ChapterPDF Available

Abstract and Figures

The so-called space-time analogy constitutes a source of inspiration to understand, engineer, and implement new systems for ultrafast optical signal processing based on concepts borrowed from the well-established field of Fourier Optics. In this review, we start by describing in a comprehensive manner the most basic notions of this analogy and discuss some recent developments with state-of-the-art technology, including the silicon-chip-based time lens and ultra-dispersive Raman devices, among others. Apart from the applications in optical communications, special emphasis is paid on the collateral benefits that the "ultra" appellative brings in fields as diverse as optical frequency comb generation, arbitrary waveform generation, optical coherence tomography, sensors, imaging, or quantum information processing.
Content may be subject to copyright.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 1 — #1
CHAPTER 1
Space-Time Analogies
in Optics
V´
ıctor Torres-Companya,b,c
, Jes ´
us Lancisc
,
and Pedro Andr ´
esd
aDepartment of Electrical and Computer Engineering,
McGill University, H3A 2A7 Montreal, QC, Canada
bCurrent address: Department of Electrical and Computer
Engineering, Purdue University, IN-47907, USA
cDepartament de F´ısica, Universitat Jaume I, 12007
Castell´o de la Plana, Spain
dDepartamento de ´
Optica, Universitat de Val`encia, 46100
Burjassot, Spain
Contents 1. Introduction 2
2. Ultrashort Light Pulse Propagation in Dispersive
Homogeneous Media 3
3. First-Order Approximation: Space-Time Analogy 5
4. Elements and their Implementations 6
4.1. Temporal ABCD Matrices 6
4.2. Spectral Dual Formalism 7
4.3. Basic Photonic Components 8
4.4. Some ABCD Matrix Properties 17
5. Coherent Ultra-High-Speed Optical Systems
and their Applications 18
5.1. Tunable Delays 18
5.2. Real-Time Fourier Transformation 20
5.3. Time-to-Frequency Converters 27
5.4. Temporal Imaging Systems 32
5.5. Spectral Imaging System 35
5.6. Ultrafast Fourier Processing Systems 37
5.7. Joint Transform Correlator 41
5.8. Temporal Talbot Effect 45
5.9. Temporal Array Illuminators 49
6. Temporal Optics in the Noncoherent Regime 52
Progress in Optics, Volume 56, c
2011, Elsevier B.V. All rights reserved.
ISSN 0079-6638, DOI: 10.1016/B978-0-444-53886-4.00001-0.
1
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 2 — #2
2Space-Time Analogies in Optics
6.1. Scalar Coherence Theory for Nonstationary
Partially Coherent Wave fields 53
6.2. Space-Time Analogy for Partially Coherent
Wave fields 55
6.3. Propagation of Nonstationary Partially
Coherent Pulses 56
6.4. Temporal van Cittert-Zernike Theorem 57
6.5. Temporal Lau Effect 62
7. Temporal Optics in the Two-Photon Regime 65
7.1. Quantum Two-Photon Correlation Functions
and Two-Photon States 66
7.2. Tailoring Two-Photon States in Spontaneous
Parametric Down-Conversion 67
7.3. Propagation of Ultrafast Two-Photon Wave
Packets 70
7.4. Two-Photon Temporal Far-Field
Phenomenon 72
7.5. Two-Photon Temporal Imaging 73
7.6. Some Remarks on Ultrafast Two-Photon
Processors 74
8. Conclusions 75
Acknowledgments 76
References 76
1. INTRODUCTION
In the last decades, the generation of pulsed beams with pulse durations in
the order of pico- and femtosecond has constituted an important topic for
the physics and engineering communities, where researchers find them-
selves continuously pushing the limits to satisfy quite radical premises
such as ultrafast, ultrabroad or ultrasmall. The characteristics of this kind
of optical radiation, that is, broadband spectrum, enormous temporal res-
olution, high peak but low average power, potentially high repetition
rate, and high spatial coherence make it an indispensable tool to develop
many applications in different fields of science and technology (Fermann,
Galvanauskas, & Sucha,2003).
At a systems level, the so-called space-time analogy constitutes a source
of inspiration to design and implement new schemes for processing
these ultrafast optical signals based on concepts borrowed from the well-
established field of Fourier optics (Goodman,1996), leading to what is
now popularly termed as Temporal Optics. The key relies on noting the
mathematical similarity between the equations that govern the parax-
ial diffraction of one-dimensional monochromatic light beams and those
describing the distortion of plane-wave pulses in a first-order dispersive
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 3 — #3
Ultrashort Light Pulse Propagation in Dispersive Homogeneous Media 3
medium (Akhmanov, Sukhorukov, & Chirkin,1969;Treacy,1969). Thanks
to the continuous advances in the optoelectronics and optical commu-
nication industries, the analogy can be extended to include other ele-
ments such as imaginglenses (Kolner,1994a) or prisms (van Howe &
Xu,2006). These optoelectronic tools have paved an avenue for creating
innovative temporal processing systems, bringing the “ultra” appella-
tive to fields as diverse as optical interconnects, optical communications,
microwave photonics, biophotonics, or quantum information processing,
among others.
In this review, we shall provide the most basic notions for the under-
standing of the space-time analogy, including the fundamental elements
and their possible implementations with state-of-the-art technology. We
provide a comprehensive approach, based on a temporal matrix formal-
ism (Nakazawa et al.,1998), to describe the characteristics of some of the
most widespread system processing architectures. Special emphasis shall
be paid on their applications in the above mentioned fields, highlighting
not only their innovative character but offering a comparative study with
respect to other, more conventional, solutions. Finally, we will review the
extension of the space-time analogy to the noncoherent (Lancis, Torres-
Company et al.,2005) and even to the nonclassical (quantum) regime (Tsang
& Psaltis,2006;Torres-Company, Lajunen et al.,2008), where this anal-
ogy still offers some potential to design – and sometimes interpret – new
physical phenomena with the aid of the notions of classical Fourier optics.
2. ULTRASHORT LIGHT PULSE PROPAGATION IN
DISPERSIVE HOMOGENEOUS MEDIA
In this section, we introduce the fundamental equations describing the
linear distortion of an ultrashort light pulse in a dispersive homogeneous
medium. Such a basic physical problem constitutes the cornerstone of
ultrafast optical signal processing.
Let us assume a scalar optical field, described by its analytic signal
U(r,t0), propagating linearly in a waveguide with translational symmetry
through the z-direction zand filled with a homogeneous lossless disper-
sive medium. The propagation constant is β(ω0)=n00/c, where c is
the speed of light in vacuum and n0)the frequency-dependent refractive
index. We can then write
U(r,t0)=A(x,y(z,t0)exp[i0t0β0z)]. (1)
where ψ(z,t0)is the pulse envelope, which modulates the monochromatic
carrier wave of angular frequency ω0; the constant β0=β(ω0); and A(x,y)
is the transversal spatial distribution of the mode, evaluated at ω0. Once
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 4 — #4
4Space-Time Analogies in Optics
A(x,y)is calculated, we only need to know ψ(z,t0)to determine U(r,t0)
at every zposition. The evolution of ψ(z,t0)is then described by a wave
equation.
Since the functional form of β(ω0)is usually unknown, it is very useful
to perform a Taylor expansion (Agrawal,2007)
β(ω0)=β0+β10ω0)+β2
2! 0ω0)2+β3
3! 0ω0)3+ · ·· (2)
where βn=dnβ(ω0)/dω0n|ω0=ω0is the nth-order dispersion coefficient of
the waveguide, with n=0, 1, 2, 3, . . . From now on, we will express the
temporal variations of the pulse in a reference framework moving at the
group velocity of the wave packet, t=t0β1z.
The wave equation describing the envelope distortion is of second
order in z. In order to reduce this equation to first order, the slowly
varying envelope approximation (SVEA) is usually invoked in the mul-
ticycle regime (Agrawal,2007). This approximation requires that the
function ψ(z,t)does not change significantly through a distance com-
pared with the carrier wavelength, and the pulse duration to be much
larger than the carrier oscillation period. Mathematically, it is translated
into |ψ(z,t)/∂z|β0|ψ (z,t)|and | ψ (z,t)/∂ t|ω0|ψ (z,t)|. Both inequali-
ties are guaranteed whenever the optical frequency bandwidth, , is
much less than the carrier frequency, ω0.
Within the SVEA, we have
iψ(z,t)
z=Hψ(z,t), (3)
where the (Hamiltonian operator) H= −Pn=2inβn
n!
n
tn. Alternatively,
Equation (3) can be rewritten in the frequency domain by Fourier trans-
formation,1
i˜
ψ(z,ω)
z=˜
H˜
ψ(z,ω), (4)
where ˜
ψ(z,ω) is the Fourier transform of ψ(z,t)and
˜
H= −X
n=2
βn
n!ωn. (5)
1In the following, the angular frequencies are referred at the baseband. They are referred to the optical
frequencies, ω0, by a shift ω0=ω+ω0.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 5 — #5
First-Order Approximation: Space-Time Analogy 5
This equation can be easily integrated as
˜
ψ(z,ω) =exp "iX
n=2
βnz
n!ωn#˜
ψ(z=0, ω). (6)
Therefore, the dispersive medium acts as a phase-only spectral filter.
Sometimes, it is useful to write the dispersive terms in the more compact
form 8n=βnz. When n=2, β2is called the group velocity dispersion
(GVD) coefficient, and 82the group delay dispersion (GDD) param-
eter. Analogously, when n=3, β3is the third-order dispersion (TOD)
coefficient.
3. FIRST-ORDER APPROXIMATION: SPACE-TIME ANALOGY
We are particularly interested in the case in which only the first term of
the operator contributes,
iψ(z,t)
z=β2
2
2ψ(z,t)
t2. (7)
Since it implies a second-order expansion in Equation (2), the medium is
said to be parabolic. This first-order approximation is physically plausi-
ble whenever 3|β23|. As an example, for standard single-mode
fiber (SMF) and a waveform centered in the telecommunication wave-
length (λ0=1.5 µm), the fiber coefficients are β2= −21.7 ps2/km and β3=
0.1 ps3/km. The signal should then have an optical bandwidth shorter
than 100 THz.
Equation (7) is a Schr¨
odinger-like equation for a free particle, ubiqui-
tous in Physics scenarios. In particular, this equation is mathematically
identical to that describing the one-dimensional (1D) scalar diffraction of a
paraxial monochromatic beam propagating in the z-direction (Goodman,
1996)
iUe(z,x)
z= − 1
2k0
2Ue(z,x)
x2, (8)
where Ue(z,x)denotes the transversal profile of the 1D beam and k0the
wave number.
The mathematical similarity between the Equations (7) and (8) is what
we know as the space-time analogy, mentioned in the Introduction. Table 1
summarizes the transfer rules connecting both domains. This connection
between diffraction and dispersion was found independently by two dif-
ferent groups at the end of the sixties (Akhmanov, Sukhorukov, & Chirkin,
1969;Treacy,1969). On the other hand, Papoulis had pointed out a formal
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 6 — #6
6Space-Time Analogies in Optics
TABLE 1 Space-Time analogy transfer rules
Space Time
Description Variable/Parameter Variable/Parameter Description
Position x t Proper time
Spatial frequency 2πuω(Baseband) angular
frequency
Wave number11/k0β2GVD coefficient
Paraxial propagation
factor (distance z)
exp[i2π2zu2/k0] exp[i82ω2/2] First-order dispersion
(GDD parameter 82)
Spatial lens factor
(focal length f)
exp[ik0x2/(2f)] exp[iKt2/2] Time lens factor
(chirping rate K)
similarity between diffraction and chirp radar, including the equivalent of
spatial lenses (Papoulis,1968a). Later on, we found in the literature some
theoretical research work performed by Saleh and Irshid (1982) about an
extension into the temporal domain of the Collet-Wolf equivalent theo-
rem regarding spatially partially coherent light (Mandel & Wolf,1995). In
the same decade, it is worth mentioning the temporal equivalents of the
Talbot effect proposed by Jannson and Jannson (1981) and spatial Fourier
transformation (Jannson,1983). However, it was not until the pioneering
work of Kolner and Nazarathy (1989) who, inspired on developing pulse
compression techniques based on electro-optic phase modulators (Kolner,
1988), developed a formal treatment of this analogy to include what we
know today as “time lenses”. It became then evident that a huge avenue
of temporal equivalent systems for ultrafast signal processing was fea-
sible to build, given the instrumentation available at that time (see e.g.,
Lohman & Mendlovic,1992;Kolner,1994a;Godil, Auld, & Bloom,1994;
Papoulis,1994;Mendlovic, Melamed, & Ozaktas,1995).
4. ELEMENTS AND THEIR IMPLEMENTATIONS
4.1. Temporal ABCD Matrices
In the previous section, we have introduced the formalism of the space-
time analogy and advanced that it can be extended to include other
photonic components apart from first-order dispersive media, such as
time lenses. In this section, we provide a unified formalism to describe
the linear distortion of the pulse envelope in a system composed by con-
catenating different elements that are susceptible to be described within
the framework of this analogy: the so-called “Gaussian” systems. Each
element in a Gaussian system is mathematically characterized by a uni-
tary 2 ×2 matrix. The whole system is quantified by a matrix calculated
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 7 — #7
Elements and their Implementations 7
by the multiplication, in the right order, of each of the elements that com-
pose it. This formalism is well known in spatial first-order Fourier optics,
which works in the paraxial regime (Siegman,1985;Collins,1970;Palma &
Bagini,1997), and has been adapted into the temporal domain (Dijaili,
Dienes, & Smith,1990;Nakazawa et al.,1998;Mookherjea & Yariv,2001).
Concretely, the action of any linear system on the input complex
envelope of a short light pulse, ψin(t), can be characterized as a linear
superposition
ψout(t)=Zψin (t0)K(t,t0)dt0, (9)
where the system is described by the kernel K(t,t0)and ψout(t)denotes
the output complex envelope. In the case in which the linear system is
Gaussian, the Kernel takes the following form
K(t,t0)=
qi
2πBexp hi
2BAt02+Dt22tt0iif B6= 0
q1
Aexp hiCt2
2Aiδ(t0t/A)if B=0
. (10)
Here the constants A,B,C, and Daccount for the system’s matrix coeffi-
cients. The reason why these systems are called Gaussian is because of the
quadratic dependence in the exponential term.
4.2. Spectral Dual Formalism
In temporal optics, there are some situations in which the Fourier trans-
form of the envelope, ˜
ψ(ω), may be the physical magnitude of interest,
rather than ψ(t). Of course, they are connected each other by a Fourier
transform relation and both carry the same quantity of information. Since
any linear system in time is linear in frequency too, it is useful to pro-
vide a similar analysis of Equations (9) and (10) in the dual space. By dual
we mean that these devices behave identically from a mathematical point
of view, but their action is performed in the Fourier domain (Papoulis,
1968b). Thus, we can write
˜
ψout(ω) =Z˜
ψin0)˜
K,ω0)dω0, (11)
where the Kernel in the frequency domain forms a Fourier transform pair
with the Kernel in time, that is,
˜
K,ω0)=ZZ K(t0,t00)exp[it0ω0t00)]dt0dt00. (12)
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 8 — #8
8Space-Time Analogies in Optics
For the particular case in which the system is Gaussian, it becomes Gaus-
sian in the spectral domain too and therefore a similar structural form of
Equation (10) holds,
˜
K,ω0)=
qi
2πBωexp hi
2BωAωω02+Dωω22ωω0iif Bω6= 0
q1
Aωexp hiCωω2
2Aωiδ(ω0ω/Aω)if Bω=0
. (13)
Of course, the dual coefficients are related to the matrix parameters A,B,
C, and Din the time domain. By inserting Equation (10) into (12), we easily
obtain
AωBω
CωDω=A B
C D1
=DC
B A . (14)
Equations (12) and (14) imply that the action of a Gaussian system in the
spectral domain is mathematically identical to the action of a Gaussian
system in the temporal domain whose matrix elements are provided by
the inverse matrix. The implications of this statement will become clearer
in the following sections.
4.3. Basic Photonic Components
Within the previous matrix formalism, we now proceed to describe the
mathematical structure of some devices that will probe very useful in tem-
poral optics applications and describe briefly their implementation with
current technology.
4.3.1. Group Delay Dispersion (GDD) Circuit
As previously advanced, a GDD circuit is an element designed to
introduce a quadratic phase factor in the spectral domain, ˜
ψout(ω) =
exp[i82ω2/2] ˜
ψin(ω), where 82is the GDD coefficient. The corresponding
matrix is
A B
C D=182
0 1 . (15)
These photonic components, as highlighted in Table 1, constitute the tem-
poral equivalent of the paraxial diffraction. However, we must note that
while diffraction only takes place for positive wave numbers in tempo-
ral optics, the GDD parameter can be positive or negative, depending
on the material component, waveguide structure, and the signal’s carrier
frequency. This subtleness certainly opens exciting new possibilities for
ultrafast signal processing.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 9 — #9
Elements and their Implementations 9
4.3.1.1. Possible Implementations
There exist many different photonic devices that can act as a GDD circuit.
Table 2 summarizes some of the most commonly used in the telecommu-
nications band. The most popular solution is the single-mode fiber. In this
case, the dispersion comes from the frequency dependence of the refrac-
tive index of the material, and the GDD parameter is 82=β2z, where β2
is the GVD parameter. Therefore, the GDD can be easily controlled with
the fiber length. Alternatively, dispersion compensating fibers (DCFs) can
be used. Thanks to the special core structure design, with shorter length
higher dispersion can be achieved. Both SMF and DCF are especially
attractive devices to achieve, when needed, high dispersion amounts.
Besides, it becomes particularly interesting that Raman amplification can
be implemented together with dispersion in the same fiber link.
Fiber solutions however present technical difficulties owing to the rel-
atively high nonlinear parameter, which makes it difficult to handle them
for dispersion management using powerful broadband pulses. In this
case, alternatives such as grating pairs or prisms can be more suitable
candidates. In other situations, it is interesting to have more compact
dispersive circuits and/or capable to operate in broader or just differ-
ent spectral windows. For this aim, different proposals including chirped
Bragg gratings inscribed in silica glass or silicon-on-insulator (SOI) have
received increasing attention in the last years. In this case, the GVD
coefficient can be engineered with a proper choice of the waveguide
dimensions and structure. The low nonlinearity and compactness of these
devices make them highly suitable for several optical communication
applications. Other alternatives such as photonic crystal fibers (PCFs)
are very attractive candidates for engineering the GDD amount over a
much broader spectral range, but the nonconventional core size, together
with the difficulty to fabricate them, imposes significant constraints that
prevent them to use in optical communication systems, where other
dispersion compensator devices easier to manage are readily accessible.
4.3.2. Temporal Lens
A temporal lens consists of any element that imprints a temporal quad-
ratic phase factor on the envelope of a light pulse without altering the
intensity profile, that is, ψout(t)=exp[iKt2/2]ψin(t). This lensing device
performs the same operation as a conventional lens does it on a spatial
light beam, that is, it modifies quadratically the phase of the input wave
front. Taking into account Equations (9) and (10), we can conclude that the
matrix of a temporal lens has the form
A B
C D=1 0
K1. (16)
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 10 — #10
TABLE 2 Several examples of photonic devices acting as GDD circuits in the telecommunications band (λ=1.5 µm). For the grating pair,
the formula parameters mean bthe spacing between gratings, dthe period and γthe incidence angle. For the linearly chirped fiber gratings
(LCFGs), ngis the group index, 3lthe grating period, and Lthe fiber length. Typical achievable numerical values are shown as example.
GDD 82Bandwidth
Parameter Typical value (ps2) Parameter Range (THz) Tuning Ref.
SMF β2z∼ −2,500 β2/(10β3)100 No [a]
DCF β2z∼ −2,500 β2/(10β3)100 No [b]
Grating pair λb
2πc2λ
d21
1λ
dsin γ2!3/20,29 c
5λ0
1λ
dsin γ2!
1+λ
dsin γsin2γ12 Yes [c]
LCFGs λ2ng3l/(πc2)[2,500 2,500] c3lL22.8 (C-band) Yes [d]
Integrated gratings n/a [9 9] n/a 0.51.5 No [e]
[a]Agrawal [2007]
[b]Gr ¨
uner-Nielsen, Wandel, Kristensen, et al. [2005]
[c]Weiner [2009]
[d]Venghaus [2006]
[e]Tan, Ikeda, Saperstein, et al. [2008]
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 11 — #11
Elements and their Implementations 11
The coefficient Kis a real number called chirping rate. This parameter
is closely related to the concept of the focal length (Kolner,1994b), and it
accounts for the quantity of spectral broadening that a pulse acquires after
time lensing. The higher the chirping rate the higher the broadening is,
regardless of the sign of K.
4.3.2.1. Possible Implementations
Electro-Optic Phase Modulation The most straightforward way to
imprint a desired quadratic phase modulation on a short light pulse
is using an electro-optic phase modulator (EOPM). An EOPM is usu-
ally built with a nonlinear crystal having a high electro-optic coefficient,
such as LiNbO3. By applying an electrical radio-frequency (RF) signal
on the crystal, the optical field propagating through acquires a phase
proportional to the applied voltage. If the electric field is just com-
posed by a single-tone RF waveform and assuming the optical pulse
is synchronized with the maximum of the RF signal, we get ψout(t)=
exp[icos(2πfrt)]ψin(t). Here, fris the frequency of the electrical wave-
form and is a real constant, called the modulation index, propor-
tional to the peak-to-peak voltage of the applied waveform, V0,=
πV0/(2Vπ). The coefficient Vπestablishes the voltage needed to acquire
aπ-phase shift and depends on the crystal material and waveguide struc-
tural parameters. For standard LiNbO3waveguide in travelling-wave
configuration, it is possible to achieve typical values around Vπ5V.
To relate these parameters with the time-lens chirping rate, we further
require the temporal pulse width to be shorter than the period. In this case,
we can approximate ψout(t)=exp[i ] exp[i (2πfrt)2/2]ψin (t). Then,
by direct inspection we get K= −4π2f2
r.
In radar literature, the capabilities of quadratic phase modulation to
increase the bandwidth of a temporal signal were very well known
(Klauder et al.,1960), even the analogy of chirping with spatial lenses was
reported as early as in the sixties (Papoulis,1968a). In optics, it was also
known that by chirping a light pulse (Giordmaine, Duguay, & Hansen,
1968) or a continuous-wave laser (Bjorkholom, Turner, & Pearson,1975)
one could enhance its optical spectrum and subsequently obtain a shorter
pulse by properly compensating for the instantaneous frequency chirp
with an adequate passive dispersive medium (Treacy,1968). The merit
of Kolner’s works on time lenses (Kolner,1988;Kolner & Nazarathy,
1989) was that of unifying these discoveries and putting them in the more
general framework of the space-time analogy, opening a door to new,
more sophisticated, ultrashort pulse processing schemes such as temporal
imaging (Bennett, Scott, & Kolner,1994).
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 12 — #12
12 Space-Time Analogies in Optics
The advantages of using the EOPM time lens include wavelength trans-
parency, high-speed operation, and easiness of integration with optical
communication components. On the other hand, a simple Taylor expan-
sion analysis on the above approximation concludes that the pulses over
which the EOPM time lens operate must satisfy σ0(πfr)1, where σ0
is a measure of the pulse duration. This restricts the EOPM time lens to
operate over low duty cycle (typically <30%) pulse trains (Kolner,1994a).
Longer pulses get affected by the higher order (nonquadratic) terms in
phase modulation. In the context of the space-time analogy, this devia-
tion from the ideal quadratic operation can be understood as the temporal
equivalent of a lens’ spatial aberrations, a subject that has been extensively
studied in the literature (Bennett & Kolner,2001;Marinho & Bernardo,
2006).
To extend the time window and chirping rate, alternative solutions
involving nonlinear phenomena have been reported, which we explain
briefly in the following paragraphs.
