Raman scattering has found renewed interest owing to the development of gas-filled hollow-core fibers, which constitute a unique platform for exploration of novel ultrafast nonlinear phenomena beyond conventional solid-core-fiber and free-space systems. Much progress has been made through models for particular interaction regimes, which are delineated by the relation of the excitation pulse duration to the time scales of the Raman response. However, current experimental settings are not limited to one regime, prompting the need for tools spanning multiple regimes. Here, we present a theoretical framework that accomplishes this goal. The theory allows us to review recent progress with a fresh perspective, makes new connections between distinct temporal regimes of Raman scattering, and reveals new degrees of freedom for controlling Raman physics. Specific topics that are addressed include transient Raman gain, the interplay of electronic and Raman nonlinearities in short-pulse propagation, and interactions of short pulses mediated by phonon waves. The theoretical model also accommodates vector effects, which have been largely neglected in prior works on Raman scattering in gases. The polarization dependence of transient Raman gain and vector effects on pulse interactions via phonon waves is investigated with the model. Throughout this Perspective, theoretical results are compared to the results of realistic numerical simulations. The numerical code that implements the new theory is freely available. We hope that the unified theoretical framework and numerical tool described here will accelerate the exploration of new Raman-scattering phenomena and enable new applications.

Raman scattering is a type of inelastic scattering in which photons undergo frequency up- or down-conversion through interactions with material vibrational or rotational excitations (phonons). It was predicted by Smekal in 19231 and observed experimentally in liquids by Raman and Krishnan in 19282–4 and independently in quartz crystals by Mandelstam and Landsherg during the same period.5–7 Raman was awarded the 1930 Nobel prize, although the full and undivided recognition of his contribution has been contentious.8–11 There was even a suggested designation of the phenomenon as “the effect of Raman, Mandelstam, and Landsherg.”12 Within two years, its discovery had led to numerous subsequent observations in 60 different liquids and gases, which were recognized as providing support for the correctness of the then-new quantum theory.13 Moreover, it has given rise to a diverse range of applications. For example, Raman spectroscopy enables identification and characterization of various materials,14–17 and Raman amplifiers and Raman lasers extend wavelengths beyond inherent limitations of natural lasing media.18–25 

With the advent of hollow-core fibers (HCFs), studies of optical interactions with atomic or molecular gases have attracted much attention.26–30 In particular, Raman scattering in molecular gases31–35 has experienced a resurgence in popularity since the pioneering work by Benabid et al.36 HCFs offer a substantial reduction of the Raman threshold through the combination of long interaction length and high intensity. Raman frequency down-conversion in H2,37–45 D2,46–52 CH4,53–59 CO2,60,61 and N262 has been demonstrated to generate a wide range of colors. Similarly, frequency up-conversion has been achieved with control of wave-vector matching63–65 and successfully applied in the generation of quantum-state-preserving photons.66 Further tunability of the converted wavelengths has been realized through the Raman-induced soliton self-frequency shift (SSFS)67–69 or its combination with photoionization-induced blue-shifting of the spectrum.70,71 Raman-enhanced supercontinuum generation has been demonstrated to cover from vacuum ultraviolet to near-infrared wavelengths,72–76 which makes possible pulse compression down to few-cycle durations.77–82 

The analysis of stimulated Raman scattering (SRS) is commonly classified into multiple regimes based on the relationship of the excitation pulse duration to the material response time (Fig. 1). The most extensively examined regime is the “steady-state” regime, in which the pulse duration (△tp) is significantly longer than the phonon dephasing time T2.25,83–85 In this regime, the Raman response of a medium depends on each temporal segment of a pulse, so the Raman interaction is determined by the pulse’s instantaneous intensity. Moreover, the independence of the temporal segments leads to an incoherent Raman pulse that originates from noise. If the pulse duration becomes shorter than the dephasing time, the interaction enters the “transient” regime. This regime allows long-lasting phonon interactions to drive the SRS process within a pulse. When the Raman gain is high and saturated, “Raman memory” enables coherent generation of Raman pulses that tend to exhibit the same temporal phase profile as the pump pulse.53 Ultrashort femtosecond Raman pulses can also be produced without Raman spectral narrowing by scattering the pulse with existing phonons.86–90 If the pulse is even shorter than the phonon oscillation period TR {equivalently, the inverse of the Raman transition frequency [ ω R / 2 π ] }, phonons are “impulsively” excited91 and persist in the medium after the pulse has left. The excited phonons allow for nonlocal interactions between the first and subsequent excitation pulses, which leads to controllable nonlinear dynamics, such as blue-shifting, red-shifting, and pulse compression, with varying delays between the pulses.92–95 By probing the nuclear motions (phonons), the ultrafast dynamics of molecules can be monitored, which underlies time-resolved Raman spectroscopy.96–100 In addition, so-called time crystals formed by periodic phonon waves establish an analogy with condense-matter physics, where Bloch oscillations and Zener tunneling are observed.101 Generalization of the “phonon” concept from solid-state physics to refer to the “coherence wave” (the off-diagonal term of the density matrix of molecular motion) in molecular gases was introduced by Russell and co-workers.42,66,102,103 The use of “phonon” to refer to the coherence wave also minimizes potential confusion in discussions about the “coherent” generation of a Raman process, which pertains to interactions involving temporal phases between pump and Raman pulses. Details of the generalization and justification are presented in Sec. 2 of the supplementary material.

FIG. 1.

Various Raman regimes depending on the temporal relation between the pulse (blue lines) and the Raman response (orange lines).

FIG. 1.

Various Raman regimes depending on the temporal relation between the pulse (blue lines) and the Raman response (orange lines).

Close modal

To establish a complete physical picture of SRS, it will be valuable to develop a unified theory that applies across temporal regimes. Despite the long history of Raman scattering and its widespread use in both the scientific and industrial communities, such a theory is still lacking. As an example of where it will be useful, we highlight techniques for ultrashort Stokes-pulse generation based on the interactions of multiple pulses with phonons.53,86–90 In this process, it is essential for the pump pulse duration to be long enough to prevent distortions and avoid competing nonlinear processes. Only with the unified theory can the interaction of impulsively excited phonons and a transient Raman process be fully understood. As another example, most prior theoretical works on Raman scattering neglect other effects that can play critical roles.102,104–110 A delicate balance between phonon annihilation (anti-Stokes processes) and creation (Stokes processes) can occur through wave-vector matching of the interaction of Stokes, pump, and anti-Stokes waves. In this situation, no new phonons are generated, resulting in no new photons as well. This phenomenon is referred to as “Raman gain suppression,” where both wave-vector matching and Kerr-induced four-wave mixing (FWM) play a non-negligible role. While its impact has been thoroughly explored in the steady-state regime,85,111–114 it has only been examined in the vicinity of perfect wave-vector matching in the transient regime.115 Even recent investigations into transient Raman gain suppression in HCFs rely on the steady-state gain model,90,102,103,110 which necessitates an explanation for its effectiveness. Unlike the case of steady-state Raman gain with FWM, a straightforward analytical expression for transient Raman gain with FWM is currently lacking.

In addition to the absence of a theory that can address multiple temporal regimes, there is a deficiency in our understanding of vector (i.e., polarization) effects in SRS. This knowledge gap can lead to uncontrolled physical phenomena, as exemplified by the depolarization observed during the SSFS process in a H2-filled anti-resonant fiber.69 Almost all prior studies employ scalar models. Vector aspects have been addressed within the steady-state formalism,84,116–121 but a model that can elucidate the temporal dynamics of vector Raman interactions does not exist. Typical Raman-scattering mechanisms involve transitions between vibrational and rotational states. Vibrational SRS in solids and liquids has been the focus of many prior studies, and its vector aspects in crystalline materials have been addressed extensively.122–126 However, vibrational SRS is isotropic in gaseous environments. On the other hand, rotational SRS exhibits pronounced anisotropy owing to its inherent connection with exchange of angular momentum, and, as a result, exhibits distinct responses to light with varying polarizations. Only recently has a tensor formulation been introduced to simulate rotational SRS in air to mitigate laser-plasma instabilities through nonlinear spectral broadening with elliptically polarized light.127 General aspects of vector SRS interactions are still not fully understood.

This Perspective aims to address these gaps in our current understanding of SRS, with a focus on processes relevant to Raman generation with picosecond- or femtosecond-duration pulses in gas-filled hollow waveguides. Although steady-state Raman gain has been investigated extensively, transient Raman gain and its associated dynamics have not been thoroughly examined. Previous studies have focused on Raman effects influenced by electronic-induced nonlinear phase modulations.128–130 However, the potentially significant contribution of the Raman response to nonlinear phases has been overlooked. Although nonlocal interactions through phonon waves have been extensively investigated,91–95,101,131–143 a complete picture only emerges through analysis that can handle different temporal regimes. Finally, the vector Raman response of gases has largely been neglected.

Here, we present a theoretical model of pulse propagation in gas-filled HCFs that covers all temporal regimes (steady-state, transient, and impulsive) of SRS as well as vector effects. Application of the theory to previous experiments will serve as a partial review of recent activities in the area and set the context for future directions. The theory aids conceptual and intuitive understanding of previously observed phenomena, in some cases from new points of view, and makes predictions about processes that are currently under investigation, or will be in the future. The results also illustrate new approaches to controlling Raman scattering for wavelength-conversion applications. Analytic expressions provide qualitative and quantitative descriptions of example phenomena, and detailed numerical simulations with representative experimental parameters back up and complement the analytical results. Although the presented model is capable of treating spontaneous Raman scattering (and all numerical results presented in this Perspective were calculated with the inclusion of spontaneous Raman scattering), the quantum nature of the initiation process does not play a key role in the processes examined here, which are in the nonlinear regime of Raman scattering (i.e., when the gain is high and saturated). Thus, we do not discuss the initiation process. For treatments of spontaneous processes and the initiation of Raman scattering, readers are referred to excellent prior works.13,107,144–149

The rest of this Perspective is organized as follows. Sec. II presents the nonlinear pulse propagation equation, which is the foundation of the unified theory, numerical simulations, and subsequent analysis presented in this Perspective. Section III applies the scalar model to phenomena in all three temporal regimes, with emphases on the behavior of the Raman gain and the interplay of Raman and electronic nonlinearities that govern a variety of ultrafast phenomena. In Sec. IV, the vector Raman response is derived and applied to polarization effects in vibrational and rotational SRS in diatomic molecules. Section V offers some perspectives on possible future directions for the field of SRS physics in gas-filled HCF.

The computational process is described in Secs. 11–13 of the supplementary material, along with some technical issues that are encountered in simulations of SRS in gases. The numerical code used to produce all of the simulation results presented in this Perspective is implemented in the MATLAB computing platform and is freely available.

Raman scattering is generally analyzed with one of the two models. One model is based on the assumption of a harmonic oscillator for the excited vibrational mode,85,139–141 while the other treats the density matrix for the vibrational or rotational states.68,73,90,101,103,105,107–110,115,142,150–153 Although they are not discussed in this Perspective, those models underlie our theoretical approach. Readers are referred to Sec. 3 of the supplementary material for detailed descriptions and comparisons of the two models.

The unidirectional pulse propagation equation (UPPE) has been commonly used to study nonlinear wave propagation in gas-filled HCF due to its ability to handle broadband phenomena and few-cycle dynamics.154 The starting point of this work is the vector UPPE [Eq. (1)],154–157 whose derivation is provided in Sec. 4 of the supplementary material.In the basis of linear polarizations,
(1a)
with an analogous equation for z A y ( z , Ω ) .
In the basis of circular polarizations,
(1b)
with an analogous equation for z A ( z , Ω ) .
The equations [Eq. (1)] include dispersion as well as instantaneous electronic and delayed Raman contributions to the nonlinearity. Ap(z, T) is the envelope of the electric field (in W ) of polarization mode p, whose Fourier transform is A p ( z , Ω ) = F A p ( z , T ) . The Fourier transform is applied with respect to angular frequency Ω = ωω0, where ω0 is the center angular frequency of the numerical frequency window required to cover the investigated physical phenomena. βp is the propagation constant of mode p, obtained from the dispersion formula either for anti-resonant fiber158 or for capillary;159, β(0) and β(1) are to reduce the propagating global-phase increment to facilitate simulations in which β(1) is the inverse group velocity of the moving reference frame that introduces the delayed time T = tβ(1)z. κ ( ω ) = 1 / ε 0 2 n eff ( ω ) 2 c 2 A eff ( ω ) and κ e ( ω ) = 3 / 4 ε 0 χ electronic ( 3 ) ( ω ) . A eff ( ω ) = 1 / F 4 d 2 x is the effective mode field area, and F(x, y) is the normalized mode profile with ∫|F|2dxdy = 1. χ electronic ( 3 ) ( ω ) is the third-order nonlinear susceptibility of the electronic response (in m2/V2) that is proportional to the gas number density; R a ( t ) and R b ( t ) are isotropic and anisotropic Raman response functions that are also proportional to the gas number density and will be derived later [Eq. (31)]. They result from the polarization160 
(2)
where
(3a)
(3b)
The supplementary material of the present article includes a detailed derivation of the UPPE for a multimode system, which includes both transverse modes and polarization modes, along with both isotropic and anisotropic Raman responses. For the sake of simplicity in illustrating the underlying physics, this Perspective considers only a single transverse mode [Eq. (1)]. The conclusions drawn here can be extended to model nonlinear dynamics involving multiple transverse modes42,78–82,103 using the multimode UPPE in the supplementary material.
The quantum nature of Raman scattering is treated by incorporating the quantum-statistical Langevin terms Γ ̃ m n R r ( z , T ) (r = a or b), which correspond to vacuum fluctuations and collisional dephasing. They obey the spectral correlation given by144,148,149,161,162
(4)
where Ω′ = ω′ − ω0. The factors resulting from overlap integrals S x x R a = S y y R a = S x x R b = S + + R a = S R a = S + R b = 1 and S x y R b = S + + R b = S R b = 1 / 2 . C F is the constant of Fourier Transform where F f ( ω ) = C F f ( t ) e i ω t d t (see Sec. 1 of the supplementary material for the tutorial regarding conventions and symbols of Fourier transform). The thermal phonons obey the Bose–Einstein distribution n th ( | Ω | ) = e | Ω | / k B T 1 1 . Θ(−Ω) is the Heaviside step function. The Langevin functions Γ ̃ m n R r ’s for different vector contributions represent different random values from the same correlation relation of the Raman response function R r . Section 4 of the supplementary material describes the numerical implementation and the symmetry requirement of Langevin terms. These Langevin terms account for spontaneous Raman scattering. Spontaneous Stokes emission results from the property that Θ(−Ω) ≠ 0 for Ω < 0, representing vacuum fluctuations of phonons. On the other hand, spontaneous anti-Stokes emission is weak unless the thermal phonon population is appreciable. Raman scattering can be described as a parametric process that involves phonons and photons.163 This was employed by von Foerster and Glauber to describe the phonon evolution164 and later used to unify spontaneous and stimulated Raman scattering within a single Maxwell–Bloch theoretical framework by Raymer and Mostowski.107,145 This framework assumes a heavily populated ground state, which aligns with the perturbative Raman regime that leads to the derived Raman response functions later in Eq. (31). A new formulation is required if the perturbative assumption is relaxed and is beyond the scope of this Perspective.165 Our UPPE formulation with Langevin terms is the generalization of the quantum-electrodynamical framework of Raymer and Mostowski to the waveguide environment.144,148,149 Equation (4) is the c-number equation that corresponds to the operator’s commutator relation. Although the equations above should be able to simulate quantum phenomena, in this Perspective, we focus on the nonlinear or highly pump-depleted regime, where quantum fluctuations from spontaneous Raman scattering are minimized. In the linear Raman regime, Stokes photons, scattered by spontaneous emission, exhibit ∼100% energy fluctuations with a negative-exponential statistical distribution.166,167 However, in the nonlinear regime, the output field is stabilized, with a distribution that is statistically peaked at the average value.168 Thus, in this Perspective, “Raman generation” predominantly denotes the manifestation of SRS, that is, Raman amplification, rather than spontaneous Raman scattering. Spontaneous Raman scattering does contribute to the fluctuations observed in the results represented in Figs. 4 and 10.

Shot noise of the input field is modeled semi-classically by including one noise photon per spectral discretization bin.162,169–171 Section 4 of the supplementary material includes more details about its numerical implementation. It is worth mentioning that there are other input-pulse noise models, obtained by either carefully modeling the buildup of lasing of the source or adding a Lorentzian noise spectrum derived from a phase-diffusion model.170 However, they aim to realistically model the input-pulse noise spectrum. Gases usually have large Raman transition frequencies, and such relatively narrowband noise spectra have insignificant impact on nonlinear processes. Thus, more-realistic noise models are unnecessary for the purposes of this work.

Two types of hollow-core fibers are considered below. In HCF, the dispersion has components of anomalous waveguide dispersion and normal gas dispersion.159 An anti-resonant fiber with a small 30-μm core diameter (and 300-nm tube-wall thickness) is assumed for analysis of processes with anomalous dispersion, while a Ag-coated capillary with a large 300-μm core diameter is assumed for those with normal dispersion.

Numerical simulations assume that the HCFs are filled with H2, N2, or O2. These gases are currently the most prevalent Raman-active gases in use. Numerical values of their parameters, as well as initial validation of our model, are provided in Secs. 11 and 12 of the supplementary material.

In this section, we aim to present Raman effects across different regimes using a unified model based on a scalar UPPE. Several known phenomena will be reexamined with this unified framework. It covers not only Raman responses to Stokes generation but also coupling with anti-Stokes waves arising from both electronic- and Raman-induced FWM. FWM underlies Raman gain suppression, Raman-induced phase modulations, and other related effects, which will be discussed later.

Raman gain is a fundamental quantity in SRS physics, so Raman gain suppression is a critical concept. Since the groundbreaking research conducted by Bloembergen and Shen on Raman gain suppression,111 there has been a significant focus on understanding parametric effects on Raman gain.112–114,116–118,120,121,172–177 The small dispersion of gases translates into small wave-vector mismatch among Stokes, pump, and anti-Stokes waves, so the effect of Raman gain suppression (which is the strongest at perfect wave-vector matching) in gas-filled HCF cannot be ignored. Argon has been added to hydrogen to modify the overall dispersion profile, creating wave-vector mismatch and mitigating gain suppression,110 for example. Higher-order transverse modes have been introduced to excite intermodal coherence waves to disrupt the Stokes/anti-Stokes balance and allow the Stokes wave to grow.42 In this paper, we focus mostly on the Raman gain of Stokes waves; for anti-Stokes evolution, please see Sec. 8 of the supplementary material for details.

To study Raman gain and other SRS effects, we start from the scalar UPPE [Ay = 0, along with corresponding Langevin terms Γ ̃ y y R a = Γ ̃ x y R b = 0 , in Eq. (1a)],
(5)
where R = R a + R b / 2 and Γ = Re Γ ̃ x x R a + Γ ̃ x x R b / 2 . To simplify the analysis, a one-component Raman response is considered: R ( t ) = Θ ( t ) R coeff e γ 2 t sin ( ω R t ) , where the dephasing rate γ2 = 1/T2. (Simulation results shown below include both vibrational and rotational SRS.) If the field is composed of Stokes (ωS = ωP − △ω), pump (ωP), and anti-Stokes (ωAS = ωP + △ω) waves,
(6)
With β(0) = β(1) = 0 for analysis purposes, we obtain from Eq. (5)
(7a)
(7b)
(7c)
where weak Stokes/anti-Stokes waves are assumed so that pump depletion is negligible. The Langevin term, before taking only its real part, consists of positive- and negative-frequency parts: Γ ( t ) = Re Γ ( Ω < 0 ) e i β S z + ω t + Γ ( Ω > 0 ) e i β A S z ω t , where △βi = βPβi (i = S or AS). △β = 2βPβSβAS is responsible for wave-vector matching among the three waves. The Raman integrals are
(8a)
(8b)

In gases, the Raman response is narrowband, and frequencies detuned from ωR do not play an important role. We assume △ω = ωR to simplify the subsequent analysis. The integral R 1 is responsible for the Raman-induced index change (which underlies phase modulations), and R 2 ; i , j , k is usually responsible for the Raman gain, where i*, j*, and k* represent the complex conjugates of the corresponding fields. Equations (7) and (8) are the basis of the analysis of distinct temporal regimes below.

We start with the steady-state regime. While this regime has been extensively explored, it serves as a reference point for subsequent discussions on the less-explored transient and impulsive regimes. Additionally, most prior studies ignore the electronic response.85,102,103,105,110 The few studies that include the electronic nonlinearity treat silica fiber, where SRS is weak.172,173,178 A thorough investigation of strong Raman response plus electronic nonlinearity in gases is still lacking.

In this regime, phonons decay on a time scale much shorter than the pulse duration, so the phonon waves are affected only by the instantaneous temporal segment of the pulse. Therefore, the Raman integrals [Eq. (8)] can be approximated as
(9a)
(9b)
where I R = t R ( t τ ) d τ = R coeff ω R ω R 2 + γ 2 2 and R ss = t R ( t τ ) e i ω R t τ   = R coeff ω R 2 i ω R γ 2 + γ 2 2 is the Fourier transform of R(t) at −ω. AiAjAk in R 2 ; i , j , k is either A P 2 A k or A P 2 A k e 2 i ϕ P ( z , t ) , where ϕP(z, t) is the phase of AP. Further approximations in Eq. (9) are made by taking ωRγ2, which is typically the case in gases at pressures less than 100 bar. Since R 1 A P 2 , the Raman-induced index change leads to an instantaneous Kerr effect. With these integrals and assuming that κSκASκ and κ e S κ e A S κ e , the Raman amplitude-gain with FWM [the quantity g in AS ∼ egzAS(z=0)] becomes (see Sec. 8 of the supplementary material for the derivation details)
(10)
To gain insights into Eq. (10), the Raman shift is assumed to be small (ωRωP), and Eq. (10) can then be reduced to
(11)
Since the gain increases monotonically with A P 2 , the Raman generation depends on the instantaneous intensity of the pump and, hence, depletes only the pulse center. The second △β term under the square-root sign in Eq. (11) corresponds to Raman gain suppression (minimum gss at △β = 0), while the first term in the square brackets corresponds to modulation instability. The latter creates a gain peak when the Raman response is weak ( Im R ss Re κ e + R ss ; blue, red, and black lines in Fig. 2).172,173,179 Such a gain shape has been widely studied in media with negligible Raman nonlinearity, creating the conventional Kerr-induced FWM gain, where both the Stokes and anti-Stokes waves are maximally generated through the parametric process at β 2 κ ω P κ e A P 2 .180 However, in Raman-active gases, due to the slow dephasing, Im R ss (∝T2 in gases) can be a few thousand times larger than Re κ e + R ss (in 20-bar H2, it is about 7 000 times for the Q(1) SRS). This results in an almost-monotonically increasing gain away from △β = 0 until it asymptotically reaches the value of Raman gain without FWM, κ ω S Im R ss A P 2 (green and pink lines in Fig. 2), which is consistent with the prior works with negligible electronic nonlinearity.85,102
FIG. 2.

Steady-state Raman gain g [Eq. (11)] with respect to wave-vector mismatch △β. The arrow indicates the trend of increasing Im R ss (blue → red → black → green → pink) with fixed Re κ e + R ss . The thick black line is the line with equal magnitudes of real and imaginary parts of κ e + R ss .

FIG. 2.

Steady-state Raman gain g [Eq. (11)] with respect to wave-vector mismatch △β. The arrow indicates the trend of increasing Im R ss (blue → red → black → green → pink) with fixed Re κ e + R ss . The thick black line is the line with equal magnitudes of real and imaginary parts of κ e + R ss .

Close modal

This section analytically calculates the steady-state Raman integrals [Eq. (9)] and Raman gain [Eq. (10)], incorporating FWM arising from electronic and Raman nonlinearities. The variation of Raman gain with the strength of the electronic and Raman responses is illustrated, ranging from negligible Raman response to the pronounced response in Raman-active gases (Fig. 2).

Because gases typically possess dephasing times greater than ∼100 ps at moderate gas pressures, experiments performed with picosecond or femtosecond pulses are commonly in the transient Raman regime. Additionally, gases usually exhibit large rotational and vibrational Raman transition frequencies. Consequently, the condition of having a pulse duration longer than the phonon oscillation period is frequently met.

The biggest challenge in analyzing transient SRS is the calculation of the Raman integrals, R 1 and R 2 ; i , j , k [Eq. (8)]. During pulse propagation, AP acquires a z-dependent phase through self-phase modulations (SPM), while Ai (i = S or AS) acquires it through cross-phase modulations (XPM) from the pump. These phase modulations are intensity- and, thus, time-dependent, which leave the Raman integrals analytically unsolvable. To simplify the problem, we assume that all three pulses have an identical flat-top temporal structure, leading to time-independent phase modulations within the pulse. In addition, it is well-known that, when the gain is high and saturated, both the Stokes and anti-Stokes signals exhibit the same temporal phase structure as the pump.105 This results from amplifying only those frequencies that can maximize the Raman integrals, which always include either A P A k * or A P * A k [Eq. (7)]. As a result, Ai = CiAP (i = S or AS, and |Ci| ≪ 1 with Ci being real-valued) is reasonably assumed. Furthermore, the flat-top assumption facilitates the analysis of temporal characteristics of the Raman processes. While we make the “flat-top” assumption here, the validity of the following discussions extends more generally to pulses with negligible nonlinear phase modulations in comparison to Raman responses.

