Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Angular dispersion: an enabling tool in nonlinear and quantum optics

Open Access Open Access

Abstract

The dispersive properties of materials, i.e., their frequency-dependent response to the interaction with light, in most situations determines whether an optical process can be observed. Although one can always search for a specific material with the sought-after properties, this material might be far from optimum or might not even exist. Therefore, it is of great interest to develop methods that could tune the dispersive properties of a medium independently of the working frequency band. Pulses with angular dispersion, or pulse-front tilt, precisely allow us to achieve this goal. In this tutorial, we show the basics of how angular dispersion can manage to tune the dispersion parameters that characterize the propagation of light in a medium, thus permitting the observation and application of various optical processes in nonlinear and quantum optics that could not be realized otherwise. To keep the focus on first principles, the list of topics addressed is not exhaustive. More specifically, we consider the role of angular dispersion for pulse stretching and compression, broadband second-harmonic generation, the generation of temporal solitons in nonlinear χ(2) media, the tunable generation of terahertz waves by means of optical rectification of femtosecond pulses, and the tuning of the frequency correlations and of the bandwidth of entangled paired photons.

© 2010 Optical Society of America

1. Introduction

When light interacts with a material, the atoms and molecules that constitute the medium respond differently depending on the frequency of the light, ω. For low intensities, the relationship between the induced polarization P(ω) and the electric field E(ω) is linear, P(ω)=ϵ0χ(1)(ω)E(ω), where the constant of proportionality is the frequency-dependent susceptibility χ(1)(ω) and ϵ0 is the permittivity of free space. The frequency dependence of the susceptibility is the so-called dispersion of the medium. One easily observable manifestation of the dispersive nature of materials was described in 1704 by Isaac Newton in his book Opticks: or a Treatise of the Reflections, Refractions, Inflections and Colours of Light [1]. Newton showed that when a prism deflects a white-light optical beam, each of the colors that make up the white light leaves the prism in a different direction (see Fig. 1). The refractive index of the medium is wavelength dependent, and therefore the angles of refraction governed by Snell’s law correspondingly change at each interface in accordance with wavelength.

Dispersion and diffraction are the main linear effects that describe the propagation of a light beam in a medium. In the linear regime, the propagation of the light beam for a distance z can be easily described by multiplying the electric field amplitude at z=0, a(q,ω,z=0), by a phase factor exp{ik(q,ω)z}, where q is the transverse wavenumber, ω is the angular frequency, and k is the longitudinal wavenumber. Normally, the bandwidth of a light pulse is much smaller than its central frequency, so the wavenumber k can be expanded in a Taylor series about the central frequency. The number of terms of the expansion to be included increases with the bandwidth, but in most situations one can safely keep the expansion to second order. In this case, the group velocity and the group-velocity dispersion (GVD) are the two parameters that characterize the linear dispersion of the medium.

With the invention of the laser, it was possible to observe optical phenomena that generate material polarizations that are no longer linearly proportional to the electric field amplitude, but that depend on its higher powers. In this case, the strength of the different nonlinear terms is determined, among other things, by the nonlinear susceptibilities of the medium, χ(n), with n=2,3,. Examples of these nonlinear effects are the generation of harmonics [2, 3] and self-phase modulation (SPM) [4].

In order to observe any of these processes, the dispersive properties of the material have to be taken into account. In certain situations, the presence of dispersion is harmful and one must look for materials that exhibit negligible dispersion over a sufficiently large bandwidth. This is the case, for instance, for the generation of ultrashort pulses by means of second-harmonic generation (SHG): owing to its intrinsic dispersion, the material can severely limit the bandwidth of the generated second-harmonic pulse, imposing an effective lower limit on the duration of the pulse that can be generated. In other situations, the dispersion is not harmful, but, to the contrary, a certain amount of dispersion is required, and in addition must be of the appropriate sign. This is the case for the observation of quadratic temporal solitons in χ(2) media and for the stretching and compression of pulses.

Although one can always search for a specific material with the sought-after linear dispersion properties, this material might be far from optimum or might not even exist. Moreover, certain applications need to be implemented in specific frequency bands. Even if an appropriate material can exist in one frequency band, it might be useless in another band. Therefore, it is of great interest to develop methods that could tune the dispersive properties of a medium independently of the material and the working frequency band.

In this tutorial, we will review how the introduction of angular dispersion into pulses, or pulse-front tilt, can tailor the dispersive response of any material in any frequency band. By applying angular dispersion to pulses, one can effectively generate a new group velocity and group-velocity dispersion as required by the specific application in mind. This capability has been used in different areas of nonlinear and quantum optics, but the different approaches, the emphases on different aspects of the process, and the different language used make it difficult to see what is the role of pulses with angular dispersion with a single view that might encompass all the different applications. The goal in this tutorial is to give an overall view of the role of angular dispersion that could be applied in the various applications of nonlinear and quantum optics that use these techniques.

Although the dispersive properties of a light beam that propagates in any given material can be changed in many different ways—for instance, by changing the temperature or by considering a different direction of propagation inside the nonlinear medium—these changes are related to modifications of the material properties of the media. Our interest here are the modifications of the dispersive properties introduced by changing the spatiotemporal structure of the light beam itself with appropriately designed dispersive elements, such as prisms or gratings.

In Section 2, we analyze the role of angular dispersion, or pulse-front tilt, in the propagation of light beams in dispersive media and derive the main equations that will be used throughout the tutorial. Sections 3, 4 will describe five optical process and applications that use pulses with angular dispersion.

In Section 3 we consider applications in the field of nonlinear optics. We begin with techniques used to obtain pulse stretching and pulse compression to show that, even in free space, with no dispersive medium present, one can induce an effective anomalous group-velocity dispersion that can be used to broaden or compress optical pulses. Next we analyze how pulses with pulse-front tilt can enlarge the bandwidth of the nonlinear process of second-harmonic generation, which allows frequency upconversion of ultrashort pulses and the generation of ultrashort pulses in new frequency bands. We also analyze why the use of angular dispersion allows the observation of quadratic temporal solitons. Without angular dispersion, the currently used materials exhibit neither the appropriate amount of dispersion nor the appropriate sign. Furthermore, the group-velocity mismatch (GVM) between all the interacting waves is to large to let us observe temporal solitons with reasonable intensities. Finally, we consider the generation of tunable terahertz (THz) waves by means of changing the amount of angular dispersion present.

Section 4 reviews how it is possible to tune the frequency correlations and the bandwidth of entangled photon pairs generated during the process of spontaneous parametric frequency downconversion, which is achieved by tuning the group velocity of the interacting waves by means of angular dispersion.

2. Angular Dispersion: How Does it Work?

The aim in this tutorial is to demonstrate that angular dispersion can be used as a tool in nonlinear and quantum optics. The key point of this demonstration is to show that, in an appropriate configuration, dispersive elements allow us to modify the dispersive response of materials; in particular, they allow us to control the group velocities and group-velocity dispersion of light beams traveling through them.

In this section, the effects of angular dispersion in a medium flanked by two dispersive elements will be discussed. We will derive analytical expressions for the effective group velocities and effective group-velocity dispersion of materials placed between two gratings. These expressions will be the key points used throughout the rest of the tutorial to exemplify how angular dispersion can be considered a tool in nonlinear and quantum optics.

The tutorial will be centered mainly on angular dispersion introduced by diffraction gratings. At the beginning of this section, the effects of a grating on a pulse beam impinging on it will be considered. In particular, it will be shown that angular dispersion produces pulse-front tilt; i.e., it tilts the front of a pulse by a certain angle. The value of the angle that describes this tilt will be used to quantify the effect of dispersive elements. Moreover, it will be shown that the tilt angle is indeed the parameter that will allow us to tune the dispersive properties of different media. For completeness, the derivation of the pulse-front tilt angle produced by a prism is also included.

Finally, a brief discussion about the terms angular dispersion and pulse-front tilt is presented. Although in most parts of the tutorial both terms will be used without distinction, we will see that indeed they represent different physical effects. In most situations they appear together, but one can find scenarios where there are pulses with pulse-front tilt but without angular dispersion and vice versa.

2.1. Effect of a Diffraction Grating on an Optical Beam

A reflection diffraction grating (see Fig. 2) is formed by a periodically corrugated reflecting surface, where the reflecting elements, the grooves, are separated by a distance comparable with the wavelength of light. An incident beam of light is separated, on reflection, into different waves; i.e., several reflected waves come off at different angles that mark the individual diffraction orders. The amount of energy that is actually reflected in each diffraction order depends on the specific shape of the grooves on the grating surface and can be appropriately designed [5]. On the other hand, the directions of propagation of the reflected waves are independent of the specific shape of the grating surface and are given by the grating equation [6]

sinϵ¯+sinθ¯=mλd,
where θ¯ is the angle of incidence, ϵ¯ is the angle of diffraction, λ is the wavelength of the radiation, d is the groove spacing, and m is the diffraction order.

Let us consider an incident wave with a spectral distribution centered at angular frequency ω0 (that corresponds to a wavelength λ0 in the medium) impinging on a grating. In this case, the expression for the angular dispersion, γ, produced by a grating can be found by expanding Eq. (1) up to first order. Let us assume that the grating is oriented in such a way that the angular dispersion occurs in the x direction. The angles of incidence and diffraction at the grating can be expressed as θ¯=θ0+θ and ϵ¯=ϵ0+ϵ, where θ0 and ϵ0 are the central angles of incidence and diffraction, and θ and ϵ are small deviations from the corresponding central values. Analogously, λ=λ0+Δλ. If we take into account that θθ0, ϵϵ0, and Δλλ0, Eq. (1) can be written up to first order as [7, 8]

ϵ=αθ+γΔλ,
where
α=cosθ0cosϵ0
and
γ=mdcosϵ0
is the angular dispersion, defined as γ=(ϵλ)λ0.

In the slowly varying envelope approximation, the amplitude of an electric field at a spatial position (x,y,z) and at time t can be written as

E(x,y,z,t)=12A(x,y,z,t)exp(ik0ziω0t)+h.c,
where ω0=2πc(λ0n0), k0=n0ω0c and h.c. stands for Hermitian conjugate. A(x,y,z,t) is the slowly varying envelope of the electric field, n0 is the refractive index of the medium at frequency ω0, and c is the velocity of light in vacuum. In general, A(x,y,z,t) is conveniently written as the transverse wavenumber variable q=(qx,qy) and the deviation from the central frequency Ω:
A(x,y,z,t)dqxdqydΩa(qx,qy,Ω,z)exp(iΩt+iqxx+iqyy).

Equation (2) allows us to find a relationship between the wave vectors of the incident and the diffracted waves. For clarity, we will use the subindex 1 to refer to the variables of the beam before the grating and the subindex 2 for the variables after the diffraction at the grating. When impinging on a grating, the incident field is diffracted, and each frequency component is dispersed in a different direction. Consider a grating that introduces angular dispersion in the x coordinate, as depicted in Fig. 3. In this case, the variables of interest become the x component of q and the x component of the transverse wavenumber after diffraction at the grating, p=(px,py). By using geometrical considerations and the fact that θ and ϵ are small, it is possible to show that px=k0ϵ and qx=k0θ. For a grating immersed in a vacuum, the index of refraction is equal to one.

Considering that (ϵλ)λ0Δλ=(ϵω)ω0Ω, qx can be rewritten as

qx=pxαtanΦαcΩ,
where
tanΦ=ck0(ϵω)ω0.
We should notice that Eq. (8) is independent of the specific element that introduces the angular dispersion, and it is valid for a functional form such as the one given by Eq. (2).

Equation (7) relates the transverse wavenumber of the diffracted and incident beams. The transformation of the optical beam due to the presence of the grating placed at z1=z2=0 can then be written as

a(qx,qy,Ω,z1=0)a(qxαtanΦαcΩ,qy,Ω,z2=0).

To further illustrate the effects of a grating on the optical beam, let us consider the amplitude of the electric field at a distance z2 from the grating. Neglecting the temporal and spatial broadening of the optical pulse due to diffraction and dispersion effects, propagation over a distance z2 corresponds to the transformation

a(qx,qy,Ω,z2=0)a(qx,qy,Ω,z2=0)exp(ik0Ωz2),
where 1k0 is the group velocity of the pulse. From Eq. (9), the electric field amplitude at a distance z2 after diffraction at the grating is
A(x2,y2,z2,t)dqxdqydΩa(qxαtanΦαcΩ,qy,Ω,z2=0)exp(ik0Ωz2)exp(iΩt+iqxx2+iqyy2).
Performing the corresponding Fourier transforms, we obtain
A(x1,y1,z1=0,t)A(αx2,y2,z2,t[k0z2+tanΦcx2]).

Equation (12) describes the transformation of the shape of the electric field amplitude by the grating. This provides us with a further physical insight into the grating effect. For A(x1,y1,z1,t) the transverse and longitudinal components are independent; therefore, at any time, we observe that the front of the pulse is perpendicular to the propagation direction z1. The left-hand part of Fig. 3 depicts, for a fixed time, an input beam whose intensity is spatially described by a Gaussian function with a beam waist in the x direction w0x. On the other hand, after the beam has been reflected by the grating, Eq. (12) tells us that the transverse and longitudinal variables are no longer independent. As a consequence, the pulse-front is no longer perpendicular to the propagation direction, as shown in the right-hand part of Fig. 3.

To clarify this idea, consider the loci of the peak intensities

t(k0z2+tanΦcx2)=0.
It is clear from this expression that the temporal and spatial variables are not independent after the pulses passes through the grating. For a fixed time (t=0), the front of the pulse is given by the line z2=tanΦx2(ck0) in the plane (x2,z2). This corresponds to a straight line with a slope given by [9, 10]
tanνtanΦck0.
It is then said that, after passing through a grating, the pulse-front acquires a tilt given by the angle ν, and consequently the front of the pulse is no longer perpendicular to the propagation direction z2. By using Eqs. (2, 8) we obtain that the tilt angle for a grating is given by
tanΦ=n0λ0γ.

From Eq. (13), we also see that for a fixed distance z2 (z2=0) the maximum of the field arrives at different times, t=x2tanΦc, for each transverse coordinate x2. Therefore, it becomes evident that the angle Φ is the angle of the loci of the peak intensities in the (x2,ct) plane, as shown in Fig. 4.

In what follows, we will see that the tilt angle can be used as a control parameter to modify the dispersive response of a material at will, and in that way angular dispersion can become a tool in nonlinear and quantum optics.

2.2. How to Use Angular Dispersion to Control Material Dispersive Properties

Angular dispersion can be effectively and expediently used to control the dispersive properties of materials. Let us consider a medium flanked by two gratings or prisms as shown in Fig. 5. As we will see below, this configuration allows us to tune at will the group velocities, group-velocity dispersion (GVD), and higher dispersive terms of the fields propagating in the medium. It is precisely this point that makes angular dispersion an enabling tool in many areas of optics. We will see particular examples in Sections 3 and 4.

The effect of a diffraction grating on the propagation of a pulse is to tilt the front of the pulse by an angle ν, as depicted in Fig. 3. Alternatively, we can also see that the line of the loci of peak intensities is tilted by an angle Φ, as shown in Fig. 4.

To see how the particular configuration depicted in Fig. 5 can help tune the dispersive properties of the material, let us start by considering what happens when a pulse with pulse-front tilt enters a medium with a different refractive index. Let us assume that the beam enters the material normally. Making use of Snell’s law to first order in the angles and wavelength, we obtain

n1ϵ1=n2ϵ2,
where ϵ1 (ϵ2) is the angle of refraction inside (outside) the material, and n1 (n2) is the refractive index. The wavenumbers inside (k1) and outside (k2) the material are different and are related by k2=n2k1n1. With these expressions in hand, we can see that neither the transverse wavenumber nor the tilt angle Φ, given by Eq. (8), changes when the beam with pulse-front tilt enters the medium. For normal incidence, the amplitude of the field at the boundary after entering the material is still a(qxαΩtanΦ(αc),qy,Ω,z2=0).