Cross-Phase Modulation (XPM) An elegant all-optical alternative for
time lensing consists of using a highly nonlinear fiber. The idea is that two
equally polarized optical fields are nonlinearly coupled due to the Kerr
effect (Agrawal,2007), which is a phenomenon based on the third-order
nonlinear susceptibility, χ(3)
, and therefore takes place in all dielectric
materials. In the simplest case where dispersion, losses, and walk-off are
ignored, the output pump envelope, ψp, out(t), and signal’s, ψs,out(t), are
analytically related to their input complex envelopes
ψs,out(t)=ψs,in (t)exp[iγL(|ψs,in(t)|2+2|ψp,in (t)|2)], (17a)
ψp,out(t)=ψp,in (t)exp[iγL(|ψp,in(t)|2+2|ψs,in (t)|2)]. (17b)
Here, γis the nonlinear Kerr coefficient and Lthe length of the nonlinear
medium. We are assuming the envelope square has dimensions of power.
We shall focus on the particular case in which the signal is weak, so that
the first term in the exponential (self-phase modulation), in Eq. (17a), can
be neglected when compared to the other one (cross-phase modulation).
In this case, ψs,out(t)=ψs,in (t)exp[i2γL|ψp,in(t)|2], so that the pump field
modifies the instantaneous phase of the signal, regardless of the shape of
the input pulse. For the particular case in which the pump field is Gaus-
sian, ψs,out(t)=ψs,in (t)exp[i2γLP0exp(t2/T2
0)], where P0is related to the
pump’s peak power and T0is a parameter related to its duration. If the
temporal width of the signal is much shorter than the temporal width
of the pump’s, ψs,out(t)ψs,in (t)exp[i2γLP0] exp[i2γLP0t2/T2
0]. In this
case, by direct inspection we get K= −4γLP0/T2
0. The temporal window
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 13 — #13
Elements and their Implementations 13
over which the phase modulation remains approximately quadratic
equals the pump pulse duration T0. The main advantage of using XPM
as a medium for time lensing is the fact that, for some materials, the Kerr
coefficient, γ, can be extremely large.
XPM as a mechanism for time lensing was firstly reported by Moura-
dian et al. (2000), who used a pump pulse with Gaussian intensity profile.
However, one can easily realize that by using preshaped parabolic-like
pump pulses (Finot et al.,2009), the temporal aberrations are substan-
tially removed. This XPM time lens with a parabolic-pump configuration
has been recently reported, achieving a nearly aberration-free operation
(Hirooka & Nakazawa,2008;Ng et al.,2008).
Sum Frequency Generation (SFG) If one relaxes the wavelength
transparency operation of the time lens, parametric phenomena can be
used to achieve the desired time lens operation. Sum-frequency genera-
tion, a second-order nonlinear phenomenon, was indeed exploited in the
first configurations for temporal imaging (Bennett, Scott, & Kolner,1994;
Bennett & Kolner,1999). The idea consists of mixing the pulse over which
we want to imprint the chirping, ψs,in(t), with a pump pulse that has
been previously dispersed in a GDD circuit, ψp,in(t). Because of the dis-
persion, and assuming a pump pulse with Gaussian profile, we can write
(Agrawal,2007)
ψp,in(t)exp "t2
2T2
z#exp "iCzt2
2#. (18)
Here, the temporal duration and chirping rate of the pump pulse are con-
trolled by the GDD amount of dispersion, that is, Tz=T0[1 +(82/T2
0)2]1/2
and Cz=82/(T4
0+82
2).T0is the initial pump pulse duration. After the
predistored pump pulse is mixed with the signal pulse, we look at the
sum-frequency component, which is isolated using a spatial or frequency
filter, depending on the chosen configuration for SFG. Assuming ideal
phase-matching conditions and ignoring pump depletion and walk-off in
the nonlinear medium, the complex envelope for the up-converted field is
ψs,out(t)ψp,in (ts,in(t)=exp "t2
2T2
z#exp "iCzt2
2#ψs,in(t). (19)
By direct inspection, we conclude that the chirping rate for a time lens
implemented with SFG is Cz, over a time window Tz. However, the tem-
poral duration of a parametric time lens and the chirping rate cannot be
tuned independently. For example, at 82=T2
0, the chirping rate achieves
its maximum possible value, but the temporal window is only Tz=2T0.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 14 — #14
14 Space-Time Analogies in Optics
On the other hand, it is very important to note that in this parametric
case, the chirping rate coefficient does not depend on the nonlinear crys-
tal parameters but on the pump pulse duration and dispersion amount. If
one is able to balance the trade-off between the chirping rate and temporal
window, there are benefits of using parametric processes for time lensing.
The first configurations demonstrated for SFG-based time lensing
involved the use of beta barium borate, leading to phase matching over
a huge bandwidth, albeit with a rather bulky setup (Bennett & Kolner,
1999). Recent advances in quasi-phase matching techniques in waveg-
uides (Yang et al.,2004) have enabled the generation of more compact
devices to produce SFG-based time lenses (Bennett et al.,2007).
Despite these advances, residual second-harmonic generation from the
pump pulse and the frequency shift of the chirped waveform make it
difficult to launch the up-converted signal into a subsequent optical sys-
tem for further processing that operates in the same spectral window. In
the next paragraphs, we review a recent alternative that overcomes these
difficulties, though still making use of a parametric nonlinear process.
Four-Wave Mixing (FWM) The configuration is very similar to the one
used for SFG, since it relies on mixing a predistorted pump pulse with the
signal over which we desire to imprint the chirp. Instead of using SFG,
which depends on the second-order nonlinear susceptibility and therefore
is only possible in certain materials, the idea is to exploit the four-wave
mixing effect, which is due to the third-order susceptibility and therefore
takes place in all dielectric materials (Salem et al.,2008). After conversion
in the nonlinear medium and assuming ideal phase-matching conditions,
the envelope of the idler wave becomes
ψs,out(t)ψ2
p,in(t
s,in(t)=exp "t2
T2
z#exp hiCzt2iψ
s,in(t). (20)
Now the chirping rate is 2Czand the time lens operates in an effective
time window Tz/2, thus leading to the same figure of merit as with
SFG. However, the benefits of FWM rely on the fact that the converted
wavelength is very near to that of the signal, so it becomes easier to per-
form further processing. Even more, with today’s technology, FWM can
be easily achieved in highly nonlinear fibers (Agrawal,2007;Saruwatari,
2000), nanophotonic wires built in silica (Foster, Turner et al.,2008), silicon
(Osgood et al.,2009), highly nonlinear chalcogenide waveguides (Pelusi
et al.,2008) or even doped silica glass (Ferrera et al.,2008), and all of them
are becoming more used for advanced optical communication applica-
tions. It is important to note that, unlike with the previous configurations,
the FWM time lens also conjugates the input complex field. This change
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 15 — #15
Elements and their Implementations 15
must be taken into account as it usually leads to different dispersion com-
pensation rules for performing optical signal processing applications like,
for example, temporal imaging (Kuzucu et al.,2009).
As shall become evident in the following sections, the possibility to
realize the time lens using FWM has permitted to implement for the first
time many schemes proposed in the framework of temporal optics for
ultrafast signal processing, with applications in modern optical commu-
nication systems. As with other parametric processes, we note that once
the phase-matching conditions are achieved, the chirping rate parameter
only depends on the settings of the pump pulse. However, a word of cau-
tion must be brought up here. Although a broad tuning range of chirping
rates can be achieved with parametric time lenses, the fact that the pump
pulse needs to be predistorted imposes a fundamental limit on the maxi-
mum speed at which devices based on these time lenses can work. This is
the reason why most of the reported schemes implement the processing
at a relatively low speed (megahertz rate). As the optical communication
field demands for higher operation speeds, further research work should
be done to upgrade the time lens operation.
Table 3 summarizes the main characteristics of the different devices
proposed here to work as time lenses. Figure 1 represents schematically
the possibilities for implementing the earlier mentioned time lenses.
4.3.3. Temporal Gaussian Modulator
This device filters an input signal in the temporal domain, so that the
output pulse envelope is given by ψout(t)=exp[02t2/2]ψin (t), where
01is a real constant related to the gate duration. This device can be
TABLE 3 Several physical principles upon which time lenses can be constructed.
Some typically achievable rates and time apertures are shown as an example, as well
as the definition of a figure of merit (FOM) aimed to account for the time aperture and
the power of the chirping lens.
Chirping KTemporal window T
Typical
Parameter value (ps2)Parameter Range (ps) FOM (|K|T2)
EOPM 4π22f2
r–0.5 (3fr)130 42
XPM 4γLP0/T2
0–5 T0100 4γLP0
SFG 82/82
2+T4
0–0.1 T0
1+ 82
T2
0!2
1/2
3 1
FWM 282/82
2+T4
0–0.2 T0
2
1+ 82
T2
0!2
1/2
2 1
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 16 — #16
16 Space-Time Analogies in Optics
Input pulse Output pulse
XPM
Input pulse
Probe pulse
Output pulse
SFG
Probe pulse
Output pulse
FWM
Input pulse
Input pulse
Probe pulse
Output pulse
GDD circuit
GDD circuit
*
(a) (c)
(
b
)(
d
)
EOPM
FIGURE 1 Different schemes for implementing a temporal lens: (a) electro-optic
phase modulation (EOPM); (b) cross-phase modulation (XPM) in a nonlinear medium;
(c) sum frequency generation (SFG); and (d) four-wave mixing (FWM). The solid lines
mean amplitude and the dashed ones phase. The star represents complex
conjugation.
understood as a temporal lens with an imaginary chirping rate. Thus, the
matrix associated is
A B
C D=1 0
i021. (21)
This action can be readily achieved with an active amplitude modulator
providing Gaussian-like modulation, such as an electro-optic modula-
tor (EOM) in a push–pull configuration or an electroabsorption modulator
(EAM). For these cases, assuming that an external clock signal with fre-
quency frdrives the devices, the typically achieved values for the pulse
gate are around 1/(3fr)for EOMs and 1/(4fr)for EAMs.
4.3.4. Spectral Gaussian Modulator
This device filters spectrally the complex field envelope as ˜
ψout(ω) =
exp[12ω2/2] ˜
ψin(ω), where 1is a real parameter related with the spectral
width. The corresponding matrix is
A B
C D=1i12
0 1 . (22)
Spectral Gaussian filters can be conveniently fabricated using fiber Bragg
grating devices, microring resonators, Fabry-Perot cavities, followed by a
long etcetera. With optical communication instrumentation, today we can
find suitable filters satisfying quite demanding constraints.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 17 — #17
Elements and their Implementations 17
4.4. Some ABCD Matrix Properties
The aim of this section is to enumerate some of the most important char-
acteristics of the temporal ABCD matrix systems, which will provide
very useful to understand the properties of the processing systems to be
described in the next section. These characteristics are adapted from the
analysis performed in the spatial domain (Siegman,1985;Collins,1970;
Palma & Bagini,1997).
1. The determinant of an ABCD matrix satisfies AD BC =1.
2. Any Gaussian system having a matrix coefficient B=0 conjugates
the input and output, meaning that the output intensity becomes a
scaled replica of the input, with a scaling factor being provided by the
matrix coefficient A, that is, Iout(t)=1
|A|Iin(t/A). These are the so-called
temporal imaging systems (TISs).
3. A Gaussian system having C=0 behaves as a temporal afocal system.
4. Combining 2 and 3, and given the spectral formalism developed in
Section 4.2, with Equation (14) in particular, we conclude that any
temporal imaging system is a spectral afocal system and vice versa.
5. Conversely to 4, any afocal system (C=0)behaves as a spectral imag-
ing system (Bω=0), meaning that the output spectrum of a signal
that has traversed a temporal afocal system constitutes a scaled replica
of the input’s optical spectrum, with a scaling factor given by D(or
Aω), that is, Sout(ω) =1
|D|Sin/D). These systems are called spectral
imaging systems (Torres-Company, Lancis, & Andr ´
es,2007a).
6. Any Gaussian system having A=0 is an optical Fourier transformer,
meaning that the output intensity pulse becomes a scaled replica of the
optical spectrum of the input pulse. In this case, the scaling factor is
provided by the matrix coefficient B, that is, Iout(t)=1
|B|Sin/B).
7. Following the dual analysis, a system having Aω=0 (D=0)consti-
tutes a time-to-frequency transformer, where the optical spectrum of
the output signal becomes a scaled replica of the input intensity profile,
with a scaling factor provided by the matrix coefficient Bω(i.e., C). In
mathematical terms, Sout(ω) =1
|C|Iin(t/C).
8. Gaussian systems having B=0 and C=0 perform simultaneously a
temporal and spectral imaging of the input signal, which is only possi-
ble if both the amplitude and phase are scaled. These are called scaling
systems and perform a temporal image of the input’s complex field.
The temporal scaling factor is provided by the matrix coefficient A.
Therefore, ψout(t)=1
|A|1/2ψin(t/A).
9. Any Gaussian system with Adifferent from zero can be described
with the combination of three equivalent systems: a temporal lens
with equivalent chirping coefficient Keq = −C/A, a scaling system with
equivalent scaling factor meq =A, and a GDD circuit with equivalent
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 18 — #18
18 Space-Time Analogies in Optics
GDD coefficient 82eq =B/A. This can be easily seen by the fact that any
2×2 matrix can be rewritten as (see, e.g., Lancis et al.,2004)
A B
C D=1 0
C/A1A0
0 1/A1B/A
0 1 . (23)
5. COHERENT ULTRA-HIGH-SPEED OPTICAL SYSTEMS
AND THEIR APPLICATIONS
In the last section, we reviewed some of the most widely used pho-
tonic components in the framework of the space-time analogy. The aim
of this section is to provide a comprehensive review of the systems built
using those photonic components and pointing out the unique capabilities
which make them so attractive in a rich variety of fields.
5.1. Tunable Delays
Many applications today require an optical delay line providing an accu-
rate and programmable temporal delay with very fine resolution. The
most widespread solution relies on using precise mechanical translation,
where the timing of a signal is controlled by varying its free-space path
length. In the framework of the space-time analogy, we can envision that
introducing a controllable delay on a short light pulse is fully equiva-
lent to introduce a controllable lateral shift on a 1D beam (van Howe &
Xu,2005a). The latter scheme can be accomplished by a sequence of two
prisms. The first prism tilts an incoming light beam which, after prop-
agating in free space a certain distance, is redirected towards the same
direction as it had in the beginning. There appears a lateral shift with
respect to the input beam, which is proportional to the travelled distance
and the angle tilt introduced by the prisms. The counterpart temporal sit-
uation is a widespread technique and it involves three main operations.
First, a wavelength conversion scheme changes the carrier frequency of
the incoming light pulse whose temporal delay is to be controlled. This
element plays the role of a “temporal prism” since a change in the carrier
frequency of a pulse is equivalent to a change in the propagating wave
number. The second element is a GDD circuit, in complete analogy with
the action of diffraction. And finally, a wavelength reconverter shifts back
the frequency of the pulse to its original value, in the same way the last
prism leads to a beam whose propagation direction is parallel to that of
the input beam. This general arrangement is known in the literature as the
“wavelength conversion and dispersion” scheme and it has been widely
used to achieve fractional delays exceeding 1000 (Sharping et al.,2005) at
high speed and in a format transparent operation (Nuccio et al.,2010).
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 19 — #19
Coherent Ultra-High-Speed Optical Systems and their Applications 19
Frequency
shifter
Time
GDD circuit Φ2
T
GDD circuit Φ2
Δω
FIGURE 2 Scheme for understanding the role of the frequency shift in an optical
delay line. Two pulses with identical complex envelopes but with different frequency
carrier suffer the same distortion in a GDD circuit but undergo a delay proportional to
the difference between their frequencies and the GDD amount.
To understand mathematically the temporal delay that can be achieved
from the previous discussion, we make use of the system sketched in
Figure 2. A short light pulse is split into two. The copy travelling through
the upper arm is shifted in frequency by a quantity , whereas the
copy travelling through the lower arm remains at the same fixed car-
rier frequency. We now consider both pulses travelling through identical
GDD circuits. Taking into account the Equations (9),(10), and (15), the
corresponding output pulse envelopes transform as
ψdown(t)=exp it2
282!Zψin(t0)exp it02
282!exp itt0
82dt0, (24)
for the wavelength-preserved pulse and
ψup(t)=exp it2
282!Zψin(t0)exp it02
282!exp it0t
82dt0,
(25)
for the wavelength converted one. By comparing Equations (24) and (25),
we conclude that, after ignoring an irrelevant constant phase factor,
ψup(t)=exp(itdown (tT). This result indicates that by shifting
the carrier frequency, a short light pulse gets delayed after propagating
in a GDD circuit. The delay, T, is measured with respect to the unshifted
version and its magnitude is given by
T=82. (26)
Note that this time delay only depends on the GDD parameter and the
frequency shift. The aim of using a second temporal prism is to shift the
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 20 — #20
20 Space-Time Analogies in Optics
carrier frequency back to the original value, so that
ψup(t)=ψdown (tT). (27)
This equation indicates that by introducing a wavelength change and let-
ting the pulse spread in a GDD medium, a controllable time delay can
be easily achieved. Unfortunately, the envelope gets distorted owing to
GDD, and for long delays (large amounts of dispersion) it needs to be
compensated for (van Howe & Xu,2005b).
To derive the above equations, we have assumed that the dispersive
medium introduces quadratic dispersion only (i.e., third- and higher dis-
persion terms can be ignored) and that the temporal prisms do not change
the envelope of the signal. These are very restrictive assumptions and,
in practice, are the main limiting factors that prevent to achieve signifi-
cant delays in the optical domain. The tuning capabilities will be satisfied
whenever the time prism introduces a controllable frequency shift or
when the dispersive medium introduces a tunable GDD amount. The first
feature can be achieved easily with an EOPM. Using a similar argument
as the one described in Section 4.3.2.1, an EOPM driven by an electrical
sinusoidal signal can introduce the required linear temporal phase if the
input optical pulse is properly aligned and is short enough in duration
(van Howe & Xu,2005a). Although this procedure introduces a highly
accurate and electrically controllable frequency shift, its magnitude is only
of few gigahertz and therefore achieving larger delays becomes challeng-
ing with this time prism implementation. An alternative procedure is to
introduce wavelength conversion making use, for example, of the FWM in
either highly nonlinear fiber (Sharping et al.,2005), a silicon nanowaveg-
uide (Okawachi et al.,2008), or SFG in a periodically poled lithium niobate
waveguide (Langrock et al.,2006;Wang et al.,2007). With these devices,
tunable delays of 0.8 ns (Sharping et al.,2005), 243 ns (Okawachi et al.,
2008), and 44 ns (Wang et al.,2007) have been reported for 10 Gb/s non-
return to zero (NRZ) signals. Distortion owing to GDD can be effectively
precompensated in these schemes.
It is worth mentioning that by cascading several wavelength converter
devices, tunable delays in the microsecond range have been recently
reported (Alic et al.,2010;Dai et al.,2010), making this technique more
attractive than other alternatives including dispersive resonances or slow
light, where dispersion management becomes a significant issue.
5.2. Real-Time Fourier Transformation
Real-time Fourier transformers (RTFTs) are optical devices capable to
modify the light intensity profile of an ultrashort light pulse so that the
achieved shape becomes a scaled version of the input coherent optical
spectrum. Two main approaches are proposed for this aim.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 21 — #21
Coherent Ultra-High-Speed Optical Systems and their Applications 21
5.2.1. Optical Fourier Transformation Technique
This system was originally proposed by Jannson (1983) and corresponds
to the temporal counterpart of the classic Fourier transformer constituted
by a spatial lens and diffraction until the focal length (Goodman,1996), as
shown in Figure 3.
Making use of the ABCD matrix approach, one can calculate the ABCD
matrix corresponding to the situation in Figure 3(b). Imposing A=0
requires having K82=1, that is, the chirp rate introduced by the time lens
must be completely compensated for. In this situation, the output intensity
profile resembles the shape of the input optical spectrum, with a scaling
Lens
x
L
x
Focal plane
(a)
(b)
GDD circuit Output intensity
Time
Input intensity
Temporal lens
Time
f
FIGURE 3 An optical Fourier transformer is a device that implements physically the
Fourier transform of an input signal. In (a), the idea is represented in the spatial
domain, where an input beam is focused by a lens. The intensity distribution at the
focal plane is the modulus square of the Fourier transform of the transversal profile of
the input beam. In (b), we have the temporal counterpart situation, where the output
pulse becomes a replica in intensity of the input optical spectrum after traversing an
ultrafast optical Fourier transformer.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 22 — #22
22 Space-Time Analogies in Optics
factor provided by the GDD amount. This system can be used to compress
and reshape ultrashort pulses, provided a suitable spectral shaping device
is introduced before the optical Fourier transformation technique. In this
case, the temporal width can be tuned by a proper selection of the tempo-
ral chirp and the GDD parameter, and even made to occupy less than one
bit period in a stream of pulses. It is important to note that this system does
not constitute a time-to-frequency converter, that is, the output spectrum
of the pulse is not, in general, a scaled replica of the input intensity pro-
file because the element Dof the resulting matrix is not zero. The physical
reason behind is that the output optical envelope appears multiplied by a
quadratic phase factor in time.
However, a simple modification of this optical Fourier transformer
can be done to achieve time-to-frequency conversion as well (Lohman &
Mendlovic,1992). It involves the concatenation of a GDD circuit with
identical GDD parameter, sandwiching the time lens. In this case, the
Fourier transformation rule remains the same, K82=1, but the chirp-
ing induced by the dispersion in the first GDD circuit compensates for
the chirp acquired through the time lens, so that the pulse at the tem-
poral focus becomes transform limited. It is easy to show that in this
case D=0, meaning that this system also performs time-to-frequency
conversion.
In any case, the most important characteristic of an optical Fourier
transformer (OFT) is the fact that the output intensity pulse is a scaled ver-
sion of the optical spectrum regardless of the input spectral phase. This
feature has provided the key to compensate for the spectral phase dis-
tortions in ultra-high-speed optical time-division multiplexing (OTDM)
linear systems (Hirooka & Nakazawa,2006). The main limitation is that
the spectral phase distortions must not lead to a pulse with a temporal
structure larger than the temporal aperture of the OFT’s time lens. At that
work, 160 Gb/s RZ differential phase-shift keying (DPSK) signals were
demultiplexed to 40 Gb/s and transmitted through a DCF-compensated
1000 km fiber link with the OFT technique. The OFT results show that
broadening due to remaining GDD and part of the TOD can be fully
compensated for, as well as reducing timing jitter (note that all these
effects imply spectral phase distortion). This regeneration property led to
a 2-dB improvement in bit error rate performance at 109, very close to
the spontaneous emission limit.2Further achievements could be obtained
if the temporal aberrations in the EOPM-based time lens were fully
compensated for.
2We remind the reader that the usual protocol in optical communication systems for considering “error
free” transmission is having a BER <109without applying forward error correction algorithms.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 23 — #23
Coherent Ultra-High-Speed Optical Systems and their Applications 23
5.2.2. Temporal Far-Field Phenomenon
In order to achieve an RTFT, the temporal lens is not really required. In the
spatial domain, it is well known that the diffraction pattern achieved at a
certain distance away from an input object, usually called the far-field dis-
tance, coalesces with the (modulus square of the) Fourier transform of the
object. In the temporal domain, this effect is well known (Fetterman et al.,
1979;Tong, Chan, & Tsang,1997), but the interpretation as a far-field or
Fraunhofer phenomenon was realized later (Muriel, Aza˜
na, & Carballar,
1999;Aza˜
na et al.,1999), an effect usually coined as frequency-to-time
mapping. The far-field or Fraunhofer condition translates to (Aza˜
na &
Muriel,2000)
82σ2
0/(4π), (28)
where σ0is a measure of the input pulse width. When this condition is
satisfied, the output intensity shape becomes a scaled replica of the input
optical spectrum too. As in the previous case, the scaling factor is also
given by the GDD coefficient. In general, the exact amount of dispersion
to reach the far-field depends on the specific input waveform, and Equa-
tion (28) is just a simple estimation. As sketched in Figure 4, once the
temporal far-field condition is practically satisfied, by adding larger dis-
persion amounts, the output intensity profile does not change its shape
but only its scale. This means that this specific RTFT brings the advantage
Input intensity
Time
GDD circuit
Output intensites
FIGURE 4 Schematic representation of the frequency-to-time mapping in a GDD
circuit or temporal far-field phenomenon. The intensity profile of a pulse at the output
of a GDD medium that satisfies the far-field condition becomes a scaled replica of the
input optical spectrum. Once the far-field condition is reached, by adding larger
dispersion amounts, the profile of the pulse does not change, only its scale becomes
larger.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 24 — #24
24 Space-Time Analogies in Optics
of resolution tuning simply by adding larger GDD amounts, since the
optical spectral resolution, δω, is provided by δω =δt/82, where δtis the
temporal resolution of the acquisition unit (assuming that the finest spec-
tral feature is adequetely mapped into the time domain). Thus, this system
is highly interesting for measuring spectra of optical pulses. The advan-
tage of using a GDD circuit and a high-speed photodiode (followed by
a high-speed digitizer) instead of a simple OSA relies on the fact that,
assuming after the photodiode there is a real-time digitizer, the optical
spectra can be measured in a pulse-by-pulse basis. In other words, the
acquisition speed is given by the repetition rate of the pulse train used.
