Al Amri2012 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 58

CHAPTER

8
Beyond the Rayleigh Limit
in Optical Lithography
Mohammad Al-Amria,b , Zeyang Liaob
and M. Suhail Zubairyb
a The National Center for Mathematics and Physics, KACST,
P.O. Box 6086, Riyadh 11442, Saudi Arabia
b Institute of Quantum Science and Engineering (IQSE) and
Department of Physics and Astronomy, Texas A&M University,
College Station, TX 77843, USA

Contents 1. Introduction 410


2. Classical Photolithography and
the Diffraction Limit 413
2.1 Mask-Based Photolithography 413
2.2 Classical Interferometric Lithography 415
3. Classical Multi-Photon Lithography 416
4. Quantum Interferometric Optical Lithography 418
4.1 Entanglement Helps to Break the
Diffraction Limit 418
4.2 A Proof-of-Principle Experiment for
Quantum Interferometric
Photolithography 421
5. Subwavelength Interferometric Lithography Via
Classical Light 423
5.1 Nonlinear Interferometric Optical
Lithography by Controlling the Phase 424
5.2 Subwavelength Lithography by Coherent
Control of Classical Light Pulses 426
5.3 Subwavelength Lithography Via Correlating
Wave Vector and Frequency 430
6. Resonant Subwavelength Lithography Via
Dark State 440
6.1 Three-Level  Type System 440
6.2 Generalization to 2× System 442
6.3 Generalization to N ×  System 443
6.4 Some Concerns 446
6.5 Experimental Demonstration of
This Scheme 447
7. Subwavelength Photolithography Via
Rabi Oscillations 450

Advances in Atomic, Molecular, and Optical Physics, Volume 61, Copyright © 2012 Elsevier Inc.
ISSN 1049-250X, http://dx.doi.org/10.1016/B978-0-12-396482-3.00002-8. All rights reserved.

409
410 Mohammad Al-Amri et al.

7.1 Achieving the Subwavelength Pattern 452


7.2 Arbitrary Subwavelength Patterns in
a Macroscopic Area 456
7.3 Potential Realizations 458
8. Summary and Outlook 460
Acknowledgments 461
References 462

Abstract It is well-known that traditional optical lithography is restricted


by the Rayleigh limit such that the smallest feature that can
be generated is restricted to half the wavelength of the light
source. Thus light beams with shorter and shorter wavelength
have been applied to print smaller and smaller circuit images.
However, when it comes to the extreme ultraviolet or X-ray,
severe problems can emerge. In the past 10–15 years, sev-
eral novel optical lithography schemes have been illustrated to
overcome the diffraction limit. In this article, we review these
schemes and explain their basic principles with possible exper-
imental realizations.

1. INTRODUCTION

Computer is one of the most important inventions in the last century.


The performance of the computer chips has increased dramatically over
the past few decades, and meanwhile the size of the integrated circuits
reduced roughly following the famous Moore’s law. Optical lithography
has been the most important driving force for these changes (Alfred, 2001;
Chris, 2007; Harry, 2005; Lu & Lipson, 2010; Mack, 2007; Seisyan, 2011). In
fact, almost any electronic equipment that uses processors or memory to
work, such as cellular phone, digital photo cameras or automobiles, is the
beneficiary of the optical lithography.
About 20 years ago, the smallest features printed with optical lithogra-
phy were twice the wavelength used to print them. Today the industry is
pressing toward the need for much smaller resolution. However, there is a
diffraction limit that restricts the smallest patterns we can print to about half
of the wavelength of light source (Abbe, 1873; Brueck et al., 1998; Rayleigh,
1879). Therefore, to make the chip smaller, we should switch to shorter
working wavelength. For example, we quote an interesting statement by
one of the experts (Rothschild, 2010):

Without some invention that significantly changes the way opti-


cal lithography is practiced, a next-generation lithography
technology—such as extreme ultraviolet lithography or
Beyond the Rayleigh Limit in Optical Lithography 411

electron projection lithography—will be required to extend the


roadmap to 45 nm node and beyond.

Nowadays, the working laser can operate in deep ultraviolet (DUV 190–
250 nm) (Chiu & Shaw, 1997; Marconi & Wachulak, 2010; Taylor
et al., 1998). Using the immersion lithography technology, the half-pitch
nodes currently obtained with 193 nm light are 45 nm and 32 nm (Ivan &
Scaiano, 2010). New tricks such as double exposure lithography (DEL) or
double patterning lithography (DPL) are possible to extend the resolution
by a factor of 2 (Lee et al., 2008). However, these technologies are not pos-
sible without the development of new material with nonlinear response
to the exposure dose. While we switch to shorter wavelength, two major
problem arise (Williams et al., 2006): First, the traditional lens and the air
absorb the light significantly. We need to invent new materials for the lens
which is hard to come, and the system should work in a vacuum system
that can be very expensive. Second, the bandgap of SiO2 is about 9 eV.
When the wavelength of the light is close to or smaller than 138 nm, it will
cause adverse charging in the SiO2 layer and destroy the substrate. This
motivates us to go beyond the Rayleigh limit and study ways to overcome
the diffraction limit.
The diffraction limit not only affects the photolithography, but also plays
an important role in the imaging system. The resolution of a far-field opti-
cal microscopy is also limited by the diffraction limit. In the past few
decades, a number of schemes have been developed to improve the reso-
lution of the microscopy. To get a better resolution, people have to switch
to shorter wavelength (e.g., electron microscope and X-ray lithography,
Rudenberg & Rudenberg, 2010; Spille & Feder, 1977; Vieu et al., 2000) which
is usually invasive to the system. While confocal microscopy introduced
optical sectioning and can get a better resolution than the conventional
ones, it did not overcome the diffraction limit (Diaspro, 2010). Near-field
scanning microscopy can obtain optical imaging with sub-diffraction res-
olution (Alkaisi et al., 2001; Betzig & Trautman, 1992; Binnig & Quate,
1986; Dryakhlushin et al., 2005; Ono & Esashi, 1998 ). These techniques
are however surface bound and are thus limited in terms of applications.
Two-photon fluorescence microscopy was first developed to achieve a
higher resolution than classical one-photon fluorescence microscopy (Denk
et al., 1990; Denk & Svoboda, 1997; Hell, 1994; Helmchen & Denk, 2005;
Strickler & Webb, 1991a, 1991b). Stimulated-emission-depletion fluores-
cence microscopy (STEP) was then developed by Hell and Wichmann
(1994) and the related concepts such as ground-state depletion (GSD) are
then developed (Hell, 2007; Hell & Kroug, 1995). A number of experiments
have also been carried out (Donnert et al., 2006; Klar & Hell, 1999; Klar et al.,
2000; Rittweger et al., 2009). Some other techniques such as Spatially Struc-
tured Illumination Microscopy (SSIM) (Gustafsson, 2005), Photoactivated
412 Mohammad Al-Amri et al.

Localization Microscopy (PALM) (Betzig et al., 2006), Stochastic Optical


Reconstruction Microscopy (STORM) (Zhuang, 2009) are also invented to
achieve super resolution. Dark state (Agarwal & Kapale, 2006; Gorshkov
et al., 2008; Kiffner et al., 2008; Yavuz & Proite, 2007) and resonance fluores-
cence (Chang et al., 2006; Chang et al., 2006; Macovei et al., 2007; Mompart
et al., 2009; Sun et al., 2011; Qamar et al., 2000) are also employed to localize
the atoms with a subwavelength resolution.
In photolithography, several schemes have been proposed to break the
diffraction limit in the past two decades. In 1992, Wu et al. pointed out that
two-photon excitation in laser scanning photolithography can allow expo-
sure of patterns not possible with conventional one-photon direct writing
(Wu et al., 1992). Unfortunately, this scheme is based on point-by-point
scanning which has limited applications. In fact the ordinary two-photon
absorption can only achieve a sharper peak but not improve the spatial
resolution. In 1999, Yablonovitch and Vrijen illustrated how to suppress
the normal resolution term and get a super-resolution image based on
two-photon absorption (Yablonovitch & Vrijen, 1999). The visibility of this
scheme is reduced due to a constant background. In 2000, Boto et al. showed
that quantum entanglement can successfully eliminate the normal resolu-
tion term without a constant background (Boto et al., 2000). After that sev-
eral papers showed that quantum entanglement is not necessary to break
the diffraction limit but the nonlinear response of the recording medium
(Bentley & Boyd, 2004; Hemmer et al., 2006; Peér et al., 2004). In 2008,
Kiffner et al. came out with a novel idea that subwavelength resolution
can be achieved by preparing the system in a position dependent trapping
state (Kiffner et al., 2008). In 2010, Liao et al. discovered that coherent Rabi
oscillations can lead to subwavelength resolution (Liao et al., 2010). This
method does not require quantum entanglement or multi-photon absorber
but just quantum coherence of the medium. In addition to these schemes,
there are some near-field subwavelength lithography methods such as pho-
tolithography based on surface plasmon (Liu et al., 2005; Luo & Ishihara,
2004a, 2004b; Martin, 2003; Schuller et al., 2010; Srituravanich et al., 2004;
Xie et al., 2011), which will not be covered in this article.
In this article, we mainly focus on optical lithography. In Section 2,
we begin by introducing traditional photolithography, the interferometric
photolithography and the diffraction limit. In Section 3, we introduce the
classical multi-photon lithography. In Section 4, we show how to achieve
subwavelength resolution via quantum entanglement. In Section 5, we will
show quantum entanglement can be mimicked by carefully manipulating
the classical light. In Section 6, we discuss a novel idea to break the diffrac-
tion limit based on dark state. In Section 7, we illustrate our novel and
simple method based on coherent Rabi oscillations. Finally we present the
summary and the outlook.
Beyond the Rayleigh Limit in Optical Lithography 413

2. CLASSICAL PHOTOLITHOGRAPHY AND


THE DIFFRACTION LIMIT

2.1 Mask-Based Photolithography


Mask-based photolithography is commonly used to print the circuit image
in the industry nowadays (Brueck, 2005; Sheats & Smith, 1998). The typical
setup is shown in Figure 1a. The light projects the image of the pattern
on the mask onto the photoresist. Some places are bright while others
are dark in the photoresist. The photoresist changes its solubility at the
place where it is shined by the light. There are two types of photoresist:
positive and negative (see Figure 1b). In a positive-tone photoresist, areas
of the material that are exposed to light are removed after development.
While in a negative-tone photoresist, the areas exposed by the light remain
behind. After development, the pattern is printed onto the photoresist.
The SiO2 layer is then etched at the place without the protection of the
photoresist.
The minimum feature size that a projection system can print is approx-
imately given by the Rayleigh criterion (Abbe, 1873; Brueck et al. 1998;
Rayleigh, 1879):
λ
CD = 0.61 , (1)
NA
where λ is the wavelength of the light used and NA is the numerical aper-
ture of the lens as seen from the wafer (Figure 2). In most areas of optics,
the numerical aperture is defined by NA = n sin θ where n is the index of

a b
Light from laser source
Photolithographic Process

Photoresist
Condenser lens SiO2 Coating
Si Substrate

Mask or Reticle
Exposure

Negative Positive
Projection lens

Transfer
Wafer TAMU & KACST
Strip

Figure 1 (a) Mask-based photolithography. (b) Positive and negative photoresist.


414 Mohammad Al-Amri et al.

Object Plane

Figure 2 Numerical aperture of a lens.

refraction of the medium in which the lens is working (1.0 for air, 1.33 for
pure water, and up to 1.56 for oils), and θ is the half-angle of the maximum
cone of light that can enter or exit the lens. Equation (1) is also called the
diffraction limit, i.e., for a working wavelength and the numerical aperture,
this is the resolution limit.
The diffraction limit can be explained by the loss of high spatial frequen-
cies due to the evanescent wave (Alkaisi et al., 1999; Brueck, 2005; Neice,
2010). According to the Fourier optics, the electric field on the imaging
plane is the summation of varies frequencies components emitted from
the object plane:

 
ε(x, y, z) = εσ (kx , ky )eikx x+iky y+ikz z dkx dky , (2)
σ kx ky

where σ is the polarization, z is the propagation direction, and kx2 +ky2 +kz2 =
n2 ω2 /c2 (n is the refractive index of the medium, ω is the angular frequency
of the light, and c is the speed of the light). For the high spatial frequency
such that kx2 + ky2 > n2 ω2 /c2 , kz is pure imaginary which means that this
component decays in the propagation direction. This corresponds to the
evanescent wave and such waves can not reach the imaging plane in the
far field.The highest spatial frequency that can reach the imaging plane
is k = kx2 + ky2 = nω/c, which corresponds to a resolution of 2π/k =
2πc/nω (Pendry, 2000). This corresponds to the maximal resolution for
the field to be equal to λ/n. The corresponding maximal resolution for
the intensity is therefore equal to λ/2n. Considering the aperture of the
lens (see Figure 2), the maximal transverse wave vector that can reach the
image plane is k = k sin θ = nω sin θ/c, and thus the maximum resolution
is λ/2n sin θ which is the well known diffraction limit.
Beyond the Rayleigh Limit in Optical Lithography 415

2.2 Classical Interferometric Lithography


Interferometric lithography is a new kind of mask-less lithography, which
uses conventional laser light (Brueck et al., 1998; Menon et al., 2005a, 2005b;
Pau et al., 2001). In interferometric lithography, two coherent plane waves
of laser radiation converge from the apposite directions and hit a surface
with θ being the angle between the direction of propagation and the sub-
strate surface (see Figure 3). The wave vectors of the two light beams are:
k1 = k( − cos θ, − sin θ), k2 = k(cos θ, − sin θ) where k = 2nπ/λ0 with n
being the refractive index of the medium and λ0 being the wavelength of
the light in the vacuum. The two coherent waves form a standing wave in
the substrate plane. Consequently, the fringes pattern on the substrate is
determined by the in-plane wave vectors k1 and k2 . The intensity distribu-
tion at the focal plane and along the x direction is:
 2
 
I(x) ∝ expik1 ·x + expik2 ·x  = 2(1 + cos 2kx cos θ). (3)

According to the criterion by Rayleigh (1879) the minimal resolvable fea-


ture size x occurs at a spacing corresponding to the distance between two
intensity maxima. From Equation (3), we obtain 2k x cos θ = 2π, which
leads to the following formula for the maximum resolution:
λ0
x = , (4)
2n cos θ
thus the resolution is wavelength dependent and has its maximum value
when x is minimum, and that happens when cos θ = 1. This leads to the
classical diffraction limit x = λ0 /2n, where we consider the grazing limit
θ = 0.

k2 k1 λ
λ

θ θ

Substrate

λ
2 cos θ

Figure 3 The scheme of interferometric lithography with classical light. Two


counter-propagating plane waves interfere on a photosensitive substrate.
416 Mohammad Al-Amri et al.

