patterns of any type of amplitude or phase DOE. For
example, the DOE design scripts presented in Chapter 2 and in the first
section of Chapter 3 can be easily integrated into these codes and the Fresnel
diffraction patterns simulated at different planes.
Table 4.2 MATLAB code for simulation of Fresnel diffraction of a binary phase axicon.
%Fresnel diffraction of binary axicon%
clear; %Clear all memory
% Defining the parameters
N¼500;% Define the matrix size
lambda¼0.633*10^-6;%Define wavelength in meters
z¼0.025;%Propagation distance ¼ 25mm and 50 m
P¼10^-4;%Radius of aperture ¼ 0.1 mm
wsamp¼10*lambda;%sampling period or width
%Creating sampled space
x¼1:N;
y¼1:N;
[X,Y]¼meshgrid(x,y);%Sampling
Rsamp¼sqrt((X-N/2).^2þ(Y-N/2).^2).*wsamp;%Define sampled
%radius
%Constructing the DOE
A¼ones(N,N);%Define matrix by assigning ones to all pixels
A(rem(Rsamp,P) , P/2)¼exp(1i*pi);
%Calculating the Fresnel diffraction pattern
PPF¼exp(1i*pi/(lambda*z).*Rsamp.*Rsamp); %Calculate the
parabolic phase factor
A1¼A.*PPF; %Multiply the circular aperture function with the
%parabolic phase factor
E¼abs(fftshift(fft2(fftshift(A1)))); %Calculate Fourier
%Transform
Figure 4.2 (a) Fresnel diffraction pattern of a binary phase axicon at z ¼ 25 mm, (b) its
intensity profile, and (c) the Fresnel diffraction pattern at z ¼ 50 m.
The axial Fresnel diffraction pattern for the above design values and focal
depth values of 0 mm, 1 mm, 2 mm, and 3 mm are shown in Figs. 4.3 (a)–(d),
respectively. They were simulated using the code in Table 4.3. The design
concept of the axilens has been utilized to design DOEs with interesting
characteristics.17–19
Figure 4.3 Images of Fresnel diffraction along the axial direction for an axilens of f ¼ 5 mm
and (a) Df ¼ 0 mm, (b) Df ¼ 1 mm, (c) Df ¼ 2 mm, and (d) Df ¼ 3 mm.
a certain intensity pattern. For example, we can consider an object such as the
logo of Indian Institute of Technology Madras. The phase of the object is
calculated using the MATLAB code given in Table 3.10, assuming input light
with uniform illumination. The object and the phase of the DOE designed
using the algorithm are shown in Figs. 4.4(a) and (b), respectively.
Now, let us study the Fresnel diffraction pattern of the phase-only DOE,
assuming uniform illumination. Due to the very different nature of the DOE,
the Fresnel diffraction pattern is quite different from any of the patterns seen
in earlier sections. Another difference is the manner in which the Fresnel
images need to be obtained, which is akin to imaging an object and studying
the image at different planes. In order to carry out this experiment, it is
necessary to multiply the phase function generated by the algorithm by a lens
function. The resulting DOE when subjected to Fresnel diffraction analysis
will yield the correct image at a plane corresponding to the focal plane of the
lens, while the image will be blurred elsewhere. Hence, in the first step, the
phase function shown in Fig. 4.5(b) is multiplied by a lens function of focal
Figure 4.4 Images of the (a) object and (b) phase matrix generated using the Gerchberg–
Saxton algorithm.
Table 4.4 MATLAB code for analyzing a DOE generated by Gerchberg–Saxton algorithm.
%% Fresnel diffraction of a DOE designed by Gerchberg-Saxton
%algorithm
%% Use table 3.10 to obtain the “DOE phase”
% Multiply the “DOE phase” with a lens function
d¼0.2;
Q¼exp(-1i*(pi/(lambda*d))*(X.^2þY.^2));
DOEphase¼DOEphase.*Q;
%Defining all parameters and sampling
lambda¼0.633e-6; %Wavelength
wsamp¼1e-6;%sampling period
x¼-N/2:N/2-1;
y¼-N/2:N/2-1;
[X,Y]¼meshgrid(x*wsamp,y*wsamp);
Rsamp¼sqrt(X.^2þY.^2);
%Calculation of Fresnel diffraction pattern
m¼100;
n¼1:m;
zs2¼0.1þ(0.2/m).*n; % Propagation distance
PPF¼zeros(N,N,m);
A1¼zeros(N,N,m);
E¼zeros(N,N,m);
Field1¼zeros(100,m);
for counter1¼1:10:m;
PPF(:,:,counter1)¼exp(1i*pi/(lambda*zs2(counter1)).
*Rsamp.*Rsamp); % Parabolic phase factor
A1(:,:,counter1)¼DOEphase.*PPF(:,:,counter1); %Multiply
%the axilens function with the parabolic phase factor
E(:,:,counter1)¼abs((fft2((A1(:,:,counter1)))));
%Calculate Fourier Transform
imagesc(E(:,:,counter1)); title (zs2)
pause(1.0)
end
Figure 4.5 Images of the Fresnel diffraction pattern of a DOE generated using Gerchberg-
Saxton algorithm with a lensing function of focal length 0.2 m recorded at (a) z ¼ 0.02 m,
(b) z ¼ 0.05 m (c) z ¼ 0.15 m and (d) z ¼ 0.2 m.
2L2
zT ¼ , (4.4)
l
where, L is the period of the grating. The pattern also repeats at the distance
zT/2 but is shifted laterally by a distance equal to half the period of the grating.
The MATLAB code for simulation of the Talbot effect, as seen with an
amplitude grating, is shown in Table 4.5.
An image of the simulated grating is shown in Fig. 4.6(a). The intensity at
the quarter, half and full Talbot planes can be seen in Figs. 4.6(b), (c,) and (d),
respectively.
At fractional Talbot distances, the grating pattern repeats, albeit with
different frequencies. This gives rise to a very beautiful phenomenon called the
Talbot carpet.26,27 The program in Table 4.4 can be modified to calculate and
create a Talbot carpet. The authors leave this to the readers as an exercise. It is
interesting to play with some of the parameters in the program to see how this
Fresnel diffraction changes. For example, decreasing the period makes the
Talbot effect difficult to observe, as the overlap of orders happens only very
close to the grating. This is schematically shown in Fig. 4.7.
The shaded triangles in Fig. 4.7 indicate the regions of overlap between
the first orders. It is clear that there is a larger overlap in the case of the second
grating, as L2 is larger; therefore, the diffraction angles are smaller. The
decreasing area of overlap at distances farther from the grating can be seen in
Figs. 4.6(b), (c), and (d).
Clearly, the Talbot effect will be easier to observe for gratings with large
periods. However, when simulating a grating with a large period, the matrix
has to be large enough that an adequate number of periods occurs in the
grating, yet the window around the grating has to be large enough to simulate
the diffraction at different distances along the optical axis. Readers are
encouraged to vary the parameters P, N, and w in the code of Table 4.4, and
to observe the resulting changes to the Talbot pattern. The same code can also
be modified to simulate the Talbot effect with a phase grating.28
Figure 4.6 (a) Image of the amplitude diffraction grating and its diffraction pattern
simulated at (b) z ¼ 0.25 zT (c) z ¼ 0.5 zT, and (d) z ¼ zT.
Figure 4.7 Schematic depiction of first orders from gratings with period (a) L1 and (b) L2.
The periods are sized such that L1 , L2.
4.3 Conclusions
This chapter describes how to model diffraction in regions other than the far
field. Although the programs presented cannot be used to study the region
immediately following the DOEs, they can be used to model the diffraction
field in an intermediate region (see Chapter 1 for validity of approximations at
different axial positions). Two other useful ideas are also discussed, namely,
the concept of simulating the axial intensity profiles and the analysis of
nonperiodic structures. The technique and procedure presented here can be
used for the analysis of the axial intensity profiles for any DOE in the region
of the validity of the approximation.
4.4 Exercises
E.4.1 Calculate the Fresnel diffraction pattern of a blazed axicon at z ¼ 50 m
with a sampling period of 10l.
References
1. J. Dyson, “Circular and spiral diffraction gratings,” Proc. R. Soc. London
Ser. A 248, 93–106 (1958).
2. J. W. Goodman, Introduction to Fourier Optics, Second Edition,
McGraw-Hill Companies Inc., New York (1996).
3. B. C. Kress and P. Meyrueis, Applied Digital Optics, John Wiley & Sons,
Chichester, UK (2009).
4. E. Hecht, Optics, Fourth Edition, Addison Wesley, San Francisco (2002).
5. S. Trester, “Computer-simulated Fresnel diffraction using the Fourier
transform,” Comput. Sci. Eng. 1, 77–83, (1999).
6. J. P. Guigay, “On Fresnel diffraction by one-dimensional periodic objects,
with application to structure determination of phase objects,” Opt. Acta
18, 677–682 (1971).
7. R. P. Muffoletto, J. M. Tyler, and J. E. Tohline, “Shifted Fresnel diffraction
for computational holography,” Opt. Express 15, 5631–5640 (2007).
8. R. F. Edgar, “The Fresnel diffraction images of periodic structures,” Opt.
Acta 16, 281–287 (1969).
9. D. Mendlovic, Z. Zalevsky, and N. Konforti, “Computation considera-
tions and fast algorithms for calculating the diffraction integral,” J. Mod.
Opt. 44, 407–414 (1997).
91
the angles involved are small, which makes it applicable only in the paraxial
case. Therefore, this equation is not valid for larger angles, which arise when
the radius of the beam equals the radius of the lens, or for optical systems with
small object and image distances.
In other words, a FZP designed for focusing a plane wavefront (infinite
conjugate mode) cannot be used for focusing a diverging wavefront (finite
conjugate mode) without aberration. In such cases where the paraxial case is
violated, it is necessary to design a FZP for the finite conjugate mode.
Finite conjugate mode elements have additional advantages when compared
to infinite conjugate mode ones. Most light sources emit diverging wavefronts
and will need additional lenses for collimation. By designing the FZP for finite
conjugate mode, it is possible to avoid using these additional lenses.
Figure 5.1 Optics configuration for focusing a diverging wavefront using a FZP (reprinted
from Ref. 26).
Figure 5.2 Plot of the radii of zones of a FZP as a function of its zone number for
finite conjugate mode (solid line) and infinite conjugate mode (dashed line) for u ¼ 1 mm,
v ¼ 5 mm, and f ¼ 0.83 mm (bottom lines); and u ¼ 5 mm, v ¼ 30 mm, and f ¼ 4.3 mm
(top lines) (reprinted from Ref. 26).
uðr2n n2 l2 Þ
v2 ¼ : (5.9)
n2 l2 þ 2nlu r2n
The MATLAB® code for performing the ray tracing is given in Table 5.1.
The ray tracing images for the above two cases (i.e., with and without
paraxial approximation) are shown in Figs. 5.3 and 5.4, respectively.
Table 5.1 MATLAB code for comparing focusing properties of FZP designed with and
without paraxial approximation.
%%program to compare the FZPs designed with and without paraxial
%%approximation
% Input parameters
u¼5000; % Object distance
v¼30000; % Image distance
l¼0.632; % Lambda-wavelength
% Designing FZPs with n zones
n¼1:200;
a¼n.*n.*l*lþ2.*n.*l*(uþv)þ2*u*v;
r¼sqrt(((a(n).*a(n))-4*u*u*v*v)./(4.*(a(n)þu*uþv*v)));
%Radius of zones
A¼(4.*n.*n.*l*lþ8.*n*l*u-4.*r(n).*r(n));
B¼(4.*n.^3*l^3þ12.*n.*n.*l*l*uþ8.*n.*l*u*u-8.*r(n).*r
(n).*n.*l-8.*r(n).*r(n).*u);
C¼(n.^4.*l.^4þ4.*n.*n.*l*l*u*uþ4.*n.^3*l^3*u-4.*r(n).*r
(n)*u*u-4.*r(n).*r(n).*n.*n*l*l-8.*r(n).*r(n).*n*l*u);
v1¼(-B(n)þsqrt(B(n).*B(n)-4.*A(n).*C(n)))./(2.*A(n));%
%Image distance
the1¼atan(r(n)./v1(n));%Angle theta
v2¼u.*(r(n).*r(n)-n.*n.*l*l)./(2.*n.*l*uþn.*n.*l*l-r(n).
*r(n));
the2¼atan(r(n)./v2(n));%Angle theta
% Ray tracing
for m¼1:100; %Ray tracing with 100 points on each ray
n¼m*2;
for p¼1:11;
v3(p)¼(p-1)*3000;
rr1(p,n)¼(v1(n)-v3(p))*tan(the1(n));
rr2(p,n)¼(v2(n)-v3(p))*tan(the2(n));
end
end
% Display results
n¼10:10:200; % Ray tracing display
plot(v3,rr1(:,n),’k’,’LineWidth’,1)
hold on
plot(v3,rr2(:,n),’b’,’LineWidth’,1)
Figure 5.3 Image of the ray tracing at the image plane for a FZP designed using
Eq. (5.6).
Figure 5.4 Image of the ray tracing at the image plane for a FZP designed using Eq. (5.7).
The magnified images of the focal regions are shown indicating the aberration
in the image plane. It can clearly be seen that the FZP designed using Eq. (5.7)
has aberrations in the focal plane, resulting in a shift in its image distance and
blurring of the spot. If neglected, this one common aberration will result in
substantial aberration in the focal plane for a FZP, resulting in focal spot blur
and shift in focal plane.
Figure 5.5 (a) Parallel rays and (b) diverging rays propagating through a glass substrate.
Figure 5.6 Optics configuration for focusing a diverging wavefront using a FZP fabricated
on a thick glass substrate (reprinted from Ref. 26).
different optical paths within the glass substrate. The light rays exit the glass–
air interface with angles equal to the angles of incidence at the air–glass
interface. Hence, the presence of the glass plate only shifts the location of the
incident rays depending on their angles of incidence.
The variable r1n is the distance from the optical axis to the position where
the ray emanating from the source with an angle u1n meets the front surface of
the glass substrate. The variable r2n is the distance from the optical axis to the
point where the ray emanating from the source with an angle u1n meets the
back surface of the glass substrate. The variable r0 2n is the distance from the
optical axis to the point where the ray emanating from the source with an
angle u1n would meet the back surface of the glass substrate if it were not
present.
From trigonometry we know that
r1n r0
tan u1n ¼ ¼ 2n : (5.10)
ut u
Applying Snell’s law of refraction at the air–glass interface, we obtain
0 ut 1 na
r2n ¼ r 2n þ t tan sin : (5.13)
u ng ð1 þ u2 ∕r022n Þ1∕2
As discussed earlier, the presence of the glass substrate does not alter the
direction of propagation but shifts the point at which the ray is incident on the
FZP. The shift in radial direction at the FZP plane is given by
Dr ¼ r0 2n r2n : (5.14)
Preserving the angle u1n, the distance between the FZP plane and the
source has to be different from u. As a consequence, the presence of the glass
substrate generates a virtual source that is shifted from the real source and has
a finite spread. The position of the different virtual sources depends on the
direction of the rays. The position of the virtual source is given by
r2n
u0 ðr0 2n Þ ¼ u: (5.15)
r0 2n
This is an interesting result. From Eq. (5.15), it can be noted that the presence
of the glass substrate generates a virtual source that is shifted away from the
source and has a finite blur.
Substituting Eq. (5.13) in Eq. (5.15), the limiting value of u0 for u1n!0 is
given by
na
lim u0 ðr0 2n Þ ¼ u t þ t : (5.16)
u1 !0 ng
Figure 5.7 Plot of wavefront shapes at z ¼ –t (solid line), z ¼ 0 without glass substrate
(dashed line), z ¼ 0 with glass substrate (dotted line), and aberration function U (dashed
and dotted line) for (a) u ¼ 5 mm and v ¼ 30 mm and (b) u ¼ 30 mm and v ¼ 5 mm
[part (a) reprinted from Ref. 26].
Figure 5.8 Plot of the locations of virtual sources as a function of radial distance for
(a) case 1 and (b) case 2 [part (a) reprinted from Ref. 26].
From Figs. 5.7(a) and (b) it can be noted that the wavefront aberration is
severe in the case where the object distance is closer to the thickness value of
glass substrate, and vice versa. Hence, the wavefront aberration will be much
higher when the optics configuration is more compact with u ¼ 1 mm and
v ¼ 5 mm.19,20 The plots of the locations of the virtual sources as a function of
the radial distances for case 1 and case 2 are shown in Figs. 5.8(a) and (b),
respectively. The virtual source is not only shifted from the real source, but it
also has a finite spread. The limiting value of the shift occurs at r ¼ 0, which
can be obtained from Eq. (5.16). The limiting values for case 1 and case 2 are
4.650 mm and 29.650 mm, respectively. The spatial spread of the virtual
sources for case 1 and case 2 are 7.65 mm and 0.22 mm, respectively. Once
again, it is shown that when the focal distances are much larger than the
thickness of glass plate, the introduced error is less.
Ray tracing was carried out for these two cases. The MATLAB code for
performing the ray tracing is given in Table 5.2. The ray tracing figures are
shown in Figs. 5.9 and 5.10 for case 1 and case 2, respectively. The magnified
images of the focusing regions are shown and indicate the spread of the virtual
sources.
By comparing Fig. 5.9 with Fig. 5.10, it is clear that the effect of
aberration is less when the object distance is much larger than the thickness of
glass plate, and vice versa. By using Eq. (5.8), it is possible to locate the
position of the images for the cases with and without the glass substrate. To
estimate the location of the image in the presence of the glass substrate, the
variable u can be replaced by the location of the virtual source u0 . Plots of the
variation in the image distance as a function of the radial distance are shown
in Fig. 5.11 and Fig. 5.12 for case 1 and case 2, respectively.
