1. Introduction
In the context of optics, space–time duality was noticed first by Tournois in 1964 [
1] and a few years later by Akhmanov et al. [
2]. It is based on the mathematical equivalence, under some conditions, of two equations governing the diffraction of beams in space and the dispersion of pulses in time. However, it was only after 1988 that space–time duality was used to develop the concept of a time lens and such lenses were used for temporal imaging [
3,
4,
5]. As the name suggests, a time lens plays the role of a conventional lens in the time domain and can be used for making the analog of a microscope for imaging a time-varying signal. For this reason, devices acting as time lenses have attracted considerable attention for optical signal processing [
6,
7,
8,
9].
In recent years, the concept of space–time duality has been extended in several directions. An example in nonlinear optics is provided by the spatial and temporal solitons. The underlying equation describing them is the nonlinear Schrödinger (NLS) equation [
10]. Spatial and temporal solitons form when this equation includes the diffractive or dispersive effects, respectively, [
11]. In the context of the Kerr frequency combs, adoption of the Lugiato–Lefever equation that was developed originally for spatial cavity solitons led to the observation of temporal cavity solitons [
12].
The temporal analog of the phenomenon of reflection, which occurs at any spatial interface separating two media of different refractive indices, constitutes another example of space–time duality [
13,
14,
15]. In this analog, a temporal interface separates two intervals of different refractive indices in the same medium. It turns out that a moving temporal interface, which changes the refractive index of a dispersive medium in a region moving at a constant speed [
16], is easier to realize in practice. When an optical pulse interacts with this moving interface inside a dispersive medium, it splits into two parts, whose frequencies are shifted such that they travel at different speeds [
17,
18,
19,
20]. These two parts correspond to reflected and transmitted pulses and are temporal analogs of the reflection and refraction at a spatial interface [
15]. It is also possible to realize the temporal analog of total internal reflection and to use it for time-domain waveguiding [
17].
Wave propagation in a medium whose refractive index varies spatially has a long history. Such an inhomogeneous medium is useful in different contexts such as graded-index fibers [
21], Bragg gratings [
22], and photonic crystals [
23]. Recent attention to time-varying media [
24,
25,
26], where the refractive index varies with time, constitutes another example of space–time duality. It has led to novel concepts such as photonic time crystals [
27,
28,
29] and spatiotemporal Bragg gratings [
30].
In this review, I discuss how the concept of space–time duality and the use of nonlinear optics has led to multiple advances in recent years.
Section 2 reviews the historical origin of space–time duality by focusing on the phenomena of spatial diffraction and temporal dispersion. Time lenses and their applications are reviewed in
Section 3. The focus of
Section 4 is on the temporal analog of reflection and refraction for optical pulses at a moving temporal boundary inside a dispersive medium. The use of nonlinear effects for creating such boundaries is discussed in
Section 5. Optical solitons, forming inside an optical fiber through the Kerr effect, are used to discuss the temporal analogs of spatial reflection, waveguides, and Fabry–Perot resonators.
Section 6 is devoted to the case of periodic temporal modulations of a medium’s refractive index, leading to the formation of spatiotemporal Bragg gratings and photonic time crystals.
2. Origin of Space–Time Duality
Diffraction and dispersion are fundamental concepts in optics [
31,
32]. Any optical beam spreads in space because of diffraction, and an optical pulse spreads in time because of dispersion [
33]. The initial concept of space–time duality was based on the mathematical equivalence of wave-propagation equations governing these two phenomena under specific conditions [
1,
2]; it has been extended further in recent years.
Consider propagation of electromagnetic waves in a linear dispersive medium. As it is simpler to solve Maxwell’s equations in the frequency domain, it is common to employ the Fourier transform of the electric field in the form
and solve the resulting Helmholtz equation,
where
,
is the refractive index of the medium at the frequency
and
c is the speed of light in vacuum. This equation can be used for continuous-wave (CW) beams or pulsed beams by choosing a suitable range of frequencies.