Let us now consider the medium to be dispersive; i.e., let us let its refractive index vary with frequency. Further, we assume that the medium fills all the space between the two gratings, as shown in Fig. 5. From now on, we will work mostly with the diffracted beam, and, in order to simplify the notation, we will drop the use of the subindex 2 used to refer to the diffracted beams, so that (x2,y2,z2x,y,z). The electric field amplitude at any propagation distance z inside the medium can be written as

a(qx,qy,Ω,z)=a(qx,qy,Ω,z=0)exp[i(k(ω)k0)z]exp(i|q|22k0z)exp(iqxztanρ),
where k0=n0ω0c is the wavenumber at the central frequency ω0. The dispersive nature of the media reflects in the frequency dependence of the wavenumber, k(ω)=n(ω)ωc. The term exp(i|q|2(2k0)z) comes from using the paraxial approximation, and the term exp(iqxztanρ) takes into account the spatial walk-off the beam may suffer when it travels in an anisotropic material. In that case, the wave vector, which is perpendicular to the wave phase front, and the Poynting vector, which determines the direction of propagation of the energy of the wave, may propagate in different directions. A beam initially centered at x=0, at z=0 would be displaced transversely to x=ztanρ at a distance z in the absence of any other effects. ρ is the Poynting vector walk-off angle, or spatial walk-off angle. In frequently used nonlinear materials, such as β-BaB2O4 (BBO), LiNbO3, or KTiOPO4 (KTP), its typical values are ρ0°5° [11].

To understand the consequences of light with pulse-front tilt traveling through a dispersive media, let us expand k(ω) up to second order about ω0. Defining the inverse group velocity k0=(kω)ω0 and the inverse GVD k0=(2kω2)ω0, it follows that

k(ω)=k0+k0Ω+12k0Ω2.
By substituting Eq. (18) into Eq. (17), we obtain the field amplitude at z=L:
a(qx,qy,Ω,z=L)a(qxαtanΦαcΩ,qy,Ω,z=0)exp(i|q|22k0L)exp(iqxLtanρ)exp(ik0ΩL+i2k0Ω2L).

To see the effect of the second grating, let us consider a grating characterized by the parameters α and Φ. This second grating introduces angular dispersion in the x direction as well, and it is oriented in such a way that it satisfies αα=1 and tanΦ=tanΦα. With these parameters, the transformation of Eq. (9) for the second grating becomes

a(qx,qy,Ω)a(αqx+ΩtanΦc,qy,Ω).
The electric field amplitude just after diffraction off the second grating, placed at z=L can be written as
A(x,y,z=L,t)dqxdqydΩa(qx,qy,Ω,z=0)exp(iqxx)exp(iqyy)exp{iΩ(tk0L+αqxLtanΦk0ctanΦtanρcL)}exp{iΩ2[k0(tanΦc)21k0]L2}exp(iαqxtanρL)exp(iα2qx22k0L)exp(iqy22k0L).

To simplify this expression and see more clearly the physics behind it, let us assume that the input beam has an elliptical spatial shape with a large beam width in the x direction, w0xw0y. Then a(qx,qy,Ω,z)a(qy,Ω,z)δ(qx), and the field of Eq. (21) becomes

A(y,z=L,t)=dqydΩa(qy,Ω,z=0)exp(iqyy)exp(iqy22k0L)exp{iΩ[t(k0+tanΦtanρc)L]}exp{iΩ2[k0(tanΦc)21k0]L2}.
Notice that in order to make this approximation valid, the size of the beam in the x transverse dimension, w0x, should be larger than the lateral displacement of the beam due to spatial walk-off, Ltanρ. This implies that the validity of Eq. (22) is restricted to propagation distances z<w0xtanρ. For example, for w0x=1mm and a walk-off angle ρ=5°, the propagation length should be L1cm.

From Eq. (22), we can see that the evolution of the spatial shape in the y transverse coordinate is the well-known expression for the diffraction of a beam. In the time domain, the effect of the presence of pulse-front tilt can be described by the introduction of two effective dispersive parameters [12, 13, 14]: an effective inverse group velocity,

k0,eff=k0+tanΦtanρc,
and an effective group-velocity dispersion (GVD),
k0,eff=k01k0(tanΦc)2.

It is important to remark that Eq. (23) refers to the effective group velocity along the propagation direction z, k0,eff. In this case, the change of group velocity requires the existence of spatial walk-off (ρ0). For this reason, optical beams with pulse-front tilt propagating in vacuum or in noncritical directions (with ρ=0) in birefringent crystals do not modify their group velocity when placed between two gratings. On the other hand, the situation is different for the effective GVD: even when no dispersive material is present, it is possible to observe an effective anomalous dispersion. This can be clearly seen by setting k0=0 in Eq. (24) so that the effective anomalous dispersion is k0,eff=tan2Φ(k0c2).

The fact that an effective anomalous dispersion always accompanies angular dispersion in a vacuum was shown in [15]. It was demonstrated that the optical transfer function for a beam traversing a pair of gratings, arranged in tandem, contains a quadratic frequency term that is responsible of the appearance of anomalous dispersion. This fact allows GVD-free propagation to be achieved in dispersive media as demonstrated in [16].

Equations (23, 24) are the key results of this section. They demonstrate that by means of the tilt angle Φ, i.e., by the amount of pulse-front tilt introduced by diffraction gratings, it is possible to modify the inverse group velocity and the GVD parameters of a material placed between the two grating as depicted in Fig. 5. It is precisely this capability that will become the crucial point to make angular dispersion a tool in nonlinear and quantum optics.

The propagation of pulses with pulse-front tilt Φ produced by angular dispersion in a dispersive medium can be described by an effective inverse group velocity and an effective inverse group-velocity dispersion (GVD) given by

k0,eff=k0+tanΦtanρc,
k0,eff=k01k0(tanΦc)2.

2.3. Angular Dispersion Produced by a Prism

Diffraction gratings are not the only optical devices that can introduce angular dispersion [17, 18]. One outstanding example is the prism [19]. Recalling Eq. (8), the tilt angle is determined solely by the amount of angular dispersion present. For completeness, here we derive the pulse-front tilt introduced by prisms following the general procedure used previously for the grating.

Let us consider a prism with an apex angle C and refractive index n(λ). Light impinges on the prism at an angle θ¯ and exits at an angle ϵ¯ after suffering refraction at the two interfaces of the prism, as depicted in Fig. 6. Using Snell’s law to describe refraction at both surfaces, one has

sinθ¯=n(λ)sinδ¯1,
n(λ)sinδ¯2=sinϵ¯.
The meaning of angles δ¯1,2 can be seen in Fig. 6.

The light impinging on the prism has a spatial and spectral distribution centered at θ0 and λ0, respectively. The spatial and spectral variables can be conveniently written as θ¯=θ0+θ, ϵ¯=ϵ0+ϵ, and λ=λ0+Δλ.

Expanding Eqs. (25) and (26) up to first order about θ0, ϵ0, and λ0, one obtains

ϵ=αθ+γΔλ,
where
α=cosθ0cosϵ0cosδ20cosδ10,
γ=sinCcosϵ0cosδ10(nλ)λ0,
with sinδ10=sinθ0n0 and sinδ20=sinϵ0n0.

Notice that Eq. (27) is equal to Eq. (2), which was obtained for a diffraction grating, although each equation was derived in a different way. For the grating, Eq. (2) comes from the grating equation that is a consequence of interference at a periodic structure, while Eq. (27) comes from Snell’s law, which is a consequence of material dispersion. When expanded to first order, both devices lead to the same effect: the linear dependence of the angle of diffraction (refraction) on the angle of incidence and the wavelength. We mention that it is possible to derive a general framework that features the common working principle of different spectroscopic devices [20]. With this approach, the dispersion produced by a prism can be treated by the same formalism as the dispersion introduced by a grating.

For minimum deviation configurations in which θ0=ϵ0, and for incidence at the Brewster angle, tanϵ0=n0, Eq. (27) becomes [18]

ϵ=θ+2(nλ)λ0Δλ.
Using the expression for the tilt angle given by Eq. (8), one readily obtains
tanΦ=2λ0(nλ)λ0.

Let us have a first glimpse at the amount of tilt induced by a prism and a grating. Let light at 800nm illuminate a commercially available grating with G=1d=1400lines/mm. Let the input angle be θ0=20°, which results in an output angle of the first diffraction order m=1 of ϵ0=51.1°. In this case, the angular dispersion is (ϵλ)λ0=0.128°nm, and the tilt angle, given by Eq. (15), is Φ=60.7°. On the other hand, for a prism [21] with a refractive index n0=1.457 and (nλ)λ0=0.002°nm at a wavelength of 600nm, the tilt angle is Φ=2.4°. The resulting values of the tilt angle give a hint that one can obtain less angular dispersion with currently used prisms than with conventional gratings. Notwithstanding, losses can be much higher in setups with gratings than in those with prisms, unless an optimized grating is designed. We should notice that a larger amount of pulse-front tilt can be obtained by combining sequences of prisms [18] or by using combinations of gratings and prisms [22].

2.4. Pulse-Front Tilt versus Angular Dispersion

Until now, we have seen that angular dispersion generates pulse-front tilt. In the cases considered, we have used both concepts indiscriminately to describe the effect of a prism and of a grating. But is the introduction of angular dispersion the only option to generate pulse-front tilt? This question was addressed in [23], where the authors showed that spatial chirp may also lead to the generation of pulses with pulse-front tilt. Although there may be alternative procedures, for the sake of simplicity we will restrict ourselves to this case.

Previously we analyzed the concept of the tilt angle in the spatial domain (x,z) and in the spatiotemporal domain (x,ct). Since the spatial chirp can be conveniently described [24] in the spatial and frequency degrees of freedom (x,Ω), in what follows we will focus our analysis on these variables. First, let us consider the amplitude of a pulse on diffraction off a grating in the (x,Ω) variables. This is obtained by Fourier transforming Eq. (12) from the time to the frequency domain and for z2=0. To obtain analytical results, we assume the input beam to have a Gaussian shape in time and space, characterized by the widths T0 and w0=w0x=w0y, respectively. With these assumptions, the complex envelope of the beam on diffraction is written as

A(x,Ω)=A0exp(α2x2w02y2w02)exp(Ω2T024+itanΦcΩx),
where A0 is an arbitrary constant. Notice that angular dispersion is characterized by the dependence exp(iμΩx), where μ=tanΦc.

Let us consider light beams that apart from angular dispersion display linear spatial chirp and GVD characterized by the parameters ξ and δ, respectively. The amplitude of the beam is now written as

A(x,Ω)=A0exp[(xξΩ)2w02]exp(Ω2T024+iδΩ22+iμΩx).
The Fourier transform of Eq. (33) gives us the spatiotemporal shape of the pulse. After some straightforward calculation, one obtains that the intensity of the beam [I(x,t)=|A(x,t)|2] is [23]
I(x,t)exp[2(tδv¯xμx)2τ¯2]exp(2x2w¯02),
where
v¯=ξξ2+T02w024,
τ¯=(T02+4ξ2w02+4δ2T02+4ξ2w02)12,
w¯0=[1w02v¯2(T024+ξ2w02)]12.
The important point to note is that, even in the absence of angular dispersion (μ=0), there is pulse-front tilt if v¯0. That requires the presence of spatial chirp (ξ0) combined with GVD (δ0). In other words, the combination of spatial chirp and GVD can also generate pulses with pulse-front tilt.

Pulses diffracted by two gratings, as described in the previous sections, are a good example of pulses that show pulse-front tilt without angular dispersion. Between the two gratings, the beam propagates in free space. Performing the Fourier transform of A(x,y,z,t) of Eq. (21) from time to frequency, we obtain

A(x,Ω)exp(Ω2T024+izΩc)exp{(xαzμΩk0)2w02+2iα2zk0}exp{iμ2z2k0Ω2},
where k0 is the wavenumber in free space and z is the distance between the gratings.

Comparing with Eq. (33), we can see that the diffraction effects arising from the propagation from one grating to the other introduce spatial chirp ξ=αzμk and an effective second-order dispersion parameter δ=μ2zk.

One can have pulses that show pulse-front tilt and no angular dispersion. This can be achieved by introducing, for instance, spatial chirp and group-velocity disperion (GVD). The combined effect of two dispersive elements and diffraction produced by the propagation of the pulse from one element to the other can effectively generate the spatial chirp and GVD required.

3. Angular Dispersion in Nonlinear Optics

Since the invention of the laser, nonlinear optics has become one of the most fruitful areas of optics. Nonlinear effects are employed in many important applications of optical technologies, such as optical fiber communications or high-resolution imaging and detection of biological tissue. Two important concepts that are at the core of nonlinear optics are considered here: the generation of optical waves at new wavelengths, analyzed in Subsections 3.2, 3.4, and the existence of an optical entity that can exist only in the realm of nonlinear optics: the soliton. Solitons are analyzed in Subsection 3.3.

Paradoxically, the observation of most effects in nonlinear optics depends on the specific linear properties of the materials used. It is here where angular dispersion, a linear effect, plays an important role, making it possible to modify the dispersive properties of materials to allow the observation of certain nonlinear effects [25].

3.1. Pulse Compression and Pulse Stretching

The first two important applications that we analyze in this tutorial that use the angular dispersion introduced by a series of gratings or prisms are pulse compression and chirped pulse amplification (CPA). In both cases, the combined use of nonlinear effects and dispersive effects aims at shortening (pulse compression) or broadening (pulse stretching) pulses in the temporal domain in order to amplify ultrashort pulses.

3.1a. Pulse Compression Techniques

The goal of pulse compression techniques considered in this subsection is to modify the properties in frequency and time of a transform-limited pulse of time duration T1 in order to generate a new pulse with a shorter duration T2<T1. This can be achieved with the scheme shown in Fig. 7. The goal of the first stage is to broaden the spectrum of the pulse. For this purpose, we can use, for instance, the nonlinear effect of self-phase modulation (SPM) in an optical fiber [4]. In SPM, the temporal width of the pulse is not changed, but because of the appearance of a quadratic temporal phase chirp, the bandwidth is enhanced. Afterward, one needs to translate the frequency broadening into pulse compression, erasing any frequency chirp introduced in the first stage. This will render the new pulse transform limited again, but this time with a shorter time duration thanks to the increased bandwidth.

Let us assume that the input pulse is a transform-limited Gaussian pulse that can be written in the temporal and frequency domains as

A(t,z=0)=A0exp(t2T12)a(Ω,z=0)exp(Ω2T124),
where Ω is the frequency deviation from the central frequency ω0 of the pulse. The full width at half-maximum (FWHM) bandwidth of the pulse is B1=(8ln2)12T1. In an optical fiber, depending on the peak power and the time duration of the pulse, many dispersive and nonlinear effects might have to be considered. For the sake of argument, let us suppose that the main effect that affects the pulse propagation in the optical fiber is SPM. In SPM, there is a time-dependent phase change, proportional to the intensity of the pulse, that is added during propagation:
A(t,z)=A(t,z=0)exp(iγfP(t)z),
where P(t)=|A(t,z)|2 is the power, z is the length of the fiber, γf=ω0n2(cAeff) describes the effective nonlinearity induced by the fiber, n2 is the nonlinear index coefficient, and Aeff is the effective mode cross section. To get further physical insight and obtain some analytical results, we expand the expression of the power about t=0 so that P(t)P0(12t2T12). In this way, we get an approximate expression of the pulse in time at the end of the nonlinear fiber that is written as
A(t,z)=A0exp[t2T12(1+2iγfP0z)],
where the constant time-independent term iγfP0z has been omitted for the sake of simplicity.

After Fourier transforming this equation, we get

a(Ω,z)exp{Ω2T124[1+(2γfP0z)2]+iΩ2γP0zT122[1+(2γfP0z)2]}.
From Eq. (42), we can see that the frequency bandwidth B2 of the pulse after propagation in the fiber increases and is given by
B2=B1[1+(2γfP0z)2]12.
As an example, for a wavelength of λ0=620nm, Aeff=50μm2, and n23.2×1020m2W one obtains γf6.5W1km1. For a peak power of P02kW and a fiber 15cm long, Eq. (43) predicts a theoretical fourfold enhancement of the bandwidth.