However, given that the output stretched pulses must not suffer tempo-
ral overlapping, the repetition rate of the pulse train cannot be arbitrarily
increased and indeed should satisfy f< σ0/(4|82|). Assuming transform-
limited pulses, σ0=4, where is the spectral bandwidth of the input
pulse, and combining the above inequality with the resolution formula,
we get 1ω/δω < (fδt)1, indicating that there is a trade-off between the
maximum acquisition rate and the spectral resolution of the system. The
number of spectral features (the optical bandwidth divided by the spectral
resolution) is bounded by a quantity depending on the temporal resolu-
tion and repetition rate of the pulse train, regardless of the GDD amount.
Nevertheless, for typical subpicosecond optical pulses and high-speed
photodetectors, it is possible to acquire optical spectra with subgigahertz
spectral resolutions at megahertz acquisition rates. This is in contrast
to the much longer acquisition speeds (kilohertz rate in a best-case sce-
nario) in commercial OSAs offering similar resolutions, where mechanical
and/or electronic artifacts limit the acquisition speeds.
The frequency-to-time mapping technique has been used in many dif-
ferent areas. For instance, real-time spectral measurements were initially
applied in absorption spectroscopy (Kelkar et al.,1999), with dispersive
fiber as the GDD medium. More recently, researchers have incorporated
Raman amplification in the dispersive medium, thus compensating for
the energy spreading of the pulse as it stretches (Solli, Chou, & Jalali,
2008). Using this amplified Fourier transformation modality, real-time and
quasi real-time absorption spectroscopies have been reported at subgiga-
hertz optical resolution and few megahertz acquisition rate (Chou, Solli, &
Jalali,2008).
This technique has also impacted the field of radio-frequency arbitrary
waveform generation (RF-AWG). Here, the aim is to tailor the intensity
profile of a short light pulse to generate electrical signals on demand
after proper opto-electronic conversion with a high-speed photodiode
(Capmany & Novak,2007). Using conventional pulse shaping techniques
(Weiner,2000), it is difficult to achieve long temporal waveforms owing to
the relatively high spectral resolution of typical pulse shapers. In other
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 25 — #25
Coherent Ultra-High-Speed Optical Systems and their Applications 25
words, while it is relatively straightforward to achieve high-frequency
electrical signals using pulse shaping, for example, in the millimeter
wave (around 60 GHz and beyond; see McKinney, Leaird, & Weiner,
2002) or terahertz regime (0.1–3 THz, see Liu, Park, & Weiner,1996), it
becomes challenging to achieve the submillimeter wave regime. However,
once the temporal far field is achieved, the waveform does not change
by adding more dispersion, and it becomes relatively easy to achieve
submillimeter-wave electrical signals with the frequency-to-time mapping
method (Chou, Han, & Jalali,2003). The concept for RF-AWG is illustrated
in Figure 5. The scheme consists on synthesizing the optical spectrum of
a coherent broadband signal with a Fourier-transform pulse shaper and
later transferring the designed spectral shape into the electrical domain by
stretching the optical pulse in a GDD circuit and subsequently detecting
it. This technique has been used to generate high-speed electrical sig-
nals (in particular ultra-wideband, UWB) with a temporal profile meeting
very stringent regulations for wireless communications (Lin, McKinney, &
Weiner,2005). An experimental example is illustrated in Figure 6, show-
ing the synthesized spectrum and the corresponding achieved electrical
signal after stretching. Due to the high frequency content (spanning the
fs laser
Electrical signal
Pulse shaper
Reflective LCOS
display
Lens
Grating
Collimator
Circulator Mirror
Synthesized
ES
Dispersive element O/E
conversion
Time
FIGURE 5 Scheme for implementing an RF-AWG using the temporal far-field
phenomenon and a Fourier-transform pulse shaper as reconfigurable spectrum
synthesizer. The optical spectrum of a broadband input pulse is synthesized in a
user-defined fashion with a Fourier-transform pulse shaper. After the synthesis, the
pulse is launched into a GDD medium that satisfies the far-field condition, so that the
output intensity is a scaled replica of the engineered spectrum. The light is detected
with a high-speed photodiode, so electrical impulses with user-defined current
variation can be easily created.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 26 — #26
26 Space-Time Analogies in Optics
21012
Times (ns)
(a)
(b)
(c)
1530 1540 1550 1560 1570
Wavelength (nm)
1.25
(dBm)
0
0
12 34 567 8 910
Frequency (GHz)
2.5 5.0
50
100
FIGURE 6 Example of RF-AWG using the scheme of Figure 6. (a) synthesized optical
spectrum; (b) achieved waveform after frequency-to-time mapping conversion. The
frequency-to-time mapping operation is evident in this figure; (c) corresponding RF
spectrum. We observe that the frequency content of the synthesized electrical signal
is indeed very broad (spanning several gigahertz). (After Lin, McKinney, & Weiner,
2005). c
[2005] IEEE.
range 0–10 GHz), this kind of electrical waveforms is extremely diffi-
cult to generate using purely electrical approaches. Given the high level
of complexity in the pulse shapes that can be achieved with this tech-
nique, interesting applications have been reported, like compensation for
antenna dispersion (McKinney & Weiner,2006) or distortion in broadband
electrical amplifiers (Bortnik et al.,2006).
Finally, this frequency-to-time mapping technique also offers interest-
ing possibilities in the field of imaging. For instance, in Fourier-domain
optical coherence tomography (FD-OCT), the depth information of a
layered object is encoded in a spectral interferogram which is usually
recorded using a regular OSA. Replacing the OSA with the GDD circuit
and a high-speed digitizer permits to obtain the depth information of
layered tissues at a very high speed (typically megahertz rate) using the
same optical source. The use of the frequency-to-time mapping method for
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 27 — #27
Coherent Ultra-High-Speed Optical Systems and their Applications 27
FD-OCT was firstly reported by Moon and Kim (2006), using a dispersive
DCF as GDD circuit, leading to an axial rate of 5 MHz and 5-µm reso-
lution at 1.5-µm illumination wavelength. However, the signal-to-noise
ratio was still low to achieve a high-quality OCT image for more demand-
ing applications (to give an example, at a different wavelength, retinal
images require sensitivities as high as 100 dB). Subsequent works used
a linearly chirped fiber Bragg grating as GDD element (Park et al.,2007;
Saperstein et al.,2007) and Raman-amplified dispersive Fourier transfor-
mation (Goda, Solli, & Jalali,2008), leading to better SNRs, but still far
from the values offered by other advanced high-speed techniques involv-
ing Fourier-domain mode-locked lasers, which have recently reported
high sensitivity and megahertz acquisition rates in OCT with similar axial
resolutions (Wieser et al.,2010).
Nevertheless, this frequency-to-time technique still shows potential as
a high-speed version of an imaging modality called spectrally encoded
endoscopy (Tearney, Shishkov, & Bouma,2002). This imaging modality
uses a scheme identical to the Fourier pulse shaper depicted in Figure 5.
The idea is to replace the SLM at the focal plane of the pulse shaper by the
object whose image is to be measured. The reflectance of the object is thus
encoded in the optical spectrum of the broadband light source used and it
can be easily measured with an OSA and decoded offline. Using 2D spec-
tral dispersers (Xiao & Weiner,2004) and the frequency-to-time mapping
technique, the reflectance of a moving object can be acquired (Goda, Tsia,
& Jalali,2009). This device can capture 2D spatial profiles of objects mov-
ing at a speed as fast as the repetition rate of the mode-locked laser used.
Experimental results of 2D dynamic imaging (a metallic particle travelling
through a tube) are shown in Figure 7.
5.3. Time-to-Frequency Converters
A time-to-frequency converter is a device operating over an incoming
ultrashort pulse in such a way that the shape of the output spectrum
is a scaled version of the input temporal intensity (Kauffman et al.,
1994;Mouradian et al.,2000;Aza˜
na,2003a). This device constitutes a
viable alternative to measure temporal intensity waveforms with a simple
OSA. It corresponds to the dual configurations presented in the previ-
ous section. Similarly, there are two different approaches to achieve this
operation.
5.3.1. GDD Circuit with Time Lens
A time-to-frequency converter has a corresponding matrix whose spectral
element satisfies Aω=0. The simplest device to achieve this operation
is presented in Figure 8. It is easy to show that the time lens and GDD
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 28 — #28
28 Space-Time Analogies in Optics
Metal microspheres
(a)
(c)
(b)
Water
Flow
20 µm
20 µm
20 µm
20 µm
20 µm
20 µm
20 µm
t=7.335 µs
t=7.172 µs
t=7.009 µs
t=0.163 µs
t=2.934 µs
t=5.705 µs
t=8.476 µs
FIGURE 7 Two-dimensional image of a moving particle acquired with the ultrafast
camera implemented with the concepts of frequency-to-time mapping. The key relies
on encoding the image information in the frequency domain. Later, the spectrum is
mapped into the temporal domain so that the object information can be acquired
with a single high-speed photodiode and subsequent digital signal processing. The
frame at which the images are acquired is equal to the pulse repetition rate and can
be as high as a few megahertz, thus opening the possibility to acquire 2D images at a
very high speed. (Image courtesy of K. Goda. After Goda, Solli, & Jalali,2008).
parameter must satisfy the same condition as an RTFT, K82=1. However,
note that the elements are disposed in the reversed order. This configura-
tion was first proposed by Kauffman et al. (1994), where the time lens
was implemented with an EOPM. More recently, using the silicon time
lens, Foster, Salem et al. (2008) have shown an on-chip time-to-frequency
transformer for measuring the intensity of optical pulses with an unprece-
dented 0.2-ps resolution in a time span larger than 100 ps, which can also
operate in a single shot. Figure 9 shows a pulse featuring a short temporal
structure in a large span captured with this device.
The main limitation of this technique is the relatively long refresh time
of the OSA, which prevents its use as a real-time optical oscilloscope for
telecommunication signals typically operating at Gb/s rate. To overcome
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 29 — #29
Coherent Ultra-High-Speed Optical Systems and their Applications 29
Time
GDD circuit
Temporal lens
K
Output ES
Frequency
Input intensity
FIGURE 8 Scheme of a time-to-frequency converter. This device encodes the
temporal intensity distribution of an input light pulse into the optical spectrum at the
output. Therefore, after proper calibration, a measurement in the spectral domain
with a regular OSA reveals the temporal intensity distribution of the input signal.
10 ps
10 ps 10 ps
0
FIGURE 9 Measurement of a complex pulse featuring ultrafast variations on a large
scale. Left part is the pulse measured with the chip-based time-to-frequency converter
and right part is the measurement done with an optical ultrafast cross-correlator.
(Adapted with permission from Macmillan Publishers Ltd. After Foster, Salem et al.,
2008). c
[2008] NPG.
this limitation, several alternative solutions have been reported recently,
enabling not only real-time operation but also amplitude and phase
retrieval information, albeit at a lower time-bandwidth product (Fontaine
et al.,2010).
The time-to-frequency conversion can be understood as the dual equiv-
alent of the previous optical Fourier transformation. In the same way that
an optical Fourier transformer is insensitive to the input spectral phase,
the time-to-frequency conversion performs its operation regardless of the
input temporal phase profile. This ensures that the device measures the
intensity distribution in such a way that the signal’s inherent random phase
fluctuations do not affect the measurement (Fern´
andez-Pousa,2006).
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 30 — #30
30 Space-Time Analogies in Optics
Input spectrum
Output spectrum
Frequency
Frequency
K
Temporal lens
X
X
FIGURE 10 Simplified time-to-frequency converter with a single time lens with high
chirping rate. This device performs the same operation as the device in Figure 9. After
introducing a huge chirp, a temporal lens maps the intensity profile of an input light
pulse into the optical spectrum. Once the spectral far-field condition is achieved, after
introducing higher chirping rates, the output spectrum distribution does not change in
shape, only in scale.
5.3.2. Spectral Far Field
By recalling that a time lens corresponds to the dual device of a GDD
circuit, it is straightforward to note that a similar far-field regime to that
reported in Section 5.2.2 also exists in the spectral domain (Aza˜
na,2003a).
This provides a route to achieve a simplified version of the previous time-
to-frequency converters by removing the input GDD parameter, as shown
in Figure 10, if the input pulse satisfies
K2/(4π ), (29)
where is the spectral bandwidth of the pulse. Analogously to the
temporal far-field phenomenon, once the above limit (also called spectral
Fraunhofer limit) is satisfied, larger chirping amounts do not change the
spectral shape but only its scale.
The first experimental demonstrations of this phenomenon made use
of a pre-distorted pump pulse in order to achieve a high chirping rate.
The idea was to use this time-to-frequency conversion to map the inten-
sity profile from a high-speed electrical signal (even a terahertz beam) into
the optical spectrum of the predistorted pulse (Jiang & Zhang,1998;Cop-
pinger, Bhushan, & Jalali,1999). By measuring the optical spectrum with
a regular OSA, one then gets access to the highly varying temporal infor-
mation of the electrical pulse. The same concept can be used for analog-to-
digital conversion, where the optical spectrum can be demultiplexed and
parallelized with lower speed detectors (Bhushan et al.,1999).
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 31 — #31
Coherent Ultra-High-Speed Optical Systems and their Applications 31
More recently, in a different context, it has been recognized that this
simplified time-to-frequency conversion can be used for generating opti-
cal frequency combs with a spectral profile with enhanced flatness when
the time lens is implemented with an EOPM (Torres-Company, Lancis, &
Andr´
es,2008). Optical frequency combs generated by sinusoidal phase
modulation are widely used for pulse shaping in the field of optical com-
munications (e.g., Miyamoto et al.,2006;Fontaine et al.,2007;Jiang et al.,
2007) because the spacing between the comb lines can be easily controlled
with an external high-speed clock signal (typically few gigahertz), and be
large enough so that the comb spectral lines are resolved and manipu-
lated with state-of-the-art pulse shapers (Jiang et al.,2005). For quite some
time, it became challenging to build an optical frequency comb with truly
flat spectral profile prior to the shaping stage. In recent years, the works
of Fujiwara et al. (2003) and Yamamoto et al. (2007) reported two archi-
tectures where an outstanding flatness was easily achieved. The process
involved the generation of a flat-top pulse train followed by an EOPM.
Figure 11 shows an example of the flat frequency comb achieved by
Yamamoto et al. (2007). The work of Torres-Company, Lancis, and Andr ´
es
(2008) pointed out that the physical phenomenon behind this was the
simplified time-to-frequency conversion, where the flat intensity profile is
mapped into the spectral domain, thanks to the action of the EOPM, which
acted as a time lens working in the spectral Fraunhofer regime. Because
the time lens operated in a pulse-by-pulse basis, the spectral envelope of
the signal (and thus that of the frequency comb) became flat. That work
explicitly pointed out that by correcting the time lens temporal aberrations
in the EOPM a better flatness than reported could be achieved. In addition,
thanks to the spectral Fraunhofer-field properties, it became clear that a
higher chirping leads to broader frequency combs without altering the
desired spectral flatness.
20
20
40
60
80
1.0 0.5
120
140
160
0
0 0.5 1.0
61 carrier light
Experiment Calculation
RIN (dB/Hz)
Relative optical frequency (THz)
Power (dBm)
RIN
FIGURE 11 Ultra-flat frequency comb achieved by time-to-frequency conversion of a
flat-top intensity pulse using an EOPM temporal lens, based on the principles of the
simplified time-to-frequency converter of Figure 10. (After Yamamoto et al., 2007).
c
[2007] IEEE.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 32 — #32
32 Space-Time Analogies in Optics
5.4. Temporal Imaging Systems
These devices provide a scaled replica in intensity of an input short light
pulse. Three different schemes are considered. When designed for mag-
nifying, the output pulse duration becomes large enough to overcome
the limited temporal resolution of current fast photodetectors, acting as
a temporal microscope for ultrashort events (Bennett & Kolner,1999).
Alternatively, TISs can also be applied for pulse compression while main-
taining the initial intensity profile (Bennett & Kolner,2000). Due to the
possibility of engineering the magnification factor to be negative, TISs
have also been proposed for the creation of anticorrelated two-photon
light in spontaneous parametric down-conversion (Tsang & Psaltis,2006),
for compensating for the TOD effect in WDM systems (Kumar,2007), or
achieving phase conjugation (Kuzucu et al.,2009). Here, three different
schemes are reviewed that can act as a TIS.
5.4.1. Conventional Temporal Imaging
The basic TIS configuration is sketched in Figure 12, where two first-order
dispersive elements are used (Kolner & Nazarathy,1989). The spatial
counterpart is also illustrated for comparison. The temporal imaging con-
dition links the chirp of the temporal lens and the GDD parameters. By
calculating the ABCD matrix corresponding to the Gaussian system of
Figure 12 and imposing B=0, we obtain
1
821 +1
822 =K. (30)
GDD circuit
Output intensity
Time
Time
K
GDD circuit
Input intensity
Temporal lens
FIGURE 12 Schematic representation of a conventional temporal imaging system.
This system scales the temporal intensity distribution of an input pulse, so that it can
be larger (like a temporal microscope), shorter (pulse compressor), or even reversed.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 33 — #33
Coherent Ultra-High-Speed Optical Systems and their Applications 33
The magnifying factor, m, is given by the matrix coefficient A, which
under the above constraint results in m=822/821. Thus, with this scheme
and selecting properly the GDD parameters, it is possible to engineer the
scaling factor of the TIS.
In the earlier demonstrations of TISs, magnification factors of 12×
(Bennett, Scott, & Kolner,1994) and 24×compression (Godil, Auld, &
Bloom,1994) were achieved using SFG and EOPM as time lenses, respec-
tively. With today’s time lenses implemented using FWM in silicon chips,
a temporal magnification factor >500 has been recently reported by Salem
et al. (2009), enabling high-speed optical sampling using a lower rate
digital sampling oscilloscope.
5.4.2. Time-Stretching Technique
As a remark, the scheme displayed in Figure 12 can be understood as the
concatenation of two different subsystems, a time-to-frequency converter
(first GDD medium and part of the time lens) and a frequency-to-time
converter (consisting of the remaining part of the time lens and the sec-
ond GDD medium). This two-fold concatenation is property inherent to
all TISs.
On the other hand, from Sections 5.2 and 5.3, we have learned that
simplified configurations for time-to-frequency and frequency-to-time
mapping exist. Then, as firstly reported by Caputti in the context of elec-
tronic systems (Caputi,1971), a simplified TIS for time magnification can
also be built. This is known in the literature as the time-stretching tech-
nique and the first implementation in the optical domain was reported by
Bhushan, Coppinger, and Jalali (1998). The operational principle is illus-
trated in Figure 13. Here, one makes use of the chirp of a predistorted
Gaussian pulse in a highly dispersive GDD medium as time lens in order
to induce the spectral Fraunhofer regime on an electrical signal (Cop-
pinger, Bhushan, & Jalali,1999) whose temporal image is to be acquired.
After this time-to-frequency conversion, the optical spectrum is mapped
into the temporal domain by a subsequent frequency-to-time operation in
a highly dispersive material (Tong, Chan, & Tsang,1997). Then, the scaling
factor for the magnified electrical signal becomes m=822/821 (Bhushan,
Coppinger, & Jalali,1998). Because the second dispersive medium must
be highly dispersive to achieve the frequency-to-time mapping opera-
tion, the typical magnification factors achievable with this technique are
very high. It is also interesting to note that a similar time-stretching tech-
nique has been recently reported. The main difference is the substitution
of the first time-to-frequency converter introduced by the dispersed pulse
with a deeply phase-modulated EOPM (Aza˜
na, Berger et al.,2005b). Some
authors have pointed out an analogy with the spatial phenomenon of
shadow casting (Aza˜
na,2003b;Aza˜
na, Lugo et al.,2005).
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 34 — #34
34 Space-Time Analogies in Optics
GDD circuit Φ21
GDD circuit Φ22
E/O device
Input
pulse
Chirped pulse
Electrical pulse to image
Stretched output pulse (temporal image)
1. Time-to-frequency conversion
2. Frequency-to-time conversion
FIGURE 13 Scheme for temporal imaging based on the time-stretching technique.
Here the signal to be imaged is the electrical impulse. The operation involves a
two-step process. First the electrical signal is mapped into the spectral domain of an
optical pulse, thanks to the chirping it acquires through propagation in a GDD circuit.
Later, the spectrum of the modified pulse is mapped to the temporal domain, thus
achieving a large magnification factor.
This time-stretching technique has become particularly useful for mag-
nification of high-speed electrical signals. Combined with high-speed
electrical analog-to-digital conversion, this method upgrades the effec-
tive sampling rate by mtimes. Because the magnification factor increases
proportionally with the GDD parameter, it is interesting to get the high-
est possible value for 822. Making use of Raman amplification in a GDD
circuit based on a DCF, a magnification factor of 250 has been recently
reported, leading to an effective sampling rate of 10 Tsample/s using a
state-of-the-art electronic digitizer (Chou et al.,2007).
5.4.3. General Temporal Imaging Phenomenon
Apart from the two previously mentioned TISs, there is another, less
known, TIS that is capable of performing an image of the temporal inten-
sity distribution of a virtual Fresnel plane (Aza˜
na & Chen,2003;Lancis,
Caraquitena et al.,2005). The idea can be understood with the help of
Figure 14. The distortion of a short light pulse with envelope ψin(t)
propagating in a GDD circuit is given by the Fresnel integral
ψun(t)=exp it2
282!Zψin(t0)exp it02
282!exp itt0
82dt0. (31)
If we consider the distortion of a chirped light pulse with envelope
exp(iKt2/2in(t)through a different GDD circuit with coefficient 82,eq,
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 35 — #35
Coherent Ultra-High-Speed Optical Systems and their Applications 35
Input pulse
Time lens
Output pulse
“virtual temporal image”
Time
Linearly distorted
pulse
Time
Time
K
GDD circuit Φ2
GDD circuit Φ2,eq
FIGURE 14 Scheme to understand the general temporal imaging phenomenon. The
distortion of a chirped light pulse can be always explained as a scaled replica of the
unchirped version.
the corresponding Fresnel integral would be
ψchirped(t)=exp it2
282,eq !Zψin(t0)exp iKt02
2!
×exp it02
282,eq !exp itt0
82,eq dt0. (32)
We note that for 82,eq =82/(1+K82), the intensity profile becomes a
scaled replica, with scaling factor 1 +K82, that would be achieved at 82
for an unchirped pulse. In other words, the distortion of a chirped light
pulse through a GDD circuit with GDD coefficient 82,eq can always be
interpreted as the temporal image of the unchirped version propagat-
ing through a GDD medium satisfying the imaging condition 82,eq =
82/(1+K82). This generalized temporal imaging law has been used to
compress electrical signals, thus increasing their electrical bandwidth a
factor given by 1 +K82, which is easily controlled by the chirp and GDD
coefficient (Aza˜
na et al.,2004;Torres-Company et al.,2006).
5.5. Spectral Imaging System
In the previous section, we have established the rules to achieve temporal
imaging, that is, a system that provides a scaled replica in intensity of an
input light pulse. However, because the scaling is done for the intensity, a
TIS does not necessarily scale the optical spectrum of the input signal. If
one desires to scale the optical spectrum of an input light pulse, regardless
of the intensity profile, an alternative procedure would be to use a spectral
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 36 — #36
36 Space-Time Analogies in Optics
GDD circuit
Temporal lens
Output ES
Frequency
Frequency
K1
K2
Input ES
Temporal lens
FIGURE 15 Scheme of a spectral imaging system. This device scales the optical
spectrum of an input light pulse, so that it can be larger (spectral microscope), shorter
(spectral compressor), or even reversed in sign.
imaging system (SIS), which is the dual of a TIS (Torres-Company, Lan-
cis, & Andr´
es,2007a). Following the ABCD matrix analysis exposed in
Section 4.4, an SIS is a Gaussian system with a matrix coefficient C=0.