Let us first consider a CW beam with a narrow spectrum (nearly monochromatic) centered at a specific frequency
. In this case,
becomes constant. Choosing the
z axis along the beam’s direction, one can introduce the slowly varying amplitude
of the electric field as
where
is a unit vector representing the beam’s state of polarization. If we use Equation (
3) in Equation (
2) and neglect the second derivative
in the paraxial approximation,
is found to satisfy
This equation governs the diffraction of CW beams in a homogeneous medium with the refractive index
at its frequency
.
Spatial spreading of a CW beam depends on its initial spot size—narrower beams spread much more rapidly than wider beams [
31]. If the spot size of a beam is elliptical such that its width is much wider in the
y direction, the beam will mostly diffract in the
x direction, and its size in the
y direction will not change. The same thing happens when the beam is confined to a planar waveguide that guides it in the
y direction. In both cases, the
y derivative can be ignored in Equation (
4) to obtain the following simpler equation:
The situation is more complicated for a beam in the form of a short pulse with a relatively wide spectrum. In the most general case, diffraction and dispersion occur simultaneously, resulting in the so-called space–time coupling, and one must solve a four-dimensional problem. The problem is simplified considerably when such pulses are launched into a single-mode waveguide such as an optical fiber. In this case, the beam does not spread spatially because of its confinement, and we can focus on the dispersion-induced spreading of pulses in time.
To include the dispersive effects, we seek solutions of Equation (
2) in the form
where
is the spatial form of the single mode supported by the waveguide. Using this form in Equation (
2) and the resulting modal solution
,
is found to satisfy
where
and
is the effective index of the single mode.
Dispersive effects are included by expanding
in a Taylor series around
and retaining terms up to second order:
where
and
for
. The parameter
is related inversely to the group velocity, while
governs its dispersion and is known as the group-velocity dispersion (GVD) parameter [
33].
Similar to the CW case, one can introduce the slowly varying amplitude
as
. Neglecting its second derivative with respect to
z,
satisfies
Using
with the expansion in Equation (
8), we obtain the simple differential equation
This equation can be converted back to the time domain by using the inverse Fourier transform
Noting that
is the Fourier transform of
, Equation (
10) in the time-domain takes the form
The
term in this equation can be eliminated using a reference frame moving with the pulse. Introducing
as a new time variable, we obtain
Equation (
13) governs the dispersion-induced spreading of pulses in the time domain. It should be compared with Equation (
5), which governs the dispersion-induced spreading of CW beams in space. The concept of space–time duality stems from the mathematical equivalence of these two equations. The parameter
appearing in Equation (
5) is replaced with the parameter
in Equation (
13). While
is always positive,
can be positive or negative, depending on the wavelength
. The important point is that one should expect similar qualitative behavior to occur in the spatial and temporal cases. As an example, it is well known that a Gaussian beam maintains its Gaussian shape as it diffracts in space, even though its width changes because of diffraction [
32]. The same statement can be made for Gaussian pulses dispersing and broadening in time inside a dispersive medium [
33].
As another example of the usefulness of the concept of space–time duality,
Figure 1a shows the diffraction pattern of a narrow slit. The solution of Equation (
5) shows that
becomes the Fourier transform of the input field
(up to a phase factor) at a distance far from the plane
, a feature known as the far-field diffraction [
31]. As seen in
Figure 1b, this feature implies that the shape of an optical pulse at the end of a long dispersive fiber would mimic the spectrum of that pulse [
8]. As a result, time-domain measurements of a pulse’s shape at the fiber’s output can provide all frequency-domain information of the optical pulse launched into it. This technique, known as dispersive Fourier transform (DFT), provides single-shot spectra of optical pulses and has proven extremely useful in several areas that include biomedical imaging for observing in real time spectral changes occurring inside a medium [
34,
35,
36].
3. Time Lens and Its Applications
As mentioned in the introduction, space–time duality was used as early as 1988 to develop the concept of a time lens and to use such lenses for temporal imaging [
3,
4,
5]. A time lens may not resemble a conventional lens, but it performs the same function as a lens in the time domain. As seen in
Figure 2a, a lens has curved surfaces such that its thickness changes along the transverse dimensions. When a plane wave passes through this lens, its phase front becomes curved because of a spatially varying phase shift
imposed by the lens that leads to its focusing. For a convex lens, this phase shift can be written as [
32]
where
is central thickness of the lens of focal length
f. Clearly, a time lens should add a time-dependent phase shift to the incoming pulse that varies as
. The important question is, what is the analog of the focal length for a time lens? To answer this and other related questions, let us solve Equation (
13).