The aberration measured as the shift in the image plane is around 14 mm
in case 1, amounting to an error of 46% in the location of the image, while it is
Table 5.2 MATLAB code for ray tracing from the real and virtual source to the FZP plane.
%%program to perform ray tracing from the real and virtual source
%to the FZP plane
% Input Parameters
u¼5000;%Object distance
t¼1050;%Thickness of glass plate
ng¼1.5;%Refractive index of glass plate
na¼1;%Refractive index of air
N¼na/ng;%Refractive index ratio
U¼u*u;
% Calculating the angles of rays and location of real/virtual
%source
for r¼1:1000;
RR(r)¼r*r;
D(r)¼(sqrt(1þ(U/RR(r))));
A(r)¼(((u-t)/u)*r);
B(r)¼asin(N/D(r));
r1(r)¼A(r)þt*tan(B(r));
u1(r)¼u*(r1(r)/r);
R1(r)¼sqrt(u*uþr*r)-u;
R3(r)¼sqrt((u-t)*(u-t)þr*r)-(u-t);
R2(r)¼sqrt(u1(r)*u1(r)þr1(r)*r1(r))-u1(r);
delR(r)¼R1(r)-R2(r);
delu(r)¼(u-u1(r));
ratio(r)¼r1(r)/r;
theta(r)¼r1(r)/u1(r);
theta1(r)¼r/u;
end
% Ray Tracing and plotting of results
for r¼1:100:1000;
for x¼1:u1(r)þ1;
y(x)¼x-1;
r11(x)¼(x-1)*tan(theta(r));
end
plot(yþ(u-u1(r)),r11,’k’,’LineWidth’,1)
hold on
end
for r¼1:100:1000;
for x¼1:uþ1;
y(x)¼x-1;
r12(x)¼(x-1)*tan(theta1(r));
end
plot(y,r12,’–k’)
hold on
end
only 10 mm with negligible error in case 2. The spreads of the image points in
case 1 and case 2 are 350 mm and 0.3 mm, respectively. The analysis can be
repeated by performing ray tracing for both cases. Ray tracing is repeated
using the procedures shown in Tables 5.1 and 5.2. Results with and without
the glass substrate for case 1 and case 2 are shown in Figs. 5.13 and 5.14,
respectively. It can be seen that the second case the shift in the image plane is
very small (10 mm). The spread is also relatively smaller.
Figure 5.9 Ray tracing of the rays emanating from the real source (dashed line) and virtual
source (solid line) generated due to the glass substrate for case 1 (reprinted from Ref. 26).
Figure 5.10 Ray tracing of the rays emanating from the real source (dashed line) and
virtual source (solid line) generated due to the glass substrate for case 2.
From the above observations it can be noted that for larger optics
configurations with object and image distances much larger than the thickness
of glass substrate, aberrations are less pronounced, while aberrations are quite
severe in compact optics configurations. Aberrations can also be studied in an
alternative way by fixing the object and image distances and varying the
refractive index or the thickness of the glass plate. Using this method will
result in the same conclusion. However, in real applications, it is necessary to
design FZPs with compact optics configurations. In such cases, if the glass
substrate is not taken into account, the DOE may introduce severe
Figure 5.11 Plot of the position of the image for different radial distances in case 1
(reprinted from Ref. 26).
Figure 5.12 Plot of the position of the image for different radial distances in case 2.
Figure 5.13 Ray tracing of the rays from the FZP plane without the glass plate (dashed
line) and with the glass plate (solid line) for case 1 (reprinted from Ref. 26).
Figure 5.14 Ray tracing of the rays from the FZP plane without the glass plate (dashed
line) and with the glass plate (solid line) for case 2.
the thickness and refractive index of the glass substrate are included in the
design of FZP. In the second scheme, a pre-calculated magnification error is
deliberately added during fabrication to compensate for the effect of the
substrate.
1∕2
C 01 2 4u0 ðrn Þ2 v2
r0n ¼ , (5.18)
4½u0 ðrn Þ2 þ v2 þ C 01
Figure 5.15 Plot of the radii of zones of a FZP designed without (solid line) and with
(dashed line) inclusion of a glass plate for case 1 (u ¼ 5 mm and v ¼ 30 mm) (reprinted from
Ref. 26).
Figure 5.16 Plot of the radii of zones of a FZP designed without (gray line) and with
(dashed line) inclusion of a glass plate for case 2 (u ¼ 30 mm and v ¼ 5 mm).
smaller widths of zones are required compared to the case when the glass plate
is not present.
Figure 5.16 shows a good overlap between the plots of the radii of zones
of the FZP designed with and without inclusion of the glass substrate, as the
aberration introduced in this case is much smaller. The rays are traced from
the FZP plane to the image plane. After aberration correction, the magnified
image of the image plane shows no aberration, as expected. The ray tracing
Figure 5.17 Ray tracing of the rays from the FZP plane with a glass plate after aberration
correction for case 1 (reprinted from Ref. 26).
image with higher magnification (as seen in the horizontal axis) is shown in
Figs. 5.17 and 5.18 for case 1 and case 2, respectively. The above results
indicate that this aberration correction method has completely cancelled the
aberration introduced by the glass substrate.
Figure 5.18 Ray tracing of the rays from the FZP plane with a glass plate after aberration
correction for case 2.
associated with each of these techniques. One of the most common errors
occurring during fabrication when using photolithography is the magnifica-
tion error, which results from the gap between the mask plate and the
substrate during exposure.29
From Eqs. (5.3), (5.4), and (5.5), the phase of the FZP can be given as
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
FFZP ¼ Fout Fin ¼ k r2 þ v2 þ k r2 þ u2 ¼ const þ 2 mp, (5.19)
where k ¼ 2p/l. The constant can be estimated by substituting m ¼ r ¼ 0:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
FF ZP ¼ k r2 þ v2 þ k r2 þ u2 ¼ kðu þ vÞ þ 2 mp: (5.20)
In an ideal situation, i.e., in the absence of any errors in design,
fabrication, and testing, the phases of the input and output waves are the same
as those of Fin and Fout, respectively. However, in a practical system, this will
not be the case. The aberrations or errors can be measured with respect to an
output reference wave given by
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
FR ðx,yÞ ¼ k r2 þ v2 : (5.21)
The phase aberration function is defined as the difference between the phase
of the reference and the phase of the output waves, and is given by
Figure 5.19 Phase aberration function plotted as a function of radial distance for radial
magnifications Mr ¼ 0.9 (solid line), Mr ¼ 0.95 (dashed and dotted), Mr ¼ 1.05 (dashed line),
and Mr ¼ 1.1 (dotted line) (reprinted from Ref. 26).
Figure 5.20 Phase of the wave plotted as a function of radial distance for radial
magnifications Mr ¼ 0.9 (solid line), Mr ¼ 0.95 (dashed and dotted line), Mr ¼ 1 (thick line),
Mr ¼ 1.05 (dashed line), and Mr ¼ 1.1 (dotted line) (reprinted from Ref. 26).
correction, as shown in Fig. 5.21. It can be noted from this figure that
this aberration correction scheme (i.e., introduction of magnification error
during fabrication) prevents only the shift of the image plane and not the
spatial spread, which is approximately 200 mm for the given experimental
parameters. This scheme can be applied to case 2 as well, and a similar effect
will be observed.
Figure 5.21 Ray tracing of rays emanating from the FZP plane to the image plane for case
1 (reprinted from Ref. 26).
5.3.3 Discussion
From the above observations, it is obvious that the aberration introduced by a
glass substrate is more pronounced when the thickness of the glass plate is on
the order of the front focal distance. This aberration is, of course, negligible in
the infinite conjugate mode or when the front focal distance is much larger
than the thickness of the substrate. While this correction may seem attractive,
one should keep in mind that configurations with higher magnification will
also have larger aberration. The optics configuration given in case 1 is more
practical and at the same time has a high magnification of 6 (v/u). This results
in an image plane shift of 14 mm. If the magnification of the system were
reduced, say from 6 to 1 (i.e., u ¼ v), the aberration would also reduce. If u ¼
v ¼ 5 mm, the image plane shift would be less than 500 mm, which has only
10% error; if u ¼ v ¼ 30 mm, the image plane shift would be only 350 mm,
which is only 1% error. Hence, by suitable choice of the front focal distance
and the magnification of the system, it is possible to make the aberration
effect almost negligible.
Figure 5.22 Plot of the location of the image plane (solid line) and error in image plane
(dotted line) for variation in object distances.
Table 5.3 MATLAB code for analysis of aberration due to error in object location along the
z direction.
%%program to analyze the aberration due to error in object
%location in z direction
% Input Parameters
u1¼5000; % Object distance in micrometers
v¼30000; % Image distance in micrometers
l¼0.632; % Wavelength in micrometers
t¼1050; % Thickness of glass substrate in micrometers
na¼1; % Refractive index of air
ng¼1.5; % Refractive index of glass
n¼1; % Zone number
a¼n*n*l*lþ2*n*l*(u1þv)þ2*u1*v;
r¼sqrt(((a*a)-4*u1*u1*v*v)/(4*(aþu1*u1þv*v))); %Radius of
first zone before aberration correction
theta¼atan(r/u1);
rr¼r*((u1-t)/u1);
r1¼rrþt*tan(asin(na/ng*(sin(atan(r/u1)))));
u2¼u1*(r1/r);
a1¼n*n*l*lþ2*n*l*(u2þv)þ2*u2*v;
rr¼sqrt(((a1*a1)-4*u2*u2*v*v)/(4*(a1þu2*u2þv*v)));
% Radius of first zone after aberration correction
%Error calculation
s¼1:1000;
u4¼s-501;
u¼5000þu4(s);
u3¼u(s).*(r1/r);
b¼(4*n*n*l*lþ8*n*l.*u3(s)-4*rr*rr);
c¼(4*n^3*l^3þ12*n*n*l*l.*u3(s)þ8*n*l.*u3(s).*u3(s)-
8*rr*rr*n*l-8*rr*rr.*u3(s));
d¼(n^4*l^4þ4*n*n*l*l.*u3(s).*u3(s)þ4*n^.3*l^3.*u3(s)-
4*rr*rr.*u3(s).*u3(s)-4*rr*rr*n*n*l*l-8*rr*rr*n*l.*u3(s));
v1¼(-c(s)þsqrt(c(s).*c(s)-4.*b(s).*d(s)))./(2.*b(s));
ve¼abs(v1(s)-v);
%Display results
plot(u4,v1/1000)
hold on
plot(u4,ve/1000,’r’)
the system is high. The same procedure may be repeated for case 2, and the
effect of aberration with variations in image distance, magnification, and 1/e2
diameter of the beam may be studied.
In the above cases, only one variation was estimated: the variation in
image location and size due to an error in the object position. It is however
necessary to include the aberration introduced in the system, as the system
was designed for one set of u and v values but was implemented using a
different set of u and v values. This not only causes a shift but also introduces
a blur that can be understood from the earlier descriptions. The ray tracing at
the image plane for errors of –500 mm and 500 mm are shown in Figs. 5.25(a)
and (b), respectively.
Figure 5.23 Plot of the magnification of the optics configuration for variation in object
distances.
Figure 5.24 Plot of the location of the 1/e2 diameter (solid line) and error in the 1/e2
diameter (dotted line) for variation in object distances.
Figure 5.25 Ray tracing of rays emanating from the FZP plane when the object distance
has an error of (a) –500 mm and (b) 500 mm.
Figure 5.26 Ray tracing of rays emanating from the FZP plane when the zones are radially
shifted by 100 nm with random zone number.
location of zones by 100 nm. This aberration will result in increase of spot
size in the image plane, resulting in a blur.
Figure 5.27 Plot of image shift from the optical axis in the image plane for a variation in the
object shift from the optical axis in the object plane.
by the product of the object location shift and the magnification of the system,
which gives an approximately linear relationship between the image and
object location shift, as plotted in Fig. 5.27.
Figure 5.28 Plot of the normalized relative intensity of the 0th (dashed line), 1st (solid line),
and 2nd (dotted line) diffraction orders as a function of error in resist thickness.
Figure 5.29 Plot of the normalized relative intensity of the 1st diffraction order for different
percentages of error in the duty ratio of the resist.
5.5 Conclusions
In the above analysis, different types of errors and their influence on various
parameters are studied independently. To characterize an optics configura-
tion, it is necessary to understand the cumulative effect of all of the above
errors.
For instance, the thickness of the resist and the element duty ratio control
only the efficiency of light in various diffraction orders but do not have any
influence on the shift of image plane or blur of spot. The location of zones and
experimental errors do not affect the efficiency of the device but introduce
aberrations such as image plane shift along the longitudinal or transverse
directions and blur.
5.6 Exercises
E.5.1 Design a ring FZP in finite conjugate mode and perform the aberration
correction for the glass substrate on which it is fabricated.
References
1. J. W. Goodman, Introduction to Fourier Optics, Second Edition,
McGraw-Hill Companies Inc., New York (1996).
2. F. A. Jenkins and H. E. White, Fundamentals of Optics, Fourth Edition,
McGraw-Hill Companies Inc., New York (2001).
3. P. L. Knight and A. Miller, The Ray and Wave Theory of Lenses,
Cambridge University Press, Cambridge (1995).
4. J. Upatnieks, A. Vander Lugt, and E. Leith, “Correction of lens
aberrations by means of holograms,” Appl. Opt. 5, 589–593 (1966).
5. M. J. Kidger, Fundamental Optical Design, SPIE Press, Bellingham,
Washington (2001) [doi: 10.1117/3.397107].
6. J. E. Ward, D. C. Auth, and F. P. Carlson, “Lens aberration correction by
holography,” Appl. Opt. 10, 896–900 (1971).
7. N. López-Gil, H. C. Howland, B. Howland, N. Charman, and R.
Applegate, “Generation of third-order spherical and coma aberrations by
use of radially symmetrical fourth-order lenses,” J. Opt. Soc. Am. A 15,
2563–2571 (1998).
8. D. C. Leiner and R. Prescott, “Correction of chromatic aberrations in
GRIN endoscopes,” Appl. Opt. 22, 383–386 (1983).
9. R. E. Hopkins, C. A. McCarthy, and R. Walters, “Automatic correction
of third-order aberrations,” J. Opt. Soc. Am. 45, 363–363 (1955).
10. U. Dutta and L. Hazra, “Monochromatic primary aberrations of a
diffractive lens on a finite substrate,” Appl. Opt. 49, 3613–3621 (2010).
11. U. Dutta and L. Hazra, “Primary chromatic aberrations of a diffractive
lens on finite substrate,” Appl. Opt. 51, 494–497 (2012).
12. J. Sung, M. Pitchumani, and E. G. Johnson, “Aberration measurement of
photolithographic lenses by use of hybrid diffractive photomasks,” Appl.
Opt. 42, 1987–1995 (2003).
13. N. Bokor and N. Davidson, “Aberration-free imaging with an aplanatic
curved diffractive element,” Appl. Opt. 40, 5825–5829 (2001).
14. T. Fujita, H. Nishihara, and J. Koyama, “Blazed gratings and Fresnel
lenses fabricated by electron-beam lithography,” Opt. Lett. 7, 578–580
(1982).
15. K. Kodate, T. Kamiya, Y. Okada, and H. Takenaka, “Focusing
characteristics of high efficiency Fresnel zone plate fabricated by deep
ultraviolet lithography,” Jpn. J. Appl. Phys. 25, 223–227 (1986).
16. T. Shiono, M. Kitagawa, K. Setsune, and T. Mitsuyu, “Reflection micro-
Fresnel lenses and their use in an integrated focus sensor,” Appl. Opt. 28,
3434–3442 (1989).
17. T. Shiono and K. Setsune, “Blazed reflection micro-Fresnel lenses
fabricated by electron-beam direct writing and dry development,” Opt.
Lett. 15, 84–86 (1990).
121
Considering two binary phase profiles, phase addition can be achieved by the
well-known X-OR (exclusive OR) binary operation. You may recall that this
technique was introduced in Chapter 2 for construction of a checkerboard
phase grating. The simplest multifunctional DOE is a checkerboard grating,
which can simultaneously split light in both the x and y directions. Using the
X-OR technique, any two binary DOEs can be combined. Before applying
this technique to other DOEs, we present the procedure for combining DOEs
using the modulo-2p phase addition method and the advantages of this
method.
Figure 6.1 Phase addition of two orthogonal 1D gratings in tandem each with an error of d
in their phase heights (reprinted from Ref. 16).
Figure 6.2 Phase profiles of the fundamental building blocks of a 2D grating (a) after
design and (b) after fabrication (reprinted from Ref. 16).
grating after design and after fabrication are shown in Figs. 6.2(a) and (b),
respectively.
Let us now compare the fundamental building blocks for the two cases in
Figs. 6.1 and 6.2. In the first case, it can be observed that the design procedure
of modulo-2p phase addition indirectly masks 25% of the area of the
fundamental building block from any phase error.16,17 In the case of a 2D
grating, from the Fourier coefficients calculation shown in Chapter 2, we have
seen that the maximum efficiency possible in the (1, 1) diffraction order is
16%. The relative intensity is the ratio of the intensity of light in a particular
diffraction order to the intensity of the incident light, and in this case it can be
given by (I1/Ii). The relative intensity of the (1, 1) diffraction order is
plotted as a function of phase error in Fig. 6.3 for the above two cases.