The most basic implementation is illustrated in Figure 15, which consists
of two temporal lenses sandwiching a GDD circuit. If we invoke the SIS
condition, the chirp rates and GDD parameter must satisfy the relation
(Torres-Company, Lancis, & Andr ´
es,2007a)
1
K1+1
K2=82, (33)
This can be easily understood if we remind that a time lens constitutes the
dual element of a GDD circuit. As a consequence, an SIS can be consid-
ered as the concatenation of a frequency-to-time mapping (i.e., the optical
Fourier transformer composed by the first time lens and part of the GDD
circuit) followed by a time-to-frequency converter (composed by part of
the GDD circuit and the second time lens). Consequently, the spectral
scaling factor would be provided by 1 K282.
Combined with a high-resolution OSA, this processing system can
be used as a spectral telescope for spectroscopy applications. The first
implementation of this original proposal involved the use of two sili-
con time lenses (Okawachi et al.,2009), achieving a magnification factor
of 105×, leading to an effective 1-GHz optical resolution. These results
appear in Figure 16. We must note however that in this SIS, the time
aperture of the first time lens will limit the shortest discernible spec-
tral feature size, regardless of the resolution of the OSA employed. This
means that by using an OSA with better resolution it might not be possi-
ble to observe low-frequency spectral features if these do not fall into the
time window imposed by the time lenses of the system. This constitutes
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 37 — #37
Coherent Ultra-High-Speed Optical Systems and their Applications 37
1.0
0.8
0.6
0.4
0.2
0
1536.1 1536.14 1536.18 1536.1 1536.14 1536.18
Transmission
1.0
0.8
0.6
0.4
0.2
0
1540 1544
Wavelength (nm)
Transmission
Wavelength (nm)
1548 1540 1544 1548
FIGURE 16 Experimental results of the measurement of the absorption spectra of
C2H2using a conventional OSA (top row) and a spectral imaging system (bottom row)
implemented using FWM time lenses. A 105×spectral magnification is observed.
From left to right there is a change in pressure of the gas under study. (Adapted from
permission of OSA. After Okawachi et al.,2009). c
[2009] OSA.
another example of how important is to minimize the temporal aberration
effects in temporal optics applications.
5.6. Ultrafast Fourier Processing Systems
In optical information processing, a Fourier filtering system is a basic com-
ponent for manipulating the spectrum of an optical signal. For ultrafast
optical pulses, the most widespread solution to do so in a reconfigurable
manner is probably the Fourier transform pulse shaper (Weiner,2000).
Briefly, the frequency components of a short light pulse are angularly dis-
persed using, typically, a grating placed at the back focal plane of a spatial
lens. At the focus of the lens, a spatial light modulator (SLM) manipulates
the amplitude and/or phase of the complex spectrum of the signal. Later,
the filtered spectrum is recombined through a reversed operation either in
a reflective configuration or by traveling through identical lens and grat-
ing. The ultimate goal of this device is the possibility for reconfiguring
the synthesized waveform, due to capability to reprogram the mask at the
SLM. With some modifications, this device can manipulate other degrees
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 38 — #38
38 Space-Time Analogies in Optics
of freedom, such as the beam profile or polarization (Feurer, Vaughan,
& Nelson,2003;Ninck et al.,2007;Esumi, Kabir, & Kannari,2009). The
simultaneous high bandwidth and high resolution obtained with these
devices have allowed for the highest complexity in pulse shaping today
(Jiang et al.,2007;Supradeepa et al.,2008). Some applications, however,
require the use of more compact and rapidly reconfigurable pulse shap-
ing processors (Geisler et al.,2009). For these applications, the space-time
analogy offers a wide range of possibilities by direct inspection into the
counterpart spatial Fourier processors.
In general terms, the Fourier processors inspired by the space-time
analogy concepts involve a three-step process, which in terms of the
temporal optics could be explained as follows. First, a frequency-to-time
converter performs physically the Fourier transform of the signal which
we want to filter. Second, the optical spectrum of the signal is filtered in
real time with an optical modulator. The point is that this active device
manipulates (in amplitude and/or phase) the spectral components of the
signal that are displayed in time. Finally a time-to-frequency conversion
reverses the operation of the first step and brings back the filtered optical
spectrum of the input signal to the temporal domain. We can distinguish
between two different systems capable of implementing this basic Fourier
processing operation.
5.6.1. 4-f Temporal Processing Systems
The scheme of this architecture is illustrated in Figure 17. An optical
Fourier transformer performs the required frequency-to-time mapping
K1K2
Φ21 Φ21
Φ22 Φ22
Input pulse
Time lens
Output pulse
GDD
circuit
GDD
circuit
Time lens
GDD
circuit
GDD
circuit
Chirp free
optical Fourier transformer 1
Chirp free
optical Fourier transformer 2
E/O
device
Electrical pulse
(filter mask)
Time Time
FIGURE 17 Scheme of a temporal 4f Fourier processing system. This device filters an
input short light pulse. The idea involves a two-step process. First, an optical Fourier
transformer performs the Fourier transform of the input signal. The spectral
components of the pulse are manipulated in real time with an electro-optical device
driven by an arbitrary electrical signal. In this way, the spectrum of the pulse can be
manipulated in a user-defined fashion. After that, a second optical Fourier transformer
undoes the operation of the first system, so that the output pulse is a filtered scaled
version of the input, as predicted by Equation (34).
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 39 — #39
Coherent Ultra-High-Speed Optical Systems and their Applications 39
operation. We note that the introduction of the first GDD circuit to achieve
optical Fourier transformation without the inherent temporal quadratic
phase factor, that is, after frequency-to-time mapping, the pulse is trans-
form limited. The optical Fourier transformation rule dictates 821K1=1.
Later, the spectrum of the pulse is manipulated directly in the time domain
using an active modulator with complex transfer function m(t). And
finally, a time-to-frequency converter undoes the operation implemented
in the first step, for which we must require 822K2=1. Consequently, the
filtered spectral complex envelope at the output of the system, ˜
ψout(ω), is
given by
˜
ψout(ω) =˜
ψin(mωω)m(822 ω). (34)
Here, ˜
ψin(ω) is the input complex spectral envelope and the spectral scal-
ing factor mω= −K1/K2. Therefore, one can scale and filter independently
the complex spectrum of the signal with the proper selection of the sys-
tem parameters. We note that in the particular case in which mω=1, the
system behaves mathematically as a pulse shaper device, in which the
complex transfer function of the modulator plays the role of the system’s
transfer function. In this configuration, the spectral resolution of the pulse
shaping system is given by δω =δt/822, where δtis the temporal resolu-
tion of the active modulator. Probably due to the complexity of the system,
no demonstration of temporal filtering with active modulation has been
reported so far.
An interesting modification of the previous device appears when there
is no active modulation, that is, m(t)=1. In this case, the system behaves
as a temporal telescope and ˜
ψout(ω) =˜
ψin(mωω), indicating that scaling
in amplitude and phase becomes possible. Although the theoretical pro-
posal of temporal telescopes was reported in the early years of temporal
imaging, it was not until very recently that the first experimental imple-
mentation has been reported, making use of two time lenses based on
FWM in silicon nano waveguides (Foster et al.,2009).
5.6.2. Simplified Configuration
If we recall that the frequency-to-time mapping operation can be achieved
using a highly dispersive GDD circuit (see Section 5.2.2), then an ultra-
fast optical filter in the time domain can be implemented following a
configuration like the one presented in Figure 18 (see Haner & Warren,
1989; Weiner & Heritage, 1990; Saperstein & Fainman,2008). The short
light pulse to be filtered is stretched in time through propagation in the
GDD circuit, so that the spectral components are decomposed in the
time domain. Then, a temporal modulator modifies the amplitude and or
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 40 — #40
40 Space-Time Analogies in Optics
Input pulse Output pulse
Electrical pulse
(filter mask)
Time Time
E/O device
GDD
circuit
Φ21
GDD
circuit
Φ22 = − Φ21
FIGURE 18 Scheme of a simplified temporal processing system. The idea is very
similar to that depicted in Figure 17. An optical pulse is stretched in time so that the
optical frequencies are displayed in the temporal domain. These optical frequencies
are manipulated in real-time by an active modulator. Then a second GDD circuit
has a reversed dispersion to undo the operation of the first stretcher element.
phase of these physical spectral components. Later, a second GDD circuit
with dispersion parameter designed to compensate for the first one (and
possibly the GDD introduced by the modulator) is inserted, with the aim
to bring back the synthesized spectral profile to the temporal domain. The
main goal of this configuration is that the temporal resolution of the whole
shaping device is determined by the duration of the pulse to be shaped
and not by the temporal response of the modulator.
In mathematical terms, the output complex envelope becomes
ψout(t)=exp it2
2822 !ZZ ψin(t00)m(t0)exp it02
21
822 +1
821 !
×exp it002
2821 !exp it0t00
821 +t
822 dt0dt00 (35)
By assuming that the temporal width of the input pulse is much shorter
than the temporal width of the modulator, invoking the far-field condi-
tion, and demanding the GDD parameters to compensate for each other,
we can then approximate the previous equation to
ψout(t)exp it2
2822 !Zψin(t0)˜
mtt0
822 dt0, (36)
where ˜
m(ω) is the Fourier transform of m(t).
The previous equation indicates that the modulation response of the
modulator manipulates the optical spectrum of the input signal. The spec-
tral resolution of the system is provided by δω =δt/822, where again δt
is the finest temporal resolution of the active modulator. By introducing
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 41 — #41
Coherent Ultra-High-Speed Optical Systems and their Applications 41
larger dispersion amounts, the system provides better resolution. It is
instructive to determine the minimum spectral resolution this system is
capable to achieve, given an input pulse train operating at a certain rep-
etition rate f. The minimum spectral resolution is given by the maximum
GDD parameter, 822,max =σ0/(4f). This maximum dispersion amount is
fixed by the condition that two neighbor pulses must not overlap after
frequency-to-time mapping conversion, that is, the system should operate
in a pulse-by-pulse basis. Thus, δωmin =4fδt0, indicating that there is
a tradeoff between the complexity of the filter and the maximum speed.
In particular, in order to shape pulses in a line-by-line basis, a tempo-
ral modulator as fast as the initial pulse duration should be used. Other
details including the deviation from ideal quadratic phase modulation in
the GDD circuits have been studied by Chi and Yao (2007).
Despite its limitations, this basic Fourier processor has been used
for several applications, where compactness at relatively high speed
rates is required. The first implementation involved the use of a hybrid
fiber/grating compressor as GDD circuits and an electro-optic modu-
lator as the active filtering device (Haner & Warren, 1989). Using an
EOPM driven by a sinusoidal signal as filter, this system can be used to
compensate for the TOD effects of short light pulses in fiber transmission
links (Pelusi, Matsui, & Suzuki,1999). The use of high-speed electro-optic
amplitude and phase modulators has triggered applications in the genera-
tion of high-speed electrical signals (Saperstein et al.,2005;Aza˜
na, Berger
et al.,2005a), generation of single-sideband modulation (Chou et al.,2009),
controllable time delays (Xin et al.,2010), and optical pulse shaping based
on binary phase coding (Thomas et al.,2009,2010). Some experimental
examples obtained with this last configuration are shown in Figure 19.
5.7. Joint Transform Correlator
A widely extended system for processing optical information in the spa-
tial domain is the joint transform correlator (JTC), firstly reported by
Goodman (1996). As sketched in Figure 20(a), this system enables the cor-
relation of two arbitrary objects, a fundamental operation for matched
filtering, pattern recognition, and encryption. The implementation of the
JTC involves a two-step process. First, we require performing the optical
Fourier transformation of the objects to correlate, which are spatially sep-
arated at the input plane. The interference pattern (or interferogram) is
recorded in intensity. The next step is to calculate the Fourier transform
of the interference pattern which contains a term proportional to the cor-
relation of the objects. We remark that in the first step the interferogram
is actually recorded in intensity prior to calculate its Fourier transform
in the second step. Thus, if one wishes to implement an in-line JTC, an
intensity-to-field conversion element is required.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 42 — #42
42 Space-Time Analogies in Optics
Normalized intensity
1.0
0.8
0.6
0.4
0.2
0.0
–20 –15 –10 –5 0
Time (ps)
51015
Normalized intensity
1.0
0.8
0.6
0.4
0.2
0.0
–15 –10 –5 0
Time (ps)
5101520
20 –20
1.6 ns
FIGURE 19 Examples of synthesized pulses using the simplified configuration
of Figure 28, with an electro-optic phase modulator as the active modulator.
The electrical signal consists of binary square pulses with amplitude leading to π
phase changes (shown in the inset of the figure). The key is that different codes can
be programmed for neighbour pulses therefore allowing for an ultrafast transition
from one synthesized pulse to the next one. (Image courtesy of A. Malacarne and
J. Aza ˜
na. After Thomas et al.,2010). c
[2010] IEEE.
In the temporal domain, there are few theoretical proposals to achieve
the correlation of two arbitrary signals (Lohman & Mendlovic,1992;Shab-
tay, Mendlovic, & Zalevsky,2000;Geraghty et al.,2009), but the temporal
equivalent of the JTC sketched in Figure 20(a) has been only recently
reported (Torres-Company & Chen,2009). The configuration is shown in
Figure 20(b). Here, two pulses with envelopes ψa(t)and ψb(t)are tem-
porally separated by a quantity T, playing the role of the spatial objects
to be correlated. The idea is to perform the frequency-to-time mapping
operation of the sequence of two pulses using a highly dispersive GDD
medium. We note that to mimic strictly the scheme in Figure 20(a), the
frequency-to-time mapping operation should be achieved by the ultra-
fast optical Fourier transformer of Section 5.2.1. However, for the pur-
poses of a JTC, the far-field phenomenon of Section 5.2.2 can be used
instead of frequency-to-time mapping. For achieving the correlation of
the pulses, one needs to perform the Fourier transformation of the inten-
sity interferogram, that is, measuring the radio-frequency spectrum of the
interferogram (Torres-Company & Chen,2009).
In mathematical terms, after dispersion in the GDD circuit with disper-
sion parameter 82, the intensity profile becomes a scaled replica of the
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 43 — #43
Coherent Ultra-High-Speed Optical Systems and their Applications 43
input spectrum,
Iout(t)= | ˜
ψa(ω)|2+ | ˜
ψb(ω)|2+˜
ψ
a(ω) ˜
ψb(ω) exp(iωT)+c.c.|ω=t/82, (37)
where ˜
ψa(ω) and ˜
ψb(ω) are the Fourier transform of ψa(t)and ψb(t),
respectively. To achieve a temporal JTC, the key is to calculate the Fourier
transform of the temporal signal given by Equation (37),
˜
Iout(f)=dc(f)+X(f+1f)+X(f1f). (38)
Δx
Lens
Lens
Computer
(a)
(b)
Correlation
peak
*
Interferogram
as object
CCD
camera
Fourier transformation Fourier transformation
Object
f
λ1
Pump
Dispersive
medium
CW laser
HNLF
OSA
TimeTime
Frequency
*
Frequency-to-time mapping
(optical) RF analyzer
T
f
FIGURE 20 Scheme of a joint transform correlator in: (a) spatial domain. An optical
Fourier transformer followed by a camera records the interference pattern between
the objects. Then, by digital or optical means, we perform the Fourier transform of
the interference, consisting of three terms, one of them containing the information
of the cross-correlation between the objects. In (b), we have the same idea presented
in the temporal domain. Two pulses for which we desire to measure their
cross-correlation are stretched in a GDD medium. The intensity profile is a scaled
replica of their spectral interferogram. Performing the Fourier transform of the
interferogram is identical to calculate the RF spectrum of this intensity profile, which
is done by a nonlinear optics method in the figure.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 44 — #44
44 Space-Time Analogies in Optics
The above equation implies that the spectrum is constituted by three
terms: dc(f)R[|˜
ψa(ω)|2+ | ˜
ψb(ω)|2] exp(i2πfω82)dω, centered exactly
at the zero-frequency component and the cross-correlation terms X(f)
Rψ
a(tb(t2πf82)dt, centered at
1f=T
2π82
. (39)
In principle, if there is no spectral overlapping between the dc and cross-
correlation terms (for which a proper choice of delay Tand dispersion
parameter 82are required), the X(f)can be isolated.
1
0
0
20
40 20 20 400
Frequency offset (GHz)
(b) (d)
40
60
0 500
Matched case
2.9 dB
Time (ps)
(a) (c)
Time (ps)
1000
Input sequence
Power JTC (dBm)
1
0
0 500
Unmatched case
1000
Input sequence
0
20
40 20 20 400
Frequency offset (GHz)
40
60
7.2 dB
Power JTC (dBm)
FIGURE 21 Experimental results for a temporal joint-transform correlator obtained
using the setup of Figure 20(b). This system can be used to identify whether two
pulses in a sequence are identical or not. Two different cases are considered and
plotted numerically, matched (left column), when the signals are identical, and
unmatched (right column), when the signals to correlate are different. Figures (b) and
(d) show the corresponding RF spectra of the Fourier transform of the sequence,
measured with the all-optical technique in Figure 20(b). We observe that the
unmatched case leads to a decreasing in power with respect to the zero-frequency
component.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 45 — #45
Coherent Ultra-High-Speed Optical Systems and their Applications 45
The recognition of ultrafast sequences is a topic of increasing rele-
vance in optical communications. A possible way to achieve this in an
all-optical and in-line manner is by using an all-optical radio-frequency
spectrum analyzer based on XPM in a highly nonlinear medium (Dorrer
& Maywar,2004;Pelusi et al.,2009). In Figure 21, we show experimental
measurements of the RF spectrum of two signals multiplexed in time
when they are the same (matched case) and when they are not (unmatched
case). By measuring the peak height relative to the zero-frequency compo-
nent in the RF spectrum of the interferogram, one can assess the similarity
of the two signals, thus opening the possibility to perform optical recog-
nition of ultrafast pulses. The advantage of using this system is that with
a suitable choice of the GDD parameter, the pulses to be correlated may
come from a sequence going at an ultra-high-speed rate (i.e., Tcould be in
the subnanosecond range).
5.8. Temporal Talbot Effect
Throughout the previous sections, we have studied the distortion of a
pulse envelope in a GDD circuit without specifying its shape. In the par-
ticular case in which the signal is a train of ultrashort light pulses, the
temporal spreading and interference between adjacent pulses may lead to
unexpected results. In this section, we focus our attention in the distortion
of periodic signals in GDD media and review an interesting phenomenon
which bares an analogy with a well-known physical effect in Fourier
optics: the Talbot effect (for a review see Patorski,1989).
Let us consider an input pulse train at repetition rate f, with periodic
envelope
ψin(t)=
X
n=−∞
cnexp(i2πnft), (40)
where, of course,
cn=f
1
2f
Z
1
2f
ψin(t)exp(i2πnft)dt. (41)
We now propagate the above signal into a GDD circuit. By substituting
the matrix elements from Equation (15) to Equations (10) and (9) and
considering Equation (40), we get
ψout(t)=
X
n=−∞
cnexp(i2π282n2f2)exp(i2πnft). (42)
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 46 — #46
46 Space-Time Analogies in Optics
This equation indicates that the output complex envelope is always a
periodic signal. And the corresponding light intensity profile
Iout(t)=
X
m=−∞
Cm(82)exp(i2πmft), (43)
Is, of course, periodic too. With
Cm(82)=exp i2πm282
82T
X
n=−∞
c
ncn+mexp i4πmn 82
82T, (44)
where we are defining the Talbot GDD parameter as
82T=1
πf2. (45)
If we now restrict our study to the specific GDD parameters that satisfy
82=P
Q82T, (46)
where Pand Qare coprime integer numbers, then, the output complex
field envelope can be written as (Berger et al.,2004)
ψout(t)=
Q1
X
L=0
G(L,Q,Pin tL
fQ. (47)
with the complex coefficients given by the summation
G(L,Q,P)=1
Q
Q1
X
q=0
exp i2πq
Q(LqP). (48)
These formulae are the basic results that allow us to study the temporal
Talbot effect or self-imaging phenomenon. This effect appears whenever
the input light sequence is recovered after propagation in a GDD circuit.
If the repetition rate is maintained, we discuss about integer Talbot effect,
whereas if the repetition rate is increased (while preserving the individual
pulse shapes), we discuss about fractional Talbot effect.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 47 — #47
Coherent Ultra-High-Speed Optical Systems and their Applications 47
5.8.1. Integer Temporal Self-Imaging
As sketched in Figure 22, the integer temporal Talbot effect or temporal
self-imaging phenomenon leads to the regeneration of an input periodic
pulse sequence propagating through a GDD circuit. The output pulse train
is an exact replica in field of the input one at those GDD amounts for which
Q=1 or in intensity for Q=2. This corresponds to the temporal counter-
part of the well-known phenomenon in the spatial domain. This effect was
discovered by Talbot in 1836, when he realized that the diffracted field of
a grating replicated itself after certain propagation distances. The tempo-
ral version of this phenomenon was proposed by Jannson and Jannson
(1981) and verified a decade later by Andrekson (1993) although with-
out pointing out the connection with the Talbot effect. Figure 23 shows an
Input train
GDD circuit
Output train
Time
Time
FIGURE 22 Schematic representation of the integer temporal Talbot effect. An input
periodic pulse train is launched into a GDD medium that satisfies the Talbot condition.
The output is an identical replica in intensity of the input, with the same period.
0
Intensity (a.u.)
Input sequence Temporal self-image
Time (ps)
(a) (b)
100 200 0
Time (ps)
100 200
FIGURE 23 Example of the temporal self-imaging or integer Talbot phenomenon.
(a) A periodic input sequence is launched into a GDD medium that satisfies the
integer Talbot condition. (b) The output sequence is a periodic sequence identical to
the input.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 48 — #48
48 Space-Time Analogies in Optics
example of a 12.5-GHz repetition rate pulse train at the input and prop-
agated through a GDD circuit satisfying the integer Talbot condition. We
observe how the pulse train is completely recovered at the output.
Giving a further step, the temporal Talbot effect has been proposed in
telecommunication systems as a simple mechanism to reduce the timing
jitter of periodic pulse trains (Fern´
andez-Pousa et al.,2004), as well as a
clock recovery device (Pudo, Depa, & Chen,2007). These configurations
have also been adapted to the temporal domain from their correspond-
ing spatial counterpart phenomena (Maheswari, Takai, Asakura,1991;
Kalestynski & Smolinska,1978). More recently, the temporal Talbot effect
has been used to compute factorial number (Bigourd et al.,2008). As an
example of the capabilities of the temporal Talbot effect for self-restoring
an imperfect sequence, with potential applications for clock recovery,
Figure 24 shows how an input sequence consisting of ones and zeros
(observed by an eye diagram with a zero base line) is transformed after
propagation into a sequence of only ones (an eye diagram without zero
base line).
5.8.2. Fractional Temporal Self-Imaging
This is the case when Qand Pare coprime integer numbers. Here, the
output intensity is constituted by pulses with identical intensity shape
but the repetition rate has increased by a factor of Q/2, as can be derived
from Equation (47). In the spatial domain, replicas in intensity of the input
grating spatial distribution can be achieved at fractional distances of the
fundamental one with a reduced period. In the temporal domain, this is
schematically represented in Figure 25. This phenomenon allows us to
achieve multiplication of the repetition rate of an input optical pulse train
just by dispersing it into a GDD circuit. This effect was firstly observed by
Shake et al. (1998) and Arahira et al. (1998), where an SMF was used as
GDD circuit. However, the interpretation with the fractional Talbot effect
was provided later (Aza˜
na et al.,1999), where the use of an LCFG as
GDD circuit was suggested. Such an experiment was later performed by
Longhi et al. (2000). It is worth mentioning that the Talbot effect can also
take place in higher order dispersion elements (Duchesne, Morandotti, &
Aza˜
na,2007).