In this regime, because phonons decay slowly, we assume γ2 ≈ 0. To solve the Raman integrals [Eq. (8)], we introduce the following mathematical approximations for an arbitrary smooth real-valued function f(t) when △tpTR:
(12a)
(12b)
These approximations are derived by partitioning one into multiple integrals, each with a small interval equal to TR. Since TR ≪ △tp, f(τ) remains nearly stationary within each interval [tn+1, tn], allowing us to approximate it as f ( t n + 1 ) + f ( t n + 1 ) τ t n + 1 , which leads to the final results [Eq. (12)]. Details are in Sec. 7 of the supplementary material.
With these approximations [Eq. (12)], the Raman integrals in the transient regime are approximated as
(13a)
(13b)
where APCk = Ak and A P C k * = A k * e 2 i ϕ P ( z , t ) . These equations are similar to Eq. (9), with the only difference lying in the imaginary part where A P 2 T 2 in the steady-state regime corresponds to t A P ( τ ) 2   d τ in the transient regime. The similarity permits the process applied in the steady-state regime to be applied to obtain an analytic expression for the Raman gain in the transient regime,
(14)

The similarity of the transient equations [Eqs. (13) and (14)] to the steady-state equations [Eqs. (9) and (10)] allows several phenomena in the steady-state regime to be transferred to the transient regime, such as the Raman gain shape and the resulting Raman gain suppression at △β = 0. This explains why the expression for the steady-state Raman gain has proven useful in predicting the qualitative features of transient Raman gain suppression.42,90,102,103,110

Despite the apparent similarity of the steady-state and transient regimes, they exhibit some important differences. First, the connection of I m R between the steady-state value A P 2 T 2 and the transient value t A P ( τ ) 2 d τ A P 2 t p shows that the transient Raman gain is smaller than the steady-state gain by a factor of △tp/T2. This leads to the potential existence of a Raman gain peak for an ultrashort pulse due to the reduced Im R tr or increased Re κ e A P 2 + R tr . In addition, monotonically increasing Im R tr with time results in not only stronger gain at the pulse trailing edge but also a varying gain shape throughout the pulse [Fig. 3(a)]. There is also a significant difference regarding where the Raman growth is the strongest. The Raman gain in the steady-state regime depends on pulse instantaneous intensity, whereas it relies on integrated pulse energy in the transient regime. This leads to the strongest Raman generation at the pulse center in the steady-state regime [Fig. 3(b)] but at the trailing edge in the transient regime [Fig. 3(c)].

FIG. 3.

(a) Varying Raman gain (extracted from Fig. 2) at different positions of a pulse in the transient Raman regime. Left represents the leading edge. Temporal profiles of the pump (P) and the Stokes (S) waves from the rotational Raman-Stokes generation in (b) the steady-state and (c) the transient regimes are shown. (b) A 20-ns and 70-μJ and (c) a 10-ps and 1-μJ Gaussian pulses at 1030 nm are launched into a 1-m-long anti-resonant fiber, filled with 30-bar H2. At this pressure, the dephasing time is 100 ps. Despite the linearly polarized fields, the dominant process is S(1) rotational SRS181 due to the stronger vibrational Raman gain suppression. All pulses in (b) and (c) are normalized to a peak power equal to one.

FIG. 3.

(a) Varying Raman gain (extracted from Fig. 2) at different positions of a pulse in the transient Raman regime. Left represents the leading edge. Temporal profiles of the pump (P) and the Stokes (S) waves from the rotational Raman-Stokes generation in (b) the steady-state and (c) the transient regimes are shown. (b) A 20-ns and 70-μJ and (c) a 10-ps and 1-μJ Gaussian pulses at 1030 nm are launched into a 1-m-long anti-resonant fiber, filled with 30-bar H2. At this pressure, the dephasing time is 100 ps. Despite the linearly polarized fields, the dominant process is S(1) rotational SRS181 due to the stronger vibrational Raman gain suppression. All pulses in (b) and (c) are normalized to a peak power equal to one.

Close modal
It has long been stated that the transient Raman gain is proportional to the square root of the integrated pulse energy,105,107,109,110 which seems to contradict the results above. By ignoring pump depletion and keeping only the term for Raman growth, the Stokes governing equation [Eq. (7b)] can be solved analytically,104,105,107
(15)
where I1(x) is the modified Bessel function of the first kind; u ( t , τ ) = 2 κ S ω S R coeff z U ( t ) U ( τ ) , where U ( t ) = t A P ( τ ) 2 d τ . By reasonably assuming ASAP and γ2 = 0, the Stokes field in Eq. (15) can be reduced to
(16)
where u = u(t, −) and the relation d I 0 ( x ) d x = I 1 ( x ) is employed. In studies of Raman gain, especially when the goal is to determine which type of SRS dominates, only the initial growth of each wave matters. Therefore, the Stokes growth is determined by the Raman gain 1 2 κ S ω S R coeff U ( t ) = κ S ω S Im R tr (1 + x ≈ ex as |x| ≪ 1), rather than u in the opposite limit. This proportionality to the integrated pulse energy aligns with our finding that the transient Raman gain without parametric suppression from the anti-Stokes wave (or equivalently with |△β| ≫ 1) is directly proportional to the integrated pulse energy. Moreover, Eq. (15) shows that, in the limit of high and saturated gain, the Stokes wave is not affected by the exact form of its initial shape AS(0, t) because the rapid variations are averaged out by the integration, which results in coherent Stokes pulses that are free from spontaneous emission noise.105 In the low-gain or unsaturated regime or in the initial growth of Raman pulses, this statement is not valid; there, the Raman pulses are strongly influenced by spontaneous quantum noise and exhibit significant fluctuations.166,167

In transient Raman scattering with high and saturated gain, the Raman pulses tend to inherit the temporal phase of the pump pulse. This results in periodic temporal modulations of participating pulses,105,106,109 self-similarity in pulse evolutions,182–184 Kerr-induced Raman suppression, and Raman pulse compression. In this paragraph, we will specifically focus on the last two effects. Konyashchenko et al. have discussed the suppression of Raman generation,128 which arises from the nonlinear wave-vector mismatch of pump and Raman pulses that do not have flat-top intensity profiles. This Kerr-induced suppression of Raman generation from SPM and XPM occurs because different rates of spectral broadening detune the frequency difference between the pump and the Stokes pulses from the Raman transition frequency (△ω = ωPωSωR), where Raman generation is most efficient. Mathematically, a differential nonlinear phase between the pump and Stokes pulses results in a reduction of the Raman temporal integral R 2 ; i , j , k , which contains either A P A i * (i = S or AS) or its complex conjugate. Since the pump pulse acquires a phase of ω P κ P κ e P + u R u coeff ω R u A P 2 through SPM and SRS (contributed by R 1 ) and the Stokes/anti-Stokes pulse acquires ω i κ i 2 κ e i + u 5 R u coeff 4 ω R u A P 2 through XPM and SRS (contributed by R 1 and Re R tr ), the difference is κ ω P 2 ω i κ e + ω P 5 4 ω i u R u coeff ω R u A P 2 , assuming that κ and κe are frequency-independent. This differential nonlinear phase increment is eliminated when the expression within the square brackets is equal to zero, a condition that can only be met in Stokes generation. To avoid suppression of Raman generation by SPM or XPM for non-flat-top pulses, there must be no differential nonlinear phase accumulation between the pump and Stokes pulses. For vibrational SRS in hydrogen with ∼125-THz transition frequency, this condition can be met with a linearly polarized pump wavelength around 970–1000 nm, which almost aligns with that of established and readily available Yb-based laser systems. Neglect of the Raman-induced nonlinear phases that arise from R 1 and Re R tr leads to the conclusion that Raman generation is most effective at 1205 nm.128 As will be discussed below, SRS in gases induces significant nonlinear phases, even larger than the electronic-induced nonlinear phases (Fig. 7) and is thus non-negligible. Here, we treat Raman suppression using a simplified model for the differential nonlinear phase. The most-accurate treatment of Raman gain suppression will come from direct incorporation of nonlinear-phase effects in the Raman gain equation [Eq. (14)]; that will require generalization of the theory beyond the assumption of “flat-top” pulses, which remains for future work.

The coherent nature of transient SRS can be illustrated by the simultaneous compression and Raman shifting of femtosecond-duration pulses. In numerical simulations, positively chirped pulses with a 200-fs transform-limited duration were launched into a H2-filled capillary. The pulse energy is fixed at 500 μJ for constant Raman gain (at huge |△β|); the varying chirped-pulse duration affects only the Kerr-induced nonlinear phase. Initially, the Stokes generation efficiency increases monotonically with increasing chirped duration due to weakening Kerr-induced Raman suppression [orange line in Fig. 4(a)]. When the gain is high and saturated, Raman pulses attempt to acquire the same phase as the pump to maximize the Raman integrals.105 This includes the acquisition of the nonlinear phase of the pump before the Kerr-induced suppression becomes significant. Spectral broadening through nonlinear phase accumulations can be used to compress pulses in a dispersive delay line. In this case, the SPM-induced nonlinear phase accumulated by the pump is transferred to the Stokes pulses, which can be compressed to durations about 3–15 times shorter than the input pump pulse [Fig. 4(b)]. In the transient Raman regime, Raman spectral narrowing occurs in the single-pulse approach, where only the trailing edge of the chirped pump pulse is transformed into Raman pulses and results in their narrower bandwidths than the pump’s. By chirping the pump appropriately, Raman spectral narrowing can be overcome by this pulse-compression effect. It has also been experimentally demonstrated by Konyashchenko et al.129,130 Figure 2 plots Raman-gain curves with a fixed real part and varying imaginary part of κ e + R ss A P 2 in the gain equation, which, in the transient regime, becomes κ e A P 2 + R tr . This is useful in visualizing the temporally varying Raman gain over a pulse [Fig. 3(a)]. On the other hand, it can be visualized differently with the varying real part and fixed imaginary part, which is the case of varying the chirp of a pulse here [Fig. 4(c)]. Because the real part of κ e A P 2 + R tr depends on peak power, whereas its imaginary part on integrated pulse energy [Eqs. (13) and (14)], reducing the chirped-pulse duration but fixing the pulse energy creates the FWM peak in the Raman gain [Fig. 4(d)]. Therefore, with a proper wave-vector mismatch, Raman gain becomes stronger with reducing duration and the resulting higher peak power. This peak-power-dependent transient Raman-gain phenomenon is observed in the highly chirped regime in Fig. 4(a), where the efficiency starts to drop with increasing chirped-pulse duration. Fluctuations in generation efficiency arise from input-pulse shot noise and spontaneous Raman scattering, with increased significance observed in regimes of weaker Raman generation. However, in our example, heightened fluctuations are only evident in the long-pulse regime. In the short-pulse regime, Stokes generation is efficient due to the prominent Raman gain peak (with proper wave-vector mismatch) before Kerr-induced spectral detuning of pump and Stokes waves becomes significant. This initially efficient Raman generation quickly amplifies the pulse beyond the linear Raman regime where spontaneous emission noise dominates. After the initial generation, the Stokes pulse stops growing due to increased spectral detuning. On the other hand, in the long-pulse regime, slow Stokes generation, despite achieving a rather high efficiency after a long propagation, renders a Stokes pulse susceptible to spontaneous emission noise. Figure 4(e) summarizes the effects on Raman generation of varying nonlinear phase modulations: in the short-pulse regime, Kerr-induced Raman suppression dominates; increasing the pulse duration reduces the Kerr-induced suppression effect, enabling Raman pulse compression, and potentially introduces a rising Raman gain (if wave-vector mismatch meets where the gain peak is); further increasing the duration sees constant or dropping Raman gain based on the amount of wave-vector mismatch. More details about peak-power dependence, pulse compression, and quantum fluctuations are provided in Sec. 9 of the supplementary material.

FIG. 4.

Nonlinear-phase effects on transient vibrational Stokes generation by chirped pulses with varying duration. Pulses at 1030 nm and with 500-μJ energy and 200-fs transform-limited pulse duration are launched into a 50-cm-long capillary, filled with H2 to 20-bar pressure. (a) Generation quantum efficiency (QE) and peak power of the dechirped output Stokes pulses, produced from the Raman process. Mean values (center lines) and 1σ standard deviations (shaded areas) are calculated from ten simulations for each duration. The x axis is plotted on logarithmic scale to expose the variation at small durations. (b) Temporal profile of the dechirped Stokes pulse with the highest peak power in (a). (TL: transform-limited pulse; D: dechirped pulse). (c) Transient Raman gain gtr of a flat-top pulse, which ignores Kerr-induced Raman suppression from intensity-dependent SPM and XPM, with respect to wave-vector mismatch △β [Eq. (14)]. The arrow indicates the trend of increasing Re κ e A P 2 + R tr (blue → red → black → green → pink) with fixed I m R tr . (d) Varying Raman gain at the trailing edge of a pulse with varying chirped duration and a fixed pulse energy. (e) Summary of the influence of nonlinear phase modulations on the Raman process with different pulse durations in the transient regime.

FIG. 4.

Nonlinear-phase effects on transient vibrational Stokes generation by chirped pulses with varying duration. Pulses at 1030 nm and with 500-μJ energy and 200-fs transform-limited pulse duration are launched into a 50-cm-long capillary, filled with H2 to 20-bar pressure. (a) Generation quantum efficiency (QE) and peak power of the dechirped output Stokes pulses, produced from the Raman process. Mean values (center lines) and 1σ standard deviations (shaded areas) are calculated from ten simulations for each duration. The x axis is plotted on logarithmic scale to expose the variation at small durations. (b) Temporal profile of the dechirped Stokes pulse with the highest peak power in (a). (TL: transform-limited pulse; D: dechirped pulse). (c) Transient Raman gain gtr of a flat-top pulse, which ignores Kerr-induced Raman suppression from intensity-dependent SPM and XPM, with respect to wave-vector mismatch △β [Eq. (14)]. The arrow indicates the trend of increasing Re κ e A P 2 + R tr (blue → red → black → green → pink) with fixed I m R tr . (d) Varying Raman gain at the trailing edge of a pulse with varying chirped duration and a fixed pulse energy. (e) Summary of the influence of nonlinear phase modulations on the Raman process with different pulse durations in the transient regime.

Close modal

In the steady-state regime, these coherent phenomena play out differently. Both Raman-enhanced SPM and nonlinear-phase-induced Raman suppression are observable in both the steady-state and the transient regimes [Eqs. (9a) and (13a)]. However, in the steady-state regime, spectral broadening can only result from the Raman process attempting to maintain a constant frequency difference ωPωS = ωR, where the highest Raman gain is, rather than from the transfer of the pump’s nonlinear phase as in the transient regime. If a Raman pulse originates from noise and lacks coherence, such incoherent spectral broadening does not lead to pulse compression. Since the steady-state Raman gain depends on κ e + R ss , determined only by materials, steady-state Raman gain in Raman-active gases has no FWM-induced gain peak and shows no varying gain shape with different chirped-pulse durations as in the transient regime.

In both the steady-state and transient regimes, the nuclear motions of a medium exhibit a response analogous to the electronic nonlinearity due to either a short phonon decay time or fast molecular oscillations. As a result, they both exhibit Kerr-like characteristics [Eqs. (9a) and (13a)] in addition to Raman gains [Eqs. (9b) and (13b)]. However, the response of the medium is no longer stationary if the pulse is shorter than the molecular oscillation period TR. The medium response is delayed relative to the pulse (Fig. 5). In addition to the delayed response, the medium is impulsively excited, which creates long-lived phonon waves after the pulse and allows for nonlocal interactions between pulses (Fig. 6).

FIG. 5.

Normalized index change (orange) induced by rotational SRS (represented by △ɛ) of hydrogen in response to a pulse of varying pulse duration (blue). The oscillation period of the S(1) rotational transition is 56.8 fs. The index change is computed from ε ( r , t ) = t R ( t τ ) E ( r , τ ) 2 d τ , with E ( r , t ) being the analytic signal of the electric field.185 

FIG. 5.

Normalized index change (orange) induced by rotational SRS (represented by △ɛ) of hydrogen in response to a pulse of varying pulse duration (blue). The oscillation period of the S(1) rotational transition is 56.8 fs. The index change is computed from ε ( r , t ) = t R ( t τ ) E ( r , τ ) 2 d τ , with E ( r , t ) being the analytic signal of the electric field.185 

Close modal
FIG. 6.

Impulsively excited index change in 1 bar of (a) H2 or (b) N2 due to rotational SRS. (c) Values of Rcoeff for 1 bar of (blue) H2, (black) N2, and (red) O2 under the linearly polarized scalar assumption. The “revivals” in N2 and O2 at large time delays arise from beating of phonon waves from different rotational transitions,186 whereas in H2, the S(1) transition at 17.6 THz dominates, which leads to a constantly oscillating sinusoidal-like index wave.

FIG. 6.

Impulsively excited index change in 1 bar of (a) H2 or (b) N2 due to rotational SRS. (c) Values of Rcoeff for 1 bar of (blue) H2, (black) N2, and (red) O2 under the linearly polarized scalar assumption. The “revivals” in N2 and O2 at large time delays arise from beating of phonon waves from different rotational transitions,186 whereas in H2, the S(1) transition at 17.6 THz dominates, which leads to a constantly oscillating sinusoidal-like index wave.

Close modal
To understand the SRS mechanism in the impulsive regime, we need to solve the Raman integrals [Eq. (8)], as in the other two regimes. The mathematical approximations in the impulsive regime are
(17)
in which △t = tt and t is where the leading edge of the pulse is. The substitution of variables τ = t + x and f ̃ ( x ) = f ( t + x ) is employed. ϕ is the phase of F f ( ω R ) . In contrast to the transient regime, the decomposition of the broadband total field into each component is not valid [Eq. (6)]. Instead, we focus on the evolution of the total field by substituting A for AP in Eq. (7a). To gain insights into the evolution within the pulse, we make the flat-top assumption to simplify the equations and obtain, for 0 < △t ≤ △tp,
(18)
where A peak 2 is the peak power of the flat-top pulse.

Due to the delayed response of the medium to the pulse, R 1 corresponds to the buildup of the Raman-induced index change, which leads to a time-dependent phase increment. This causes the pulse spectrum to red-shift ( ω = d ϕ d t t < 0 ) .139 Traditionally, red-shifting is treated as a consequence of impulsive phonon-wave generation after the pulse, whereas here it is explained by the nonlinear dynamics resulting from the index change. Greater red-shifting at the trailing edge of the pulse induces the negative chirp.140 If the dispersion is anomalous, the pulse will stretch temporally. On the other hand, in the normal-dispersion regime, this nonlinear phase will compete with the positive chirp from dispersion and SPM, potentially leading to pulse compression. However, in initial simulations, we find that SPM-induced chirp dominates over Raman-induced nonlinear phases. More investigations into this effect are thus required.

The Raman-enhanced SPM effect in the transient regime, resulting from R 1 and Re R tr , has been exploited to generate ultra-broadband continua and few-cycle pulses.73,75 However, with shorter pulses, the Raman-induced index changes undergo a transition from SPM and XPM in the transient regime [Eqs. (9a) and (13a)] to red-shifting in the impulsive regime [Eq. (18)], which diminishes the SPM enhancement. Figure 5 illustrates the transition from pulse-following (∝|A|2) to a rising index [ t 2 ] with reducing pulse duration. In H2, the transition occurs at around 50 fs, where the dominant S(1) rotational transition starts to become impulsive, while in N2, it occurs at 500 fs [Figs. 6(c) and 7]. Because N2 has more available Raman transitions, which are also at smaller frequencies [Fig. 6(c)], the Raman-induced SPM enhancement in N2 is about 11.3 2.50 3.2 1.7 6 times as strong as in H2, leading to the final 11.3 3.2 3.5 times stronger Raman-enhanced SPM. This effect has been studied with 30 and 280-fs pulses in H273 and N2,75 respectively, and greater enhancements should be achievable by using pulses longer than the transition durations. On the other hand, simultaneous spectral broadening with red-shifting can produce few-cycle pulses at long wavelengths.78 The optimal pulse duration for this process lies between the transient and impulsive regimes, where the distinctive characteristics of Kerr and red-shifting effects both come into play. It is worth noting that the Raman-induced SPM is 4.5 times larger than the electronic-induced SPM in N2, while they have nearly the same magnitude in H2, values that are consistent with measurements of the total and electronic-induced nonlinear refractive indices.186 

FIG. 7.

Accumulation of a nonlinear phase from numerical simulations with varying pulse duration in (a) H2 and (b) N2. A linearly polarized Gaussian pulse at 800 nm and with a fixed 6.67-MW peak power is launched into a 2-m-long anti-resonant fiber pressurized to 1 bar. The peak power is chosen so that the nonlinearity is sufficiently weak to avoid further amplitude-related Raman-gain effects, and the dispersion is artificially set to zero to isolate the nonlinear effects. The nonlinear phase ϕmax is calculated from the spectral broadening of the pulse.187,188

FIG. 7.

Accumulation of a nonlinear phase from numerical simulations with varying pulse duration in (a) H2 and (b) N2. A linearly polarized Gaussian pulse at 800 nm and with a fixed 6.67-MW peak power is launched into a 2-m-long anti-resonant fiber pressurized to 1 bar. The peak power is chosen so that the nonlinearity is sufficiently weak to avoid further amplitude-related Raman-gain effects, and the dispersion is artificially set to zero to isolate the nonlinear effects. The nonlinear phase ϕmax is calculated from the spectral broadening of the pulse.187,188

Close modal

It is common to model the Raman response of a medium effectively as a single damped harmonic oscillator, with amplitude determined by the “Raman fraction” fr of the total nonlinear response, which can be found from Fig. 7: R ( t ) f r n 2 total e γ 2 t sin ( ω R eff t ) .77–79,189,190 We are now in position to examine the validity of the model while shedding some insights on Raman processes. This model is especially useful if the Raman response can be approximated as a harmonic oscillator with strong damping, which mitigates the potential numerical aliasing resulting from long-lived phonon waves. This is the case in N2 and O2. Due to their reviving indices, sufficiently short pulses essentially experience only the first spike of index change [Fig. 6(b)], which can be approximated by a highly damped harmonic oscillator. When the pulse becomes long enough to be affected by further reviving indices, this simplified model fails. In other words, the model operates by smearing the clustered sharp spectral Raman lines around 2 THz into one broadband Raman spectrum [Fig. 6(c)] and will fail when the smearing effect becomes invalid.

Weaker Raman-induced nonlinear phase accumulation can be advantageous for some processes. An example is SSFS, which has been widely studied in solid glass fibers.191–197 Since the soliton number is proportional to γefftp,198 where γeff is the effective or total nonlinear coefficient, a pulse in the impulsive regime has a smaller soliton number than that in the transient regime due to the reduced pulse duration. A smaller soliton number mitigates energy loss during soliton fission and thus enhances the efficiency of generating the reddest Raman soliton during SSFS.199 Furthermore, the decrease in Raman enhancement of SPM leads to a reduction in γeff, which also decreases the soliton number. The decrease can be as much as a factor of 4.5 times in N2, for example [Fig. 7(b)]. Finally, the reduced nonlinear phase is transformed to more-pronounced red-shifting through the rising Raman-induced index change [Eq. (18)]. Together, these factors make the SSFS a compelling option for wavelength-shifting in gases when it can be driven in the impulsive regime.

The temporal behavior [Figs. 6(a) and 6(b)] of Raman-induced index change can play an important role in Raman processes. We continue with the example of the SSFS. In gases with ultra-narrowband Raman responses, only pulses in the impulsive regime have the required bandwidth for SSFS to occur. The Raman soliton continues to red-shift due to the rising Raman-induced index change. Moreover, the soliton duration increases and its bandwidth decreases due to the soliton area theorem.198 The SSFS significantly slows down when the Raman soliton becomes long enough that the Raman-induced index change starts to catch up with the pulse, and the nonlinear-phase-induced red-shifting ceases. As a result, the Raman-soliton duration is approximately bounded above by 1–2 times the duration of the first spike of the Raman-induced index change. In this situation, the Raman soliton can also become too narrowband for intrapulse SSFS to occur. In H2 where the dominant S(1) rotational transition has TR = 57.8 fs [Fig. 6(a)], the soliton duration quickly evolves to 63 fs when the index starts to follow the pulse [Fig. 8(a)]. Similarly in N2, where the first spike is 118-fs-long [Fig. 6(b)], the soliton duration quickly evolves to 183 fs [Fig. 8(b)]. A slow increase of the duration follows due to a weakly delayed Raman-induced index change. In H2, the increase of the duration slows down also at its transition duration between two Raman regimes, where SSFS basically stops [Fig. 7(a)]. However, in N2, it slows down at a duration smaller than its transition duration of 500 fs [Fig. 7(b)]. Since the duration of temporal spikes of Raman response is inversely correlated with the collective spectral width of clustered Raman responses (as in N2 and O2), the ability of pulse-following is determined by not only the spectral positions of clustered Raman responses but also their collective spectral width. Higher frequencies and larger collective spectral width of clustered Raman responses correspond to a sharper temporal spike of the overall Raman response, leading to a shorter soliton duration after SSFS. On the contrary, since only one S(1) rotational transition dominates in H2, there is no such thing as collective spectral width and only the Raman frequency determines the temporal behavior of the Raman response that affects the Raman-soliton duration.

FIG. 8.

Evolutions of the (full-width-at-half-maximum) pulse duration during the SSFS process in a 100-m-long anti-resonant hollow-core fiber, filled with 10-bar (a) H2 or (b) N2. Black dashed lines are where the increase of the duration drops to 1% for every 1-m propagation. Insets are temporal profiles of the Raman-induced index changes (△ɛ; orange) and Raman solitons (blue) when the increase of the soliton duration significantly slows down (after black dashed lines). The injected linearly polarized fundamental solitons at 1030 nm have 10-fs duration, with 350-nJ energy in H2 and 270-nJ energy in N2 due to different dispersion and electronic nonlinearities.