A third case is considered that involves two 1D gratings with orthogonal
periodicity in tandem, but it is assumed that the phase error is 0 in only one of
the gratings. The other useful advantage of having a single DOE generated via
the modulo-2p phase addition method over having two independent DOEs is
now clearly visible. The variation in the relative intensity in the (1, 1)
diffraction order is relatively slower for the multifunctional DOE compared to
two independent DOEs. It is seen from Fig. 6.3 that the 2D grating generated
by modulo-2p phase addition method behaves like two 1D gratings with
phase error in only one of the gratings. This is a consequence of masking off
regions by the design used in the modulo-2p phase addition method.
In the above analysis, the performance of two binary DOEs in tandem
was compared to a single binary DOE generated by the modulo-2p phase
addition of the phase profiles of the two binary DOEs. Although the
Figure 6.3 Plot of relative intensity in the (1, 1) diffraction order for different phase errors
in the gratings for three cases: case 1 is two orthogonal binary gratings in tandem with phase
error in both gratings; case 2 is a 2D checkerboard binary grating; and case 3 is two orthogonal
binary gratings in tandem with phase error in one of the gratings (reprinted from Ref. 16).
comparison was made with binary DOEs, the modulo-2p technique can be
utilized to combine the functions of two or more binary DOEs, one binary
DOE with a blazed DOE, two blazed DOEs, as well as phase and amplitude
DOEs. In the following sections, the techniques and MATLAB® codes for
combining the aforementioned types of DOEs are discussed.
Table 6.2 MATLAB code for design of a multifunctional DOE from a circular grating and a
1D grating.
%Multifunctional DOE – CG and 1D grating
clear; %Clear all memory
% Defining Grating Parameters
N=500; %Define Matrix sizes
A1=zeros(N,N); %Define Matrices by assigning 1 to all pixels
A2=zeros(N,N);
P1=50;%Define the period of the binary axicon grating
P2=25;%Define the period of the 1D grating
FFr=0.5;%Define fill factor for radial periodicity
FFy=0.5;%Define fill factor for periodicity along y direction
% Constructing the grating
x=1:N;
y=1:N;
[X,Y]=meshgrid(x,y);
r=sqrt((X-N/2).*(X-N/2)þ(Y-N/2).*(Y-N/2));
A1(rem(r,P1),P1*FFr)=1;
A2(rem(Y,P2),P2*FFy)=1;
A3=exp(1i*pi*xor(A1,A2));
A3(r.150)=0;
%Observing the grating output in the far-field
E=fftshift(fft2(A3));
I=(abs(E)/(N*N)).*(abs(E)/(N*N));
Figure 6.4 (a) Image of the phase profile of the multifunctional DOE designed by modulo-
2p phase addition of a 1D grating (period ¼ 25 mm) with a circular grating (period ¼ 50 mm).
(b) Far-field diffraction pattern of the multifunctional DOE designed by modulo-2p phase
addition of a 1D grating with a circular grating.
circumference of the ring for FCG ¼ p. In other words, the combined DOE
has an effective efficiency of only 20%.
The ring patterns having the same radii and with varying intensity values
correspond to the different diffraction orders of the 1D grating. The larger
ring patterns [seen as faint circles in Fig. 6.4(b)] surrounding the first ring
patterns are the higher orders of the circular grating. Higher orders are not
visible when using ‘colormap (gray)’ due to their poor intensity. However, it
is possible to view even the lower-intensity profiles by choosing a different
colormap.
In the above case, only two DOEs were combined. However, the same
procedure can be extended to combine more DOEs. For instance, one could
combine two orthogonal 1D gratings and a circular grating. This is same as
combining a 2D grating (checkerboard) with a circular grating. The
MATLAB code can be modified to first generate the binary phase profile
of a 2D grating from two binary 1D gratings, followed by the generation of
the phase profile of the binary multifunctional DOE from the phase profiles of
the binary 2D grating and the circular grating. In this combination, the
multifunctional DOE will generate four ring patterns along both the x and y
directions, each with an intensity of 16% 50% of the input light and ring
patterns with less intensity corresponding to the higher diffraction orders of
the 2D grating. Hence, the effective efficiency in the (1, 1) order of the 2D
grating and 1st order (þ1 and –1 combined) of the circular grating is only
8%. The circular grating has a period L1 ¼ 50 mm, and the binary 1D
gratings have periodicity L2x ¼ 25 mm and L2y ¼ 25 mm in the x and y
directions, respectively. An image of the phase profile of the 2D phase-
modulated circular grating and its far-field diffraction pattern are shown in
Figs. 6.5(a) and (b), respectively. The four ring patterns correspond to the
Figure 6.5 Image of the phase profile of the multifunctional DOE designed by modulo-2p
phase addition of two orthogonal 1D gratings (period ¼ 25 mm) and a circular grating
(period ¼ 50 mm). (b) Far-field diffraction pattern of the multifunctional DOE designed by
modulo-2p phase addition of two orthogonal 1D gratings with a circular grating.
(1, 1) diffraction orders of the 2D grating and the þ1 and –1 combined
order of the circular grating.
The radius of the ring pattern and the spacing between the ring patterns
can be determined in both of the above cases using the grating equation and
trigonometry, as discussed in Chapter 2. The radius of the ring pattern is
given by
1 l
r ¼ z tan sin . (6.9)
L1
Equation (6.9) can be used to determine the spacing between the spots by
replacing L1 with L2. The radius of the ring pattern and the spacing between
the ring patterns at a distance of 50 mm is found to be 632 mm and 2.5 mm,
respectively.
The above procedure can also be extended to combine two binary DOEs
such as an FZP and different gratings, which can be used to generate 2 2,
3 3 arrays of focused spots, etc., using the modulo-2p phase addition
method. Array generation is required in applications that demand identical
spots or patterns repeated along the x and y directions. One such application
is micro-drilling using an axicon array.18 A well-known element used for array
generation is the Dammann grating.19,20
Using the above technique, it is also possible to tailor the intensity of
different spots. Arrays of focused spots and ring patterns can also be easily
generated by using an array of identical FZPs and circular gratings,
respectively. However, there are many advantages to using a multifunctional
DOE to generate the array compared to using an array of identical DOEs.
First, in the case of DOEs containing an array of elements, different sections
of the input beam are incident upon different elements, which may result
Table 6.3 MATLAB code for design of a multifunctional DOE from a blazed FZP and a 1D
grating.
%Multifunctional DOE – Blazed FZP and 1D grating
clear;%Clear all memory
%Define grating and FZP parameters
N=500;%Define Matrix size
f=10000;
lambda=0.632;
P=200;
FFy=0.5;
A1=zeros(N,N);%Define the matrices assigning ones to all pixels
A2=zeros(N,N);
A3=zeros(N,N);
%Grating and FZP construction
x=1:N;
y=1:N;
[X,Y]=meshgrid(x,y);
r=sqrt((X-N/2).*(X-N/2)þ(Y-N/2).*(Y-N/2));
A1=(f-sqrt(f*fþr.*r))*(2*pi)/(0.632);
A2(rem(Y,P),P*FFy)=pi;
A=rem(A1þA2,2*pi);
B1=exp(1i*A1);
B2=exp(1i*A2);
B3=exp(1i*A);
Figure 6.6 Plot of phase profiles of (a) a blazed FZP, (b) a binary 1D grating, (c) a
multifunctional DOE after normal addition of phase profiles of FZP and 1D grating, and
(d) binary 1D grating and multifunctional DOE after modulo-2p addition of phase profiles of
FZP and binary 1D grating at y ¼ 250 (adapted from Ref. 21).
Figure 6.7 (a) Image of the phase profile of a multifunctional DOE generated by combining
the phase functions of a blazed circular grating and a 2D binary checkerboard grating.
(b) Far-field diffraction pattern of the same multifunctional DOE.
Table 6.4 MATLAB code for design of a multifunctional DOE from a blazed circular grating
and a blazed 1D grating.
%Multifunctional DOE – Blazed circular grating and blazed 1D
%grating clear
% Clear all memory
%Define grating parameters
N=500;%Define Matrix size
Pr=25;
Px=10;
%Blazed axicon and grating construction
x=1:N;%Sampling
y=1:N;
[X,Y]=meshgrid(x,y);
r=sqrt((X-N/2).*(X-N/2)þ(Y-N/2).*(Y-N/2));
P1=rem(r,Pr); %Construction of blazed axicon
A1=(P1/Pr)*2*pi;
P2=rem(X,Px);%Construction of blazed 1D grating
A2=(P2/Px)*2*pi;
A=rem(A1þA2,2*pi); %Modulo-2pi phase addition of the two phase
%profiles
B=exp(1i*A);
B(r.150)=0;
Figure 6.8 Phase profiles of (a) a blazed circular grating, (b) a blazed 1D grating, and (c) a
multifunctional DOE.
Figure 6.9 (a) Image of the phase profile of the multifunctional blazed DOE containing the
functions of a blazed circular grating and a blazed 1D grating. (b) Far-field diffraction pattern
of the same multifunctional blazed DOE.
the functions of an amplitude FZP with a blazed spiral phase plate. The
MATLAB code for generation of the phase profile of this multifunctional
DOE is shown in Table 6.5. The image of the phase profile of the
multifunctional DOE is shown in Fig. 6.10.
Table 6.5 MATLAB code for design of a multifunctional DOE from a blazed spiral phase
plate and an amplitude FZP.
%Multifunctional DOE – Blazed circular grating and blazed 1D
%grating clear;
%Clear all memory
%Define grating parameters
N=500; %Define Matrix size
M=32; %Define number of zones
A1=zeros(N,N); %Define Matrices by assigning 0 or 1 to all
%elements
r1=zeros(M,M);
r=zeros(N,N);
f=3000; %Define focal length and wavelength (in micrometers)
lambda=0.633;
L=1;%Define topological charge
% Construction of the spiral FZP
for n=1:M; %Calculate the widths of the zones
r1(n)=sqrt(n*f*lambda);
end
for n=1:2:M; %Scan element by element using two for loops
for p=1:N;
for q=1:N;
r(p,q)=sqrt((p-N/2)*(p-N/2)þ(q-N/2)*(q-N/2));
if r(p,q) . r1(n) && r(p,q) , r1(nþ1);
A1(p,q)=exp(1i*L*(atan2((q-N/2),(p-N/2))));
end
end
end
end
Figure 6.10 Image of the phase profile of a multifunctional DOE containing the functions of
an amplitude FZP and a blazed spiral phase plate.
at a particular plane, when the wavelength varies, the spacing between the
diffraction spots varies linearly with wavelength. According to Eq. (6.12),
assuming that r2n ≫ n2 l2 , the focal length is inversely proportional to the
variation in wavelength. These equations, therefore, tell us that d and f change
in opposite ways with respect to the wavelength. Now let us combine
Eqs. (6.11) and (6.12) by substituting Eq. (6.12) in Eq. (6.11), giving rise to
l r2n n2 l2
dðlÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi . (6.13)
L l2
2 2nl
Equation (6.14) does not depend on the wavelength. In other words, the
spacing between the spots is independent of the wavelength. How does this
independence come about? When the wavelength of light increases, the
spacing between the diffraction spots increases at the plane that is a focal
length distance away. On the other hand, when the wavelength of light
increases, the focal length of the multifunctional DOE decreases. This results
in the diffraction spots being focused at a plane that is closer to the DOE
plane when compared to the situation before the wavelength change. Hence,
these two effects cancel each other out, rendering a wavelength-independent
spot spacing. There is an alternate way to think about this effect. In the case of
a diffraction grating, when wavelength increases, the spacing between the
spots increases, and vice versa, at a particular plane of interest. However,
when the diffraction grating is used together with a FZP, the location of the
plane of interest is controlled by the FZP and varies to compensate for the
change in the spot spacing. This effect, when regarded from the application
point of view, proves quite useful. If the 1D grating is replaced by a binary
axicon in the multifunctional DOE, a focused ring pattern will be generated at
the focal plane of FZP. The ring pattern with a static ring diameter can be
moved in space just by varying the wavelength of light. Although the ring
diameter stays constant, the focal plane position changes; therefore, such
combined elements can be said to be quasi-achromatic.
The above effect is valid only when the features of the grating and the
FZP are much larger than the wavelength of light used, or in other words,
when both of the DOEs are scalar (refer to Chapter 1). It is also useful to note
that the above effect is true for only the 1st diffraction order.
What happens for the higher diffraction orders? Modifying Eq. (6.11) to
represent the 2nd and 3rd orders yields Eqs. (6.15) and (6.16), respectively:
2lf ðlÞ
dðlÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi , (6.15)
L2 4l2
3lf ðlÞ
dðlÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi , (6.16)
L2 9l2
Figure 6.11 Plot of the focal length of the FZP with variation in the wavelength of light
(Reprinted from Ref. 29).
Figure 6.12 Plot of the spacing between the 0th and 1st diffraction order for the
multifunctional DOE (black line) and the 1D binary grating (gray line) with variation in
wavelength of light (reprinted from Ref. 29).
6.1.4 Conclusion
In Section 6.1, different types of multifunctional DOEs generated by the
modulo-2p phase addition are presented, and different diffractive features of
the combined DOEs are analyzed. The efficiency of binary multifunctional
DOEs is found to be less sensitive to phase errors. In the last section, the
wavelength-independent property arising due to combining two DOEs is
discussed. These interesting features can be used to our advantage in the design
and fabrication of DOEs for different applications. One important point should
not be overlooked when dealing with multifunctional DOEs: In the generation
of the phase profile of the multifunctional DOE, modulo-2p phase addition is
carried out on the phase profiles of the two or more independent DOEs. This
can give rise to subwavelength features, as seen in Figs. 6.4, 6.6, and 6.9. Such
features are quite difficult to fabricate, and diffraction by such features cannot
be explained by scalar diffraction formulation.
Figure 6.13 Path difference profile of a negative axicon (reprinted from Ref. 16).
Figure 6.14 Phase profile of the multifunctional DOE containing the functions of a negative
axicon and a FZP.
Figure 6.15 Plot of the radius of zones of the multifunctional DOE with an FZP and a
negative axicon (solid line) and the FZP alone (dashed line) (reprinted from Ref. 16).
noted that the presence of the DOE increases the zone width within the
diameter of the device. A closer look will reveal that this facility of larger
zone width at the outermost region has been created at the expense of the
f-number of the DOE. Therefore, after addition of the phase, if the device is
fabricated with the previous radius R, then the f-number of the device is less
than that of the FZP alone. Secondly, the efficiency of the DOE is preserved
as in the case of the aberration correction technique. The effective efficiency
of the binary multifunctional DOE is 40% (32% if it is through the modulo-
2p phase addition technique). When light is incident on the multifunctional
DOE, a focused ring pattern is formed at the focal plane of the FZP
function. The width of the ring pattern is given by 1.65 the diffraction-
limited spot size.27,28
The same procedure can be used to combine any refractive phase with any
DOE to generate a multifunctional DOE.
6.3 Conclusions
In this chapter, the art of designing a multifunctional DOE by combining
one or more DOE functions is presented in detail. Two schemes, namely,
the modulo-2p phase addition method and the analog method, are
presented to combine two or more DOEs. The two main disadvantages
associated with the modulo-2p phase addition method are the poor
efficiency and the fine features that are generated during design. The fine
features decrease the focal depth of the device, and if they are
subwavelength, they introduce polarization effects as well. The analog
method seems to have higher efficiency and results in relatively easier
fabrication. This is because the feature sizes increase within the radius of
the element after inclusion of the refractive phase, although at the expense
of the f-number of the DOE. This increases the focal depth of the device
with an increase in the ring width. The width of the ring pattern is 1.65
the diffraction-limited spot size.27,28
Table 6.6 provides a comparison of the characteristics of multifunctional
DOE created by three different techniques: (1) a circular grating (or axicon)
and a FZP in tandem, (2) a circular grating and a FZP combined by the
modulo-2p phase addition method, and (3) a circular grating and an FZP
combined by the analog method.
6.4 Exercises
E.6.1 Design a multifunctional DOE to generate a focused ring pattern with a
diameter of 1 mm at a distance of 30 mm from the DOE using both the
modulo-2p phase addition method29 and the analog method. Extract some
useful parameters. Assume that the fabrication system is capable of
fabricating DOEs with a maximum size of 2 mm 2 mm.
20 mm, and propagating along the same direction from the DOE plane as
shown in the optical configuration given in the figure below.
E.6.3 Design a helical axicon containing the functions of a spiral phase plate
with charge L ¼ 10 and an amplitude binary axicon with a period L ¼ 10 mm
using the modulo-2p phase addition method. Calculate the far-field
diffraction pattern with an aperture having a 100-mm diameter.
References
1. J. W. Goodman, Introduction to Fourier Optics, Second Edition,
McGraw-Hill Companies Inc., New York (1996).
2. B. C. Kress and P. Meyrueis, Applied Digital Optics, John Wiley & Sons,
Chichester, UK (2009).
3. C. Zhang, N. R. Quick, and A. Kar, “Diffractive optical elements for
pitchfork beam shaping,” Opt. Eng. 48, 078001 (2009) [doi: 10.1117/
1.3183919].
4. M. Honkanen and J. Turunen, “Tandem systems for efficient generation
of uniform-axial-intensity Bessel fields,” Opt. Commun. 154, 368–375
(1998).
5. J. M. Herrera-Fernandez, L. M. Sanchez-Brea, and E. Bernabeu, “Near-
field shaping with two binary diffractive optical elements in tandem,” Opt.
Commun. 297, 182–189 (2012).