The achievement of ultra-high repetition rate pulse sequences is an
important requisite in optical time-division multiplexing (OTDM) sys-
tems. What makes the fractional Talbot effect so interesting is its simplicity.
However, it is important to note two main limitations which make it diffi-
cult to use as an effective mechanism for OTDM applications: the high
sensitivity to slight mismatches of the fractional Talbot condition and
the fact that after multiplication, not every pulse has the same phase.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 49 — #49
Coherent Ultra-High-Speed Optical Systems and their Applications 49
(a)
(b)
100 ps
100 ps
2.5 mW
2.5 mW
FIGURE 24 Experimental demonstration of the temporal Talbot effect for restoring an
imperfect sequence. (a) Eye diagram of the input signal. The zero base line indicates
that the sequence is constituted by 1s and 0s, that is, it is not a perfect periodic
signal. (b) Output after travelling a GDD medium satisfying the integer Talbot
condition. The zero base line has disappeared, indicating that the output signal is
conformed by 1s only, that is, it is a periodic sequence. (After Pudo, Depa, & Chen,
2007). c
[2007] IEEE.
To overcome this last problem, Atkins and Fischer (2003) demonstrated a
method that involves the cross-gain modulation effect in a semiconductor
optical amplifier to induce intensity-to-field conversion.
5.9. Temporal Array Illuminators
The electro-optic pulse generation method is a technique that produces
high-repetition-rate pulse trains while preserving throughput (Kobayashi
et al.,1988;Murata et al.,2000;Komukai, Yamamoto, & Kawanishi,2005).
As sketched in Figure 26, the basic system consists of a continuous-wave
narrowband laser modulated in phase by a periodic microwave signal.
Pulsed light is achieved by temporal spreading in a GDD circuit, which
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 50 — #50
50 Space-Time Analogies in Optics
Time
Time
Input train at f
Output train at 2
f
GDD circuit
Φ2
2T/4
FIGURE 25 Schematic representation of the fractional temporal Talbot effect. An
input periodic sequence can have its repetition rate multiplied by Q/2 times after
launching the train onto a GDD circuit that satisfies the fractional condition given by
Equation (46).
Time
Time
CW laser
EOPM
GDD circuit
Output intensity
RF signal
FIGURE 26 Schematic representation of the electro-optic pulse generation method.
A continuous wave signal is modulated periodically in phase with a sinusoidal signal.
The modulated signal is dispersed in a GDD circuit to induce phase-to-intensity
conversion and thus achieving a high-repetition-rate pulse train.
converts the input phase modulation into intensity modulation. Electro-
optic phase modulation is commonly related to sinusoidal profiles only.
This is because commercially available single-tone RF modulators can
produce signal bandwidths >80 GHz, whereas arbitrary RF-waveform
generators have a limited maximum bandwidth 10 GHz. Pulse gener-
ation by sinusoidal phase modulation is usually understood by means of
the so-called bunching parameter (Kobayashi et al.,1988). This approach
establishes the required GDD amount to focus the parabolic region of
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 51 — #51
Coherent Ultra-High-Speed Optical Systems and their Applications 51
the sinusoidal profile. Then, it is usually accepted that the pulsed light
is optimally compressed if 82K=1, where Kis the chirping rate of the
approximation to a parabolic profile. However, this compression rule can-
not explain the rich panoply of achievable pulse profiles at different GDD
parameters. On the other hand, the space–time analogy provides some
tools to analyze the problem of ultrashort light pulse generation based
on propagation of periodically phase-modulated light (Torres-Company,
Lancis, & Andr´
es,2006a). The key relies on noticing the equivalence with
the spatial diffraction of phase structures. Here, the so-called Fresnel array
illuminators use pure phase diffraction gratings for producing a two-
dimensional (2D) array of bright spots with equal amounts of light by
free-space propagation at a finite distance. In this direction, an important
amount of research work has been devoted to understand the properties
of light diffracted by periodic phase structures and, above all, to opti-
mize their design in order to achieve a set of high-contrast two-level light
intensity dots by diffraction (see, for example, Winthrop & Worthington,
1965;Guigay,1971;Patorski & Parfjanowicz,1981;Leger & Swanson,1990;
Arriz´
on & Ojeda-Casta˜
neda,1992).
When translated into the time domain, we note that the analysis car-
ried out in the previous section is still valid even when the input sequence
is phase-only modulated (as long as it is periodic in time). However,
thanks to the phase-only restriction, some analytical formulae previously
reported in the field of spatial Fresnel array illuminators can be derived
for particular GDD amounts (Torres-Company et al.,2006). For example,
for P=1 and Q=4, the summation in Equation (48) can be performed
analytically (Guigay,1971), and the corresponding intensity profile can be
written as
Iout(t)=1sin[V(t1/(2f)) V(t)], (49)
where V(t)is the periodic phase profile. This result has provided the
key to explain the a priori counterintuitive finding reported by Komukai,
Yamamoto, and Kawanishi (2005), who achieved a flat-top-pulse profile
from a sinusoidally phase-modulated CW laser at a GDD amount different
from the bunching condition. With a careful insight in their experimental
settings, we note that they selected a GDD parameter at the one-quarter
Talbot condition. By substituting V(t)=sin(2πft)in Equation (49) we
achieve
Iout(t)=1+sin[2sin(2πft)]. (50)
This output pulse profile is schematically represented in Figure 27 for
=π/4 and f=40 GHz. This waveform fits perfectly to the experimen-
tal results previously reported. This constitutes another example of how
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 52 — #52
52 Space-Time Analogies in Optics
20 40 60 80 1000
Time (ps)
Intensity (a.u.)
FIGURE 27 Example of a flat-top pulse generated using the temporal array
illumination theory. A sinusoidally phase-modulated signal is launched into a GDD
circuit satisfying a fractional temporal Talbot dispersion. The intensity profile
achieved in this case can be analytically expressed using Equations (49) and (50).
the space-time analogy can be useful to interpret ultrafast phenomena
from a completely different perspective.
6. TEMPORAL OPTICS IN THE NONCOHERENT REGIME
The previous section focused on ultrafast signal processing applications
operating on fully coherent signals. However, realistic pulses are subject
to random fluctuations, either of fundamental nature (due to the quan-
tum origin of the radiation) or technical (like the spontaneous emission
radiation in amplifier media or thermal or mechanical variations). Opti-
cal coherence theory takes into account the random variations of the
electric field based on the statistical theory of stochastic signals, through
the so-called correlation functions. The modern formulation of optical
coherence theory was developed in the 1950s (Mandel & Wolf,1995).
Since then, the research has been mainly focused on the case of sta-
tistically stationary light fields, for which the averaged light intensity
remains constant in time. This model is suitable to describe the coher-
ence properties of thermal-like light, such as that emitted by diode (LED)
or amplified spontaneous emission (ASE) sources. However, it fails to
deal with the coherence properties of pulsed radiation, where the average
intensity does not remain constant in time. For this aim, optical coher-
ence theory has been recently extended to the nonstationary case (see, e.g.,
Paakonen et al.,2002;Lajunen, Vahimaa, & Tervo,2005;Torres-Company,
Lajunen, & Friberg,2007). In this section, we provide a brief summary
of the basics of the scalar optical coherence theory applied to nonstation-
ary radiation and extend the concepts of the space-time analogy beyond
the bounds of coherent processing. This will be done highlighting the
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 53 — #53
Temporal Optics in the Noncoherent Regime 53
stochastic nature of light not only from a pernicious point of view, but
also pointing out some possible technical advantages in ultrafast signal
processing applications.
6.1. Scalar Coherence Theory for Nonstationary Partially
Coherent Wave fields
Let us consider an ensemble of stochastic complex scalar fields, {U(t)},
whose spatial structure is assumed to be a plane-wave. In this case, the
meaningful physical magnitude becomes the mutual coherence function,
0(t1,t2)=U(t1)U(t2), (51)
which contains the first-order statistical information of the complex field
associated with the light source. The angle brackets denote statistical
ensemble average. From the above function, one can recover important
physical quantities like the average intensity
I(t)=0(t,t), (52)
or the complex degree of coherence
γ (t1,t2)=0(t1,t2)
I(t1)I(t2), (53)
which determines the strength of the field correlations at the two par-
ticular instant times t1and t2. An optical pulse is said to be first-order
coherent if |γ (t1,t2)| = 1, t1,t2, and incoherent if |γ (t1,t2)| = 0, t1,t2.
Intermediate cases deal with partially coherent radiation.
A similar analysis can be done in the spectral domain. In this case, we
study the first-order correlations among the frequency components of the
optical field through the cross-spectral density function
W1,ω2)=D˜
U1)˜
U2)E, (54)
where ˜
U(ω) denotes the Fourier transform of U(t). Important physi-
cal magnitudes can also be obtained from this analysis, like the optical
spectrum
S(ω) =W,ω), (55)
or the complex degree of spectral coherence
µ(ω1,ω2)=W1,ω2)
S1)S2). (56)
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 54 — #54
54 Space-Time Analogies in Optics
Time and frequency representations contain the same quantity of infor-
mation and are linked each other through the so-called generalized
Wiener-Khintchine theorem
0(t1,t2)=ZZ W1,ω2)exp [i1t1ω2t2)]dω1dω2, (57)
and, conversely,
W1,ω2)=1
4π2ZZ 0(t1,t2)exp [i1t1ω2t2)]dt1dt2. (58)
In the limit of a stationary light field, we have 0(t1,t2)=0s(t2t1)and,
accordingly,
W1,ω2)=Ss1)δ(ω2ω1), (59)
indicating that a stationary light field has completely uncorrelated fre-
quencies. Here, according to the Wiener-Khintchine theorem, Ss(ω) forms
a Fourier transform pair with 0s(τ ).
6.2. Space-Time Analogy for Partially Coherent Wave fields
In the noncoherent regime, the meaningful physical quantity becomes
the mutual coherence function, rather than the individual complex
field realizations. Our aim in this section is to get the set of equations
describing the distortion of the mutual coherence function in a dispersive
medium with GVD coefficient β2. To extend the problem studied in
Section 2 to the partially coherent regime, we must remind that the
complex field U(z,t)is now a complex random function. We assume the
randomness can be fully accounted for in the complex envelope ψ(z,t).
Then, the corresponding mutual coherence function is 0(z1,z2;t1,t2)=
hU(z1,t1)U(z2,t2)i=hψ(z1,t1 (z2,t2)iexp{−i[ω0τβ0(z2z1)]} =
0e(z1,z2;t1,t2)exp{−i[ω0τβ0(z2z1)]}, where τ=t2t1. Since each of
the individual realizations of the envelope satisfies Equation (7), we easily
find that the mutual coherence function associated with the envelope,
0e(z1,z2;t1,t2), must satisfy the following pair of equations
"2
t2
j2i(1)j1
β2
zj#0e(z1,z2;t1,t2)=0, (60)
with j=(1, 2). The key is to note that these equations are mathemati-
cally identical to those describing the free-space propagation of a spatially
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 55 — #55
Temporal Optics in the Noncoherent Regime 55
partially coherent light beam at frequency ω, under the paraxial approxi-
mation (Lancis, Torres-Company et al.,2005)
"2
x2
j+2ik(1)j
zj#Ws,e (r1,r2;ω)=0, (61)
where r=(x,z),Ws,e(r1,r2;ω) is the spatial envelope part of the cross-
spectral density function defined in the space-frequency domain, and
k=ω/cis the free-space wave number. We note that the transfer rules
highlighted in Table 1 remain valid in the partially coherent regime. The
goal of this formal extension is that it paves the way to design new ultra-
fast signal processors based on the concepts of Fourier optics incoherent
processing.
6.3. Propagation of Nonstationary Partially Coherent Pulses
To study the propagation of the mutual coherence function through tem-
poral Gaussian systems, one can simply take the results from Section 4.1
and evaluate the corresponding propagated mutual coherence function by
calculating the ensemble average on Equation (9) and assuming the enve-
lope is a complex random function. Taking into account the definition of
the input and output mutual coherence functions, it is straightforward
to get
0e,out (t1,t2)=ZZ K(t1,t0
1)K(t2,t0
2)0e,in t0
1,t0
2dt0
1dt0
2. (62)
Alternatively, one can study the evolution of the cross-spectral density
function3
Wb,out (ω1,ω2)=ZZ ˜
K1,ω00
1)˜
K2,ω00
2)Wb,in ω00
1,ω00
2dω00
1dω00
2. (63)
The key of the above analysis relies on noting that the kernel operators are
exactly the same as those used for the fully coherent case.
Although a rich variety of noncoherent processing systems can be par-
ticularized from the previous analysis, we shall concentrate in a couple of
interesting arrangements. We proceed to propagate an arbitrary mutual
coherence function inside a GDD circuit. For this aim, we take the corre-
sponding matrix coefficients from Equation (15), particularize the kernel
3The subindex bindicates that the cross-spectral density function is referred at the baseband, that is,
Wb1,ω2)=W0
1ω0,ω0
2ω0).
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 56 — #56
56 Space-Time Analogies in Optics
of Equation (10), and insert them into Equation (62) to get
0e,out (t1,t2)=I0exp "it2
1t2
2
282#ZZ exp "it02
1t02
2
282#
×exp it0
1t1t0
2t2
820e,in t0
1,t0
2dt0
1dt0
2, (64)
where I0is an irrelevant constant. We provide a detailed analysis for the
case in which the input mutual coherence function corresponds to a source
like the one reported by Lajunen et al. (2003). It is a spectrally incoherent
stationary light source (e.g., ASE radiation) that is externally modulated
by a fast modulator, as sketched in Figure 6.1. The optical spectral band-
width of the source fixes the coherence time of the light (by an inverse
relation) and the temporal duration of the modulator fixes the duration of
the partially coherent signal (Lajunen et al.,2003). By inserting the corre-
sponding expression of the mutual coherence function for this light source
into Equation (64) and doing some straightforward algebra, the averaged
intensity profile at the output of the GDD circuit is given by
Iout(t)=Sb(t/82)Icoh(t), (65)
where irrelevant constant factors have been dropped for clarity. Here,
Sb(ω) is the spectral density of the source centered at the baseband and
Icoh(t)is the intensity that would be achieved if the spectral source was a
perfect Dirac’s delta, that is,
Icoh(t)=Zexp i82
2ω2˜
M(ω) exp(iωt)dω
2, (66)
where ˜
M(ω) is the Fourier transform of the external complex modulation
m(t). This general result is known as the extension of the Collett–Wolf
equivalent theorem to the Fresnel region, and it was reported first in the
spatial domain by Saleh (1979) and formulated recently in the temporal
domain by Chantada, Fern´
andez-Pousa, and G´
omez-Reino (2006). The rel-
evance of the previous result is that by a proper optical spectrum synthesis
of the light source one can either shape an output intensity profile or filter
the signal that would be achieved in the coherent regime.
6.4. Temporal van Cittert-Zernike Theorem
Let us come back to Equation (64). In this section, we aim to focus on the
equivalent far-zone expression for the output mutual coherence function,
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 57 — #57
Temporal Optics in the Noncoherent Regime 57
which is achieved when the quadratic factor inside the integrand can be
removed. Following the analysis performed by Gori (2005) in the spatial
domain and having in mind the space-time analogy transfer rules, we eas-
ily find that the far-zone is reached in partially coherent temporal optics
whenever
|82|σ
4πmin[2σ,tc], (67)
where σis a measure of the temporal intensity width at the input; min[a,b]
means the smaller of the real numbers a,b; and tcis a measure of the coher-
ence time at the input. Once the above inequality is satisfied, the output
mutual coherence function becomes
0e,out (t1,t2)=I0exp "it2
1t2
2
282#Wb,in(t1/82,t2/82), (68)
that is, a scaled replica of the cross-spectral density function at the input,
in close resemblance to the results presented in Section 5.2.2.
Now, we consider again the same optical source used to achieve Equa-
tion (65). In particular, we further assume that the coherence time is much
smaller than the temporal duration of the modulator, so that the mutual
coherence function behaves as a quasi-homogeneous light pulse (Lajunen,
Friberg, & Ostlund,2006),
0e,in (t1,t2)=Iin t1+t2
2γe,in (t2t1), (69)
where Iin(t)and γe,in(t2t1)denote the averaged intensity and the enve-
lope of the complex degree of coherence of the input partially coherent
light pulse, respectively. Substituting these into Equation (68) and after
some straightforward algebra, one can easily find that the output aver-
aged temporal intensity is given by the Fourier transform of the input
complex degree of coherence and vice versa, the output complex degree
of coherence is given by the Fourier transform of the input averaged
intensity (Lajunen, Friberg, & Ostlund,2006). This double Fourier trans-
formation rule is what is known as the generalized temporal van Cittert-
Zernike theorem, which was firstly proposed and experimentally verified
by Dorrer (2004).
6.4.1. Ultrafast Spectrometer
The temporal van Citter-Zernike may have practical applications. From
the previous section, we have learned that if we consider the source from
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 58 — #58
58 Space-Time Analogies in Optics
Spectrally incoherent light
Deterministic
temporal modulator
ASE light
Partiall coherent
nonstationary pulses
FIGURE 28 Practical scheme to generate partially coherent pulses. A broadband
stationary light source is externally modulated by a deterministic modulator. The
modified mutual coherence function corresponds to an ensemble of non-stationary
light pulses. The coherence time of these pulses is fixed by the coherence time of the
stationary light, while the spectral correlations are fixed by the temporal gate of
the deterministic modulator. (After Lajunen et al.,2003).
Figure 28 having a coherence time much narrower than the temporal
duration of the temporal modulator, the complex degree of coherence of
the input light source determines the output intensity profile, leading to
(Dorrer,2004)
Iout(t)Sin (t/82), (70)
which can be alternatively interpreted as an approximation of Equa-
tion (65). The important point here is that the averaged output intensity is
a scaled replica of the input’s spectral density. This can be considered as an
extension of the frequency-to-time mapping operation from Section 5.2.2
to the incoherent regime (Torres-Company, Lancis, & Andr´
es,2007b).
Figure 29 represents an experimental verification of this frequency-to-time
mapping operation. As spectrally incoherent source we used the ASE from
the emission spectrum of an Erbium-doped fiber amplifier. The spectral
density is measured with an OSA and sketched in Figure 29(a). Typical
coherence times of EDFAs are around 100 fs. The deterministic modulator
was a 10 Gb/s LiNbO3electro-optic modulator (EOM) biased at quadra-
ture and driven with a short electrical Gaussian impulse. The resulting
pulse has a Gaussian shape with 88 ps full-width at half maximum
(FWHM) duration. In this way, the quasi-homogeneous requirement is
satisfied. Finally, the modulated light is launched into an SMF of 1.44-km
length. Note that this GDD amount is much larger than the minimum
required by Equation (67), assuming β2= −21 ps2/km. The output inten-
sity shape is measured again with the same sampling scope and the result
is shown in Figure 29(b). The frequency-to-time mapping predicted by
Equation (70) is thus achieved.
The relevance of this mapping is very similar to that previously
exposed in the coherent regime (see Section 5.2.2). There we learned that
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 59 — #59
Temporal Optics in the Noncoherent Regime 59
1520 1530 1540
Wavelength (nm)
(a) (b)
Time (ns)
Optical spectrum (a.u.)
Optical intensity (a.u.)
1550 1560 1570 0 0.5 1.0 1.5 2.0
FIGURE 29 Experimental verification of the incoherent frequency-to-time mapping.
The optical spectrum of an incoherent light source can be mapped into the time
domain by modulation and subsequent dispersion in a GDD medium, in a similar
manner to the coherent effect explained in Section 5.2.2. After statistical averaging,
the output intensity profile becomes a scaled replica of the optical spectrum. Once the
mapping is achieved, adding more dispersion does not change the intensity profile,
only its scale.
the coherent frequency-to-time conversion is allowed for capturing opti-
cal spectra in a high-speed manner using broadband coherent pulses,
with potential applications in absorption spectroscopy, imaging, arbitrary
RF waveform generation, sensing, and reflectometry. In the noncoherent
regime, Equation (70) establishes that high-speed spectral measurements
can be done too using an incoherent light source instead.
However, a word of caution must be brought up here. Unlike in the
coherent regime, where the spectral measurements can be done in sin-
gle shot (i.e., the mapping takes place for a single individual pulse), in
the incoherent regime we note that Equation (70) is valid only for aver-
age quantities. In other words, both the spectral density of the input light
source and the acquired intensity are constructed upon an ensemble aver-
age operation. In practical terms, this means that one needs a statistically
significant amount of data (i.e., individual intensity shots) to construct
the quantity to be measured, which in turn reduces the operation speed.
In general, the SNR of the acquired spectra asymptotically increases as
N(Dorrer,2009), while the spectrum acquisition speed gets reduced
by N, thus indicating a trade-off between SNR and acquisition speed in
incoherent frequency-to-time mapping. As an example, for an SNR of 20
dB and an initial clock rate of 1 GHz (i.e., the active modulator repeti-
tion rate), it would be possible to achieve optical spectra varying at a
rate of 100 kHz, which is nevertheless quite competitive with the per-
formance acquired with the ultrafast coherent spectrometer presented at
Section 5.2.2.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 60 — #60
60 Space-Time Analogies in Optics
6.4.2. Arbitrary RF-Waveform Generator
A possible application of the generalization of the frequency-to-time map-
ping operation to the incoherent case lies in the field of arbitrary RF
waveform generation. Let us consider a scheme like the one presented in
Figure 30. A spectrally incoherent stationary light source is manipulated
in amplitude with a reconfigurable filter. Similar to the previous section,
the modified optical source is actively modulated with an electro-optic
modulator providing optical pulses with a temporal duration larger than
the inverse of the spectral resolution of the shaper in order to satisfy the
condition of quasi-homogeneous light pulse. After this, the modulated
light signal is launched into a GDD medium satisfying condition (67).
Then, as predicted by Equation (70), the synthesized spectral density is
transferred to the averaged temporal profile. This pulse is finally mapped
into the electrical domain by using a high-speed photodiode (Torres-
Company, Lancis, & Andr´
es,2006b). A possible implementation of this
scheme can be done with the setup of Figure 30, where the reconfig-
urable spectral filter is achieved with a Fourier transform pulse shaper
(Torres-Company et al.,2008a). Two examples of the synthesized spec-
tra and the achieved intensity profiles acquired with a sampling scope
are displayed in Figure 31, when using a short length of fiber for the
frequency-to-time mapping operation. The fiber length is the minimum
FIGURE 30 Schematic representation of arbitrary RF waveform generation based on
incoherent pulse shaping. In the optical implementation a reflective Fourier-transform
pulse shaper is used as reconfigurable filter.
Electrical signal
Applied mask
Dispersive element
O/E
conversion
Time
Deterministic
modulator
Spectrally
incoherent
source
Pulse shaper
Reflective LCOS
display
Lens
Grating
Collimator
Circulator Mirror
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 61 — #61
Temporal Optics in the Noncoherent Regime 61
1530 1532
(a)
1534
ES linear scale (a.u.)
λ (nm)
1536 1538 1540
(g)
(c)(b)
(
f
)
500 ps 500 ps
500 ps
500 ps
02468
f (GHz)
(d)
RF spectrum (10dB/div)
10 12 14
(
h
)
RF spectrum (10dB/div)
02468
f
(GHz)
10 12 14
(
e
)
ES linear scale (a.u.)
1525 1530 1535
λ (nm)
1540
FIGURE 31 Two examples (one per row) of high-bandwidth electrical signals
synthesized with the incoherent technique of Figure 30. The synthesized optical
spectra with a Fourier-transform pulse shaper are shown in (a) and (e). The achieved
intensity profiles are shown in (b) and (f) without performing average. A reduced SNR
is apparent from the figure. For 4×average, the waveform fidelity increases as seen
from (c) and (g). The corresponding electrical RF spectra are displayed in (d) and (h),
reflecting the high-bandwidth content of the synthesized signals.
to guarantee the far-zone operation while at the same time maximizes the
achievable RF bandwidth. The corresponding RF spectra synthesized with
this method are also displayed in the figure, leading to a relatively broad
frequency content (0–10 GHz). As mentioned in the previous section,
we also note the reduced signal-to-noise ratio (SNR) for the synthesized
waveforms when no statistical averaging is performed on the signal. This
indicates that this technique is more suitable for RF-AWG applications not
requiring single-shot high waveform fidelity, like spectral measurements
of broadband optoelectronic equipment (Torres-Company et al.,2008b).