FIG. 8.

Evolutions of the (full-width-at-half-maximum) pulse duration during the SSFS process in a 100-m-long anti-resonant hollow-core fiber, filled with 10-bar (a) H2 or (b) N2. Black dashed lines are where the increase of the duration drops to 1% for every 1-m propagation. Insets are temporal profiles of the Raman-induced index changes (△ɛ; orange) and Raman solitons (blue) when the increase of the soliton duration significantly slows down (after black dashed lines). The injected linearly polarized fundamental solitons at 1030 nm have 10-fs duration, with 350-nJ energy in H2 and 270-nJ energy in N2 due to different dispersion and electronic nonlinearities.

Close modal
One of the most important features of impulsive SRS is its ability to create long-lasting phonon waves [Figs. 6(a) and 6(b)], which allows for nonlocal interactions between temporally separated pulses. This phenomenon has been extensively studied in situations with either two ultrashort pulses in the impulsive regime91–95,101,131–143,200,201 or two long pulses in the transient regime.86–90 However, a general description of this process covering pulses of all time scales has not been reported. Investigation of multi-pulse interactions through phonons can provide deeper understanding of the underlying physics and help optimize applications such as ultrafast time-resolved spectroscopy and controllable wavelength conversion. The following discussion builds on our initial attempts toward the unified theory of SRS.90 From Eq. (17), the Raman response to a short pulse is
(19)
The strength of the excited phonon waves is governed by the power spectral density F | A 1 | 2 at −ωR. Even if the pulse is initially in the transient regime, the generated Stokes or anti-Stokes waves can become sufficiently strong to induce beating with the pump. Short pulses that result from the beating then impulsively excite phonon waves.73 This is the basis of a technique designed to avoid Raman spectral narrowing and generate femtosecond Stokes pulses with excitation pulses in the transient regime (Fig. 9).86–90 
FIG. 9.

Double-pulse approach to Raman generation that avoids Raman spectral narrowing. The first pulse excites the phonon waves, which scatter the second pulse starting with its leading edge so that the entire pulse undergoes Raman generation, rather than only the trailing edge that results in Raman spectral narrowing for a chirped pulse.

FIG. 9.

Double-pulse approach to Raman generation that avoids Raman spectral narrowing. The first pulse excites the phonon waves, which scatter the second pulse starting with its leading edge so that the entire pulse undergoes Raman generation, rather than only the trailing edge that results in Raman spectral narrowing for a chirped pulse.

Close modal
Phonons can interact with light pulses through multiple mechanisms. If the second pulse is weak and long such that it is in the transient Raman regime [Figs. 10(a) and 10(b)], it can undergo either Stokes or anti-Stokes generation depending on the wave-vector mismatch,
(20a)
(20b)
where β ( j ) i = β j P β j i is the difference of the propagation constants of the i (Stokes or anti-Stokes) waves in the jth pulse, βph is the wave vector of the excited phonon waves, and C is the strength of the phonon waves. [See Sec. 10 of the supplementary material for the derivation of Eq. (20).] This wave-vector-matching effect can be treated as phonon amplification for Stokes generation and phonon absorption for anti-Stokes generation.64,90 Phonon amplification is the strongest when the incoming phonons perfectly seed the process by having the same wave vector as the generated phonons, β ph = β ( 2 ) ph β ( 2 ) S . On the other hand, phonon absorption is most efficient if the incoming phonons satisfy β ph + β ( 2 ) A S = 0 . The response of the second pulse to the incoming phonons is linear owing to the absence of A 2 S / A S on the right-hand side of Eq. (20) and negligible pump depletion. If both wave-vector-matching relations are simultaneously satisfied, Stokes and anti-Stokes waves experience growth; this occurs for β ( 2 ) S + β ( 2 ) A S = β ( 2 ) = 0 . This observation highlights the fact that phonon waves play a linear driving role in the SRS process and thus can overcome (nonlinear) Raman gain suppression at △β(2) = 0. During the phonon-absorption process [ β ph + β ( 2 ) A S = 0 , but, in general, β ph β ( 2 ) S ], phonons can still, to some extent, drive the generated anti-Stokes waves back to the pump through Stokes scattering. This results in oscillatory energy exchange between the anti-Stokes and pump waves [Fig. 10(b)]. Furthermore, the beating of phonon waves resulting from multiple Raman processes in gases creates non-uniform back-conversion of the anti-Stokes pulse to the pump, which degrades the quality of the anti-Stokes pulse [Figs. 10(e) and 10(f)]. Since the second pulse is weak, one might expect the generated Raman pulses to exhibit up to ∼100% energy fluctuations.166,167 However, this is not observed in phonon amplification [Fig. 10(a)] and the initial evolution of phonon absorption [Fig. 10(b)]. Existing coherent phonons linearly drive the Raman generation process and effectively dominate noisy thermal phonons and vacuum fluctuations, which stabilizes Raman pulses. This process enables quantum-state-preserving frequency conversion,66 for example. On the other hand, significant fluctuations are observed in the latter evolution of phonon absorption. This results from oscillatory evolutions between the anti-Stokes and pump waves, caused by phonon-induced back-conversion. Since the temporal pattern of phonon beating is strongly influenced by quantum fluctuations, back-conversion induces significant fluctuations to the Raman processes. Things turn out differently if the second pulse is strong. In this situation, self-induced SRS dominates [Figs. 10(c) and 10(d)]. The phonon-driven Stokes wave acts as the seed for the latter SRS process [Figs. 10(c) and 10(g)], but the growth of the phonon-driven anti-Stokes wave is suppressed by the aforementioned back-conversion, and eventually, the Stokes wave dominates [Fig. 10(d)]. Despite the domination of the Stokes process under the condition of phonon absorption ( β ph + β ( 2 ) A S = 0 ) in the strong-pump situation, the non-uniform anti-Stokes back-conversion process in the initial process modifies the phase structure of the pump wave, leaving the latter-generated Stokes pulse temporally structured [Fig. 10(h)]. Unlike the low-energy regime where significant fluctuations exist in the phonon-absorption process, large nonlinearity of the self-induced SRS stabilizes evolutions and all participating pulses regardless of which phonon process to occur.
FIG. 10.

Effects of phonon dynamics in the transient regime. Two 50-fs pulses of different wavelengths are chirped to 10-ps duration and are launched into a 1-m-long capillary filled with 20-bar H2. Their wavelengths are chosen to yield the process of interest, with the varying first-pulse wavelength and fixed second-pulse 2-μm wavelength. Because of the unique S-shaped dispersion curve of gas-filled HCF, pumps of different wavelengths create different phonon wave vector βph.64,90 Evolutions of photon numbers of Stokes (red), pump (black), and anti-Stokes (blue) waves under different wave-vector-matching conditions are displayed for the cases of weak (10 μJ) and strong (5 mJ) second-pulse energy; the first pulse is fixed at 2 mJ. (a) and (c) satisfy the phonon amplification condition, where the first pulse is at 0.7 μm, while (b) and (d) satisfy the phonon absorption condition, where the first pulse is at 1.09 μm. The second pulse is weak in (a) and (b), while it is sufficiently strong to induce its own SRS in (c) and (d). Mean values (center lines) and 1σ standard deviations (shaded areas) are calculated from ten repeated simulations. (e) and (f) The temporal profiles of the dechirped anti-Stokes pulses from (b); in (f), the pulse quality is degraded by back-conversion. (f) The result from one of the ten simulations, which fluctuates significantly due to the fluctuating phonon-beating pattern. (g) and (h) The temporal profiles of the dechirped Stokes pulses from (c) and (d), respectively. The Stokes pulse in (g) has more than twice the peak power as that in (h). Despite being not shown here, the pulse quality of the dechirped Stokes pulse in (a) is the same as that in (g) from (c) (TL: transform-limited pulse, D: dechirped pulse).

FIG. 10.

Effects of phonon dynamics in the transient regime. Two 50-fs pulses of different wavelengths are chirped to 10-ps duration and are launched into a 1-m-long capillary filled with 20-bar H2. Their wavelengths are chosen to yield the process of interest, with the varying first-pulse wavelength and fixed second-pulse 2-μm wavelength. Because of the unique S-shaped dispersion curve of gas-filled HCF, pumps of different wavelengths create different phonon wave vector βph.64,90 Evolutions of photon numbers of Stokes (red), pump (black), and anti-Stokes (blue) waves under different wave-vector-matching conditions are displayed for the cases of weak (10 μJ) and strong (5 mJ) second-pulse energy; the first pulse is fixed at 2 mJ. (a) and (c) satisfy the phonon amplification condition, where the first pulse is at 0.7 μm, while (b) and (d) satisfy the phonon absorption condition, where the first pulse is at 1.09 μm. The second pulse is weak in (a) and (b), while it is sufficiently strong to induce its own SRS in (c) and (d). Mean values (center lines) and 1σ standard deviations (shaded areas) are calculated from ten repeated simulations. (e) and (f) The temporal profiles of the dechirped anti-Stokes pulses from (b); in (f), the pulse quality is degraded by back-conversion. (f) The result from one of the ten simulations, which fluctuates significantly due to the fluctuating phonon-beating pattern. (g) and (h) The temporal profiles of the dechirped Stokes pulses from (c) and (d), respectively. The Stokes pulse in (g) has more than twice the peak power as that in (h). Despite being not shown here, the pulse quality of the dechirped Stokes pulse in (a) is the same as that in (g) from (c) (TL: transform-limited pulse, D: dechirped pulse).

Close modal

Control of the Raman processes through the wave-vector matching relies on a fixed phonon wave vector βph. This assumption holds only when phonons are excited through Stokes generation with a long pulse in the transient regime, resulting in phonons with a fixed wave vector equal to the difference between the pump and the Stokes fields, β ph = β ( 1 ) S = β ( 1 ) P β ( 1 ) S [Fig. 11(a)]. If the excitation pulse is short enough to enter the impulsive regime, the phonon wave vector is no longer fixed, but is instead determined by the nonlinear red-shifting process [Fig. 11(b)]. Since the phonon wave vector is not constant, the wave-vector-matching condition cannot be consistently met throughout the evolution, which results in an uncontrolled scenario of Stokes and anti-Stokes generation from the second pulse. This knowledge not only directs us toward controllable Raman generation through nonlocal phonon interactions but also offers more information of various Raman processes previously investigated, such as the phonon amplification conducted by Bustard et al.202 If phonons are impulsively excited by an ultrashort pulse, the upcoming pulse for amplifying the phonons with transient SRS must be sufficiently strong to guarantee the Stokes generation.

FIG. 11.

Evolutions of phonon wave vector βph during the SRS process in two different Raman regimes. Black, red, and blue lines represent total wave vectors of pump, Stokes, and phonons, respectively. Since the total momentum should be conserved, phonon wave vector is the difference between the pump and the Stokes waves in the transient regime: βtotal = NPβP + NSβS + Nphβph, where NP + NS + Nph = Ntotal represents the number of photons/phonons. In the impulsive regime, no discrete Stokes wave is generated and the pump pulse shifts to the red with varying propagation constant: β total = N P β P + i = 1 N ph β ph , i , where NP + Nph = Ntotal. In (b), βP ∼ cos(NtotalNP) is assumed to artificially create a varying βP to simulate the red-shifting of the pump. Because of the red-shifting process, phonon wave vector is not a constant, in contrast to the transient regime.

FIG. 11.

Evolutions of phonon wave vector βph during the SRS process in two different Raman regimes. Black, red, and blue lines represent total wave vectors of pump, Stokes, and phonons, respectively. Since the total momentum should be conserved, phonon wave vector is the difference between the pump and the Stokes waves in the transient regime: βtotal = NPβP + NSβS + Nphβph, where NP + NS + Nph = Ntotal represents the number of photons/phonons. In the impulsive regime, no discrete Stokes wave is generated and the pump pulse shifts to the red with varying propagation constant: β total = N P β P + i = 1 N ph β ph , i , where NP + Nph = Ntotal. In (b), βP ∼ cos(NtotalNP) is assumed to artificially create a varying βP to simulate the red-shifting of the pump. Because of the red-shifting process, phonon wave vector is not a constant, in contrast to the transient regime.

Close modal

The story is different when the second pulse is in the impulsive Raman regime. In this situation, there is no discrete generation of Stokes or anti-Stokes fields by the second pulse. However, control of the delay between the two pulses permits for the control of the index change experienced by the second pulse. This can lead to controlled red-shifting through the rising index, as observed with R 1 in the discussed impulsive regime [Eq. (18)], or blue-shifting for the falling index.92–95 

Vector (i.e., polarization) effects in SRS from molecules have been largely neglected. Vibrational SRS is almost isotropic and thus exhibits good consistency between experiments and scalar models.203,204 The isotropy results from E r ̂ r ̂ E | E | 2 / 3 in the perturbative regime, where E and r ̂ are shown in Fig. 12. (Details are in Sec. 5 of the supplementary material.) Rotational SRS in gases requires a vector model due to its anisotropic nature arising from the exchange of angular momentum. Although there are a few prior works deriving the scalar model for rotational SRS in gases, they neglect its anisotropic nature.76,186,190,205–208 Here, we aim to address this gap by introducing a vector UPPE for gases.

FIG. 12.

Illustration of spherical coordinates, where the diatomic molecular axis is aligned along r ̂ and the field propagates ( k ) along z ̂ . The field is polarized in the xy-plane.

FIG. 12.

Illustration of spherical coordinates, where the diatomic molecular axis is aligned along r ̂ and the field propagates ( k ) along z ̂ . The field is polarized in the xy-plane.

Close modal

There is a long history of investigation of rotational SRS in gases. Since Yoshikawa and Imasaka proposed the idea of phase-locking an SRS-induced multi-frequency spectrum to produce femtosecond pulses,209 simultaneous generation of multiple orders of Stokes/anti-Stokes waves has received substantial attention.72,76,141,151–153,210–213 One of the promising approaches is through molecular phase modulation, which requires impulsive excitation of gas molecules through SRS.63,92,141,151–153,210,212–215 While vibrational SRS can potentially produce sub-femtosecond pulses, rotational SRS is usually preferred in experiments because the longer period makes it easier to reach the impulsive regime. In addition, it can fill the spectral gap between vibrational Stokes/anti-Stokes wavelengths to create a smooth supercontinuum76 for potential generation of a single ultrashort pulse. It has long been known that rotational SRS is sensitive to the polarization of the field. A circularly polarized field preferentially drives rotational transitions, while linearly polarized light preferentially drives vibrational transitions.84,203,211 How these two scattering processes compete under different polarization conditions, and how to quantify the outcomes, have remained unclear. Rotational SRS, particularly in N2, is commonly observed with a linearly polarized field,37–40,46,47,68,72,75–78,90,95,138,143,151,153,201,211,214 but how its anisotropic nature affects the polarization and the subsequent nonlinear dynamics is unknown.

Derivation of the vector UPPE begins with the (real-valued) polarization P R , N g rot , averaged over microscopic molecular orientations and summed over many molecules (Ng is the number density of gas molecules in 1/m3). Here, we consider only diatomic molecules to simplify the derivation (Fig. 12), but an analytic expression should also be possible for more-complicated molecules by following the same process.

Here, we outline the derivation of the rotational Raman response; details are included in Sec. 6 of the supplementary material. The derivation generalizes the perturbative density-matrix approach of Chen et al.207 by including the tensor polarizability α = α r ̂ r ̂ + α θ ̂ θ ̂ + ϕ ̂ ϕ ̂ of a diatomic molecule.127 With the identity relation r ̂ r ̂ + θ ̂ θ ̂ + ϕ ̂ ϕ ̂ = I , the Raman-induced polarization P R rot = α E from a single molecule becomes
(21)
where △α = αα is the polarizability anisotropy and E is the real-valued electric field. The SRS-induced change of populations in different energy levels can be obtained as a perturbation of the density matrix,216 
(22)
where the density matrix ρ = ρ(0) + ρ(1) with the unperturbed density matrix ρ(0) satisfying the Boltzmann distribution.217  Hint is the perturbed Hamiltonian of the system in the interaction picture. γab and ωab = ωaωb are the dephasing rate and frequency of the ab transitions, respectively. The solution to the density matrix allows for the determination of the microscopic-orientation-averaged SRS-induced polarization through
(23)
where the microscopic orientational averaging is applied over multiple rotational states.

The unperturbed part of Eq. (23) can be solved with a simple integral over solid angle. By observing that r ̂ r ̂ E ( 0 ) = sin 2 θ 2 ( 0 ) E = E 2 1 4 π sin 2 θ d Ω = E 3 , the unperturbed polarization P R rot ( 0 ) = α + 2 α 3 E corresponds to the linear susceptibility, which is usually replaced by a Sellmeier equation.218–221 

To solve the perturbed part of Eq. (23), we need ρ a b ( 1 ) , which leads us to first finding H int ( r , t ) ,
(24)
Therefore,
(25)
where ρ a b ( 0 ) = ρ a ( 0 ) δ a b and δab is the Kronecker delta function. Equation (22) then becomes
(26)
where E r ̂ r ̂ E plays the most important role as it determines the interaction between the gas molecule and the field. While E r ̂ r ̂ E = | E | 2 / 3 + r ̂ r ̂ , where r ̂ r ̂ | E | 2 / 3 , the deviation from the equilibrium value | E | 2 / 3 results in rotational SRS. The small perturbation explains why vibrational SRS weakly depends on the molecular orientation and is thus isotropic. Prior works that assume linearly polarized fields obtain a simple | E | 2 cos 2 ( π 2 θ ) dependence for E r ̂ r ̂ E ,68,207 where θ is the angle between the molecule r ̂ and the z axis, and the field is polarized in the xy-plane. The cos 2 ( π 2 θ ) dependence underlies what is referred to as “molecular alignment” in response to an external field.75,77,138,143,190,201,207 However, the response becomes complicated when E is not linearly polarized. Similarly, we need r ̂ r ̂ E to solve Eq. (23). After a lengthy derivation, we obtain
(27a)
(27b)
in which B = E + + E * . The analytic signal of the real-valued electric field E = 1 2 E + c.c. is E = E + ε ̂ + + E ε ̂ , where ε ̂ ± = x ̂ ± i y ̂ 2 is the basis of circular polarization. C0, C−2, and C2 are constants with spherical harmonics, Y 2 0 ( θ ) , Y 2 2 ( θ , ϕ ) , and Y 2 2 ( θ , ϕ ) , respectively. Note that | E | 2 = 1 2 | B | 2 1 2 | E | 2 .
With ρ(1) from Eq. (26), we can solve for the orientational average of the perturbed part of Eq. (23), r ̂ r ̂ E ( 1 ) ,
(28)
which includes the product of J M | E r ̂ r ̂ E | J M and J M | r ̂ r ̂ E | J M [Eq. (26)]. With the help of Gaunt coefficients222,223 and the Racah formula,185,224 we obtain the following relation for Cμ (μ = 0, ±2) [Eq. (27)], which is used to solve for the product:
(29)
where σ = 1/9 if μ = ν = 0 or 2/3 if |μ| = |ν| = 2. C′ is a constant in a term that will not be considered further because only off-diagonal terms contribute to SRS. Using the fact that ρ a b ( 1 ) r ̂ r ̂ E b a * = ρ b a ( 1 ) r ̂ r ̂ E a b in the summation and with Eqs. (27)(29), we obtain the analytic signal of P R , N g rot = 1 2 P R , N g rot + c.c. ,
(30)
where R J ( t ) = Θ ( t ) e γ 2 t sin ( ω J + 2 , J t ) . The dephasing rate γJ+2,J = γ2 is assumed to be the same for all transitions.
Equation (30) is the vector rotational Raman response, but it is not in a form that provides straightforward information about its isotropy and anisotropy. In addition, its formulation with circularly polarized fields necessitates a vector transformation if the field is not circularly polarized, hindering us from straightforward understanding of the nonlinear dynamics of different polarizations. With the help of Eq. (1b), we can identify the isotropic and anisotropic parts of the rotational response [Eq. (31)]. Including the contributions of both the vibrational68 and rotational SRS, the final Raman response functions in the vector UPPE [Eq. (1)] become
(31a)
where
(31b)
d α d Q 0 is the polarizability derivative at equilibrium, and μ is the reduced mass of the molecule (μ = m/2 for homonuclear diatomic molecules, where m is the atomic mass). γ 2 vib and γ 2 rot are the dephasing rates of vibrational and rotational transitions, respectively. ω ν 2 , J 2 ; ν 1 , J 1 = ω ν 2 , J 2 ω ν 1 , J 1 is the angular frequency of the transition between states (ν1, J1) and (ν2, J2). ν and J are the vibrational and rotational quantum numbers, respectively.
To facilitate the subsequent analysis, we explicitly show the nonlinear terms under circular [Eq. (32a)] and linear bases [Eq. (32b)]. For a linearly polarized field [Ay = 0 in Eq. (32b)], the rotational Raman response is 4Rrot, which has a coefficient 4 60 = 1 15 . This is consistent with the derivation of the scalar linearly polarized UPPE,68 
(32a)
(32b)
To understand the dynamics of the vector Raman response and the underlying competition among different scattering processes, we derive and interpret the vector Raman gain, just as was done previously in the scalar case. The vector field A = A 1 e ̂ 1 + A 2 e ̂ 2 is assumed to be
(33a)
(33b)
where the pump is polarized along e ̂ 2 . To simplify the analysis, we consider only the most-widely used circular and linear bases. Assuming weak pump depletion, dominant terms include two factors of A 2 P or combinations of its complex conjugate [ A 2 P A 2 P , A 2 P A 2 P * , or A 2 P * A 2 P * ]. Inspection of Eq. (32) reveals that the governing equations of the Stokes/anti-Stokes waves, A i k (i = 1 or 2, k = S or AS), contain only terms with A 2 P A 2 P A 1 * and A 2 P A 2 P * A 1 for the evolution of A 1 k , while they contain only |A2|2A2 for the evolution of A 2 k . Thus, the pump transfers energy only to Stokes and anti-Stokes waves that both have the same polarization as the pump or both have the orthogonal polarization to the pump. For example, an x-polarized pump does not generate or couple with x-polarized Stokes and y-polarized anti-Stokes simultaneously through SRS or FWM.
Since the evolutions of Stokes/anti-Stokes waves co-polarized with the pump have been thoroughly investigated in the scalar situation, here we examine the cross-polarized case. From Eq. (33), the coupled equation of the Stokes wave A 1 S , cross-polarized with the pump A 2 P , is
(34)
where again β(0) = β(1) = 0. The Langevin function consists of positive- and negative-frequency parts Γ 12 b = Γ ̃ 12 R b / 2 = Γ 12 b ( Ω < 0 ) e i β S z + ω t + Γ 12 b ( Ω > 0 ) e i β A S z ω t . R a ; 1 and R b ; 2 ; i , j , k are defined similarly to Eq. (8) with the subscripts P and S/AS corresponding to A 2 P and A 1 S / A S , respectively. The pump is governed by
(35)
For a circularly polarized field, κ 1 = 4 3 , κ2 = 0, r1 = 1, r 2 = 1 2 , r3 = 1, r4 = 0, κ 3 = 2 3 , r5 = 1, and r 6 = 1 2 ; for a linearly polarized field, κ 1 = 2 3 , κ 2 = 1 3 , r1 = 1, r2 = 0, r 3 = 1 2 , r 4 = 1 2 , κ3 = 1, r5 = 1, and r6 = 1. The process used to derive the scalar Raman gain yields the steady-state cross-polarized Raman gain,
(36a)
(36b)
where R b ; ss is defined analogously to Eq. (9b) for the anisotropic Raman response function Rb, and
(37)
The steady-state cross-linearly polarized Raman gain asymptotically approaches 1 2 κ ω S Im R b ; ss A 2 P 2 as |△β| becomes large [Eq. (36b)], half of what can be achieved in cross-circularly polarized Raman gain [Eq. (36a)]. Unlike co-polarized situations where both isotropic and anisotropic Raman responses play an important role in the co-polarized Raman gain, only the anisotropic Raman response contributes to cross-polarized Raman gain.
To gain insights into Eq. (36b), we assume that ωRωP to obtain
(38)

where β = β + 2 κ ω P 2 κ e 3 A 2 P 2 + R b ; 1 . The gain shape and its evolution with varying imaginary part of R b ; ss are basically the same as those found in the scalar scenario (Fig. 2), but translated to negative △β (Fig. 13). Perfect Raman gain suppression occurs at β = 2 κ ω P 2 κ e 3 A 2 P 2 + R b ; 1 = 2 κ ω P 2 κ e 3 + R b coeff ω R b A 2 P 2 , which depends on pump wavelength and peak power.

FIG. 13.

Raman gain shape gcross-linear [Eq. (38)] vs wave-vector mismatch △β. The arrow indicates the trend of increasing Im R b ; ss 2 (blue → red → black → green → pink) with fixed Re κ e 3 + R b ; ss 2 . The thick black line is the Raman gain with equal magnitudes of real and imaginary parts of the quantity κ e 3 + R b ; ss 2 .

FIG. 13.

Raman gain shape gcross-linear [Eq. (38)] vs wave-vector mismatch △β. The arrow indicates the trend of increasing Im R b ; ss 2 (blue → red → black → green → pink) with fixed Re κ e 3 + R b ; ss 2 . The thick black line is the Raman gain with equal magnitudes of real and imaginary parts of the quantity κ e 3 + R b ; ss 2 .