Equation (42) shows that SPM also introduces a positive quadratic frequency chirp of the form exp(iαSPMΩ22) with

αSPM=γfP0T12z[1+(2γfP0z)2].
This must be compensated in order to translate the increase of the frequency spectrum into the generation of a shorter pulse.

The angular dispersion introduced by gratings or prisms can be used to compensate such a quadratic chirp. As was shown in Subsection 2.2, a pair of gratings separated by a distance L in a vacuum introduces a quadratic negative frequency chirp keff=tan2Φ(kc2). Therefore, to achieve compensation it is required that αSPM=keff.

The combined effect of self-phase modulation, which broadens the bandwidth of a pulse while it propagates down an optical fiber, and of the angular dispersion introduced by dispersive elements (gratings and/or prisms), which renders the resulting pulse transform limited, allows the generation of a new, compressed pulse with a shorter time duration.

As an example, in [26] 90fs input pulses at the wavelength of 619nm were focused into a polarization-preserving fiber 15cm long. The authors observed a factor of 3 increase in the frequency spectrum, from 6nm to about 20nm. A pair of gratings with 600 linesmm and a slant angle of 30° set 6.4cm apart were used to compress the pulse after propagation in the fiber, measuring the output pulses with a time duration of 30fs.

For the sake of simplicity, we have considered only the effects of the presence of quadratic phase terms in the frequency domain. Notwithstanding, apart from SPM, other dispersive and nonlinear effects might cause the appearance of nonquadratic frequency chirp terms, especially when we are dealing with ultrashort pulses. The effects to be included are higher-order dispersion, cross-phase modulation, self-steepening, and the self-induced Raman effect [4]. On the other hand, pairs of gratings also introduce nonquadratic phase terms that should also be taken into account. With this information in hand, one can use appropriately engineered combinations of prisms and gratings to erase the frequency chirp generated in the nonlinear fiber. In [27], the authors used combinations of prisms and gratings to compensate not only for the quadratic but also for the cubic phase of ultrashort optical pulses. They obtained compressed pulses as short as 6fs.

In general, as one moves to shorter pulses and higher peak powers, the phase terms, induced by nonlinear and dispersive effects during propagation in the fiber, present more complicated shapes in an increasingly larger bandwidth. In order to compensate for these phases for generating ultrashort pulses, one needs to consider appropriately tailored arrangements of prisms and gratings that are correspondingly more sophisticated [19].

3.1b. Chirped Pulse Amplification

Until now, we have analyzed a two-stage pulse compression scheme: the first stage broadens the spectrum, and the second stage renders the pulse transform limited. Certain applications require not only pulse compression at a certain stage, but also pulse stretching in another stage. This is the case for chirped pulse amplification (CPA), a scheme to amplify pulses that avoids the serious damage that high peak powers of several gigawatts per square centimeter can cause to gain media due to the effect of self-focusing [28]. To avoid this damage, one should reduce the peak power of the pulse before the pulse enters the amplifying stage.

The general scheme of CPA is shown in Fig. 8(a). The first stage is intended to reduce the peak power of the input pulse by introducing a quadratic phase term. Although this can be achieved by propagation in an optical fiber, the high powers used can nonetheless generate other undesirable effects that can be avoided by employing angular dispersion with pairs of gratings. If we again consider an input pulse given by Eq. (39) with peak power P1, the pulse at the end of the pulse stretching stage is written as

A(t,L)=A0T1T122ikeffLexp(t2T122ikeffL),
where keff is the effective GVD introduced by the pair of gratings of the pulse stretching stage and L is the separation between them. The peak power P2 of the pulse after the first stage is
P2=P11+(2keffLT12)2,
and its time duration T2 is
T2=T11+(2keffLT12)2.
As an example, let us consider a pair of gratings with 1400 linesmm, separated by a distance L=20cm. For an angle of incidence of 20°, the effective GVD at 800nm is k=4495fs2mm. A 100fs pulse is stretched to 18ps, and the peak power after the pulse stretching stage is P26×103P1.

The second stage of CPA is the amplification of the pulse. Assuming that the gain bandwidth is broad enough to amplify the whole frequency spectrum of the pulse and that it does not introduce any frequency chirp, the final stage should compensate for the angular dispersion introduced in the first stage. Therefore, it should introduce an effective quadratic chirp of value keffL.

The role of angular dispersion in chirped pulse amplification (CPA) is to stretch a pulse before it is amplified in order to reduce its peak power and to recompress it afterwards.

We have seen in Subsection 2.2 that a pair of gratings separated by a distance L can introduce only an effective anomalous dispersion (negative dispersion). To generate effective GVD values with a positive sign, we have to introduce elements that modify the propagation of the pulse between the two gratings. One example is shown in Fig. 8(b). Two lenses are separated a distance d from each grating. The separation between the two lenses is 2f, where f is the focal lens of each lens. After some straightforward calculation [29], which takes into account the paraxial propagation of the pulse from one grating to the other, one finds that the transfer function of the system can be written as

H(q)=exp{ifdk0|q|2}.
Using the transformation of Eq. (20), one can see that a quadratic phase term in frequency of the form exp(iβΩ2) appears as
β=fdk0(tanΦc)2.
For d<f, the pair of gratings and the two lenses introduce an effective positive dispersion (β>0), while for d>L the dispersion introduced is negative (β<0). Therefore, one system can be used in the pulse stretching stage, and the other in the final pulse compression stage. If the amplification process introduces any additional quadratic phase frequency term, the pulse compression scheme of the final stage can be properly tailored to remove any quadratic phase term present. As an example, in [30] several compression and stretching stages were used to generate transform-limited pulses with 100μJ of energy and 340MW peak power, using mismatched grating stretcher–compressors.

3.2. Achromatic Phase Matching

3.2a. Broadband Second-Harmonic Generation

Second-harmonic generation (SHG) is a nonlinear process in which light at a certain frequency ω1 (fundamental frequency, FF) interacts with a nonlinear crystal and generates a nonlinear polarization in the medium at double frequency ω2=2ω1 that causes the generation of optical radiation at the same frequency ω2 (the second harmonic, SH). At a more fundamental level, SHG is a process in which the atoms or molecules that make up the nonlinear material absorb two photons at frequency ω1 and re-emit a single photon at frequency ω2. To do so, the interacting waves have to satisfy the energy and momentum conservation laws.

In the field of nonlinear optics, momentum conservation is usually called the phase-matching condition, which comes from the requirement to match the phases of the interacting waves [29, 31, 32]. The electric field of the fundamental and second-harmonic waves can be written as Ei(x,y,t,z)=12Ai(x,y,t,z)exp(ikiziωit)+h.c., i=1,2, and the consequent phase-matching condition is

k2=2k1.
Normally we try to work with optical materials that are transparent at the wavelengths of interest. Due to Kramers–Kronig relations that relate the real and imaginary parts of susceptibility (i.e., relate dispersion to absorption), the absence of losses results in normal dispersion. Since ki=ni(ωi)ωic, the condition imposed on refractive indices, i.e., n2(ω2)=n1(ω1) that results from Eq. (50), would be difficult to satisfy.

On the other hand, for one wavelength it is relatively easy to satisfy the condition given by Eq. (50) in birefringent media where the polarization of the fundamental and second-harmonic waves are chosen to be mutually orthogonal and their refractive indices follow different dispersion curves. It is, however, more complicated to satisfy the phase-matching condition for a broader range of frequencies, which is necessary for the conversion of short light pulses that exhibit a correspondingly broader spectrum.

The k vectors of the interacting fields can be expanded in a Taylor series about the central frequency, i.e., ki(ωi+Ωi)=ki+kiΩi+12kiΩi2, where ωi are the central frequencies and Ωi are frequency deviations. The fulfilment of the phase-matching condition given by Eq. (50) over a broader range of frequencies can then be satisfied when the group velocities of the interacting waves are equal,

k2=k1,
i.e., when the group velocity mismatch (GVM) is zero, GVM=k2k1=0. Nevertheless, in most materials this is usually not the case for the wavelengths of interest, which are determined by the lasers at our disposal and by the wavelengths where available detectors exhibit high efficiencies.

The GVM could be neglected when the crystal length L is much smaller than the temporal walk-off length Lgvm, LLgvm, which is defined as

Lgvm=T0|k2k1|,
where T0 is the duration of the fundamental pulse. In this case, when the conversion efficiency is small and the FF intensity can be considered constant (undepleted approximation), it can be found that for large-area beams the efficiency of the nonlinear conversion in a nonlinear crystal of length L is proportional to [33, 34]
I2(L)I1(0)I1(0)L2sinc2(ΔkL2)=I1(0)L2sin2(ΔkL2)(ΔkL2)2,
where I1,2 are the corresponding intensities and Δk=k2(ω2+Ω2)2k1(ω1+Ω1). If we design the SHG configuration so that perfect phase matching is achieved at the central frequencies, i.e., k2=2k1, the efficiency of the SHG process decreases with frequency as sinc2(ΔkL2), since Δk0 for frequencies different from the central frequencies.

It is possible to increase the bandwidth of spectral acceptance by using a shorter nonlinear crystal, because the spectral acceptance is inversely proportional to (k2k1)L. A major disadvantage is a serious reduction of the conversion efficiency, because the signal energy at the central frequency is proportional to the square of the crystal length. The decrease of the crystal length would require a corresponding increase of the input power to compensate for the reduction of conversion efficiency. By extension, the reduction of the spectral acceptance of the nonlinear crystal effectively sets a minimum duration of the SH pulse that can be achieved. For example, in [35], the spectral acceptance of a 1mm long BBO crystal was measured at a FF of 496nm. The obtained acceptance of 0.52nm imposes an approximate minimum pulse duration of 700fs, assuming a Gaussian pulse shape.

It is here where the angular dispersion comes to help. It was shown in Subsection 2.2 that with the help of angular dispersion it is possible to control the group velocities and higher-order dispersion terms. Simply changing the angle of incidence of light at a diffraction grating, we can tune the dispersive properties to the desired values.

Let us try to get a further insight into the effects of the use of pulses with angular dispersion in the process of SHG [36, 37]. Let us consider the full frequency dependence of the FF and SH wave vectors k1(ω1+Ω1) and k2(ω2+Ω2), respectively. In order to enhance the efficiency of SHG over a larger frequency range, the phase-matching condition must be fulfilled over a broader range of frequencies. If the beam passes through a prism or a grating that introduces angular dispersion, each frequency of the outgoing diverging beam propagates in a different direction. One can choose such angular dispersion so that each frequency enters the nonlinear medium at such an angle that the phase-matching condition is satisfied for every frequency along its direction of propagation. The phase-matching condition, given by Eq. (50), requires now that

n1(ϵ(λ1),λ1)=n2(ϵ(λ2),λ2),
where ϵ(λi) is the propagation direction of each frequency inside the nonlinear crystal, λ1 is the wavelength of the FF wave in a vacuum, and λ2 is the wavelength of the SH wave in a vacuum.

We let λ10 and λ20 denote the central wavelengths of the FF and SH waves in vacuum, respectively, and ϵ0 the angle of the direction of propagation of the central frequencies. The refractive indices of both waves are denoted n1 and n2. If we expand Eq. (54) to first order and use the expression for the total derivative (dn1dλ)λ10=(n1λ)λ10+(n1ϵ)ϵ0(ϵλ)λ10, and similarly for the SH wave, we obtain

(ϵλ)λ10=(n1λ)λ1012(n2λ)λ20(n2θ)θ0(n1θ)θ0,
which gives us the angular dispersion necessary for achieving the group-velocity-matching condition.

The use of pulse-front tilt allows us to fulfill the condition LLgvm by effectively increasing the temporal walk-off length Lgvm. Its maximum value is obtained when the effective group velocities are equal, k1,eff=k2,eff. Without loss of generality, let us consider a type-I SHG process, where both the FF and the SH waves might be extraordinary waves. Let us designate the spatial walk-off angles as ρ1 and ρ2. The effective group velocities of the interacting waves are (see Subsection 2.2)

k1,eff=k1+tanΦtanρ1c,
k2,eff=k2+tanΦtanρ2c.
Using the expression of the tilt angle given by Eq. (8), one can show that the group-velocity matching (k1,eff=k2,eff) requires that
(ϵλ)λ10=(n1λ)λ1012(n2λ)λ20n1(tanρ2tanρ1),
where we use the phase-matching condition n1=n2 and ck1,2=n1,2λ1,2(n1,2λ1,2). Recall that the derivatives for the SH are taken at half the fundamental wavelength, i.e., λ20=λ102.

Let us consider SHG in a uniaxial birefringent crystal. The walk-off angle (ρρ1,ρ2) can be written as [11]

ρ(θ)=θtan1{no2ne2(θ)tanθ},
where θ is the angle of the direction of propagation of the wave with respect to the optic axis of the crystal and no and ne(θ) are the ordinary and extraordinary refractive indices, respectively. After some tedious but straightforward calculations, one finds that
n1,2θ=n1,2tanρ1,2.
Substituting Eq. (59) into Eq. (57), we again obtain the required angular dispersion for the group-velocity matching [36] given by Eq. (55).

We should notice that Eq. (55) does not explicitly show the need to use a configuration with a nonzero spatial walk-off for the FF and/or SH waves. However, using Eq. (59), we conclude that a configuration where the FF or SH waves experience spatial walk-off is required. It should be noted that this is true in collinear configurations only, where the FF and SH waves propagate in the same direction inside the nonlinear crystal. Group-velocity matching can also be obtained without spatial walk-off in noncollinear geometries, where the FF and SH waves propagate in different directions, as described in [25, 38]. In most of these configurations, even if the spatial walk-off is nonzero, it plays a minor role, because the main correction to the pulse group velocity results from the angle of the noncollinear interaction.

The previous analysis gives us a clear picture of how the use of beams with pulse-front tilt allows us to increase the spectral acceptance bandwidth of the SHG process. Through selection of the appropriate amount of angular dispersion, one generates new effective group velocities of the interacting waves, making the group velocities of the FF and SH waves equal. This is equivalent to selecting directions of propagations for each frequency such that each frequency fulfills the condition of phase-matching in its own direction.

Figure 9(a) shows the evolution of a SH pulse with GVM compensation. Thanks to the perfect phase matching over a broader range of frequencies, the SH quickly and efficiently builds up. By comparison, in Fig. 9(b) without GVM compensation, we can see the SH wave hardly appears, and after a few millimeters of propagation, it disappears completely owing to backconversion that transfers the energy back into the fundamental wave. The evolution of the slowly varying envelopes depicted in Fig. 9 is calculated by use of the evolution equations described in Subsection 3.2b.

Achromatic phase matching uses angular dispersion to modify the group velocity of the fundamental and the second-harmonic (SH) waves in the process of second-harmonic generation (SHG). When the effective group velocities are made equal, the spectral acceptance of the frequency-doubling crystal increases, which allows the generation of SH pulses with enhanced bandwidths and thus shorter time durations.

One of the advantages of the method described here is the possibility to control the group velocities of the interacting waves in any frequency region of interest and in any nonlinear crystal, which significantly extends the range of frequencies and materials that can be used for the generation of ultrashort pulses through the SHG of short input pulses. It is especially important when no materials are available that can directly be used at specific wavelengths of interest [39]. For instance, broadband SH pulses have been generated in type-I BBO at 258nm in a 7mm long crystal [40], at 330nm in a 4mm long crystal [41], or around 456nm and at 527nm in a type-II 3.77mm long BBO crystal [39]. In all of these cases, if the nonlinear crystals were used without pulse-front tilt, the bandwidth and the efficiency of the SH wave would have been severely reduced.