6.5. Temporal Lau Effect
The temporal Talbot effect explained in Section 5.8 can also be extended
to the noncoherent regime. The spatial noncoherent Talbot effect is known
in the Fourier optics literature as the Lau effect (Jahns & Lohmann,1979).
This effect consists of overlapping several temporal self-images coming
from different light sources. The temporal counterpart of this effect was
firstly reported by Zalvidea, Duchowicz, and Sicre (2004), who proposed
the use of different mode-locked pulse trains as different light sources.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 62 — #62
62 Space-Time Analogies in Optics
An alternative implementation is by Lancis et al. (2006), who proposed
the use of several CW light sources that are externally modulated by a
common electro-optic modulator. This proposal has been recently verified
experimentally by Torres-Company, Fern ´
andez-Pousa, and Chen (2009),
highlighting the potential benefits of the temporal Lau effect for ultrafast
photonic applications.
Such an implementation can be described as follows. In the one-
dimensional (1D) spatial case, sketched in Figure 32(a), a quasi-
monochromatic spatially incoherent source is imaged onto a 1D grating
(G1) thus producing an array of incoherent spots, equidistant in the
transversal direction to the propagation. Since G1 is located at the focal
plane of a lens, each of the spots produces a tilted plane wave that
illuminates a second grating (G2). By diffracting a distance corresponding
to an integer Talbot plane, each of the plane waves produces a self-image
shifted in the transversal coordinate. Owing to the incoherent nature of
the illuminating points, the Talbot images are summed in intensity. If
the distance between consecutive spots satisfies a certain relation, the
intensities overlap. Far from being a simple curiosity, this phenomenon
(a) (b)
GDD circuit
Modulator
T
Laser 0
Laser 1
Laser 2
Laser L1
Input pulse
train
ω0
ω0
Δω
φ2
ω1ω2ωL1
1 st self-image
T
+
1
2
3
L
G1 f
Incoherent
point sources
Lens
1st
Self-image
(incoherent
superposition)
G2
k1
k2
k3
kL
zT
ω1
ω2
ωL1
FIGURE 32 Schematic representation of the Lau effect in (a) spatial domain. Here, an
input array of equidistant spatially incoherent emitters (G1) at the back focal plane of
a lens emits rays in different directions illuminating a grating (G2). Each emitter
produces a self-image of the G2 after propagating a Talbot distance. If the distance
between emitters satisfies a certain criterion, each of the self-images will overlap in
intensity. (b) Shows the corresponding scheme in the temporal domain. An array of
CW lasers equally spaced in frequency plays the role of the incoherent spots at the
focal plane in (a). The lasers are modulated externally by a temporal periodic function
emulating the role of G2. Each of these lasers generates a temporal self-image. When
the spectral distance between lasers satisfies a certain criterion, the superposition of
the temporal self-images becomes in intensity.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 63 — #63
Temporal Optics in the Noncoherent Regime 63
has been used for spatial signal processing applications (see, e.g., Patorski,
1989). Following Lancis et al. (2006), we now consider the temporal analog
sketched in Figure 32(b). An array of LCW lasers with equally spaced
frequencies plays the role of the tilted plane waves, with being the
spectral separation. The monochromatic carriers are modulated externally
by a repetitive sequence with repetition rate f, playing the role of G2. Then
the light is launched into a GDD circuit that satisfies the integer temporal
self-imaging condition [Equation (46) with Q=2]. In view of the analysis
carried out at Section 5.1, we know that consecutive carriers undergo a
relative shift provided by Equation (26). Thus, if
f T =m, (71)
where mis an integer number, there will be an overlapping in the out-
put temporal envelopes. This overlapping can be in intensity as long as
the photodetector bandwidth is shorter than the possible beat between
the highest and lowest modulation sidebands of any pair of consecutive
lasers, thus leading to
Iout(t)=
L1
X
n=0
PnIin(tnT), (72)
where Iin(t)is the temporal intensity profile provided by the external
modulator and Pnis the relative power of the nth carrier.
An illustrative example of this effect is shown in Figure 33. For sim-
plicity, only two carriers are used. The corresponding input and output
0 100
ω1
ω2
ω1
ω1and ω2
ω2
200
Time (ps)
(a) (b) (c)
0 100 200
Time (ps)
0 100 200
Time (ps)
Input intensity (a.u.)
Temporal
self-images (a.u.)
Intensity out (a.u.)
FIGURE 33 Experimental results for the temporal Lau effect. (a) Input sequence for
each of the corresponding frequencies. (b) Achieved first temporal self-image for the
corresponding carriers. We observe how the frequencies have been properly selected
so that the self-images are perfectly synchronized in time. (c) Intensity superposition
of the self-images.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 64 — #64
64 Space-Time Analogies in Optics
0 100 200
Time (ps)
Intensity out (a.u.)
FIGURE 34 A small detuning from the frequency spacing in the temporal Lau effect
can lead to completely different shapes by intensity superposition, opening the way
to alternative electrical pulse shaping mechanisms.
intensity profiles are shown for each frequency. We observe that each
of the sources generates a perfect self-image. The separation between
the light sources is such that it ensures intensity superposition, as illus-
trated in Figure 33. This way to achieve “incoherence” guarantees that
the generated electrical signal is free of phase-induced noise, thus offer-
ing nice features for the generation of arbitrary RF electrical impulses.
Indeed, thanks to the dispersion-wavelength controllable delay, together
with the fact that the individual lasers’ power can be easily controlled,
this temporal Lau effect offers a tremendous potential for arbitrary RF-
waveform generation (Torres-Company, Fern ´
andez-Pousa, & Chen,2009).
As an example, Figure 34 shows how a slight detuning from the perfect
self-imaging superposition condition can lead to a high-speed pulse train
consisting of triangle-like pulses.
7. TEMPORAL OPTICS IN THE TWO-PHOTON REGIME
The previous sections dealt with light sources whose properties can be
described using the laws of classical physics. However, some recent exper-
iments show that there are other properties and effects which are better
understood within the framework of a quantum theory (Mandel & Wolf,
1995). A particularly relevant example of this is the entanglement between
particles.
In optics, the most versatile method for producing entangled photon
pairs (or two-photon states) has been spontaneous parametric down-
conversion (SPDC), where an intense pump beam impinges onto a
second-order nonlinear medium, eventually producing a pair of lower
frequency photons in a nonseparable state (Shih,2003). These photon
pairs may be entangled in polarization, momentum, space, frequency, and
time. There exist alternative schemes for producing entangled photons, for
example through electromagnetically induced transparency, where two
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 65 — #65
Temporal Optics in the Two-Photon Regime 65
pump beams interact with a cloud of cold atoms with Raman-type transi-
tions (Balic et al.,2005). More compact solutions have been reported in the
last few years, including spontaneous four-wave mixing in microstruc-
tured fibers (Alibart et al.,2006) or silicon nano waveguides (Sharping
et al.,2006), and two-photon emission from semiconductor quantum wells
(Hayat, Ginzburg, & Orenstein,2008).
Two-photon states are a key resource for quantum information process-
ing applications, including cryptography (Ekert et al.,1992), teleportation
(Bouwmeester et al.,1997), or metrology (Giovannetti, Lloyd, & Maccone,
2001), just to name a few.
The aim of this chapter is to establish some counterpart ultrafast sig-
nal processors for broadband two-photon light emitted from SPDC. The
similarity with the previous schemes and their relevance with quantum
information applications will be pointed out.
7.1. Quantum Two-Photon Correlation Functions
and Two-Photon States
For light in an arbitrary quantum state, the probability of observing a pho-
ton at a time t1and another at t2is proportional to the intensity correlation
function (Glauber,1963)
G(2)
12 (t1,t2)=D9|ˆ
E(t1)ˆ
E(t2)ˆ
E+(t2)ˆ
E+(t1)|9E, (73)
where ˆ
E(t)and ˆ
E+(t)denote the negative- and positive-frequency com-
ponents of the optical field operator at time instant t, and |9irepresents
the vector state.
For the particular case of light in a two-photon state, the right-
hand side of Equation (73) factorizes as (Shih,2003)h9|ˆ
E(t1)ˆ
E(t2)
ˆ
E+(t2)ˆ
E+(t1)|9i=h9|ˆ
E(t1)ˆ
E(t2)|0ih0|ˆ
E+(t1)ˆ
E+(t2)|9i, with |0ibeing
the vacuum state. Therefore, we can write as
G(2)
12 (t1,t2)=|912(t1,t2)|2, (74)
with
912(t1,t2)=D0|ˆ
E+(t1)ˆ
E+(t2)|9E. (75)
Then, 912(t1,t2)can be interpreted as the probability amplitude distribu-
tion of the two-photon state or two-photon wave packet.
The above discussion has a dual interpretation in the frequency
domain. In this case, the meaningful quantity is the frequency-domain
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 66 — #66
66 Space-Time Analogies in Optics
two-photon probability amplitude, ˜
9121,ω2), whose modulus square
determines the probability of detecting a photon at frequency ω1and
the other at frequency ω2. This function forms with 912(t1,t2)a Fourier
transform pair, that is,
912(t1,t2)=ZZ ˜
9121,ω2)exp [i1t1+ω2t2)]dω1dω2, (76)
and
˜
9121,ω2)=1
4π2ZZ 912(t1,t2)exp [i1t1+ω2t2)]dt1dt2. (77)
Obviously, when we fix the two-photon probability amplitude in the time
domain, the properties in the dual (frequency) space are automatically
settled.
7.2. Tailoring Two-Photon States in Spontaneous Parametric
Down-Conversion
In this section, because of its high versatility and easy tuning, we concen-
trate in the particular method of SPDC for generating two-photon states.
As mentioned before, in SPDC, a high-power pump beam impinges onto
a nonlinear crystal, eventually producing two photons at lower frequency,
usually referred for historical reasons as signal and idler (Shih,2003).
Energy and momentum are conserved in this process. The two-photon
probability amplitude can be written as 912(t1,t2)=D0|ˆ
E+
s(t1)ˆ
E+
i(t2)|9E,
where ˆ
E+
s,i(t)is the positive-frequency electric-field operator at time t
for the signal and idler photon, respectively. We are concentrating in
the temporal distribution only because we consider a single-mode spa-
tial structure (which could be achieved by imaging the down-converted
beams onto a waveguide with translational symmetry, like, for example,
a single-mode fiber). But we remark that a similar analysis can be car-
ried out in the spatial domain. The structure of the two-photon state, |9i,
depends on the particular arrangement used for the SPDC process, which
includes the spatiotemporal distribution of the pump beam, the material
intrinsic and synthesized properties, and the optical system to bring the
pump source onto the crystal. The interpretation in the spectral domain is
straightforward. The spectral two-photon probability amplitude becomes
˜
9121,ω2)= h0|ˆ
˜
E+
s1)ˆ
˜
E+
i2)|9i, where ˆ
˜
E+
s,i(ω) is the positive-frequency
field operator for the signal and idler photons, respectively. From Equa-
tions (76) and (77), it is clear that 912(t1,t2)and ˜
9121,ω2)still form a
Fourier transform pair.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 67 — #67
Temporal Optics in the Two-Photon Regime 67
In the same way that the synthesis of the electric field through pulse
shaping techniques constitutes the workhorse for numerous applications
based on signal processing, in quantum information processing based on
two-photon states, the important magnitude to be tailored is the two-
photon wave packet, 912(t1,t2). Certainly, one could write another review
paper about the current techniques that permit to synthesize this mag-
nitude. We will just briefly mention, for example, the use of appropriate
materials and wavelengths (Kuzucu et al.,2005;Mosley et al.,2008); intro-
ducing angular dispersion into the pump and down-converted beams
(Hendrych, Micuda, & Torres,2007); inserting phase (Pe’er et al.,2005)
and/or amplitude (Zah, Halder, & Feurer,2008) spectral filters in sig-
nal and idler beams; using chirped quasi-phase-matched materials (Nasr
et al.,2008); the so-called spatial-to-spectral mapping technique (Valencia
et al.,2007); and the combination of angular dispersion and noncollinear
geometries (Shi et al.,2008).
We will not go into the details on how to achieve them, but rather
we shall concentrate on the remarkable implications derived from the
912(t1,t2)structures that lead to unique photon correlations. For exam-
ple, when 912(t1,t2)912 (t2t1), the signal and idler photons are said
to be temporally correlated. This can be better understood by looking into
Figure 35, which depicts the two-photon probability amplitude in both
time and frequency domains. Once the signal photon “clicks” its cor-
responding detector at time instant t1, the idler photon will then click
the other within a very narrow temporal uncertainty given by the two-
photon coherence time [which is the temporal width of the antidiagonal
of 912(t1,t2)]. This temporal structure is the natural shape appearing in
SPDC when pumped by a powerful beam with a very narrow line width
(see, e.g., Hong & Mandel,1985;Shih et al.,1994;Gatti et al.,2003).
Based on the Fourier properties, temporal correlation implies frequency
anticorrelation, that is, ˜
9121,ω2)81)δ(ω1+ω2), indicating that the
sum of the frequencies in the photon pair gives back the frequency of
the pump beam, which is consistent with energy conservation in the
SPDC process [see Figure 35(a)]. Here, the function 8(ω) is the two-
photon spectrum (often called biphoton spectrum) an forms a Fourier
transform pair with 912=t2t1)in the variable τ. Thus, the inverse
of its spectral width determines the coherence time of the second-order
intensity correlation. On the other hand, the width of the diagonal of
˜
9121,ω2)in the 1,ω2)plane is related with the pump line width
and therefore approaches a Dirac delta. An unequivocal signature of the
quantum character of these photon correlations is that the width of the
temporal correlation and the width of the frequency anticorrelation are
not bounded by a Fourier relation (Khan & Howell,2006). For example,
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 68 — #68
68 Space-Time Analogies in Optics
1
Time t1 (a.u.)
0
(a)
1
Time t2 (a.u.)
0
1
1
1
Frequency ω2 (a.u.)
0
1
1
Frequency ω1 (a.u.)
0
(b)
1
1
1
Time t1 (a.u.)
Time t2 (a.u.)
0
0
(c)
1
1
1
Frequency ω2 (a.u.)
0
1
1
1
Time t1 (a.u.)
Time t2 (a.u.)
0
0
(e)
1
1
1
Frequency ω2 (a.u.)
0
1
1
Frequency ω1 (a.u.)
0
(d)
11
Frequency ω1 (a.u.)
0
(f)
1
FIGURE 35 Schematic representation of different two-photon correlations. Upper
row shows the two-photon wavepacket, 912(t1,t2), and lower row shows the
corresponding frequency-domain representation, ˜
9121,ω2), for: (a) (b) temporally
correlated biphoton; (c) (d) frequency-correlated biphoton; and (e) (f) uncorrelated
biphoton.
having strong and simultaneous frequency anticorrelation and temporal
correlation may lead to important implications in clock synchronization
(Valencia, Scarcelli, & Shih,2004) and cryptography (Khan & Howell,
2006).
Another interesting shape of the two-photon probability amplitude is
the dual of the previous case. Here, the signal and idler photons are
correlated in frequency instead, meaning that ˜
9121,ω2)81ω2).
Now, if we were able to detect the signal photon at frequency ω1, then
the idler would have the same frequency within an uncertainty given
by the width of the antidiagonal of ˜
9121,ω2)in the 1,ω2)plane, as
can be understood from Figure 35(b). Based on the Fourier relations, fre-
quency correlation leads to temporal anticorrelation, that is,912(t1,t2)
0(t1(t1+t2). Here, the temporal function 0(t)forms a Fourier trans-
form pair with 8(1ω =ω1ω2)in the variable . The most remarkable
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 69 — #69
Temporal Optics in the Two-Photon Regime 69
feature of these two-photon states is that the temporal anticorrelation
width is independent of the biphoton function 8(1ω =ω1ω2). These
two-photon states are a valuable resource for timing and positioning
metrology because the precision of the measurements can be enhanced
with respect to the standard quantum limit by a factor of 2 (Giovannetti,
Lloyd, & Maccone,2001). A possible scheme for achieving frequency cor-
related photon pairs from SPDC involves the use a broadband pump pulse
with proper engineering of the phase-matching conditions (Price, U’ Ren,
& Walmsley,2001). The first demonstration of this technique was reported
by Kuzucu et al. (2005).
Finally, another increasingly interesting shape of the probability ampli-
tude is the corresponding to a separable state, for which 912(t1,t2)=
f(t1)g(t2). Obviously, separability in the time domain also implies sepa-
rability in the spectral domain. This means that the two photons in the
pair are fully uncorrelated. This kind of light source is very interest-
ing for the implementation of single-photon sources based on a concept
known as heralding, where the idea relies on detecting one photon,
thereby revealing the presence of its twin. Poorly speaking, if the shape
of ˜
9121,ω2)is the closest to a circle in the 1,ω2)plane, the two
photons would have the same spectral shape, thus becoming spectrally
indistinguishable. Besides, the separability in the two-photon probability
amplitude ensures that each photon is in a single spectral mode (ensuring
a high purity). The typical approach to achieve this shape was by narrow-
band filtering (Ou, Rhee, & Wang,1999). However, having a separable
two-photon probability amplitude with broad bandwidth (i.e., directly
achieved without filtering) remained a challenge for quite a long time. The
generation of ultrabroad pure heralded single photons constitutes a major
step towards the implementation of some quantum information protocols.
It is known that these appreciated wave packets can be achieved by a care-
ful manipulation of the phase-matching conditions in type-II SPDC when
pumped with a broadband pulse (Price, U’ Ren, & Walmsley,2001). The
first experimental report of these states was recently presented by Mosley
et al. (2008).
7.3. Propagation of Ultrafast Two-Photon Wave Packets
From the previous section, we have learned that there are several impor-
tant applications that rely on a particular shape of the two-photon wave
packet, 912(t1,t2), and that this can be adequately tailored with a careful
engineering of the SPDC process. In this section, we show how the two-
photon wave packet propagates, after its generation, through the sort of
processing systems that we have studied in Section 5. We shall find that
they can assist us to manipulate further the engineered two-photon states.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 70 — #70
70 Space-Time Analogies in Optics
Ki(t, t)
G(2)(t1, t2)
Ks(t, t)
Nonlinear
crystal
Signal
Idler
Pump
FIGURE 36 General scheme for two-photon propagation through linear systems. The
two-photon pair is generated through interaction in a nonlinear crystal. Signal and
idler are physically separated and travel different linear systems, which are
mathematically described by the corresponding kernel. Later the photons are
detected in coincidence with ideally ultrafast detectors, leading to the measurement
of the Glauber intensity correlation function.
Let us consider a scheme like the one presented in Figure 36. Each of
the photons from a pair in a properly tailored SPDC process is launched
into a different linear processing system. To study the propagation of
the two-photon probability amplitude, 912(t1,t2)= h0|ˆ
E+
s(t1)ˆ
E+
i(t2)|9i,
we just have to assess the change of the electric-field operator, ˆ
E+
s,i(t). Giv-
ing a further step, we can rewrite this electric-field operator as ˆ
E+
s,i(t)=
ˆ
ψs,i(t)exp[i0s,i tβ0s,iz)], where ω0s,i is the central frequency of the sig-
nal (idler) photon and β0s,i is the corresponding phase factor. Analogously
to what we did in Section 4.1, we just have to study the evolution of the
envelope operator. Since the systems are linear, the transformation of the
complex envelope operators is described by
ˆ
ψs,out(t)=Zˆ
ψs,in(t0)Ks(t,t0)dt0, (78a)
ˆ
ψi,out(t)=Zˆ
ψi,in(t0)Ki(t,t0)dt0, (78b)
which is an obvious quantum generalization of Equation (9). Here, the ker-
nel Ks(i)(t,t0)corresponds to the system through which the signal (idler)
photon propagates. An important consequence of these equations, is that
the system operates individually on the electric field of the corresponding
photon. The implications of this statement will be clear later. In any case,
the measurable quantity is the joint-probability detection, which is pro-
vided by the modulus square of the propagated two-photon probability
amplitude, 912,out(t1,t2). From the previous equations, it is easy to find
that this magnitude evolves as
912e,out(t1,t2)=ZZ Ks(t1,t0
1)Ki(t2,t0
2)912e,in(t0
1,t0
2)dt0
1dt0
2, (79)
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 71 — #71
Temporal Optics in the Two-Photon Regime 71
where the subscript e refers to the envelope part of 912(t1,t2). Of course,
alternatively, the two-photon wave packet can be tracked in the spectral
domain,
˜
912b,out (ω1,ω2)=ZZ ˜
Ks1,ω0
1)˜
Ki2,ω0
2)˜
912b,in ω0
1,ω0
2dω0
1dω0
2, (80)
where the subscript b refers to the baseband two-dimensional function
˜
9121,ω2).
7.4. Two-Photon Temporal Far-Field Phenomenon
This section deals with the extension of the technique presented in
Section 5.2.2 to the two-photon regime. For simplicity, we shall consider
the case of spectrally anticorrelated photons that naturally arise from
CW-pumped SPDC. We assume a non-collinear geometry, so that after
generation each photon from the pair is launched onto a different GDD
medium. The output two-photon probability amplitude can be calcu-
lated using Equation (79), considering the kernel for the corresponding
GDD medium and reminding that the generated photons are frequency
anticorrelated,
912e,out(t1,t2)=Zexp i82eff
2ω28(ω) exp[i(t2t1]dω, (81)
where 8(ω) is the biphoton spectrum and the GDD parameters are
grouped into the coefficient 82eff =82s +82i, with 82s(i) being the GDD
parameter of the dispersive circuit that the signal (idler) photon traverses.
We note the striking similarity of this equation with the one that governs
the distortion of an ultrashort light pulse. Thanks to this similarity that we
can adapt some of the learned notions in the previous sections to this new
exciting field. For example, it is clear the two-photon probability ampli-
tude will spread by increasing 82eff. If this parameter keeps increasing to
the point that the inequality [which is just an extension of Equation (28)]
82eq t2
c/(4π) (82)
is satisfied, with tcbeing the biphoton coherence time, the output wave
packet becomes
912e,out (t1,t2)=exp "i(t2t1)2
282eff #8t2t1
82eff , (83)
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 72 — #72
72 Space-Time Analogies in Optics
800
600
400
200
025 30 35 40 45 50
t (ns)
Number of counts
FIGURE 37 Experimental demonstration of the two-photon temporal far-field
phenomenon. The measured two-photon correlation function in a CW-pumped type-II
SPDC becomes a scaled replica of the expected biphoton spectrum. The
frequency-to-time mapping occurs thanks to the distortion of in the shape introduced
by the GDD circuits (SMF in the experiments). (Reprinted with permission from
Valencia et al., 2002). c
[2002] APS.
which is a scaled replica of the biphoton spectrum. Here, the scaling factor
is provided by the effective coefficient 82eff. This frequency-to-time map-
ping was first discovered by Valencia et al. (2002), and it can be used to
measure the biphoton spectrum without the need of scanning monochro-
mators, with a spectral resolution given by δω =δt/82eff, where δtis the
temporal resolution of the photon coincidence detectors (again ensuring
that the minimum spectral feature is properly mapped into the spectral
domain). An example of this technique is shown in Figure 37, where the
two-photon coincidence detection is plotted together with the expected
biphoton spectrum calculated from the phase-matching conditions for the
particular SPDC arrangement.
While we have focused on the case of CW-pumped SPDC, the technique
can be easily applied for other schemes with different two-photon probabil-
ity amplitudes. Then, if the joint-probability detection is measured in the
whole (t1,t2)plane and not only in the line (t2t1), it becomes possible
to build a spectrogram that calculates the joint-probability detection at
two frequencies using the same equipment for calculating the probability
detection in time (Avenhaus et al.,2010).
The challenge in this technique relies on overcoming the timing jit-
ter limited resolution in state-of-the-art photon detectors and introducing
high dispersion with minimal photon losses in order to get a spectral
resolution competitive with more conventional spectral devices.