As Equation (
13) is linear, we can solve it with the Fourier-transform method. Using the Fourier transform relation in Equation (
11),
satisfies an ordinary differential equation
where we replaced
with
for notational simplicity. Its solution is given by
Thus, GVD modifies the phase of each spectral component of the pulse by an amount that scales with frequency as
. Even though such phase changes do not affect the optical spectrum, they can modify the pulse shape. It is useful to think of any dispersive medium of length
L as a spectral filter and write Equation (
16) in the form
where
governs the action of this spectral filter.
By taking the inverse Fourier transform indicated in Equation (
11), the solution of Equation (
13) is found to be
where
is obtained from
Evoking the convolution theorem, the preceding solution can also be written in the time domain as
where the impulse response of the filter
is obtained from
Using the form of
in Equation (
17), the impulse response
is found to be
where the frequency integration was carried out using
The impulse response in Equation (
22) depends on a single parameter
, representing the net group-delay dispersion (GDD) over the medium’s length
L. This suggests that all properties of a dispersive medium should depend on
D, including the temporal analog of the focal length of a lens. Thus, the phase shift imposed by a time lens, the temporal analog of Equation (
14), should have the form [
9]
where
is a constant phase and
is the focal GDD of the time lens. A time-varying phase can be related to the frequency chirp imposed on the pulse. Using
, a quadratic time dependence implies a linear chirp
.
The device used in 1988 as a time lens was a phase modulator [
3]. Such devices impose a sinusoidal phase shift on an optical wave, when a microwave signal is used to modulate the refractive index of a suitable electro-optic material, such as a lithium niobate (LiNbO
3) crystal (or waveguide). In this situation, the input and output fields are related as
where
is the maximum phase shift,
is the modulation frequency, and
is a phase that depends on how the modulator is biased. Choosing this phase as
, and expanding the cosine function in a Taylor series around
, we obtain
Thus, a short pulse whose intensity peaks at
undergoes a phase shift in the form given in Equation (
14) such that
. The phase shift is quadratic only if the pulse width
is considerably shorter than the modulation period
. For a modulator operating at 10 GHz,
should not exceed 10 ps. Several nonlinear techniques have been used in recent years for making time lenses with improved properties [
9].
Consider the temporal analog of focusing shown in
Figure 2b. In the temporal case, a relatively wide pulse is chirped first by a time lens and then transmitted through a dispersive medium whose length is chosen to satisfy the relation
. In this case, the output field is obtained from Equation (
20) using
, where
governs the pulse shape and the quadratic phase is imposed by the time lens. The integration can be conducted for Gaussian-shape pulses. Using
, the input can be written in the form of a chirped Gaussian pulse as
where the chirp parameter
is negative and
is related to the full-width at half-maximum of the pulse as
. By taking the Fourier transform of
, we obtain
The propagation of a chirped Gaussian pulse through a dispersive medium has been studied in the context of fiber-optic communication systems [
33]. Substituting Equation (
28) into Equation (
18) and using Equation (
23), frequency integration can be performed to obtain:
where
. This equation shows that a Gaussian pulse remains Gaussian on propagation, but its width, chirp, and amplitude change as dictated by the factor
Q. Specifically, the chirp changes from its initial value
C to become
. The width of the output pulse becomes minimum when the pulse becomes unchirped. Using
with
and
, this condition is reduced to
assuming that
. At the focal distance, the pulse is compressed by a factor of
. It is easy to show that the minimum width of the pulse is given by
. Pulses can be compressed by a large factor by a time lens for small values of the ratio
.