Figure 4 Spatial distribution of excitation near the focus of a diffraction limited laser
beam for one-photon absorption (a) and two-photon absorption (b). Figure reprinted
with permission from Wu et al. (1992). Copyright 1992 by SPIE.

It is important to note that, in this analysis above, we have made no


mention of the atomic structure of the photoresist. It is implicitly assumed
that the atomic system is initially in the ground state and the ionizing rate
is proportional to the intensity of the light incident on the photoresist. Thus
the atomic response of the photoresist is considered to be linear.

3. CLASSICAL MULTI-PHOTON LITHOGRAPHY

The concept of two-photon excitation was first described by Goeppert–


Mayer (1931) in her doctoral dissertation, and first observed in cesium
vapor using laser excitation by Abella (1962). Two-photon excitation is
found to be able to improve the resolution in the fluorescence microscopy
(Denk et al., 1990; Strickler & Webb, 1991b) and increase the data capacity
of the storage (Strickler & Webb, 1991a). Such a process depends quadrat-
ically on the photon intensity I of the incident beam (Goeppert-Mayer,
1931; McClain, 1974), i.e., the excitation rate, or the number of photon
being absorbed per unit time, W = δI 2 , where δ is the absorption cross
section, typically of the order of 10−58 m4 s/photon. The quadratic depen-
dence of the intensity improves the resolution. In Figure 4, we show the one-
photon and two-photon excitation rate of a Gaussian shape pulse. From the
figure, it is clear that the two-photon process provides a sharper feature.
Similar idea is proposed to improve the resolution of photolithography
(Kawata et al., 2001; Wu et al., 1992). By scanning the focal volume in a pro-
grammed 3D pattern through a thick positive photoresist, it is possible to
produce patterns with high aspect ratio trenches and multilayered under-
cut. However, the point-by-point scanning is time-consuming which limits
its applications.
Beyond the Rayleigh Limit in Optical Lithography 417

Figure 5 (a) The ordinary intensity fringe pattern, proportional to (1 + cos 2kx ),
produced by converging rays as in Figure 3. (b) The intensity squared fringe pattern
(3/2 + 2 cos 2kx + 1/2 cos 4kx ), which consists of a normal-resolution spatial harmonic
at 2k , a super-resolution spatial harmonic at 4k , and a constant term. (c) Here
(3/2 + 1/2 cos 4kx ), where the normal-resolution spatial harmonic at 2k was canceled.
The super-resolution component at 4k remains, on a constant background. (d) By
employing a photoresist with a sharp atomic transition the constant background is
eliminated, leaving a pure super-resolution image. Figure reprinted with permission
from Yablonovitch and Vrijen (1999). Copyright 1999 by SPIE.

Can we generate a super-resolution image by two-photon absorption in


one step? Let us look at the two-photon process of a standing field. The
excitation rate is proportional to

3
(1 + cos 2κx)2 = + 2 cos 2κx + cos 4κx, (5)
2
where κ = k cos θ. Comparing the fringe pattern with the one-photon
process, we see that the pattern of two-photon absorption is a mixture
of a normal-resolution image represented by cos 2κx term and a super-
resolution image represented by cos 4κx. Indeed, cos 4κx, represents a dou-
bling of the spatial resolution over the one-photon process. The cos 2κx
term destroys this super resolution. From Figure 5a and b, we see that the
two-photon process does have a sharper peak, but the period of the struc-
ture is the same as that for one-photon process. The resolution of the whole
pattern does not improve!
418 Mohammad Al-Amri et al.

The second term in Equation (5) comes from the absorption of photons
from different paths, i.e., one photon from the left and other photon from
the right. If we can eliminate this term, we can obtain a super-resolution
image. Yablonovitch and Vrijen (1999) showed that the normal resolution
term can be entirely suppressed, using a classical frequency modulation
scheme, where simultaneous absorption of a pair of photons is accompa-
nied by a twofold spatial-resolution enhancement. Their scheme is shown
in Figure 6. The incident rays on one edge of the lens have frequency ω0 ,
while rays on the other edge consist of two frequencies (i.e., ω1 = ω0 + δ,
ω2 = ω0 − δ). Fringes resulting from the interference of rays from opposite
edge oscillate rapidly at the different frequency δ and the normal-resolution
image is washed away, forming a constant background. Provided that the
frequencies 2ω0 and ω1 + ω2 are coherently related, the super-resolution
image is stationary. The resulting fringes are indicated in Figure 5c.
The constant background can be entirely eliminated in principle. If the
atomic transition at 2ω0 is sufficiently sharp, the background two-photon
transitions of the frequency combinations ω0 + ω1 , ω0 + ω2 , 2ω1 , and 2ω2
do not occur. In this case, the background should vanish and leave only
the double frequency component (Figure 5d).

4. QUANTUM INTERFEROMETRIC OPTICAL LITHOGRAPHY

4.1 Entanglement Helps to Break the Diffraction Limit


Quantum entanglement can effectively eliminate the normal resolution
term in Equation (5) and achieve sub-diffraction limited pattern (Boto
et al., 2000; Lee & Lee, 2008; Williams et al., 2006). The system is described
schematically in Figure 7, where two photon beams are incident on a sym-
metric, loss-less beam splitter BS at ports A and B. The output beams
get reflected off by a mirror pair into the substrate. The two beams get
converged on the imaging plane. The photoresist consists of two-photon
absorber.
Now, let us first look at what happens when we start with the input
state | 11 = |1A 1B , i.e., one photon from the upper arm and one photon
from the lower arm, see Figure 7. Interference effect upon passage through
a symmetric, loss-less beam splitter can cause the product state √ |1A 1B  to
become the quantum entanglement state (|2C 0D  + |0C 2D )/ 2. Hence,
after the beam splitter, the two photons emerge either both in the upper
arm C or both in the lower arm D, but never one photon in each arm.
Absorption of one photon from upper path and one photon from lower
path never happens. The normal resolution term is therefore completely
washed away and leaves only the super-resolution term. This is the basic
principle of this scheme.
Beyond the Rayleigh Limit in Optical Lithography 419

ω0+ ω0 ω1+ ω2

Lens

2ω0 ω + ω2
1
θ θ

Fringe Pattern

Figure 6 The fringe pattern produced by two-photon excitation of a photoresist, in


which the incident rays on opposite sides of the lens are separated into distinct
frequency grouping. Figure reprinted with permission from Yablonovitch and Vrijen
(1999). Copyright 1999 by SPIE.

Two photons are incident on the photoresist in such a way that both of
them are either in the upper path or in the lower path. The deposition rate
is then proportional to
 2
 2ikx 
e + e−2ikx  = 2(1 + cos 4kx). (6)
420 Mohammad Al-Amri et al.

Mirror

Substrate

Α C

BS
D
Β
θ

Mirror

Figure 7 Interferometric lithography setup where two photon beams hitting a beam
splitter at ports A and B, and then get reflected by two mirrors. The two photon
beams get interfere on the substrate.

We can see clearly that the slowly oscillating term cos (2kx) has been com-
pletely eliminated and we are left with only cos (4kx) terms which gives the
resolution of λ/4, half of the usual diffraction limit. The essential physics is
simple: The two photons incident on a two-photon absorbing substrate cor-
responds to an effective doubling of the frequency or reducing the effective
wavelength by a factor of 2, thus giving rise to a λ/4 resolution.
This result can be generalized by considering the entangled state at ports
C and D in Figure 7 to be in the so-called NOON state, i.e., | (N) =
√1 ( |NC |0D + |0C |ND ). The deposition rate on an N-photon absorbing
2
substrate is then proportional to N = 1+cos 2Nkx, with resolution λ0 /2N.
The generation of NOON state is not simple. This is a maximally entan-
gled photon number state and it was suggested (Boto et al., 2000) that this
state can be created using optical components such as parametric down
converters, particularly for the case when N = 2. The schemes for the gen-
eration of NOON states with higher values of N in a deterministic manner
remains a challenge. This proposed method for subwavelength lithogra-
phy was generalized further to arbitrary patterns in one and two dimen-
sions (Kok et al., 2001). This requires arbitrary entangled Fock states of the
form | (N, M) = √1 (|NC MD  + |MC ND ). The application of these ideas
2
to quantum imaging lithography is discussed by Shih (2007).
Beyond the Rayleigh Limit in Optical Lithography 421

Besides the obvious difficulty of generating higher order NOON states,


this method for subwavelength lithography suffers from some other seri-
ous problems as well (Agarwal et al., 2001). On the one hand, one needs to
produce weak light field to contain only two photons per mode, and at the
same time, use this particular field to excite two-photon absorption which
requires a strong field. When one photon is localized, the momentum of
the other photon becomes completely delocalized, and thus this photon
can end up anywhere (Tsang, 2007). Hence, the general usefulness of this
method is quite limited.
The limitations of this N-photon entanglement technique are summa-
rized in the following quote by Anisimov and Dowling (2009):
Nearly ten years ago, a quantum optical approach to imaging
quantum imaging, which uses path-entangled states was sug-
gested as a solution for this requirement of ever-shorter wave-
lengths and a way to beat the classical limit (Abella, 1962). This
turned out to be easier said than done. While quantum lithog-
raphy is viewed as one of the killer apps for the nascent field of
quantum imaging, the bugaboo in its implementation has been
the continued lack of the right kind of multi-photon photore-
sists that would operate at the low flux levels required for a real
proof-of-principle experiment.

4.2 A Proof-of-Principle Experiment for Quantum Interferometric


Photolithography
A proof-of-principle experiment was reported (D’Angelo et al., 2001) uti-
lizing the entangled photon pairs in a Young double-slit experiment. A
similar experiment in the context of measuring the de Broglie wavelength
of two-photon wave packets in a Young double-slit experiment was also
reported by Fonseca et al. (1999).
The schematic setup of the actual experiment is illustrated in Figure 8.
However, in order to explain the essential physics of the subwavelength
lithography scheme based on entangled light we first look at the setup
shown in Figure 9. Here the two slits are placed symmetrically on the left
and right sides of the entangled photon source. Region V at the center of
Figure 9a is the place where an entangled photon pair can be generated such
that the photons of the same pair propagate horizontally and in opposite
direction. They are indicated by straight lines as seen in the figure. There
is one photon counting detector on each side. These two detectors scan
symmetrically in x direction for the arrival of the photon pair, and regis-
ter the coincident “clicks." The joint detection counting rate is sin c2 (2β)
which has a narrower feature than the classical pattern by a factor of 2
(here β = πa sin θ/λ, a is the width of the slit and θ is the scattering angle)
(Pittman et al., 1995; Strekalov et al., 1995). This proof-of-principle
422 Mohammad Al-Amri et al.

Figure 8 Schematic of the experimental setup. The 458 nm line of an argon ion laser
is used to pump a 5 mm BBO crystal that produce pairs of orthogonally polarized
signal and idler photons, which emerges collinearly. The pump is separated from the
signal–idler pair by a mirror M that reflects the pump while transmits the signal–idler
pair. A cutoff filter F is used to suppress the pump further. The signal–idler beam
passes through a double slit, which is placed close to the output side of the crystal,
and is reflected by two mirrors, M1 and M2 , onto a pinhole P followed by a polarizing
beam splitter PBS. The signal and idler photons are separated by PBS and are
detected by the photon counting detectors D1 and D2 , respectively. The output pulses
of each detector are sent to a coincidence counting circuit for the signal–idler joint
detection. Figure reprinted with permission from D’Angelo et al. (2001). Copyright
2001 by the American Physical Society.

experimental setup can indicate that entanglement can break the one-slit
diffraction limit.
If we replace the single slit in the above setup with a double slit, see
Figure 9b, we can observe the two-photon interference corresponding to
the case when N = 2. The entangled photon pairs will either pass through
the upper two slits or through the lower two slits. In this situation, we
can find that the double-slit two-photon spatial interference pattern has
a higher modulation frequency than the classical double-slit interference
pattern.
In the actual setup shown in Figure 8, the two-photon state N = 2
is generated via spontaneous parametric down conversion (SPDC). In the
experiment, the double slit must be placed as close as possible to the output
surface of the BBO crystal. In this case, the two entangled photons travel
through the same slit and we can erase the first-order interference. Figure 10
shows the experimental results, where Figure 10a shows the distribution
of coincident detection events versus the rotation angle θ of mirror M1 . In
Figure 10b we present the first-order interference-diffraction pattern of a
classical light by the same double slit in the same experimental setup. When
comparing the two-photon interference-diffraction pattern result with that
of the first-order interference-diffraction pattern of a classical light, one can
Beyond the Rayleigh Limit in Optical Lithography 423

Figure 9 Schematic of a two-photon diffraction-interference thought experiment.


Detectors D1 , D2 perform the joint coincidence detection for entangled photon pair
“signal and idler” which are represented by the right and left sides of the sketch.
Scheme with a single slit in (a) and with a double slit in (b). Figure reprinted with
permission from D’Angelo et al. (2001). Copyright 2001 by the American Physical
Society.

see that the two-photon interference-diffraction pattern has a higher spatial


interference modulation period and a narrower diffraction pattern width.

5. SUBWAVELENGTH INTERFEROMETRIC LITHOGRAPHY VIA


CLASSICAL LIGHT

The practical difficulties of the experiment presented in the previous


section have stimulated several elegant high resolution schemes without
entanglement. In these schemes the resolution enhancement can result
solely from the properties of the N-photon absorption process if we can
somehow eliminate the low spatial frequency components while retaining
the high frequency components (Bentley & Boyd, 2004; Boyd & Bentley,
2006a, 2006b; Chang et al., 2006; Chang et al., 2006; Hammer et al., 2006;
Kawabe et al., 2007; Peér et al., 2004). In the following subsections, we
introduce these subwavelength lithography schemes.
424 Mohammad Al-Amri et al.

Figure 10 (a) Experimental measurement of the coincidences for the two-


photon double-slit interference-diffraction pattern. (b) Measurement of the
interference-diffraction pattern for classical light in the same experimental setup.
Figure reprinted with permission from D’Angelo et al. (2001). Copyright 2001 by the
American Physical Society.