7.5. Two-Photon Temporal Imaging
Recalling Figure 36, we have already commented that each linear tempo-
ral system acts individually on the envelope of the electric-field operator
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 73 — #73
Temporal Optics in the Two-Photon Regime 73
of the corresponding photon. Then, we can use the notions exposed in
Section 5 for Gaussian temporal systems. In particular, let us consider
a temporal imaging system introduced in the arm of the signal pho-
ton, while the remaining idler beam unaltered. The transformation of the
complex envelope operator will be (Tsang & Psaltis,2006)
ˆ
ψs,out(t)=exp hiK0t2/2iˆ
ψs,in(t/m), (84a)
ˆ
ψi,out(t)=ˆ
ψi,in(t0). (84b)
Thus, apart from an irrelevant quadratic phase factor with chirping K0,
the output complex envelope operator becomes a scaled replica of the
input one. The scaling factor is the same as the one provided by a classical
temporal imaging system, m=822/821 . In other words, it is possible to
change the scaling of the photon correlations in one of the variables (t1or
t2)while leaving unaffected the other one. As pointed out by Tsang and
Psaltis (2006), by using a temporal imaging system with m= −1, one can,
in principle, achieve a temporally anticorrelated biphoton starting from a
temporally correlated two-photon wave packet.
7.6. Some Remarks on Ultrafast Two-Photon Processors
It is not difficult to think about different alternative configurations to tune
the correlations of the photons generated from SPDC using the ultrafast
processing schemes that we have studied in Section 5. For example, with
a proper engineering of the spectral scaling factor, the spectral imaging
system from Section 5.5 could be analogously used to create a spec-
trally anticorrelated biphoton from a spectrally correlated one and vice
versa. Or the 4f ultrafast Fourier processing systems could be used to
process in-line the temporal and spectral structure of more complicated
two-photon wave packets.
However, no operation of time lensing at a single-photon scale has
been reported yet. At a low-photon level, implementing time lens config-
urations based on nonlinear optics constitutes a major challenge, owing
to the intrinsic low-power regime operation. The EOPM time lens con-
stitutes in this case the most reliable alternative. In this direction, there
exist some ultrafast Fourier processing schemes based on the use of an
EOPM (Kolchin et al.,2008;Sensarn, Yin, & Harris,2009), but they are
not strictly operating as a time lens. Nevertheless, the impressive results
achieved in these experiments indicate that there is huge avenue to adapt
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 74 — #74
74 Space-Time Analogies in Optics
the notions that we have acquired in classical temporal optics to process
temporal two-photon wave packets.
8. CONCLUSIONS
The space-time analogy establishes a mathematical connection between
the equations that govern the paraxial diffraction of a light beam and
those describing the linear distortion of a short light pulse in a dispersive
medium. Far from being a simple curiosity, this analogy has permitted
us to identify and design new photonic devices and systems, as well as
reinterpret physical phenomena, bringing enhanced capabilities for pro-
cessing ultrafast optical pulses. The key of this success relies on the fact
that we can borrow well-established concepts from the field of classical
Fourier optics and adapt them to the particular problem or situation in
the temporal domain.
We have also seen that while most of the research efforts have been
devoted to explore the coherent temporal counterpart situations, the anal-
ogy is not at all restricted to coherent signals. Actually, in the last years,
we have seen increasing efforts to perform incoherent processing of short
pulses. In this direction, there are important challenges associated with
the fact that incomplete coherence necessarily implies dealing with the
stochastic nature of light. From a practical perspective, this means that
one has to balance the advantages of having incoherent sources with the
extra noise associated with the system’s noise performance.
Finally, we have also observed that the space-time analogy can be
formally extended to the field of quantum optics. In particular, owing
to the ultrafast nature of the temporal correlations of entangled two-
photon states, this kind of light source can be manipulated using simi-
lar arrangements to those used for ultrafast coherent signal processors.
In this direction, some processing schemes have already been experi-
mentally realized, while some exciting proposals are still waiting to be
achieved.
The continuous and impressive advances in the fabrication of new
nanophotonic devices, having ultra-high nonlinearities with low peak
power and offering the possibility for dispersion engineering; ultrafast
lasers, with ever more compact and robust schemes, delivering ultrashort
pulses with wavelength tuning and high stability; ultra-high-speed digi-
tal acquisition systems, with a continuous scaling in electronic bandwidth
and real-time performance; and finally quantum light sources with inte-
gration capabilities and high brightness, permit us to conclude that there
is still plenty of room to keep extending and performing new processing
systems conceived within the concepts of this fruitful analogy.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 75 — #75
References 75
ACKNOWLEDGMENTS
We are very grateful to many people with whom we have commented
the ideas exposed in this review and have directly or indirectly partici-
pated in some of the experimental results provided. Lawrence R. Chen
(McGill University, Canada) deserves special gratitude, as the experimen-
tal results presented in this review were taken in his fiber optics laboratory
(unless otherwise stated). We thank Jos´
e Aza˜
na (Institut National de la
Recherche Scientifique, INRS, Canada) and Alejandra Valencia (Institute
of Photonic Sciences, ICFO, Spain) for the careful reading and insightful
comments on the contents of the chapter. Hanna Lajunen, Enrique Sil-
vestre (Universitat de Valencia, Spain), Ari T. Friberg (Royal Institute of
Technology, KTH, Sweden), Juan P. Torres, (Institute of Photonic Sciences,
ICFO, Spain), J. Ojeda-Casta˜
neda (University of Guanajuato, Mexico) and
Carlos R. Fern´
andez-Pousa (University of Elche, Spain) are also gratefully
acknowledged for sharing comments about part of the contents exposed
in this review. Thanks also to Keisuke Goda (University of California,
Los Angeles, USA) for providing us with the material of Figure 7 and
to Antonio Malacarne and Jos´
e Aza˜
na (Institut National de la Recherche
Scientifique, INRS, Canada) for providing Figure 19. We also thank the
principal investigators for each of the works explicitly highlighted in
figures for granting us permission to use their results.
Victor Torres acknowledges the Spanish Ministry of Science (MICINN)
and the Spanish Foundation for Science and Technology (FECYT) for
a postdoctoral fellowship. This work has been partially funded by
the National Science and Engineering Research Council (Canada) and
MICINN, Spain (under projects FIS2010-15746 and Consolider program
CSD2007-00013).
REFERENCES
Agrawal, G. P. (2007). Nonlinear fiber optics (4th ed.). New York: Academic Press.
Akhmanov, S. A., Sukhorukov, A. P., & Chirkin, A. S. (1969). Sov. Phys. JETP, 28, 748.
Alibart, O., Fuconis, J., Wong, G. K. L., Murdoch, S. G., Wadsworth, W. J., & Rarity, J. G.
(2006). New J. Phys., 8, 67.
Alic, N., Myslivets, E., Moro, S., Kuo, B. P.-P., Jopson, R. M., McKinstrey, C. J., & Radic, S.
(2010). J. Lightwave Technol., 28, 448.
Andrekson, P. A. (1993). Opt. Lett., 18, 1621.
Arahira, S., Kutsuzawa, S., Matsui, Y., Kunimatsu, D., & Ogawa, Y. (1998). J. Lightwave
Technol., 16, 405.
Arriz´
on, V., & Ojeda-Casta˜
neda, J. (1992). J. Opt. Soc. Am. A, 9, 1801.
Atkins, S., & Fischer, B. (2003). IEEE Photon. Technol. Lett., 15, 132.
Avenhaus, M., Eckstein, A., Mosley, P. J., & Silberhorn, C. (2010). Opt. Lett., 34, 2873.
Aza˜
na, J. (2003a). Opt. Commun., 217, 205.
Aza˜
na, J. (2003b). Opt. Lett., 28, 501.
Aza˜
na, J., Berger, N. K., Levit, B., & Fischer, B. (2005a). Opt. Lett., 30, 3228.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 76 — #76
76 Space-Time Analogies in Optics
Aza˜
na, J., Berger, N. K., Levit, B., & Fischer, B. (2005b). IEEE Photon. Technol. Lett., 17, 94.
Aza˜
na, J., Berger, N. K., Levit, B., Smulakovsky, V., & Fischer, B. (2004). Opt. Lett., 29, 2849.
Aza˜
na, J., & Chen, L. R. (2003). J. Opt. Soc. Am. B, 20, 1447.
Aza˜
na, J., Chen, L. R., Muriel, M. A., & Smith, P. W. E. (1999). Electron. Lett., 35, 2223.
Aza˜
na, J., Lugo, J. E., Kira, A. G., & Plant, D. V. (2005). Opt. Commun., 247, 257.
Aza˜
na, J., & Muriel, M. A. (1999). Opt. Lett., 24, 1672.
Aza˜
na, J., & Muriel, M. A. (2000). IEEE J. Quantum Electron., 36, 517.
Balic, V., Braje, D. A., Kolchin, P., Yin, G. Y., & Harris, S. E. (2005). Phys. Rev. Lett., 94, 183601.
Bennett, C. V., & Kolner, B. H. (1999). Opt. Lett., 24, 783.
Bennett, C. V., & Kolner, B. H. (2000). IEEE J. Quantum Electron., 36, 430.
Bennett, C. V., & Kolner, B. H. (2001). IEEE J. Quantum Electron., 37, 20.
Berger, N. K., Levit, B., Bekker, A., & Fischer, B. (2004). IEEE Photon. Technol. Lett., 16, 1855.
Bennett, C. V., Moran, D. V., Langrock, C., Fejer, M. M., & Ibsen, M. (2007). Proceedings of
Conference Lasers and Electro-Optics, doc. CFF1.
Bennett, C. V., Scott, R. P., & Kolner, B. H. (1994). Appl. Phys. Lett., 65, 2513.
Bhushan, A. S., Coppinger, F., & Jalali, B. (1998). Electron. Lett., 34, 1081.
Bhushan, A. S., Coppinger, F., Yegnanarayanan, S., & Jalali, B. (1999). Opt. Lett., 24, 738.
Bigourd, D., Chatel, B., Schleich, W. P., & Girard, B. (2008). Phys. Rev. Lett., 100, 030202.
Bjorkholom, J. E., Turner, E. H., & Pearson, D. B. (1975). Appl. Phys. Lett., 26, 564.
Bortnik, B., Poberezhskiy, I. Y., Chou, J., Jalali, B., & Fetterman, H. R. (2006). J. Lightwave
Technol., 24, 2752.
Bouwmeester, D., Pan, J.-, Mattle, K., Eibl, M., Weinfurter, H., & Zeilinger, A. (1997). Nature,
390, 575.
Capmany, J., & Novak, D. (2007). Nature Photon., 1, 319.
Caputi, W. J. (1971). IEEE Trans. Aero. Electron. Syst., AES7, 269.
Chantada, L., Fern´
andez-Pousa, C. R., & G´
omez-Reino, C. (2006). Opt. Commun., 266, 393.
Chi, H., & Yao, J. P. (2007). J. Lightwave Technol., 25, 3528.
Chou, J., Boyraz, O., Solli, D., & Jalali, B. (2007). Appl. Phys. Lett., 91, 161105.
Chou, J., Han, Y., & Jalali, B. (2003). IEEE Photon. Technol. Lett., 15, 581.
Chou, J., Rose, T. S., Conway, J. A., Valley, G. C., & Jalali, B. (2009). Opt. Lett., 34, 869.
Chou, J., Solli, D. R., & Jalali, B. (2008). Appl. Phys. Lett., 92, 111102.
Collins, S. A. (1970). J. Opt. Soc. Am., 60, 1168.
Coppinger, F., Bhushan, A. S., & Jalali, B. (1998). Electron. Lett., 38, 399.
Coppinger, F., Bhushan, A. S., & Jalali, B. (1999). IEEE Trans. Microwave Theory Techn., 47,
1309.
Dai, Y. T., Okawachi, Y., Turner-Foster, A. C., Lipson, M., Gaeta, A. L., & Xu, C., (2010). Opt.
Express, 18, 333.
Dijaili, S. P., Dienes, A., & Smith, S. J. (1990). IEEE J. Quantum Electron., 26, 1158.
Dorrer, C. (2004). J. Opt. Soc. Am. B, 21, 1417.
Dorrer, C. (2009). Opt. Express, 17, 3341.
Dorrer, C., & Maywar, D. N. (2004). J. Lightwave Technol., 22, 266.
Duchesne, D., Morandotti, R., & Aza˜
na, J. (2007). J. Opt. Soc. Am. B, 24, 113.
Ekert, A. K., Rarity, J. G., Tapster, P. R., & Palma, G. M. (1992). Phys. Rev. Lett., 69, 1293.
Esumi, Y., Kabir, M. D., & Kannari, F. (2009). Opt. Express, 17, 19153.
Fermann, M. E., Galvanauskas, A., & Sucha, G., (Eds.). (2003). Ultrafast lasers: technology and
applications. New York: Marcel Dekker.
Fern´
andez-Pousa, C. R. (2006). Opt. Lett., 31, 3049.
Fern´
andez-Pousa, C. R., Mateos, F., Chantada, L., Flores-Arias, M. T., Bao, C., P´
erez, M. V., &
G´
omez-Reino, G. (2004). J. Opt. Soc. Am. B, 21, 1170.
Ferrera, M., Razzari, L., Duchesne, D., Morandotti, R., Yang, Z., Liscindi, M., Sipe, J. E., Chu,
S., Little, B. E., & Moss, D. J. (2008). Nature Photon., 2, 737.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 77 — #77
References 77
Fetterman, H. R., Tannenwald, P. E., Parker, C. D., Melngailis, J., Williamson, R. C.,
Woskoboinikow, P., Praddaude, H. C., & Mulligan, W. J. (1979). Appl. Phys. Lett., 34, 123.
Feurer, T., Vaughan, J. C., & Nelson, K. A. (2003). Science, 299, 374.
Finot, C., Dudley, J. M., Kibler, B., Richardson, D. J., & Millot, G. (2009). IEEE J. Quantum
Electron., 45, 1482.
Fontaine, N. K., Scott, R. P., Cao, J., Karalar, A., Jiang, W., Okamoto, K., Heritage, J. P., Kolner,
B. H., & Yoo, S. J. B. (2007). Opt. Lett., 32, 865.
Fontaine, N. K., Scott, R. P., Zhou, L. J., Soares, F. M., Heritage, J. P., & Yoo, S. J. B. (2010).
Nature Photon., 4, 248.
Foster, M., Salem, R., Geraghty, D. F., Turner-Foster, A. C., Lipson, M., & Gaeta, A. L. (2008).
Nature, 456, 81.
Foster, M. A., Salem, R., Okawachi, Y., Turner-Foster, A. C., Lipson, M., & Gaeta, A. L. (2009).
Nature Photon., 3, 581.
Foster, M. A., Turner, A. C., Lipson, M., & Gaeta, A. L. (2008). Opt. Express, 16, 1300.
Fujiwara, M., Teshima, M., Kani, J., Suzuki, H., Takachio, N., & Iwatsuki, K. (2003). J.
Lightwave Technol., 21, 2705.
Gatti, A., Zambrini, R., San Miguel, M., & Lugiato, L. A. (2003). Phys. Rev. A, 68, 05387.
Geisler, D. J., Fontaine, N. K., He, T. T., Scott, R. P., Paraschis, L., Heritage, J. P., & Yoo, S. J. B.
(2009). Opt. Express, 17, 15911.
Geraghty, D. F., Salem, R., Foster, M. A., & Gaeta, A. L. (2008). IEEE Photon. Technol. Lett., 20,
487.
Giordmaine, J. A., Duguay, M. A., & Hansen, J. W. (1968). IEEE J. Quantum Electron., QE-4,
252.
Giovannetti, V., Lloyd, S., & Maccone, L. (2001). Nature, 412, 417.
Glauber, R. J. (1963). Phys. Rev., 130, 2529.
Goda, K., Solli, D. R., & Jalali, B. (2008). Appl. Phys. Lett., 93, 031106.
Goda, K., Tsia., K. K., & Jalali, B. (2009). Nature, 458, 1145.
Godil, A. A., Auld, B. A., & Bloom, D. M. (1994). IEEE J. Quantum Electron., 30, 827.
Goodman, J. W. (1996). Introduction to fourier optics (2nd ed.). New York: McGraw-Hill.
Gori, F. (2005). Opt. Lett., 30, 2840.
Gr ¨
uner-Nielsen, L., Wandel, M., Kristensen, P., Jorgensen, C., Jorgensen, L., V., Edvold, B.,
P´
alsdottir, B., & Jakobsen, D. (2005). J. Lightwave Technol., 23, 3566.
Guigay, J. P. (1971). Opt. Acta, 18, 677.
Haner, M., & Warren, W. S. (1988). Appl. Phys. Lett., 52, 1458.
Hayat, A., Ginzburg, P., & Orenstein, M. (2008). Nature Photon., 2, 238.
Hendrych, M., Micuda, M., & Torres, J. P. (2007). Opt. Lett., 32, 2339.
Hirooka, T., & Nakazawa, M. (2006). J. Lightwave Technol., 24, 2530.
Hirooka, T., & Nakazawa, M. (2008). IEEE Photon. Technol. Lett., 20, 1869.
Hong, C. K., & Mandel, L. (1985). Phys. Rev. A, 31, 2409.
Jahns, J., & Lohmann, A. W. (1979). Opt. Commun., 28, 263.
Jannson, T. (1983). Opt. Lett., 4, 232.
Jannson, T., & Jannson, J. (1981). J. Opt. Soc. Am., 71, 1373.
Jiang, Z., Huang, C. B., Leaird, D. E., & Weiner, A. M. (2007). Nature Photon., 1, 463.
Jiang, Z., Seo, D. S., Leaird, D. E., & Weiner, A. M. (2005). Opt. Lett., 30, 1557.
Jiang, Z. P., & Zhang, X. C. (1998). Appl. Phys. Lett., 72, 1945.
Kalestynski, A., & Smolinska, B. (1978). Opt. Acta, 25, 125.
Kauffman, M. T., Banyai, W. C., Godil, A. A., & Bloom, D. M. (1994). Appl. Phys. Lett., 64, 270.
Kelkar, P. V., Coppinger, F., Bhushan, A. S., & Jalali, B. (1999). Electron. Lett., 35, 1661.
Khan, I. A., & Howell, J. C. (2006). Phys. Rev. A, 73, 031801(R).
Klauder, J. R., Price, A. C., Darlington, S., & Albersheim, W. J. (1960). Bell Syst. Techn. J., 39,
745.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 78 — #78
78 Space-Time Analogies in Optics
Kobayashi, Y., Yao H., Amano, K., Fukushima, Y., Morimoto, A., & Sueta, T. (1988). IEEE J.
Quantum Electron., 24, 382.
Kolchin, P., Belthangady, C., Du, S. W., Yin., G. Y., & Harris, S. E. (2008). Phy. Rev. Lett., 101,
103601.
Kolner, B. H. (1988). Appl. Phys. Lett., 52, 1122.
Kolner, B. H. (1994a). IEEE J. Quantum Electron., 30, 1951.
Kolner, B. H. (1994b). J. Opt. Soc. Am. A, 11, 3229.
Kolner, B. H., & Nazarathy, M. (1989). Opt. Lett., 14, 630.
Komukai, T., Yamamoto, T., & Kawanishi, S. (2005). IEEE Photon. Technol. Lett., 17, 1746.
Kumar, S. (2007). Opt. Lett., 32, 346.
Kuzucu, O., Fiorentino, M., Albota, M. A., Wong, F. N. C., & Kartner, F. X. (2005). Phys. Rev.
Lett., 94, 083601.
Kuzucu, O., Okawachi, Y., Salem, R., Foster, M. A., Turner-Foster, A. C., Lipson, M., & Gaeta,
A. L. (2009). Opt. Express, 17, 20605.
Lajunen, H., Friberg, A. T., & Ostlund, P. (2006). J. Opt. Soc. Am. A, 23, 2530.
Lajunen, H., Tervo, J., Turunen, J., Vahimaa, P., & Wyrowski, F. (2003). Opt. Express, 11, 1894.
Lajunen, H., Vahimaa, P., & Tervo, J. (2005). J. Opt. Soc. Am. A, 22, 1536.
Lancis, J., Caraquitena, J., Andr´
es, P., & Muriel, M. A. (2005). Opt. Commun., 253, 156.
Lancis, J., G´
omez-Sarabia, C. M., Ojeda-Casta˜
neda, J., Fern´
andez-Pousa, C. R., & Andr´
es, P.
(2006). J. Eur. Opt. Soc. R. P., 1, 06018.
Lancis, J., M´
ınguez-Vega, G., Tajahuerce, E., Climent, V., Andr´
es, P., & Caraquitena, J. (2004).
J. Opt. Soc. Am. A, 21, 1875.
Lancis, J., Torres-Company, V., Silvestre, E., & Andr´
es, P. (2005). Opt. Lett., 30, 2973.
Langrock, C., Kumar, S., McGeehan, J. E., Willner, A. E., & Fejer, M. M. (2006). J. Lightwave
Technol., 24, 2579.
Leger, J. R., & Swanson, G. J. (1990). Opt. Lett., 15, 288.
Lin, I. S., McKinney, J. D., & Weiner, A. M. (2005). IEEE Microwave Wirel. Comp. Lett., 15, 226.
Liu, Y. Q., Park, S. G., & Weiner, A. M. (1996). IEEE J. Sel. Top. Quantum Electron., 2, 709.
Lohman, A. W., & Mendlovic, D. (1992). Appl. Opt., 31, 6212.
Longhi, S., Marano, M., Laporta, P., Svelto, O., Belmonte, M., Agogliati, L., Arcangeli, L.,
Pruneri, V., Zervas, M. N., & Ibsen, M. (2000). Opt. Lett., 25, 1481.
Maheswari, R. U., Takai, N., & Asakura, T. (1991). Opt. Commun., 81, 259.
Mandel, L., & Wolf, E. (1995). Optical coherence and quantum optics. Cambridge: Cambridge U.
Press.
Marinho, F. J., & Bernardo, L. M. (2006). J. Opt. Soc. Am. A, 23, 2331.
McKinney, J. D., Leaird, D. E., & Weiner, A. M. (2002). Opt. Lett., 27, 1345.
McKinney, J. D., & Weiner, A. M. (2006). IEEE Trans. Microw. Theory Tech., 54, 1681.
Mendlovic, D., Melamed, O., & Ozaktas, H. M. (1995). Appl. Opt., 34, 4113.
Miyamoto, D., Mandai, K., Kurokawa, T., Takeda, S., Shioda, T., & Tsuda, H. (2006). IEEE
Photon. Technol. Lett., 18, 721.
Mookherjea, S., & Yariv, A. (2001). Phys. Rev. E, 64, 016611.
Moon, S., & Kim, D. Y. (2006). Opt. Express, 24, 11575.
Mosley, P. J., Lundeen, J. S., Smith, J. S., Wasylczyk, P., U’Ren, A. B., Silberhorn, C., &
Walmsley, I. A. (2008). Phys. Rev. Lett., 100, 133601.
Mouradian, L. K., Louradour, F., Messager, V., Barthelemey, A., & Froehly, C. (2000). IEEE J.
Quantum Electron., 36, 795.
Murata, H., Morimoto, A., Kobayashi, T., & Yamamoto, S. (2000). IEEE J. Sel. Top. In Quantum
Electron., 6, 1325.
Muriel, M. A., Aza˜
na, J., & Carballar, A. (1999). Opt. Lett., 24, 1.
Nakazawa, M., Kubota, H., Sahara, A., & Tamura, K. (1998). IEEE J. Quantum Electron., 34,
1075.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 79 — #79
References 79
Nasr, M. B., Carrasco, S., Saleh, B. E. A., Sergienko, A. V., Teich, M. C., Torres, J. P., Torner, L.,
Hum, D. S., & Fejer, M. M. (2008). Phys. Rev. Lett., 100, 183601.
Ng, T. T., Parmigiani, F., Ibsen, M., Zhang, Z., Petropoulos, P., & Richardson, D. J. (2008).
IEEE Photon. Technol. Lett., 20, 1097.
Ninck, M., Galler, A., Feurer, T., & Brixner, T. (2007). Opt. Lett., 32, 3379.
Nuccio, S. R., Yilmaz, O. F., Wang., X., Wang, J., Wu, X., & Willner, A. E. (2010). Opt. Lett., 35,
1819.