Another example of the capability of the pulse-front tilt technique to enhance the bandwidth of SHG was demonstrated in [42]. The authors used 25nm wide FF pulses at 1550nm that were to be upconverted in a 1cm long periodically poled LiNbO3 crystal (PPLN). A schematic of their configuration is shown in Fig. 10. In their case, the vector phase-matching condition required an angle of 61° between the FF beam and the grating wave vector of the poling, as shown in Fig. 10(b). The beam was focused elliptically to achieve a spot 530μm wide in the x direction in which the pulse-front tilt was introduced, and 180μm wide in the y direction. Figure 11 shows the main results. 8.3nm wide SH pulses at 775nm were generated that correspond to a pulse duration of 170fs if measured by autocorrelation, considering a Gaussian pulse. For the sake of comparison, the spectrum of the upconverted wave in collinear SHG with a crystal of identical length without GVM compensation was also measured. The measured spectrum was 0.61nm, which corresponds to 2.9ps in time. A noncollinear quasi-phase-matched configuration in combination with angular dispersion demonstrated a 14-fold increase of the spectral acceptance.

3.2b. Evolution Equations

Let us consider type-I SHG pumped by pulsed FF light. The electric field of the FF and SH waves can be written as Ei(x,y,t,z)=12Ai(x,y,t,z)exp(ikiziωit)+h.c. with i=1,2 and h.c. standing for Hermitian conjugate. The evolution of the slowly varying amplitudes Ai can be described by two coupled equations [12]

iA1z+ik1A1tk122A1t2itanρ1A1x+12k1[2A1x2+2A1y2]+iΓ12A1+K1A1*A2exp(iΔkz)=0,
iA2z+ik2A2tk222A2t2itanρ2A2x+12k2[2A2x2+2A2y2]+iΓ22A2+K2A12exp(iΔkz)=0,
where z is the longitudinal coordinate, K1=ω12χ(2)(k1c2), and K2=ω22χ(2)(k2c2), χ(2) is the second-order nonlinear coefficient, Δk=2k1k2 is the wave vector mismatch, and Γi are the absorption coefficients. The spatial walk-off parameters are given by the angles ρi, ki are the inverse group velocities, and ki are the GVD parameters.

If the input FF signal has a large beam waist and is tilted, it can be shown that the temporal evolution of the FF and SH waves can be described by [12]

iA1z+ik1,effA1tk1,eff22A1t2+iΓ12A1+K1A1*A2exp(iΔkz)=0,
iA2z+ik2,effA2tk2,eff22A2t2+iΓ22A2+K2A12exp(iΔkz)=0.
To derive Eqs. (61), we have to assume that the pulse-front tilt of the FF wave is mirrored in the generated SH wave with the same value of the tilt angle (tanΦ). The validity of this assumption can be verified by solving Eqs. (60) with an input FF beam with pulse-front tilt. In Fig. 12 we show the SH intensity as a function of the crystal length under different conditions. Two conclusion can be readily drawn. First, the reduced Eqs. (61) are a valid approximation of the more general Eqs. (60). Second, GVM severely diminishes SH efficiency conversion.

3.3. Solitons in χ(2) Media

An initially short pulse propagating along the longitudinal direction z in a medium will broaden in time owing to the chromatic dispersion (GVD) of the material. The pulse will also broaden in the transverse spatial dimensions (x,y) because of diffraction. The nonlinear interaction of the light beam with the atoms of the material modifies the main features of light propagation and can even allow, under appropriate circumstances, the temporal and spatial broadening caused by dispersion and diffraction to be overcome. The interplay among dispersion, diffraction, and nonlinearity can produce localized wave packets whose features in space or time (or both) do not change during propagation. These localized objects are called spatial or temporal solitons, respectively. When solitons present a bell-shaped intensity profile, they are called bright solitons. The localized structures that correspond to a null of the optical signal are called dark solitons. The existence of bright or dark solitons depends on the specific dispersive properties of the nonlinear media and on the type of nonlinearity that produce them.

In certain cases, when wave propagation can be described by very specific equations, such as the Korteweg–de Vries equation or the nonlinear Schrödinger equation, one can give a very precise mathematical definition of what a soliton is [43, 44]. In more real physical scenarios, many effects should be considered that make the evolution equations more complicated. Nevertheless, the equations that describe the evolution of the relevant physical parameters still predict the existence of localized waves with solitonlike behavior. To encompass such solitary waves, one only needs to generalize the definition of what a soliton is [45, 46]. In general, a soliton is a localized structure that propagates undistorted over long distances because of the balance between dispersion and diffraction on the one side and the nonlinearity of the material on the other side.

As has already been said, optical solitons can be considered in the temporal or spatial domains. Consider a light beam that propagates in a medium. If in some way the transverse spatial shape of the optical field is confined (for example in a waveguide), the degrees of freedom of interest that describe the light propagation are the direction z and the temporal variable t. In this case, the solitons are referred to as temporal. On the other hand, if the temporal variable can be considered constant, for example by using long pulses or continuous-wave (cw) lasers, the spatial variables will govern the wave propagation and the solitons are referred to as spatial.

The nonlinear effect responsible for the generation of optical solitons depends on the particular scheme being used. For example, parametric interactions in χ(2) media allow the generation of multicolor solitons formed by waves with different frequencies. The formation of these so-called quadratic solitons is mediated by the interaction of the FF and the SH waves in a SHG geometry. On the other hand, self-focusing due to cubic χ(3) Kerr nonlinearity allows the observation of temporal solitons when single-mode optical fibers are used [47], and the photorefractive effect in electro-optic materials can create a saturable nonlinear refractive index, where photorefractive solitons can be observed [48].

In this subsection of the tutorial, we will concentrate on quadratic solitons, i.e., the solitons produced in χ(2) materials in a SHG geometry. The key point to understand the enabling role of angular dispersion in the generation of solitons can be seen by defining the effective lengths that determine the relevant quantities that enter into play and by writing the equations that describe the evolution of the FF and SH waves as a function of these new variables [49]. If we consider a typical temporal width T0, beam waists W0x and W0y, and a value of the peak amplitude N, it is possible to define a normalized field amplitude, ai=Ai/N, and spatial and temporal dimensionless variables τ=tT0, s=xW0x, η=yW0x, ξ=z(2Ldis). With these definitions, and using the subindex 1 for the FF wave and the subindex 2 for the SH, the evolution equations (60) become

ia1ξ+122a1τ2+12LdisLdif[α22a1s2+2a1η2]+i2LdisLabsa1+2LdisLnla1*a2exp(i2πLdisLcohξ)=0,
ia2ξ+12LdisLdis2a2τ2+14LdisLdif[α22a2s2+2A2η2]i2LdisLgvma2τi2LdisLwA2s+i2LdisLabsa2+2LdisLnla12exp(i2πLdisLcohξ)=0.
The definition of all the characteristic lengths is given in the following list:
  • Dispersion length (FF), Ldis=T02(2|k1|)
  • Dispersion length (SH), Ldis=T02(2|k2|)
  • Diffraction length (FF), Ldif=k1W0y22
  • Diffraction length (SH), Ldif=k2W0y22=2Ldif
  • Spatial walk-off length, Lw=W0y|tanρ2tanρ1|
  • Temporal walk-off length, Lgvm=T0|k1k2|
  • Absorption length (FF), Labs=1Γ1
  • Absorption length (SH), Labs=1Γ2
  • Nonlinear length (FF), Lnl=1K1N
  • Nonlinear length (SH), Lnl=1K2N
  • Coherence length, Lcoh=π|Δk|
  • α=W0yW0x

For the sake of simplicity, notice that the diffraction length Ldif is defined with the value of the waist W0y, corresponding to the size in the dimension in which the soliton will be formed. Angular dispersion is introduced in the transverse dimension x. On the other hand, the spatial walk-off length Lw is defined with the same value of the waist, because the walk-off angle lies in the plane of angular dispersion, i.e., in the direction x. For both Ldif and Lw, the quantity of interest is W0y.

The balance between the linear effects included in Eqs. (62) and the nonlinear parametric interaction between the FF and SH waves is mediated by the ratio of the characteristic lengths corresponding to the dispersion and diffraction lengths. The observation of different types of solitons implies that the relationship between all of these characteristic lengths is tailored accordingly.

In order to observe one-dimensional and two-dimensional spatial quadratic solitons, there are three main requisites that have to be satisfied: (a) material dispersion must be small enough to render the temporal effects on the propagation negligible, (b) the crystal length must be larger than, or at least comparable with the diffraction length in order make the effects of nonlinear induced focusing observable, and (c) the spatial walk-off length must be larger than, or at most comparable with the diffraction length (2LdifLw1) so that solitons can be excited with the currently available peak powers [50, 51].

Spatial quadratic solitons have been observed in one dimension [52]. Materials with appropriate dispersive properties were chosen such that the use of cw, or even picosecond pulses, renders negligible all the temporal effects. For the case of one dimension, a 47mm planar waveguide of LiNbO3 was used to confine the beam in one transverse dimension. A cw beam with a waist in the nonconfined dimension of 70μm yields a diffraction length of Ldif=19mm and LLdif2.5. In addition, the spatial walk-off is negligible, and hence the conditions for the generation of spatial solitons are satisfied.

Spatial quadratic solitons have also been observed in two dimensions [53]. They were observed in a 1cm type-II KTP bulk sample pumped by a 15ps laser pulse at 1064nm. The beam waist of the input FF beam was W0=20μm, leading to Ldif=2mm, and the walk-off angles were ρ1=0.19° and ρ2=0.28°, giving Lw13mm. With these values it is easy to see that it is possible to observe spatial solitons, since LLdif5 and 2LdifLw0.3.

For observing temporal solitons, restrictions analogous to the ones described for spatial solitons have to be satisfied: (a) spatial effects on the beam propagation must be minimized, for example, by using beams with a large waist size, (b) the length of the nonlinear crystal must be larger than, or at least comparable with the dispersion length, (c) the temporal walk-off length must be larger than, or at most comparable with [50, 51] the dispersion length (2LdisLgvm1), and (d) the dispersion must have the appropriate sign of anomalous dispersion (i.e., k1,k2<0) to support the existence of bright solitons. The conditions to obtain spatial and temporal solitons are summarized in Table 1.

From the conditions described above for the observation of temporal solitons, it is clear that the possibility of controlling the group velocities k1 and k2 and GVDs k1 and k2 is crucial for the observation of temporal solitons. As shown in Subsection 2.2, this is precisely what angular dispersion allows us to do, and it thus gives us a tool for the generation of temporal quadratic solitons.

For a quantitative explanation of the use of angular dispersion for the generation of solitons, let us consider the generation of spatiotemporal solitons in a type-I LiIO3 crystal pumped by a T0=100fs laser pulse centered at 800nm [54]. For LiIO3, the refractive index at the phase-matching angle is n1=n2=1.87; the GVM between the fundamental and the SH waves is k2k1=566.9fsmm; the GVD for the FF wave is k1=197.3fs2mm; the GVD for the SH wave is k2=600.5fs2mm; and, finally, the FF and SH walk-off angles that lie in the x direction are ρ1=0° and ρ2=4.9°, respectively. For the generation of spatiotemporal solitons, the spatial pump profile is elliptical. In the spatial dimension in which the soliton will be formed, let us say y, the waist is chosen such that the diffraction length is smaller than the length of the crystal. For this type of experiment, typical crystal lengths are of the order of a few millimeters; so the waist in the y direction is set to be W0y20μm.

All of these values allow us to calculate the effective length parameters and determine the feasibility of observing quadratic spatiotemporal solitons. The numerical values for the effective lengths relevant to Table 1 are

  • Ldis=25.3mm
  • Ldis=8.3mm
  • Ldif=2.9mm
  • Ldif=5.8mm
  • Lgvm=0.18mm
The spatial walk-off length Lw has not been written explicitly, since the spatial walk-off lies in the x direction where the beam has a very large beam width, so the condition 2LdifLw1 is not relevant for soliton formation in the y transverse dimension.

With these numbers, it is possible to see that for a 10mm long crystal, LLdif=3.4 and condition (b) for spatial solitons is satisfied. On the other hand, in the temporal domain, LLdis=0.4, 2LdisLgvm281, and the dispersion is normal (k1,k2>0). It is not then possible to observe quadratic temporal solitons, since the conditions listed above for the observation of temporal solitons are not satisfied.

To observe temporal solitons in the discussed example, it is necessary to reverse the sign of the GVD parameters, reduce the effects of the GVM, and make the dispersion length smaller than the length of the material. As was mentioned before, and from the discussion in Subsection 2.2, one way to accomplish this is by using angular dispersion. Consider the case in which the LiIO3 crystal is flanked by two gratings of 1400 linesmm. When the pump beam impinges on the grating at an angle θ0=20° (the output angle of the first-order diffraction would be ϵ0=51°), angular dispersion is introduced in the x dimension, and the front of the pulse becomes tilted by a tilt angle Φ=60.7°. In order for Eqs. (23, 24) to be applicable, the pump beam has an elliptical spatial distribution: W0x is a few millimeters, and W0y20μm.

According to the second term of Eq. (24), the tilt angle introduces an additional anomalous dispersion of 2400fs2mm. The new effective GVM is k2,effk1,eff=55.6fsmm. And the new GVDs at the FF and SH waves are k1,eff=2213.8fs2mm and k2,eff=605fs2mm, respectively. The new effective characteristic lengths obtained by inserting the gratings are

  • Ldiseff=2.3mm(Ldis=25.3mm)
  • Ldiseff=8.3mm(Ldis=8.3mm)
  • Ldif=2.9mm(Ldif=2.9mm)
  • Ldif=5.8mm(Ldif=5.8mm)
  • Lgvmeff=1.8mm(Lgvm=0.18mm)

The values previously obtained with no pulse-front tilt are shown again for the sake of comparison. It can now be seen that the conditions for the observation of spatiotemporal solitons are satisfied: LLdis=4.3, the magnitude of the GVM between the FF and SH waves has been highly reduced so that 2LdisLgvm2.5, and the dispersion has become anomalous, thus enabling the excitation of bright solitons.

In the experimental implementations of spatiotemporal quadratic solitons, BBO and LiIO3 are some of the materials most widely used. The first temporal solitons were observed in a 7mm type-I BBO crystal, where pulses of 200fs duration at 527nm were injected [55]. Shortly after [54], spatiotemporal quadratic solitons were observed in a 1cm type-I LiIO3 crystal with highly elliptical 110fs pulses at 795nm. In later work [56], the generation of spatiotemporal solitons in BBO and LiIO3 was studied in detail.

A typical experimental setup used to generate spatiotemporal solitons with the help of angular dispersion [56] is shown in Fig. 13. In this setup, the gratings that introduce angular dispersion and modify the dispersive properties of the nonlinear crystal are clearly seen. Cylindrical lenses to control the spatial shape of the beam and guarantee the necessary ellipticity that allows the tailoring of group velocities and GVD are also depicted.

The presence of spatiotemporal solitons generated with the setup of Fig. 13 was demonstrated by measuring the temporal pulse duration and the beam waist of the soliton wave in the y transverse dimension at different propagation distances [57]. Figure 14 shows the experimental results: the dashed curves represent the temporal and spatial broadening that the optical wave would suffer if the χ2 nonlinearity were not active, i.e., working at a low pump power. On the other hand, the experimental black points were measured for a high pump power and reflect the fact that the temporal and spatial widths of the soliton wave do not change during propagation thanks to the interplay among the nonlinear effect, dispersion, and diffraction. This is precisely the fact that reveals the presence of spatiotemporal solitons.

3.4. Generation of Terahertz Waves

The generation of electromagnetic pulses at THz frequencies is of great interest in various fields. One of the main areas of application of THz waves is the probing and detection of materials, since the characteristic energies of many interactions in molecules occur in this region [58]. These waves with submillimeter wavelengths at the crossing of the far-infrared and microwaves are, in general, not that easy to produce because their wavelength is too long for optical devices and too short for electronic circuits.

One of the ways to produce THz waves is to use a special case of frequency difference generation, called “optical rectification.” In this process, a photon at frequency ω1 is absorbed by an atom of a nonlinear medium, and two new photons are generated: one at the optical frequency ω2, and another one at a much lower frequency ω1ω2 (THz). For the process of optical rectification to be efficient, the phase-matching conditions between all the interacting waves have to be satisfied.