5.1 Nonlinear Interferometric Optical Lithography by Controlling the


Phase
Bentley and Boyd (2004) pointed out that super-resolution of quantum
lithography results primarily from the nonlinear response of the
recording medium and not from quantum features of the light field.
The basic principle of their proposal is shown in Figure 11. The key
feature of this setup is that one component of the light beam is shifted
by a phase and this phase is incremented by a fixed amount for the suc-
cessive laser pulse. The successive mth laser pulse will have relative phase
given by

φm = 2πm/M, (7)
Beyond the Rayleigh Limit in Optical Lithography 425

Mirror

Δφm
N−photon
absorber

BS

Mirror

Figure 11 Sketch of the used technique, where intense laser pulse is separated into
two equal beams at a beam splitter BS. One of them gets phase shifted, while the
other is not. The two beams are brought together on a recording medium that
functions by means of N-photon absorption. Figure reprinted with permission from
Bentley and Boyd (2004). Copyright 2004 by the Optical Society of America.

where M is the total number of the pulse. The deposition rate of the N-
photon absorber is

M
I(N, M) = (Em E∗m )N , (8)
m=1

where
Em = eikx cos θ + e−ikx cos θ ei φm . (9)

If the phase shifts φm were not introduced, the resulting deposition rate
is simply I(N, M) = [1 + cos (2kx cos θ)]N which generate the same spa-
tial period as one-photon interference pattern, but with a sharper fea-
tures. However, if the phase is introduced as appearing in Equation (9),
the slowly varying terms can be averaged out, leaving only a spatial com-
ponent cos (2Mkx sin θ), and possibly harmonics of this component if N
is at least twice as large as M. Therefore, the pattern generated in this
way can be M times better than that allowed by the normal interfero-
metric lithography. To achieve the optimum minimum feature size, we
can take θ as 0◦ , which is the situation of two beams striking the record-
ing plate near the grazing incidence. The visibility of this pattern is given
426 Mohammad Al-Amri et al.

by 
AN,M + AMHo
V=  , (10)
AN,0 + AMHe
where
2N! 2N!
AN,M = (M = 0), AN,0 = , (11)
(N − M)!(N + M)! 2(N!)2
where AN,M is the desired component, AN,0 the dc component of the depo-
sition pattern, AMHo the odd harmonics, and AMHe the even harmonics of
the desired frequency.
A proof-of-principle experiment was conducted to demonstrate the res-
olution enhancement in this scheme (Figure 12). In this experiment, the
properties of an N-photon absorber are simulated by an Nth-harmonic gen-
erator followed by a CCD camera. This meant that the desired harmonic
is recorded by the CCD and any other light was spectrally filtered out.
The repeated M laser shots data is collected and summed by a computer.
Another technical point is that the prism is mounted on a micrometer-
controlled translation stage, which is used as the phase shifter.
When N = M = 1, we obtain the usual interference pattern that would
be recorded on a linear absorber (see Figure 13a). As seen in Figure 13b
and c, the fringes become narrow but the spacing between them remains
constant as N is increased. When M  2, the resolution is enhanced. This
is seen in Figure 13d and e where the resolution is enhanced by a factor
of 2 when M = 2, while the resolution enhancement is three times when
M = 3. Moreover, when N = M = 3, the data shows a fringe spacing that
is one-third of that for N = 1. The fringe visibility is reduced as predicted
by the theory.
The conclusion is that this technique that uses classical light can lead to
an improvement in the resolution. The procedure is quite straightforward
to implement. However, the primary drawback of this technique is the
lack of suitable N-photon absorbing media, especially for large N. Also, for
increasing M, each individual phase shift becomes smaller, which requires
greater phase resolution.

5.2 Subwavelength Lithography by Coherent Control of Classical


Light Pulses
In another subwavelength resolution scheme Peér et al. (2004) showed
that super resolution can be achieved by coherent control of the classical
light pulses without entanglement. The basic idea is as follows: If, in a
N-photon absorbing medium, the excitation lifetime is much longer than
the pulse duration, the two excitations induced by two pulses can still inter-
fere to each other even if they do not overlap in time. Moreover, because
two excitation pulses do not overlap in time, the cross absorption terms
Beyond the Rayleigh Limit in Optical Lithography 427

Figure 12 The experimental setup. Here, the output of a picosecond Nd:YAG laser is
directed onto a thin plate beam splitter. The transmitted component propagates to a
right-angle prism, where it is translated and reflected back to the beam splitter, while
the reflected component from the beam splitter propagates to a plane mirror and is
reflected back at an angle such that it will overlap the translated component in the
detection plane. Figure reprinted with permission from Bentley and Boyd (2004).
Copyright 2004 by the Optical Society of America.

are eliminated. For two exciting pulses (E1 and E2 ), the intensity of the
excitation is

 2
 
I(x) ∝ EN N
1 (x, ωA ) + E2 (x, ωA ) , (12)

where x is the spatial coordinate. We note that the mixed terms such as
p N−p
E1 E2 are absent and the resolution is enhanced by a factor N.
To achieve a desired narrow lithography spot, we should shape the
spatial phase fronts of the excitation pulses at the focus such that they
interfere constructively in the center of the spot and destructively near the
428 Mohammad Al-Amri et al.

Figure 13 Measured intensity distributions for (a) M = N = 1, (b) M = 1, N = 2,


(c) M = 1, N = 3, (d) M = N = 2, (e) M = 2, N = 3, and (f) M = N = 3. Note that the
first three patterns have the same period (because M = 1) but that the fringes become
sharper with increasing N. Note also the doubling of the fundamental frequency in (d)
and (e) and the tripling of the frequency in (f). Figure reprinted with permission from
Bentley and Boyd (2004). Copyright 2004 by the Optical Society of America.

edges. For example, we should search for M pulse fields such that
M 2
 
 N 
I(xf ) =  Ek (xf ) , (13)
 
k=1

where I(xf ) is the desired spot.


Figure 14 is a schematic setup where a glass plate delays a segment of
the pulse with respect to the other. If the delay between the pulses is tuned
correctly, it will lead to the desired constructive interference at the focal
spot while destructive interference at the edge. A Fourier equivalent of
the segmentation scheme was suggested for doubling the resolution with
two-photon absorption (Korobkin & Yablonovitch, 2002; Yablonovitch &
Vrijen, 1999). While this scheme works for two-photon case, generalization
to the N-photon case is not straightforward and its implementation is more
complicated.
An experiment is conducted to prove the resolution enhancement of this
scheme. The experimental setup is presented in Figure 15 which utilize the
technique in Figure 14. The pulses (100 fs around 778 nm) are emitted from
a Ti:Sapphire laser. The cylindrical telescope weakly focuses the laser into
the Rb cell. The delay line introduces a relative delay between the two
halves of the pulse. The cylindrical lens in front of the cell is used to tightly
focuses the beam in the perpendicular dimension in order to increase the
Beyond the Rayleigh Limit in Optical Lithography 429

Figure 14 Schematic setup for generating quantum interference in lithographic


medium. A glass plate delays half of a planar pulse with respect to the other half. As a
result, the nonlinear medium at the focus is excited by two instant pulses. The delay
between the pulses can be fine tuned by a small tilt of the glass, in order to generate
two-photon interference. Figure reprinted with permission from Bentley and Boyd
(2004). Copyright 2004 by the Optical Society of America.

Figure 15 Experimental configuration and relevant level diagram for atomic Rb.
Figure reprinted with permission from Peér et al. (2004). Copyright 2004 by the Optical
Society of America.

signal. The right part of Figure 15 is the energy level diagram of Rb in


which the 5S–5D two-photon transition centered at 778 nm was used in
this experiment. To avoid the one-photon excitation of the intermediate 5P
level (at about 780 nm), a pulse shaper is placed in front of the cylindrical
telescope to block the resonant frequency at 780 nm. Moreover, a π phase
shift to frequencies above and below the resonance can be induced by the
pulse shaper to maximize the two-photon excitation (Dudovich et al., 2001).
Finally, the two-photon excitation can be detected by imaging the resulting
fluorescence at 420 nm onto an enhanced CCD camera.
In Figure 16, results are shown where double resolution is observed.
Two CCD images of dark spot (destructive interference in the center of
the spot) and the corresponding transverse line cross sections are shown
430 Mohammad Al-Amri et al.

Figure 16 Experimental results. (a) Images and transverse cross sections of “dark
spots” (destructive at the center) for a short relative delay (crosses: data, gray line:
theoretical fit) and a long relative delay (circles: data, line: theoretical fit),
demonstrating the double resolution of two-photon interference compared to
one-photon interference. (b) is the corresponding two-photon “bright spot” as
compared to the diffraction limited one-photon spot (dashed). Figure reprinted with
permission from Peér et al. (2004). Copyright 2004 by the Optical Society of America.

in Figure 16a. Here, two distinct cases should be noticed. First, a regular
one-photon interference is observed when the delay was tuned shorter
than the coherence length of the pulse. Second, two-photon interference is
observed when the delay was tuned far beyond the coherence length. We
can see that the dark spot in the second case is about half of that of first
case. In Figure 16b “bright spot" (constructive at the center) is observed in
the two-photon interference regime. Here we also see that the central lobe
is about twofold narrower than the one-photon case.
This experiment verified a scheme for sub-diffraction limit that relies
on the quantum nature of the lithographic material and not of the excit-
ing field. In order for this method to be practical, a nonlinear lithographic
material with a narrow excitation line is required. However, similar to pre-
vious scheme, it suffers from the low efficiency of N-photon absorption and
the high intensity requirement, especially for large N. This is a significant
obstacle in making use of this scheme in real lithography.

5.3 Subwavelength Lithography Via Correlating Wave Vector and


Frequency
Hemmer et al. (2006) showed that the diffraction limit can be broken by
correlating wave vector and frequency in a narrow band, multi-photon
detection process that uses Doppleron-type resonances. The motivation
for this work comes from two points of view. First, the early theoretical
Beyond the Rayleigh Limit in Optical Lithography 431

suggestion (Berman & Ziegler, 1977; Haroche & Hartmann, 1972; Kyröla &
Stenholm, 1977), and experimental observation (Freund et al., 1975; Reid &
Oka, 1977), of directional multi-photon resonances, called “Doppleron,” in
saturated absorption spectroscopy. Second, the recently initiated research
(Herkommer et al., 1997; Qamar et al., 2000; Zubairy et al.„ 2002), where
an atom (or molecule) can be localized to subwavelength precision based
on the conditional detection of fluorescence photons as the atom passes
through a standing-wave field.
The basic idea of this scheme is shown in Figure 17. Two counter-
propagating plane waves consisting of signal frequencies ν± interfere on a
photosensitive substrate. The drive fields ω± assist a directional resonance
for pairs of signal photons, i.e.,

ωab = 2ν± − ω± . (14)

If the detection bandwidth is narrow, the ± channels will realize distinct


resonances (Mollow, 1968). The atoms will absorb two photons from the left
beam or from the right beam, but never one photon from each beam, which
is similar to the path-number entanglement in quantum field lithography.
As a consequence, the one-photon interference term will be suppressed and
keep only the pure two-photon interference term which has a resolution
half of the diffraction limit.

5.3.1 Illustrative Calculation for the Case N = 2


Doppleron-type resonances can be observed in a two-level system, pro-
vided that the one-photon detunings and field strengths dominate the
linewidths in a saturated absorption process (Pritchard & Gould, 1985).

ω−

ω
ν ν−
λ
θ θ

Δ1+ Δ1−
ν ν−
ω λ/ ω−
ν Δ2+ θ
ν− Δ2−

Figure 17 Subwavelength interference with classical light. Two counter-propagating


plane waves consisting of signal frequencies ν± interfere on a photosensitive
substrate. The drive fields ω± assist a directional resonance for pairs of signal
photons. Figure reprinted with permission from Hemmer et al. (2006). Copyright 2006
by the American Physical Society.
432 Mohammad Al-Amri et al.

The theory of photoelectron counting has been developed in the semiclas-


sical (Mandel et al., 1964) and quantum field (Scully & Lamb, 1969) regimes.
In Figure 17, two signal frequencies and two drive frequencies are used
to complete three-photon resonance for each direction. The intermediate
levels cj are off-resonant by detunings 1± = ωc1 b −ν± and 2± = ωac2 −ν± .
Assuming that the signal Rabi frequency S is the same for the two tran-
sitions, the interaction Hamiltonian in the rotating-wave approximation
(S , D ν± , ω0 ) is given by
 
HI = S |c1 
b|ei 1± t + |a
c2 |ei 2± t + h.c.
 
+ D |c1 
c2 |ei( 1± + 2± )t + h.c. , (15)

where the second term was written make use of Equation (14). Next we
derive the Schrödinger equations for the state amplitudes. For large one-
photon detunings j± S , D , the intermediate levels cj can be adia-
batically eliminated by setting the time derivatives of the slowly varying
amplitudes, c̃j = cj exp ( − i j± t), to zero. This furnishes an effective cou-
pling between levels a and b:

2S 2 D
iȧ − a=− S b, (16)
2± 1± 2±
and similarly with a ↔ b and 1 ↔ 2. Apart from dispersive phase shifts,
the effective coupling is thus described by a three-photon Rabi frequency,
eff = (2S D )/( 1± 2± ). In the usual perturbative regime, 1/ j± t
1/eff , the rate of excitation from b to a is given to lowest order by a third-
order Fermi Golden rule:
 
  2  2
 D 
R(3) = 2π  S  δ(ωab + ω± − 2ν± ). (17)
 1± 2± 

This gives the effective rate of two-photon absorption of the signal field ν±
when assisted by the drive field ω± .
The application to subwavelength interference proceeds as follows:

– As the ± channels realize distinct resonances, the atoms will absorb


two photons from one signal beam or the other, but never one photon
from each beam.
– As a consequence, the spatial period of the fringes will carry the two-
photon wavelength, which is one-half the wavelength of each photon,
the same as achieved by a quantum, entangled state of the form |2, 0+
|0, 2.
Beyond the Rayleigh Limit in Optical Lithography 433

The net electric field as seen by the atoms on the surface of the photoresist
consist of two pairs of counter-propagating signal fields (of same intensity)
as well as the normally incident drive fields (see Figure 17) can be written
as:

E(x, t) = E S ei(k+ x−ν+ t) + ei(k− x−ν− t) + ED e−iω+ t + e−iω− t + c.c., (18)

where k± = ±(ν± /c) cos θ. Hence, the third-order excitation rate of the
atoms takes the general form
  t3  t2 2
d  t 
(3)
R (x, t) ∝  dt3 dt2 dt1
a|HI (x, t3 )HI (x, t2 )HI (x, t1 )|b ,
dt 0 0 0
(19)
where the interaction Hamiltonian is given in Equation (15).
Under conditions of three-photon resonance, the leading contributions
to the above integral will comprise exactly the two channels for the
frequency-selective excitation shown in Figure 17, whose rates were calcu-
lated in Equation (17). One ends up with the only two significant terms in
the field product where the same beam, + or −, contributes twice:

d  i2k+ x (3) 
(3) 2
R(3) (x, t) ∝ e r+ (t) + ei2k− x r− (t) ; (20)
dt
 t  t3  t2

(3)
r± (t) = dt3 dt2 dt1 ES ei 1± t1 ED e−i( 1± + 2± )t2 ES ei 2± t3 ,
0 0 0
(21)
where the dipole moments have been suppressed. If the one-photon detun-
ings are large, j± ν+ − ν− , then the excitation amplitudes r± (t) are
approximately equal, and the single beam, two-photon spatial frequen-
cies 2k± make up the interference pattern, i.e., the inter-beam cross terms
exp[i(k+ + k− )x] are absent because they are out of three-photon resonance.
Here one can see that the two-beam semiclassical lithography exactly
simulates quantum field lithography (Boto et al., 2000) with unlimited spa-
tial coherence. Moreover, the visibility is only limited by the small differ-
ence in excitation amplitudes of the two channels in Equation (20).