Okawachi, Y., Foster, M. A., Chen, X. P., Turner-Foster, A. C., Salem, R., Lipson, M., Xu, C., &
Gaeta, A. L. (2008). Opt. Express, 16, 10349.
Okawachi, Y., Salem, R., Foster, M. A., Turner-Foster, A. C., Lipson, M., & Gaeta, A. L. (2009).
Opt. Express, 17, 5691.
Osgood, R. M. Jr., Panoiu, N. C., Dadap, J. I., Liu, X., Chen, X., Hsieh, I.- W., Dulkeith, E.,
Green, W. M. J., & Vlasov, Y. A. (2009). Adv. Opt. Photon., 1, 162.
Ou, Z. Y., Rhee, J. K., & Wang, L. J. (1999). Phys. Rev. A, 60, 593.
Paakonen, P., Turunen, J., Vahimaa, P., Friberg, A. T., & Wyrowski, F. (2002). Opt. Commun.,
204, 53.
Palma, C., & Bagini, V. (1997). J. Opt. Soc. Am. A, 14, 1774.
Papoulis, A. (1968a). Systems and transforms with applications in optics. New York:McGraw-Hill.
Papoulis, A. (1968b). J. Opt. Soc. Am., 58, 653.
Papoulis, A. (1994). J. Opt. Soc. Am. A, 11, 3.
Park, Y., Ahn, T. J., Kieffer, J. C., & Aza ˜
na, J. (2007). Opt. Express, 15, 4597.
Patorski, K. (1989). Prog. in Opt., 27, 3.
Patorski, K., & Parfjanowicz, G. (1981). Opt. Acta, 28, 357.
Pe’er, A., Dayan, B., Friesem, A. A., & Silberberg, Y. (2005). Phys. Rev. Lett., 94, 073601.
Pelusi, M. D., Luan, F., Vo, T. D., Lamont, M. R. E., Madden, S. J., Bulla, D. A., Choi, D. Y.,
Luther-Davies, B., & Eggleton, B. J. (2009). Nature Photon., 3, 139.
Pelusi, M. D., Matsui, Y., & Suzuki, A. (1999). IEEE Photon. Technol. Lett., 11, 1461.
Pelusi, M. D., Ta’eed, V. G., Fu, L. B., Magi, E., Lamont, M. R. E., Madden, S., Choi, D. Y., Bulla,
D. A. P., Luther-Davies, B., & Eggleton, B. J. (2008). IEEE J. Sel. Top. Quantum Electron., 14,
529.
Price, W. P., U’ Ren, A. B., & Walmsley, I. A. (2001). Phys. Rev. A, 64, 063815.
Pudo, D., Depa, M., & Chen, L. R. (2007). J. Lightwave Technol., 25, 2898.
Saleh, B. E. A. (1979). Opt. Commun., 30, 135.
Saleh, B. E. A., & Irshid, M. I. (1982). Opt. Lett., 7, 342.
Salem, R., Foster, M. A., Turner, A. C., Geraghty, D. F., Lipson, M., & Gaeta, A. L. (2008). Opt.
Lett., 33, 1047.
Salem. R., Foster, M. A., Turner-Foster, A. C., Geragthy, D. F., Lipson, M., & Gaeta, A. L.
(2009). Opt. Express, 17, 4324.
Saperstein, R. E., Alic, N., Panasenko, D., Rokitski, R., & Fainman, Y. (2005). J. Opt. Soc. Am.
B, 22, 2427.
Saperstein, R. E., Alic, N., Zamek, S., Ikeda, K., Slutsky, B., & Fainman, Y. (2007). Opt. Express,
15, 15464.
Saperstein, R. E., & Fainman, Y. (2008). Appl. Opt., 47, A21.
Saruwatari, M. (2000). IEEE J. Sel. Top. Quantum Electron., 6, 1363.
Sensarn, S., Yin, G. Y., & Harris, S. E. (2009). Phys. Rev. Lett., 103, 163601.
Shabtay, G., Mendlovic, D., & Zalevsky, Z. (2000). Appl. Opt., 39, 6556.
Shake, I., Takara, H., Kawanishi, S., & Saruwatari, M. (1998). Electron. Lett., 34, 792.
Sharping, J. E., Lee, K. F., Foster, M. A., Turner, A. C., Schmidt, B. S., Lipson, M., Gaeta, A. L.,
& Kumar, P. (2006). Opt. Express, 14, 12388.
Sharping, J. E., Okawachi, Y. Van Howe, J., Xu, C., Wang. Y., Willner, A. E., & Gaeta, A. L.
(2005). Opt. Express, 13, 7872.
To protect the rights of the author(s) and publisher we inform you that this PDF is an uncorrected proof for internal business
use only by the author(s), editor(s), reviewer(s), Elsevier and typesetter diacriTech. It is not allowed to publish this proof online
or in print. This proof copy is the copyright property of the publisher and is confidential until formal publication.
“Wolf 04-ch01-001-080-9780444538864” — 2011/11/18 — page 80 — #80
80 Space-Time Analogies in Optics
Shi, X., Valencia, A., Hendrych, M., & Torres, J. P. (2008). Opt. Lett., 33, 875.
Shih, Y. (2003). Rep. Prog. Phys., 66, 1009.
Shih, Y. H., Sergienko, A. V., Rubin, M. H., Kiess, T. E., & Alley, C. O. (1994). Phys. Rev. A, 50,
23.
Siegman, A. E. (1985). Lasers. Mill Valley: Univ. Science Books.
Solli, D. R., Chou, J., & Jalali, B. (2008). Nature Photon., 2, 48.
Supradeepa, V. R., Huang, C. B., Leaird, D. E., & Weiner, A. M. (2008). Opt. Express, 16, 11878.
Tan, D. T. H., Ikeda, K., Saperstein, R. E., & Fainman, Y. (2008). Opt. Lett., 24, 3013.
Tearney, G. J., Shishkov, M., & Bouma, B. E. (2002). Opt. Lett., 27, 412.
Thomas, S., Malacarne, A., Fresi, F., Poti, L., & Aza ˜
na, J. (2010). J. Lightwave Technol., 28, 1832.
Thomas, S., Malacarne, A., Fresi, F., Poti, L., Bogoni, A., & Aza ˜
na, J. (2009). Opt. Lett., 34, 545.
Tong, Y. C., Chan, L. Y., & Tsang, H. K. (1997). Electron. Lett., 33, 983.
Torres-Company, V., & Chen, L. R. (2009). Opt. Lett., 34, 3385.
Torres-Company, V., Fern´
andez-Alonso, M., Lancis, J., Barreiro, J. C., & Andr´
es, P. (2006).
Opt. Express, 14, 9617.
Torres-Company, V., Fern´
andez-Pousa, C. R., & Chen, L. R. (2009). Opt. Lett., 34, 1885.
Torres-Company, V., Lancis, J., & Andr´
es, P. (2006a). Opt. Express, 14, 3171.
Torres-Company, V., Lancis, J., & Andr´
es, P. (2006b). IEEE Photon. Technol. Lett., 18, 2626.
Torres-Company, V., Lancis, J., & Andr´
es, P. (2007a). Opt. Lett., 32, 2849.
Torres-Company, V., Lancis, J., & Andr´
es, P. (2007b). J. Opt. Soc. Am. A, 24, 888.
Torres-Company, V., Lancis, J., & Andr´
es, P. (2008). Opt. Lett., 33, 1822.
Torres-Company, V., Lancis, J., Andr´
es, P., & Chen, L. R. (2008). J. Lightwave Technol., 26, 2476.
Torres-Company, V., Lancis, J., Andr´
es, P., & Chen, L. R. (2008b). Opt. Express, 16, 21564.
Torres-Company, V., Lajunen, H., & Friberg, A. T. (2007). J. Opt. Soc, Am. B, 24, 1441.
Torres-Company, V., Lajunen, H., Lancis, J., & Friberg, A. T. (2008). Phys. Rev. A, 77, 043811.
Treacy, E. B. (1968). Phys. Lett. A, 28, 34.
Treacy, E. B. (1969). IEEE J. Quantum Electron., QE-5, 454–458.
Tsang, M., & Psaltis, D. (2006). Phys. Rev. A, 71, 043806.
Valencia, A., Cere, A., Shi, X. J., Molina-Terriza, G., & Torres, J. P. (2007). Phys. Rev. Lett., 99,
243601.
Valencia, A., Chekhova, M. V., Trifonov, A., & Shih, Y. H. (2002). Phys. Rev. Lett., 88, 183601.
Valencia, A., Scarcelli, G., & Shih, Y. (2004). Appl. Phys. Lett., 85, 2655.
van Howe, J., & Xu, C. (2005a). Opt. Lett. 30, 99.
van Howe, J., & Xu, C. (2005b). Opt. Express, 13, 1138.
van Howe, J., & Xu, C. (2006). J. Lightwave Technol., 24, 2649.
Venghaus, H. (Ed.). (2006). Wavelength filters in Fibre Optics. Berlin: Springer.
Wang, Y., Yu, C., Yan, L., Willner, A. E., Roussev, R., Langrock, C., Fejer, M. M., Sharping, J.
E., & Gaeta, A. L. (2007). IEEE Photon. Technol. Lett., 19, 861.
Weiner, A. M. (2000). Rev. Sci. Inst., 71, 1929.
Weiner, A. M. (2009). Ultrafast optics. New Jersey: Wiley Interscience.
Weiner, A. M., & Heritage, J. P. (1989). US Patent 4928316.
Wieser, W., Biedermann, B. R., Klein, T., Eigenwilling, C. M., & Huiber, R. (2010). Opt. Express,
18, 14685.
Winthrop, J. T., & Worthington, C. R. (1965). J. Opt. Soc. Am., 55, 373.
Xiao, S. J., & Weiner, A. M. (2004). Opt. Express, 12, 2895.
Xin, M., Yin, F. F., Lei, C., Chen, M. H., Chen, H. W., Tang, C., & Xie, S. Z. (2010). Opt. Lett.,
35, 1596.
Yamamoto, T., Komukai, T., Takada, K., & Suzuki, A. (2007). Electron. Lett., 43, 1040.
Yang, S. D., Weiner, A. M., Parameswaran, K. R., & Fejer, M. M. (2004). Opt. Lett., 17, 2070.
Zah, F., Halder, M., & Feurer, T. (2008). Opt. Express, 16, 16452.
Zalvidea, D., Duchowicz, R., & Sicre, E. E. (2004). Appl. Opt., 43, 3005.
... The exploration of optimal pulse shape using this expression will be discussed in another publication [32]. This also shows that the spectral phase that precedes the temporal phase in several experiments [11][12][13][33][34][35][36][37][38] has no impact except via a different ξ R (t). ...
... Note that, due to the significantly greater impact of quadratic phase modulation on pulse bandwidth compared to sinusoidal modulation, we have chosen k = 0.1 in Fig. 2a to ensure that all pulse shapes are clearly visible within the same plot. Conversely, in Fig. 2b, we have opted for k = 1, which is the parameter employed in the results for quadratically phased pulses in Fig. 4. The values selected for k = 1 and Ω = 1 in this paper are considerably smaller compared to the typical values used in experimental studies [11][12][13][33][34][35][36][37][38]. They suggest the potential direction of future experimental efforts to realize the benefits of chirping in single-photon spectroscopy. ...
... Analytical expressions for the classical and quantum contributions to the FI of a quadratically phase-modulated Gaussian pulse, as described by Eqns. (33) and (35), at asymptotically long times t ≫ 1/max(Γ, Γ ⊥ ) is presented in Appendix D. As Eqns. (D1) and (D2) show, the sole pulse characteristic they depend on is its bandwidth. ...
... We are unaware of any C 2 n model that includes dispersion. Even ignoring the likely ω dependence of C 2 n and applying the common Taylor series expansion approach for k (see Refs. [25,[58][59][60][61] and the citations therein), the problem is still not tractable because of the fractional power ω 1 and ω 2 terms in D [see Equation (9)]. To proceed further, we are left with little recourse other than neglecting dispersion and assuming, at most, a linear dependence of k on ω, such that all space-frequency terms in Equation (10) are no greater than second order. ...
... Before proceeding with the derivation, it is important to briefly discuss the physical consequences of ignoring dispersion. The behavior of a pulsed beam in a dispersive medium is very similar to diffraction in free space [25,61]. For a beam propagating in free space, it is customary to report the propagation distance z relative to the Rayleigh range z Rx = k(ω c )σ 2 ...
Article
Full-text available
We investigate the behaviors of spatiotemporal optical vortex (STOV) beams propagating through atmospheric turbulence. Starting with the extended Huygens-Fresnel integral, we derive an expression for the mutual coherence function (MCF) of a STOV beam, of any vortex order, propagating through turbulence of arbitrary strength. We specialize our general MCF and compare it to existing results from the literature. In addition, we analyze, in detail, first-and second-order STOV beams, focusing on how turbulence degrades both the beams' shapes and vortices. Lastly, we simulate the propagation of a second-order STOV beam through atmospheric turbulence and compare the results to predictions from our theoretical MCF. We find the results to be in good agreement thereby validating our analysis.
... As described in Ref. [47], a twisted STOV beam can be synthesized using a Fourier transform pulse shaper (FTPS) [8,18,51,[57][58][59][60][61], which is depicted in Fig. 1. Light from a coherent, femtosecond pulsed laser is spread by a grating (G), and in combination with the cylindrical lens (CL) of focal length ps , has its spectrum mapped to physical space at the plane of the spatial light modulator (SLM). ...
Article
Full-text available
We derive a closed-form expression for the mutual coherence function (MCF) of a twisted spatiotemporal optical vortex (STOV) beam after propagating a distance z in a linear dispersive medium. A twisted STOV beam is a partially coherent optical field that possesses a coherent STOV and a stochastic twist coupling its space and time dimensions. These beams belong to a special class of space-time-coupled light fields that carry transverse (to the direction of propagation) orbital angular momentum, making them potentially useful in numerous applications including quantum optics, optical manipulation, and optical communications. After presenting the derivation, we validate our new general MCF by showing that it simplifies to the MCFs for two example beams from the literature. Lastly, we present and analyze space-time beam profiles and complex arguments (phases) of the MCF at multiple z and differing amounts of material dispersion, field correlation or coherence, and beam twist to gain insight into how twisted STOV beams evolve in dispersive media. We conclude with a brief summary.
... The sequence and the CWF at different stages have been schematically depicted in Fig. 1 for an initial state of two coherent, temporally separated Gaussian pulses. Notably, these operations are completely analogous to a series of a propagation, passing through a lens, and a propagation of a single transverse spatial dimension in the paraxial approximation [43][44][45]. The Chronocyclic Wigner function (CWF) for the initial state of two coherent temporally separated pulses is depicted at the stages of the transformation. ...
Article
Full-text available
The Fractional Fourier Transform (FRT) corresponds to an arbitrary-angle rotation in the phase space, e.g., the time-frequency (TF) space, and generalizes the fundamentally important Fourier Transform. FRT applications range from classical signal processing (e.g., time-correlated noise optimal filtering) to emerging quantum technologies (e.g., super-resolution TF sensing) which rely on or benefit from coherent low-noise TF operations. Here a versatile low-noise single-photon-compatible implementation of the FRT is presented. Optical TF FRT can be synthesized as a series of a spectral disperser, a time-lens, and another spectral disperser. Relying on the state-of-the-art electro-optic modulators (EOM) for the time-lens, our method avoids added noise inherent to the alternatives based on non-linear optical interactions (such as wave-mixing, cross-phase modulation, or parametric processes). Precise control of the EOM-driving radio-frequency signal enables fast all-electronic control of the FRT angle. In the experiment, we demonstrate FRT angles of up to 1.63 rad for pairs of coherent temporally separated 11.5 ps-wide pulses in the near-infrared (800 nm). We observe a good agreement between the simulated and measured output spectra in the bright-light and single-photon-level regimes, and for a range of pulse separations (20 ps to 26.7 ps). Furthermore, a tradeoff is established between the maximal FRT angle and optical bandwidth, with the current setup accommodating up to 248 GHz of bandwidth. With the ongoing progress in EOM on-chip integration, we envisage excellent scalability and vast applications in all-optical TF processing both in the classical and quantum regimes.
... Space-time duality has been an everlasting topic in physics, which inspired ingenious contributions. The analogy of the space and time in classic optics has led to significant discoveries in optical research [1,2]. For examples, the well-known Kramers-Kronig relation, which arises from the causal temporal response of the material [3], has its spatial counterpart [4]; the spatial domain reflection and refraction can happen in the time domain as well [5]. ...
Article
Full-text available
A spatiotemporal diffractive deep neural network (STD2NN) is proposed for spatiotemporal signal processing. The STD2NN is formed by gratings, which convert the signal from the frequency domain to the spatial domain, and multiple layers consisting of spatial lenses and space light modulators (SLMs), which conduct spatiotemporal phase modulation. An all-optical backpropagation (BP) algorithm for SLM phase tuning is proposed, with the gradient of the loss function computed by the inner product of the forward propagating optical field and the backward propagating conjugated error field. As a proof of concept, a spatiotemporal word “OPTICA” is generated by the STD2NN. Afterwards, a spatiotemporal optical vortex (STOV) beam multiplexer based on the STD2NN is demonstrated, which converts the spatially separated Gaussian beams into the STOV wave-packets with different topological charges. Both cases illustrate the capability of the proposed STD2NN to generate and process the spatiotemporal signals.
... converting picosecond-scale pulses in the telecommunication band, optimal for high-rate fiber transmission, to nanosecondscale pulses in the visible range processed by quantum memories. An efficient tool for such transformations is provided by the technique of temporal imaging [3,4] allowing one to stretch and compress optical waveforms in time. This technique requires a device imprinting a quadraticin-time phase modulation on the input waveform, known as the time lens [5,6]. ...
Article
Full-text available
A time lens allows one to stretch or compress optical waveforms in time, similar to the conventional lens in space. However, a single-time-lens imaging system always imparts a residual temporal chirp on the image, which may be detrimental for quantum networks, where the temporal image interacts with other fields. We show that a two-time-lens imaging system satisfying the telescopic condition, a time telescope, is necessary and sufficient for creating a chirpless image. We develop a general theory of a time telescope, find the conditions for loss minimization, and show how an erecting time telescope creating a real image of a temporal object can be constructed. We consider several applications of such a telescope to making indistinguishable the photons generated by spontaneous parametric downconversion or single emitters such as quantum dots.
Article
Advances in Fourier optics and information processing technology have facilitated a variety of new research areas in telecommunications, biology, quantum information science, spectroscopy, and atomic and molecular physics. The concept of space-time duality plays an important role as a system design tool for ultrafast optical processing and characterization. In particular, time-lens-based temporal imaging systems enable additional features such as waveform expansion and compression that are particularly useful for ultrafast optical signal processing. The time-equivalent optical setup of a spatial 2-f system, which acts as a Fourier processor (time-to-frequency and frequency-to-time conversion),can be used as a basic component of more complex temporal imaging systems. This special issue focuses on timestretch spectroscopy, which uses frequency chirp ultrafast laser pulses to capture ultrafast phenomena in real time.
Chapter
Due to the ultrahigh bandwidth capacity of photonics, high-speed optical imaging, and sensing technologies can capture and manipulate extremely fast and complex signals. However, the ultimate performance of these sensing systems is entirely constrained by the information throughput of digitization and processing electronics. This digitization bottleneck can be alleviated using the principles of compressed sensing to implement data compression in the optical domain prior to digitization. Continuous High-Rate Photonically Enabled Compressed Sensing (CHiRP-CS) is a flexible photonic architecture to implement such high-speed compressed sensing in continuously recording scenarios. To do so CHiRP-CS leverages high-speed optical signal processing in robust fiber optic hardware based on the principles of the space-time duality of electromagnetic waves. This chapter first provides a brief introduction to the background for CHiRP-CS including a description of compressed sensing and the space-time duality of electromagnetic waves. Subsequently, the working principles of CHiRP-CS are described and several example applications are highlighted.
Article
Full-text available
In the case of a periodic object, the wave amplitude in every point of a Fresnel diffraction pattern may be obtained from the values of the wave amplitude at a finite number of points in the object plane. This results is applied to the case of pure phase objects. A simple condition on the object structure is derived, which would allow the object phase distribution to be determined directly from a measured intensity profile for a certain defocusing distance.
Article
Half Title Wiley Series Page Title Copyright Dedication Contents Preface
Article
We extend the well-known analogy between the problems of paraxial diffraction in space and dispersion in time to optical pulse compression and propose a time-domain analog to spatial imaging that allows for the distortionless expansion or compression of optical power waveforms. We call this new concept temporal imaging and derive equivalent expressions for the focal length and the f-number of a time lens and the magnification of an imaging system. It should now become possible, with a temporal microscope, to expand ultrafast optical phenomena to a time scale that is accessible to conventional high-speed photodiodes.
Article
The behaviour of the temporal Talbot effect under carrier partial coherence is analysed. This effect finds application as a passive method for multiplying the repetition rate of the intensity of pulse trains in first-order dispersive media. The study is based on the representation of the average intensity of partially coherent modulated light after first order dispersion as a coherence-dependent low pass filter acting on the intensity distribution obtained with a monochromatic carrier. The resulting coherence limitations are more restrictive in Talbot devices with high indices, due to the necessity of highly coherent carriers to produce a faithful reconstruction of trains by multiple pulse-to-pulse interference. Coherence limitations for architectures composed of externally modulated cw sources are analysed. Generalizations of the theory to several spectral lines, and its relation with the Collett–Wolf theorem are also reported.
Article
We review the field of femtosecond pulse shaping, in which Fourier synthesis methods are used to generate nearly arbitrarily shaped ultrafast optical wave forms according to user specification. An emphasis is placed on programmable pulse shaping methods based on the use of spatial light modulators. After outlining the fundamental principles of pulse shaping, we then present a detailed discussion of pulse shaping using several different types of spatial light modulators. Finally, new research directions in pulse shaping, and applications of pulse shaping to optical communications, biomedical optical imaging, high power laser amplifiers, quantum control, and laser-electron beam interactions are reviewed. (C) 2000 American Institute of Physics. [S0034-6748(00)02005-0].
Article
The axial scan rate of optical frequency-domain reflectometry and optical coherence tomography can be increased to megahertz frequencies by dispersive Fourier transformation. However, the fundamental connection between dispersion and loss creates a trade-off between detection sensitivity and acquisition speed. Here we circumvent this predicament by using distributed Raman postamplification of the reflection from the sample. The Raman amplification enables measurement of weak signals, which are otherwise buried in detector noise. It extends the depth range without sacrificing the acquisition speed. Single-shot imaging with improved sensitivity at an axial scan rate of 36.6 MHz is demonstrated.
Article
The ultrashort pulses generated by a Nd:glass laser have frequency swept carriers and can therefore be compressed by a pair of gratings to a length much closer to the reciprocal of the bandwidth. The author acknowledges stimulating discussions with M.J. Brienza and A.J. DeMaria, and the technical assistance of R. Bodurtha.
Article
To every optical system S consisting of thin lenses illuminated by narrow beams, there corresponds a dual system Su with the following property: If the complex amplitudes of the fields on the input and output planes of S and Su are a(x,y), au(x,y) and b(x,y), bu(x,y), respectively, and au(x,y) equals the Fourier transform of a(x,y), then bu(x,y) equals the Fourier transform of b(x,y). This dualism can be used to simplify the analysis and synthesis of certain idealized systems.
Article
Using the space–time duality theory, we introduce and analyze theoretically the temporal equivalent of the spatial self-imaging phenomena under point-source illumination (spherical wave-front illumination). This temporal effect can be produced by passing a given periodic optical pulse sequence through a time lens followed by a dispersive medium. The time lens (quadratic phase modulator) implements the time-domain equivalent of spherical wave-front illumination. Based on this temporal effect, we demonstrate that a system composed of a time lens followed by a dispersive medium can be configured to operate over a periodic optical pulse sequence of finite duration (i) as a conventional temporal imaging system, providing a distortionless temporal compression or expansion of the original pulse sequence, or (ii) as an advanced temporal imaging system, combining the capabilities of a conventional imaging system with those of the fractional temporal self-imaging effect, i.e., multiplication of the original pulse-repetition rate.