Spatial lenses are used routinely for imaging in devices such as cameras and microscopes. In the simplest situation, a single lens of focal length
f images an object placed at a distance
on its one side. The image may be magnified (or reduced) and is produced at a distance
that satisfies the following imaging condition:
Time lenses were used for time-domain imaging starting in 1988 and follow an analogous scheme. The object takes the form of a time-varying signal. This signal is passed through a dispersive medium with
, before it is chirped by a time lens with the focal GDD
. Its imaging occurs when the chirped signal passes though a second dispersive medium with
. As one would expect from space–time duality, the imaging condition satisfies a relation that is identical to that in Equation (
31):
Other imaging properties such as magnification factor follow the same duality scheme. One application of temporal imaging is to stretch ultrashort pulses so much that their shape can be measured using a photodetector. It is important to mention that higher-order dispersion terms, neglected in Equation (
8), can produce “imaging abberations” for ultrashort pulses. As the applications of time lenses related to imaging and optical signal processing have been discussed in several past reviews [
7,
8,
9], I focus on more recent advances in what follows.
5. Solitons as Moving Boundaries
A moving index boundary can be realized using the optical Kerr effect. In this case, intense pump pulses are launched into a nonlinear dispersive medium, such as an optical fiber [
10]. Owing to the Kerr effect, the fiber’s refractive index increases with intensity by a small amount
, where
is the Kerr coefficient. Thus, pump pulses increase the refractive index in a time window set by their width, and this window moves at the speed of pump pulses. When a probe pulse, moving at a different speed because of its different wavelength, interacts with this index window, a reflected pulse is generated at a wavelength shifted from that of the probe [
18].
In general, pump pulses launched into an optical fiber become distorted because of the dispersive and nonlinear effects, resulting in a high-index region that does not maintain itself over the fiber’s length. This problem can be solved by making use of optical solitons, forming when pump pulses are launched at a wavelength longer than the zero-dispersion wavelength of the fiber so that the GVD is anomalous at the pump’s wavelength [
10]. Formation of solitons requires that the width and peak power (
and
) of pump pulses are chosen such that
, where
N is the soliton’s order and
is the GVD a the pump’s wavelength. The nonlinear parameter
is related to the Kerr coefficient
as
, where
is the effective area of the single mode supported by the fiber.
The shape of the soliton inside the fiber does not change under such conditions, and its power varies with time as
. The fiber’s refractive index
n increases by a small amount (typically
) over the soliton’s duration and is the largest at the peak of the soliton. This increase in
n creates a spatiotemporal boundary, moving at the speed of pump pulses. A probe pulse sees this increase through
, where
is the probe’s nonlinear parameter [
10]. The factor of two results from the nonlinear phenomenon of cross-phase modulation (XPM).
5.1. Soliton-Induced Temporal Reflection
The probe’s evolution is governed by Equation (
34) after
on its right side is replaced by the XPM term as follows:
This equation can be written in a normalized form of Equation (
35) as
where
and
. The parameter
compares the soliton’s width to that of the probe. The parameter
governs the XPM-induced interaction between the two pulses. This equation can be solved numerically to study the interaction of a probe pulse with a pump soliton inside a dispersive fiber. In the moving fame, the soliton’s peak remains fixed at
.
When a soliton is used to form a moving boundary, the refractive index increases only over its duration. The situation differs considerably from that of a sharp boundary discussed in
Section 4.2 because a soliton provides two interfaces at its leading and trailing edges with finite rise and fall times. In spite of this, temporal reflection exhibits similar features in the two cases. As an example,
Figure 6 shows the temporal reflection and refraction of a Gaussian pulse from a soliton 10 times narrower than the pulse (
) using
and
. Time-domain TIR occurs for values of
larger than 400. The impact of a boundary’s sharpness on temporal reflection was studied in 2021 using a transfer-matrix approach with a staircase model [
20]. The results show that TIR persists even for shallow boundaries with long rise times.
It turns out that the reflectivity of a soliton can be found in an analytic form with a simple trick. We can remove the second term in Equation (
41) by shifting
from the probe’s central frequency to the one for which
. With this shift in the reference frequency, Equation (
42) takes the form of the Schrödinger equation:
where
is the wave function and
plays the role of a potential barrier. Notice that
plays the role of time and
plays the role of a spatial coordinate. Even though Equation (
43) does not contain
ℏ, as expected for a classical problem, it is useful because one can use relevant quantum results with only minor changes. It shows that the time-reflection problem is analogous to the scattering of a quantum particle from a potential barrier. This analogy is another type of space–time duality where the time evolution of a quantum particle is mapped to the spatial evolution of optical waves.