5.3.2 Generalization to N Photons


We now turn to a multi-photon resonance. The schematics for the system
are given in Figure 18. Two bunches of signal fields counter-propagate
along the substrate (θ = π/2) and a drive field is incident normally. Either
bunch of fields together
 with the drive field can excite the multi-photon
transition from level b to |a. The photons from the signal fields of frequen-
cies νn± are absorbed and the photons of the drive field with frequency ω0
434 Mohammad Al-Amri et al.

a
N+ N-

1+ 1-
0

2Nsin
b a
c3
N
c1 3
1 0
2
0 (2N-2)
1 c2N-2
4
2
c2
b

Figure 18 (a) The scheme of interferometric lithography. Two bunches of signal


fields counter-propagate (θ = π/2) and the drive field incidents normally. (b) The level
structure of the substrate atom. Either bunch of fields together with the drive field
satisfies the multi-photon resonance. n± is the detuning of intermediate level cn .
Figure reprinted with permission from Sun et al. (2007). Copyright 2007 by the
American Physical Society.

are emitted. The N signal photons satisfy a frequency summation resonance


condition

N
νn± = ωab + (N − 1)ω0 = Nν0 , (22)
n=1

such that the N-photon wave vector, Nν0 /c = 2π/(λ0 /N) is the same for
both bunches. We further require that any interchange of photons between
bunches, νn+ ↔ νn − , results in a loss of resonance. Therefore only two
resonant processes make up the interference.
The electric field on the surface is:

N


E(x, t) = ESn+ ei(kn+ x−νn+t) + ESn− ei(kn− x−νn− t) + ED e−iω0 t + c.c.,


n=1

where kn± = ±(νn± /c). In the level structure of the substrate, the inter-
mediate levels cj are off-resonant by detunings 1± = ωc1 b − ν1± , 2± =
ν1± − ω0 − ωc2 b , . . . , (2N−2)± = ωac2N−2 − νN± .
Beyond the Rayleigh Limit in Optical Lithography 435

Under the conditions of multi-photon resonance, the leading contribu-


tions to the multi-photon excitation rate come from the two resonant pro-
cesses, i.e.,
 2
(2N−1) d  i Nν0 x (2N−1) Nν x
−i c0 (2N−1) 

R (x, t) ∝ e r+ (t) + e r− (t) . (23)
dt 
c

If the one-photon detunings are large and ESn± are suitably chosen, the
excitation amplitudes r± (t) can be made approximately equal with a phase
difference. Factoring them out we find that the remaining expression looks
like the interference of single photon absorption with k = Nν0 /c. So the
exposure pattern are fringes with distance λ0 /2N.
The semiclassical scheme to multiple beams can be generalized as seen
in Figure 19 for N = 2. Each point on the slit plane is associated with two
complementary frequencies, ν1k and ν2k , that satisfy a sum frequency res-
onance achieved through opposing spatial chirps created using inverted
prisms. Then, photon pairs from a single spatial point on the slit plane will
be absorbed collinearly (i.e., same wave vector) in the focal plane. This sim-
ulates the multi-mode state vector |2, 0, . . . , 0 + · · · + |0, . . . , 0, 2. As shown
for quantum field lithography (D’Angelo et al., 2001), this would achieve
subwavelength resolution not only in the carrier fringe (double-slit inter-
ference), but also in the envelope (single slit diffraction). The ratio of the

ν ν ω

ν
ω
ν

Figure 19 Subwavelength diffraction for classical light. Two laser pulses are given
opposite spatial chirps using inverted prisms, and the resulting beams are combined
by a beam splitter (BS) to illuminate the slit plane with a position-dependent
frequency doublet, such that ν1k + ν2k = const. This creates a correlation between
wave vector and frequency pairs in the focal plane of lens L3, and writes a two-photon
pattern onto the Doppleron substrate in both carrier and envelope. Figure reprinted
with permission from Hemmer et al. (2006). Copyright 2006 by the American Physical
Society.
436 Mohammad Al-Amri et al.

pulse bandwidth to detector bandwidth determines the effective number


of wave vectors constituting the diffraction pattern, or equivalently, the
number of discrete partitions of the slit apertures.
We note that the above semiclassical approach can be adapted to imag-
ing, for example, two lenses in an f − f − f − f configuration. Here, one
introduces a correlation between wave vector and frequency in the focal
plane of the first lens, after the light has passed through the object, i.e., once
the angular spectrum of the light is prescribed by the diffracting apertures.
This can be accomplished by using a filter array in the focal plane that
selects the desired spatial chirp from a broadband input. Using a dual fil-
ter array, one can associate a frequency pair (ν1k , ν2k ) with each wave vector
such that the sum frequency is fixed: ν1k + ν2k = const. The result is a sub-
diffraction image spot (airy disk) created on the substrate in the image plane
when vignetting due to the lens apertures is taken into account. As in the
diffraction scheme, the bandwidth of the multi-photon process effectively
discretizes the angular spectrum on the substrate, which in turn determines
the resolution needed for the filter array in this imaging scheme.
A concern was expressed by Cho (2006) that this scheme may generate
only a tight pattern of parallel lines. We now turn to the possibility of
generating arbitrary subwavelength patterns using this scheme.

5.3.3 Generation of Arbitrary Patterns


Here we discuss the procedures to obtain arbitrary patterns in both one and
two dimensions (Sun et al., 2007). The pattern is described by a truncated
Fourier series and the scheme is based on multiple exposures. In order
to enable subwavelength resolution, we need to have the fundamental
frequency much larger than the signal frequencies. This can be done by
modifying the resonance condition.

(a) Arbitrary 1D Pattern


According to Equation (22), the fundamental frequency ν0 is the average
of νn± . This means that the fringes can be subwavelength, however, for
arbitrary pattern we need more harmonic components. Another point to
note is that ν0 is also limited by the level separation ωab .
To remove these limitations, we change the resonance condition to


nN
νn± = ωab + (nN − 1)ωN = Nν0 . (24)
n=1

This is the key equation. The main difference is that the number of sig-
nal fields involved in a multi-photon resonance changes from N into nN .
As far as the frequency summation equals Nν0 , there is no requirement
Beyond the Rayleigh Limit in Optical Lithography 437

this process must absorb N photons. The fundamental frequency ν0 can be


much larger than νn± to obtain a subwavelength resolution. ν0 can even be
much larger than ωab , which is a unique property due to the Doppleron-
type resonances of this scheme (Hemmer et al., 2006; Sun et al., 2007).
The drive frequency is also changed from ω0 to ωN , which can be differ-
ent for each N. The arbitrary photon number nN and the drive frequency
ωN provides more freedom to choose the fields.
In order to write an arbitrary 1D pattern, we start with a high frequency
ν0 . For any N (=1, 2, 3, …) we can always find some suitable nN and ωN to
achieve the multi-photon wave vector Nν0 /c. Two bunches of signal fields
grazing from +x and −x directions give the excitation rate

 2
(2nN −1) d  i Nν0 x (2nN −1) Nν x
−i c0 (2nN −1) 

R (x, t) ∝  e c r+ (t) + e r− (t) . (25)
dt

For each N we can use the above method to make an exposure. After mul-
tiple exposures we get fringes corresponding to N = 1, 2, 3, . . . , Nmax. The
final pattern is


Nmax  tN
P(x) = R(2nN −1) (x, t)dt
N=1 0


Nmax  2  Nν0 x 2

 (2nN −1)   i c Nν x
−i c0 iθN 
= cN r (tN ) e +e e 
N=1

Nmax
2Nν0 x 2Nν0 x
=Q+ aN cos + bN sin , (26)
c c
N=0

where cN is the ratio coefficient in R(2nN −1) . Here P(x) is a truncated Fourier
series with a penalty deposition Q. Such a series can approximate any 1D
pattern if enough components are included. The coefficients and phases of
each component can be controlled by ESn± and tN .

(b) Arbitrary 2D Pattern


Next, we would like to see how this method can be applied to generate
arbitrary pattern in two dimensions. Such a generalization still relies on
multiple exposures, using two bunches of signal fields each time. However,
the direction of these two bunches is not limited to the x axis.
In this approach, we firstsend in the two bunches from the oppo-
site directions ±(N x̂ + Mŷ)/ N 2 + M2 and make an exposure. The sum
438 Mohammad Al-Amri et al.

frequency of either bunch should satisfy


n√

N 2 +M2 
νn± = N 2 + M2 ν0 . (27)
n=1

Like the 1D case, we have the multi-photon excitation rate


 2
(2n√ −1) d  i Nν0 x+Mν0 y (2n−1) Nν x+Mν y
−i 0 c 0 (2n−1) 

R N 2 +M2 (x, t) ∝
 e c r+ (t) + e r− (t)  .
dt
(28)
Then
 change the directions of these two bunches into ±(N x̂ − M ŷ)/
N 2 + M2 and make another exposure. Note the field coefficients and
phases for the twoexposures could be different. For each nonzero (N, M)
pair that satisfies N 2 + M2  Nmax we make two exposures like this. If
N or M = 0 then only one exposure is needed. From Equation (21) we find
the final exposure pattern
 
2ν0 (Nx + My)
P(x, y) = aNM cos
√ c
0< N 2 +M 2 Nmax

2ν0 (Nx + My)
+ bNM sin A + a2NM + b2NM
c
2ν0 (Nx − My) 2ν0 (Nx − My)
+cNM cos + dNM sin
 c c
 
+ cNM
2 + d2NM = Q +

0 N 2 +M 2 Nmax

2ν0 Nx 2ν0 My
× (aNM + cNM ) cos cos
c c
2ν0 Nx 2ν0 My
+( − aNM + cNM ) sin sin
c c
2ν0 Nx 2ν0 My
+ (bNM + dNM ) sin cos
c c 
2ν0 Nx 2ν0 My
+(bNM − dNM ) cos sin . (29)
c c

This is a truncated 2D Fourier series with a penalty deposition. It can


approximate any two-dimensional pattern in principle.
As an example, we consider the test function

h if − π2 < 2νc0 x , c0 < π2 ,
2ν y
F(x, y) = (30)
0 elsewhere.
Beyond the Rayleigh Limit in Optical Lithography 439

Figure 20 Approximation of a 2D pattern using direction variation and multiple


exposures. Here the penalty deposition is already subtracted to get the truncated
Fourier series. The upper limit is Nmax = 10. Figure reprinted with permission from
Sun et al. (2007). Copyright 2007 by the American Physical Society.

For the upper limit Nmax = 10 we get the truncated Fourier series P(x) − Q
as shown in Figure 20. Except for the abrupt ramp and the corners of the hat,
the error to the test function is within ±0.1h, which is acceptable consider-
ing the number of components included. The penalty deposition Q = 2.15h.
Q and h are in arbitrary unit. This unit has to be chosen carefully to ensure
the photoresist threshold dose falls between Q and Q + h (Levinson, 2001).
For the places with exposure dose close to Q, the photoresist only has a
small loss after the development. While for the places with exposure dose
close to Q + h, the photoresist is completely removed.
The scheme due to Sun et al. (2007) has many advantages: It requires
neither superposition of entangled Fock states, nor broadband sensitive
substrate. It can approximate any 2D pattern in principle. In all the interfer-
ometric schemes discussed so far, we require N-photon absorbers.
However, whereas the earlier schemes require the signal frequency sum-
mation equal to the level separation of the substrate, in the present scheme
the summation can be much larger than the level separation. The way to
get larger summation is by increasing νn± or add another signal field. As
a result, higher fundamental frequency can be obtained which basically
means smaller pattern, and more Fourier components. This is the main
advantage of this scheme.
The main limitation of this scheme is the same as before: Due to multi-
photon absorption, we require highly intense fields which may make the
experimental realization quite difficult.
With these difficulty in mind, we now turn to possible schemes for sub-
wavelength lithography that require resonant atom field interaction.
440 Mohammad Al-Amri et al.

6. RESONANT SUBWAVELENGTH LITHOGRAPHY VIA


DARK STATE

In 2008, Kiffner et al. presented an alternative and novel scheme for reso-
nant subwavelength lithography without the requirement of an N-photon
absorption process (Kiffner et al., 2008). This scheme relied on the phe-
nomenon of coherent population trapping (CPT) (Arimondo, 1996;
Scully & Zubairy, 1997). Atoms are prepared in a position dependent state,
the subwavelength spatial distribution coming from the phase shifted
standing wave patterns in a multi-level resonant atom-field system.

6.1 Three-Level  Type System


It is known that CPT occurs in a three-level  type system as shown
in Figure 21a. The two ground states are represented by |b1  and |b2 ,
which are resonantly coupled to the excited state |a1  by laser fields with
Rabi frequencies R1 and S1 , respectively. In such configuration, we can
get the dark state once the system is optically pumped into a coherent

a b
a2
a1 a1

1 1 2
1 1 2

b3
b2 b2
b1 b1

c aN
a2
a1
N N

1 1 2 2
b N+1
bN
b3
b2
b1

Figure 21 Considered level schemes of the substrate. The ground states |bn  and
|bn+1  are resonantly coupled to the excited state |an  via Rabi frequencies Rn and
Sn , respectively. Each excited state |an  decays to the ground states |bn  and |bn+1 
by spontaneous emission. (a) Single  system. In (b), a sequence of two  systems is
displayed. (c) General level scheme with N excited and N + 1 ground states as a
sequence of N -type systems. Figure reprinted with permission from Kiffner et al.
(2008). Copyright 2008 by the American Physical Society.
Beyond the Rayleigh Limit in Optical Lithography 441

superposition of the two ground states which is then decoupled from the
applied light fields. The dark state is given by

|D  = (S1 |b1  − R1 |b2 )/ |S1 |2 + |R1 |2 . (31)

In this scheme, R1 and S1 represent standing waves in the z direction with


wave number k0 = 2π/λ0 that are formed by plane waves incident on the
substrate, see Figure 22a. We can see from Equation (32) that the population
of these two ground states in |D  depend very much on the ratio of the
Rabi frequencies R1 and S1 . If the standing waves corresponding to R1

Figure 22 (a) The standing wave patterns R1 and S1 are formed by two plane
waves Xi , Yi with wavelength λi . The period of each intensity pattern is given by λ0 /2,
where λ0 = λi / cos θi , and θi is chosen such that the effective wavelength in the
substrate plane is equal to λ0 for both R1 and S1 . Subfigures (b) and (c) correspond to
the  system shown in Figure 21b. Part (b) illustrates the intensity profiles of the
standing waves R1 and S1 according to Equation (31). Note that |R1 |2 and |S1 |2 are
not drawn to scale. The solid line in (c) shows the population of state |b1 
corresponding to R1 = (0 /10) cos (k0 z) and S1 = 0 sin (k0 z).
442 Mohammad Al-Amri et al.

and S1 are phase shifted with respect to each other, then the ratio R1 /S1
becomes position dependent. Hence the population in any ground level
(|b1  or |b2 ) can be made position dependent. The question we address is
whether a subwavelength population distribution can be obtained in one
of the ground levels (say |b1 ).
Let us look at the case that the two standing waves are phase shifted by
π/2, i.e.,
R1 = 0 cos (k0 z), S1 = 0 sin (k0 z). (32)
The populations of |b1  and |b2  in |D  are then given by

|S1 |2
|
b1 |D |2 = = [1 − cos (2k0 z)]/2, (33a)
|R1 |2 + |S1 |2
|R1 |2
|
b2 |D |2 = = [1 + cos (2k0 z)]/2. (33b)
|R1 |2 + |S1 |2
The two ground states populations show the same spatial modulation as
the intensity profiles of the standing waves corresponding to S1 and R1 ,
respectively. It is important to note that the populations do not depend
on the maximal Rabi frequency |0|, but rather on the ratio of the Rabi
frequencies R1 and S1 . Here the atomic population in (say level |b1 ) is
modulated with spatial frequency 2k0 giving a resolution of λ0 /2 which
gives the same result as the Rayleigh limit. This same limit is obtained
by assuming a linear response of a two-level atomic system. Here, in the
three-level atomic system, we have recovered the same limit but with very
different physics.
A question is whether we can obtain subwavelength resolution beyond
Rayleigh limit using the dark state physics used here. We will address this
question in the following sections.
Before moving to the more complicated system, we point out another
interesting feature in the present three-level system. For unequal ampli-
tudes of the Rabi frequencies of the two standing-wave fields, a single
very narrow spatial structure at a controllable position within a range of
λ/2 can be generated. For example, if we choose

R1 = (0 /10) cos (k0 z), S1 = 0 sin (k0 z), (34)

the population in the ground state |b1  is shown in Figure 22c. Here the
population in level |b1  is unity everywhere except at those point where
S1 = 0. This can be used to write desired structures point by point.