Let us first consider a case in which the optical and the THz waves propagate in the same direction z (collinear configuration). The pump beam is an intense optical pulse with central frequency ω0, and a large beam area (plane wave). The electric field of the pump writes E(z,t)=12A(z,t)exp(ikopt0ziω0t)+h.c., where kopt0=ω0noptphc is the wavenumber of the optical pulse, noptph is the refractive index at the central frequency, and h.c. stands for Hermitian conjugate. The slowly varying amplitude A(z,t) can be written as

A(z,t)=dωa(ω)exp{i[kopt(ω0+ω)kopt0]ziωt},
where ω is the optical frequency deviation from the central frequency ω0. Notice that here we slightly modify the notation with respect to the previous sections. The optical frequency deviation is now denoted ω instead of Ω, the refractive index at the central frequency is denoted noptph instead of n0, and the optical wavenumber at ω0 is written as kopt0 instead of k0. The goal is not to confuse the reader, but to adapt our notation to the symbols generally used in the literature when dealing with combinations of optical and THz waves.

When the conditions for optical rectification are satisfied, the pump beam that interacts with a nonlinear crystal generates a nonlinear polarization at the THz frequency Ω, which is written as [59]

PTHzNL(Ω)=ϵ0χ(2)dωa(ω+Ω)a*(ω)exp{i[kopt(ω+Ω)kopt(ω)]z}.
Inspection of Eq. (64) reveals that the phase of the nonlinear polarization goes as kopt(ω+Ω)kopt(ω). So for the newly generated THz wave to build up, the phase-matching condition
kTHz(Ω)=kopt(ω+Ω)kopt(ω)
has to be fulfilled. The wavenumber of the THz wave inside the medium is
kTHz(Ω)=ΩnTHzphc,
where nTHzph is the refractive index at the THz frequency. Expanding kopt(ω+Ω) of Eq. (65) in a Taylor series to first order, we obtain [60]
nTHzph=noptgr,
where noptgr=cvg is the optical group index and vg is the optical group velocity.

In materials with a large nonlinear coefficient, such as LiNbO3, nTHzph is much larger than noptgr in the frequency band of interest [61]. For instance, for a pump at 800nm, noptgr=2.23 and nTHzph=5.16 [62]; so collinear phase matching is not possible in this case. However, if we could modify the optical group index it would then be possible to satisfy the phase-matching condition given by Eq. (67). This is precisely what can be achieved by introducing angular dispersion in the pump beam as described in Subsection 2.2: if a laser pulse acquires pulse-front tilt, the group index noptgr is changed. To clarify this idea we can rewrite Eq. (23) in terms of the group index. Noticing that noptgr=ckopt, we obtain

nopt,effgr=noptgr+tanΦtanρ.

In a more general case, THz generation occurs in a noncollinear geometry; i.e., the optical and THz waves do not propagate along the same direction. Figure 15 depicts the general configuration considered where the THz wave is generated at an angle γ¯ with respect to the pump beam direction of propagation. Different optical frequencies, ω and ω+Ω, propagate inside the nonlinear crystal in different directions, kopt(ω) and kopt(ω+Ω), determined by the amount of angular dispersion ν introduced in the pump. In the noncollinear case, the phase-matching condition must be written in the vector format [10]

kTHz(Ω)=kopt(ω+Ω)kopt(ω).
Rewriting this equation in the x and z components, we obtain [60]
nTHzphcosγ¯=noptgr,
nTHzphsinγ¯=noptphω0(ϵω)ω0.
Dividing these two expressions, we find that the angle of propagation of the excited THz wave is given by
tanγ¯=noptphnoptgrω0(ϵω)ω0.

Equation (71) is the key expression that highlights the role of angular dispersion in the generation of THz waves by means of optical rectification. By changing (ϵω)ω0 with the help of dispersive elements, we can tune the noncollinear angle γ¯ and satisfy the phase-matching conditions. Furthermore, the introduction of different amounts of angular dispersion allows the tuning of the frequency of the generated THz wave [62].

In a dispersive medium, the tilt angle Φ is given by Eq. (8), and the relationship between the angles ν and Φ is given by Eq. (14), which if rewritten in terms of the refractive index and the group index of the optical wave yield [9]

tanν=tanΦnoptgr.
Using Eqs. (8, 72), we see that γ¯=ν; i.e., the angle of propagation of the THz wave is perpendicular to the pulse front that is tilted at an angle ν in the xz plane (dashed lines in Fig. 15).

The efficiency of the generation of THz waves by means of the process of optical rectification of femtosecond pulses depends on the fulfillment of the phase-matching condition between the optical and THz waves. By using optical pulses with pulse-front tilt, not only do we make it possible to comply with the phase-matching condition, but also we can tune the THz frequency at which this happens, which adds tunability to the scheme.

The full potential of using tilted pulses to tune the frequency of the generated THz waves was demonstrated in [62]. This is shown in Figs. 16, 17. The THz output was tuned between 1 and 4.4THz in LiNbO3 at a temperature of 10K by changing the tilt angle ν between 59° and 64°. Figure 16 shows the spectra of the THz waves when the tilt angle ν is changed. Figure 17 plots the change of the central frequency of the THz waves as a function of the tilt angle. The measured energy is also shown. The capability of generating even higher frequencies would require the use of pulses with pulse durations below 100fs in less absorbing materials.

We remark that the use of pulse-front techniques allows the use of pumps in different frequency bands that may be more optimal. For instance, pulse energies of up to 100nJ with a spectral bandwidth of up to 2.5THz were obtained by optical rectification of 1030nm laser pulses with 400μJ energy and 300fs pulse duration [63], achieving a conversion efficiency of 2.5×104, an order of magnitude higher than the one measured when using other materials in an optimized geometry.

4. Angular Dispersion as an Enabling Tool in Quantum Optics

Until now, we have seen the role of angular dispersion in applications in the field of nonlinear optics. We may regard these applications as belonging to classical optics in the sense that we are dealing with large amounts of photons whose properties can accurately be described by the classical Maxwell equations. In this section, on the other hand, we will see that angular dispersion is also useful in quantum optics when, for example, pairs of photons are created.

We will focus on the generation of pairs of photons by means of the process of spontaneous parametric downconversion (SPDC). SPDC is a nonlinear optical process in which an intense pump impinges on a nonlinear material and occasionally creates a pair of photons of lower frequencies. The correlations of the photon pairs produced by SPDC are of particular interest from both the fundamental and the practical points of view. Fundamentally, SPDC photon pairs are present at the core of many experiments to test the validity of the foundations of quantum mechanics. And, practically, new technologies based on their correlations promise improvements over their classical counterparts, among them quantum communications and information processing or clock synchronization.

In this tutorial, we will concentrate on the frequency properties of the SPDC photons, namely, on the bandwidth and the type of frequency correlations between the two photons. The appropriate frequency content of paired photons depends on the particular application under consideration [64, 65, 66, 67]. Some applications require frequency-correlated photons, some require frequency-anticorrelated photons, and some require frequency-uncorrelated photons. For this reason in recent years various methods to tailor at will the spectral properties of paired photons have been developed [68, 69, 70, 71].

In this section, we will see how the use of light beams with pulse-front tilt in SPDC allows us to control the bandwidth and the type of frequency correlation. These two frequency properties can be tuned by a proper tailoring of the group velocities and GVD parameters of all the waves that interact in the nonlinear process [72, 73]. As we have seen throughout the tutorial, this tailoring is precisely what angular dispersion enables. Indeed, it will be the pulse-front tilt angle that will play the role of a control parameter to tune, at will, the bandwidth and the type of frequency correlations of the SPDC photons. Remarkably, when using pulse-front tilt, there is no need for any particular engineering of the SPDC source. Moreover, the method is independent of the material and can be used in any frequency range where other methods do not work.

As we will see in what follows, angular dispersion allows us to increase the bandwidth of the joint spectrum of paired photons, an important point for the generation of very narrow temporal biphotons. In addition, the use of pump beams with pulse-front tilt makes possible the generation of frequency-correlated, frequency-anticorrelated, and even uncorrelated photon pairs. The latter case offers a very attractive applicability of angular dispersion for the generation of heralded indistinguishable and pure single photons with a tunable frequency bandwidth [74].

4.1. Angular Dispersion in Spontaneous Parametric Downconversion

Let us consider the generation of SPDC photons in a collinear configuration depicted in Fig. 18 when the two downconverted photons propagate in the same direction. Unlike the typical SPDC setups, the nonlinear medium is placed between two diffraction gratings that introduce angular dispersion. It can be shown that after the two gratings, the quantum state of the downconverted photons can be written as [75]

|Ψ=dΩsdΩidqsdqiΨ(Ωs,Ωi,qs,qi)|ωs0+Ωs,qss|ωi0+Ωi,qii,
where the subindices s,i denote the signal and idler photons, respectively, Ωj is the frequency detuning from the central frequency ωj0, qj are the transverse wavenumbers, and
Ψ(Ωs,Ωi,qs,qi)=Ep(Ωs+Ωi,qs+qi)sinc(ΔkL2)exp(iskL2)
is the joint spectral amplitude that contains all the information about the bandwidth and type of frequency correlations of the two-photon state, also called the biphoton. From Eq. (74), we can see that the frequency content of the downconverted light is determined by the length of the crystal L, by the spectral characteristics of the pump beam Ep(ωp0+Ωp), and by the dispersive properties of the nonlinear material expressed by the phase-mismatch term along the longitudinal direction, Δk=kpkski, with kj=[(nj(ωj)ωjc)2|qj|2]12, where nj is the refractive index and sk=kp+ks+ki.

The role of angular dispersion is to modify Ψ(Ωs,Ωi,qs,qi) due to the dependence of Δk on q. Inspection of Eq. (20) shows us that angular dispersion can thus modify the frequency shape of the phase-matching function Δk. After the second grating, downconverted photons are detected by using strong spatial filters, e.g., single-mode fibers, so that qs,qi0. To get a further physical insight, let us expand Δk up to second order. Considering walk-off for the pump, the signal, and the idler,

Δk=(kp,effks,eff)Ωs+(kp,effki,eff)Ωi12ks,effΩs212ki,effΩi212kp,eff(Ωs+Ωi)2,
where the effective inverse group velocity kj,eff and effective GVD kj,eff are given by Eqs. (23, 24): kj,eff=kj+tanΦtanρjc and kj,eff=kj[tanΦc]2kj. We assume that there is perfect phase matching for Ωs=Ωi=0. From these expressions, it is clear that, as described in Subsection 2.2, angular dispersion allows us to modify the effective inverse group velocity and effective GVD. Employing pump pulses tilted by an angle Φ in SPDC configurations, it is possible to control the frequency properties of paired photons.

4.2. Tunable Control of Frequency Correlations of Paired Photons

To see how angular dispersion can be used to control the type of frequency correlation of SPDC photons, let us consider the first-order terms of Eq. (75). One obtains frequency-anticorrelated photons (Ωs=Ωi) when ks,eff=ki,eff and frequency-correlated photons (Ωs=Ωi) when kp,eff=(ks,eff+ki,eff)2. Frequency-uncorrelated photons are obtained if the effective group velocity of the pump is equal to that of the signal or the idler, kp,eff=ks,eff and/or kp,eff=ki,eff.

Figure 18 shows the experimental arrangement used to measure the joint spectrum S(ωs,ωi)=|Ψ(Ωs,Ωi)|2 of the entangled paired photons. Before being detected, each of the photons forming a pair passes through its respective monochromator, which is scanned to measure the joint spectrum S(ωs,ωi). Figure 19 shows the experimental results that demonstrate the feasibility to fully control the frequency correlations in SPDC via angular dispersion. In particular, the measurements were performed by using a 3.5mm type-II BBO crystal cut for collinear degenerate phase matching.

In Fig. 19, different types of frequency correlation are observed by varying the tilt angle Φ. The first row of Fig. 19 shows the case with no tilt. As expected for a pulsed pump and type-II phase-matching, the spectra of the signal and idler photons are different, one being narrower than the other, which is a consequence of their different group velocities. The following rows correspond to different values of the tilt angle. The second and fourth rows of Fig. 19 depict the cases of highly frequency-anticorrelated and highly frequency-correlated photons. The third row shows the interesting case of frequency-uncorrelated pairs.

The frequency uncorrelation observed in the third row of Fig. 19 is indeed a signature of the presence of a separable quantum state. To demonstrate full separability of the two-photon state, it is also required that there be no phase entanglement [76]. Theoretically, we can calculate the Schmidt decomposition of the state given by Eq. (73) when the separability condition of the group velocities is fulfilled. In that case, the entropy of entanglement is nearly 1, and the Schmidt decomposition contains only one mode, revealing the separability of the quantum state [68]. In this way, the scheme offers a possibility to generate paired photons in a separable state, a quantum state so desired in quantum information processing applications. Frequency-uncorrelated photons can well serve as a source of heralded single photons: the detection of one of the photons heralds the presence of its twin photon in the setup without in any way changing its state. Experimentally, a tomographic analysis or a four-photon experiment such as the one described in [71] would be needed to fully demonstrate the separability and therefore the purity of the generated single photons.

The frequency correlations of entangled paired photons generated in the process of spontaneous parametric downconversion can be tuned independently of the frequency band and of the nonlinear crystal used. The photons can exhibit frequency anticorrelation, frequency correlation, and even frequency uncorrelation only through control of the amount of angular dispersion.

4.3. Controlling Bandwidth of Paired Photons

As we have mentioned before, Eq. (74) tells us that the bandwidth of the downconverted light is determined by the length and the dispersive properties of the nonlinear material, by the geometry of the SPDC configuration (collinear or noncollinear, type I or type II) and by the spectral characteristics of the pump beam. In the following, we will see that the use of SPDC with a pump beam with angular dispersion allows us to control the SPDC bandwidth as well [77].

To further clarify the ideas, we consider the case of a narrowband pump, for example, a picosecond pulsed laser. In this case, energy conservation dictates ΩsΩi, and Eq. (75) reduces to

Δk(ki,effks,eff)Ωs12(ks,eff+ki,eff)Ωs2.
From this expression, it is easy to see that if we introduce angular dispersion with a tilt angle Φ such that ki,effks,eff=0, the bandwidth will increase because the first nonzero terms that contribute to Δk are terms of second order or higher. For example, in a type-II process the bandwidth of generated photons is inversely proportional to the length of the nonlinear material L [72]. If the first-order terms of Δk are removed by using angular dispersion, the dependence of the bandwidth on the length will go as 1L. In addition, the higher dispersion terms are much weaker, which further broadens the spectrum. In a type-I SPDC process, if the tilt Φ is such that ks,eff=ki,eff, the first nonzero terms in Δk are of fourth order Ωs4, and the dependence of the bandwidth on the length of the crystal goes as 1L14.

The values of the tilt angle that maximize the bandwidth for type-II and type-I processes are [77]

ΦIImax=tan1{c(kiks)tanρstanρi},
ΦImax=tan1c2ksks0,
respectively, where kj0=kj(ωj0) in the nonlinear medium.

The effect of introducing angular dispersion to increase the bandwidth of the SPDC photons was demonstrated experimentally by using the setup of Fig. 18 with a 2mm BBO crystal cut for degenerate type-II collinear phase matching. Figure 20(a) depicts the joint spectrum for the case without angular dispersion, and Fig. 20(b) shows the joint spectrum when a tilt Φ=ΦIImax=38° was introduced.

To get a quantitative value of the increase in the bandwidth, let us examine Fig. 21. The solid curves represent the theoretical predictions, and the dots are the experimental results. Figures 21(a), 21(b) correspond to the spectra of single counts of the signal and idler photons. The curve on the left corresponds to no tilt, Φ=0°, and the one on the right to the case of Φ=38°. For the case with no gratings, a FWHM bandwidth of Δλs5.2nm is obtained, while for the case with gratings, the FWHM bandwidth is Δλs37nm, leading to a sevenfold increase of the bandwidth. Figures 21(c), 21(d) depict the coincidence counts. The widths along the antidiagonal line (at 45°) of the joint spectra, Figs. 20(a), 20(b), are plotted. The coincidence width of 7.5nm broadens to 52nm.