Close modal

The Raman gain for the cross-circularly polarized case behaves distinctly from all other cases: it is unaffected by FWM [Eq. (36a)] because only phase-modulation terms, O ( A 2 P 2 A 1 S ) , appear in its evolution equation. In contrast to the other Raman gains, cross-circularly polarized Raman generation cannot be FWM-suppressed. This effect has been experimentally observed117,118,121 and modeled with a different theoretical formalism that considers only the rotational Raman response.116,120 In gases, wave-vector mismatch is usually small and the Raman gain is generally reduced substantially. Without a means to suppress it, the cross-circularly polarized Raman gain can be 100 times stronger than the Raman gain for other combinations of polarizations (Fig. 14).

FIG. 14.

Transient S(1) rotational Raman gain of H2 at the pulse trailing edge for co-circularly polarized [Eq. (14)] and cross-circularly polarized [Eq. (40b)] pump and Stokes waves. A 1.5-mJ and 1-ps pulse is launched into a capillary under the same conditions as in Fig. 4. To calculate the Raman gain, the Raman-induced index change must include contributions from all Raman responses in the sum u R u coeff ω R u , which appears in R 1 and Re R tr , while only one is required in I m R tr due to the narrowband feature of Raman response functions. Suppression of the co-polarized gain occurs at 2.04 μm, where △β = 0. The ratio of the gain magnitudes is 90 at 1030 nm.

FIG. 14.

Transient S(1) rotational Raman gain of H2 at the pulse trailing edge for co-circularly polarized [Eq. (14)] and cross-circularly polarized [Eq. (40b)] pump and Stokes waves. A 1.5-mJ and 1-ps pulse is launched into a capillary under the same conditions as in Fig. 4. To calculate the Raman gain, the Raman-induced index change must include contributions from all Raman responses in the sum u R u coeff ω R u , which appears in R 1 and Re R tr , while only one is required in I m R tr due to the narrowband feature of Raman response functions. Suppression of the co-polarized gain occurs at 2.04 μm, where △β = 0. The ratio of the gain magnitudes is 90 at 1030 nm.

Close modal
With the Raman-gain formulas, we can calculate the ratio of Raman gains for different combinations of pump and Stokes polarizations, with negligible FWM. Although the actual Raman gain is affected by the parametric gain-suppression effect, we can gain physical intuition by knowing the approximate gain values without the suppression effect (i.e., at huge wave-vector mismatch). For co-polarized SRS, the Raman gain without FWM g ( | β | 1 ) | I m R | [Eqs. (10) and (14)], while for cross-polarized SRS, the ratio of Raman gains without FWM g + ( | β | 1 ) : g x y ( | β | 1 ) = Im R b : 1 2 Im R b [Eq. (36)]. As a result, with Eq. (32), we obtain
(39)
The ratio of rotational Raman gains is 1:4:6:3, which is consistent with prior works.118,119 This consistency also supports the newly derived rotational Raman response [Eq. (31)].
For the transient regime, we can obtain the cross-polarized Raman gain by substituting R b ; tr for R b ; ss A 2 P 2 ,
(40a)
(40b)
where
(41)

Overall, the gain behaves similarly to the steady-state gain but with the differences pointed out in the scalar case, such as the varying gain shape throughout the pulse and the coherent properties.

Cross-polarized SRS can play an important role in experiments, with behavior that deviates from the predictions of a scalar model. As an example, for a 300-μm-core capillary, filled with H2 to 20-bar pressure and exposed to a 1.5-mJ and 1-ps pump pulse at 1030 nm, the cross-linearly polarized rotational Raman gain is ten times stronger than the vibrational Raman gain [Fig. 15(a)]. This finding challenges the scalar prediction that vibrational SRS dominates. To further examine this phenomenon, we investigated the Raman gain for varying pulse duration and energy. Vibrational SRS exceeds the co-linearly polarized rotational SRS, as predicted by scalar calculations.90 However, the scalar calculation does not foresee the dominance of cross-polarized rotational SRS for wide ranges of pulse duration and energy [Fig. 15(b)]. The distinct Raman-suppression conditions for cross-linearly polarized (which depends on the pump peak power and wavelength) and co-polarized SRS (which occurs at △β = 0) underlie the possible control for dominant vibrational SRS [Fig. 15(c)] or cross-linearly polarized rotational SRS [Fig. 15(d)].

FIG. 15.

Vector aspects of Raman gain. (a) Transient Raman gain at the pulse trailing edge for S(1) rotational and Q(1) vibrational SRS co-polarized with a linearly polarized pump [Eq. (14)], as well as for S(1) rotational SRS cross-polarized with it [Eq. (40a)]. An x-polarized, 1.5-mJ and 1-ps pulse is launched into a capillary filled with H2 to a pressure of 20 bar. The inset shows the variation of co-polarized Raman gains. (b) Ratio of vibrational transient Raman gain to cross-polarized rotational transient Raman gain at 1030 nm. The vibrational and rotational gains are equal along the black dashed line. Stars indicate conditions where (c) the vibrational SRS (0.3 mJ, 8 ps) or (d) the cross-polarized rotational SRS (1.5 mJ, 1 ps) dominate. PSD: power spectral density (spectral intensity) of the pulse. The lengths of the capillary in (c) and (d) are 1 m and 10 cm, respectively, chosen to produce enough Raman generation for visualization.

FIG. 15.

Vector aspects of Raman gain. (a) Transient Raman gain at the pulse trailing edge for S(1) rotational and Q(1) vibrational SRS co-polarized with a linearly polarized pump [Eq. (14)], as well as for S(1) rotational SRS cross-polarized with it [Eq. (40a)]. An x-polarized, 1.5-mJ and 1-ps pulse is launched into a capillary filled with H2 to a pressure of 20 bar. The inset shows the variation of co-polarized Raman gains. (b) Ratio of vibrational transient Raman gain to cross-polarized rotational transient Raman gain at 1030 nm. The vibrational and rotational gains are equal along the black dashed line. Stars indicate conditions where (c) the vibrational SRS (0.3 mJ, 8 ps) or (d) the cross-polarized rotational SRS (1.5 mJ, 1 ps) dominate. PSD: power spectral density (spectral intensity) of the pulse. The lengths of the capillary in (c) and (d) are 1 m and 10 cm, respectively, chosen to produce enough Raman generation for visualization.

Close modal

The generation of a cross-polarized Stokes or anti-Stokes wave can naturally produce depolarization of a pulse in propagation. For a linearly polarized pulse, the growth of the cross-linearly polarized Raman pulses is slow due to gain suppression. As a consequence, the linearly polarized pulse can reasonably maintain its polarization [Fig. 16(a)]. In contrast, the growth of cross-circularly polarized waves is unimpeded and is unaffected by FWM-induced gain suppression. In N2, with its numerous low-frequency transitions [Fig. 6(c)], SRS becomes an intrapulse effect and ultimately produces significant depolarization [Fig. 16(b)].

FIG. 16.

Depolarization effects in SRS. A 100-μJ and 300-fs pulse at 1030 nm is launched into a 2-m-long capillary, filled with N2 to a pressure of 3 bar. Output spectra of (a) linearly and (b) circularly polarized input pulses. PSD: power spectral density (spectral intensity) of the pulse. The inset shows the initial growth of the circularly polarized Stokes pulse in ε ̂ cross-polarized with the pump.

FIG. 16.

Depolarization effects in SRS. A 100-μJ and 300-fs pulse at 1030 nm is launched into a 2-m-long capillary, filled with N2 to a pressure of 3 bar. Output spectra of (a) linearly and (b) circularly polarized input pulses. PSD: power spectral density (spectral intensity) of the pulse. The inset shows the initial growth of the circularly polarized Stokes pulse in ε ̂ cross-polarized with the pump.

Close modal
Vector effects are manifested not only in the Raman gain but also in nonlocal interactions mediated by phonon waves. The vector nature of the interaction offers a new degree of freedom, the polarization, for controlling Raman generation through phonon amplification or absorption processes [Eq. (20)]. Due to the different vector properties of vibrational and rotational SRS, selective excitation of one scattering process becomes possible. To investigate this effect, we assume that the second of the two pulses is weak and focus on its “linear” response to the phonon waves while ignoring its self-induced SRS. In addition, to simplify the analysis, the first pulse is polarized either in x ̂ or in ε ̂ + : A 1 = A 1 , x x ̂ or A 1 , + ε ̂ + . From Eq. (32), the Raman term of the UPPE for the second pulse A 2 becomes
(42a)
(42b)

These equations illustrate that when excited by a circularly polarized pulse, the phonon waves exhibit an isotropic behavior in relation to the second pulse [Eq. (42b)]. In contrast, if the excitation pulse is linearly polarized, the phonon waves are anisotropic for the second pulse. In this situation, the second pulse experiences a twofold increase in the strength of the rotational phonon waves, along with a π phase delay, when it is co-polarized as opposed to when it is cross-polarized [Eq. (42a)]. In addition, the impact of rotational phonon waves is the weakest when they are excited by a circularly polarized field. The polarization dependence that results from these effects can be exploited to drive vibrational or rotational transitions. As an explicit example, an intense ultrashort pulse was launched into a H2-filled capillary to impulsively excite rotational phonon waves, followed by a weak delayed pulse. The results exhibit the twofold enhancement in rotational SRS for the second pulse if the excitation pulse is linearly polarized [Fig. 17(a)], with no significant difference if it is circularly polarized [Fig. 17(b)].

FIG. 17.

Vector effects on propagating rotational phonon waves. Two pulses are launched into a capillary filled with H2. The first pulse drives the S(1) rotational transition impulsively, and the second pulse scatters from the phonon waves. The energy of the first pulse is 50 μJ in (a) and 200 μJ in (b) to compensate for the four times weaker rotational SRS in circular polarization. The output spectrum of the second pulse is shown for the cases where the first pulse is (a) linearly polarized ( x ̂ ) or (b) circularly polarized ( ε ̂ + ) . The second pulse is either x-polarized or y-polarized. PSD: power spectral density (spectral intensity) of the pulse. In (b), the two polarizations do not have exactly equal magnitudes because both the Stokes and anti-Stokes waves grow from noise, and the wave-vector matching is not specifically controlled [Eq. (20)].

FIG. 17.

Vector effects on propagating rotational phonon waves. Two pulses are launched into a capillary filled with H2. The first pulse drives the S(1) rotational transition impulsively, and the second pulse scatters from the phonon waves. The energy of the first pulse is 50 μJ in (a) and 200 μJ in (b) to compensate for the four times weaker rotational SRS in circular polarization. The output spectrum of the second pulse is shown for the cases where the first pulse is (a) linearly polarized ( x ̂ ) or (b) circularly polarized ( ε ̂ + ) . The second pulse is either x-polarized or y-polarized. PSD: power spectral density (spectral intensity) of the pulse. In (b), the two polarizations do not have exactly equal magnitudes because both the Stokes and anti-Stokes waves grow from noise, and the wave-vector matching is not specifically controlled [Eq. (20)].

Close modal

Femtosecond wavelength conversion is investigated to extend various ultrafast science applications beyond natural lasing wavelengths. As just one example, nonlinear microscopy requires high peak power at 1300 or 1700 nm for three-photon imaging.225,226 Raman generation in gas-filled HCF is a promising candidate for the generation of wavelengths from the ultraviolet to the mid-infrared regimes. In addition, it enables power scaling due to low nonlinearity of gases, high damage threshold, and low propagation loss.26,227 However, current research is mostly restricted to pulse durations from hundreds of picoseconds to a few nanoseconds, where SRS can dominate over competing nonlinearities. A relatively small number of works have targeted ultrafast (femtosecond) Raman generation.42,48,57,68,78,86–90,128–130,228 Detrimental competition from the Kerr nonlinearity can be avoided by appropriately stretching an ultrashort pulse in time (Fig. 4) and removing the frequency chirp from the generated Raman pulse, albeit with Raman spectral narrowing. Although femtosecond Raman generation without Raman spectral narrowing has been achieved in a two-pulse approach, it might also be possible with a simpler single-pulse approach through interference with a continuous-wave field.53 Additionally, Raman spectral narrowing in a single-pulse approach can be overcome by operating in the temporal regime where Raman-induced pulse compression occurs (Fig. 4). Gas-based optical parametric amplification (OPA) is another candidate for femtosecond wavelength conversion.229,230 Raman-enhanced SPM can potentially boost the OPA process. It also provides a new degree of freedom for controlling the process through the pulse duration, which, for example, stabilizes OPA against variations of pulse duration in a regime where increasing the duration leads to competition from the Kerr nonlinearity; the effects of lower peak power can be offset by stronger SPM enhancement.

With a deeper understanding of SRS in N2 and O2, gas-based nonlinear optics may advance toward air photonics, first envisioned in the terahertz regime.231,232 To date, most complex physical phenomena, such as phonon dynamics, have been studied in H2 due to its simple Raman response.63–65,87,90,92–94,101,153,210,214,233,234 On the other hand, investigations of N2 and O2 allow for progress toward photonics based on ambient-air-filled HCFs that can be handled the same as solid-core fibers. Prior studies in N2 or air include spectral broadening75–77,127,208 and the resulting ultrashort pulse generation,78,79 phonon-driven spectral control143,200 and transient Raman generation,228 soliton compression235,236 and self-frequency shift.189 Only recently has vibrational Stokes generation in N2 at around 1.3–1.4 μm been realized in the nanosecond regime.62 Optimization and extension of these studies into various temporal regimes with varying pulse energies, as well as under ambient pressure, will be interesting. As an example, pulse compression in ambient-air-filled hollow-core fiber covering a broad range of pulse duration and energy can potentially offer the advantages of a waveguide (e.g., alignment-free operation and compactness) over multipass cells237–242 and multiplate compressors.243–246 Most current gas-filled pulse compressors use inert gases to avoid SRS.239–242,247–252 However, with specific combinations of pulse energy and duration, it will be possible to exploit SPM enhancement from SRS or to generate ultrashort spatially stabilized pulses through the formation of Raman-induced multidimensional solitary states.79,80,82 It will be straightforward to extend the theoretical framework of this Perspective to these other gas-based nonlinear optical platforms.

The gas pressure is a degree of freedom with HCF. This enables novel physical phenomena based on longitudinally varying pressure, which is not realizable in free-space or solid-core-fiber platforms. Self-focusing can be prevented with a negative pressure gradient, and ionization-induced defocusing can be prevented with a positive pressure gradient.253,254 Moreover, pressure gradients can extend the physics of tapered solid-core fiber192,255 to gas-filled HCF. As an example, SSFS with a positive pressure gradient exhibits several advantages over constant pressure. A flattened anomalous dispersion profile during SSFS suppresses pulse temporal broadening through the soliton area theorem Ns = γtp/|β2|, if the pulse experiences constant anomalous dispersion (|β2|) despite red-shifting, due to higher gas pressure. Increasing nonlinearity can compensate the loss of the Raman process; furthermore, it can reduce the pulse duration by maintaining Ns = 1. These effects combined might help push SSFS performance beyond what can be achieved with constant pressure.68,69 Moreover, the excitation of higher-order modes, which has been observed in recent SSFS experiments,69 at the fiber input can be suppressed. It is worth noting that although peak-power-induced ionization can be suppressed by gradient pressures, scaling of average power is ultimately limited by thermal effects in Raman processes,256 which can be resolved by reducing the repetition rate or working with pulse bursts.

Regarding technological developments, gas-filled HCF may have impact beyond research laboratories if they are pressurized and sealed, so gas-handling apparatus is not required.257 Such fiberized cells may enhance SRS by splicing both ends to fiber Bragg gratings to form resonant cavities.258,259 With continuous-wave light, 99.99% of the output light at the Stokes wavelength has been obtained,259 and Stokes waves at 1.7 μm have been generated in H2260 or D2.50 Beyond the previous single-pass structure, an all-fiber gas Raman oscillator has been recently implemented to produce 1.8 W of continuous Stokes waves at 1693 nm through rotational SRS in H2.261 With further development of the fusion splicing technology,262–268 a simple and monolithic gas-based all-fiber system that supports the aforementioned SRS phenomena will be possible.

Recent years have witnessed a surge of effort in the field of quantum information science.269–273 Quantum frequency conversion (QFC) serves the role of connecting quantum nodes274,275 that operate at diverse wavelengths. It has been achieved through three-wave mixing in quadratic (χ(2)) nonlinear crystals276–280 and through Bragg-scattering FWM in solid-core photonic crystal fibers.281–283 Although solid-core fibers overcome several challenges in crystals, such as stringent phase-matching requirements and undesirable optical background, further improvements have recently been reached with a H2-filled hollow-core fiber.66 In hydrogen, long-lasting phonon waves are established through SRS, which can frequency up-convert a signal pulse with proper wave-vector matching. The theoretical framework presented above will facilitate the optimization of phonon-driven nonlocal interactions and competing Raman gains. Temporal overlap of the pump and signal pulses may be unnecessary, both frequency down- and up-conversions are possible by changing gas pressure,284 and controllable use of either rotational or vibrational SRS for different amounts of frequency shift is possible. To date, QFC in gases is limited to hydrogen, which offers both the huge transition frequency and simple Raman response [Fig. 6(c)];66,284 implementations with other gases may be desirable for various reasons. Ultimately, we foresee a promising future where ultra-tunable QFC with hollow-core fiber, filled with a wide range of gases, is achieved.

In optical quantum communication, it is crucial to create photonic quantum states with a controlled degree of entanglement and preserved coherence among the modes over long-distance transmission. Among the four degrees of freedom available for encoding a photon (polarization, two transverse spatial dimensions, and time/frequency), time-frequency encoding not only spans an unbounded high-dimensional Hilbert space but is also compatible with existing single-mode waveguide platforms. During studies of spontaneous Raman scattering, fluctuations of the generated Stokes pulse were explained as resulting from the number of excited statistically independent “coherent temporal modes,”146,147 or so-called “time-frequency Schmidt modes.”285 Recently, temporal modes have garnered increasing interest in quantum information science because they constitute an orthogonal broadband wave-packet/pulsed basis and enable the use of the time/frequency degree of freedom.286–301 Photon pairs are generated in various platforms, such as bulk nonlinear crystals,288 solid-core photonic crystal fibers,302–306 and tapered,307 birefringent,308 dispersion-shifted,309–311 or highly nonlinear312 fibers and waveguides.313,314 However, they are limited by spontaneous Raman scattering, which creates uncorrelated Raman photons.309–311,314 Liquid-filled315,316 and Xe-filled317,318 HCFs have been used for Raman-free photon-pair generation. As demonstrated for quantum frequency conversion by Tyumenev et al.,66 Raman interactions in gases are controllable and can exhibit (quantum-state-preserving) coherence in the transient regime. Raman interactions within gas-filled HCFs present an opportunity to avoid the imposition of adverse noisy effects on quantum communication and conversion. For example, phonon-driven processes might be considered for controllable coherent generation and conversion of temporal modes. Aforementioned ambient-air-filled or sealed pressurized HCF is a promising platform for the simplest operation. All-fiber operations based on both solid-core and hollow-core fibers might be possible with a fiberized quantum pulse gate and pulse shaper.293,300,301 As HCF proves its utility in achieving high-fidelity and low-latency single-photon transmission319 and photon-pair generation,317,318 the exploration of SRS in gases holds the promise of unveiling quantum applications based on temporal modes in the future.

In this Perspective, we have only considered forward SRS. However, a Stokes pulse can be generated in the opposite direction from the pump pulse. The counter-propagating Stokes pulse can not only extract energy throughout the pump pulse but also experiences strong temporal compression.320–322 This has been observed in several liquids,320,321,323–325 methane,326–328 and D2329 in free-space geometries. Investigations of backward SRS have also been conducted in solid-core silica fiber330–333 and H2-filled HCF.334–338 The transient self-similar nature of the evolution, which exhibits in forward SRS as well,182–184 and the corresponding periodically modulated Stokes pulse have been studied.335,336 Prior works have relied on a seeded process, but only recently has noise-initiated backward SRS been demonstrated to dominate the Raman process.337 This occurs when the pump spectral linewidth is much smaller than the Raman linewidth so that the forward and backward Raman gains are comparable. Equivalently, this corresponds to the steady-state Raman regime if the pulse is transform-limited (Fig. 1). The temporal phase profile of the backward-SRS Stokes pulse can be retained in silica fiber, as manifested by successful dechirping of the Stokes pulse from ∼23 ps to 500 fs.333 These observations suggest that backward SRS will have similar coherence properties as forward SRS, which motivates extension of the model presented in this Perspective to the backward situation. To date, most of the investigations of backward SRS have been confined to the steady-state Raman regime. However, the transient regime, as previously shown, can unveil diverse, unexplored, coherent physical phenomena. As demonstrated by Abdolvand et al. in a H2-filled HCF, transient amplification through backward SRS is possible.334 This observation raises questions about the phase relation between the pump and Stokes waves within the transient regime of backward SRS. For example, if either wave is produced by chirping a femtosecond pulse, can a femtosecond Stokes pulse be generated through backward SRS (after dechirped)? An affirmative answer could lead to more-efficient Raman generation without Raman spectral narrowing than the forward situation. Due to the dominance of forward Raman gain, it will be a challenge to demonstrate backward SRS in the transient regime, which might be overcome with gas-filled HCF.

The analysis of the vector properties of SRS in gases sheds insights on the evolution of the polarizations of the interacting fields. On the other hand, preservation of the polarization is essential to various applications, such as interferometric sensors,339–342 frequency metrology,343–345 and quantum communications.346–350 Hollow-core fibers with a symmetric structure can exhibit exceptional polarization purity, with up to 70-dB polarization extinction ratio,351 but only for linear pulse propagation. Nonlinear interactions, prevalent in ultrashort pulse propagation, can introduce significant polarization coupling through XPM, FWM, or SRS as illustrated in this Perspective. Highly birefringent hollow-core fiber preserves the polarization state by introducing a short beat length between polarization modes, effectively suppressing any polarization coupling through either linear or nonlinear interactions. Remarkably, it has achieved birefringence comparable to commercial solid-core polarization-maintaining fibers, accompanied by ultra-low loss.352–356 Further development of such fibers will be desirable for applications that must avoid polarization effects in Raman-active gases.

The vector physics with two polarization modes described here can be considered the initial phase of a broader exploration of multimode gas-based nonlinear optics. Multimode nonlinear effects have spurred numerous studies and potential applications. The trend began in solid-core fibers176,357–365 and has recently been extended to hollow-core fibers. Interested readers are referred to a recent review for multimode research.366 To date, multimode nonlinear effects in hollow-core fibers have been primarily studied in capillaries due to their large core size, where extreme red-shifted spectra,78 multidimensional solitary states,79–82 and few-cycle visible pulse generation367 have been observed. There are also recent developments in multimode hollow-core photonic crystal fibers,355,368–373 as well as the exploration of multimode nonlinear physics, such as circumvention of Raman gain suppression through multimode propagation.42,103 The unified theory covering all temporal regimes can be a starting point for understanding the potential complexities that arise from multimode interactions. As an example, we expect that the Raman-gain equations for co- and cross-polarized multimode fields will resemble those discussed here, but may exhibit unique gain-suppression relations. Intermodal SRS may play out differently in the different temporal regimes. As an example, the wave-vector-matching relations for intermodal nonlocal phonon interactions will be different from the single-mode versions. These differences may open new possibilities for controlling Raman scattering and inspire new applications.

The analytic theory described in this Perspective relies solely on two fundamental Raman integrals [Eq. (8)]. The theory helps elucidate a wide range of Raman phenomena, including Raman gain suppression in diverse regimes, nonlinear-phase-induced Raman suppression and Raman-pulse compression, and the interplay of Raman-enhanced SPM and its transitional behavior. In addition, interactions of pulses with Raman-generated phonons in various Raman regimes can be investigated. The vector model of Raman interactions presented here facilitates the exploration of nonlinear dynamics influenced by the anisotropic nature of SRS, such as cross-polarized Raman generation, cross-polarized Raman gain, and vector nonlocal interactions. The development of the theoretical framework presented here was motivated by current directions in the interaction of ultrashort pulses with Raman-active gases, and we believe that it will be a valuable tool for future investigations in this area. We hope that this Perspective will be instructive for individuals who are just beginning the investigation of Raman phenomena and will also foster a deeper understanding of Raman physics in experienced researchers, all with the goal of advancing the basic science and applications of Raman scattering.

See the supplementary material for supporting content.

This work was supported by the Office of Naval Research (Grant No. N00014-19-1-2592) and the National Institutes of Health (Grant Nos. R01EB033179 and U01NS128660). The authors acknowledge motivating discussions with Yishai Eisenberg.

The authors have no conflicts to disclose.

Yi-Hao Chen: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Investigation (lead); Software (lead); Validation (lead); Visualization (lead); Writing – original draft (lead). Frank Wise: Funding acquisition (lead); Supervision (lead); Writing – review & editing (lead).

The code used in this work has been made publicly available at https://github.com/AaHaHaa/gas_UPPE. It allows for modeling not only scalar and vector but also single-mode and multimode (transverse modes + polarization modes) situations, as well as in inert (Raman-inactive) and Raman-active gases with either a constant or gradient pressure. In addition, photoionization of a single-mode scalar field is supported. Scripts to generate simulation results in this paper are also included.