6. S. Tibuleac and R. Magnusson, “Narrow-linewidth bandpass filters with
diffractive thin-film layers,” Opt. Lett. 26, 584–586 (2001).
7. T. Stone and N. George, “Hybrid diffractive-refractive lenses and
achromats,” Appl. Opt. 27, 2960–2971 (1988).
8. N. Davidson, A. A. Friesem, and E. Hasman, “Analytic design of hybrid
diffractive–refractive achromats,” Appl. Opt. 32, 4770–4774 (1993).
25. B. Zhang and D. Zhao, “Focusing properties of Fresnel zone plates with
spiral phase,” Opt. Express 18, 12818–12823 (2010).
26. A. Vijayakumar and S. Bhattacharya, “Design, fabrication, and evalua-
tion of a multilevel spiral-phase Fresnel zone plate for optical trapping,”
Appl. Opt. 51, 6038–6044 (2012).
27. P.-A. Bélanger and M. Rioux, “Ring pattern of a lens–axicon doublet
illuminated by a Gaussian beam,” Appl. Opt. 17, 1080–1088 (1978).
28. A. Vijayakumar, “Design, Fabrication, and Evaluation of Diffractive
Optical Elements for the Generation of Focused Ring Patterns,” Doctoral
Thesis, Indian Institute of Technology Madras, Chennai, India (2015)
[doi: 10.13140/RG.2.2.25996.51847].
29. A. Vijayakumar and S. Bhattacharya, “Quasi-achromatic Fresnel zone
lens with ring focus,” Appl. Opt. 53, 1970–1974 (2014).
30. A. Vijayakumar and S. Bhattacharya, “Design, fabrication, and evalua-
tion of diffractive optical elements for generation of focused ring
patterns,” Proc. SPIE 9449, 944902 (2015) [doi: 10.1117/12.2077407].
31. J. J. Zambuto, R. E. Gerber, J. K. Erwin, and M. Mansuripur, “Ring-lens
focusing and push-pull tracking scheme for optical disk systems,” Appl.
Opt. 33, 7987–7994 (1994).
145
neutral density filters (NDFs). CGHs can overcome the difficulties faced
when physically recording a hologram.8–11
The basic principle of the generation of any HOE involves the
superposition of two optical waves A1exp(jF1) and A2exp(jF2), and the
resulting interference pattern modulated as 2A1A2 cos(F1 – F2) is a HOE.1,8,9
In the case of CGHs, this interference is simulated, and the resulting fringe
pattern is processed and fabricated using lithography techniques or is printed
out on transparent sheets. The basic optics configuration for recording and
reconstructing a HOE is shown in Figs. 7.1 and 7.2, respectively.
The intensity of the interference pattern created by the superposition of
the object wave A1exp(iF1) and the reference wave A2exp(iF2) is given by
Figure 7.1 Optics configuration for recording the interference pattern (hologram) formed by
the superposition of an object wave with a reference wave.
Figure 7.2 Optics configuration for reconstruction of the object wave by illuminating the
hologram with the reference wave.
I ðx,yÞ ¼ jA1 ðx,yÞj2 þ jA2 ðx,yÞj2 þ 2jA1 ðx,yÞjjA2 ðx,yÞj cos½F1 ðx,yÞ F2 ðx,yÞ.
(7.1)
x
C2 ðx,yÞ ¼ exp j tanðuÞ2p . (7.3)
l
As both waves are considered at a specific plane, the z dependence need not
appear in the equation. The plane wave propagates at an angle of u with the
z direction, so the tilted-plane wavefront makes an angle of u with the
x direction. Hence, the phase of the wave varies linearly along the x direction.
In the discussion presented in Chapter 2 on the refractive equivalents of DOEs,
it was shown that the refractive prism is equivalent to a diffractive grating.
A similar connection may be found in CGHs as well. To generate a 1D grating,
the phase profile variation along the x direction is the phase of a refractive
prism. The optics configuration is shown in Fig. 7.4, and the MATLAB code is
given in Table 7.1.
It can be noted from Eq. (7.3) that for larger values of u, it may be difficult
to view the fringe patterns, as the width of the fringe pattern can be less than
the smallest currently available pixel size of approximately 1 mm. Hence, in
order to expand the fringe pattern, the pixel size is decreased by several
factors. From Eq. (1.7) we find that the period of the grating corresponding to
Figure 7.4 Optics configuration for superposition of two plane waves propagating in
different directions.
Table 7.1 MATLAB code for the design of a 1D grating using the computer-generated
holographic method.
%%1D grating
%Define parameters
N=500;%%Define size of the matrix
Angle=1;%Define angle of the second plane wave
V=0.5;%%Visibility controller
lambda=0.632*1e-6;%Define wavelength
%Create sampled space
del=1*1e-6;%sampling
x=-N/2:N/2-1;
y=-N/2:N/2-1;
[X,Y]=meshgrid(x*del,y*del);
%Simulate interference
A=V*exp(1i*(Y/lambda)*tand(Angle)*2*pi);
B=V*exp(1i*2*pi);
D=A+B;%Interference of the object and reference wave
I=abs(D).*abs(D);
%%Construct grating
I1=im2bw(I);%Binarize the matrix
Grating=exp(1i*pi*I1);%Generate the phase grating
a diffraction angle u ¼ 1 deg for the 1st diffraction order is L ¼ 36.2 mm. The
period of the sinusoidal grating obtained from the MATLAB program is
36 pixels, which matches the calculated value.
The period of the grating is 3.6 mm if the angle between the two plane
waves is 10 deg. However, it is not possible to view 3.6 mm accurately with a
mesh of pixel size 1 mm. Hence, the sampling period must be decreased 10,
resulting in a magnification of the fringe pattern. The nonbinarized
hologram (fringe pattern) generated due to the interference between two
plane waves, with an angle of 10 deg between them, and a cross section of its
intensity profile are shown in Figs. 7.5(a) and (b), respectively. The period of
the grating generated by MATLAB is 36 pixels. This has to be multiplied by
the sampling period, 0.1 mm in this case, to obtain 3.6 mm. The orientation
of the grating can be modified by changing the object waves’ propagation
direction.
Figure 7.5 (a) Image and (b) intensity profile of the 1D sinusoidal grating.
Figure 7.6 Optics configuration for superposition of a plane wave with a spherical wave.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2p
C2 ðx,yÞ ¼ exp j f 2 þ ðx2 þ y2 Þ . (7.5)
l
The optics configuration is shown in Fig. 7.6, and the MATLAB code is given
in Table 7.2.
The binarized hologram (fringe pattern) generated due to the interference
between a plane wave and a spherical wave, and a plot of its binary phase
profile are shown in Figs. 7.7(a) and (b), respectively. The radius of the first
zone is measured to be 154 mm, which matches the value calculated using the
formula given in Eq. (2.14). The same procedure can be used to design an FZP
for off-axis focusing light.
Table 7.2 MATLAB code for design of a FZP using the computer-generated holographic
method.
%%Fresnel zone plate
%Define parameters
N=500;%%Define size of the matrix
Angle=1;%Define angle of the second plane wave
V=0.5;%%Visibility controller
lambda=0.632*1e-6;%Define wavelength
%Create sampled space
del=1*1e-6;%sampling
x=-N/2:N/2-1;
y=-N/2:N/2-1;
[X,Y]=meshgrid(x*del,y*del);
f=0.01; %Focal length of 1 cm
%Simulate interference
A=V*exp(1i*((2*pi)/(lambda))*sqrt(X.^2+Y.^2+f*f));
B=V*exp(1i*2*pi);
D=A+B;%Interference of the object and reference wave
I=abs(D).*abs(D);
figure (2)
colormap gray
imagesc(IN)
%%Construct FZP
I1=im2bw(I);%Binarize the matrix
Grating=exp(1i*pi*I1);%Generate the FZP
Figure 7.7 (a) Binarized fringe pattern and (b) plot of the phase profile of the FZP.
Figure 7.8 Optics configuration for superposition of a plane wave with a conical wave.
Table 7.3 MATLAB code for design of a diffractive axicon using the computer-generated
holographic method.
%Diffractive axicon
%Simulate interference
A=V* exp(1i*(sqrt(X.^2+Y.^2)/lambda)*tand(Angle)*2*pi);
B=V*exp(1i*2*pi);
spaced fringe pattern in this direction. The reference wave is identical to that
given in Eq. (7.4), and the equation for the conical wave at the interference
plane is given by
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2p
C2 ðx,yÞ ¼ exp j tanðuÞ x2 þ y2 . (7.6)
l
The optics configuration for superposition of a conical wave with a plane
wave is shown in Fig. 7.8. The MATLAB code is similar to that given in
Table 7.1 for a 1D grating, except that the object wave must be generated with
radial instead of linear (x or y) dependence. The code for design of the object
wave is shown in Table 7.3. In this case, it is necessary to increase the
sampling period to view the fringe pattern with larger diffraction angles,
similar to the procedure shown in the case of a 1D grating.
The image of the generated sinusoidal axicon is shown in Fig. 7.9. The
period of the measured grating is 36 mm, which matches the calculation result
for a diffraction angle of 1 deg.
The procedure presented here can be used to design any CGH, irrespective
of its phase profile, and the design section of the MATLAB code can be
modified to generate the fringe pattern.
Figure 7.9 Image of the sinusoidal axicon generated by the superposition of a plane wave
with a conical wave of having a diffraction angle of 1 deg (reprinted from Ref. 33).
Figure 7.10 Hologram recording and reconstruction using a reference wave and a plane
wave.
2p
C1 ðx,yÞ ¼ exp j tanðuÞx , (7.7)
l
2p
C2 ðx,yÞ ¼ exp j tanðuÞx . (7.8)
l
The MATLAB code given in Table 7.1 can be modified by replacing the
equation of the reference wave with that given in Eq. (7.8). The resulting
fringe pattern has a period of 18 mm, which will be the same when a plane
wave tilted 2 deg is superposed with a plane wave with no tilt.
Figure 7.11 Images of the off-axis binary axicons with tilt angles of (a) 1 deg and (b) 2 deg,
and (c), (d), their respective far-field diffraction patterns (reprinted from Ref. 33).
The same procedure can be used to generate the phase profile of off-axis
axicons with different orientations. The same procedure can also be used
for generation of tilted beams with other types of CGHs such as an
off-axis FZP. The twin images occur because the phase profile is binary
(or sinusoidal) and can be reduced to one image by modifying the profile to a
blazed one.
Table 7.4 MATLAB code for the design of a binary forked grating using the computer
generated-holographic method.
%Forked grating
L=1;
A=V*exp(1i*L*(atan2(Y,X)));
B=V*exp(1i*(2*pi/lambda)*tand(Angle)*Y);
Figure 7.12 Images of the binary forked gratings with tilt angle u ¼ 1 deg and topological
charge (a) L ¼ 1 and (b) L ¼ 3, and (c), (d) their respective far-field diffraction patterns.
Figure 7.13 Images of the phase profile of the helical axicon with topological charges
(a) L ¼ 3 and (b) L ¼ 5, and (c), (d) their respective far-field diffraction patterns.
where
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Df
f ðx,yÞ ¼ f 0 þ x2 þ y2 . (7.11)
R
The equation of a tilted plane wave is given in Eq. (7.7). The MATLAB code
for design of the axilens is given in Table 7.5. The element is designed with
f0 ¼ 10 mm and Df ¼ 3 mm, and the tilt angle u ¼ 1 deg.
The binary phase profile of the generated CGH is shown in Fig. 7.14(a), and
its diffraction pattern is shown in Fig. 7.14(b). The diffraction pattern shows twin
Figure 7.14 (a) Phase profile of an off-axis binary axilens and (b) the far-field diffraction
pattern of the axilens.
Table 7.5 MATLAB code for design of an off-axis axilens using the holographic method.
%Off axis axilens
%Define parameters
N=500;%%Define size of the matrix
Angle=1;%Define angle of the second plane wave
V=0.5;%%Visibility controller
lambda=0.632*1e-6;%Define wavelength
L=3;
f0=0.01;
delf=0.003;
%Create sampled space
del=1*1e-6;%sampling
x=-N/2:N/2-1;
y=-N/2:N/2-1;
[X,Y]=meshgrid(x*del,y*del);
f=f0+delf*sqrt(X.^2+Y.^2);
%Simulate interference
A=V*exp(1i*((2*pi)/lambda)*(f-sqrt(f.*f-(X.^2+Y.^2))));
B=V*exp(1i*(2*pi/lambda)*tand(Angle)*Y);
D=A+B;%Interference of the object and reference wave
IN=abs(D).*abs(D);
(Continued )
beams, as the phase profile of the CGH is either binary or sinusoidal. Otherwise,
the diffraction pattern is similar to that of an off-axis axicon, as expected.
Figure 7.15 Optics configuration for focusing light using a refractive axicon (reprinted from
Ref. 33).
Figure 7.16 Optics configuration for generation of a tilted Bessel-like beam using a
refractive axicon and a thin prism (BPSE) in tandem (reprinted from Ref. 33).
Figure 7.17 Optics configuration for generation of an accelerating Airy beam using a
refractive axicon and a thin prism (BPSE) in tandem (reprinted from Ref. 33).
the angle of the axicon is selected as 0.1 deg, and the thickness profile of the
BPSE is selected as
t2 ðx,yÞ ¼ 10x0.5 þ 10y0.5 . (7.17)
The MATLAB code is similar to that given in Tables 7.1 and 7.2. The design
section of the MATLAB code is shown in Table 7.6. The phase profile of the
CGH is shown in Fig. 7.18(a), and the far-field diffraction pattern is shown in
Fig. 7.18(b).
The two first diffraction orders result due to the binary or sinusoidal phase
profile of the CGH. The above procedure can be repeated to generate
accelerating Airy beams with interesting path profiles. As with any diffractive
element, the beam very close to the element is not easily accessible and hence
is not useful for many applications. Additionally, the intensity at each plane is
less because the light intensity is spread over the entire focal depth. These
issues can be resolved by replacing the axicon with an axilens.
Table 7.6 MATLAB code for design of a CGH for generating accelerating beams.
%Accelerating beams
%Define parameters
N=500;%%Define size of the matrix
Angle=0.1;%Define angle of the second plane wave
V=0.5;%%Visibility controller
lambda=0.632*1e-6;%Define wavelength
%Create sampled space
del=1*1e-6;%sampling
x=-N/2:N/2-1;
y=-N/2:N/2-1;
(Continued )
Figure 7.18 (a) Image of the binary phase profile of the CGH generated by the interference
of a conical wave and a BPSE and (b) its far-field diffraction pattern (reprinted from Ref. 33).
x
C1 ðx,yÞ ¼ exp i tanðuÞ2p , (7.18)
l
x
C2 ðx,yÞ ¼ exp i tanðuÞ2p , (7.19)
l
y
C3 ðx,yÞ ¼ exp i tanðuÞ2p , (7.20)
l
y
C4 ðx,yÞ ¼ exp i tanðuÞ2p . (7.21)
l
For normalizing the matrices, a visibility factor of 0.25 is selected for each
optical wave. The MATLAB code for generation of an interference pattern
with 2D periodicity by the superposition of the above four optical wavefronts
is given in Table 7.7. The image of the interference pattern is shown in
Fig. 7.19(a). The procedure can be repeated by replacing the tilted-plane
wavefronts by other wavefronts with interesting profiles.38–41 The change in
the interference pattern when the orientation of two of the tilted plane waves
is varied is shown in Fig. 7.19(b). These interference patterns can have
subwavelength features and, hence, are useful as antireflection coatings and in
photonic crystal waveguide fabrication.42
Figure 7.19 (a) Interference pattern obtained by the superposition of four tilted plane
waves described by Eqs. (7.18)–(7.21). (b) Interference pattern obtained by the
superposition of four tilted plane waves with different orientations.
The result of the integration yields four terms. The first term corresponds
to the Fourier transform of a constant. The second term is the
autocorrelation function of the object wave with itself. The third and fourth
terms are of interest and correspond to the interference between the object
and reference waves from the point source.1 The steps for simulating the
recording and reconstruction of a Fourier hologram are presented in
Fig. 7.21.
Figure 7.21 Flow chart for simulating the recording and reconstruction of a Fourier
hologram.
Table 7.8 MATLAB code for the design of computer-generated Fourier holograms.
%Fourier Hologram
% Load the object and create the object and reference matrices
N=500;%Define the size of the matrix
A=imread(‘F:\IITM.jpg’,’jpg’);%Loading the image
A=double(A);
A=A(1:N,1:N);
A=imresize(A,[N,N]);
B=zeros(N,N);%Dirac delta function
B(250,250)=100;
%Calculate the far field diffraction patterns of the object and
%reference
A1=fftshift(fft2(A));%Calculate the diffraction pattern for
the object
I1=abs(A1).*abs(A1);
B1=fftshift(fft2(B));%Calculate the diffraction pattern for
the reference
I2=abs(B1).*abs(B1);
%Create the hologram by interfering the far field diffracted
%fields of object and reference
D1=A1+B1;
I3=abs(D1).*abs(D1);
%Filtering
I4=(I3-I1);%Filtering of autocorrelation term
%Hologram reconstruction
D2=fftshift(fft2(I4));
I5=abs(D2).*abs(D2);
The reconstructed image shows two images of the object, one of which is
inverted, as expected. The image of the object can be easily created in
Microsoft PowerPoint or Paint. The MATLAB code for design of Fourier
holograms and their reconstruction is given in Table 7.8. The image of the
object is shown in Fig. 7.22(a). The image of the hologram is shown in
Fig. 7.22(b). The reconstructed images without and with filtering are shown in
Figs. 7.23(a) and (b), respectively.