6.2 Generalization to 2× System


The above result for a single  type three-level system can be generalized
to 2 ×  or so-called M system, see Figure 21b. The generalization of the
Beyond the Rayleigh Limit in Optical Lithography 443

dark state Equation (32) for this system is given by Zubairy et al. (2002)

|D2×  = (S1 S2 |b1  − R1 S2 |b2  + R1 R2 |b3 )/ C2 , (35)

where S1,2 and R1,2 are the driving fields, and C2 is the normalization
constant and given by:

 
3 n−1 
2
C2 = |Rk |2 |Sj |2 . (36)
n=1 k=1 j=n

The probability to find the system in state |b1  is proportional to |S1 S2 |2 .


This involves the product of the fields S1 and S2 . If both S1 and S2 have a
sinusoidal oscillation behavior with respect to position, i.e., S1 ∼ sin (k0 z)
and S2 ∼ sin (k0 z + φ), we obtain

|S1 S2 |2 ∼ [cos (φ) − cos (2k0 z + φ)]2 . (37)

It is crucial to choose the relative phase shift of the two standing waves as
φ = π/2, in order to get:

|S1 S2 |2 ∼ [1 − cos (4k0 z)]/2. (38)

The population oscillations with wave number 4k0 are obtained, while the
contribution with wave number 2k0 has been canceled, see Figure 23c. The
spatial resolution is half of the classical limit! However, we should note
that for the moment we have neglected the normalization constant C2 in
Equation (36) which is also position dependent.
Here, unlike previous schemes by Yablonovitch and Vrijen (1999),
Bentley and Boyd (2004), Peér et al. (2004), and Hemmer et al. (2006) where
one needs high light field intensities, the scheme can work at very low laser
intensities. This happens because there is no need for nonlinear transition
amplitudes between different states but rather one exploits the nonlinear
dependence of the ground state population probabilities on the Rabi fre-
quencies, which only depends on relative field strengths.

6.3 Generalization to N ×  System


In this section we extend the generalization to level schemes with an N × 
structure, see Figure 21c. In the interaction picture and in rotating-wave
approximation, the interaction Hamiltonian of the N ×  system takes the
form:
N
 
HN× =  Rn |an 
bn | + Sn |an 
bn+1 | + H.c. , (39)
n=1
444 Mohammad Al-Amri et al.

Figure 23 (a) Each standing wave pattern Rn , Sn is formed by two plane waves
Xi , Yi with wavelength λi . The period of each intensity pattern is given by λ0 /2, where
λ0 = λi / cos θi , and θi is chosen such that the effective wavelength in the substrate
plane is equal to λ0 for all Rn and Sn . Subfigures (b) and (c) correspond to the M
system shown in Figure 21b. Part (b) illustrates the intensity profiles of the standing
waves Rn and Sn (n = 1, 2) according to Equation (36). Note that |Rn |2 and |Sn |2 are
not drawn to scale. The solid line in (c) shows the population of state |b1 
corresponding to Equations (37) and (39) with η = 1/20. It varies with wave number
4k0 . The dotted line is the corresponding result with nonzero ground state
decoherence rates γcoh . We set γcoh = γ , where γ is the full decay rate on the
|an  ↔ |bn±1  transition. Figure reprinted with permission from Kiffner et al. (2008).
Copyright 2008 by the American Physical Society.

where H.c. denotes the Hermitian conjugate. One key assumption is that
the resonance frequencies of the various transitions are sufficiently distinct
such that the Rabi frequencies Rn and Sn can be chosen individually.
The dynamics of the atomic density operator can be described by a
master equation

∂t  = −i[Hint , ]/ + Lγ , (40)


Beyond the Rayleigh Limit in Optical Lithography 445

where Lγ is just the spontaneous emission of the N excited states to the


ground states. We assume that each excited state |an , in this system, decays
to both ground states |bn  and |bn+1 .
One needs to pay attention to the case of steady state in order not to
have singular cases. The reason is that the system, in this case, depends
very much on the initial condition. This can be avoided by setting either all
Rabi frequencies Rn or all Sn (1  n  N) to be different from zero at any
point in space. The steady state of the system is then evaluated by Matsko
et al. (2003) to be

1   
N+1 n−1 N
|DN×  = √ ( − 1)n+1 Rk Sj |bn , (41)
CN n=1
k=1 j=n

where
 n−1
N+1  
N
CN = |Rk |2 |Sj |2 (42)
n=1 k=1 j=n
 
is the normalization constant and we set 0k=1 = N j=N+1 = 1. Thus as in
the case of the single  system the atoms are optically pumped into a dark
state |DN× . From now on, we suppose that the atoms have reached this
steady state.
As discussed in the previous subwavelength schemes, the main key
point here is to have a product of N sinusoidal waves with wave number
k0 to display spatial oscillations with wave number Nk0 only. However, the
question is what about all other harmonics with wave number nk0 with
n  N? The answer simply is that they can be canceled with a suitable
choice of the relative phase shifts of the standing waves. This property is
described by the trigonometric identities

N
sin (Nk0 z)
sin[k0 z + (n − 1)π/N] = , (43a)
2N−1
n=1

N
cos (Nk0 z)
sin[k0 z + (2n − 1)π/(2N)] = . (43b)
2N−1
n=1

It is straightforward to see the applicability of these identities to this sys-


tem. For this, we notice that the coefficient of |b1  in the expansion of
the dark state in Equation (42) is proportional to the product of all Rabi

frequencies Sn , i.e.,
b1 |DN×  ∼ N n=1 Sn . Similarly, the matrix element


bN+1 |DN×  ∼ N n=1 nR involves the product of all Rabi frequencies Rn .


If we choose the position dependence of Rn and Sn according to
Sn (z) = Sn sin[k0 z + (n − 1)π/N], (44a)
Rn (z) = Rn sin[k0 z + (2n − 1)π/(2N)], (44b)
446 Mohammad Al-Amri et al.

it follows from Equations (42) and (43) that we have

|
b1 |DN× |2 = A1 [1 − cos (2Nk0 z)]/2, (45a)
2
|
bN+1 |DN× | = AN+1 [1 + cos (2Nk0 z)]/2. (45b)

This equation is the main result of this scheme and it shows that the pop-
ulation of ground state is modulated by a spatial frequency N times of
the classical diffraction limit. We again need to note that the amplitudes
 N
A1 = N n=1 |Sn | /(CN 4
2 N−1 ) and A
N+1 = n=1 |Rn | /(CN 4
2 N−1 ) also depend

on position which will be discussed in next subsection.

6.4 Some Concerns


In the previous subsections, we mentioned that the normalization constant
is position dependent. This may, in general lead to undesirable spatial
oscillations for the atomic population. In order to have a high lithography
contrast, a full population oscillation amplitude is required. In principle,
it is best to set the amplitudes A1 and AN+1 to be equal to unity, and that
can be obtained if the parameters Rn and Sn in Equation (42) are chosen
according to

|R1 | = |SN | = η0 , 0 < η 1, (46a)


|RN | = |S1 | = |Rn | = |Sn | = 0 , 1 < n < N, (46b)

where 0 is an arbitrary positive Rabi frequency. Thus, the laser fields


driving the outermost ground states should be much weaker than all other
fields. In this case, the amplitudes A1 and AN+1 are then given by

A1 = AN+1 = 1/ 1 + η2 fN (z) , (47)

where the function fN is independent of η. As η is much smaller than unity,


A1 ≈ AN+1 ≈ 1 which is what we want. We also note that the population
of the remaining ground state (1 < n  N) are suppressed by a factor of η2 .
The second concern for this scheme is that CPT mechanism relies very
much on the preservation of the ground state coherence in order to evolve
into the stationary dark state. The question then is: Is this scheme still valid
for large ground state decoherence rate? The answer is yes, even a large
ground state decoherence rate γcoh does not affect the applicability of this
scheme as one can see in Figure 21c, the dotted curve coincides with the
solid line curve. Here dotted curve shows the result for the same parameters
as the solid line, but with γcoh set equal to the population decay rate γ on
the dipole-allowed transitions from the excited state to the ground states.
The third concern is that, in all our calculations, we assume Rn and Sn
have the same wave numbers k0 , but we also require that the frequencies
Beyond the Rayleigh Limit in Optical Lithography 447

of the fields Rn and Sn are distinct for individual addressing. These two
assumptions seems to be inconsistent. However, both conditions can be
met by choosing the appropriate incident angles θi , such that k0 = ki cos θi .
Despite of that, for the case of N ×  systems with larger N, it can be a
challenging task. However, Kiffner et al. (2008) estimate the influence of
small wave vector mismatch and find that the scheme also works for that
mismatch.
In this proposed scheme, a desired 2D final pattern can be achieved via
multiple exposure with different harmonics based on a Fourier decomposi-
tion as discussed before. We would need a medium that supports the gen-
eration of oscillations with maximal wave number 2Nk0 , where all smaller
wave numbers 2nk0 with 0 < n  N can be generated by appropriately
modifying the incident angle θ (Bentley & Boyd, 2004). The required har-
monics can also be generated without changing θ using different n × 
(n  N) subsystems of the same full level structure.

6.5 Experimental Demonstration of this Scheme


Here we discuss an experiment where the scheme presented above based
on dark state can be used to obtain sub-diffraction imaging. The experiment
approach of Li et al. (2008) is based on coherent population trapping in Rb
vapor.
The experimental setup is schematically shown in Figure 24. Before
describing the setup, we briefly discuss the three-level  system that was
used, see the inset in Figure 24. Probe and drive fields are applied to
the three-level  atoms which, in this case, are 87 Rb atoms with |a =
|52 P1/2 , F = 1, m = 0, |b = |52 S1/2 , F = 2, m = −1, and |c = |52 S1/2,
F = 2, m = +1. The system  evolves to the dark state which is given by
|D = (p |c − d |b)/ 2p + 2d , where d is the Rabi frequency of the
drive field, whereas p is of the probe filed. Usually, d p . When
drive field is nonzero, the dark state practically is |b and the medium is
transparent to the probe beam. However, if the drive field is zero, then the
dark state is |c, and the probe beam can be absorbed at these positions.
Therefore, the intensity profile of the transmitted probe beam is modulated
by the spatial intensity of the drive beam.
An external cavity diode laser is sent through a polarization-preserving
single-mode optical fiber. The output laser is vertically polarized and split
into two beams: drive and probe. The probe beam carries a small portion
of the laser intensity, and its polarization is rotated to be horizontal.
The drive beam is split into two beams that cross at a small angle, using
a Mach–Zehnder interferometer, see the dashed square in Figure 24. This
trick is used to generate a double-peak spatial distribution for the drive
field. Figure 24a shows neatly a two-peak interference pattern of crossing
448 Mohammad Al-Amri et al.

Figure 24 Experimental schematic. λ/2: half-wave plate; λ/4: quarter-wave plate; L1,
L2, L3: lenses; Mach–Zehnder interferometer MZ; piezoelectric transducer PZT;
polarizing beam splitter PBS, photodiode PD; CCD camera CCD. Image (a) is the
spatial intensity distribution of the drive field. Image (b) is the beam profile of the
parallel probe beam without the lens L1. Image (c) is the beam profile of the
diffraction limited probe beam with the lens L1. The inset is the energy diagram of the
three-level  Rb atom. Figure reprinted with permission from Li et al. (2008).
Copyright 2008 by the American Physical Society.

beams. Now, the probe and drive beams combine on a polarizing beam
splitter. This leads the probe field and the interference pattern of the drive
field to be overlapped in a Rb cell. Just before the cell, a quarter-wave plate
converts probe and drive beams into left and right circularly polarized
beams. The output beam is directed through Rb cell which is filled with
87 Rb and has a length of 4 cm. Magnetic shield is applied to isolate the cell

from the environmental magnetic fields. Inside the cell a solenoid provides
an adjustable, longitude magnetic field. The cell is installed in an oven that
heats the cell to reach an atomic density of 1012 cm3 . The laser is tuned to
the D1 line of 87 Rb at the transition 52 S1/2, F = 2 → 52 P1/2 , F = 1. After
passing through the cell, the probe and drive beams are converted back to
linear polarizations by another quarter-wave plate and then get separated
by a polarizing beam splitter. A photodiode PD is used to monitor the
power of transmitted probe field, while the spatial intensity distribution
of probe field is recorded by an imaging system. This system is consisting
of the lens L3 and a CCD camera.
Two experiments have been done. In the first experiment, the lenses L1
and L2 are removed. Figure 25a shows the spatial intensity distribution of
Beyond the Rayleigh Limit in Optical Lithography 449

Figure 25 The results of the first experiment, the lenses L1 and L2 are not used.
Image (a) shows snap shot of the intensity distribution of the drive field in the Rb cell.
While (b) shows the intensity distribution of the transmitted probe beam. Figure (c)
and (d) are the corresponding intensity profiles. Figure reprinted with permission
from Li et al. (2008). Copyright 2008 by the American Physical Society.

the drive beam, while Figure 25b shows the intensity distribution of the
transmitted probe beam. Both beams have the same spacing between two
peaks, but the probe intensity distribution has sharper peaks than the drive
intensity. Figure 25c and d are the horizontal cross sections of the drive and
the transmitted probe distributions. In the drive intensity profile, the width
(FWHM) of the peaks of the drive intensity is 0.4 mm (Figure 25c), while
it is 0.1 mm for the transmitted probe intensity (Figure 25d). The finesse
(ratio of spacing between peaks to the width of peaks) of the transmit-
ted probe intensity distribution is smaller than that of the drive intensity
distribution by a factor of 4.
In the second experiment, the lenses L1 and L2 are used. A parallel
probe beam with a diameter of 1.4 mm is focused by the lens L1. The focal
length of lens L1 is 750 mm, and the beam size at the waist has a diffraction
limited size 0.5 mm. The lens L2 is used to make the drive beam smaller in
the Rb cell, where the pattern of drive field is spatially overlapped with the
waist of the probe beam. Figure 26 shows the experimental result where
in (a) the drive field still has a double-peak intensity distribution. Again,
450 Mohammad Al-Amri et al.