1.
A.
Smekal
, “
Zur Quantentheorie der Dispersion
,”
Naturwissenschaften
11
,
873
875
(
1923
).
2.
C. V.
Raman
and
K. S.
Krishnan
, “
A new type of secondary radiation
,”
Nature
121
,
501
502
(
1928
).
3.
C. V.
Raman
, “
A new radiation
,”
Indian J. Phys.
2
,
387
398
(
1928
), available at http://repository.ias.ac.in/70648/.
4.
C. V.
Raman
and
K. S.
Krishnan
, “
The optical analogue of the Compton effect
,”
Nature
121
,
711
(
1928
).
5.
G. S.
Landsherg
and
L. I.
Mandelstam
, “
New phenomenon in scattering of light (preliminary report)
,”
J. Russ. Phys.-Chem. Soc.
60
,
335
(
1928
), available at http://jetp.ras.ru/cgi-bin/dn/e_037_01_0001.pdf.
6.
G. S.
Landsherg
and
L. I.
Mandelstam
, “
Eine neue Erscheinung bei der Lichtzerstreuung in Krystallen
,”
Naturwissenschaften
16
,
557
558
(
1928
)
7.
G. S.
Landsherg
and
L. I.
Mandelstam
, “
Über die Lichtzerstreuung in Kristallen
,”
Z. Phys.
50
,
769
780
(
1928
)
8.
S.
Sri Kantha
, “
Raman’s prize
,”
Nature
340
,
672
(
1989
).
9.
I. L.
Fabelinskiĭ
, “
Priority and the Raman effect
,”
Nature
343
,
686
(
1990
).
10.
I. L.
Fabelinskiĭ
, “
Seventy years of combination (Raman) scattering
,”
Phys.-Usp.
41
,
1229
(
1998
).
11.
R.
Singh
and
F.
Riess
, “
The 1930 Nobel prize for physics: A close decision?
,”
Notes Rec. R. Soc. London
55
,
267
283
(
2001
).
12.
D.
Khvol’son O
,
Fizika Nashikh Dneĭ (Physics of Our Days)
, 4th ed. (
GTTI
,
Leningrad-Moscow
,
1932
).
13.
M. G.
Raymer
and
I. A.
Walmsley
, “
III The quantum coherence properties of stimulated Raman scattering
,” in
E.
Wolf
(Ed.)
Progress in Optics
,
28
(
Elsevier
,
1990
), pp.
181
270
.
14.
R. S.
Das
and
Y. K.
Agrawal
, “
Raman spectroscopy: Recent advancements, techniques and applications
,”
Vib. Spectrosc.
57
,
163
176
(
2011
).
15.
H. J.
Butler
,
L.
Ashton
,
B.
Bird
,
G.
Cinque
,
K.
Curtis
,
J.
Dorney
,
K.
Esmonde-White
,
N. J.
Fullwood
,
B.
Gardner
,
P. L.
Martin-Hirsch
,
M. J.
Walsh
,
M. R.
McAinsh
,
N.
Stone
, and
F. L.
Martin
, “
Using Raman spectroscopy to characterize biological materials
,”
Nat. Protoc.
11
,
664
687
(
2016
).
16.
R. R.
Jones
,
D. C.
Hooper
,
L.
Zhang
,
D.
Wolverson
, and
V. K.
Valev
, “
Raman techniques: Fundamentals and Frontiers
,”
Nanoscale Res. Lett.
14
,
231
(
2019
).
17.
A.
Orlando
,
F.
Franceschini
,
C.
Muscas
,
S.
Pidkova
,
M.
Bartoli
,
M.
Rovere
, and
A.
Tagliaferro
, “
A comprehensive review on Raman spectroscopy applications
,”
Chemosensors
9
,
262
(
2021
).
18.
A. J.
Stentz
, “
Applications of Raman lasers and amplifiers in fiber communication systems
,”
Proc. SPIE
3263
,
91
94
(
1998
).
19.
D.
Bayart
,
P.
Baniel
,
A.
Bergonzo
,
J.-Y.
Boniort
,
P.
Bousselet
,
L.
Gasca
,
D.
Hamoir
,
F.
Leplingard
,
A.
Le Sauze
,
P.
Nouchi
,
F.
Roy
, and
P.
Sillard
, “
Broadband optical fibre amplification over 17.7 THz range
,”
Electron. Lett.
36
,
1569
1571
(
2000
).
20.
M. N.
Islam
, “
Raman amplifiers for telecommunications
,”
IEEE J. Sel. Top. Quantum Electron.
8
,
548
559
(
2002
).
21.
O.
Boyraz
and
B.
Jalali
, “
Demonstration of 11 dB fiber-to-fiber gain in a silicon Raman amplifier
,”
IEICE Electron. Express
1
,
429
434
(
2004
).
22.
K.
Rottwitt
and
J. H.
Povlsen
, “
Analyzing the fundamental properties of Raman amplification in optical fibers
,”
J. Lightwave Technol.
23
,
3597
(
2005
).
23.
Y.
Feng
,
L.
Taylor
, and
D. B.
Calia
, “
Multiwatts narrow linewidth fiber Raman amplifiers
,”
Opt. Express
16
,
10927
10932
(
2008
).
24.
A. K.
Sridharan
,
J. E.
Heebner
,
M. J.
Messerly
,
J. W.
Dawson
,
R. J.
Beach
, and
C. P. J.
Barty
, “
Brightness enhancement in a high-peak-power cladding-pumped Raman fiber amplifier
,”
Opt. Lett.
34
,
2234
2236
(
2009
).
25.
G.
Agrawal
, “
Chapter 8—Stimulated Raman scattering
,” in
Nonlinear Fiber Optics
, 5th ed. (
Academic Press
,
Boston
,
2013
), pp.
295
352
.
26.
P. St. J.
Russell
, “
Photonic-crystal fibers
,”
J. Lightwave Technol.
24
,
4729
4749
(
2006
).
27.
J. C.
Travers
,
W.
Chang
,
J.
Nold
,
N. Y.
Joly
, and
P. St. J.
Russell
, “
Ultrafast nonlinear optics in gas-filled hollow-core photonic crystal fibers (Invited)
,”
J. Opt. Soc. Am. B
28
,
A11
A26
(
2011
).
28.
F.
Poletti
,
M. N.
Petrovich
, and
D. J.
Richardson
, “
Hollow-core photonic bandgap fibers: Technology and applications
,”
Nanophotonics
2
,
315
340
(
2013
).
29.
C.
Markos
,
J. C.
Travers
,
A.
Abdolvand
,
B. J.
Eggleton
, and
O.
Bang
, “
Hybrid photonic-crystal fiber
,”
Rev. Mod. Phys.
89
,
045003
(
2017
).
30.
B.
Debord
,
F.
Amrani
,
L.
Vincetti
,
F.
Gérôme
, and
F.
Benabid
, “
Hollow-Core fiber technology: The rising of ‘gas photonics
,’”
Fibers
7
,
16
(
2019
).
31.
R. W.
Minck
,
R. W.
Terhune
, and
W. G.
Rado
, “
Laser-stimulated Raman effect and resonant four-photon interactions in gases H2, D2, and CH4
,”
Appl. Phys. Lett.
3
,
181
184
(
1963
).
32.
E. E.
Hagenlocker
and
W. G.
Rado
, “
Stimulated Brillouin and Raman scattering in gases
,”
Appl. Phys. Lett.
7
,
236
238
(
1965
).
33.
G. G.
Bret
and
M. M.
Denariez
, “
Study of hydrogen stimulated Raman emission
,”
Phys. Lett.
22
,
583
585
(
1966
).
34.
T. R.
Loree
,
C. D.
Cantrell
, and
D. L.
Barker
, “
Stimulated Raman emission at 9.2 μm from hydrogen gas
,”
Opt. Commun.
17
,
160
162
(
1976
).
35.
A. Z.
Grasyuk
,
L. L.
Losev
,
D. N.
Nikogosyan
, and
A. A.
Oraevskiĭ
, “
Generation of single picosecond pulses of up to 0.6 mJ energy and of 9.2 μm wavelength by stimulated Raman scattering
,”
Sov. J. Quantum Electron.
14
,
1257
1258
(
1984
).
36.
F.
Benabid
,
J. C.
Knight
,
G.
Antonopoulos
, and
P. St. J.
Russell
, “
Stimulated Raman scattering in hydrogen-filled hollow-core photonic crystal fiber
,”
Science
298
,
399
402
(
2002
).
37.
Z.
Wang
,
F.
Yu
,
W. J.
Wadsworth
, and
J. C.
Knight
, “
Efficient 1.9 μm emission in H2-filled hollow core fiber by pure stimulated vibrational Raman scattering
,”
Laser Phys. Lett.
11
,
105807
(
2014
).
38.
A. V.
Gladyshev
,
A. N.
Kolyadin
,
A. F.
Kosolapov
,
Y. P.
Yatsenko
,
A. D.
Pryamikov
,
A. S.
Biryukov
,
I. A.
Bufetov
, and
E. M.
Dianov
, “
Efficient 1.9-μm Raman generation in a hydrogen-filled hollow-core fibre
,”
Quantum Electron.
45
,
807
812
(
2015
).
39.
A. V.
Gladyshev
,
A. N.
Kolyadin
,
A. F.
Kosolapov
,
Y. P.
Yatsenko
,
A. D.
Pryamikov
,
A. S.
Biriukov
,
I. A.
Bufetov
, and
E. M.
Dianov
, “
Low-threshold 1.9 μm Raman generation in microstructured hydrogen-filled hollow-core revolver fibre with nested capillaries
,”
Laser Phys.
27
,
025101
(
2016
).
40.
A. V.
Gladyshev
,
A. F.
Kosolapov
,
M. M.
Khudyakov
,
Y. P.
Yatsenko
,
A. N.
Kolyadin
,
A. A.
Krylov
,
A. D.
Pryamikov
,
A. S.
Biriukov
,
M. E.
Likhachev
,
I. A.
Bufetov
, and
E. M.
Dianov
, “
4.4-μm Raman laser based on hollow-core silica fibre
,”
Quantum Electron.
47
,
491
494
(
2017
).
41.
H.
Li
,
W.
Huang
,
Y.
Cui
,
Z.
Zhou
, and
Z.
Wang
, “
Pure rotational stimulated Raman scattering in H2-filled hollow-core photonic crystal fibers
,”
Opt. Express
28
,
23881
23897
(
2020
).
42.
S.
Loranger
,
P. St. J.
Russell
, and
D.
Novoa
, “
Sub-40 fs pulses at 1.8 μm and MHz repetition rates by chirp-assisted Raman scattering in hydrogen-filled hollow-core fiber
,”
J. Opt. Soc. Am. B
37
,
3550
3556
(
2020
).
43.
W.
Huang
,
Z.
Li
,
Y.
Cui
,
Z.
Zhou
, and
Z.
Wang
, “
Efficient, watt-level, tunable 1.7 μm fiber Raman laser in H2-filled hollow-core fibers
,”
Opt. Lett.
45
,
475
478
(
2020
).
44.
Y.
Wang
,
M. K.
Dasa
,
A. I.
Adamu
,
J. E.
Antonio-Lopez
,
M. S.
Habib
,
R.
Amezcua-Correa
,
O.
Bang
, and
C.
Markos
, “
High pulse energy and quantum efficiency mid-infrared gas Raman fiber laser targeting CO2 absorption at 4.2 μm
,”
Opt. Lett.
45
,
1938
1941
(
2020
).
45.
A. I.
Adamu
,
Y.
Wang
,
M. S.
Habib
,
M. K.
Dasa
,
J. E.
Antonio-Lopez
,
R.
Amezcua-Correa
,
O.
Bang
, and
C.
Markos
, “
Multi-wavelength high-energy gas-filled fiber Raman laser spanning from 1.53 μm to 2.4 μm
,”
Opt. Lett.
46
,
452
455
(
2021
).
46.
A. V.
Gladyshev
,
A. F.
Kosolapov
,
M. M.
Khudyakov
,
Y. P.
Yatsenko
,
A. K.
Senatorov
,
A. N.
Kolyadin
,
A. A.
Krylov
,
V. G.
Plotnichenko
,
M. E.
Likhachev
,
I. A.
Bufetov
, and
E. M.
Dianov
, “
Raman generation in 2.9–3.5 μm spectral range in revolver hollow-core silica fiber filled by H2/D2 mixture
,” in
Conference on Lasers and Electro-Optics
(
Optica Publishing Group
,
2017
), p.
STu1K.2
.
47.
Y.
Cui
,
W.
Huang
,
Z.
Li
,
Z.
Zhou
, and
Z.
Wang
, “
High-efficiency laser wavelength conversion in deuterium-filled hollow-core photonic crystal fiber by rotational stimulated Raman scattering
,”
Opt. Express
27
,
30396
30404
(
2019
).
48.
A. V.
Gladyshev
,
Y.
Yatsenko
,
A.
Kolyadin
,
V.
Kompanets
, and
I.
Bufetov
, “
Mid-infrared 10-μJ-level sub-picosecond pulse generation via stimulated Raman scattering in a gas-filled revolver fiber
,”
Opt. Mater. Express
10
,
3081
3089
(
2020
).
49.
H.
Li
,
W.
Huang
,
Y.
Cui
,
Z.
Zhou
, and
Z.
Wang
, “
3 W tunable 1.65 μm fiber gas Raman laser in D2-filled hollow-core photonic crystal fibers
,”
Opt Laser. Technol.
132
,
106474
(
2020
).
50.
W.
Pei
,
H.
Li
,
W.
Huang
,
M.
Wang
, and
Z.
Wang
, “
All-fiber gas Raman laser by D2-filled hollow-core photonic crystal fibers
,”
Photonics
8
,
382
(
2021
).
51.
Z.
Li
,
W.
Pei
,
H.
Li
,
W.
Huang
,
X.
Li
,
Z.
Wang
, and
J.
Chen
, “
D2-filled hollow-core fiber gas Raman laser at 2.15 μm
,”
Photonics
9
,
753
(
2022
).
52.
W.
Huang
,
Z.
Wang
,
Z.
Zhou
,
W.
Pei
,
Y.
Cui
,
X.
Li
,
Z.
Li
, and
J.
Chen
, “
Tunable fiber gas Raman laser of 6 W at 2.9 μm by deuterium-filled hollow-core fiber
,”
IEEE J. Sel. Top. Quantum Electron.
(published online)
(
2023
).
53.
A.
Chugreev
,
A.
Nazarkin
,
A.
Abdolvand
,
J.
Nold
,
A.
Podlipensky
, and
P. St. J.
Russell
, “
Manipulation of coherent Stokes light by transient stimulated Raman scattering in gas filled hollow-core PCF
,”
Opt. Express
17
,
8822
8829
(
2009
).
54.
Y.
Chen
,
Z.
Wang
,
Z.
Li
,
W.
Huang
,
X.
Xi
, and
Q.
Lu
, “
Ultra-efficient Raman amplifier in methane-filled hollow-core fiber operating at 1.5 μm
,”
Opt. Express
25
,
20944
20949
(
2017
).
55.
Z.
Li
,
W.
Huang
,
Y.
Cui
, and
Z.
Wang
, “
Efficient mid-infrared cascade Raman source in methane-filled hollow-core fibers operating at 2.8 μm
,”
Opt. Lett.
43
,
4671
4674
(
2018
).
56.
L.
Cao
,
S.-F.
Gao
,
Z.-G.
Peng
,
X.-C.
Wang
,
Y.-Y.
Wang
, and
P.
Wang
, “
High peak power 2.8 μm Raman laser in a methane-filled negative-curvature fiber
,”
Opt. Express
26
,
5609
5615
(
2018
).
57.
A. V.
Gladyshev
,
M. S.
Astapovich
,
Y. P.
Yatsenko
,
A. F.
Kosolapov
,
A. G.
Okhrimchuk
, and
I. A.
Bufetov
, “
SRS generation of femtosecond pulses in a methane-filled revolver hollow-core optical fibre
,”
Quantum Electron.
49
,
1089
(
2019
).
58.
W.
Huang
,
Y.
Cui
,
Z.
Li
,
Z.
Zhou
,
Y.
Chen
, and
Z.
Wang
, “
Diode-pumped single-pass tunable mid-infrared gas Raman source by methane-filled hollow-core fiber
,”
Laser Phys. Lett.
16
,
085107
(
2019
).
59.
A. M.
Lanari
,
H. C. H.
Mulvad
,
S. M.
Abokhamis Mousavi
,
I. A.
Davidson
,
Q.
Fu
,
P.
Horak
,
D. J.
Richardson
, and
F.
Poletti
, “
High power Raman second Stokes generation in a methane filled hollow core fiber
,”
Opt. Express
31
,
41191
41201
(
2023
).
60.
K.
Krupa
,
K.
Baudin
,
A.
Parriaux
,
G.
Fanjoux
, and
G.
Millot
, “
Intense stimulated Raman scattering in CO2-filled hollow-core fibers
,”
Opt. Lett.
44
,
5318
5321
(
2019
).
61.
Y.
Wang
,
O. T. S.
Schiess
,
R.
Amezcua-Correa
, and
C.
Markos
, “
CO2-based hollow-core fiber Raman laser with high-pulse energy at 1.95 μm
,”
Opt. Lett.
46
,
5133
5136
(
2021
).
62.
L.
Hong
,
C.
Zhang
,
J.
Wahlen
,
J. E.
Antonio-Lopez
,
R.
Amezcua-Correa
,
C.
Markos
, and
Y.
Wang
, “
High energy and narrow linewidth N2-filled hollow-core fiber laser at 1.4 μm
,” arXiv:2312.03581 (
2023
).
63.
S. T.
Bauerschmidt
,
D.
Novoa
,
B. M.
Trabold
,
A.
Abdolvand
, and
P. St. J.
Russell
, “
Supercontinuum up-conversion via molecular modulation in gas-filled hollow-core PCF
,”
Opt. Express
22
,
20566
20573
(
2014
).
64.
S. T.
Bauerschmidt
,
D.
Novoa
,
A.
Abdolvand
, and
P. St. J.
Russell
, “
Broadband-tunable LP01 mode frequency shifting by Raman coherence waves in a H2-filled hollow-core photonic crystal fiber
,”
Optica
2
,
536
539
(
2015
).
65.
M. K.
Mridha
,
D.
Novoa
,
P.
Hosseini
, and
P. St. J.
Russell
, “
Thresholdless deep and vacuum ultraviolet Raman frequency conversion in hydrogen-filled photonic crystal fiber
,”
Optica
6
,
731
734
(
2019
).
66.
R.
Tyumenev
,
J.
Hammer
,
N. Y.
Joly
,
P. St. J.
Russell
, and
D.
Novoa
, “
Tunable and state-preserving frequency conversion of single photons in hydrogen
,”
Science
376
,
621
624
(
2022
).
67.
D. G.
Ouzounov
,
F. R.
Ahmad
,
D.
Müller
,
N.
Venkataraman
,
M. T.
Gallagher
,
M. G.
Thomas
,
J.
Silcox
,
K. W.
Koch
, and
A. L.
Gaeta
, “
Generation of megawatt optical solitons in hollow-core photonic band-gap fibers
,”
Science
301
,
1702
(
2003
).
68.
Y.-H.
Chen
,
P.
Sidorenko
,
E.
Antonio-Lopez
,
R.
Amezcua-Correa
, and
F.
Wise
, “
Efficient soliton self-frequency shift in hydrogen-filled hollow-core fiber
,”
Opt. Lett.
47
,
285
288
(
2022
).
69.
F.
Tani
,
J.
Lampen
,
M.
Butryn
,
M. H.
Frosz
,
J.
Jiang
,
M. E.
Fermann
, and
P. St. J.
Russell
, “
Temporal self-compression and self-frequency shift of submicrojoule pulses at a repetition rate of 8 MHz
,”
Phys. Rev. Appl.
18
,
064069
(
2022
).
70.
M. F.
Saleh
and
F.
Biancalana
, “
Tunable frequency-up/down conversion in gas-filled hollow-core photonic crystal fibers
,”
Opt. Lett.
40
,
4218
4221
(
2015
).
71.
M. F.
Saleh
and
F.
Biancalana
, “
Ultra-broadband supercontinuum generation in gas-filled photonic-crystal fibers: The epsilon-near-zero regime
,”
Opt. Lett.
46
,
1959
1962
(
2021
).
72.
F.
Couny
,
F.
Benabid
,
P. J.
Roberts
,
P. S.
Light
, and
M. G.
Raymer
, “
Generation and photonic guidance of multi-octave optical-frequency combs
,”
Science
318
,
1118
1121
(
2007
).
73.
F.
Belli
,
A.
Abdolvand
,
W.
Chang
,
J. C.
Travers
, and
P. St. J.
Russell
, “
Vacuum-ultraviolet to infrared supercontinuum in hydrogen-filled photonic crystal fiber
,”
Optica
2
,
292
300
(
2015
).
74.
P.
Hosseini
,
A.
Ermolov
,
F.
Tani
,
D.
Novoa
, and
P. St. J.
Russell
, “
UV soliton dynamics and Raman-enhanced supercontinuum generation in photonic crystal fiber
,”
ACS Photonics
5
,
2426
2430
(
2018
).
75.
J. E.
Beetar
,
M.
Nrisimhamurty
,
T.-C.
Truong
,
G. C.
Nagar
,
Y.
Liu
,
J.
Nesper
,
O.
Suarez
,
F.
Rivas
,
Y.
Wu
,
B.
Shim
, and
M.
Chini
, “
Multioctave supercontinuum generation and frequency conversion based on rotational nonlinearity
,”
Sci. Adv.
6
,
eabb5375
(
2020
).
76.
S.-F.
Gao
,
Y.-Y.
Wang
,
F.
Belli
,
C.
Brahms
,
P.
Wang
, and
J. C.
Travers
, “
From Raman frequency combs to supercontinuum generation in nitrogen-filled hollow-core anti-resonant fiber
,”
Laser Photonics Rev.
16
,
2100426
(
2022
).
77.
G.
Fan
,
R.
Safaei
,
O.
Kwon
,
V.
Schuster
,
K.
Légaré
,
P.
Lassonde
,
A.
Ehteshami
,
L.
Arias
,
A.
Laramée
,
J.
Beaudoin-Bertrand
,
J.
Limpert
,
Z.
Tao
,
M.
Spanner
,
B. E.
Schmidt
,
H.
Ibrahim
,
A.
Baltuška
, and
F.
Légaré
, “
High energy redshifted and enhanced spectral broadening by molecular alignment
,”
Opt. Lett.
45
,
3013
3016
(
2020
).
78.
P. A.
Carpeggiani
,
G.
Coccia
,
G.
Fan
,
E.
Kaksis
,
A.
Pugžlys
,
A.
Baltuška
,
R.
Piccoli
,
Y.-G.
Jeong
,
A.
Rovere
,
R.
Morandotti
,
L.
Razzari
,
B. E.
Schmidt
,
A. A.
Voronin
, and
A. M.
Zheltikov
, “
Extreme Raman red shift: Ultrafast multimode nonlinear space-time dynamics, pulse compression, and broadly tunable frequency conversion
,”
Optica
7
,
1349
1354
(
2020
).
79.
R.
Safaei
,
G.
Fan
,
O.
Kwon
,
K.
Légaré
,
P.
Lassonde
,
B. E.
Schmidt
,
H.
Ibrahim
, and
F.
Légaré
, “
High-energy multidimensional solitary states in hollow-core fibres
,”
Nat. Photonics
14
,
733
739
(
2020
).
80.
L.
Arias
,
A.
Longa
,
G.
Jargot
,
A.
Pomerleau
,
P.
Lassonde
,
G.
Fan
,
R.
Safaei
,
P. B.
Corkum
,
F.
Boschini
,
H.
Ibrahim
, and
F.
Légaré
, “
Few-cycle Yb laser source at 20 kHz using multidimensional solitary states in hollow-core fibers
,”
Opt. Lett.
47
,
3612
3615
(
2022
).
81.
T.-C.
Truong
,
J. E.
Beetar
, and
M.
Chini
, “
Light-field synthesizer based on multidimensional solitary states in hollow-core fibers
,”
Opt. Lett.
48
,
2397
2400
(
2023
).
82.
M.
Kumar
,
M.
Arshadipirlar
,
R.
Safaei
,
H.
Ibrahim
, and
F.
Légaré
, “
Generating ultrashort visible light pulses based on multidimensional solitary states in gas-filled hollow core fiber
,”
APL Photonics
8
,
056104
(
2023
).
83.
N.
Bloembergen
, “
The stimulated Raman effect
,”
Am. J. Phys.
35
,
989
1023
(
1967
).
84.
D. A.
Long
,
The Raman Effect: A Unified Treatment of the Theory of Raman Scattering by Molecules
(
John Wiley & Sons, Ltd.
,
2002
).
85.
R. W.
Boyd
, “
Chapter 10—Stimulated Raman scattering and stimulated Rayleigh-wing scattering
,” in
Nonlinear Optics
, 3rd ed. (
Academic Press
,
Burlington
,
2008
), pp.
473
509
.
86.
A. V.
Konyashchenko
,
L. L.
Losev
, and
S. Y.
Tenyakov
, “
Raman frequency shifter for laser pulses shorter than 100 fs
,”
Opt. Express
15
,
11855
11859
(
2007
).
87.
A. V.
Konyashchenko
,
L. L.
Losev
,
V. S.
Pazyuk
, and
S. Y.
Tenyakov
, “
Frequency shifting of sub-100 fs laser pulses by stimulated Raman scattering in a capillary filled with pressurized gas
,”
Appl. Phys. B
93
,
455
461
(
2008
).
88.
N. V.
Didenko
,
A. V.
Konyashchenko
,
P. V.
Kostryukov
,
L. L.
Losev
,
V. S.
Pazyuk
,
S. Y.
Tenyakov
,
V. Y.
Molchanov
,
S. I.
Chizhikov
, and
K. B.
Yushkov
, “
40-fs hydrogen Raman laser
,”
Quantum Electron.
45
,
1101
1104
(
2015
).
89.
C.
Vicario
,
M.
Shalaby
,
A. V.
Konyashchenko
,
L. L.
Losev
, and
C. P.
Hauri
, “
High-power femtosecond Raman frequency shifter
,”
Opt. Lett.
41
,
4719
4722
(
2016
).
90.
Y.-H.
Chen
,
J.
Moses
, and
F.
Wise
, “
Femtosecond long-wave-infrared generation in hydrogen-filled hollow-core fiber
,”
J. Opt. Soc. Am. B
40
,
796
806
(
2023
).
91.
Y.-X.
Yan
and
K. A.
Nelson
, “
Impulsive stimulated light scattering. I. General theory
,”
J. Chem. Phys.
87
,
6240
6256
(
1987
).
92.
N.
Zhavoronkov
and
G.
Korn
, “
Generation of single intense short optical pulses by ultrafast molecular phase modulation
,”