The hologram can be used for the study of low- and high-pass filtering
computational experiments. In the hologram, the high spatial frequencies
of the object are at the outer parts due to high diffraction angles, and the
low spatial frequencies of the object are at the center. Hence, by zero
padding either the outer or inner regions of the hologram before
reconstruction, the areas of different frequency content can be identified.
The center part of the hologram (100 pixels 100 pixels) is zero padded in
the first case, and the outer parts with length of 100 pixels are zero padded
in the second case. The results of reconstruction in each case are shown in
Figs. 7.24(a) and (b).
For the high-pass-filtered case, it can be seen that only the edges of the
letters are visible, indicating that light is diffracted at higher angles at the
edges and at lower angles in the center. In the case of low-pass filtering, the
Figure 7.22 (a) Image of the object used for the generation of the computer-generated
Fourier hologram and (b) image of the Fourier hologram obtained by the interference of the
object and reference waves.
Figure 7.23 Reconstruction of the hologram (a) without filtering and (b) with filtering of the
autocorrelation factor.
Figure 7.24 Reconstructed images from the hologram after (a) high-pass and (b) low-pass
filtering.
ð þ`
þ` ð
ejkz ej 2zðx þy Þ
k 2 2
k 2
Uðx,yÞ ¼ T o ðj,hÞ exp j ðx þ y Þ
2
jlz 2z
` ` (7.25)
2p
exp j ðxj þ yhÞ djdh.
lz
N ¼ lz. (7.27)
Equation 7.28 shows that the distance z can be controlled by varying the
matrix size N. This also solves the problem of scaling, as Eqs. (7.25) and (7.26)
are equal. From Eq. (7.25), it can be understood that the Fresnel diffraction
calculation involves three steps. The object matrix must be multiplied by the
parabolic phase factor in the first step, followed by the Fourier transform
operation, and once again multiplication by the parabolic phase factor. The
reference wave is a plane wave; hence, a matrix with constant values can be
used to represent it. The hologram can be generated by adding the two
matrices and performing the filtering operations discussed in Section 7.3. The
computer reconstruction has to be done in the same sequence to obtain the
reconstructed image. The MATLAB code for the design of a Fresnel
hologram is given in Table 7.9. The image of the object is shown in Fig. 7.25.
Table 7.9 MATLAB code for the design of computer-generated Fresnel holograms.
%Fresnel Hologram
%Load object and create the object and reference matrices
N=500;% Define the matrix size
B=ones(N,N);%Reference wave
A=imread(‘C:\Vijayakumar\mesh.jpg’,’jpg’);%Loading the
%image
A=double(A);%Convert symbolic object to numeric object
A=A(1:N,1:N);
A=A/max(max(A));%Normalizing the matrix
%Sampling
del=1;%
x=-N/2:N/2-1;
y=-N/2:N/2-1;
[X,Y]=meshgrid(x*del,y*del);
%Calculate the Fresnel diffraction pattern of the object
A1=A.*exp(-1i*(pi/N)*(X.^2+Y.^2));
A2=fftshift(fft2(fftshift(A1)));
A3=A2.*exp(1i*(pi/N)*(X.^2+Y.^2));
%Create interference between the object and reference waves
H1=B+A3;%Interference pattern
%Filtering
I1=abs(A3).*abs(A3);
I2=abs(H1).*abs(H1);
H2=I2-I1;%Filtering
H2=H2/max(max(H2));%Normalizing the hologram
%Reconstruct the hologram
H3=H2.*exp(-1i*(pi/N)*(X.^2+Y.^2));
H4=fftshift(fft2(fftshift(H3)));
H5=H4.*exp(1i*(pi/N)*(X.^2+Y.^2));
H6=(abs(H5).*abs(H5));
H7=rot90(H6,2);
colormap(gray)%Display reconstructed image
imagesc(H7)
Figure 7.25 Image of the object used for generation of the Fresnel hologram.
Figure 7.26 (a) Image of the hologram and (b) reconstructed image of the object.
The image of the hologram and the reconstructed image are shown in
Figs. 7.26(a) and (b), respectively. Using the same method, it is also possible
to design holograms for objects with depth information. Objects at different
distances can be realized by constructing them in matrices with different
numbers of pixels.45 During reconstruction, different objects will be
reconstructed at different distances.
7.7 Conclusions
The design of DOEs using the computer-generated holographic method is
discussed in detail. CGHs ranging from simple 1D gratings to recently
developed elements that generate accelerating Airy beams are covered. The
procedures discussed in this chapter can be used for designing CGHs to be
7.8 Exercises
E.7.1 Design a CGH that focusus a diverging wavefront with a wavelength of
633 nm emanating from a point source a distance 10 mm away from the CGH
plane to a point a distance 20 mm from the FZP plane.
E.7.2 Design an off-axis FZP with a focal length of 25 mm and a tilt angle of
3 deg for a wavelength of 633 nm.
E.7.3 Using the holographic method, design a modified axicon with two
angles a1 and a2 such that a1 is twice that of a2 as shown in the figure below.
E.7.4 Design a CGH that can generate a focused ring pattern with a radius of
100 mm at a distance of 5 mm when illuminated by a uniform plane wave.
E.7.5 Design a CGH that can generate a helical wave front with topological
charge L ¼ 5 and focus it at a distance of 5 mm from the CGH plane.
References
1. J. W. Goodman, Introduction to Fourier Optics, Second Edition,
McGraw-Hill Companies Inc., New York (1996).
2. B. C. Kress and P. Meyrueis, Applied Digital Optics, John Wiley & Sons,
Chichester, UK (2009).
3. P. Hariharan, Optical Holography: Principles, Techniques and Applications,
Second Edition, Cambridge University Press, Cambridge, UK (1996).
177
number of zones, say, around 1000, then the procedure becomes tiresome and
time consuming, and the chances of making an error during design are high. In
some cases, such as multifunctional DOEs, it is not possible to create the DOE’s
design with the basic structures available in the software.
In advanced lithography systems such as the RAITH150TWO, Quanta™
400F (FEI), or the Quanta 3D FEG (FEI), to name a few, the CAD software
is fairly sophisticated. For example, it is possible to generate matrix copies in
the RAITH150TWO, which makes the design of periodic structures possible
with a single command. Additional Boolean operations such as OR, AND,
etc., also exist and are useful for designing multifunctional DOEs. In the case
of the Quanta 3D FEG, there is provision to load a bitmap file of a DOE
generated using another source, and this file can be scaled in the software
before fabrication. In older versions of some lithography systems, the DOE
design can be broken down and can be fabricated in sections. For instance, a
grating can be fabricated line-by-line. This technique is quite useful, as the
DOE patterning is possible with limited computational space and at the same
time with high resolution. The same technique can be extended to
multifunctional DOEs as well, by breaking them into a set of line data.
A major difference between a DOE design file created as an image file and
one created using CAD software is that in the former the circular objects are
piece-wise continuous, while in the latter they are polygons. As a result, the
file sizes are larger when they are created as image files. It is not
straightforward to create lithography files within MATLAB®. Hence, the
image files created using MATLAB are converted into lithography files using
software such as LinkCAD™. However, in software such as Python™, it is
possible to directly create output files in a lithography format such as GDSII.
Figure 8.1 Images of rings with a width of 10 mm with (a) pixel size of 1 mm and (b) pixel
size of 0.01 mm.
The same equations that were used for simulation can be employed again
for designing the DOEs but with an increased spatial resolution for the
lithography processes. The resolution of a design is given by the ratio of the
physical lengths of the pattern to the number of pixels in the length design.
For instance, if the length of a design file is 1000 pixels, and the physical
dimension is 10 mm, then the resolution of the design is 10 nm. In our designs,
most of the elements2,3 were generated with a matrix size of 20,000 20,000
corresponding to a device size of 2 mm 2 mm and, hence, a resolution of
100 nm. In some lithography systems such as the Quanta 3D FEG, the bitmap
file can be directly used for fabrication. The physical dimension of each pixel
of a bitmap file can be defined within the system. However, in most
lithography systems, including mask writers for photolithography, the input
design file must be in specific CAD formats such as DXF, GDSII, etc., as
discussed earlier.
For complex structures such as HOEs or IFTA-based DOEs, which
cannot be designed with CAD software, the following procedure can be used.
The DOE is designed as a bitmap file and converted to either DXF or GDSII
formats using LinkCAD software. The time of execution of the MATLAB
codes with a computer with a 3.4-GHz processor and a RAM of 16 GB varied
from few minutes to a couple of hours, depending on the different operations
involved. The choice of the scheme for designing DOEs depends on the
geometric composition of the structure and the required resolution.
For most of the examples in this book, the designs were generated using
MATLAB and converted to GDSII or DXF formats using LinkCAD. This
method of generating the design in a format accepted by the lithography
system may not be required when using advanced software that can be
purchased separately with systems like the RAITH150TWO. This software
accepts many file formats and allows one to create designs with high
resolution. However, such software is quite expensive. Therefore, alternative
procedures are needed. One such procedure for converting a matrix
representing a DOE to a GDSII file using LinkCAD is shown in Fig. 8.2.
Figure 8.2 Schematic of the procedure for converting a matrix or image generated in
MATLAB into a GDSII file using LinkCAD.
The pixel size of the element is the ratio of the element’s length in millimeters
to the size of the matrix. The pixel size is the same as the user units in this case.
Using LinkCAD is easy. The import format can be selected as ‘Image’,
and the export format can be selected as ‘GDSII’. There are a wide variety of
export formats such as DXF, CIF, etc. Next, the pixel size is selected. If the
matrix size is 20,000 20,000 for a device of size 2 mm 2 mm, then the pixel
size is 100 nm. The scaling factor can be selected, which in this case is 1. Then
the file can be loaded and converted. There are numerous other facilities
within the software that are useful for conversion. The converted file can be
viewed either in LinkCAD or with any other CAD software before patterning.
The output from LinkCAD can be directly used with any lithography system.
This is a far less expensive method of creating lithography files than using the
in-built special software available with some e-beam systems.
Table 8.1 MATLAB code for generating lithography data of a grating as lines.
%Generating line data for lithography
N=2000; % Size of the matrix
A=zeros(N,N); % Define matrices
B=zeros(N,N);
P=100; % Define period of the grating
p1=0; % Initialize line number
for q=1:N;
if rem(q,P),=P/2;
p1=p1þ1;
X1(p1)=1;
Y1(p1)=q;
X2(p1)=N;
Y2(p1)=q;
end
end
fid ¼ fopen(‘F:\1D grating.txt’,’wt’);%Open a text file
for s=1:p1;
fprintf(fid,’Draw Line %d,%d’,X1(s),Y1(s));
fprintf(fid,’\n’);
fprintf(fid,’ %d,%d’,X2(s),Y2(s));
fprintf(fid,’\n’);
end
fclose(fid);%Close the text file
Table 8.2 MATLAB code for generating lithography data of a circular grating as polygon-
approximated circles.
%Generating polygon data for lithography
N=2000; % Size of the DOE
A=zeros(N,N); % Define matrices
B=zeros(N,N);
theta1=zeros(M);
P=100; % Define period of the grating
M=1000; % Define the number of points in the circle
p1=0; % Initialize the iteration variable
for r=1:300; % Generate point data of every circle
for theta=1:M;
if rem(r,P),P/2;
theta1(theta)=((2*pi)/M)*theta;
p1=p1þ1;
X1(p1)=(r-1)*cos(theta1(theta));
Y1(p1)=(r-1)*sin(theta1(theta));
end
end
end
fid = fopen(‘F:\circular grating.txt’,’wt’);%Open text file
for s=1:p1;
fprintf(fid,’Draw Line %d,%d’,X1(s),Y1(s));
fprintf(fid,’\n’);
end
fclose(fid);%Close text file
Figure 8.3 (a) Plot of the points of the polygons approximated on the circles with 1000
vertices for a circular grating. (b) Magnified central portion of the image in part (a).
In this case, the DOE has been designed with 1000 points for each circle.
However, the number of points needs to be optimized based on the element
radius and the memory limit of the system. The central parts of polygonal-
approximated circular gratings with 10 and 25 vertices are shown in Figs. 8.4(a)
and (b), respectively.
The MATLAB code given in Table 8.1 for 1D gratings follows raster
scanning. The MATLAB code given in Table 8.2 for circular gratings follows
Figure 8.4 Plot of the polygonal path scanning of circular gratings with (a) 10 vertices and
(b) 25 vertices.
8.2 Photolithography
A brief description of the fabrication of devices using photolithography4,5
with basic fabrication recipes is presented in this section. A mask writer such
as Heidelberg Instruments DWL66 is used to prepare the mask pattern on a
glass plate with a chromium layer and a resist layer. The laser head can be
selected based on the minimum feature of the design, which is created using
MATLAB and then converted into a GDSII file using LinkCAD. The GDSII
file is converted into a binary file within the system for patterning.6,7 The
pattern is transferred to the resist layer of the mask plate by laser exposure.
The mask plate is developed using potassium hydroxide solution, during
which, the regions of the resist layer exposed to laser are removed. The mask
plate is then immersed in chrome etchant solution to remove the chromium
layer in the regions where the resist is removed. The chrome etchant is
prepared by combining a solution made from ammonium ceric nitrate and
acetic acid in de-ionized (DI) water. The mask plate is then rinsed in acetone
(80 °C) to remove the unexposed and undeveloped resist layer.
The mask plate can now be used for fabrication of a device on a resist-
coated glass plate. The size and shape of the glass substrate can be customized,
but in most cases the substrates used are ready-made large circular wafers. The
wafer can be diced into smaller sizes as required. In our case, we used
borosilicate glass wafers with a refractive index of 1.5 and with a diameter and
thickness of 3 inches and 500 mm, respectively. The glass wafer is diced using
diamond cutters or a dicing machine (Ultratec’s ULTRASLICE). The diced
glass wafer is cleaned in an ultrasonic bath in isopropyl alcohol (IPA) and
acetone for 2 min each to remove any chemical contaminants and dust
particles, followed by a rinse in DI water. It is then dried with nitrogen gas.
S1813 (Microchem) photoresist is used for fabrication of the device. The resist is
spin-coated on the glass substrate with a thickness of 1.5 mm. It is prebaked,
followed by an exposure in the UV lithography system (l ¼ 365 nm). The resist
is developed with sodium hydroxide developer and post-baked. The baking
conditions are optimized to harden the S1813 resist for etching processes.
The pattern in the resist layer can be transferred to the glass substrate by
wet8 or dry etching.9 Wet etching can be done using hydrofluoric acid (HF)
with the etch rates controlled by the concentration of the acid. The plasma
etching is carried out using Ar and SF6 gases. The etch rate in this case can be
precisely controlled by controlling the temperature, pressure, and power of the
system. If the period of the DOE is uniform throughout the pattern, then
constant etch depth can be obtained. However, this is not the scenario for
many DOEs. Let us consider a DOE such as a FZP, where the period of the
structure decreases radially outward with the minimum features occurring at
the outermost edges of the device. In this case, the etch rate will be higher for
larger openings than that for smaller ones. There is a strong etch depth
dependence with feature size. For demonstration purposes, a FZP with a
negative axicon phase is considered.
The DOE is designed for a wavelength of l ¼ 1064 nm to generate a ring
pattern with a radius of r0 ¼ 25 mm. The DOE is designed with a diameter of
8 mm. The inner maximum zone width and outermost minimum zone width
of the DOE are 67 mm and 4 mm, respectively. The optical microscope
images of the central and outermost parts of the DOE in chromium mask are
shown in Figs. 8.5(a) and (b), respectively. In this case, due to a slight
overdevelopment of resist in the mask plate, an error of 7% was found in the
width of the zones. The piece-wise continuous approximation of circles is
visible in the outermost zones in Fig. 8.5(b). The thickness of the chromium
pattern is measured using a confocal microscope (Olympus) as shown in
Fig. 8.6. This is the first evaluation step. The image of the DOE on the resist is
shown in Fig. 8.7(a). The profile of the DOE after wet etching with 10% HF
with an etch rate of 12.2 nm/min is shown in Fig. 8.7(b). The profile of the
DOE after plasma etching is shown in Fig. 8.7(c). In wet etching the control of
the thickness is not very accurate and besides it gives isotropic etching.
Figure 8.5 Optical microscope images of the (a) central part and (b) outermost part of the
DOE on a chromium layer.
Figure 8.6 Profile of the pattern measured using the Olympus confocal microscope
system.
Figure 8.7 (a) Optical microscope image of the DOE in the resist (S1813) layer, (b) profile
of the DOE fabricated using wet etching with 10% hydrofluoric acid, and (c) profile of the
DOE fabricated using plasma etching with Ar and SF6 (reprinted from Ref. 7).
In the case of wet as well as dry etching, S1813 was used as the mask. The
experiment was repeated to etch a larger depth of 1064 nm. The etch depth
was 1034 nm at the central part and 930 nm for the outermost part of the
Figure 8.8 Profile of the outermost part of the DOE after plasma etching with a recipe for
etching 1064 nm (reprinted from Ref. 7).
Figure 8.9 Surface profile images of sections of (a) a FZP, (b) a fractal zone plate, and
(c) a photon sieve after development and UV exposure (reprinted from Ref. 6).
This discussion has been limited to binary DOEs. Many publications are
available that address fabrication of multilevel DOEs using photolithography
with multiple masks using alignment markers and consecutive expo-
sures.1,11–13
After fabrication on resist and etching, the element must be characterized
using an optical microscope and a surface profiler for feature measurements
along the x, y, and z directions. The surface profile images obtained from an
interferometry-based profiler (such as one from Veeco Instruments) of sections
of a FZP, a fractal zone plate, and a photon sieve soon after development and
UV exposure are shown in Figs. 8.9(a)–(c), respectively.6,14,15 The final
investigation is performed through optical means by passing light through the
DOE and observing its diffraction pattern at the plane of interest using a 2D
detector such as a CCD.