The effect of spectrum broadening can be employed to generate a very short temporal biphoton that is given by the Fourier transform of the joint spectral amplitude [72]. It was shown in [77] that when the pulse-front tilt is used, the spectral phase profile remains smooth and very flat, which translates into ultrashort nearly transform-limited biphotons with a temporal correlation of a few femtoseconds. The detection of one of the photons of the pair determines the detection of the other photon located in a distant place within a time window given by the biphoton’s temporal width. This phenomenon could expediently be used, e.g., for clock synchronization. The pulse-front-tilt technique contrasts with other methods in which the increase in bandwidth is not directly accompanied by a decrease of the correlation time. That can be caused by a particular shape of the spectral phase or when the phase relationship between individual frequencies, and thus coherence, is lost. This is, for example, the case of white-light continuum generated by Kerr self-phase modulation.

Up to this point we have considered the effects of angular dispersion when a collinear geometry is used. However, it is worth mentioning that the combination of the pulse-tilt techniques described above with the use of noncollinear geometries further expands the possibilities to control the joint spectrum of paired photons [74]. In noncollinear geometries it is possible to map the spatial characteristics of the pump beam into the spectra (spatial-to-spectral mapping) [78, 79], providing another way to manipulate the joint spectral amplitude of the biphoton.

The bandwidth of paired photons generated in the process of spontaneous parametric downconversion can be enhanced or reduced, independently of the frequency band and of the nonlinear crystal used. It is possible to generate spectral bandwidths of up to hundreds of nanometers in the visible optical range that translate into temporal correlations between the two photons of just a few femtoseconds.

5. Conclusions

The angular dispersion of light is an old physical phenomenon that was already discussed at the beginning of the 18th century by Isaac Newton in his book Opticks [1], where he describes how white light decomposes into colors and diverges after passing through a prism.

In this tutorial we have shown how, with the appearance of the laser, angular dispersion has become an important enabling tool in different areas of nonlinear and quantum optics. The key tool is the possibility to modify the dispersive properties of materials by using light pulses with suitable amounts of angular dispersion. The use of these pulses in many applications has been described with such an unifying view.

Most times, this common perspective is absent in scientific papers and technical reports, or at least is not clearly seen, because of the use of different notation in each field or because emphasis is put on diverse aspects each time.

Section 2 of this tutorial hopefully offers such a view, and each application considered in Sections 3, 4 is analyzed under the general unifying framework developed in Section 2. The first two applications, pulse compression and CPA, described in Subsection 3.1, are nowadays routinely used in commercial systems for compressing and stretching optical pulses. The next application is achromatic phase matching, considered in Subsection 3.2, which enables us to enhance the capability of frequency doublers for efficiently doubling ultrashort pulses.

The excitation of quadratic temporal solitons is described in Subsection 3.3, where it is shown that the introduction of angular dispersion permits the observation of temporal solitons. In Subsection 3.4, we have described how the use of pulses with pulse-front tilt allows us to satisfy the condition of phase matching, a requisite not easily achievable, but notwithstanding necessary, for the efficient generation of THz waves in the process of optical rectification of femtosecond laser pulses. Even more, the method allows us to tune the frequency of the generated THz wave, allowing the implementation of tunable generators of THz waves. Finally, the generation of entangled paired photons with tunable bandwidth and frequency correlations is analyzed in Section 4.

Subsection 3.3 and Section 4 are two outstanding cases that exemplify the role of light beams with angular dispersion. By altering the unfavorable conditions offered by most natural materials, the use of angular dispersion makes it possible to observe physical effects that would not be possible otherwise.

The necessary conditions for the observation of temporal solitons in quadratic nonlinear media are not met in commonly used nonlinear crystals. This is also the case for the observation of frequency-entangled photons that show frequency correlation or frequency uncorrelation. In both cases, pulses with angular dispersion allow us to modify the dispersive properties of media, effectively engineering new materials that meet the necessary requirements in terms of new effective group velocity and GVD parameters.

Summarizing, light beams with angular dispersion, or pulse-front tilt, allow us to perform tasks in nonlinear and quantum optics not possible otherwise, highlighting the role of angular dispersion as enabling tool.

Acknowledgments

We thank F. Wise, A. Schober, and J. Hebling for providing some of the figures that appear in this tutorial, helping to greatly improve its clarity. We give special thanks to L. Torner and S. Carrasco for long and helpful discussions during many years about the role of angular dispersion in nonlinear optics, especially concerning the generation of quadratic spatiotemporal solitons and the control of the frequency correlations of entangled photons. We thank X. Liu, F. Wise, and P. Di Trapani for occasional, but illuminating, discussions about solitons. We thank P. Loza and D. Artigas for helpful discussions concerning the process of SHG. Finally, we also thank R. Trebino and S. Akturk for discussions about the relationship between angular dispersion and pulse-front tilt. This work was supported by the European Commission (Qubit Applications, contract 015848), by the Government of Spain (Consolider Ingenio CSD2006-00019, FIS2007-60179), and was supported in part by FONCICYT project 94142.

Tables and Figures

Tables Icon

Table 1. Conditions to Obtain Spatial, Temporal, and Spatiotemporal Solitons

 figure: Fig. 1

Fig. 1 Decomposition of a white-light beam into different colors due to the angular dispersion introduced by a prism, as originally depicted by I. Newton in Opticks[1].

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 Schematic of a grating surface and definition of angles: (+) refers to positives angles, and (−) to negative ones; m>0 corresponds to positive diffraction orders, and m<0 to negative diffraction orders.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Tilting of the front of the pulse. After the grating, the front of the pulse is no longer perpendicular to the direction of propagation.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 The line of the loci of peak intensities in the xct plane is tilted by an angle Φ.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 The material whose dispersive properties are to be tailored is located between two gratings Gr1 and Gr2. Light enters the dispersive medium perpendicularly to its input face.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Schematic of a prism and definition of angles.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 General scheme for pulse compression.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 (a) General scheme for CPA and compression. (b) Device that can introduce positive or negative dispersion, depending on the length d. f is the focal length of the two lenses.

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 Detailed evolution of the SH pulse (a) with and (b) without GVM compensation. With GVM compensation with the help of pulse-front tilt, the SH quickly and efficiently builds up. Without GVM compensation, the nonlinear process is very inefficient, and the SH hardly appears. In addition, the lack of broadband phase-matching results in backconversion. Conditions: input FF peak intensity, 10MWcm2; FF input beam width, 3mm; FF input pulse duration, 100fs; wavelength of the fundamental wave,1.6μm; length of the NPP crystal [N-(4-nitrophenyl)-L-prolinol], 3mm. Figure courtesy of J. P. Torres [12].

Download Full Size | PDF

 figure: Fig. 10

Fig. 10 Schematic diagram of a typical SHG configuration that uses pulses with pulse-front tilt. (a) The FF beam acquires pulse-front tilt and is focused into a PPLN crystal with a lens. A second grating is used to remove the angular dispersion introduced by the first grating. (b) Close-up view that illustrates the tilted quasi-phase-matching grating used. Figure courtesy of A. Schober [42].

Download Full Size | PDF

 figure: Fig. 11

Fig. 11 SHG obtained with angular dispersion in PPLN. (a) Measured autocorrelation and (b) spectrum with angular dispersion. For the sake of comparison, (c) and (d) show the measured autocorrelation and spectrum when a crystal of identical length in an collinear configuration, with no pulse-front tilt, is used. Figure courtesy of A. Schober [42].

Download Full Size | PDF

 figure: Fig. 12

Fig. 12 Simulated evolution of the peak intensity of the SH beam as a function of the crystal length. Solid curve, evolution according to Eqs. (61) with GVM compensation and no loss; dotted-dashed curve, evolution according to Eqs. (60) with GVM compensation and no loss; dashed curve, evolution according to Eqs. (61) with GVM compensation and loss; dotted curve, evolution according to Eqs. (61) with no GVM compensation. Inset, SH output pulse. Conditions: input FF peak intensity, 10MWcm2; FF input beam width, 3mm; FF input pulse duration, 100fs. Figure courtesy of J. P. Torres [12].

Download Full Size | PDF

 figure: Fig. 13

Fig. 13 Experimental setup to observe quadratic temporal solitons in BBO. (a) Schematic of the experiment. (b), (c) Highly elliptical spatial profiles of the pump beam. A cylindrical lens focuses the beam in the y direction. Figure courtesy of F. Wise [56]. © 2004 by the American Physical Society.

Download Full Size | PDF

 figure: Fig. 14

Fig. 14 Experimental (a) temporal and (b) spatial widths of the solitons during propagation. The dashed curves represent the temporal and spatial widths of the wave if the propagation were dictated only by dispersion and diffraction that produce temporal and spatial broadening. The experimental black diamonds confirm that the temporal and spatial widths of the generated soliton remained constant. The insets show the temporal and spatial profiles at some selected distances. The peak intensity is 8GWcm2, and Δk=60π25mm. Figure courtesy of F. Wise [57].

Download Full Size | PDF

 figure: Fig. 15

Fig. 15 Schematic of the configuration for noncollinear phase matching of optical and THz waves with pulse-front tilt.

Download Full Size | PDF

 figure: Fig. 16

Fig. 16 Spectra of the THz pulses measured for different tilt angles ν. The maxima of the spectra are normalized. Figure courtesy of J. Hebling [62]. © 2004 by Springer.

Download Full Size | PDF

 figure: Fig. 17

Fig. 17 Dependence of the energy and the frequency of the THz pulses on the tilt angle ν. The solid curves are guides to the eye. Figure courtesy of J. Hebling [62]. © 2004 by Springer.

Download Full Size | PDF

 figure: Fig. 18

Fig. 18 Experimental setup used to demonstrate the control of frequency correlation and the bandwidth enhancement in SPDC by means of angular dispersion. G denotes gratings; PBS, polarization beam splitter; Mono, monochromators; D, single-photon counting modules; M, mirrors; &, coincidence electronics.

Download Full Size | PDF

 figure: Fig. 19

Fig. 19 Shape S(ωs,ωi) of the frequency correlations measured experimentally (left) and predicted theoretically (right). (a), (b) no tilt, Φ=0°; (c), (d) anticorrelated photons, Φ=38°; (e), (f) uncorrelated photons, Φ=20°; (g), (h) correlated photons, Φ=52°. Pump-beam bandwidth, Δλp=2nm; nonlinear crystal length, L=3.5nm. Figure courtesy of M. Hendrych [77]. © 2009 by the American Physical Society.

Download Full Size | PDF

 figure: Fig. 20

Fig. 20 Measured joint spectral density S(ωs,ωi) for a 2mm type-II BBO crystal: (a) tilt angle Φ=0° and (b) tilt angle Φ=ΦIImax=38°. The joint spectrum broadens sevenfold. Figure courtesy of M. Hendrych [77]. © 2009 by the American Physical Society.

Download Full Size | PDF

 figure: Fig. 21

Fig. 21 (a) Signal single counts for Φ=0°; Δλs=5.2nm. (b) Signal single counts for Φ=ΦIImax=38°; Δλs=37nm. (c) Coincidences along the antidiagonal for Φ=0°; ΔΛ=7.5nm. (d) Coincidences along the antidiagonal for Φ=ΦIImax=38°; ΔΛ=52nm. Solid curves represent the theoretical prediction; squares are the experimental data.

Download Full Size | PDF

1. I. Newton, Opticks: or a Treatise of the Reflections, Refractions, Inflections and Colours of Light, 4th ed. (William Innys, 1730), http://books.google.com/books?id=XXu4AkRVBBoC.

2. Y. R. Shen, Principles of Nonlinear Optics (Wiley-Interscience, 2002).

3. P. N. Butcher and D. Cotter, Elements of Nonlinear Optics (Cambridge Univ. Press, 1991).

4. G. P. Agrawal, Nonlinear Fiber Optics (Academic, 2001).

5. Richardson Gratings, Diffraction Grating Handbook (Newport, 2000), http://gratings.newport.com/information/handbook/handbook.asp.

6. H. Haus, Waves and Fields in Optoelectronics (Prentice-Hall, 1983).

7. O. E. Martínez, “Pulse distortions in tilted pulse schemes for ultrashort pulses,” Opt. Commun. 59, 229–232 (1986). [CrossRef]  

8. O. E. Martínez, “Grating and prism compressors in the case of finite beam size,” J. Opt. Soc. Am. B 3, 929–934 (1986). [CrossRef]  

9. J. Hebling, “Derivation of the pulse front tilt caused by angular dispersion,” Opt. Quantum Electron. 28, 1759–1763 (1996). [CrossRef]  

10. J. Hebling, K. L. Yeh, M. C. Hoffmann, B. Bartal, and K. A. Nelson, “Generation of high-power terahertz pulses by tilted-pulse-front excitation and their application possibilities,” J. Opt. Soc. Am. B 25, B6–B19 (2008). [CrossRef]  

11. V. G. Dimitriev, G. G. Gurzadyan, and D. N. Nikogosyan, Handbook of Nonlinear Crystals (Springer, 1997). [CrossRef]  

12. J. P. Torres, S. Carrasco, L. Torner, and E. W. Van Stryland, “Frequency doubling of femtosecond pulses in walk-off-compensated N-(4-nitrophenyl)-L-prolinol,” Opt. Lett. 25, 1735–1737 (2000). [CrossRef]  

13. R. Danielius, A. Piskarskas, P. Di Trapani, A. Andreoni, C. Solcia, and P. Foggi, “Matching of group velocities by spatial walk-off in collinear three-wave interaction with tilted pulses,” Opt. Lett. 21, 973–975 (1996). [CrossRef]   [PubMed]  

14. R. Danielius, A. Piskarskas, P. Di Trapani, A. Andreoni, C. Solcia, and P. Foggi, “A collinearly phase-matched parametric generator/amplifier of visible femtosecond pulses,” IEEE J. Quantum Electron. 34, 459–464 (1998). [CrossRef]  

15. E. B. Treacy, “Optical pulse compression with diffraction gratings,” IEEE J. Quantum Electron. 5, 454–458 (1969). [CrossRef]  

16. S. Szatmari, P. Simon, and M. Feuerhake, “Group-velocity-dispersion-compensated propagation of short pulses in dispersive media,” Opt. Lett. 21, 1156–1158 (1996). [CrossRef]   [PubMed]  

17. O. E. Martínez, J. P. Gordon, and R. L. Fork, “Negative group-velocity dispersion using refraction,” J. Opt. Soc. Am. A 1, 1003–1006 (1984). [CrossRef]  

18. R. L. Fork, O. E. Martínez, and J. P. Gordon, “Negative dispersion using pairs of prisms,” Opt. Lett. 9, 150–152 (1984). [CrossRef]   [PubMed]  

19. J. C. Diels and W. Rudolph, Ultrashort Laser Pulse Phenomena (Academic, 1995).

20. J. Hebling and Z. Márton, “Theory of spectroscopic devices,” J. Opt. Soc. Am. B 23, 966–972 (2006). [CrossRef]  

21. J. P. Gordon and R. L. Fork, “Optical resonator with negative dispersion,” Opt. Lett. 9, 153–155 (1984). [CrossRef]   [PubMed]  

22. O. E. Martínez, “Hybrid prism-grating ultrashort pulse compressor,” Opt. Commun. 83, 117–122 (1986). [CrossRef]  

23. S. Akturk, X. Gu, E. Zeek, and R. Trebino, “Pulse-front tilt caused by spatial and temporal chirp,” Opt. Express 12, 4399–4410 (2004). [CrossRef]   [PubMed]  

24. X. Gu, S. Akturk, and R. Trebino, “Spatial chirp in ultrafast optics,” Opt. Commun. 242, 599–604 (2004). [CrossRef]  

25. J. A. Fullop and J. Hebling, “Applications of tilted-pulse-front excitation,” in Recent Optical and Photonic Technologies, K. Y. Kim, ed. (Inteh, 2010), chap. 11, http://sciyo.com/books/show/title/recent-optical-and-photonic-technologies?PHPSESSID=0imuiv8409jhevei6sb4toh0d5