Phys. Rev. Lett.
88
,
203901
(
2002
).
93.
M. F.
Saleh
,
A.
Armaroli
,
A.
Marini
, and
F.
Biancalana
, “
Strong Raman-induced noninstantaneous soliton interactions in gas-filled photonic crystal fibers
,”
Opt. Lett.
40
,
4058
4061
(
2015
).
94.
M. F.
Saleh
and
F.
Biancalana
, “
Soliton dynamics in gas-filled hollow-core photonic crystal fibers
,”
J. Opt.
18
,
013002
(
2015
).
95.
F.
Belli
,
A.
Abdolvand
,
J. C.
Travers
, and
P. St. J.
Russell
, “
Control of ultrafast pulses in a hydrogen-filled hollow-core photonic-crystal fiber by Raman coherence
,”
Phys. Rev. A
97
,
013814
(
2018
).
96.
S. E. J.
Bell
, “
Tutorial review. Time-resolved resonance Raman spectroscopy
,”
Analyst
121
,
107R
120R
(
1996
).
97.
D. W.
McCamant
,
P.
Kukura
, and
R. A.
Mathies
, “
Femtosecond time-resolved stimulated Raman spectroscopy: Application to the ultrafast internal conversion in β-carotene
,”
J. Phys. Chem. A
107
,
8208
8214
(
2003
).
98.
H.
Yan
,
D.
Song
,
K. F.
Mak
,
I.
Chatzakis
,
J.
Maultzsch
, and
T. F.
Heinz
, “
Time-resolved Raman spectroscopy of optical phonons in graphite: Phonon anharmonic coupling and anomalous stiffening
,”
Phys. Rev. B
80
,
121403
(
2009
).
99.
S. K.
Sahoo
,
S.
Umapathy
, and
A. W.
Parker
, “
Time-resolved resonance Raman spectroscopy: Exploring reactive intermediates
,”
Appl. Spectrosc.
65
,
1087
1115
(
2011
).
100.
H.
Kuramochi
and
T.
Tahara
, “
Tracking ultrafast structural dynamics by time-domain Raman spectroscopy
,”
J. Am. Chem. Soc.
143
,
9699
9717
(
2021
).
101.
M. F.
Saleh
,
A.
Armaroli
,
T. X.
Tran
,
A.
Marini
,
F.
Belli
,
A.
Abdolvand
, and
F.
Biancalana
, “
Raman-induced temporal condensed matter physics in gas-filled photonic crystal fibers
,”
Opt. Express
23
,
11879
11886
(
2015
).
102.
S. T.
Bauerschmidt
,
D.
Novoa
, and
P. St. J.
Russell
, “
Dramatic Raman gain suppression in the vicinity of the zero dispersion point in a gas-filled hollow-core photonic crystal fiber
,”
Phys. Rev. Lett.
115
,
243901
(
2015
).
103.
P.
Hosseini
,
M. K.
Mridha
,
D.
Novoa
,
A.
Abdolvand
, and
P. St. J.
Russell
, “
Universality of coherent Raman gain suppression in gas-filled broadband-guiding photonic crystal fibers
,”
Phys. Rev. Appl.
7
,
034021
(
2017
).
104.
C.-S.
Wang
, “
Theory of stimulated Raman scattering
,”
Phys. Rev.
182
,
482
494
(
1969
).
105.
R. L.
Carman
,
F.
Shimizu
,
C. S.
Wang
, and
N.
Bloembergen
, “
Theory of Stokes pulse shapes in transient stimulated Raman scattering
,”
Phys. Rev. A
2
,
60
72
(
1970
).
106.
J. N.
Elgin
and
T. B.
O’Hare
, “
Saturation effects in transient stimulated Raman scattering
,”
J. Phys. B: At. Mol. Phys.
12
,
159
168
(
1979
).
107.
M. G.
Raymer
and
J.
Mostowski
, “
Stimulated Raman scattering: Unified treatment of spontaneous initiation and spatial propagation
,”
Phys. Rev. A
24
,
1980
1993
(
1981
).
108.
M. D.
Duncan
,
R.
Mahon
,
L. L.
Tankersley
, and
J.
Reintjes
, “
Transient stimulated Raman amplification in hydrogen
,”
J. Opt. Soc. Am. B
5
,
37
52
(
1988
).
109.
G.
Hilfer
and
C. R.
Menyuk
, “
Stimulated Raman scattering in the transient limit
,”
J. Opt. Soc. Am. B
7
,
739
749
(
1990
).
110.
P.
Hosseini
,
D.
Novoa
,
A.
Abdolvand
, and
P. St. J.
Russell
, “
Enhanced control of transient Raman scattering using buffered hydrogen in hollow-core photonic crystal fibers
,”
Phys. Rev. Lett.
119
,
253903
(
2017
).
111.
N.
Bloembergen
and
Y. R.
Shen
, “
Coupling between vibrations and light waves in Raman laser media
,”
Phys. Rev. Lett.
12
,
504
507
(
1964
).
112.
C. T.
Hsieh
,
N. D.
Foltz
, and
C. W.
Cho
, “
Production of the stimulated Raman lines in H2 gas with a focused laser beam
,”
J. Opt. Soc. Am.
64
,
202
205
(
1974
).
113.
B. N.
Perry
,
P.
Rabinowitz
, and
D. S.
Bomse
, “
Stimulated Raman scattering with a tightly focused pump beam
,”
Opt. Lett.
10
,
146
148
(
1985
).
114.
M. D.
Duncan
,
R.
Mahon
,
J.
Reintjes
, and
L. L.
Tankersley
, “
Parametric Raman gain suppression in D2 and H2
,”
Opt. Lett.
11
,
803
805
(
1986
).
115.
A. P.
Hickman
and
W. K.
Bischel
, “
Theory of Stokes and anti-Stokes generation by Raman frequency conversion in the transient limit
,”
Phys. Rev. A
37
,
2516
2523
(
1988
).
116.
Y. A.
Il’inskii
and
V. D.
Taranukhin
, “
Characteristics of stimulated Raman scattering in the presence of spatial degeneracy of levels
,”
Sov. J. Quantum Electron.
4
,
224
(
1974
).
117.
V. S.
Averbakh
,
A. I.
Makarov
, and
V. I.
Talanov
, “
Stimulated Raman scattering on rotational and vibrational transitions in nitrogen gas
,”
Sov. J. Quantum Electron.
8
,
472
(
1978
).
118.
G. V.
Venkin
,
Y. A.
Il’inskiĭ
, and
G. M.
Mikheev
, “
Influence of the polarization of radiation on the energy characteristics and threshold of stimulated Raman scattering due to rotational transitions
,”
Sov. J. Quantum Electron.
15
,
395
(
1985
).
119.
M.
Rokni
and
A.
Flusberg
, “
Stimulated rotational Raman scattering in the atmosphere
,”
IEEE J. Quantum Electron.
22
,
1102
1108
(
1986
).
120.
A.
Flusberg
and
R.
Holmes
, “
Polarization dependence of gain in stimulated Raman scattering
,”
Phys. Rev. Lett.
58
,
2039
2042
(
1987
).
121.
K.
Leung
,
M.
Oron
,
D.
Klimek
,
R.
Holmes
, and
A.
Flusberg
, “
Observation of parametric gain suppression in rotational Raman transitions of N2 and H2
,”
Opt. Lett.
13
,
33
35
(
1988
).
122.
R.
Loudon
, “
The Raman effect in crystals
,”
Adv. Phys.
13
,
423
482
(
1964
).
123.
V. L.
Strizhevskii
,
V. V.
Obukhovskii
, and
A. M.
Panarin
, “
Theory of stimulated Raman scattering of light in anisotropic crystals
,”
Sov. Phys. JETP
32
,
908
913
(
1971
).
124.
S.
De Silvestri
,
J. G.
Fujimoto
,
E. P.
Ippen
,
E. B.
Gamble
,
L. R.
Williams
, and
K. A.
Nelson
, “
Femtosecond time-resolved measurements of optic phonon dephasing by impulsive stimulated Raman scattering in α-perylene crystal from 20 to 300 K
,”
Chem. Phys. Lett.
116
,
146
152
(
1985
).
125.
T. P.
Dougherty
,
G. P.
Wiederrecht
, and
K. A.
Nelson
, “
Impulsive stimulated Raman scattering experiments in the polariton regime
,”
J. Opt. Soc. Am. B
9
,
2179
2189
(
1992
).
126.
B.
Xu
,
N.
Mao
,
Y.
Zhao
,
L.
Tong
, and
J.
Zhang
, “
Polarized Raman spectroscopy for determining crystallographic orientation of low-dimensional materials
,”
J. Phys. Chem. Lett.
12
,
7442
7452
(
2021
).
127.
R. H.
Lehmberg
,
M. F.
Wolford
,
J. L.
Weaver
,
D.
Kehne
,
S. P.
Obenschain
,
D.
Eimerl
, and
J. P.
Palastro
, “
Stimulated rotational Raman scattering of arbitrarily polarized broadband light
,”
Phys. Rev. A
102
,
063530
(
2020
).
128.
A. V.
Konyashchenko
,
P. V.
Kostryukov
,
L. L.
Losev
, and
V. S.
Pazyuk
, “
Minimum possible laser pulse duration for SRS
,”
Quantum Electron.
47
,
593
596
(
2017
).
129.
A. V.
Konyashchenko
,
L. L.
Losev
, and
V. S.
Pazyuk
, “
SRS converter—Compressor of femtosecond ytterbium laser pulses
,”
Quantum Electron.
49
,
845
849
(
2019
).
130.
A. V.
Konyashchenko
,
L. L.
Losev
, and
V. S.
Pazyuk
, “
SRS of ultrashort laser pulses with nonlinear phase modulation
,”
Quantum Electron.
51
,
217
221
(
2021
).
131.
Y.-X.
Yan
,
E. B.
Gamble
,
K. A.
Nelson
, and
K. A.
Nelson
, “
Impulsive stimulated scattering: General importance in femtosecond laser pulse interactions with matter, and spectroscopic applications
,”
J. Chem. Phys.
83
,
5391
5399
(
1985
).
132.
Y.-X.
Yan
and
K. A.
Nelson
, “
Impulsive stimulated light scattering. II. Comparison to frequency-domain light-scattering spectroscopy
,”
J. Chem. Phys.
87
,
6257
6265
(
1987
).
133.
S.
Ruhman
,
A. G.
Joly
, and
K. A.
Nelson
, “
Time-resolved observations of coherent molecular vibrational motion and the general occurrence of impulsive stimulated scattering
,”
J. Chem. Phys.
86
,
6563
6565
(
1987
).
134.
S.
Ruhman
,
A. G.
Joly
,
B.
Kohler
,
L. R.
Williams
, and
K. A.
Nelson
, “
Intramolecular and intermolecular dynamics in molecular liquids through femtosecond time-resolved impulsive stimulated scattering
,”
Rev. Phys. Appl.
22
,
1717
1734
(
1987
).
135.
S.
Ruhman
,
B.
Kohler
,
A. G.
Joly
, and
K. A.
Nelson
, “
Molecular dynamics in liquids from femtosecond time-resolved impulsive stimulated scattering
,”
IEEE J. Quantum Electron.
24
,
470
481
(
1988
).
136.
A. M.
Weiner
,
D. E.
Leaird
,
G. P.
Wiederrecht
, and
K. A.
Nelson
, “
Femtosecond multiple-pulse impulsive stimulated Raman scattering spectroscopy
,”
J. Opt. Soc. Am. B
8
,
1264
1275
(
1991
).
137.
L.
Dhar
,
J. A.
Rogers
, and
K. A.
Nelson
, “
Time-resolved vibrational spectroscopy in the impulsive limit
,”
Chem. Rev.
94
,
157
193
(
1994
).
138.
M. Z.
Hoque
,
M.
Lapert
,
E.
Hertz
,
F.
Billard
,
D.
Sugny
,
B.
Lavorel
, and
O.
Faucher
, “
Observation of laser-induced field-free permanent planar alignment of molecules
,”
Phys. Rev. A
84
,
013409
(
2011
).
139.
G.
Korn
,
O.
Dühr
, and
A.
Nazarkin
, “
Observation of Raman self-conversion of fs-pulse frequency due to impulsive excitation of molecular vibrations
,”
Phys. Rev. Lett.
81
,
1215
1218
(
1998
).
140.
A.
Nazarkin
and
G.
Korn
, “
Raman self-conversion of femtosecond laser pulses and generation of single-cycle radiation
,”
Phys. Rev. A
58
,
R61
R64
(
1998
).
141.
A.
Nazarkin
,
G.
Korn
,
M.
Wittmann
, and
T.
Elsaesser
, “
Generation of multiple phase-locked Stokes and anti-Stokes components in an impulsively excited Raman medium
,”
Phys. Rev. Lett.
83
,
2560
2563
(
1999
).
142.
A. M.
Zheltikov
,
A. A.
Voronin
,
R.
Kienberger
,
F.
Krausz
, and
G.
Korn
, “
Frequency-tunable multigigawatt sub-half-cycle light pulses from coupled-state dynamics of optical solitons and impulsively driven molecular vibrations
,”
Phys. Rev. Lett.
105
,
103901
(
2010
).
143.
Y.
Mo
,
W.
Cao
,
Z.
Wang
,
X.
Sun
,
Q.
Zhang
, and
P.
Lu
, “
Spectral tuning of a broadband optical pulse via stimulated Raman scattering of a prealigned molecule
,”
Phys. Rev. A
105
,
043504
(
2022
).
144.
P. D.
Drummond
, “
Quantum theory of fiber-optics and solitons
,” in
Coherence and Quantum Optics VII
, edited by
J. H.
Eberly
,
L.
Mandel
, and
E.
Wolf
(
Springer US
,
Boston, MA
,
1996
), pp.
323
332
.
145.
J.
Mostowski
and
M. G.
Raymer
, “
The buildup of stimulated Raman scattering from spontaneous Raman scattering
,”
Opt. Commun.
36
,
237
240
(
1981
).
146.
M. G.
Raymer
,
I. A.
Walmsley
,
J.
Mostowski
, and
B.
Sobolewska
, “
Quantum theory of spatial and temporal coherence properties of stimulated Raman scattering
,”
Phys. Rev. A
32
,
332
344
(
1985
).
147.
M. G.
Raymer
,
Z. W.
Li
, and
I. A.
Walmsley
, “
Temporal quantum fluctuations in stimulated Raman scattering: Coherent-modes description
,”
Phys. Rev. Lett.
63
,
1586
1589
(
1989
).
148.
P. D.
Drummond
and
J. F.
Corney
, “
Quantum noise in optical fibers. I. Stochastic equations
,”
J. Opt. Soc. Am. B
18
,
139
152
(
2001
).
149.
S.
Lefrançois
and
F. W.
Wise
, “
Stimulated Raman scattering in chirped-pulse amplification: The role of vibrational dephasing
,” in
CLEO: 2011—Laser Applications to Photonic Applications
(
Optica Publishing Group
,
2011
), p.
CThU2
.
150.
N. J.
Everall
,
J. P.
Partanen
,
J. R. M.
Barr
, and
M. J.
Shaw
, “
Threshold measurements of stimulated Raman scattering in gases using picosecond KrF laser pulses
,”
Opt. Commun.
64
,
393
397
(
1987
).
151.
V.
Kalosha
and
J.
Herrmann
, “
Phase relations, quasicontinuous spectra and subfemtosecond pulses in high-order stimulated Raman scattering with short-pulse excitation
,”
Phys. Rev. Lett.
85
,
1226
1229
(
2000
).
152.
F. L.
Kien
,
K.
Hakuta
, and
A. V.
Sokolov
, “
Pulse compression by parametric beating with a prepared Raman coherence
,”
Phys. Rev. A
66
,
023813
(
2002
).
153.
A. V.
Sokolov
and
S. E.
Harris
, “
Ultrashort pulse generation by molecular modulation
,”
J. Opt. B: Quantum Semiclassical Opt.
5
,
R1
R26
(
2002
).
154.
F.
Tani
,
J. C.
Travers
, and
P. St. J.
Russell
, “
Multimode ultrafast nonlinear optics in optical waveguides: Numerical modeling and experiments in kagome photonic-crystal fiber
,”
J. Opt. Soc. Am. B
31
,
311
320
(
2014
).
155.
M.
Kolesik
and
J. V.
Moloney
, “
Nonlinear optical pulse propagation simulation: From Maxwell’s to unidirectional equations
,”
Phys. Rev. E
70
,
036604
(
2004
).
156.
F.
Poletti
and
P.
Horak
, “
Description of ultrashort pulse propagation in multimode optical fibers
,”
J. Opt. Soc. Am. B
25
,
1645
1654
(
2008
).
157.
A.
Couairon
,
E.
Brambilla
,
T.
Corti
,
D.
Majus
,
O.
de J Ramírez-Góngora
, and
M.
Kolesik
, “
Practitioner’s guide to laser pulse propagation models and simulation
,”
Eur. Phys. J. Spec. Top.
199
,
5
76
(
2011
).
158.
M.
Bache
,
M. S.
Habib
,
C.
Markos
, and
J.
Lægsgaard
, “
Poor-man’s model of hollow-core anti-resonant fibers
,”
J. Opt. Soc. Am. B
36
,
69
80
(
2019
).
159.
E. A. J.
Marcatili
and
R. A.
Schmeltzer
, “
Hollow metallic and dielectric waveguides for long distance optical transmission and lasers
,”
Bell Syst. Tech. J.
43
,
1783
1809
(
1964
).
160.
Q.
Lin
and
G. P.
Agrawal
, “
Raman response function for silica fibers
,”
Opt. Lett.
31
,
3086
3088
(
2006
).
161.
F. X.
Kärtner
,
D. J.
Dougherty
,
H. A.
Haus
, and
E. P.
Ippen
, “
Raman noise and soliton squeezing
,”
J. Opt. Soc. Am. B
11
,
1267
1276
(
1994
).
162.
J. M.
Dudley
,
G.
Genty
, and
S.
Coen
, “
Supercontinuum generation in photonic crystal fiber
,”
Rev. Mod. Phys.
78
,
1135
1184
(
2006
).
163.
J. A.
Giordmaine
and
W.
Kaiser
, “
Light scattering by coherently driven lattice vibrations
,”
Phys. Rev.
144
,
676
688
(
1966
).
164.
T.
von Foerster
and
R. J.
Glauber
, “
Quantum theory of light propagation in amplifying media
,”
Phys. Rev. A
3
,
1484
1511
(
1971
).
165.
B. D.
Strycker
, “
Time-ordering in Heisenberg’s equation of motion as related to spontaneous radiation
,”
Sci. Rep.
11
,
22177
(
2021
).
166.
M. G.
Raymer
,
K.
Rzażewski
, and
J.
Mostowski
, “
Pulse-energy statistics in stimulated Raman scattering
,”
Opt. Lett.
7
,
71
73
(
1982
).
167.
I. A.
Walmsley
and
M. G.
Raymer
, “
Observation of macroscopic quantum fluctuations in stimulated Raman scattering
,”
Phys. Rev. Lett.
50
,
962
965
(
1983
).
168.
M.
Lewenstein
, “
Fluctuations in the nonlinear regime of stimulated Raman scattering
,”
Z. Phys. B
56
,
69
75
(
1984
).
169.
J.
Eggleston
and
R.
Byer
, “
Steady-state stimulated Raman scattering by a multimode laser
,”
IEEE J. Quantum Electron.
16
,
850
853
(
1980
).
170.
M. H.
Frosz
, “
Validation of input-noise model for simulations of supercontinuum generation and rogue waves
,”
Opt. Express
18
,
14778
14787
(
2010
).
171.
E.
Genier
,
P.
Bowen
,
T.
Sylvestre
,
J. M.
Dudley
,
P.
Moselund
, and
O.
Bang
, “
Amplitude noise and coherence degradation of femtosecond supercontinuum generation in all-normal-dispersion fibers
,”
J. Opt. Soc. Am. B
36
,
A161
A167
(
2019
).
172.
E. A.
Golovchenko
,
P. V.
Mamyshev
,
A. N.
Pilipetskii
, and
E. M.
Dianov
, “
Mutual influence of the parametric effects and stimulated Raman scattering in optical fibers
,”
IEEE J. Quantum Electron.
26
,
1815
1820
(
1990
).
173.
F.
Vanholsbeeck
,
P.
Emplit
, and
S.
Coen
, “
Complete experimental characterization of the influence of parametric four-wave mixing on stimulated Raman gain
,”
Opt. Lett.
28
,
1960
1962
(
2003
).
174.
V.
Couderc
,
P.
Leproux
,
V.
Tombelaine
,
L.
Grossard
, and
A.
Barthélémy
, “
Raman cascade suppression by using wide band parametric conversion in large normal dispersion regime
,”
Opt. Express
13
,
8584
8590
(
2005
).
175.
P.
Tchofo Dinda
,
G.
Millot
, and
S.
Wabnitz
, “
Polarization switching of stimulated Raman scattering in optical fibers by dual-frequency pumping
,”
Opt. Lett.
22
,
1595
1597
(
1997
).
176.
H.
Pourbeyram
,
G. P.
Agrawal
, and
A.
Mafi
, “
Stimulated Raman scattering cascade spanning the wavelength range of 523 to 1750 nm using a graded-index multimode optical fiber
,”
Appl. Phys. Lett.
102
,
201107
(
2013
).
177.
M.
Huang
,
S.
Sun
,
T. S.
Saini
,
Q.
Fu
,
L.
Xu
,
D.
Wu
,
H.
Ren
,
L.
Shen
,
T. W.
Hawkins
,
J.
Ballato
, and
A. C.
Peacock
, “
Raman amplification at 2.2 μm in silicon core fibers with prospects for extended mid-infrared source generation
,”
Light Sci. Appl.
12
,
209
(
2023
).
178.
S.
Coen
,
D. A.
Wardle
, and
J. D.
Harvey
, “
Observation of non-phase-matched parametric amplification in resonant nonlinear optics
,”
Phys. Rev. Lett.
89
,
273901
(
2002
).
179.
A. S. Y.
Hsieh
,
G. K. L.
Wong
,
S. G.
Murdoch
,
S.
Coen
,
F.
Vanholsbeeck
,
R.
Leonhardt
, and
J. D.
Harvey
, “
Combined effect of Raman and parametric gain on single-pump parametric amplifiers
,”
Opt. Express
15
,
8104
8114
(
2007
).
180.
G.
Agrawal
, “
Chapter 10—Four-wave mixing
,” in
Nonlinear Fiber Optics
, 5th ed. (
Academic Press
,
Boston
,
2013
), pp.
397
456
.
181.
G. K.
Teal
and
G. E.
MacWood
, “
The Raman spectra of the isotopic molecules H2, HD, and D2
,”
J. Chem. Phys.
3
,
760
764
(
1935
).
182.
C. R.
Menyuk
,
D.
Levi
, and
P.
Winternitz
, “
Self-similarity in transient stimulated Raman scattering
,”
Phys. Rev. Lett.
69
,
3048
3051
(
1992
).
183.
A.
Nazarkin
,
A.
Abdolvand
,
A. V.
Chugreev
, and
P. St. J.
Russell
, “
Direct observation of self-similarity in evolution of transient stimulated Raman scattering in gas-filled photonic crystal fibers
,”
Phys. Rev. Lett.
105
,
173902
(
2010
).
184.
K.
Krupa
,
A.
Parriaux
,
G.
Millot
, and
G.
Fanjoux
, “
Self-similarity in transient stimulated Raman scattering in CO2-filled hollow-core fiber
,”
J. Opt. Soc. Am. B
40
,
637
644
(
2023
).
185.
A.
Böhm
and
M.
Loewe
,
Quantum Mechanics: Foundations and Applications, Theoretical and Mathematical Physics
(
Springer
,
Berlin, Heidelberg
,
2013
).
186.
E. T. J.
Nibbering
,
G.
Grillon
,
M. A.
Franco
,
B. S.
Prade
, and
A.
Mysyrowicz
, “
Determination of the inertial contribution to the nonlinear refractive index of air, N2, and O2 by use of unfocused high-intensity femtosecond laser pulses
,”
J. Opt. Soc. Am. B
14
,
650
660
(
1997
).
187.
S. C.
Pinault
and
M. J.
Potasek
, “
Frequency broadening by self-phase modulation in optical fibers
,”
J. Opt. Soc. Am. B
2
,
1318
1319
(
1985
).
188.
G.
Agrawal
, “
Chapter 4—Self-phase modulation
,” in
Nonlinear Fiber Optics
, 5th ed. (
Academic Press
,
Boston
,
2013
), pp.
87
128
.
189.
A. V.
Gorbach
and