Employing an interferometer-based surface profiler for characterization of
fabricated DOEs requires a certain level of expertise in optical alignment to
obtain the fringe pattern. Additionally, it is not possible to view features
smaller than the wavelength of light. In the recent advanced profilers such as
Olympus confocal microscopes, most of the features are automated, enabling
high precision and at the same time allowing a layman to analyze fabricated
devices. Apart from these profilers, there are mechanical contact stylus
profilers that can measure the profile of the fabricated DOEs by moving a pin
head over the surface of the DOE. (Pin heads with 0.1-mm diameter are
available.) However, this method of measurement can damage the DOE
pattern, and give only one line of data at a time. Hence, it is advantageous to
use optical surface profilers that can obtain the 3D information of the pattern
in a short duration.
pattern, as it is time consuming to fabricate DOEs with large areas. In the case
of EBL, DOEs are fabricated by deflection of the electron beam using electric
and magnetic fields. The spot size of the electron beam is controlled by the
different apertures in the lithography system. Aberrations such as astigma-
tism, etc., that are present in optical systems also exist in EBL systems.
However, most of the recently developed EBL systems have built-in tools to
minimize such aberrations. The results presented are from DOEs fabricated
with a RAITH150TWO EBL system. Most other EBL systems will have
similar procedures and operations.
harder substrate and can be used in the resist layer itself. As discussed in
earlier chapters, the phase (relating to the height of the DOE features) is
crucial when obtaining maximum efficiency in the 1st diffraction order. The
phase corresponds to a particular optical path length. For example, the
thickness of a resist layer that is equivalent to a phase value of p is given by
l/2(nr – 1), where nr is the refractive index of the resist. The refractive index of
the resist layer can be calculated from Cauchy’s coefficients, which are
provided in the resist data sheet or obtained by ellipsometry measurements.
The anisole concentration of the resist must be selected such that the spin-
coating speed for obtaining the necessary resist thickness is higher than
1000 rpm. This is necessary to obtain good uniformity of the resist layer. For a
large sample, a spin speed .5000 rpm requires higher pressure in the chuck to
hold the sample. High speeds may result in the sample been thrown from the
chuck. Given these constraints, PMMA A8 was selected, as the spin-coating
speed decreases to the desired range for l ¼ 633 nm and for binary DOEs. For
PMMA A8, the refractive index is such that the resist thickness value is
approximately the wavelength of the source used. Hence, for a wavelength of
633 nm, the resist height required is also 633 nm. PMMA A4 can be used in a
two-step spin-coating process. However, in this case, a resist layer with one-
half the thickness value is coated and baked, and the process is repeated to
obtain the full thickness value. There is no measurable difference between the
output intensity profiles achieved using A4 twice and A8 once, but the latter is
preferred due to the reduced number of steps required.
In order to repeatedly obtain the desired resist height, a calibration chart for
spin speed versus resist thickness (and acceleration) needs to be generated for
each resist to be used. One such graph for PMMA A8 is shown in Fig. 8.10.
The system was set at a fixed acceleration of 300 rpm/s and a coating duration
of 45 s. The thickness of the resist was measured using a confocal microscope.
Figure 8.10 Plot of resist thickness measured using a confocal microscope versus coating
speed. Acceleration was fixed at 300 rpm/s and spin coating time at 45 s.
Figure 8.10 is an initial crude estimate of the resist thickness versus spin
speed. However, since phase is very sensitive to errors in height, a finer
calibration curve is required. The range chosen for the more accurate calibration
is shown in the dashed box. The experiment is repeated, varying the spin speed in
steps of 100 rpm from 3500 rpm to 4500 rpm with constant acceleration, resist
quantity, resist temperature, and baking temperature. The spin speed that
generates a resist thickness of 633 nm was found to be 4300 rpm. The resist
thickness measurement values were found to have an error of 10 nm.
During electron beam patterning, some of the ITO layer needs to be
exposed to enable grounding and avoid charging effects. Therefore, after
baking, a section of the resist at the edge of the substrate was removed using
acetone. However, this method was found to affect other parts of the substrate
as well, causing some parts of the patterned region to peel off during
development. In order to avoid this problem, a region of the top of the sample
was masked using adhesive tape prior to spin coating the resist. The bottom of
the sample was also completely masked using tape to avoid any resist getting
coated on the back of the substrate due to suction because of the pressure in
the chuck. Both pieces of tape were removed immediately after spin coating
and prior to baking.
Figure 8.11 Schematic of the ITO sample mounted in the RAITH150TWO system (HMDS is
hexamethyldisilizane, an adhesion promoter) (adapted from Ref. 23).
Figure 8.12 SEM image of the ITO layer after focus and stigmation correction.
Figure 8.13 SEM images of the contamination spot burnt on the ITO layer with: (a) perfect
focus and stigmation correction, and (b) perfect focus without stigmation correction.
size, the resolution is higher, but stitching with adjacent fields is required,
resulting in stitching error. If the write field equals the design size, the
maximum resolution of the system can be utilized without any stitching. These
three cases are shown in Figs. 8.14(a), (b), and (c), respectively. A FZP
designed with front and back focal distances of 5 mm and 30 mm,
respectively, and l ¼ 633 nm is considered for analysis.
In the first case [Fig. 8.14(a)], it is clear that the resolution of the system is
wasted by choosing a larger write field. The possible resolution of patterning
in this case is N/65536, which is higher than the available resolution of
M/65,536. Therefore, this case should not be used to write any structures. In
the second case [Fig. 8.14(b)], the resolution is higher, as the write field is
much smaller than the design size. However, stitching is required, and the
stitching error was found to increase with the size of the write field. An
alternative method of increasing the resolution is to design different sections
of the element and then integrate during fabrication. However, this procedure
may not be successful due to the drift error present in the stage, as will be
further discussed in the following sections. Hence, the best scenario for
fabrication of DOEs is to exactly match the write field to the design size.
Whichever method is chosen, exposure and development times need to be
Figure 8.14 Write field with size M M and design size N N configuration for three
cases: (a) M . N, (b) M , N, and (c) M ¼ N.
Figure 8.15 Optical microscope images of the DOEs fabricated using electron beam direct
writing with (a) write field ¼ 20 mm and design size ¼ 2 mm, and (b) write field ¼ 4 mm and
design size ¼ 8 mm.
Figure 8.16 Optical microscope image of a DOE fabricated using electron beam direct
writing and overdeveloped.
Figure 8.17 Optical microscope images of the DOEs fabricated using electron beam direct
writing with no HMDS layer: (a) with modified baking temperatures and (b) without modified
baking temperatures.
the outermost part of the device are shown in Fig. 8.18. This problem was
solved by improving the manual focus and stigmation correction procedure.
Optical microscope images of the outermost part of the device and the full
device fabricated with optimized conditions are shown in Figs. 8.19(a) and (b),
respectively.
Naturally, when optimizing the fabrication procedure using EBL, users
will need to follow a version of this procedure that helps optimize fabrication
based on the system and resist that they are using. An example of a repeatable,
high-resolution fabrication technique developed after optimization of EBL
(RAITH150TWO) is given in Table 8.4. The outermost zones of a 2-mm device
are found to have a slight zig-zag of around 50 nm due to high beam
deflection [Fig. 8.19(a)] that cannot not be solved. In the RAITH150TWO
system, the smaller displacements of the stage are controlled by piezoelectric
devices. When glass substrates were used, it was noted that some charge
accumulation in that device resulted in a slight drift that was corrected by
Figure 8.18 Optical microscope images of the DOEs fabricated using electron beam direct
writing when the dose value was decreased at the outermost part of the devices.
Figure 8.19 Optical microscope images of the (a) outermost part of the DOE and (b) the
full device fabricated using electron beam direct writing with optimized fabrication
parameters (reprinted from Ref. 2).
turning off the joystick controller during patterning. There were other minor
technical problems due to the varying life of the chemicals, varying
temperature and humidity conditions, etc., that varied the fabrication results.
In most cases, the glass sample was not reused, as the quality of the resist layer
deteriorated with each re-use. The process flow for fabrication of DOEs using
EBL (RAITH150TWO) is given in Table 8.4.
There are different modes for patterning. For small patterns, the stage
must be kept fixed and the beam is deflected using electric and magnetic fields
(high resolution). For large patterns, the beam is kept fixed and the stage is
moved (low resolution) in what is called the fixed-beam moving stage (FBMS)
mode. In the DOEs presented in this book, the patterns were written with a
single write field without stitching and without FBMS. The acceleration
voltage, which controls the sharpness of the electron beam spot, the aperture,
which controls the electron beam spot size, and the working distance (the
distance between the column and the substrate), which controls the
magnification of the system, are set to the optimized values. The dose values
are optimized to calculate the clearing dose for every resist thickness for
binary patterns. For gradient patterns, the resist thickness is optimized for
different values of electron beam dose. Further fabrication of DOEs using
RAITH150TWO reveals that it is possible to fabricate elements with diameters
≤ 6.66 mm 6.66 mm without stitching and with a minimum feature of
$50 nm. The patterning time was reasonable, with 30–40 min for pattern sizes
of 2 mm 2 mm for a 120-mm aperture and 10-kV acceleration voltage.
8.3.4 Fabrication of multilevel structures
Fabrication of multilevel structures using an electron beam system has been
reported.24–26 One method to carry out this type of fabrication is to use a
grayscale resist such as PMMA 35K, which has a linear dose-to-resist-
thickness profile. In this section, we present techniques to fabricate multilevel
structures with a binary electron beam resist. The design of multilevel
structures and the calculation of resist thickness have already been presented
in detail in previous chapters.
Four-level and eight-level circular gratings with a period of 250 mm were
fabricated using electron beam direct writing in the RAITH150TWO system
using the fabrication parameters given in Table 8.4. The spin-coating
conditions and baking temperatures were optimized to obtain a resist
Figure 8.20 Plot of resist thickness after developing for different values of electron beam
dose.
thickness as close as possible to the calculated value of 1266 nm. The electron
beam resist selected for use is a binary resist with a very sharp resist-thickness-
versus-dose profile. The profile is also very sensitive to temperature variations
during processing. To improve the accuracy of the height, the resist-thickness-
versus-dose characterization must be carried out immediately before device
fabrication. Dose optimization is carried out by varying the dose from
10 mC/cm2 to 50 mC/cm2 in steps of 0.5 mC/cm2. In each case, the
development time is kept at 5 min. With such a high developing time, the dose
requirement is less, and the patterning time is considerably reduced. The
resist-thickness-versus-dose profile is plotted in Fig. 8.20.
The dose values corresponding to the different resist thickness values for a
four-level and an eight-level structure are noted from the graph. The device is
fabricated using two schemes. In the first scheme, the design is generated using
MATLAB, a much faster technique than the second scheme, where the design
is directly made using RAITH design software. In the first scheme, the design
is split into different layers with different dose values as a stack. During
fabrication, these layers are combined by assigning the same location. For the
four-level and eight-level circular gratings, the stack consists of four and eight
images, respectively. The images are generated for each dose value, as shown
in Fig. 8.21 for a four-level grating. During fabrication these images are
Figure 8.21 Images of sections of four-level axicons with different dose values.
Figure 8.22 Optical microscope images of the fabricated (a) four-level and (b) eight-level
axicons using the stack method.
stacked at the same location. A similar procedure is carried out for the
fabrication of the eight-level grating.
Optical microscope images of the four- and eight-level circular gratings
fabricated by the stack method are shown in Figs. 8.22(a) and (b), respectively.
Due to a stage error of few microns, the stack’s center shifted by a fixed amount
in the same direction for every layer of the stack. This stage error is prevalent in
many electron beam systems. Many lithography systems possess an in-built
correction for this error. In FEI systems, which we will discuss in the final
section, it can be noted that the drift correction can be completely nullified.
In the second method, the design is created in RAITH design software.
Each ring is designed individually with a different dose factor. The design was
initially carried out with no spacing between the two levels, resulting in
proximity errors,27 as shown in Fig. 8.23.
These proximity errors are minimized by giving spaces with a dose value
of 0 between each level. The optimized value of the manual proximity
correction space is 100 nm between every level. Optical microscope images of
the four- and eight-level circular gratings are shown in Figs. 8.24(a) and (b),
respectively. Resist profiles of the four- and eight-level axicons measured
using a confocal microscope are shown in Figs. 8.25(a) and (b), respectively.
The profiles clearly show that the proximity errors were reduced. The
roughness measurement using a profiler show a roughness variation from
Figure 8.23 Resist thickness profile variation over the period of the circular grating.
Figure 8.24 Optical microscope images of the fabricated (a) four-level and (b) eight-level
circular gratings designed using RAITH design software.
Figure 8.25 Resist profiles of (a) an eight-level and (b) a four-level circular grating
measured using a confocal microscope.
23 nm to 47 nm across the resist profile. The roughness value is higher for the
levels fabricated with higher electron beam dose. An average resist height
error of ,11% was obtained. The four-level and eight-level axicons were
evaluated using a diode laser with a wavelength of 633 nm. The light from the
laser source was collimated using a 10X objective. The Bessel beams generated
by the four-level and eight-level axicons and their corresponding intensity
profiles are shown in Figs. 8.26 and 8.27, respectively.
The calculated value of the 1/e2 diameter was 150 mm, while the
experimental value was found to be 162 mm. The average transmittivity of the
ITO layer is 85% for l ¼ 633 nm. The efficiencies of the four-level and
eight-level circular grating were found to be 41% and 75%, respectively.
Figure 8.26 (a) Image of the Bessel beam and its (b) intensity profile generated by a four-
level circular grating.
Figure 8.27 (a) Image of the Bessel beam and its (b) intensity profile generated by an
eight-level circular grating.
The decrease in efficiency was partly due to the transmittivity of the ITO layer
and partly due to the resist height errors.
Multilevel axicons with four and eight levels were designed and fabricated
using electron beam direct writing. The fabrication was carried out using a
binary electron beam resist with a very steep resist-thickness-versus-dose
profile. The evaluation results show the generation of Bessel beams within the
focal depth of the devices. The efficiency of the eight-level structure was close
to the theoretical efficiency, although this was not the case for the four-level
structure. The repeatability in fabrication was improved, as the fabrication
was carried out immediately after dose calculation. Some other possible
reasons for the decrease in repeatability might be the life of the developer
solution, variation in electron beam current, manual errors in fabrication
processes with identical conditions, etc. Also, color changes in the resist were
noted; these changes corresponded to different thicknesses of the resist, which
might be an additional reason for the decrease in efficiency.
Figure 8.28 Schematic of the optical testing set up for evaluation of DOEs.
Figure 8.29 Optical microscope image of the FZP designed in finite conjugate mode after
aberration correction (reprinted from Ref. 2).
Figure 8.30 Normalized intensity profile at the image plane for the FZP without aberration
correction (dotted line), with aberration correction using scheme 1 (solid line), and with
aberration correction using scheme 2 (dashed line) (reprinted from Ref. 2).
Figure 8.31 (a) Optical microscopic image of the multifunctional DOE containing the
functions of a binary FZP and a binary circular grating, and images of the ring pattern
generated by the DOE for the wavelengths (b) 635 nm and (c) 532 nm (reprinted from
Ref. 2).
Figure 8.32 Optical microscope images of the multifunctional DOE containing the
functions of a ring FZP (f ¼ 30 mm, r0 ¼ 100 mm) with (a) a 1D grating and (b) a 2D
checkerboard grating (L ¼ 50 mm).
Figure 8.33 Images of the diffraction patterns of the ring pattern arrays generated by the
DOEs shown in (a) Fig. 8.32(a) and (b) Fig. 8.32(b).
distance of 30 mm from the DOEs in Figs. 8.32(a) and (b) are shown in
Figs. 8.33(a) and (b), respectively.
Optical microscope images of the multifunctional DOE designed by
combining a binary circular grating and a binary 1D grating and 2D grating
(Section 6.1.2) are shown in Figs. 8.34(a) and (b), respectively. The far-field
diffraction patterns of Figs. 8.34(a) and (b) measured at a distance of 200 mm
are shown in Figs. 8.35(a) and (b), respectively.
Figure 8.34 Optical microscope images of the multifunctional DOE containing the
functions of a circular grating (L ¼ 200 mm) and a (a) 1D grating and (b) 2D checkerboard
grating (L ¼ 100 mm).
Figure 8.35 Image of the 1st and 3rd order diffraction patterns of the ring pattern arrays
generated from the DOEs shown in (a) Fig. 8.34(a) and (b) Fig. 8.34(b) recorded in a CCD.
Figure 8.36 (a) Optical microscope image of the 3 3 circular grating array and (b) the
Bessel intensity interference pattern recorded using a CCD.
Figure 8.37 Optical microscope images of the binary multifunctional DOE generated
by combining the FZP (f ¼ 30 mm) containing the phase of a negative axicon with
(a) X ¼ 1.32 and (c) X ¼ 2.64. (b) Image of the ring patterns generated by the same DOE
radius values of r0 ¼ 25 mm and (d) r0 ¼ 25 mm recorded by a CCD (reprinted from
Ref. 29).