Figure 26 The results of the experiment with the diffraction limited probe beam, the
lenses L1 and L2 are used. Images (a) and (b) show the image of the intensity
distribution of the drive field and the intensity distribution of the transmitted probe
field, respectfully, in the Rb cell. Curves (c) and (d) are the corresponding profiles.
Figure reprinted with permission from Li et al. (2008). Copyright 2008 by the American
Physical Society.

(b) shows similar double-peak intensity distribution for the transmitted


probe beam. Figure 26c and d are the horizontal cross sections of the drive
and transmitted probe profiles respectively. The width of the peaks in the
drive beam is 165 µm, while it is 93 µm for transmitted probe beam. The
structure created within the diffraction limit, for the probe beam, has a size
five times smaller than that of the diffraction limited size 500 µm. Thus,
this experiment successfully demonstrate that the diffraction limit can be
broken using the dark state physics.

7. SUBWAVELENGTH PHOTOLITHOGRAPHY VIA


RABI OSCILLATIONS

So far all the schemes for overcoming the Rayleigh limit are based on multi-
photon absorption or multi-level multi-beam systems. In 2010, Liao et al.
presented a novel and simple scheme for subwavelength lithography based
Beyond the Rayleigh Limit in Optical Lithography 451

Pulse laser 2 Pulse laser 1 a

b
τ12

Figure 27 Schematics for the proposed lithographic scheme. In the first step a laser
pulse induces the Rabi oscillation between the ground state and the excited state of
the molecules in the photoresist. Then a second laser pulse is applied to dissociate
the atoms in the excited state, cutting the chemical bound of the molecules. The
molecules change its solubility in the photoresist developer and the required pattern
can then be formed in the photoresist. Figure reprinted with permission from Liao et
al. (2010). Copyright 2010 by the American Physical Society.

on Rabi oscillations (Liao et al., 2010). This method is similar to the tradi-
tional photolithography but adding a critical step before dissociating the
chemical bound of the photoresist. The subwavelength pattern is achieved
by inducing the multi-Rabi-oscillation between the ground state and one
intermediate state.
In Figure 27, the molecules are simplified as a three-level system. In the
traditional optical lithography only one light beam is used to dissociate
the molecules. Here we sequentially turn on two different frequencies of
lights. The first light pulse induces Rabi oscillations between the ground
state and the intermediate excited state. Then the second light only dis-
sociates the molecules in the excited states but not those in the ground
state. Initially, the valence electrons of the chemical bond are in the ground
state |b. The first light has frequency ν1 which is resonant with the energy
difference ωab between |a and |b. The atoms or molecules undergo Rabi
oscillations between state |a and |b. At some predetermined time, the
molecules occupy the excited state |a with spatially modulated probabili-
ties. The second light pulse of frequency ν2 dissociates the molecules which
are in the excited state |a. The dissociation of the molecules cuts the chem-
ical bond and changes the chemical properties of the photoresist. We can
then use photoresist developer to wash out the dissociated molecules or
undissociated molecules (Wayne & Wayne, 1996). The resulting patterns
of the photoresist should then depend on the spatial distribution of the
excited state induced by the first light pulse. If the spatial modulation of
the probability to find the molecules at excited state has subwavelength pat-
tern, then the resulting patterns of the photoresist is also subwavelength.
We next show that our method can potentially lead to subwavelength pat-
terns of almost arbitrary accuracy, much easier than any other proposed
methods.
452 Mohammad Al-Amri et al.

7.1 Achieving the Subwavelength Pattern


The first step is very critical in order to achieve the subwavelength pattern.
We illustrate it in more detail and show how to prepare the molecules in a
subwavelength position dependent state. Two beams of light from opposite
directions are incident on the photoresist and they form a standing wave
on the surface of the photoresist. The standing light field interacts with
the molecules in the photoresist, for which we consider two kinds of light
sources: Continuous wave and a Gaussian pulse.

7.1.1 Continuous Wave Analysis


For simplicity, we first consider the continuous wave with frequency reso-
nant to the two atomic levels. The standing electric field on the
surface is
E(r, t) = E0 cos (ν1 t)eik·r + E0 cos (ν1 t)e−i(k·r+2φ)
= 2E e−iφ cos (kx cos θ + φ) cos (ν t)
0 1 (48)
in which E0 is the field amplitude, ν1 is the frequency, θ is the angle between
the incident light and the surface, and 2φ is the phase difference of the two
beams. Considering the dipole interaction between the electric field and
the atoms, and for the resonant case where ν1 = ω, the probability for the
atoms to be in the excited state |a at time T is:
 
1 − cos R T cos (kx cos θ + φ)
Pa (x, T) = , (49)
2
where we assume that the atoms are initially in the ground state, R =
(2|℘ba |E0 )/ is the Rabi frequency at the peak electric intensity, and |℘ba | is
the amplitude of the electric dipole moment. From this equation, we can see
that the probability in the excited state is spatially dependent and the shape
depends on the field area R T. As the molecules that are in the excited
state are dissociated, the spatial pattern also depends on the field area. We
now look at the spatial pattern in more detail. For simplicity, we choose
θ = 0 and φ = 0 which does not change the overall properties. Then we
have Pa (x, T) = (1 − cos[R T cos (kx)])/2 which is a double cosine function
and we can calculate the positions of the valleys and the peaks. First, we
note that the usual Rayleigh limit is obtained in the linear approximation
corresponding to R T 1. In this case Pa (x, T) ≈ α(1 + cos (2kx)) with
α = (R T)2 /8 leading to a resolution of λ/2. Next we look at the situation
where we are not restricted by the linear approximation and various Rabi
oscillations during the interaction time T are allowed.
When cos (kx) = 2mπ/R T, the probability Pa (x, T) is 0 which corre-
sponds to the valleys, and when cos (kx) = (2m + 1)π/R T, the proba-
bility is 1 which corresponds to the peaks, where m = 0, ±1, ±2, . . .. It is
Beyond the Rayleigh Limit in Optical Lithography 453

Figure 28 Subwavelength patterns generated by fields of different R T . (a)


R T = π ; (b) R T = 2π ; (c) R T = 3π ; (d) R T = 4π . The solid line is when the decay
is not included whereas the green dashed line shows the results with γ = ωab /1000.
Figure reprinted with permission from Liao et al. (2010). Copyright 2010 by the
American Physical Society.

readily seen that when R T = π, there are two valleys (x = λ/4, 3λ/4)
and three peaks (x = 0, λ/2, λ) within one wavelength (Figure 28a), which
gives the same result as the classical interference lithography. However,
when R T  2π, more valleys and peaks appear and the classical limita-
tion is broken. For example, when R T = 2π, there are five valleys (x =
0, λ/4, λ/2, 3λ/4, λ) and four peaks (x = λ/6, λ/3, 2λ/3, 5λ/6) within one
wavelength (Figure 28b). When R T becomes larger, the pattern becomes
smaller (Figure 28c and d). Therefore it is, in principle, possible to achieve
arbitrarily smaller subwavelength patterns by using stronger field or
lengthening the interaction time to induce more Rabi oscillations.
The physics behind the subwavelength pattern is the nonlinearity asso-
ciated with the Rabi oscillations. If the field is intense enough then the
Rabi oscillations are induced that help to modulate the population in level
|a, thus leading to the subwavelength oscillations for the population. For
example, when R T = π, one photon is absorbed and we are in the linear
regime. The corresponding resolution is the same as that obtained in the
classical lithography (Figure 28a). When R t = 2π, one photon is absorbed
454 Mohammad Al-Amri et al.

then another photon is emitted, leading to a full Rabi cycle. The resulting
resolution is half of the classical case (Figure 28b). And so on.
In order to see clearly the advantage of the present method over any
previous method for precision lithography, we refer to Equation (50). In
the case when φ = π/2, and kx cos θ 1, Equation (50) reduces to

1 − cos (2keff x)
Pa = , (50)
2

where keff = R Tk cos θ or

λeff = λ/(R T cos θ). (51)

Thus, a large number of Rabi oscillations in the interaction time can lead
to an arbitrarily small effective wavelength. Therefore a novel feature of
our scheme is that it should be possible to generate a nano-scale pattern
using a microwave field. For example, if two sublevels of a system have
energy difference of about 3 GHz and the coherence time is of the order
of 1 s, we can use a microwave pulse with wavelength 10 cm and pulse
duration 0.1 s to induce the Rabi oscillations between these two levels. If
R = 0.1 GHz, the resolution could be of the order 10 nm.

7.1.2 Gaussian Pulse Analysis


So far we considered the light field to be a continuous wave. However, in
practical applications we usually use laser pulses instead. Our study shows
that the result of the pulses is similar to that of the continuous wave.
Two beams of Gaussian pulses with the same frequency ν1 , same max-
imal amplitude
√ E0 and same full width at half maximum of the intensity
tFWHM = 2 ln 2σ are incident on the photoresist from opposite directions
with angle θ, and σ is the width of the pulse. They then form a standing
electric field described by

 
t2
E(x, t) = 2E0 exp − 2 cos (kx cos θ + φ) cos (ν1 t), (52)

where φ is the phase difference between these two pulses. The electric
field couples to the molecules in the photoresist. If ν1 is resonant to the
two energy levels |a and |b, the electric field drives Rabi oscillations
between these two levels. The Rabi frequency is R (x, t) = 2|℘ba |E0 exp
2
(− 2σt 2 ) cos (kx cos θ + φ)/. According to the Area theorem, the upper-level
Beyond the Rayleigh Limit in Optical Lithography 455

probability after the pulse is


∞
1 − cos[ −∞ R (x, t)dt]
Pa (x) 
2

1 − cos[2 2π σ |℘ba|E0 cos (kx cos θ + φ)]
=
2
1 − cos[0 t0 cos (kx cos θ + φ)]
= , (53)
2
where 0 = 2|℘ba |E0 / is the maximal Rabi frequency and we define t0 =

π
t
2 ln 2 FWHM
. From the equation, we see that the pattern generated by
the Gaussian pulse is the same as that of the continuous wave, but just
replace T by t0 . For example, when 0 t0 = 2π, one Rabi cycle is driven
and the pattern has a resolution of λ/4 (Figure 29b) which is the same as
Figure 28b. When 0 t0 = 4π, two Rabi cycles are driven and the resolution
is λ/8 (Figure 29b) which is the same as Figure 28d.
In the real system, decoherence time is an important factor we should
consider. When the pulse time exceed the decoherence time, the visibility
reduces dramatically. Usually the dephasing time is much smaller than
the decay time. Therefore here we only consider the effect of dephasing
time. Numerical simulation shows that when tFWHM = τ/2 (where τ is
the dephasing time), the visibility is reduced to about 80% but the total

Figure 29 (a) The Gaussian pulse. The red dash line is the amplitude profile and the
thick dark
 line is the intensity profile; (b) The pattern produced by the Gaussian pulse
π  t
when 2 ln 2 0 FWHM = 2π ; (c) The pattern produced by the Gaussian pulse when

π  t
2 ln 2 0 FWHM = 4π . The solid line is the result without the decoherence while the
green dashed line shows the results with tFWHM = τ/2. Figure reprinted with
permission from Liao et al. (2010). Copyright 2010 by the American Physical Society.
456 Mohammad Al-Amri et al.

patterns are almost the same as the result without decoherence. Therefore,
if tFWHM  τ , our scheme still works well.

7.2 Arbitrary Subwavelength Patterns in a Macroscopic Area


In the previous section we have shown how to achieve a simple subwave-
length pattern via coherent Rabi oscillations. For any practical applications
we should produce more complicated patterns (Kok et al., 2001; Pau et al.,
2001; Sun et al., 2007). In the following, we will discuss how to produce arbi-
trary subwavelength patterns in a macroscopic area. For one-dimensional
case, any functions in the range L can be expanded as a Fourier series:
∞  
a0  2nπx 2nπx
f (x) = + an cos + bn sin . (54)
2 L L
n=1

For the components with periods L/n larger than optical wavelength λ,
we just use the traditional way, i.e., shine two dissociative lasers with fre-
quency large enough to dissociate the molecules directly and they form
a standing wave correspond to the component and with strength related
to the Fourier coefficient. For the components with L/n < λ, we apply
our subwavelength scheme to realize them. We shine two phase locked
pulses with amplitude E0 from angle θ to form a standing wave and the
third one with amplitude E1 from the right angle to form a constant back-
ground. The resulting electric field is E(x, t) = [2E0 cos (kx cos θ + φ) +
2
E1 ] exp (− 2σt 2 ) cos (ν1 t). When nπ −   kx cos θ  nπ +  (n is an integer
and  is a small number),
 
t2
E(x, t)  ±[2E0 kx cos θ + E1 ] exp − 2 cos (ν1 t), (55)

where φ is set to be 90◦ . Then the Rabi frequency is
2|℘ab | 2|℘ab |E0 k cos θ 2|℘ab |E1
R (x) = [E0 cos (kx cos θ + φ) + E1 ] ≈ x+ .
  