The observation of temporal reflection from a soliton requires a short pump pulse and a probe pulse traveling at nearly the same speed at a different wavelength. In a 2012 experiment [
40], performed using a short microstructured fiber (only 1.1 m long), the zero-dispersion wavelength of the fiber was near 710 nm. This feature allowed the use of 105 fs pulses, emitted by a Ti–sapphire laser operating at 810 nm. Each pulse was launched such that it formed an optical soliton in the anomalous-GVD region of the fiber. Probe pulses were launched in the normal-GVD region of the fiber at wavelengths near 620 nm and traveled at nearly the speed of pump pulses. When the wavelength of probe pulses was varied from 595 to 645 nm, either a blue shift or a red shift was observed for the “reflected” pulse at the output end of the fiber, depending on whether the probe was traveling slower or faster than the soliton. The observed frequency shifts matched predictions based on Equation (
38).
5.2. Time-Domain Waveguides
Optical waveguides are devices that confine light spatially to their core region through TIR at the core-cladding interfaces [
32,
41]. Similarly to the spatial case, time-domain TIR can be used to provide the temporal analog of optical waveguides. In the temporal case, a pulse would be confined within a moving time window, where the refractive index differs from the regions outside of that window [
17]. When a probe pulse is located in the middle of two fundamental solitons acting as mirrors, it travels first toward one of these solitons and is totally reflected from it. The spectrum of the reflected pulse is shifted such that it slows down and moves away from this soliton. When the pulse arrive at the second soliton, it is reflected again through TIR, and its center frequency shifts back to the original value. This process repeats itself, trapping the pulse between the two solitons.
Figure 7 reveals how such a waveguide functions by solving Equation (
42) numerically, after it was modified to include the impact of both solitons:
It shows the temporal (left) and spectral (right) evolution along the length of a dispersive medium when a Gaussian-shaped probe pulse of width
is located initially in the middle of two solitons, separated by
. In normalized units, Equation (
44) was solved with the initial amplitude
using parameter values
,
, and
. Two solitons, with their peaks located at
(vertical dashed lines in
Figure 7), were 10 times shorter than
. As expected, the spectrum of pulse confined within the waveguide’s core shifts back and forth after TIR occurring at the high-index boundaries provide by two solitons. As a practical example, when
ps for probe pulses at a wavelengths near 1.1 μm, the required fiber’s length is around 1 km. The wavelength of 0.5 ps pump pulses should be near 1.5 μm to ensure anomalous GVD needed for the two solitons.
It is important to emphasize that the solitons are not essential for making temporal waveguides. A waveguide is formed whenever two boundaries form a time window where the refractive index differs from that outside the window. One should analyze whether such waveguides support temporal modes that are analogous to the spatial modes of planar waveguides. Modifying the right side of Equation (
34) to include the two boundaries located at
, we obtain
To find the temporal modes, we look for solutions of Equation (
45) within the time window
in the form
where
is the temporal shape of the mode,
K is the eigenvalue for this mode, and
is a frequency shift occurring because of the
term in Equation (
45).
Substituting Equation (
46) into Equation (
45) and equating the real and imaginary parts, we obtain
From Equation (
47), the frequency shift for all modes is found to be
. Using this value of
in Equation (
48), the modes are found by solving
This eigenvalue equation provides temporal shapes of modes for specific eigenvalues
K.
At this point, one can follow the same procedure used for spatial waveguides to find the temporal profiles of different modes [
17]. Similarly to the spatial case, one can introduce a dimensionless parameter as
This parameter determines the number of modes supported by a temporal waveguide of width
. In analogy with the spatial case, the waveguide supports
m modes when
. In particular, a temporal waveguide supports only a single mode (
) if it is designed such that
.
A clear evidence of the formation of a temporal waveguide by two solitons was seen in a 2015 experiment through a pump–probe type experiment [
42]. A 29 m long photonic crystal fiber was employed with its zero-dispersion wavelength near 980 nm. Pump pulses were 250 fs wide, and their wavelength was tunable from 1000 to 1500 nm. Probe pulses were considerably wider, and their 802 nm wavelength was in the normal-GVD region of the fiber. Each pump pulse was split into two pulses, separated by 3.6 ps, using the setup similar to a Michelson interferometer. Probe pulses were synchronized such that each was located in the center of a pair of pump pulses at the input end of the fiber.