Figure 8.38 (a) Optical microscope image of a HOE used for the generation of an
accelerating Airy beam with a path profile of t2(x, y) ¼ 10(x)0.5 þ 10(y)0.5. (b) Far-field
diffraction pattern of the HOE generated by the superposition of a curved plane wavefront
and a conical wavefront with t2(x, y) ¼ 10(x)0.5 þ 10(y)0.5. (c) A magnified version of the far-
field diffraction pattern in (b) (reprinted from Ref. 30).
Figure 8.39 (a) Schematic of ITO-coated samples mounted in a Quanta 3D FEG system
and (b) photograph of the fabrication chamber of a Quanta 3D FEG.
order to obtain the correct functionality. The FIB system consists of different
application files containing material parameters corresponding to substrates
such as silicon, gold, etc. However, there is no built-in application file for glass
or for the ITO-coated substrates. Hence, the system is optimized by varying the
current, while maintaining the parameters acceleration voltage, volume per
dose, and dwell time at 30kV, 0.02 mm3/nC, and 1 ms, respectively. The
software of the FIB converts the 24-bit bitmap files into dwell time and beam
position information. White pixels correspond to the user-defined dwell time,
while black pixels correspond to a minimum dwell time of 100 ns. The current is
varied between 0.1 and 1 nA, with 1 nA providing good results.
These numbers can be taken as a guide as to where to begin the
optimization of the milling process. However, in general, optimization of
milling take into account the material being milled and the structure sizes. For
example, re-deposition will be a major problem, especially with finer features.35
Most researchers will carry out an extensive optimization36–38 before
fabricating the final structures. In this system, multilevel structures are possible
using the stack method (Fig. 8.21), which failed with the electron beam system.
The stack method was employed for fabrication of multilevel spiral phase plates
(SSPs).39 An electron beam image of the fabricated FZP is shown in Fig. 8.40.
The design files of the stack for fabrication of a four-level SPP and the
electron beam image of the fabricated device are shown in Figs. 8.41(a)
and (b), respectively. The design files of a blazed 1D grating and a SPP along
with their respective electron beam images are shown in Fig. 8.42. The
parameters for the blazed structures are 30 kV, 3 nA, set depth of 3 mm, and
volume per dose of 0.15 mm3/nC. For the four-level SPP, the values are 30 kV,
0.3 nA, set depth varying in each wedge pattern, and 0.15 mm3/nC. Both
elements were fabricated using the Si application file.
FIB milling has many advantages over EBL and photolithography in
terms of the number of processing steps involved. However, FIB milling has
Figure 8.40 Electron beam image of the FZP fabricated using Quanta 3D FEG.
Figure 8.41 (a) Images of the stack files for fabrication of four-level SPP and (b) electron
beam image of the fabricated four-level SPP.
Figure 8.42 (a) Image of the blazed grating, (b) electron beam image of the fabricated
blazed grating, (c) Image of the gradient SPP, and (d) electron beam image of the fabricated
gradient SPP.
Figure 8.43 Images of the fiber tip along the x and y directions (a) after stripping the buffer
layer and (b) after cleaving.
Figure 8.44 (a) Holder with v-grooves to hold four fibers. The top part (shown in the
dashed rectangle) fits into the FIB system. When coating the fibers, the top part along with
the holder, sit in the coating chamber of the sputtering unit. (b) Close-up SEM photograph of
one fiber in the FIB system (reprinted from Ref. 45).
Figure 8.45 Images of the fiber (a) before processing, (b) after stripping away its buffer
layer, (c) after cleaving, and (d) after metallization. (Figures not drawn to scale.)
Figure 8.46 Electron beam images of the fiber tip (a) with and (b) without contamination.
Fig. 8.44(b). The processing steps before FIB milling are shown in Fig. 8.45.
Electron beam images of the fiber tip with and without contamination are
shown in Figs. 8.46(a) and (b), respectively.
A unique problem faced when milling fiber relates to the placement of the
structure with respect to the core of the fiber. The structure must be milled over
the core. However, due to the thin metal coating, the core may not be visible.
If the patterns of the structure vary only in one direction or two orthogonal
Figure 8.47 Cross section of a length of fiber (a) before and (b) after etching and
metallization. The central lighter color represents the core.
directions, one can overcome the problem of locating the center in the following
way. Initially, the center of the fiber can be geometrically located and a structure
larger than the core can be written. However, for structures with circular
symmetry, it is important for the center of the structure to match the center of
the core, which is difficult to achieve when the core is not visible. One way to
solve this problem is to etch the fiber tip before metallization.44 The slightly
higher etch rate of the core compared to that of the cladding creates a step that is
visible even after metallization. The center of the step can then be geometrically
located. Cross sections of the core and cladding are shown in Fig. 8.47.
Alternatively, the gold layer around the core region can be removed by
milling a circle of 20 mm with an ion beam current of 3 nA for 4 min. This initial
patterning exposes the core, which helps to accurately position the pattern.
Depth optimization is achieved as before. The parameters acceleration voltage,
volume per dose, and dwell time are kept constant at 30 kV, 0.15 mm3/nC, and
1 ms, respectively. The current is varied between 0.1 and 0.3 nA, with 0.3 nA
providing good results. The set depth is varied for these parameters from 0.5 to
1.5 mm. To test the depth obtained in each case, an extra line of platinum is
deposited using the gas injector needle available with the system. An extra
portion of the milled structure is removed by further milling in order to measure
the cross section (and thereby obtain the actual depth milled). A SEM image of
the structure from which the cross section can be measured is shown in Fig. 8.48.
For the actual fabrication of the structures on the fiber tips, 300 pA of
beam current and 30-kV acceleration voltage with a maximum dwell time of
1 ms, 10-nm steps, and an antiparallel beam movement are utilized. Since the
substrate is a dielectric, accumulation of charge is possible, especially if
the core is exposed by milling out a circle. This, in turn, can cause a drift of the
ion beam during the milling process. The effects are clear in Fig. 8.49(a),
which shows a 1D grating milled with no drift correction applied. With each
layer milled, the beam position changes slightly, and the final grating appears
to be smeared out. In contrast, the grating is written without this error when
drift correction is applied, as seen in Fig. 8.49(b).
Drift correction is carried out by first milling a pattern that will act as an
alignment marker. Typically, a cross with good contrast is milled as shown in
Figure 8.48 SEM image showing the depth milled for a structure consisting of four parallel
lines (reprinted from Ref. 45).
Figure 8.49 Image of a 1D grating milled (a) with beam drift and (b) with drift correction.
Fig. 8.50. In order to achieve the better contrast, a region of platinum is deposited
and the cross milled into that area. A binary axicon (of diameter 10 mm with a
period of 1 mm and fill factor of 50%) written on the core of a fiber is shown in
Fig. 8.51. Drift correction is applied, and the obtained pattern is clear.
Recent research by Juodkazis et al. shows that illumination of the
substrate with UV light can help to discharge ions from the surface, indicating
Figure 8.50 Image of a fiber with an alignment mark (to the left of the core) that
subsequently will have a diffractive structure written on it. The alignment mark is shown
enclosed within the dashed ellipse.
Figure 8.51 Image of a binary axicon written over the core of a single-mode fiber
(Reprinted from Ref. 45).
that FIB milling can be done even without a metal coating.50 This would be
extremely useful when working with fibers.
8.5 Conclusions
This chapter studies the fabrication of DOEs using photolithography, EBL,
and FIB lithography. Of particular importance is the transition from the
diffraction equations to a design (CAD file) that can be understood by the
writing tool. The simplest methods for generating the CAD files required for
fabrication are discussed in detail. The basic recipes for fabrication of DOEs
with all three types of lithography tools are also presented. As there is
already a lot of literature on photolithography, the focus of this chapter was
more on fabrication using electron beam and focused ion beam lithography
systems.
Fabrication of DOEs using FIB is a relatively new area compared to
photolithography and EBL. However, the advantages of fabrication of DOEs
using FIB are enormous. In particular, FIB makes it almost easy to fabricate
elements on a fiber tip, which has immense potential in the newly evolving
fields of nanophotonics, plasmonics, and biomedicine. The different problems
associated with the fabrication of DOEs using FIB are addressed, and step-by-
step procedures for such are presented.
References
1. B. C. Kress and P. Meyrueis, Applied Digital Optics, John Wiley & Sons,
Chichester, UK (2009).
2. A. Vijayakumar and S. Bhattacharya, “Characterization and correction
of spherical aberration due to glass substrate in the design and fabrication
of Fresnel zone lenses,” Appl. Opt. 52, 5932–5940 (2013).
Acknowledgment
We thank Dr. Pramitha Vayalamkuzhi, Department of Electrical Engineer-
ing, IIT Madras, Chennai, India for her special contributions to and review of
this chapter.
Table A.1 MATLAB function to generate normalized output for amplitude structures.
function Norm_outputA(x, N)
% This function generates the output intensity pattern with
%intensity normalized for amplitude DOEs
% function arguments (required when calling the function) are
% DOE amplitude (x) and matrix size (N)
E¼fftshift(fft2(A)); %fftshift is used to re-order the terms in
%their natural order
IN¼(abs(E)/(N*N)).*(abs(E)/(N*N)); % Calculating intensity
figure (1)
colormap(gray); %colormap(gray) is used to display grayscale
%image
imagesc(A);% imagesc is used to display a high constrast image
figure (2)
colormap(gray);
imagesc(IN);
217
Table A.3 MATLAB function to generate normalized output for phase structures.
function Norm_outputP(x, N)
% This function generates the output intensity pattern with
%intensity normalized for phase DOEs
% function arguments (required when calling the function) are
% DOE amplitude (x) and matrix size (N)
E¼fftshift(fft2(x)); %fftshift is used to re-order the terms in
%their natural order
IN¼(abs(E)/(N*N)).*(abs(E)/(N*N)); % Calculating intensity
figure (1)
colormap(gray); %colormap(gray) is used to display grayscale
%image
imagesc(angle(x));% imagesc is used to display a high constrast
%image
figure (2)
colormap(gray);
imagesc(IN);
A. Vijayakumar and Shanti Bhattacharya,
Design and Fabrication of Diffractive Optic Elements
with MATLAB®
SPIE Press, Bellingham, Washington (2017).
Click here to download supplementary material.
In case the above button does not work, the supplemental files for this eBook are
also available for download by copying and pasting the following link to a
browser:
http:/spie.org/Samples/Pressbook_Supplemental/TT109_sup.zip
Chapter 2 Solutions
E.2.1 Equate Eqs. (2.4) and (2.5).
I 1 ¼ 0.5I 0 ,
4
sin2 F ¼ 0.5cos2 F ,
p2 2 2
F ¼ 1.67 rad:
Displayed values in figure window 2 are 0.22 and 0.45 for the 1 and 0th
diffraction orders, respectively.
219
Figure E.2.2 Image of the diffraction spots seen by zooming in on figure window 2.
E.2.3 The simplest method is to draw a triangle of size 100 100 pixels and
generate the full matrix with the ‘repmat’ command.
E.2.4 Define two focal length values fx ¼ 3000 mm and fy ¼ 6000 mm for the
grating lines along the x and y directions, respectively, in MATLAB code.
The MATLAB code for generation of a 2D FZP is shown in Table E.2.4. The
image of the 2D grating seen in figure window 1 is shown in Fig. E.2.4.
Figure E.2.5 Image of the elliptical FZP generated with focal length values fx ¼ 3000 mm
and fy ¼ 4500 mm.
and 2.8. A single axicon is designed and is subsequently replicated to form the
complete device. The MATLAB code is given in Table E.2.6. The image of the
axicon array generated using MATLAB is shown in Fig. E.2.6.
Figure E.2.6 Image of the axicon array generated with a period of L ¼ 100 mm.
Chapter 3 Solutions
E.3.1 Calculate the phase values in steps of p/8 and the corresponding resist
thickness values. The values are shown in Fig. E.3.1a.
Figure E.3.1a Phase values and resist thickness values of a 16-level 1D phase grating.
The MATLAB code shown in Table 3.2 can be modified by changing the
number of phase levels and the phase increment as given in Table E.3.1. The
phase profile of the 16-level 1D phase grating is shown in Fig. E.3.1b. The
displayed efficiency value is 0.987, which matches with the value (0.987)
calculated using Eq. 3.1. The image of the phase profile looks similar to that
of a blazed grating.
The image of the phase profile of a negative blazed FZP and a four-level FZP
are shown in Fig. E.3.2.
Figure E.3.2 Phase profile of (a) a blazed negative FZP (dotted line) and (b) its four-level
approximation (solid line).
The phase profile of the binary SPP with charge L ¼ 5 before applying
radius constraint and its far-field diffraction pattern are shown in Figs. E.3.3
(a) and (b), respectively.
Figure E.3.3 (a) Phase profile of a binary SPP with topological charge of L ¼ 5 and (b) its
far-field diffraction pattern.
E.3.4 To design the three-level DOE, the same logo was considered with ten
iterations. The MATLAB code is same as that given in Table 3.9 except that
the grayscale phase profile of the DOE has to be converted into a three-level
phase profile. The MATLAB code for design of a three-level DOE is shown in
Table E.3.4.
Table E.3.4 MATLAB code for design of a three-level DOE using the IFTA.
DOE1¼zeros(N,N);
g¼3;%Define the number of phase levels
delphase¼2*pi/3;%Define the phase increment
for p¼1:N;%Convert the greyscale phase profile into a 3-level
%phase profile
for q¼1:N;
for n¼1:g;
if DOE(p,q).-piþ(n-1)*delphase && DOE(p,q),¼-piþ(n)*
delphase;
DOE1(p,q)¼(n-1)*delphase;
end
end
end
end
%Verification of result
DOE2¼exp(1i*DOE1);
I¼abs(fft2(DOE2)); %Calculate the Fourier transform
colormap(gray)
Imagesc(I)
Figure E.3.4 Image of the logo generated from the three-level DOE.
E.3.5 The design of a ring lens is similar to that described in Chapter 2. The
optics configuration for focusing light on a ring using a ring lens is shown in
Fig. E.3.5a.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2p
FRing lens ðrÞ ¼ f f 2 þ ðr r0 Þ2 : (E.3.5)
l 2p
The MATLAB code is similar to that of a gradient FZP with the replacement
of the phase equation, as shown in Table E.3.5. The image of the ring lens and
the plot of its phase profile are shown in Figs. E.3.5b and Fig. E.3.5c,
respectively.
Figure E.3.5b Image of the phase profile of the gradient ring lens.
Figure E.3.5c Plot of the phase profile of the gradient ring lens at y ¼ 250 pixels.
Chapter 4 Solutions
E.4.1 The MATLAB code is similar to that shown in Table 4.2 while the DOE
design script has to be replaced with that in Table E.4.1.
The diffraction pattern at z ¼ 50 m is shown in Fig. E.4.1. It can be noted
that the other higher-order ring patterns vanished, with only a high-efficiency
first-order ring pattern remaining.
Table E.4.1 MATLAB code for design and analysis of blazed axicon using the Fresnel
diffraction formula.
%Fresnel diffraction of blazed axicon%
clear; %Clear all memory
% Defining the parameters
N¼500;% Define the matrix size
lambda¼0.633*10^-6;%Define wavelength in meters
z¼50;%Propagation distance ¼ 50 m
P¼10^-4;%Period of axicon ¼ 0.1 mm
wsamp¼10*lambda;%sampling period or width
%Sampling the space
x¼1:N;
y¼1:N;
[X,Y]¼meshgrid(x,y);%Sampling
Rsamp¼sqrt((X-N/2).^2þ(Y-N/2).^2).*wsamp;%Define sampled
%radius
%Constructing the DOE
A¼ones(N,N);%Define matrix by assigning ones to all pixels
A¼exp(1i*(rem(Rsamp,P))*(2*pi)/P);
A(Rsamp.N/2*wsamp)¼0;
% Calculating the Fresnel diffraction
PPF¼exp(1i*pi/(lambda*z).*Rsamp.*Rsamp); %Calculate the
%parabolic phase factor
A1¼A.*PPF; %Multiply the circular aperture function with
%the parabolic phase factor
E¼abs(fftshift(fft2(fftshift(A1)))); %Calculate Fourier
%transform
E.4.2 The MATLAB code for designing the phase element is given in
Table E.4.2. The images of the diffracted fields at z ¼ 5 mm, 10 mm, 20 mm,
and 50 mm are shown in Fig. E.4.2a, and the plots of the cross section in the
same figure are shown in Fig. E.4.2b. There is a shift in the location of the
maximum intensity. The shift in the peaks shows a nonlinear response that
matches the phase profile of the DOE.
Table E.4.2 MATLAB code for design and analysis of a DOE using the Fresnel diffraction
formula.
%Fresnel diffraction of DOE%
%clear; %Clear all memory
% Defining the parameters
N¼500;% Define the matrix size
lambda¼0.633*10^-6;%Define wavelength in meters
z¼10*1e-3;%Propagation distance
del¼1*1e-6;%sampling period or width
%Sampling the space
x¼-N/2:N/2-1;
y¼-N/2:N/2-1;
[X,Y]¼meshgrid(x*del,y*del);%Sampling
R¼sqrt(X.^2þY.^2);%Define sampled radius
%Constructing the DOE
A¼exp(1i*(Xþ250*1e-6).^3*2*pi*5*10^11);
% Calculating the Fresnel diffraction
PPF¼exp(1i*pi/(lambda*z)*R.*R); %Calculate the parabolic
%phase factor
A1¼A.*PPF; %Multiply the circular aperture function with
%the parabolic phase factor
E¼abs(fftshift(fft2(fftshift(A1)))); %Calculate Fourier
%transform
E¼E/max(max(E));
figure (1)
imagesc(E);
figure (2)
%plot(E(N/2,:)/(max(max(E(N/2,:)))))
Figure E.4.2a Images of the intensity profile at z ¼ 5 mm, 10 mm, 20 mm, and 50 mm.