(56)
The Rabi frequency is approximately a linear function of the position, and
the gradient of intensity is approximately a constant in the region (nπ −
)/k cos θ  x  (nπ + )/k cos θ. Then the pattern produced in this linear
region is
1 − cos (Ax + B)
Pa (x, T)  , (57)
2
 
where A = 2π/ ln 20 tFWHM k cos θ and B = 2π/ ln 21 tFWHM . The
coefficients A and B can be controlled by the field strength and the pulse
time. The effective wavelength

λeff = λ/( 2π/ ln 20 tFWHM cos θ) (58)
Beyond the Rayleigh Limit in Optical Lithography 457

can be arbitrary small by using stronger field or longer pulse time. We note
that ignoring the constant background 1/2, when B = 0, the pattern is a
cosine function; when B = π/2, the pattern is a sine function.
For example, if we want to produce sine pattern with λ/5 resolution
in a large region, we can do it in two steps (Figure 30): First, we etch
the pattern in the linear region as shown in Figure 30a. We then shift the
standing wave by a phase π/2 such that the linear region shifts by a dis-
tance of λ/2. This allows us to write the sine pattern in the remaining
region (Figure 30b) thus leading to the resulting sine pattern in the entire
region as shown in Figure 30c. The peak power for E0 is about 15 MW/cm2
(cos (θ) = 1/4, |℘ab | = 10 D, tFWHM = 1 ps) (Becker et al., 1988) and the
peak power for E1 is about 0.37 MW/cm2 . For larger resolution, the peak
power should increase. For example, to reach λ/10 resolution, the peak
power for E0 is about 60 MW/cm2 and the peak power for E1 is about
0.37 MW/cm2 . In addition, for the Fourier coefficients an and bn , we can
control the strength and time of the dissociation pulse to control the disso-
ciation rate or we can use different wavelengths with different absorption
rates.
We can also generalize our method to two-dimensional patterns. Arbi-
trary 2D periodic function with f (x + λ, y + λ) = f (x, y) can be simulated
by the truncated Fourier series:

 N 
M     
2π(mx + ny) 2π(mx − ny)
f (x, y) = amn cos + bmn cos
λ λ
m=0 n=0
   
2π(mx + ny) 2π(mx − ny)
+cmn sin + dmn sin
λ λ

a b

Figure 30 A proposed scheme to print a sine pattern in an arbitrary large region.


Figure reprinted with permission from Liao et al. (2010). Copyright 2010 by the
American Physical Society.
458 Mohammad Al-Amri et al.

   

M 
N
m2 + n2 2π cos (θ) (mx + ny) π
≈ amn cos cos  +
cos (θ) λ m2 + n2 2
m=0 n=0
  
m2 + n2 2π cos (θ) (mx − ny) π
+bmn cos cos  +
cos (θ) λ m2 + n2 2
  
m2 + n2 2π cos (θ) (mx + ny) π
+cmn sin cos  +
cos (θ) λ m2 + n2 2
  
m2 + n2 2π cos (θ) (mx − ny) π
+dmn sin cos  + (59)
cos (θ) λ m2 + n2 2

in which θ is near 90◦ . In the practical application, we should realize


each Fourier component one by one. For the first and third components

in Equation (6) we shine the pulses from directions (mx̂ + nŷ)/ m2 + n2
while for the other two components we shine the pulses from directions
(mx̂ − nŷ)/ m + n and 0 t0 = m2 + n2 /cos (θ). Besides, due to the
2 2
constant 1/2 appears in Equation (3), there is an additional penalty depo-
sition Q which depends on the Fourier coefficients. For example, apply-
ing the numerical simulation we print characters “TAMU-KACST" within
one wavelength (Figure 31). In the simulation, we take θ = 80◦ and M =
N = 15. Q = 0.24h where h is the height of the pattern. We have a total of
15×15×4 = 900 components and each component needs 4 pulses (three for
standing wave and one for dissociation). Therefore we need 3600 pulses in
total. Each component takes about 1 ms and the whole process takes about
1 s. In our example with the region λ × λ, the required maximal power is
about 200 MW/cm2 for a pulse duration of t0 = 5 ps.

7.3 Potential Realizations


The scheme shown in Figure 27 is a simplified model. In the following we
introduce two possible realizations of our scheme in two different systems.
The first one is in the organic molecular photochemistry. The typical
state energy diagram for the chemical bound is shown in Figure 32 (Turro
et al., 2009). Here S0 and S1 are the ground singlet state and the first
excited singlet state, respectively and T1 is the first excited triplet state.
KF is the fluorescence decay rate from S1 to S0 ; KP is the phosphores-
cence decay rate from T1 to S0 ; while KST is the intersystem crossing rate
from S1 to T1 . To induce Rabi oscillation, the system should be kept coher-
ently. Therefore, the decoherence time is an important parameter in our
scheme. The typical decoherence time τ is about 1∼5 ps at room temper-
ature (Fischer & Laubereau, 1975). To realize our subwavelength scheme,
the requirements for these parameters are tFWHM  τ and KST KF KP .
For tFWHM  τ , the system keeps coherent. For KST KF , intersystem
Beyond the Rayleigh Limit in Optical Lithography 459

Figure 31 A 2D pattern “TAMU-KACST” printed within one wavelength using the


present method. Parameters are M = N = 15, θ = 80◦ . Figure reprinted with
permission from Liao et al. (2010). Copyright 2010 by the American Physical Society.

a S1
k ST
T1
kF
kP

b S0

Figure 32 The schematics for the state energy diagram for molecular organic
photochemistry. Figure reprinted with permission from Liao et al. (2010). Copyright
2010 by the American Physical Society.

crossing from S1 to T1 dominates, which means that most of the molecules


at S1 will transfer to T1 instead of decaying to S0 . As the transition from T1
to S0 is spin forbidden, the lifetime (or phosphorescence time) of T1 is long.
Within the phosphorescence time, we shine the second pulse to dissociate
the molecules in state T1 . Indeed, the requirements can be satisfied in some
real systems. Usually, the time scale for KF : 105 –109 Hz; KST : 105 –1011 Hz;
KP : 10−2 –103 Hz. The Rabi frequency can be chosen as 1012 –1014 Hz. One
example is 1-Bromonaphthalene (Turro et al., 2009) for which KF ∼ 106 Hz,
KST ∼ 109 Hz, KP ∼ 30 Hz. The lifetime of the intermediate state T1 is about
460 Mohammad Al-Amri et al.

30 ms which is long enough for us to shine the second pulse. It is worth-


while to mention that the dipole–dipole interaction or exchange interaction
may induce energy transfer between neighboring molecules which limits
the resolution in our scheme (Turro et al., 2009). However these effects can
be ignored for the following reasons. The dipole–dipole energy transfer
rate is of the order of fluorescence rate when the distance between two
molecules is in the range of 1–5 nm. However, as we require KST to be
much larger than KF , the intersystem crossing to T1 occurs in times shorter
than that required for the dipole–dipole energy transfer to the neighboring
molecules. Also when the molecules are in the triplet state the dipole–
dipole energy transfer between the two molecules is forbidden. Therefore
the energy transfer due to the dipole–dipole interaction can be ignored
in our scheme. While the triplet–triplet energy transfer is allowed by the
electron exchange interaction, it can only happen at a distance within 1 nm
which is about the size of the molecules. Usually we cannot reach such
small patterns in the photoresist lithography.
The second possible realization is to generate a nanopattern using a
microwave. For example, the solid state system such as the NV-diamond
has a long dephasing time. The ground triplet state is split into two sub-
levels (ms = 0 and ms = ±1). The energy difference between these two
sublevels are about 2.9 GHz, which corresponds to a microwave with
wavelength of about 0.1 m. The dephasing time at room temperature can
reach 1.8 ms (Balasubramanian et al., 2009). Let tFWHM = 1 ms and R =
0.1 GHz, then we can reach a resolution of about 300 nm. At the low tem-
perature, the dephasing time can be even larger, and the pattern can be
smaller.

8. SUMMARY AND OUTLOOK

The diffraction limit is one of the major obstacles for the resolution of opti-
cal microscope and the current photolithography techniques. Researchers
have been struggling to increase the numerical aperture to improve the
resolution, but until now the improvement is not significant. Although the
working wavelength is reduced to print finer pattern, the light source,
lens and the photoresist working for high energy photons are hard to
find. Atomic and electron beam lithography are possible candidates of the
nanometer lithography, but they are restricted by the secondary electron
scattering and low throughput problem. Therefore, it is very interesting
and useful if we can somehow overcome the diffraction limit.
In the last two decades several ways to go beyond the diffraction limit
have been illustrated. Two-photon process and its generalization to multi-
photon process are first studied to increase the resolution of microscope
and later illustrated to shrink the pattern of photolithography. The
Beyond the Rayleigh Limit in Optical Lithography 461

principle of this method is straightforward, but the requirements of


extremely high laser intensity and low efficiency are the main concerns.
Photon entanglement successfully suppress the normal resolution term
and keep only the super-resolution term in the absorption rate. How to
produce ultrapure NOON quantum entanglement state is the biggest chal-
lenge of this scheme. Actually, quantum entanglement is not required
to suppress the normal resolution term. If we can carefully control the
light source by either controlling the phase relationship between pulses
or matching the wave vector and frequency with the material energy lev-
els in a narrow band, we can achieve similar result of quantum entan-
glement. However, multi-photon absorber is required to generate higher
harmonic patterns which also requires extremely high laser intensity and
also subject to low visibility. Spatial dependent dark state is a novel idea
to produce subwavelength resolution either in the microscopy or in the
photolithography. This scheme does not require quantum entanglement
or multi-photon absorber, but it requires additional levels and beams for
higher harmonic generation. Subwavelength resolution can be also simply
achieved by inducing Rabi oscillations between two energy levels in pho-
toresist. The advantages of this scheme are that it does not require quan-
tum entanglement, multi-photon absorber, multi-level and multi-beam.
Moreover, it is also very straightforward to produce higher harmonic com-
ponents in which we just need stronger pulse or longer pulse time. The
resolution limitation of this scheme is mainly due to the relaxation time
of the material. To achieve a higher resolution, we should find a material
which has a relatively long relaxation time.
Neither quantum entanglement nor multi-photon absorber is required
for subwavelength photolithography, but nonlinearity is somehow
involved in every scheme invented to break the diffraction limit until
now. Every scheme has its own advantages and disadvantages. In the near
future, we should find a suitable light source and a suitable material that
match all the requirements of one of the promising schemes. For industry
applications, effective way to generate arbitrary 2D pattern and through-
put are also important issues that we should consider in the future.

ACKNOWLEDGMENTS
We would like to thank many colleagues with whom we discussed the
subject matter of this article over the years. In particular we thank Joerg
Evers, Phil Hemmer, Martin Kiffner. Ashok Mutukrishnan, Marlan Scully,
and Qingqing Sun with whom we collaborated on different aspects of sub-
wavelength lithography. This work is supported by a grant from the King
Abdul Aziz City for Science and Technology (KACST). The research of MSZ
is supported by NPRP grant 08-043-1-011 by the Qatar National Research
Fund (QNRF).
462 Mohammad Al-Amri et al.

REFERENCES
Abbe, E. (1873). Beiträge zur Theorie des Mikroskops und der mikroskopischen. Archiv für
Mikroskopische Anatomie, 9, 413.
Abella, I. D. (1962). Optical double-photon absorption in cesium vapor. Physical Review Letters,
9, 453.
Agarwal, G. S., Boyd, R. W., Nagasako, E. M., & Bentley, S. J. (2001). Comment on quantum
interferometric optical lithography: Exploiting entanglement to beat the diffraction limit.
Physical Review Letters, 86, 1389.
Agarwal, G. S., & Kapale, K. T. (2006). Subwavelength atom localization via coherent popu-
lation trapping. Journal of Physics B: Atomic, Molecular and Optical Physics, 39, 3437.
Alfred, K-K. W. (2001). Resolution enhancement techniques in optical lithography. SPIE Press.
Alkaisi, M. M., Blaikie, R. J., & McNab, S. J. (2001). Nanolithography in the evanescent near
field. Advanced Materials, 13, 877.
Alkaisi, M. M., Blaikie, R. J., McNab, S. J., Cheung, R., & Cumming, D. R. S. (1999). Sub-
diffraction-limited patterning using evanescent near-field optical lithography. Applied
Physics Letters, 75, 3560.
Anisimov, P. M., & Dowling, J. P. (2009). Super resolution with superposition. Physics, 2, 52.
Arimondo, E. (1996). Coherent population trapping in laser spectroscopy. Progress in Optics,
35, 259.
Balasubramanian, G., Neumann, P., Twitchen, D., Markham, M., Kolesov, R., Mizuochi, N.,
et al. (2009). Ultralong spin coherence time in isotopically engineered diamond. Nature
Materials, 8, 383.
Becker, P. C., Fork, R. L., Brito Cruz, C. H., Gordon, J. P., & Shank, C. V. (1988). Optical stark
effect in organic dyes probed with optical pulses of 6-fs duration. Physical Review Letters,
60, 2462.
Bentley, S. J., & Boyd, R. W. (2004). Nonlinear optical lithography with ultra-high sub-Rayleigh
resolution. Optics Express, 12, 5735.
Berman, P. R., & Ziegler, J. (1977). Generalized dressed-atom approach to atom-strong-field
interactions—Application to the theory of lasers and Bloch–Siegert shifts. Physical Review
A, 15, 2042.
Betzig, E., Patterson, G. H., Sougrat, R., Lindwasser, O. W., Scott Olenych, S., Bonifacino, J. S.,
et al. (2006). Imaging intracellular fluorescent proteins at nanometer resolution. Science,
313, 1642.
Betzig, E., & Trautman, J. K. (1992). Near-field optics: Microscopy, spectroscopy, and surface
modification beyond the diffraction limit. Science, 257, 189.
Binnig, G., & Quate, C. F. (1986). Atomic force microscope. Physical Review Letters, 56, 930.
Boto, A., Kok, P., Abrams, D., Braunstein, S., Williams, C. P., Dowling, J. P. (2000). Quantum
interferometric optical lithography: Exploiting entanglement to beat the diffraction limit.
Physical Review Letters, 85, 2733.
Boyd, R. W., & Bentley, S. J. (2006). Recent progress in quantum and nonlinear optical lithog-
raphy. Journal of Modern Optics, 53, 713.
Boyd, R. W., & Bentley, S. J. (2006). Recent progress in quantum and nonlinear optical lithog-
raphy. Erratum Journal of Modern Optics, 53, 1529.
Brueck, S. R. J. (2005). Optical and interferometric lithography—Nanotechnology enablers.
Proceedings of the IEEE, 93, 1704.
Brueck, S. R. J., Zaidi, S. H., Chen, X., & Zhang, Z. (1998). Interferometric lithography: From
periodic arrays to arbitrary patterns. Microelectronic Engineering, 41–42, 145.
Chang, J., Evers, J., Scully, M. O., & Zubairy, M. S. (2006). Measurement of the separation
between atoms beyond diffraction limit. Physical Review A, 73, 031803(R).
Chang, H. J., Shin, H., O’Sullivan-Hale, M. N., & Boyd, R. W. (2006). Implementation of sub-
Rayleigh-resolution lithography using an N-photon absorber. Journal of Modern Optics, 53,
2271.
Chiu, G. L. -T., & Shaw, J. M. (1997). Optical lithography: Introduction. IBM Journal of Research
and Development, 41, 3.
Cho, A. (2006). A new way to beat the limits on shrinking transistors?. Science, 312, 672.
Chris, A. M. (2007). Fundamental principles of optical lithography: The science of microfabrication.
John Wiley & Sons.
Beyond the Rayleigh Limit in Optical Lithography 463