26. C. V. Shank, R. L. Fork, R. Yen, R. H. Stolen, and W. J. Tomlinson, “Compression of femtosecond optical pulses,” Appl. Phys. Lett. 40, 761 (1982). [CrossRef]  

27. R. L. Fork, C. H. Brito Cruz, P. C. Becker, and C. V. Shank, “Compression of optical pulses to six femtoseconds by using cubic phase compensation of femtosecond optical pulses,” Opt. Lett. 12, 483–485 (1987). [CrossRef]   [PubMed]  

28. D. Strickland and G. Mourou, “Compression of amplified chirped optical pulses,” Opt. Commun. 56, 219–221 (1985). [CrossRef]  

29. B. E. A. Saleh and M. C. Teich, Fundamentals of Photonics (Wiley, 1991). [CrossRef]  

30. Y. Zaouter, J. Boullet, E. Mottay, and E. Cormier, “Transform-limited 100μJ340MW pulses from a nonlinear-fiber chirped-pulse amplifier using a mismatched grating stretcher-compressor,” Opt. Lett. 33, 1527–1529 (2008). [CrossRef]   [PubMed]  

31. A. Yariv, Optical Electronics in Modern Communications (Oxford Univ. Press, 1997).

32. R. W. Boyd, Nonlinear Optics (Academic, 2008).

33. A. Yariv, Quantum Electronics (Wiley, 1975).

34. P. W. Milonni and J. H. Eberly, Lasers (Academic, 1988).

35. G. Szabó and Zs. Bor, “Frequency conversion of ultrashort pulses,” Appl. Phys. B 58, 237–241 (1994). [CrossRef]  

36. O. E. Martínez, “Achromatic phase matching for second harmonic generation of femtosecond pulses,” IEEE J. Quantum Electron. 25, 2464–2468 (1989). [CrossRef]  

37. V. D. Volosov, S. G. Karpenko, N. E. Kornienko, and V. L. Strizhevskii, “Method for compensating the phase-matching dispersion in nonlinear optics,” Sov. J. Quantum Electron. 4, 1090–1098 (1975). [CrossRef]  

38. T. R. Zhang, H. R. Choo, and M. C. Downer, “Phase and group velocity matching for second harmonic generation of femtosecond pulses,” Appl. Opt. 29, 3927–3933 (1990). [CrossRef]   [PubMed]  

39. A. Dubietis, G. Valiulis, G. Tamosauskas, R. Danielius, and A. Piskarskas, “Nonlinear second-harmonic pulse compression with tilted pulses,” Opt. Lett. 22, 1071–1073 (1997). [CrossRef]   [PubMed]  

40. G. Szabó and Zs. Bor, “Broadband frequency doubler for femtosecond pulses,” Appl. Phys. B 50, 51–54 (1990). [CrossRef]  

41. B. Richman, S. Bisson, R. Trebino, E. Sidick, and A. Jacobson, “Efficient broadband second-harmonic generation by dispersive achromatic nonlinear conversion using only prisms,” Opt. Lett. 23, 497–499 (1998). [CrossRef]  

42. A. Schober, M. Charbonneau-Lefort, and M. Fejer, “Broadband quasi-phase-matched second-harmonic generation of ultrashort optical pulses with spectral angular dispersion,” J. Opt. Soc. Am. B 22, 1699–1713 (2005). [CrossRef]  

43. M. J. Ablowitz, Solitons and the Inverse Scattering Transform (Society for Applied and Industrial Mathematics (SIAM), 1981). [CrossRef]  

44. P. G. Drazin and R. S. Johnson, Solitons: an Introduction (Cambridge Univ. Press, 1989). [CrossRef]  

45. N. N. Akhmediev and A. Ankiewicz, Solitons: Nonlinear Pulses and Beams (Chapman & Hall, 1997).

46. A. Hasewaga and Y. Kodama, Solitons in Optical Communications (Clarendon, 1995).

47. L. F. Mollenauer, R. H. Stolen, and J. P. Gordon, “Experimental observation of picosecond pulse narrowing and solitons in optical fibers,” Phys. Rev. Lett. 45, 1095–1098 (1980). [CrossRef]  

48. M. Segev, B. Crosignani, A. Yariv, and B. Fischer, “Spatial solitons in photorefractive media,” Phys. Rev. Lett. 68, 923–926 (1992). [CrossRef]   [PubMed]  

49. C. R. Menyuk, R. Schieck, and L. Torner, “Solitary waves due to χ(2):χ(2) cascading,” J. Opt. Soc. Am. B 11, 2434–2443 (1992). [CrossRef]  

50. L. Torner, D. Mazilu, and D. Mihalache, “Walking solitons in quadratic nonlinear media,” Phys. Rev. Lett. 77, 2455–2458 (1996). [CrossRef]   [PubMed]  

51. S. Carrasco, J. P. Torres, L. Torner, and F. Wise, “Walk-off acceptance for quadratic soliton generation,” Opt. Commun. 191, 363–370 (2001). [CrossRef]  

52. R. Schieck, Y. Baek, and G. I. Stegeman, “One-dimensional spatial solitary waves due to cascaded second-order nonlinearities in planar waveguides,” Phys. Rev. E 53, 1138–1141 (1996). [CrossRef]  

53. W. E. Torruellas, Z. Wang, D. J. Hagan, E. W. Van Stryland, G. I. Stegeman, L. Torner, and C. R. Menyuk, “Observation of two-dimensional spatial solitary waves in a quadratic medium,” Phys. Rev. Lett. 74, 5036–5039 (1995). [CrossRef]   [PubMed]  

54. X. Liu, L. J. Qian, and F. W. Wise, “Generation of optical spatiotemporal solitons,” Phys. Rev. Lett. 82, 4631–4634 (1999). [CrossRef]  

55. P. Di Trapani, D. Caironi, G. Valiulis, A. Dubietis, R. Danielius, and A. Piskarskas, “Observation of temporal solitons in second harmonic generation with tilted pulses,” Phys. Rev. Lett. 81, 570–573 (1998). [CrossRef]  

56. X. Liu, K. Beckwitt, and F. W. Wise, “Two-dimensional optical spatiotemporal solitons in quadratic media,” Phys. Rev. E 62, 1328–1340 (2000). [CrossRef]  

57. F. Wise and P. Di Trapani, “The hunt for light bullets–spatiotemporal solitons,” Opt. Photon. News 13(2), 29–32 (February 2002).

58. J. Hebling, K.-L. Yeh, K. A. Nelson, and M. C. Hoffmann, “High-power THz generation, THz nonlinear optics, and THz nonlinear spectroscopy,” IEEE J. Sel. Top. Quantum Electron. 14, 345–353 (2008). [CrossRef]  

59. K. L. Vodopyanov, “Optical generation of narrow-band terahertz packets in periodically-inverted electro-optics crystals: conversion efficiency and optimal laser pulse format,” Opt. Express 14, 2263–2276 (2006). [CrossRef]   [PubMed]  

60. J. Hebling, G. Almasi, I. Z. Kozma, and J. Kuhl, “Velocity matching by pulse front tilt for large area THz-pulse generation,” Opt. Express 10, 1161–1166 (2002). [CrossRef]   [PubMed]  

61. A. G. Stepanov, J. Kuhl, I. Z. Kozma, E. Riedle, G. Almasi, and J. Hebling, “Scaling up the energy of THz pulses created by optical rectification,” Opt. Express 13, 5762–5768 (2005). [CrossRef]   [PubMed]  

62. J. Hebling, A. G. Stepanov, G. Almasi, B. Bartal, and J. Kuhn, “Tunable THz pulse generation by optical rectification of ultrashort pulses with tilted pulses,” Appl. Phys. B 78, 593–599 (2004). [CrossRef]  

63. M. C. Hoffmann, K. L. Yeh, J. Hebling, and K. A. Nelson, “Efficient terahertz generation by optical rectification at 1035nm,” Opt. Express 15, 11706–11713 (2007). [CrossRef]   [PubMed]  

64. V. Giovannetti, S. Lloyd, and L. Maccone, Science 306, 1330–1336 (2004). [CrossRef]   [PubMed]  

65. V. Giovannetti, L. Maccone, and S. Lloyd, “Quantum-enhanced positioning and clock synchronization,” Nature 412, 417–419 (2001). [CrossRef]   [PubMed]  

66. A. Valencia, G. Scarcelli, and Y. H. Shih, “Distant clock synchronization using entangled photon pairs,” Appl. Phys. Lett. 85, 2655–2657 (2004). [CrossRef]  

67. T. Aichele, A. I. Lvovsky, and S. Schiller, “Optical mode characterization of single photons prepared by means of conditional measurements on a biphoton state,” Eur. Phys. J. D 18, 237–245 (2002). [CrossRef]  

68. J. P. Torres, F. Macià, S. Carrasco, and L. Torner, “Engineering the frequency correlations of entangled two-photon states by achromatic phase matching,” Opt. Lett. 30, 314–316 (2005). [CrossRef]   [PubMed]  

69. O. Kuzucu, M. Fiorentino, M. A. Albota, F. N. C. Wong, and F. X. Kärtner, “Two-photon coincident-frequency entanglement via extended phase matching,” Phys. Rev. Lett. 94, 083601 (2005). [CrossRef]   [PubMed]  

70. M. Hendrych, M. Mičuda, and J. P. Torres, “Tunable control of the frequency correlations of entangled photons,” Opt. Lett. 32, 2339–2341 (2007). [CrossRef]   [PubMed]  

71. P. J. Mosley, J. S. Lundeen, B. J. Smith, P. Wasylczyk, A. B. U’Ren, C. Silberhorn, and I. A. Walmsley, “Heralded generation of ultrafast single photons in pure quantum states,” Phys. Rev. Lett. 100, 133601 (2008). [CrossRef]   [PubMed]  

72. T. E. Keller and M. H. Rubin, “Theory of two-photon entanglement for spontaneous parametric down-conversion driven by a narrow pump pulse,” Phys. Rev. A 56, 1534–1541 (1997). [CrossRef]  

73. W. P. Grice, A. B. U’Ren, and I. A. Walmsley, “Eliminating frequency and space-time correlations in multiphoton states,” Phys. Rev. A 64, 063815 (2001). [CrossRef]  

74. X. Shi, A. Valencia, M. Hendrych, and J. Torres, “Generation of indistinguishable and pure heralded single photons with tunable bandwidth,” Opt. Lett. 33, 875–877 (2008). [CrossRef]   [PubMed]  

75. J. P. Torres, M. W. Mitchell, and M. Hendrych, “Indistinguishability of entangled photons generated with achromatic phase matching,” Phys. Rev. A 71, 022320 (2005). [CrossRef]  

76. K. W. Chan, J. P. Torres, and J. H. Eberly, “Transverse entanglement migration in Hilbert space,” Phys. Rev. A 75, 050101(R) (2007). [CrossRef]  

77. M. Hendrych, X. Shi, A. Valencia, and J. P. Torres, “Broadening the bandwidth of entangled photons: a step towards the generation of extremely short biphotons,” Phys. Rev. A 79, 023817 (2009). [CrossRef]  

78. S. Carrasco, J. P. Torres, L. Torner, A. V. Sergienko, B. E. Saleh, and M. C. Teich, “Spatial-to-spectral mapping in spontaneous parametric down-conversion,” Phys. Rev. A 70, 043817 (2004). [CrossRef]  

79. A. Valencia, A. Ceré, X. Shi, G. Molina-Terriza, and J. P. Torres, “Shaping the waveform of entangled photons,” Phys. Rev. Lett. 99, 243601 (2007). [CrossRef]  

aop-2-3-319-i001 Juan P. Torres leads a research group in Optics at the ICFO-Institute of Photonic Sciences in Barcelona, Spain, since 2003. He is also associate professor at the Universitat Politècnica de Catalunya since 1996, where he teaches courses on electromagnetic theory and several aspects of Photonics. He did his undergraduate studies in physics at the Universitat de Barcelona from 1982 to 1987, and received a Ph.D. in Sciences from the Universitat Politècnica de Catalunya in 1994. Juan P. Torres' main area of interest is Nonlinear and Quantum optics. In particular, he is interested in exploring theoretically and experimentally the unique features of new types of optical waves, such as solitons and vortex beams, and basic concepts of quantum theory, such as entanglement and decoherence. Juan P. Torres has co-authored numerous scientific papers published in international peer-reviewed journals and has established collaborations with research groups in many different countries. Nowadays, he is very interested in applying concepts and techniques born and developed in nonlinear and quantum optics to the life sciences.

aop-2-3-319-i002 Martin Hendrych received his Ph.D. degree in quantum optics from Palacky University, Olomouc, Czech Republic, in 2003. During his study, he experimentally implemented quantum key distribution, quantum identification, and quantum secret-sharing schemes. In 2000, he was awarded a NATO Advanced Science Fellowship to fund his stay in the Quantum Imaging Laboratory at Boston University, Boston. Upon completion of his degree, he worked as a Research Scientist at the Joint Laboratory of Optics of Palacky University and the Institute of Physics of the Czech Academy of Sciences, Olomouc, Czech Republic. Since 2005, he is a Research Fellow at ICFO—Institute of Photonic Sciences, Barcelona, Spain. Among his main research areas are quantum and nonlinear optics, design and implementation of sources of entangled photons in the fields of quantum information and quantum communications, frequency entanglement, dispersion control, and Bragg-reflector waveguides.

aop-2-3-319-i003 Alejandra Valencia works at ICFO—The Institute of Photonic Sciences in Barcelona, Spain, since October 2005. She did her undergraduate studies in physics at Universidad de los Andes, Bogotá, Colombia (1994–1999). In 2002, A. Valencia received her Master in Sciences degree and in 2005 her Ph.D., both from the University of Maryland Baltimore County (UMBC), USA. The topic of her dissertation was the study of protocols for clock synchronization based on the characteristics of entangled photon pairs. As a postdoctoral researcher, her interest has been mainly oriented towards the engineering of the frequency correlations of entangled photon pairs and the generation of pure single photons. A. Valencia has coauthored various scientific papers published in international peer-reviewed journals and has established collaborations with research groups in different countries. Nowadays, she works in the knowledge and technology transfer unit (KTT) of ICFO developing all the outreach and scientific divulgation activities of the institute.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (21)