Figure E.4.2b Plots of the intensity profile at z ¼ 5 mm (solid line), 10 mm (dashed line),
20 mm (dotted line), and 50 mm (dashed and dotted line).
E.4.3 When a linear phase is used, there is no shift of the maximum intensity
with respect to the propagation distance due to the scaling factor that
compensates for the shift that appears in a lab experiment.
Chapter 5 Solutions
E.5.1 The optical path length equation for focusing a diverging wave on a ring
of radius r0 is given by
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u2 þ r2n þ v2 þ ðrn r0 Þ2 u2 þ r20 v ¼ nl, (E.5.1)
where, u and v are the object and image distances, respectively. Assuming that
r0 ,, (r, u and v), Eq. (E.5.1) can be expressed as
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r20
u þ rn þ v þ ðrn r0 Þ u 1 þ 2 v ¼ nl:
2 2 2 2
(E.5.2)
2u
For (r02/2u2) ,,1, using binomial expansion,
sffiffiffiffiffiffiffiffiffiffiffiffiffi
r20 r20
u 1þ 2 ≅u 1þ 2 : (E.5.3)
u 2u
Solving Eq. (E.5.3) for rn yields
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
b b2 4ac
rn ¼ , (E.5.4)
2a
where
r4 r2
a ¼ u2 þ v2 þ n2 l2 þ 2uv þ 2nlðu þ vÞ þ 0 2 þ 0 ðv þ nlÞ,
4u u
4 2
r0 r0
b ¼ r0 n l þ 2 þ 2uv þ 2nlðu þ vÞ þ
2 2
ðv þ nlÞ þ 2u ,
2
4u u
2
1 2 2 r0 4 r0 2
c ¼ n l þ 2 þ 2uv þ 2nlðu þ vÞ þ ðv þ nlÞ þ u2 ðv2 þ r0 2 Þ:
4 4u u
The technique discussed in Section 5.2 is used to find the location of the
virtual sources u’ generated by the glass substrate:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
0
b b02 4a0 c0
r0n ¼ , (E.5.5)
2a0
where
r4 r2
a0 ¼ u02 þ v2 þ n2 l2 þ 2u0 v þ 2nlðu0 þ vÞ þ 0 02 þ 00 ðv þ nlÞ,
4u u
4 2
0 r0 0 0 r0 02
b ¼ r0 n l þ 02 þ 2u v þ 2nlðu þ vÞ þ 0 ðv þ nlÞ þ 2u ,
2 2
4u u
2
1 2 2 r0 4 r0 2
c ¼ n l þ 02 þ 2u v þ 2nlðu þ vÞ þ 0 ðv þ nlÞ þ u02 ðv2 þ r0 2 Þ:
0 0 0
4 4u u
E.5.2 The radius of a FZP can be calculated with the inclusion of the glass
substrate and the additional polymer holder using Eq. (5.18). The thickness
function is not constant as in the earlier case, but is binary with two values,
namely, 1 mm and 3 mm, respectively at different radial sections. The plot of
the radii of zones as a function of zone number for case 1, t ¼ 1 mm (constant)
and case 2, t ¼ 3 mm (0.5 mm , r , 0.7 mm), and t ¼ 1 mm elsewhere is
shown in Fig. E.5.2. The radii values are different in the region (0.5 mm , r ,
0.7 mm), as expected.
Figure E.5.2 Plot of the radii of zones for case 1: t ¼ 1 mm (solid line) and case 2: t ¼ 3 mm
(0.5 mm , r , 0.7 mm), and t ¼ 1 mm elsewhere (dotted line).
E.5.3 The radius of a FZP can be calculated with the inclusion of the glass
substrate whose thickness varies with respect to the radial coordinate. The
radii values are plotted as a function of zone number for two cases—case 1:
t ¼ 1 mm and case 2: t ¼ 1 þ r, as shown in Fig. E.5.3.
Figure E.5.3 Plot of the radii of zones for case 1: t ¼ 1 mm (dotted line) and case 2: t ¼ 1 þ
r (solid line).
Chapter 6 Solutions
E.6.1 The two DOEs of choice are a FZP and a circular grating or axicon to
generate a focused ring pattern with a diameter of 1 mm at a distance of
30 mm. A circular grating generates a ring pattern in its far field and when it is
used together with a FZP, and it generates a focused ring pattern at the focal
plane of the FZP. Hence, the focal length of f ¼ 30 mm and the period of the
circular grating can be estimated from the diameter of the ring pattern, which
is 1 mm.
l d
sinu ¼ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
: (E.6.1a)
L d2
þf2
4
After simplification,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
l d 2 þ 4f 2
L¼ ¼ 38 mm: (E.6.1b)
d
The MATLAB code for generation of the binary phase profile of a FZP
and a circular grating using the modulo-2p phase addition method for
generation of the multifunctional DOE are given in Table E.6.1.
Table E.6.1 MATLAB code for design of a multifunctional DOE from a binary circular
grating and a binary FZP using the modulo-2p phase addition method.
%Multifunctional DOE – Circular grating and FZP – Modulo-2p phase
%addition method
clear; %Clear all memory
%Define grating parameters
N¼1000;%Define matrix size
M¼10;%Define number of half period zones of FZP
f¼30000; %Define focal length of FZP
lambda¼0.633;%Define wavelength
A1¼zeros(N,N); %Define a Matrix by assigning 0 to all
%elements
A2¼zeros(N,N);
r¼zeros(N,N);
r1¼zeros(M);
Pr¼38; %Define the period of the grating
FFr¼0.5; %Define fill factors for x and y periodicity
%Construct the binary FZP and binary circular grating
for n¼1:M;
r1(n)¼sqrt(n*f*lambda);
end
for n¼1:2:M;
for p¼1:N;
for q¼1:N;
r(p,q)¼sqrt((p-N/2)*(p-N/2)þ(q-N/2)*(q-N/2));
if r(p,q),N/2;
if rem(r(p,q),Pr),Pr*FFr;
A1(p,q)¼1;
end
if r(p,q) . r1(n) && r(p,q) , r1(nþ1);
A2(p,q)¼1;
end
end
end
end
end
A3¼exp(1i*pi*xor(A1,A2)); %XOR operation between A1 and A2
Figure E.6.1a Image of the phase profile of the multifunctional DOE containing the
functions of a circular grating with a period of 38 mm and a FZP with a focal length of
f ¼ 30 mm.
d∕2
b ¼ sin1 ðng sinaÞ a ¼ ¼ 0.0167 rad: (E.6.1c)
f
Assuming that ng ¼ 1.5, the base angle of the axicon is given by 0.0335 rad.
Hence, from the relationship between the thickness and X, we find that X ¼
26.5. The radii of zones of the multifunctional DOE can be estimated using
Eq. (6.19).
The MATLAB code for designing the multifunctional DOE using the
analog method is similar to the code given in Table E.6.1. However, the
equation of radii of zones must be replaced by Eq. (6.19). The plot of radii of
zones as a function of zone number is shown in Fig. E.6.1b. It can be noted
that the radius of the first zone is . 1 mm. Hence, it is not possible to design
the element using the analog method for the above design values.
Figure E.6.1b Plot of the radii of zones of the multifunctional DOE as a function of the zone
number.
E.6.2 There are many interesting optics configurations for superposing two
signals with slightly different wavelengths. A simple configuration is as
follows. The multifunctional DOE needs to have an axicon and a grating. The
axicon converts the Gaussian intensity profile into a Bessel intensity profile,
while the grating can be used for beam combining. The period of the grating
must be selected to obtain a diffraction angle of 5 deg (0.0872 rad). As the
wavelengths are close to one another, a grating with same period can be used
for the average wavelength lavg ¼ 602 nm. The period of the grating is
estimated as L1 ¼ 7 mm. The distance between the central maximum and the
first minimum is given by
1.22l
r0 ¼ : (E.6.2)
psinb
Hence, the period of the axicon is L1 ¼ 51 mm. The element can be designed
using the MATLAB code given in Table E.6.2.
Table E.6.2 MATLAB code for design of multifunctional DOE from a blazed axicon and a
1D grating.
%%Multifunctional DOE- Blazed axicon and binary grating
clear;%Clear all memory
%Define grating parameters
N¼500;%Define Matrix size
Pr¼51;
Px¼7;
FFx¼0.5;
A2¼zeros(N,N);%Define the matrices assigning ones to all
%pixels
%Construction of blazed axicon and binary grating
x¼1:N;
y¼1:N;
[X,Y]¼meshgrid(x,y);
r¼sqrt((X-N/2).*(X-N/2)þ(Y-N/2).*(Y-N/2));
P1¼rem(r,Pr);
A1¼(P1/Pr)*2*pi;
A2(rem(X,Px),Px*FFx)¼pi;
A¼exp(1i*rem(A1þA2,2*pi));
The grating needs to be binarized such that the þ1st diffraction order of
one of the beams will superpose the –1st diffraction order of the other beam, as
shown in Fig. E.6.2a. The magnified central region of the DOE is shown in
Fig. 6.2b.
Figure E.6.2a Optics configuration for superposing light beams with slightly different
wavelengths propagating at þ5 deg and –5 deg.
Figure E.6.2b Image of the phase profile of a multifunctional DOE containing the functions
of a blazed axicon and a binary grating used for converting a Gaussian intensity profile into a
Bessel intensity profile and for beam combining.
E.6.3 The phase profiles of a blazed spiral phase plate with charge L ¼ 10 and
a binary amplitude axicon with period L ¼ 10 mm are added. The MATLAB
code is similar to that given in Table 6.5, and the FZP phase profile must be
replaced by the phase profile of a binary axicon or a circular grating. The
image of the multifunctional DOE and its far-field diffraction pattern are
shown in Fig. E.6.3(a) and (b), respectively.
Figure E.6.3 (a) Image of the phase profile of a helical axicon and (b) the far-field
diffraction pattern of a helical axicon.
E.6.4 Two DOEs, namely, a ring lens29,30 and a checkerboard grating, are
selected for generation of 2 2 ring patterns. The period of the grating
calculated using trigonometry and the diffraction equation for generation of
ring patterns with a spacing of 758 mm is 50 mm. The binary phase profile of
the ring FZP can be generated using the path length equation, given by27
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðrn r0 Þ2 þ f 2 f ¼ nl: (E.6.4a)
The MATLAB code is similar to Table 6.2, where the phase profiles of the
binary axicon and 1D grating must be replaced by those of the ring FZP and
the 2D checkerboard grating. The image of the DOE is shown in Fig. E.6.4. It
can be observed that the diameter of the ring pattern is independent of
wavelength, as the element is a ring lens. The distance between the rings is also
independent of wavelength, as the variation in the focal length of the ring lens
due to wavelength variation compensates for the variation in distance between
the ring patterns due to wavelength variation.
Figure E.6.4 Image of a multifunctional DOE containing the functions of a binary ring FZP
and a binary checkerboard grating.
Chapter 7 Solutions
E.7.1 The CGH that focuses a diverging wavefront into a point is nothing but
a FZP designed in finite conjugate mode. The object and image distances are
10 mm and 20 mm, respectively. This CGH can be designed by superposition
of the two wavefronts described by Eqs. (E.7.1.a) and (E.7.1.b). The design
part of the MATLAB code is given in Table E.7.1. The image of the FZP
generated using the above MATLAB code is given in Fig. E.7.1.
Table E.7.1 MATLAB code for design of a FZP in finite conjugate mode.
%FZP finite conjugate mode
u¼0.01;%Define object distance
v¼0.02;%Define image distance
A¼V*exp(1i*((2*pi)/lambda)*(u-sqrt(u*u-sqrt(X.^2þY.
^2))));
B¼V*exp(1i*((2*pi)/lambda)*(-vþsqrt(v*v-(X.^2þY.^2);
D¼AþB;%Interference
Figure E.7.1 Image of the sinusoidal phase profile of a CGH that can focus a diverging
wavefront at a distance of 10 mm from a FZP plane to a point a distance of 20 mm from the
FZP plane.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2p
C1 ðx,yÞ ¼ exp j u u2 ðx2 þ y2 Þ , (E.7.1.a)
l
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2p
C2 ðx,yÞ ¼ exp j v2 ðx2 þ y2 Þ v : (E.7.1.b)
l
of Table 7.5; the wavefront profile of the axilens must be replaced by the
wavefront profile of a FZP. The image of the off-axis FZP with a tilt angle of
3 deg and focal length of 25 mm is shown in Fig. E.7.2.
Figure E.7.2 Image of the sinusoidal phase profile of a CGH that can focus a plane
wavefront a distance of 25 mm from the FZP plane with an angle of 3 deg.
Table E.7.3 MATLAB code for the design of an axicon with two base angles.
%Axicon with two base angles
N¼500;
Angle¼1;
V¼0.5;
lambda¼0.632;
for p¼1:N;%%Design of a conical wave with two angles and a plane
%reference wave
for q¼1:N;
r(p,q)¼sqrt((p-N/2)*(p-N/2)þ(q-N/2)*(q-N/2));
if r(p,q),125;
A(p,q)¼V*exp(1i*(r(p,q)/lambda)*tand(Angle*2)*2*pi);
else
A(p,q)¼V*exp(1i*(((125/lambda)*tand(Angle*2)*2*pi)þ
(((r(p,q)- 125)/lambda)*tand(Angle)*2*pi)));
end
B(p,q)¼V*exp(1i*0);
end
end
D¼AþB;%Interference
Figure E.7.3 Image of a binarized CGH generated by superposing a plane wave with a
wave generated by an axicon with two base angles.
E.7.4 An optical element that generates a ring pattern in its far field is an
axicon. If this axicon is used together with a FZP, then it generates a focused
ring pattern at the focal plane of the FZP. The divergence angle of the axicon
necessary to generate a ring pattern of radius 100 mm at the focal plane at a
distance of 5 mm is given by
r 1 1 100
b ¼ tan ¼ tan ≅ 1.15 deg :
f 5000
Table E.7.4 MATLAB code for design of a CGH for generation of a focused ring pattern.
%CGH with ring focus
N¼500;
Angle¼1.15;%Define divergence angle
V¼0.5;%%Visibility controller
lambda¼0.632*1e-6;%Define wavelength
f¼0.005;%Define wavelength
del¼1*1e-6;%
x¼-N/2:N/2-1;
y¼-N/2:N/2-1;
[X,Y]¼meshgrid(x*del,y*del);
R¼sqrt(X.^2þY.^2);
A¼V*exp(1i*((2*pi)/lambda)*(f-sqrt(f*f-R.*R)));
B¼V*exp(1i*(R/lambda)*tand(Angle)*2*pi);
D¼AþB;%Interference of the object and reference wave
Figure E.7.4 Image of a sinusoidal CGH that can focus light on a ring with a radius of
100 mm at a distance of 5 mm.
Table E.7.5 MATLAB code for design of a CGH for generation of a helical wavefront with
L ¼ 5 and focused at a distance of 5 mm.
%CGH for generation and focusing of a helical wavefront
N¼500;
L¼5;
V¼0.5;%%Visibility controller
lambda¼0.632*1e-6;%Define wavelength
f¼0.005;%Define wavelength
del¼1*1e-6;%
x¼-N/2:N/2-1;
y¼-N/2:N/2-1;
[X,Y]¼meshgrid(x*del,y*del);
R¼sqrt(X.^2þY.^2);
B¼V*exp(1i*((2*pi)/lambda)*(f-sqrt(f*f-R.*R)));
A¼V*exp(1i*L*(atan2(X,Y)));
D¼AþB;%Interference of the object and reference wave
Figure E.7.5 Image of a sinusoidal CGH that can focus a helical wavefront with L ¼ 5 at a
distance of 5 mm.
247
I P
infinite conjugate mode, 92 paraxial approximation, 76
interferograms, 143 phase grating, 29
inverse Fourier transform phase multiplexing, 119
algorithm, 60 phase-only DOEs, 122
iterative Fourier transform photolithography, 175
algorithm, 44 photon sieve, 187
pixel size, 27
L plasma etching, 186
Laguerre–Gaussian beam, 157 polygon approximation, 181
line focus, 39 polygonal path scanning, 181
lithography file design, 175 prism, 8
low- and high-pass filtering, 167
Q
M Quanta 3D FEG (FEI), 178
MATLAB®, 2, 17, 48, 76, 94, 127, Quanta™ 400F (FEI), 178
148, 178 quasi-achromatic, 135
maximum efficiency, 31
mesh generation technique, 63 R
mesh nodes, 63, 68, 70 RAITH150TWO, 178
metallization, 188 raster scanning, 182
micro-optical element ray tracing, 94
fabrication, 175 reflective–diffractive
modulo-2p phase addition combinations, 119
technique, 122 refractive–diffractive
multifunctional DOEs, 119, 134, combinations, 119
152, 203 resist thickness error, 115
multilevel DOEs, 46 resist-thickness-versus-dose
multilevel FZP, 52 characterization, 197
multilevel SPP, 55, 57, 206 ring pattern, 128
multilevel structure fabrication, 196
multiple-beam interference, 163 S
multiplexed diffractive sampling criteria, 27
optics, 119 sampling periods, 78
scalar diffraction, 24
N scalar theory, 2
negative axicon, 138 scanning electron microscope, 191
self-image, 85
O serpentine scanning, 183
off-axis axicon, 154 sinusoidal axicon, 152
off-axis axilens, 158 sinusoidal grating, 31, 149
off-axis Bessel beams, 154 Snell’s law, 3, 96
optical microscope, 187 spatial frequencies, 24