D’Angelo, M., Chekhova, M. V., & Shih, Y. (2001). Two-photon diffraction and quantum lithog-
raphy. Physical Review Letters, 87, 013602.
Denk, W., Strickler, J. H., & Webb, W. W. (1990). Two-photon laser scanning fluorescence
microscopy. Science, 248, 73.
Denk, W., Svoboda, K. (1997). Photon upmanship: Why multiphoton imaging is more than a
gimmick. Neuron, 18, 351.
Diaspro, A. (Ed.). (2010). Nanoscopy and multidimensional optical fluorescence microscopy. Boca
Raton: CRC Press.
Donnert, G., Keller, J., Medda, R., Alexandra Andrei, M., Rizzoli, S. O., Lührmann, R., et al.
(2006). Macromolecular-scale resolution in biological fluorescence microscopy. Proceedings
of the National Academy of Sciences, 103, 11440.
Dryakhlushin, V. F., Klimov, A. Y., Rogov, V. V., & Vostokov, N. V. (2005). Near-field optical
lithography method for fabrication of the nanodimensional objects. Applied Surface Science,
248, 200.
Dudovich, N., Dayan, B., Gallagher Faeder, S. M., & Silberberg, Y. (2001). Transform-limited
pulses are not optimal for resonant multiphoton transitions. Physical Review Letters, 86, 47.
Fischer, S. F., & Laubereau, A. (1975). Dephasing processes of molecular vibrations in liquids.
Chemical Physics Letters, 35, 6.
Fonseca, E. J. S., Monken, C. H., & Pádua, S. (1999). Measurement of the de Broglie wavelength
of a multiphoton wave packet. Physical Review Letters, 82, 2868.
Freund, S. M., Römheld, M., & Oka, T. (1975). Infrared-radio-frequency two-photon and mul-
tiphoton lamb dips for CH3 F. Physical Review Letters, 35, 1497.
Goeppert-Mayer, M. (1931). Ueber elementarakte mit zwei quantenspruengen. Annalen der
Physik, 9, 273.
Gorshkov, A. V., Jiang, L., Greiner, M., Zoller, P., & Lukin, M. D. (2008). Coherent quantum
optical control with subwavelength resolution. Physical Review Letters, 100, 093005.
Gustafsson, M. G. L. (2005). Nonlinear structured-illumination microscopy: Wide-field flu-
orescence imaging with theoretically unlimited resolution. Proceedings of the National
Academy of Sciences, 102, 13081.
Haroche, S., & Hartmann, F. (1972). Theory of saturated-absorption line shapes. Physical
Review A, 6, 1280.
Harry, J. L. (2005). Principles of lithography. SPIE Press.
Hell, S. W. (1994). Improvement of lateral resolution in far-field light microscopy using two-
photon excitation with offset beams. Optics Communications, 106, 19.
Hell, S. W. (2007). Far-field optical nanoscopy. Science, 316, 1153.
Hell, S. W., & Kroug, M. (1995). Ground-state-depletion fluorescence microscopy: A concept
for breaking the diffraction resolution limit. Applied Physics B, 60, 495.
Hell, S. W., & Wichmann, J. (1994). Breaking the diffraction resolution limit by stimulated
emission: Stimulated-emission-depletion fluorescence microscopy. Optics Letters, 19, 780.
Helmchen, F., & Denk, W. (2005). Deep tissue two-photon microscopy. Nature Methods, 2, 932.
Hemmer, P. R., Muthukrishnan, A., Scully, M. O., & Zubairy, M. S. (2006). Quantum lithog-
raphy with classical light. Physical Review Letters, 96, 163603.
Herkommer, A. M., Schleich, W. P., & Zubairy, M. S. (1997). Autler–Townes microscopy of a
single atom. Journal of Modern Optics, 44, 2507.
Ivan, M. G., Scaiano, J. C. (2010). Photoimaging and lithographic processes in polymers. In
N. S. Allen (Ed.), Photochemistry and photophysics of polymer materials (pp. 479–507). New
Jersey: John Wiley & Sons.
Kawabe, Y., Fujiwara, H., Okamoto, R., Sasaki, K., & Takeuchi, S. (2007). Quantum interference
fringes beating the diffraction limit. Optics Express, 15, 14244.
Kawata, S., Sun, H. -B., Tanaka, T., & Takada, K. (2001). Finer features for functional microde-
vices. Nature (London), 412, 697.
Kiffner, M., Evers, J., & Zubairy, M. S. (2008). Resonant interferometric lithography beyond
the diffraction limit. Physical Review Letters, 100, 073602.
Klar, T. A., & Hell, S. W. (1999). Subdiffraction resolution in far-field fluorescence microscopy.
Optics Letters, 24, 954.
Klar, T. A., Jakobs, S., Dyba, M., Egner, A., & Hell, S. W. (2000). Fluorescence microscopy with
diffraction resolution barrier broken by stimulated emission. Proceedings of the National
Academy of Sciences, 97, 8206.
464 Mohammad Al-Amri et al.

Kok, P., Boto, A. N., Abrams, D. S., Williams, C. P., Braunstein, S. L., Dowling, J. P. (2001).
Quantum-interferometric optical lithography: Towards arbitrary two-dimensional pat-
terns. Physical Review A, 63, 063407.
Korobkin, D. V., & Yablonovitch, E. (2002). Twofold spatial resolution enhancement by two-
photon exposure of photographic film. Optical Engineering, 41, 1729.
Kyröla, E., & Stenholm, S. (1977). Velocity tuned resonances as multi-Doppleron processes.
Optics Communications, 22, 123.
Lee, S., Byers, J., Jen, K., Zimmerman, P., Rice, B., Turro, N. J., et al. (2008). An analysis of
double exposure lithography options. Proceedings of SPIE, 6924, 69242A.
Lee, S. K., & Lee, H. -W. (2008). The role of entanglement in quantum lithography. Journal of
the Physical Society of Japan, 77, 124001.
Levinson, H. J. (2001). Principles of lithography. Washington: SPIE Press.
Liao, Z., Al-Amri, M., & Zubairy, M. S. (2010). Quantum lithography beyond the diffraction
limit via Rabi oscillations. Physical Review Letters, 105, 183601.
Li, H., Sautenkov, V. A., Kash, M. M., Sokolov, A. V., Welch, G. R., Rostovtsev, Y. V., et al.
(2008). Optical imaging beyond the diffraction limit via dark states. Physical Review A, 78,
013803.
Liu, Z., Wei, Q. H., & Zhang, X. (2005). Surface plasmon interference nanolithography. Nano
Letters, 5, 957.
Lu, C., & Lipson, R. H. (2010). Interference lithography: A powerful tool for fabricating peri-
odic structures. Laser and Photonics Reviews, 4, 568.
Luo, X. G., & Ishihara, T. (2004a). Subwavelength photolithography based on surface-plasmon
polariton resonance. Optics Express, 12, 3055.
Luo, X. G., & Ishihara, T. (2004b). Surface plasmon resonant interference nanolithography
technique. Applied Physics Letters, 84, 4780.
Mack, C. (2007). Fundamental optical principles of lithography: The science of microfabrication. West
Sussex, England: John Wiley & Sons.
Macovei, M., Evers, J., Keitel, C. H., & Zubairy, M. S. (2007). Localization of atomic ensembles
via superfluorescence. Physical Review A, 75, 033801.
Mandel, L., Sudarshan, E. C. G., & Wolf, E. (1964). Theory of photoelectric detection of light
fluctuation. Proceedings of the Physical Society, 84, 435.
Marconi, M. C., & Wachulak, P. W. (2010). Extreme ultraviolet lithography with table top
lasers. Progress in Quantum Electronics, 34, 173.
Martin, O. J. F. (2003). Surface plasmon illumination scheme for contact lithography beyond
the diffraction limit. Microelectronic Engineering, 67, 24.
Matsko, A. B., Novikova, I., Zubairy, M. S., & Welch, G. R. (2003). Nonlinear magneto-optical
rotation of elliptically polarized light. Physical Review A, 67, 043805.
McClain, W. M. (1974). Two-photon molecular spectroscopy. Accounts of Chemical Research, 7,
129.
Menon, R., Patel, A., Gil, D., & Smith, H. I. (2005). Maskless lithography. Materials Today, 7021,
26. ISSN:1369
Menon, R., Walsh, M., Galus, M., Chao, D., Patel, A., & Smith, H. I. (2005). Maskless lithog-
raphy using diffractive-optical arrays. In Frontiers in optics, OSA technical digest series.
Optical Society of America (Paper FWU5).
Mollow, B. R. (1968). Two-photon absorption and field correlation functions. Physical Review,
175, 1555.
Mompart, J., Ahufinger, V., & Birkl, G. (2009). Coherent patterning of matter waves with
subwavelength localization. Physical Review A, 79, 053638.
Neice, A. (2010). Methods and limitations of subwavelength imaging. Advances in imaging and
Electron Physics, 163, 117.
Ono, T., & Esashi, M. (1998). Subwavelength pattern transfer by near-field photolithography.
Japanese Journal of Applied Physics, Part 1, 37, 6745.
Pau, S., Watson, G. P., & Nalamasu, O. (2001). Writing an arbitrary pattern using interference
lithography. Journal of Modern Optics, 48, 1211.
Peér, A., Dayan, B., Vucelja, M., Silberberg, Y., & Friesem, A. A. (2004). Quantum lithography
by coherent control of classical light pulses. Optics Express, 12, 6600.
Pendry, J. B. (2000). Negative refraction makes a perfect lens. Physical Review Letters, 85, 3966.
Beyond the Rayleigh Limit in Optical Lithography 465

Pittman, T. B., Shih, Y. H., Strekalov, D. V., & Sergienko, A. V. (1995). Optical imaging by
means of two-photon quantum entanglement. Physical Review A, 52, R3429.
Pritchard, D. E., & Gould, P. L. (1985). Experimental possibilities for observation of unidirec-
tional momentum transfer to atoms from standing-wave light. Journal of the Optical Society
of America B – Optical Physics, 2, 1799.
Qamar, S., Zhu, S. -Y., & Zubairy, M. S. (2000). Atom localization via resonance fluorescence.
Physical Review A, 61, 063806.
Rayleigh, L. (1879). Investigations in optics, with special reference to the spectroscope. Philo-
sophical Magazine, 8, 261.
Reid, J., & Oka, T. (1977). Direct observation of velocity-tuned multiphoton processes in the
laser cavity. Physical Review Letters, 38, 67.
Rittweger, E., Han, K. Y., Irvine, S. E., Eggeling, C., & Hell, S. W. (2009). STED microscopy
reveals crystal colour centres with nanometric resolution. Nature Photonics, 3, 144.
Rothschild, M. (2010). A roadmap for optical lithography. Optics and Photonics News, 21, 27.
Rudenberg, H. G., & Rudenberg, P. G. (2010). Origin and background of the invention of the
electron microscope: Commentary and expanded notes on Memoir of Reinhold Rüden-
berg. Advances in Imaging and Electron Physics, 160, 207.
Schuller, J. A., Barnard, E. S., Cai, W., Jun, Y. C., White, J. S., Brongersma, M. L. (2010). Plas-
monics for extreme light concentration and manipulation. Nature Materials, 9, 193.
Scully, M. O., & Lamb, W. E. (1969). Quantum theory of an optical maser. III. Theory of
photoelectron counting statistics. Physical Review, 179, 368.
Scully, M. O., & Zubairy, M. S. (1997). Quantum optics. Cambridge, England: Cambridge
University Press.
Seisyan, R. P. (2011). Nanolithography in microelectronics: A review. Technical Physics, 56,
1061.
Sheats, J. R., & Smith, B. W. (Eds.). (1998). Microlithography science and technology (1st ed.). New
York: Marcel Dekker Inc..
Shih, Y. (2007). Quantum imaging. IEEE Journal of Selected Topics in Quantum Electronics, 13,
1016.
Spille, E., & Feder, R. (1977). X-ray lithography. Topics in Applied Physics, 22, 35.
Srituravanich, W., Fang, N., Sun, C., Luo, Q., & Zhang, X. (2004). Plasmonic nanolithography.
Nano Letters, 4, 1085.
Strekalov, D. V., Sergienko, A. V., Klyshko, D. N., & Shih, Y. H. (1995). Observation of two-
photon “Ghost" interference and diffraction. Physical Review Letters, 74, 3600.
Strickler, J. H., & Webb, W. W. (1991). Three-dimensional optical data storage in refractive
media by two-photon point excitation. Optics Letters, 16, 1780.
Strickler, J. H., & Webb, W. W. (1991). Two-photon excitation in laser scanning fluorescence
microscopy. Proceedings of SPIE, 1398, 107.
Sun, Q., Al-Amri, M., Scully, M. O., & Zubairy, M. S. (2011). Subwavelength optical microscopy
in the far field. Physical Review A, 83, 063818.
Sun, Q., Hemmer, P. R., & Zubairy, M. S. (2007). Quantum lithography with classical light:
Generation of arbitrary patterns. Physical Review A, 75, 065803.
Taylor, J. S., Sommargren, G. E., Sweeney, D. W., & Hudyma, R. M. (1998). Fabrication and
testing of optics for EUV projection lithography. Proceedings of SPIE, 3331, 580.
Tsang, M. (2007). Relationship between resolution enhancement and multiphoton absorption
rate in quantum lithography. Physical Review A, 75, 043813.
Turro, N. J., Ramamurthy, V., & Scaino, J. C. (2009). Principles of molecular photochemistry: An
introduction. . Sausalito, CA: University Science Books.
Vieu, C., Carcenac, F., P’ipin, A., Chen, Y., Mejias, M., Lebib, A., et al. (2000). Electron beam
lithography: Resolution limits and applications. Applied Surface Science, 164, 111.
Wayne, C. E., & Wayne, R. P. (1996). Photochemistry.. New York: Oxford University Press.
Williams, C., Kok, P., Lee, H., & Dowling, J. P. (2006). Quantum lithography: A non-computing
application of quantum information. Informatik – Forschung und Entwicklung, 21, 73.
Wu, E. S., Strickler, J. H., Harrell, W. R., & Webb, W. W. (1992). Optical/laser microlithography
V. Proceedings of SPIE, 1674, 776.
Xie, Z., Yu, W., Wang, T., Zhang, H., Fu, Y., Liu, H., et al. (2011). Plasmonic nanolithography:
A review. Plasmonic, 6, 565.
466 Mohammad Al-Amri et al.

Yablonovitch, E., & Vrijen, R. B. (1999). Optical projection lithography at half the Rayleigh
resolution limit by two-photon exposure. Optical Engineering, 38, 334.
Yavuz, D. D., & Proite, N. A. (2007). Nanoscale resolution fluorescence microscopy using
electromagnetically induced transparency. Physical Review A, 76, 041802.
Zhuang, X. (2009). Nano-imaging with STORM. Nature Photonics, 3, 365.
Zubairy, M. S., Matsko, A. B., & Scully, M. O. (2002). Resonant enhancement of high-order
optical nonlinearities based on atomic coherence. Physical Review A, 65, 043804.

You might also like