D. V.
Skryabin
, “
Soliton self-frequency shift, non-solitonic radiation and self-induced transparency in air-core fibers
,”
Opt. Express
16
,
4858
4865
(
2008
).
190.
D.
Langevin
,
J. M.
Brown
,
M. B.
Gaarde
, and
A.
Couairon
, “
Determination of molecular contributions to the nonlinear refractive index of air for mid-infrared femtosecond laser-pulse excitation
,”
Phys. Rev. A
99
,
063418
(
2019
).
191.
J. P.
Gordon
, “
Theory of the soliton self-frequency shift
,”
Opt. Lett.
11
,
662
664
(
1986
).
192.
X.
Liu
,
C.
Xu
,
W. H.
Knox
,
J. K.
Chandalia
,
B. J.
Eggleton
,
S. G.
Kosinski
, and
R. S.
Windeler
, “
Soliton self-frequency shift in a short tapered air–silica microstructure fiber
,”
Opt. Lett.
26
,
358
360
(
2001
).
193.
B. R.
Washburn
,
S. E.
Ralph
,
P. A.
Lacourt
,
J. M.
Dudley
,
W. T.
Rhodes
,
R. S.
Windeler
, and
S.
Coen
, “
Tunable near-infrared femtosecond soliton generation in photonic crystal fibres
,”
Electron. Lett.
37
,
1510
1512
(
2001
).
194.
I. G.
Cormack
,
D. T.
Reid
,
W. J.
Wadsworth
,
J. C.
Knight
, and
P. St. J.
Russell
, “
Observation of soliton self-frequency shift in photonic crystal fibre
,”
Electron. Lett.
38
,
167
169
(
2002
).
195.
J. H.
Lee
,
J.
van Howe
,
C.
Xu
, and
X.
Liu
, “
Soliton self-frequency shift: Experimental demonstrations and applications
,”
IEEE J. Sel. Top. Quantum Electron.
14
,
713
723
(
2008
).
196.
K.
Wang
and
C.
Xu
, “
Tunable high-energy soliton pulse generation from a large-mode-area fiber and its application to third harmonic generation microscopy
,”
Appl. Phys. Lett.
99
,
071112
(
2011
).
197.
S. A.
Dekker
,
A. C.
Judge
,
R.
Pant
,
I.
Gris-Sánchez
,
J. C.
Knight
,
C. M.
de Sterke
, and
B. J.
Eggleton
, “
Highly-efficient, octave spanning soliton self-frequency shift using a specialized photonic crystal fiber with low OH loss
,”
Opt. Express
19
,
17766
17773
(
2011
).
198.
G. P.
Agrawal
, “
Chapter 5—Optical solitons
,” in
Nonlinear Fiber Optics
, 5th ed. (
Academic Press
,
Boston
,
2013
), pp.
129
191
.
199.
G.
Agrawal
, “
Chapter 12—Novel nonlinear phenomena
,” in
Nonlinear Fiber Optics
, 5th ed. (
Academic Press
,
Boston
,
2013
), pp.
497
552
.
200.
Y.
Kida
and
T.
Imasaka
, “
Four-wave optical parametric amplification in a Raman-active gas
,”
Photonics
2
,
933
945
(
2015
).
201.
Z.
Rodnova
,
T.
Saule
,
G.
Gibson
, and
C. A.
Trallero-Herrero
, “
Molecular alignment-assisted spectral broadening and shifting in the near-infrared with a recycled depleted pump from an optical parametric amplifier
,”
Opt. Express
31
,
42327
42337
(
2023
).
202.
P. J.
Bustard
,
B. J.
Sussman
, and
I. A.
Walmsley
, “
Amplification of impulsively excited molecular rotational coherence
,”
Phys. Rev. Lett.
104
,
193902
(
2010
).
203.
R. W.
Minck
,
E. E.
Hagenlocker
, and
W. G.
Rado
, “
Stimulated pure rotational Raman scattering in deuterium
,”
Phys. Rev. Lett.
17
,
229
231
(
1966
).
204.
L. D.
Landau
,
J. S.
Bell
,
M. J.
Kearsley
,
L. P.
Pitaevskii
,
E. M.
Lifshitz
, and
J. B.
Sykes
,
Electrodynamics of Continuous Media
,
Course of Theoretical Physics
(
Elsevier Science
,
2013
).
205.
W. E.
Martin
and
R. J.
Winfield
, “
Nonlinear effects on pulsed laser propagation in the atmosphere
,”
Appl. Opt.
27
,
567
577
(
1988
).
206.
A. M.
Zheltikov
, “
Raman response function of atmospheric air
,”
Opt. Lett.
32
,
2052
2054
(
2007
).
207.
Y.-H.
Chen
,
S.
Varma
,
A.
York
, and
H. M.
Milchberg
, “
Single-shot, space- and time-resolved measurement of rotational wavepacket revivals in H2, D2, N2, O2, and N2O
,”
Opt. Express
15
,
11341
11357
(
2007
).
208.
S. A.
Mousavi
,
H. C. H.
Mulvad
,
N. V.
Wheeler
,
P.
Horak
,
J.
Hayes
,
Y.
Chen
,
T. D.
Bradley
,
S.-u.
Alam
,
S. R.
Sandoghchi
,
E. N.
Fokoua
,
D. J.
Richardson
, and
F.
Poletti
, “
Nonlinear dynamic of picosecond pulse propagation in atmospheric air-filled hollow core fibers
,”
Opt. Express
26
,
8866
8882
(
2018
).
209.
S.
Yoshikawa
and
T.
Imasaka
, “
A new approach for the generation of ultrashort optical pulses
,”
Opt. Commun.
96
,
94
98
(
1993
).
210.
S. E.
Harris
and
A. V.
Sokolov
, “
Subfemtosecond pulse generation by molecular modulation
,”
Phys. Rev. Lett.
81
,
2894
2897
(
1998
).
211.
H.
Kawano
,
Y.
Hirakawa
, and
T.
Imasaka
, “
Generation of high-order rotational lines in hydrogen by four-wave Raman mixing in the femtosecond regime
,”
IEEE J. Quantum Electron.
34
,
260
268
(
1998
).
212.
A. V.
Sokolov
,
D. R.
Walker
,
D. D.
Yavuz
,
G. Y.
Yin
, and
S. E.
Harris
, “
Raman generation by phased and antiphased molecular states
,”
Phys. Rev. Lett.
85
,
562
565
(
2000
).
213.
E.
Sali
,
K. J.
Mendham
,
J. W. G.
Tisch
,
T.
Halfmann
, and
J. P.
Marangos
, “
High-order stimulated Raman scattering in a highly transient regime driven by a pair of ultrashort pulses
,”
Opt. Lett.
29
,
495
497
(
2004
).
214.
P. J.
Bustard
,
B. J.
Sussman
, and
I. A.
Walmsley
, “
Phase-stable molecular phase modulation
,”
Central Laser Facility Annual Report
,
2008
.
215.
R.
Tyumenev
,
D.
Novoa
, and
P.
St.
J. Russell
, “
Narrowband VUV light by molecular modulation in dual-pumped H2-filled hollow-core photonic crystal fiber
,” in
Conference on Lasers and Electro-Optics
(
Optica Publishing Group
, 2020), p.
FM2P.6
.
216.
R. W.
Boyd
, “
Chapter 3—Quantum-mechanical theory of the nonlinear optical susceptibility
,” in
Nonlinear Optics
, 3rd ed. (
Academic Press
,
Burlington
,
2008
), pp.
135
206
.
217.
R. K.
Pathria
,
Statistical Mechanics
(
Elsevier Science
,
2016
).
218.
W.
Sellmeier
, “
Ueber die durch die Aetherschwingungen erregten Mitschwingungen der Körpertheilchen und deren Rückwirkung auf die ersteren, besonders zur Erklärung der Dispersion und ihrer Anomalien
,”
Ann. Phys.
223
,
386
403
(
1872
).
219.
E. R.
Peck
and
B. N.
Khanna
, “
Dispersion of nitrogen
,”
J. Opt. Soc. Am.
56
,
1059
1063
(
1966
).
220.
E. R.
Peck
and
S.
Huang
, “
Refractivity and dispersion of hydrogen in the visible and near infrared
,”
J. Opt. Soc. Am.
67
,
1550
1554
(
1977
).
221.
J.
Zhang
,
Z. H.
Lu
, and
L. J.
Wang
, “
Precision refractive index measurements of air, N2, O2, Ar, and CO2 with a frequency comb
,”
Appl. Opt.
47
,
3143
3151
(
2008
).
222.
J. A.
Gaunt
and
R. H.
Fowler
, “
IV. The triplets of helium
,”
Philos. Trans. R. Soc. London
228
,
151
196
(
1929
).
223.
E.
Merzbacher
,
Quantum Mechanics
(
Wiley
,
1998
).
224.
G.
Racah
, “
Theory of complex spectra. II
,”
Phys. Rev.
62
,
438
462
(
1942
).
225.
D. G.
Ouzounov
,
T.
Wang
,
M.
Wang
,
D. D.
Feng
,
N. G.
Horton
,
J. C.
Cruz-Hernández
,
Y.-T.
Cheng
,
J.
Reimer
,
A. S.
Tolias
,
N.
Nishimura
, and
C.
Xu
, “
In vivo three-photon imaging of activity of GCaMP6-labeled neurons deep in intact mouse brain
,”
Nat. Methods
14
,
388
390
(
2017
).
226.
D. M.
Chow
,
D.
Sinefeld
,
K. E.
Kolkman
,
D. G.
Ouzounov
,
N.
Akbari
,
R.
Tatarsky
,
A.
Bass
,
C.
Xu
, and
J. R.
Fetcho
, “
Deep three-photon imaging of the brain in intact adult zebrafish
,”
Nat. Methods
17
,
605
608
(
2020
).
227.
F.
Poletti
, “
Nested antiresonant nodeless hollow core fiber
,”
Opt. Express
22
,
23807
23828
(
2014
).
228.
A. V.
Konyashchenko
,
P. V.
Kostryukov
,
L. L.
Losev
, and
V. S.
Pazyuk
, “
Femtosecond compressed-nitrogen Raman laser
,”
Quantum Electron.
47
,
1
4
(
2017
).
229.
D.
Faccio
,
A.
Grün
,
P. K.
Bates
,
O.
Chalus
, and
J.
Biegert
, “
Optical amplification in the near-infrared in gas-filled hollow-core fibers
,”
Opt. Lett.
34
,
2918
2920
(
2009
).
230.
O.
Zurita-Miranda
,
C.
Fourcade-Dutin
,
F.
Fauquet
,
F.
Darracq
,
J.-P.
Guillet
,
P.
Mounaix
,
H.
Maillotte
, and
D.
Bigourd
, “
Tunable ultrafast infrared generation in a gas-filled hollow-core capillary by a four-wave mixing process
,”
J. Opt. Soc. Am. B
39
,
662
670
(
2022
).
231.
J.
Dai
,
B.
Clough
,
I.-C.
Ho
,
X.
Lu
,
J.
Liu
, and
X.-C.
Zhang
, “
Recent progresses in terahertz wave air photonics
,”
IEEE Trans. Terahertz Sci. Technol.
1
,
274
281
(
2011
).
232.
B.
Clough
,
J.
Dai
, and
X.-C.
Zhang
, “
Laser air photonics: Beyond the terahertz gap
,”
Mater. Today
15
,
50
58
(
2012
).
233.
S. E.
Harris
and
A. V.
Sokolov
, “
Broadband spectral generation with refractive index control
,”
Phys. Rev. A
55
,
R4019
R4022
(
1997
).
234.
R.
Tyumenev
,
P. St. J.
Russell
, and
D.
Novoa
, “
Narrowband vacuum ultraviolet light via cooperative Raman scattering in dual-pumped gas-filled photonic crystal fiber
,”
ACS Photonics
7
,
1989
1993
(
2020
).
235.
M.
Maurel
,
F.
Amrani
,
B.
Debord
,
F.
Gérôme
, and
F.
Benabid
, “
Soliton self-compression in air filled kagome HCPCF
,” in
2019 Conference on Lasers and Electro-Optics Europe and European Quantum Electronics Conference
(
Optica Publishing Group
,
2019
).
236.
J.
Luan
,
P. St. J.
Russell
, and
D.
Novoa
, “
Efficient self-compression of ultrashort near-UV pulses in air-filled hollow-core photonic crystal fibers
,”
Opt. Express
29
,
13787
13793
(
2021
).
237.
J.
Schulte
,
T.
Sartorius
,
J.
Weitenberg
,
A.
Vernaleken
, and
P.
Russbueldt
, “
Nonlinear pulse compression in a multi-pass cell
,”
Opt. Lett.
41
,
4511
4514
(
2016
).
238.
M.
Hanna
,
X.
Délen
,
L.
Lavenu
,
F.
Guichard
,
Y.
Zaouter
,
F.
Druon
, and
P.
Georges
, “
Nonlinear temporal compression in multipass cells: Theory
,”
J. Opt. Soc. Am. B
34
,
1340
1347
(
2017
).
239.
L.
Lavenu
,
M.
Natile
,
F.
Guichard
,
Y.
Zaouter
,
X.
Delen
,
M.
Hanna
,
E.
Mottay
, and
P.
Georges
, “
Nonlinear pulse compression based on a gas-filled multipass cell
,”
Opt. Lett.
43
,
2252
2255
(
2018
).
240.
M.
Hanna
,
F.
Guichard
,
N.
Daher
,
Q.
Bournet
,
X.
Délen
, and
P.
Georges
, “
Nonlinear optics in multipass cells
,”
Laser Photonics Rev.
15
,
2100220
(
2021
).
241.
M.
Müller
,
J.
Buldt
,
H.
Stark
,
C.
Grebing
, and
J.
Limpert
, “
Multipass cell for high-power few-cycle compression
,”
Opt. Lett.
46
,
2678
2681
(
2021
).
242.
L.
Silletti
,
A.
bin Wahid
,
E.
Escoto
,
P.
Balla
,
S.
Rajhans
,
K.
Horn
,
L.
Winkelmann
,
V.
Wanie
,
A.
Trabattoni
,
C. M.
Heyl
, and
F.
Calegari
, “
Dispersion-engineered multi-pass cell for single-stage post-compression of an ytterbium laser
,”
Opt. Lett.
48
,
1842
1845
(
2023
).
243.
C.-H.
Lu
,
W.-H.
Wu
,
S.-H.
Kuo
,
J.-Y.
Guo
,
M.-C.
Chen
,
S.-D.
Yang
, and
A. H.
Kung
, “
Greater than 50 times compression of 1030 nm Yb:KGW laser pulses to single-cycle duration
,”
Opt. Express
27
,
15638
15648
(
2019
).
244.
S.
Zhang
,
Z.
Fu
,
B.
Zhu
,
G.
Fan
,
Y.
Chen
,
S.
Wang
,
Y.
Liu
,
A.
Baltuska
,
C.
Jin
,
C.
Tian
, and
Z.
Tao
, “
Solitary beam propagation in periodic layered Kerr media enables high-efficiency pulse compression and mode self-cleaning
,”
Light Sci. Appl.
10
,
53
(
2021
).
245.
G.
Barbiero
,
H.
Wang
,
M.
Graßl
,
S.
Gröbmeyer
,
D.
Kimbaras
,
M.
Neuhaus
,
V.
Pervak
,
T.
Nubbemeyer
,
H.
Fattahi
, and
M. F.
Kling
, “
Efficient nonlinear compression of a thin-disk oscillator to 8.5 fs at 55 W average power
,”
Opt. Lett.
46
,
5304
5307
(
2021
).
246.
M.
Seidel
,
P.
Balla
,
C.
Li
,
G.
Arisholm
,
L.
Winkelmann
,
I.
Hartl
, and
C. M.
Heyl
, “
Factor 30 pulse compression by hybrid multipass multiplate spectral broadening
,”
Ultrafast Sci.
2022
,
2022
.
247.
A. I.
Adamu
,
M. S.
Habib
,
C. R.
Petersen
,
J. E. A.
Lopez
,
B.
Zhou
,
A.
Schülzgen
,
M.
Bache
,
R.
Amezcua-Correa
,
O.
Bang
, and
C.
Markos
, “
Deep-UV to mid-IR supercontinuum generation driven by mid-IR ultrashort pulses in a gas-filled hollow-core fiber
,”
Sci. Rep.
9
,
4446
(
2019
).
248.
J. C.
Travers
,
T. F.
Grigorova
,
C.
Brahms
, and
F.
Belli
, “
High-energy pulse self-compression and ultraviolet generation through soliton dynamics in hollow capillary fibres
,”
Nat. Photonics
13
,
547
554
(
2019
).
249.
S. I.
Hwang
,
S. B.
Park
,
J.
Mun
,
W.
Cho
,
C. H.
Nam
, and
K. T.
Kim
, “
Generation of a single-cycle pulse using a two-stage compressor and its temporal characterization using a tunnelling ionization method
,”
Sci. Rep.
9
,
1613
(
2019
).
250.
F.
Köttig
,
D.
Schade
,
J. R.
Koehler
,
P. St. J.
Russell
, and
F.
Tani
, “
Efficient single-cycle pulse compression of an ytterbium fiber laser at 10 MHz repetition rate
,”
Opt. Express
28
,
9099
9110
(
2020
).
251.
U.
Elu
,
L.
Maidment
,
L.
Vamos
,
F.
Tani
,
D.
Novoa
,
M. H.
Frosz
,
V.
Badikov
,
D.
Badikov
,
V.
Petrov
,
P. St. J.
Russell
, and
J.
Biegert
, “
Seven-octave high-brightness and carrier-envelope-phase-stable light source
,”
Nat. Photonics
15
,
277
280
(
2021
).
252.
D.
Schade
,
F.
Köttig
,
J. R.
Koehler
,
M. H.
Frosz
,
P. St. J.
Russell
, and
F.
Tani
, “
Scaling rules for high quality soliton self-compression in hollow-core fibers
,”
Opt. Express
29
,
19147
19158
(
2021
).
253.
M.
Nurhuda
,
A.
Suda
,
K.
Midorikawa
,
M.
Hatayama
, and
K.
Nagasaka
, “
Propagation dynamics of femtosecond laser pulses in a hollow fiber filled with argon: Constant gas pressure versus differential gas pressure
,”
J. Opt. Soc. Am. B
20
,
2002
2011
(
2003
).
254.
K. F.
Mak
,
J. C.
Travers
,
N. Y.
Joly
,
A.
Abdolvand
, and
P. St. J.
Russell
, “
Two techniques for temporal pulse compression in gas-filled hollow-core kagomé photonic crystal fiber
,”
Opt. Lett.
38
,
3592
3595
(
2013
).
255.
A. C.
Judge
,
O.
Bang
,
B. J.
Eggleton
,
B. T.
Kuhlmey
,
E. C.
Mägi
,
R.
Pant
, and
C. M.
de Sterke
, “
Optimization of the soliton self-frequency shift in a tapered photonic crystal fiber
,”
J. Opt. Soc. Am. B
26
,
2064
2071
(
2009
).
256.
J. E.
Beetar
,
M.
Nrisimhamurty
,
T.-C.
Truong
,
Y.
Liu
, and
M.
Chini
, “
Thermal effects in molecular gas-filled hollow-core fibers
,”
Opt. Lett.
46
,
2437
2440
(
2021
).
257.
F.
Benabid
,
F.
Couny
,
J. C.
Knight
,
T. A.
Birks
, and
P. St. J.
Russell
, “
Compact, stable and efficient all-fibre gas cells using hollow-core photonic crystal fibres
,”
Nature
434
,
488
491
(
2005
).
258.
F.
Couny
,
F.
Benabid
, and
O.
Carraz
, “
Enhanced SRS in H2 filled hollow core photonic crystal fibre by use of fibre Bragg grating
,”
J. Opt. A: Pure Appl. Opt.
9
,
156
(
2007
).
259.
F.
Couny
,
F.
Benabid
, and
P. S.
Light
, “
Subwatt threshold cw Raman fiber-gas laser based on H2-filled hollow-core photonic crystal fiber
,”
Phys. Rev. Lett.
99
,
143903
(
2007
).
260.
W.
Pei
,
H.
Li
,
W.
Huang
,
M.
Wang
, and
Z.
Wang
, “
All-fiber tunable pulsed 1.7 μm fiber lasers based on stimulated Raman scattering of hydrogen molecules in hollow-core fibers
,”
Molecules
26
,
4561
(
2021
).
261.
H.
Li
,
W.
Huang
,
W.
Pei
,
Z.
Zhou
,
Y.
Cui
,
M.
Wang
, and
Z.
Wang
, “
All-fiber gas Raman laser oscillator
,”
Opt. Lett.
46
,
5208
5211
(
2021
).
262.
R.
Thapa
,
K.
Knabe
,
K. L.
Corwin
, and
B. R.
Washburn
, “
Arc fusion splicing of hollow-core photonic bandgap fibers for gas-filled fiber cells
,”
Opt. Express
14
,
9576
9583
(
2006
).
263.
P. S.
Light
,
F.
Couny
, and
F.
Benabid
, “
Low optical insertion-loss and vacuum-pressure all-fiber acetylene cell based on hollow-core photonic crystal fiber
,”
Opt. Lett.
31
,
2538
2540
(
2006
).
264.
J. T.
Kristensen
,
A.
Houmann
,
X.
Liu
, and
D.
Turchinovich
, “
Low-loss polarization-maintaining fusion splicing of single-mode fibers and hollow-core photonic crystal fibers, relevant for monolithic fiber laser pulse compression
,”
Opt. Express
16
,
9986
9995
(
2008
).
265.
N. V.
Wheeler
,
M. D. W.
Grogan
,
P. S.
Light
,
F.
Couny
,
T. A.
Birks
, and
F.
Benabid
, “
Large-core acetylene-filled photonic microcells made by tapering a hollow-core photonic crystal fiber
,”
Opt. Lett.
35
,
1875
1877
(
2010
).
266.
X.
Zheng
,
B.
Debord
,
L.
Vincetti
,
B.
Beaudou
,
F.
Gérôme
, and
F.
Benabid
, “
Fusion splice between tapered inhibited coupling hypocycloid-core kagomé fiber and SMF
,”
Opt. Express
24
,
14642
14647
(
2016
).
267.
C.
Wang
,
R.
Yu
,
B.
Debord
,
F.
Gérôme
,
F.
Benabid
,
K. S.
Chiang
, and
L.
Xiao
, “
Ultralow-loss fusion splicing between negative curvature hollow-core fibers and conventional SMFs with a reverse-tapering method
,”
Opt. Express
29
,
22470
22478
(
2021
).
268.
Z.
Zhang
,
R.
Li
,
C.
Wang
,
M.
Zhou
,
Y.
Liu
, and
Y.
Pang
, “
Fusion splicing of hollow-core to standard single-mode fibers using a gradient-index bridge fiber
,”
J. Opt. Technol.
90
,
42
45
(
2023
).
269.
N.
Gisin
and
R.
Thew
, “
Quantum communication
,”
Nat. Photonics
1
,
165
171
(
2007
).
270.
T. D.
Ladd
,
F.
Jelezko
,
R.
Laflamme
,
Y.
Nakamura
,
C.
Monroe
, and
J. L.
O’Brien
, “
Quantum computers
,”
Nature
464
,
45
53
(
2010
).
271.
R.
Horodecki
, “
Quantum information
,” arXiv:2103.07712 (
2021
).
272.
S. K.
Sood
and
Pooja
, “
Quantum computing review: A decade of research
,”
IEEE Trans. Eng. Manage.
(published online)
(
2023
).
273.
H.-H.
Lu
,
M.
Liscidini
,
A. L.
Gaeta
,
A. M.
Weiner
, and
J. M.
Lukens
, “
Frequency-bin photonic quantum information
,”
Optica
10
,
1655
1671
(
2023
).
274.
H. J.
Kimble
, “
The quantum internet
,”
Nature
453
,
1023
1030
(
2008
).
275.
I. A.
Walmsley
and
J.
Nunn
, “
Editorial: Building quantum networks
,”
Phys. Rev. Appl.
6
,
040001
(
2016
).
276.
J.
Huang
and
P.
Kumar
, “
Observation of quantum frequency conversion
,”
Phys. Rev. Lett.
68
,
2153
2156
(
1992
).
277.
N.
Maring
,
D.
Lago-Rivera
,
A.
Lenhard
,
G.
Heinze
, and
H.
de Riedmatten
, “
Quantum frequency conversion of memory-compatible single photons from 606 nm to the telecom C-band
,”
Optica
5
,
507
513
(
2018
).
278.
M.-Y.
Zheng
,
Q.
Yao
,
B.
Wang
,
X.-P.
Xie
,
Q.
Zhang
, and
J.-W.
Pan
, “
Integrated multichannel lithium niobate waveguides for quantum frequency conversion
,”
Phys. Rev. Appl.
14
,
034035
(
2020
).
279.
H.
Fan
,
Z.
Ma
,
J.
Chen
,
Z.
Li
,
C.
Tang
,
Y. M.
Sua
, and
Y.
Huang
, “
Photon conversion in thin-film lithium niobate nanowaveguides: A noise analysis
,”
J. Opt. Soc. Am. B
38
,
2172
2179
(
2021
).
280.
X.
Wang
,
X.
Jiao
,
B.
Wang
,
Y.
Liu
,
X.-P.
Xie
,
M.-Y.
Zheng
,
Q.
Zhang
, and
J.-W.
Pan
, “
Quantum frequency conversion and single-photon detection with lithium niobate nanophotonic chips
,”
npj Quantum Inf.
9
,
38
(
2023
).
281.
H. J.
McGuinness
,
M. G.
Raymer
,
C. J.
McKinstrie
, and
S.
Radic
, “
Quantum frequency translation of single-photon states in a photonic crystal fiber
,”
Phys. Rev. Lett.
105
,
093604
(
2010
).
282.
T. A.
Wright
,
C.
Parry
,
O. R.
Gibson
,
R. J. A.
Francis-Jones