5.3. Temporal Fabry–Perot Resonators
A conventional Fabry–Perot resonator (FPR) consists of two partially reflecting mirrors enclosing a medium of constant refractive index
n. The transmission spectrum of such a device is in the form of a frequency comb and exhibits a periodic resonances occurring at frequencies
, where
m is an integer and
is the resonator’s length. Such devices are routinely used for spectral analysis in many applications. As we have seen, solitons inside an optical fiber act effectively as mirrors in the time domain. One can construct the temporal analog of an FPR by using two such solitons separated in time by constant interval. Such a device was analyzed in a 2021 study using a matrix approach [
43].
When a probe pulse is incident at such an FPR, it can be transmitted or reflected, depending on whether its spectrum falls within or outside of a resonance peak.
Figure 8 shows this behavior for an FPR made with two 80 fs solitons, separated by 1.4 ps. The temporal evolution of a relatively wide Gaussian pulse (width 20 ps) was simulated numerically in two cases. The probe’s spectrum fits within a transmission peak in case (a) but falls in between two peaks in case (b). Part (d) shows the location of the pulse’s spectrum in the two cases within the transmission spectrum of the FPR. Time-varying index changes induced by two solitons are shown in part (c). As expected, the probe pulse is transmitted when its spectrum falls within a transmission peak. In contrast, it is mostly reflected when its spectrum falls in the middle of two transmission peaks. Experimental observation of a such a soliton-based time-domain FPR was lacking at the time of writing but would be of considerable interest.
7. Concluding Remarks
This review has focused on the concept of space–time duality in optics and its applications. Although this concept was originally based on the analogy between the diffraction of beams in space and the dispersion of pulses in time, it has been extended considerably in recent years. The first part of the review used time lenses and their applications as a simple example of the temporal analog of conventional lenses.
Novel phenomena emerge when optical pulses propagate through a nonlinear dispersive medium whose refractive index is modulated, both in space and time, in a traveling-wave fashion. Using optical fibers as an example of such a medium, we discussed the temporal analog of reflection at a moving index interface. A probe pulse incident at such an interface splits into two parts with different optical spectra such that the reflected part never crosses the interface. When the index change is large enough, the temporal analog of total internal reflection occurs, which allows one to construct time-domain waveguides that confine pulses to within a time window of fixed duration.
The use of nonlinear optics for creating moving index boundaries has allowed novel temporal analogs to be observed experimentally. The use of solitons through the Kerr effect indicates that such effects can be observed in silica fibers by employing a pump–probe configuration. A single solitons acts as a time-domain mirror that can be used to produce large spectral shifts through temporal reflection. Two closely spaced solitons can be used to make a temporal waveguide that confines probe pulses through multiple total internal reflections at both solitons. Two such solitons can also produce the temporal analog of a Fabry–Perot resonator by acting as partially reflecting mirrors.
An even more exotic temporal structure can be created by launching a periodic train of pump pulses that travel as fundamental solitons inside an optical fiber. Such a train of solitons induces periodic temporal modulations of the fiber’s refractive index, leading to the formation of spatiotemporal Bragg gratings. Such a device exhibits the so-called momentum gaps that are temporal analogs of the frequency band gaps forming in spatially periodic gratings. A probe pulse is totally reflected from this type grating when the momentum of its photons lies inside of a momentum gap. A photonic analog of Anderson localization can also occur when some disorder is introduced into such periodic photonic time crystals.
It is important to stress that it was not possible to include all examples of space–time duality in this review because of space limitations. For example, spatial and temporal solitons are space–time duals, and this duality has been exploited in the context of Kerr frequency combs [
12]. Some aspects of self-imaging in graded-index fibers [
21] and the formation of multimode solitons in such fibers [
48] are also related to space–time duality if we recall that spatial self-imaging is the dual of temporal oscillations of a harmonic oscillator.