Fig. 1
Fig. 1 Decomposition of a white-light beam into different colors due to the angular dispersion introduced by a prism, as originally depicted by I. Newton in Opticks[1].
Fig. 2
Fig. 2 Schematic of a grating surface and definition of angles: (+) refers to positives angles, and (−) to negative ones; m > 0 corresponds to positive diffraction orders, and m < 0 to negative diffraction orders.
Fig. 3
Fig. 3 Tilting of the front of the pulse. After the grating, the front of the pulse is no longer perpendicular to the direction of propagation.
Fig. 4
Fig. 4 The line of the loci of peak intensities in the x c t plane is tilted by an angle Φ.
Fig. 5
Fig. 5 The material whose dispersive properties are to be tailored is located between two gratings Gr1 and Gr2. Light enters the dispersive medium perpendicularly to its input face.
Fig. 6
Fig. 6 Schematic of a prism and definition of angles.
Fig. 7
Fig. 7 General scheme for pulse compression.
Fig. 8
Fig. 8 (a) General scheme for CPA and compression. (b) Device that can introduce positive or negative dispersion, depending on the length d. f is the focal length of the two lenses.
Fig. 9
Fig. 9 Detailed evolution of the SH pulse (a) with and (b) without GVM compensation. With GVM compensation with the help of pulse-front tilt, the SH quickly and efficiently builds up. Without GVM compensation, the nonlinear process is very inefficient, and the SH hardly appears. In addition, the lack of broadband phase-matching results in backconversion. Conditions: input FF peak intensity, 10 MW cm 2 ; FF input beam width, 3 mm ; FF input pulse duration, 100 fs ; wavelength of the fundamental wave, 1.6 μ m ; length of the NPP crystal [N-(4-nitrophenyl)-L-prolinol], 3 mm . Figure courtesy of J. P. Torres [12].
Fig. 10
Fig. 10 Schematic diagram of a typical SHG configuration that uses pulses with pulse-front tilt. (a) The FF beam acquires pulse-front tilt and is focused into a PPLN crystal with a lens. A second grating is used to remove the angular dispersion introduced by the first grating. (b) Close-up view that illustrates the tilted quasi-phase-matching grating used. Figure courtesy of A. Schober [42].
Fig. 11
Fig. 11 SHG obtained with angular dispersion in PPLN. (a) Measured autocorrelation and (b) spectrum with angular dispersion. For the sake of comparison, (c) and (d) show the measured autocorrelation and spectrum when a crystal of identical length in an collinear configuration, with no pulse-front tilt, is used. Figure courtesy of A. Schober [42].
Fig. 12
Fig. 12 Simulated evolution of the peak intensity of the SH beam as a function of the crystal length. Solid curve, evolution according to Eqs. (61) with GVM compensation and no loss; dotted-dashed curve, evolution according to Eqs. (60) with GVM compensation and no loss; dashed curve, evolution according to Eqs. (61) with GVM compensation and loss; dotted curve, evolution according to Eqs. (61) with no GVM compensation. Inset, SH output pulse. Conditions: input FF peak intensity, 10 MW cm 2 ; FF input beam width, 3 mm ; FF input pulse duration, 100 fs . Figure courtesy of J. P. Torres [12].
Fig. 13
Fig. 13 Experimental setup to observe quadratic temporal solitons in BBO. (a) Schematic of the experiment. (b), (c) Highly elliptical spatial profiles of the pump beam. A cylindrical lens focuses the beam in the y direction. Figure courtesy of F. Wise [56]. © 2004 by the American Physical Society.
Fig. 14
Fig. 14 Experimental (a) temporal and (b) spatial widths of the solitons during propagation. The dashed curves represent the temporal and spatial widths of the wave if the propagation were dictated only by dispersion and diffraction that produce temporal and spatial broadening. The experimental black diamonds confirm that the temporal and spatial widths of the generated soliton remained constant. The insets show the temporal and spatial profiles at some selected distances. The peak intensity is 8 GW cm 2 , and Δ k = 60 π 25 mm . Figure courtesy of F. Wise [57].
Fig. 15
Fig. 15 Schematic of the configuration for noncollinear phase matching of optical and THz waves with pulse-front tilt.
Fig. 16
Fig. 16 Spectra of the THz pulses measured for different tilt angles ν. The maxima of the spectra are normalized. Figure courtesy of J. Hebling [62]. © 2004 by Springer.
Fig. 17
Fig. 17 Dependence of the energy and the frequency of the THz pulses on the tilt angle ν. The solid curves are guides to the eye. Figure courtesy of J. Hebling [62]. © 2004 by Springer.
Fig. 18
Fig. 18 Experimental setup used to demonstrate the control of frequency correlation and the bandwidth enhancement in SPDC by means of angular dispersion. G denotes gratings; PBS, polarization beam splitter; Mono, monochromators; D, single-photon counting modules; M, mirrors; &, coincidence electronics.
Fig. 19
Fig. 19 Shape S ( ω s , ω i ) of the frequency correlations measured experimentally (left) and predicted theoretically (right). (a), (b) no tilt, Φ = 0 ° ; (c), (d) anticorrelated photons, Φ = 38 ° ; (e), (f) uncorrelated photons, Φ = 20 ° ; (g), (h) correlated photons, Φ = 52 ° . Pump-beam bandwidth, Δ λ p = 2 nm ; nonlinear crystal length, L = 3.5 nm . Figure courtesy of M. Hendrych [77]. © 2009 by the American Physical Society.
Fig. 20
Fig. 20 Measured joint spectral density S ( ω s , ω i ) for a 2 mm type-II BBO crystal: (a) tilt angle Φ = 0 ° and (b) tilt angle Φ = Φ II max = 38 ° . The joint spectrum broadens sevenfold. Figure courtesy of M. Hendrych [77]. © 2009 by the American Physical Society.
Fig. 21
Fig. 21 (a) Signal single counts for Φ = 0 ° ; Δ λ s = 5.2 nm . (b) Signal single counts for Φ = Φ II max = 38 ° ; Δ λ s = 37 nm . (c) Coincidences along the antidiagonal for Φ = 0 ° ; Δ Λ = 7.5 nm . (d) Coincidences along the antidiagonal for Φ = Φ II max = 38 ° ; Δ Λ = 52 nm . Solid curves represent the theoretical prediction; squares are the experimental data.

Tables (1)

Tables Icon

Table 1 Conditions to Obtain Spatial, Temporal, and Spatiotemporal Solitons

Equations (85)

Equations on this page are rendered with MathJax. Learn more.

sin ϵ ¯ + sin θ ¯ = m λ d ,
ϵ = α θ + γ Δ λ ,
α = cos θ 0 cos ϵ 0
γ = m d cos ϵ 0
E ( x , y , z , t ) = 1 2 A ( x , y , z , t ) exp ( i k 0 z i ω 0 t ) + h.c ,
A ( x , y , z , t ) d q x d q y d Ω a ( q x , q y , Ω , z ) exp ( i Ω t + i q x x + i q y y ) .
q x = p x α tan Φ α c Ω ,
tan Φ = c k 0 ( ϵ ω ) ω 0 .
a ( q x , q y , Ω , z 1 = 0 ) a ( q x α tan Φ α c Ω , q y , Ω , z 2 = 0 ) .
a ( q x , q y , Ω , z 2 = 0 ) a ( q x , q y , Ω , z 2 = 0 ) exp ( i k 0 Ω z 2 ) ,
A ( x 2 , y 2 , z 2 , t ) d q x d q y d Ω a ( q x α tan Φ α c Ω , q y , Ω , z 2 = 0 ) exp ( i k 0 Ω z 2 ) exp ( i Ω t + i q x x 2 + i q y y 2 ) .
A ( x 1 , y 1 , z 1 = 0 , t ) A ( α x 2 , y 2 , z 2 , t [ k 0 z 2 + tan Φ c x 2 ] ) .
t ( k 0 z 2 + tan Φ c x 2 ) = 0 .
tan ν tan Φ c k 0 .
tan Φ = n 0 λ 0 γ .
n 1 ϵ 1 = n 2 ϵ 2 ,
a ( q x , q y , Ω , z ) = a ( q x , q y , Ω , z = 0 ) exp [ i ( k ( ω ) k 0 ) z ] exp ( i | q | 2 2 k 0 z ) exp ( i q x z tan ρ ) ,
k ( ω ) = k 0 + k 0 Ω + 1 2 k 0 Ω 2 .
a ( q x , q y , Ω , z = L ) a ( q x α tan Φ α c Ω , q y , Ω , z = 0 ) exp ( i | q | 2 2 k 0 L ) exp ( i q x L tan ρ ) exp ( i k 0 Ω L + i 2 k 0 Ω 2 L ) .
a ( q x , q y , Ω ) a ( α q x + Ω tan Φ c , q y , Ω ) .
A ( x , y , z = L , t ) d q x d q y d Ω a ( q x , q y , Ω , z = 0 ) exp ( i q x x ) exp ( i q y y ) exp { i Ω ( t k 0 L + α q x L tan Φ k 0 c tan Φ tan ρ c L ) } exp { i Ω 2 [ k 0 ( tan Φ c ) 2 1 k 0 ] L 2 } exp ( i α q x tan ρ L ) exp ( i α 2 q x 2 2 k 0 L ) exp ( i q y 2 2 k 0 L ) .
A ( y , z = L , t ) = d q y d Ω a ( q y , Ω , z = 0 ) exp ( i q y y ) exp ( i q y 2 2 k 0 L ) exp { i Ω [ t ( k 0 + tan Φ tan ρ c ) L ] } exp { i Ω 2 [ k 0 ( tan Φ c ) 2 1 k 0 ] L 2 } .
k 0 , eff = k 0 + tan Φ tan ρ c ,
k 0 , eff = k 0 1 k 0 ( tan Φ c ) 2 .
k 0,eff = k 0 + tan Φ tan ρ c ,
k 0,eff = k 0 1 k 0 ( tan Φ c ) 2 .
sin θ ¯ = n ( λ ) sin δ ¯ 1 ,
n ( λ ) sin δ ¯ 2 = sin ϵ ¯ .
ϵ = α θ + γ Δ λ ,
α = cos θ 0 cos ϵ 0 cos δ 20 cos δ 10 ,
γ = sin C cos ϵ 0 cos δ 10 ( n λ ) λ 0 ,
ϵ = θ + 2 ( n λ ) λ 0 Δ λ .
tan Φ = 2 λ 0 ( n λ ) λ 0 .
A ( x , Ω ) = A 0 exp ( α 2 x 2 w 0 2 y 2 w 0 2 ) exp ( Ω 2 T 0 2 4 + i tan Φ c Ω x ) ,
A ( x , Ω ) = A 0 exp [ ( x ξ Ω ) 2 w 0 2 ] exp ( Ω 2 T 0 2 4 + i δ Ω 2 2 + i μ Ω x ) .
I ( x , t ) exp [ 2 ( t δ v ¯ x μ x ) 2 τ ¯ 2 ] exp ( 2 x 2 w ¯ 0 2 ) ,
v ¯ = ξ ξ 2 + T 0 2 w 0 2 4 ,
τ ¯ = ( T 0 2 + 4 ξ 2 w 0 2 + 4 δ 2 T 0 2 + 4 ξ 2 w 0 2 ) 1 2 ,
w ¯ 0 = [ 1 w 0 2 v ¯ 2 ( T 0 2 4 + ξ 2 w 0 2 ) ] 1 2 .
A ( x , Ω ) exp ( Ω 2 T 0 2 4 + i z Ω c ) exp { ( x α z μ Ω k 0 ) 2 w 0 2 + 2 i α 2 z k 0 } exp { i μ 2 z 2 k 0 Ω 2 } ,
A ( t , z = 0 ) = A 0 exp ( t 2 T 1 2 ) a ( Ω , z = 0 ) exp ( Ω 2 T 1 2 4 ) ,
A ( t , z ) = A ( t , z = 0 ) exp ( i γ f P ( t ) z ) ,
A ( t , z ) = A 0 exp [ t 2 T 1 2 ( 1 + 2 i γ f P 0 z ) ] ,
a ( Ω , z ) exp { Ω 2 T 1 2 4 [ 1 + ( 2 γ f P 0 z ) 2 ] + i Ω 2 γ P 0 z T 1 2 2 [ 1 + ( 2 γ f P 0 z ) 2 ] } .
B 2 = B 1 [ 1 + ( 2 γ f P 0 z ) 2 ] 1 2 .
α SPM = γ f P 0 T 1 2 z [ 1 + ( 2 γ f P 0 z ) 2 ] .
A ( t , L ) = A 0 T 1 T 1 2 2 i k eff L exp ( t 2 T 1 2 2 i k eff L ) ,
P 2 = P 1 1 + ( 2 k eff L T 1 2 ) 2 ,
T 2 = T 1 1 + ( 2 k eff L T 1 2 ) 2 .
H ( q ) = exp { i f d k 0 | q | 2 } .
β = f d k 0 ( tan Φ c ) 2 .
k 2 = 2 k 1 .
k 2 = k 1 ,
L gvm = T 0 | k 2 k 1 | ,
I 2 ( L ) I 1 ( 0 ) I 1 ( 0 ) L 2 sinc 2 ( Δ k L 2 ) = I 1 ( 0 ) L 2 sin 2 ( Δ k L 2 ) ( Δ k L 2 ) 2 ,
n 1 ( ϵ ( λ 1 ) , λ 1 ) = n 2 ( ϵ ( λ 2 ) , λ 2 ) ,
( ϵ λ ) λ 10 = ( n 1 λ ) λ 10 1 2 ( n 2 λ ) λ 20 ( n 2 θ ) θ 0 ( n 1 θ ) θ 0 ,
k 1 , eff = k 1 + tan Φ tan ρ 1 c ,
k 2 , eff = k 2 + tan Φ tan ρ 2 c .
( ϵ λ ) λ 10 = ( n 1 λ ) λ 10 1 2 ( n 2 λ ) λ 20 n 1 ( tan ρ 2 tan ρ 1 ) ,
ρ ( θ ) = θ tan 1 { n o 2 n e 2 ( θ ) tan θ } ,
n 1 , 2 θ = n 1 , 2 tan ρ 1 , 2 .
i A 1 z + i k 1 A 1 t k 1 2 2 A 1 t 2 i tan ρ 1 A 1 x + 1 2 k 1 [ 2 A 1 x 2 + 2 A 1 y 2 ] + i Γ 1 2 A 1 + K 1 A 1 * A 2 exp ( i Δ k z ) = 0 ,
i A 2 z + i k 2 A 2 t k 2 2 2 A 2 t 2 i tan ρ 2 A 2 x + 1 2 k 2 [ 2 A 2 x 2 + 2 A 2 y 2 ] + i Γ 2 2 A 2 + K 2 A 1 2 exp ( i Δ k z ) = 0 ,
i A 1 z + i k 1 , eff A 1 t k 1 , eff 2 2 A 1 t 2 + i Γ 1 2 A 1 + K 1 A 1 * A 2 exp ( i Δ k z ) = 0 ,
i A 2 z + i k 2 , eff A 2 t k 2 , eff 2 2 A 2 t 2 + i Γ 2 2 A 2 + K 2 A 1 2 exp ( i Δ k z ) = 0 .
i a 1 ξ + 1 2 2 a 1 τ 2 + 1 2 L dis L dif [ α 2 2 a 1 s 2 + 2 a 1 η 2 ] + i 2 L dis L abs a 1 + 2 L dis L nl a 1 * a 2 exp ( i 2 π L dis L coh ξ ) = 0 ,
i a 2 ξ + 1 2 L dis L dis 2 a 2 τ 2 + 1 4 L dis L dif [ α 2 2 a 2 s 2 + 2 A 2 η 2 ] i 2 L dis L gvm a 2 τ i 2 L dis L w A 2 s + i 2 L dis L abs a 2 + 2 L dis L nl a 1 2 exp ( i 2 π L dis L coh ξ ) = 0.
A ( z , t ) = d ω a ( ω ) exp { i [ k opt ( ω 0 + ω ) k opt 0 ] z i ω t } ,
P THz NL ( Ω ) = ϵ 0 χ ( 2 ) d ω a ( ω + Ω ) a * ( ω ) exp { i [ k opt ( ω + Ω ) k opt ( ω ) ] z } .
k THz ( Ω ) = k opt ( ω + Ω ) k opt ( ω )
k THz ( Ω ) = Ω n THz ph c ,
n THz ph = n opt gr ,
n opt , eff gr = n opt gr + tan Φ tan ρ .
k THz ( Ω ) = k opt ( ω + Ω ) k opt ( ω ) .
n THz ph cos γ ¯ = n opt gr ,
n THz ph sin γ ¯ = n opt ph ω 0 ( ϵ ω ) ω 0 .
tan γ ¯ = n opt ph n opt gr ω 0 ( ϵ ω ) ω 0 .
tan ν = tan Φ n opt gr .
| Ψ = d Ω s d Ω i d q s d q i Ψ ( Ω s , Ω i , q s , q i ) | ω s 0 + Ω s , q s s | ω i 0 + Ω i , q i i ,
Ψ ( Ω s , Ω i , q s , q i ) = E p ( Ω s + Ω i , q s + q i ) sinc ( Δ k L 2 ) exp ( i s k L 2 )
Δ k = ( k p , eff k s , eff ) Ω s + ( k p , eff k i , eff ) Ω i 1 2 k s , eff Ω s 2 1 2 k i , eff Ω i 2 1 2 k p , eff ( Ω s + Ω i ) 2 ,
Δ k ( k i , eff k s , eff ) Ω s 1 2 ( k s , eff + k i , eff ) Ω s 2 .
Φ II max = tan 1 { c ( k i k s ) tan ρ s tan ρ i } ,
Φ I max = tan 1 c 2 k s k s 0 ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.