, and
P. J.
Mosley
, “
Resource-efficient frequency conversion for quantum networks via sequential four-wave mixing
,”
Opt. Lett.
45
,
4587
4590
(
2020
).
283.
K. A. G.
Bonsma-Fisher
,
P. J.
Bustard
,
C.
Parry
,
T. A.
Wright
,
D. G.
England
,
B. J.
Sussman
, and
P. J.
Mosley
, “
Ultratunable quantum frequency conversion in photonic crystal fiber
,”
Phys. Rev. Lett.
129
,
203603
(
2022
).
284.
A.
Aghababaei
,
C.
Biesek
,
F.
Vewinger
, and
S.
Stellmer
, “
Frequency conversion in pressurized hydrogen
,”
Opt. Lett.
48
,
45
48
(
2023
).
285.
C. K.
Law
,
I. A.
Walmsley
, and
J. H.
Eberly
, “
Continuous frequency entanglement: Effective finite Hilbert space and entropy control
,”
Phys. Rev. Lett.
84
,
5304
5307
(
2000
).
286.
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
).
287.
A. B.
U’Ren
,
C.
Silberhorn
,
R.
Erdmann
,
K.
Banaszek
,
W. P.
Grice
,
I. A.
Walmsley
, and
M. G.
Raymer
, “
Generation of pure-state single-photon wavepackets by conditional preparation based on spontaneous parametric downconversion
,” arXiv: quant-ph/0611019 (
2006
).
288.
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
).
289.
O.
Morin
,
C.
Fabre
, and
J.
Laurat
, “
Experimentally accessing the optimal temporal mode of traveling quantum light states
,”
Phys. Rev. Lett.
111
,
213602
(
2013
).
290.
B.
Brecht
,
A.
Eckstein
,
R.
Ricken
,
V.
Quiring
,
H.
Suche
,
L.
Sansoni
, and
C.
Silberhorn
, “
Demonstration of coherent time-frequency Schmidt mode selection using dispersion-engineered frequency conversion
,”
Phys. Rev. A
90
,
030302
(
2014
).
291.
B.
Brecht
,
D. V.
Reddy
,
C.
Silberhorn
, and
M. G.
Raymer
, “
Photon temporal modes: A complete framework for quantum information science
,”
Phys. Rev. X
5
,
041017
(
2015
).
292.
V.
Ansari
,
J. M.
Donohue
,
B.
Brecht
, and
C.
Silberhorn
, “
Tailoring nonlinear processes for quantum optics with pulsed temporal-mode encodings
,”
Optica
5
,
534
550
(
2018
).
293.
D. V.
Reddy
and
M. G.
Raymer
, “
High-selectivity quantum pulse gating of photonic temporal modes using all-optical Ramsey interferometry
,”
Optica
5
,
423
428
(
2018
).
294.
M.
Kues
,
C.
Reimer
,
J. M.
Lukens
,
W. J.
Munro
,
A. M.
Weiner
,
D. J.
Moss
, and
R.
Morandotti
, “
Quantum optical microcombs
,”
Nat. Photonics
13
,
170
179
(
2019
).
295.
M. G.
Raymer
and
I. A.
Walmsley
, “
Temporal modes in quantum optics: Then and now
,”
Phys. Scr.
95
,
064002
(
2020
).
296.
C.
Fabre
and
N.
Treps
, “
Modes and states in quantum optics
,”
Rev. Mod. Phys.
92
,
035005
(
2020
).
297.
V.
Ansari
,
B.
Brecht
,
J.
Gil-Lopez
,
J. M.
Donohue
,
J.
Řeháček
,
Z.
Hradil
,
L. L.
Sánchez-Soto
, and
C.
Silberhorn
, “
Achieving the ultimate quantum timing resolution
,”
PRX Quantum
2
,
010301
(
2021
).
298.
Z.
Wang
,
R.
Malaney
, and
R.
Aguinaldo
, “
Temporal modes of light in satellite-to-Earth quantum communications
,”
IEEE Commun. Lett.
26
,
311
315
(
2022
).
299.
C.
Joshi
,
B. M.
Sparkes
,
A.
Farsi
,
T.
Gerrits
,
V.
Verma
,
S.
Ramelow
,
S. W.
Nam
, and
A. L.
Gaeta
, “
Picosecond-resolution single-photon time lens for temporal mode quantum processing
,”
Optica
9
,
364
373
(
2022
).
300.
A.
Eckstein
,
B.
Brecht
, and
C.
Silberhorn
, “
A quantum pulse gate based on spectrally engineered sum frequency generation
,”
Opt. Express
19
,
13770
13778
(
2011
).
301.
B.
Brecht
,
A.
Eckstein
,
A.
Christ
,
H.
Suche
, and
C.
Silberhorn
, “
From quantum pulse gate to quantum pulse shaper—Engineered frequency conversion in nonlinear optical waveguides
,”
New J. Phys.
13
,
065029
(
2011
).
302.
O.
Cohen
,
J. S.
Lundeen
,
B. J.
Smith
,
G.
Puentes
,
P. J.
Mosley
, and
I. A.
Walmsley
, “
Tailored photon-pair generation in optical fibers
,”
Phys. Rev. Lett.
102
,
123603
(
2009
).
303.
M.
Halder
,
J.
Fulconis
,
B.
Cemlyn
,
A.
Clark
,
C.
Xiong
,
W. J.
Wadsworth
, and
J. G.
Rarity
, “
Nonclassical 2-photon interference with separate intrinsically narrowband fibre sources
,”
Opt. Express
17
,
4670
4676
(
2009
).
304.
C.
Söller
,
B.
Brecht
,
P. J.
Mosley
,
L. Y.
Zang
,
A.
Podlipensky
,
N. Y.
Joly
,
P. St. J.
Russell
, and
C.
Silberhorn
, “
Bridging visible and telecom wavelengths with a single-mode broadband photon pair source
,”
Phys. Rev. A
81
,
031801
(
2010
).
305.
R. J. A.
Francis-Jones
,
R. A.
Hoggarth
, and
P. J.
Mosley
, “
All-fiber multiplexed source of high-purity single photons
,”
Optica
3
,
1270
1273
(
2016
).
306.
J.
Su
,
L.
Cui
, and
X.
Li
, “
Wavelength tunable source of correlated photon pairs based on photonic crystal fiber
,”
IEEE Photonics Technol. Lett.
30
,
907
910
(
2018
).
307.
A. A.
Shukhin
,
J.
Keloth
,
K.
Hakuta
, and
A. A.
Kalachev
, “
Heralded single-photon and correlated-photon-pair generation via spontaneous four-wave mixing in tapered optical fibers
,”
Phys. Rev. A
101
,
053822
(
2020
).
308.
B. J.
Smith
,
P.
Mahou
,
O.
Cohen
,
J. S.
Lundeen
, and
I. A.
Walmsley
, “
Photon pair generation in birefringent optical fibers
,”
Opt. Express
17
,
23589
23602
(
2009
).
309.
X.
Li
,
J.
Chen
,
P.
Voss
,
J.
Sharping
, and
P.
Kumar
, “
All-fiber photon-pair source for quantum communications: Improved generation of correlated photons
,”
Opt. Express
12
,
3737
3744
(
2004
).
310.
H.
Takesue
and
K.
Inoue
, “
1.5-μm band quantum-correlated photon pair generation in dispersion-shifted fiber: Suppression of noise photons by cooling fiber
,”
Opt. Express
13
,
7832
7839
(
2005
).
311.
S. D.
Dyer
,
B.
Baek
, and
S. W.
Nam
, “
High-brightness, low-noise, all-fiber photon pair source
,”
Opt. Express
17
,
10290
10297
(
2009
).
312.
Y. M.
Sua
,
J.
Malowicki
,
M.
Hirano
, and
K. F.
Lee
, “
Generation of high-purity entangled photon pair in a short highly nonlinear fiber
,”
Opt. Lett.
38
,
73
75
(
2013
).
313.
A.
Eckstein
,
A.
Christ
,
P. J.
Mosley
, and
C.
Silberhorn
, “
Highly efficient single-pass source of pulsed single-mode twin beams of light
,”
Phys. Rev. Lett.
106
,
013603
(
2011
).
314.
C.
Xiong
,
G. D.
Marshall
,
A.
Peruzzo
,
M.
Lobino
,
A. S.
Clark
,
D.-Y.
Choi
,
S. J.
Madden
,
C. M.
Natarajan
,
M. G.
Tanner
,
R. H.
Hadfield
,
S. N.
Dorenbos
,
T.
Zijlstra
,
V.
Zwiller
,
M. G.
Thompson
,
J. G.
Rarity
,
M. J.
Steel
,
B.
Luther-Davies
,
B. J.
Eggleton
, and
J. L.
O’Brien
, “
Generation of correlated photon pairs in a chalcogenide As2S3 waveguide
,”
Appl. Phys. Lett.
98
,
051101
(
2011
).
315.
M.
Barbier
,
I.
Zaquine
, and
P.
Delaye
, “
Spontaneous four-wave mixing in liquid-core fibers: Towards fibered Raman-free correlated photon sources
,”
New J. Phys.
17
,
053031
(
2015
).
316.
M.
Cordier
,
A.
Orieux
,
R.
Gabet
,
T.
Harlé
,
N.
Dubreuil
,
E.
Diamanti
,
P.
Delaye
, and
I.
Zaquine
, “
Raman-tailored photonic crystal fiber for telecom band photon-pair generation
,”
Opt. Lett.
42
,
2583
2586
(
2017
).
317.
M.
Cordier
,
P.
Delaye
,
F.
Gérôme
,
F.
Benabid
, and
I.
Zaquine
, “
Raman-free fibered photon-pair source
,”
Sci. Rep.
10
,
1650
(
2020
).
318.
W.
Larson
,
T. L.
Courtney
, and
C.
Keyser
, “
Spectrally pure photons generated in a quasi-phase matched xenon-filled hollow-core photonic crystal fiber
,”
Opt. Express
30
,
5739
5757
(
2022
).
319.
X.
Chen
,
W.
Ding
,
Y.-Y.
Wang
,
S.-F.
Gao
,
F.
Xu
,
H.
Xu
,
Y.-F.
Hong
,
Y.-Z.
Sun
,
P.
Wang
,
Y.-Q.
Lu
, and
L.
Zhang
, “
High-fidelity, low-latency polarization quantum state transmissions over a hollow-core conjoined-tube fiber at around 800 nm
,”
Photonics Res.
9
,
460
470
(
2021
).
320.
M.
Maier
,
W.
Kaiser
, and
J. A.
Giordmaine
, “
Intense light bursts in the stimulated Raman effect
,”
Phys. Rev. Lett.
17
,
1275
1277
(
1966
).
321.
M.
Maier
,
W.
Kaiser
, and
J. A.
Giordmaine
, “
Backward stimulated Raman scattering
,”
Phys. Rev.
177
,
580
599
(
1969
).
322.
J.
Murray
,
J.
Goldhar
,
D.
Eimerl
, and
A.
Szoke
, “
Raman pulse compression of excimer lasers for application to laser fusion
,”
IEEE J. Quantum Electron.
15
,
342
368
(
1979
).
323.
R. R.
Alfano
and
G. A.
Zawadzkas
, “
Observation of backward—Stimulated Raman scattering generated by picosecond laser pulses in liquids
,”
Phys. Rev. A
9
,
822
824
(
1974
).
324.
S. M.
Arrivo
,
K. G.
Spears
, and
J.
Sipior
, “
Picosecond transient backward stimulated Raman scattering and pumping of femtosecond dye lasers
,”
Opt. Commun.
116
,
377
382
(
1995
).
325.
N.
Tcherniega
,
A.
Sokolovskaia
,
A. D.
Kudriavtseva
,
R.
Barille
, and
G.
Rivoire
, “
Backward stimulated Raman scattering in water
,”
Opt. Commun.
181
,
197
205
(
2000
).
326.
V. V.
Akulinichev
,
V. A.
Gorbunov
, and
E. G.
Pivinskii
, “
Competition between forward and backward stimulated Raman scattering in gases
,”
Quantum Electron.
27
,
427
(
1997
).
327.
J.
Leng
,
G.
Sha
,
X.
Hua
,
H.
Yang
, and
C.
Zhang
, “
Study of the competition between forward and backward stimulated Raman scattering in methane
,”
Appl. Phys. B
82
,
463
468
(
2006
).
328.
R. R.
Jacobs
,
J.
Goldhar
,
D.
Eimerl
,
S. B.
Brown
, and
J. R.
Murray
, “
High-efficiency energy extraction in backward-wave Raman scattering
,”
Appl. Phys. Lett.
37
,
264
266
(
1980
).
329.
J. O.
White
, “
High-efficiency backward Stokes Raman conversion in deuterium
,”
J. Opt. Soc. Am. B
7
,
785
789
(
1990
).
330.
C.-H.
Lin
,
B. R.
Marshall
,
M. A.
Nelson
, and
J. K.
Theobald
, “
Backward stimulated Raman scattering in multimode fiber
,”
Appl. Opt.
17
,
2486
2488
(
1978
).
331.
M.
Nakazawa
,
M.
Tokuda
,
Y.
Negishi
, and
N.
Uchida
, “
Active transmission line: Light amplification by backward-stimulated Raman scattering in polarization-maintaining optical fiber
,”
J. Opt. Soc. Am. B
1
,
80
85
(
1984
).
332.
K.
Mochizuki
,
N.
Edagawa
, and
Y.
Iwamoto
, “
Amplified spontaneous Raman scattering in fiber Raman amplifiers
,”
J. Lightwave Technol.
4
,
1328
1333
(
1986
).
333.
E. M.
Dianov
,
P. V.
Mamyshev
,
A. M.
Prokhorov
, and
D. G.
Fursa
, “
Stimulated-Raman-effect amplification of femtosecond light pulses in counterpropagating and copropagating pump beams
,”
JETP Lett.
46
,
482
485
(
1987
), available at http://jetpletters.ru/ps/1233/article_18619.pdf.
334.
A.
Abdolvand
,
A.
Nazarkin
,
A. V.
Chugreev
,
C. F.
Kaminski
, and
P. St. J.
Russell
, “
Solitary pulse generation by backward Raman scattering in H2-filled photonic crystal fibers
,”
Phys. Rev. Lett.
103
,
183902
(
2009
).
335.
N.
Thang
and
T.
Thanh
, “
Self-similarity in transient backward stimulated Raman scattering by gas filled photonic crystal fibres
,”
The 9th International Conference on Photonics and Applications
(
2016
), available at https://www.researchgate.net/publication/311806450.
336.
T.
Nguyen Manh
and
T. T.
Doan
, “
A train of backward Raman soliton pulses in hollow-core photonic crystal fibres filled with hydrogen gas
,”
Optik
127
,
10259
10265
(
2016
).
337.
M. K.
Mridha
,
D.
Novoa
, and
P. St. J.
Russell
, “
Dominance of backward stimulated Raman scattering in gas-filled hollow-core photonic crystal fibers
,”
Optica
5
,
570
576
(
2018
).
338.
C.
Feng
,
S.
Jiang
, and
Y.
Feng
, “
Study on backward stimulated Raman scattering in HC-PCF
,” in
2022 IEEE 10th Joint International Information Technology and Artificial Intelligence Conference (ITAIC)
(
IEEE
,
2022
), Vol.
10
, pp.
149
153
.
339.
B.
Culshaw
, “
The optical fibre Sagnac interferometer: An overview of its principles and applications
,”
Meas. Sci. Technol.
17
,
R1
(
2006
).
340.
M. A.
Terrel
,
M. J. F.
Digonnet
, and
S.
Fan
, “
Resonant fiber optic gyroscope using an air-core fiber
,”
J. Lightwave Technol.
30
,
931
937
(
2012
).
341.
M.
Mičuda
,
E.
Doláková
,
I.
Straka
,
M.
Miková
,
M.
Dušek
,
J.
Fiurášek
, and
M.
Ježek
, “
Highly stable polarization independent Mach-Zehnder interferometer
,”
Rev. Sci. Instrum.
85
,
083103
(
2014
).
342.
J. N.
Chamoun
and
M. J. F.
Digonnet
, “
Noise and bias error due to polarization coupling in a fiber optic gyroscope
,”
J. Lightwave Technol.
33
,
2839
2847
(
2015
).
343.
L.-S.
Ma
,
P.
Jungner
,
J.
Ye
, and
J. L.
Hall
, “
Delivering the same optical frequency at two places: Accurate cancellation of phase noise introduced by an optical fiber or other time-varying path
,”
Opt. Lett.
19
,
1777
1779
(
1994
).
344.
R.
Salem
,
G.
Tudur
,
T. U.
Horton
,
G. M.
Carter
, and
T. E.
Murphy
, “
Polarization-insensitive optical clock recovery at 80 Gb/s using a silicon photodiode
,”
IEEE Photonics Technol. Lett.
17
,
1968
1970
(
2005
).
345.
S.
Falke
,
M.
Misera
,
U.
Sterr
, and
C.
Lisdat
, “
Delivering pulsed and phase stable light to atoms of an optical clock
,”
Appl. Phys. B
107
,
301
311
(
2012
).
346.
D.
Rosenberg
,
J. W.
Harrington
,
P. R.
Rice
,
P. A.
Hiskett
,
C. G.
Peterson
,
R. J.
Hughes
,
A. E.
Lita
,
S. W.
Nam
, and
J. E.
Nordholt
, “
Long-distance decoy-state quantum key distribution in optical fiber
,”
Phys. Rev. Lett.
98
,
010503
(
2007
).
347.
C.
Antonelli
,
M.
Shtaif
, and
M.
Brodsky
, “
Sudden death of entanglement induced by polarization mode dispersion
,”
Phys. Rev. Lett.
106
,
080404
(
2011
).
348.
M.
Brodsky
,
E. C.
George
,
C.
Antonelli
, and
M.
Shtaif
, “
Loss of polarization entanglement in a fiber-optic system with polarization mode dispersion in one optical path
,”
Opt. Lett.
36
,
43
45
(
2011
).
349.
M.
Shtaif
,
C.
Antonelli
, and
M.
Brodsky
, “
Nonlocal compensation of polarization mode dispersion in the transmission of polarization entangled photons
,”
Opt. Express
19
,
1728
1733
(
2011
).
350.
D. E.
Jones
,
B. T.
Kirby
, and
M.
Brodsky
, “
Tuning quantum channels to maximize polarization entanglement for telecom photon pairs
,”
npj Quantum Inf.
4
,
58
(
2018
).
351.
A.
Taranta
,
E.
Numkam Fokoua
,
S. M.
Abokhamis Mousavi
,
J. R.
Hayes
,
T. D.
Bradley
,
G. T.
Jasion
, and
F.
Poletti
, “
Exceptional polarization purity in antiresonant hollow-core optical fibres
,”
Nat. Photonics
14
,
504
510
(
2020
).
352.
B. J.
Mangan
,
J. K.
Lyngso
, and
P. J.
Roberts
, “
Realization of low loss and polarization maintaining hollow core photonic crystal fibers
,” in
2008 Conference on Lasers and Electro-Optics and 2008 Conference on Quantum Electronics and Laser Science
(
IEEE
,
2008
), pp.
1
2
.
353.
J. M.
Fini
,
J. W.
Nicholson
,
B.
Mangan
,
L.
Meng
,
R. S.
Windeler
,
E. M.
Monberg
,
A.
DeSantolo
,
F. V.
DiMarcello
, and
K.
Mukasa
, “
Polarization maintaining single-mode low-loss hollow-core fibres
,”
Nat. Commun.
5
,
5085
(
2014
).
354.
S.
Yerolatsitis
,
R.
Shurvinton
,
P.
Song
,
Y.
Zhang
,
R. J. A.
Francis-Jones
, and
K. R.
Rusimova
, “
Birefringent anti-resonant hollow-core fiber
,”
J. Lightwave Technol.
38
,
5157
5162
(
2020
).
355.
G.
Stȩpniewski
,
D.
Dobrakowski
,
D.
Pysz
,
R.
Kasztelanic
,
R.
Buczyński
, and
M.
Klimczak
, “
Birefringent large-mode-area anti-resonant hollow core fiber in the 1.9 μm wavelength window
,”
Opt. Lett.
45
,
4280
4283
(
2020
).
356.
Y.-f.
Hong
,
S.-f.
Gao
,
W.
Ding
,
X.
Zhang
,
A.-q.
Jia
,
Y.-l.
Sheng
,
P.
Wang
, and
Y.-y.
Wang
, “
Highly birefringent anti-resonant hollow-core fiber with a Bi-thickness fourfold semi-tube structure
,”
Laser Photonics Rev.
16
,
2100365
(
2022
).
357.
A.
Hasegawa
, “
Self-confinement of multimode optical pulse in a glass fiber
,”
Opt. Lett.
5
,
416
417
(
1980
).
358.
R. H.
Stolen
,
J. E.
Bjorkholm
, and
A.
Ashkin
, “
Phase-matched three-wave mixing in silica fiber optical waveguides
,”
Appl. Phys. Lett.
24
,
308
310
(
1974
).
359.
S.
Fan
and
J. M.
Kahn
, “
Principal modes in multimode waveguides
,”
Opt. Lett.
30
,
135
137
(
2005
).
360.
L. G.
Wright
,
D. N.
Christodoulides
, and
F. W.
Wise
, “
Controllable spatiotemporal nonlinear effects in multimode fibres
,”
Nat. Photonics
9
,
306
310
(
2015
).
361.
K.
Krupa
,
A.
Tonello
,
B. M.
Shalaby
,
M.
Fabert
,
A.
Barthélémy
,
G.
Millot
,
S.
Wabnitz
, and
V.
Couderc
, “
Spatial beam self-cleaning in multimode fibres
,”
Nat. Photonics
11
,
237
241
(
2017
).
362.
L. G.
Wright
,
D. N.
Christodoulides
, and
F. W.
Wise
, “
Spatiotemporal mode-locking in multimode fiber lasers
,”
Science
358
,
94
97
(
2017
).
363.
F. O.
Wu
,
A. U.
Hassan
, and
D. N.
Christodoulides
, “
Thermodynamic theory of highly multimoded nonlinear optical systems
,”
Nat. Photonics
13
,
776
782
(
2019
).
364.
J.
Demas
,
L.
Rishøj
,
X.
Liu
,
G.
Prabhakar
, and
S.
Ramachandran
, “
Intermodal group-velocity engineering for broadband nonlinear optics
,”
Photonics Res.
7
,
1
7
(
2019
).
365.
H.
Pourbeyram
,
P.
Sidorenko
,
F. O.
Wu
,
N.
Bender
,
L.
Wright
,
D. N.
Christodoulides
, and
F. W.
Wise
, “
Direct observations of thermalization to a Rayleigh–Jeans distribution in multimode optical fibres
,”
Nat. Phys.
18
,
685
690
(
2022
).
366.
L. G.
Wright
,
F. O.
Wu
,
D. N.
Christodoulides
, and
F. W.
Wise
, “
Physics of highly multimode nonlinear optical systems
,”
Nat. Phys.
18
,
1018
1030
(
2022
).
367.
R.
Piccoli
,
J. M.
Brown
,
Y.-G.
Jeong
,
A.
Rovere
,
L.
Zanotto
,
M. B.
Gaarde
,
F.
Légaré
,
A.
Couairon
,
J. C.
Travers
,
R.
Morandotti
,
B. E.
Schmidt
, and
L.
Razzari
, “
Intense few-cycle visible pulses directly generated via nonlinear fibre mode mixing
,”
Nat. Photonics
15
,
884
889
(
2021
).
368.
B.
Winter
,
T. A.
Birks
, and
W. J.
Wadsworth
, “
Multimode hollow-core anti-resonant optical fibres
,” in
Frontiers in Optics + Laser Science APS/DLS
(
Optica Publishing Group
,
2019
), p.
JTu4A.18
.
369.
C.
Goel
and
S.
Yoo
, “
Multimode nested antiresonant hollow core fiber
,”
J. Lightwave Technol.
39
,
6592
6598
(
2021
).
370.
Z.
Dong
,
X.
Zhang
,
P.
Yao
,
J.
Yao
,
S.
Wan
, and
P.
Wang
, “
Development of multi-mode rod-type hollow-core antiresonant fiber
,”
Proc. SPIE
12280
,
1228005
(
2022
).
371.
W.
Shere
,
E. N.
Fokoua
,
G. T.
Jasion
, and
F.
Poletti
, “
Designing multi-mode anti-resonant hollow-core fibers for industrial laser power delivery
,”
Opt. Express
30
,
40425
40440
(
2022
).
372.
D.
Wu
,
F.
Yu
,
C.
Wu
,
M.
Zhao
,
J.
Zheng
,
L.
Hu
, and
J.
Knight
, “
Low-loss multi-mode anti-resonant hollow-core fibers
,”
Opt. Express
31
,
21870
21880
(
2023
).
373.
M.
Zhao
,
F.
Yu
,
D.
Wu
,
X.
Zhu
,
S.
Chen
,
M.
Wang
,
M.
Liu
,
K.
Zhao
,
R.
Zhai
,
Z.
Jia
, and
J.
Knight
, “
Delivery of nanosecond laser pulses by multimode anti-resonant hollow core fiber at 1 μm wavelength
,” arXiv:2312.06234 (
2023
).

Supplementary Material