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

Scaling rules for high quality soliton self-compression in hollow-core fibers

Open Access Open Access

Abstract

Soliton dynamics can be used to temporally compress laser pulses to few fs durations in many different spectral regions. Here we study analytically, numerically and experimentally the scaling of soliton dynamics in noble gas-filled hollow-core fibers. We identify an optimal parameter region, taking account of higher-order dispersion, photoionization, self-focusing, and modulational instability. Although for single-shots the effects of photoionization can be reduced by using lighter noble gases, they become increasingly important as the repetition rate rises. For the same optical nonlinearity, the higher pressure and longer diffusion times of the lighter gases can considerably enhance the long-term effects of ionization, as a result of pulse-by-pulse buildup of refractive index changes. To illustrate the counter-intuitive nature of these predictions, we compressed 250 fs pulses at 1030 nm in an 80-cm-long hollow-core photonic crystal fiber (core radius 15 µm) to ∼5 fs duration in argon and neon, and found that, although neon performed better at a repetition rate of 1 MHz, stable compression in argon was still possible up to 10 MHz.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Over the last few years, progress in solid-state laser technology has led to ultrafast systems that can provide up to a few kW average power and turn-key operation [1,2]. However, for many applications in ultrafast science and material processing [35], the few-hundred-fs pulses generated by these lasers are too long, and external pulse compression is required. Typically, this relies on optical nonlinearity to broaden the spectrum of laser pulses, followed by phase compensation. In optical systems with anomalous dispersion at the laser frequency, the interplay between nonlinearity and dispersion can cause laser pulses to shorten without need for additional phase compensation. This effect, known as soliton self-compression, was initially observed in conventional solid-core fibers and exploited to obtain pulses hundreds of fs in duration [6,7]. Since then, soliton self-compression has been used in a variety of different systems [812]. Among these, gas-based systems such as hollow-core waveguides and multi-pass cells, are particularly attractive for compression to durations of a few fs at high peak and average power [1115]. Furthermore, the great flexibility of gas-filled hollow-core fibers is very convenient, as it allows few-fs pulses to be obtained over a wide spectral range (from the UV to the mid-IR) from pump pulse energies from sub-µJ up to several mJ [10,12,13,1619]. This is because, in common with many phenomena in gas-based nonlinear optics [20], soliton dynamics in gas-filled hollow-core fibers is scale-invariant [8]. As a result, by using proper scaling relations, similar propagation patterns can be obtained for a plethora of different parameter sets. This flexibility comes, however, with the challenge of identifying optimal experimental parameters for high-quality compression. Soliton self-compression can be impaired in several different ways, depending on the system properties, pulse parameters, and laser repetition rate. It is therefore crucial to have a thorough understanding of these limitations and how optimal parameters can be identified.

Here we study, theoretically and experimentally, the scaling of soliton dynamics in noble gas-filled hollow-core photonic crystal fibers (HC-PCFs). We investigate the limits set by modulational instability, higher-order dispersion, self-focusing and photoionization, and identify an optimal parameter region for pulse compression. We then discuss practical constraints such as repetition rate scaling, and relate the theoretical quantities to experimental parameters. Finally, we present experimental results on soliton self-compression and the dependence on laser repetition rate. For this, we compress 250 fs, 1030 nm laser pulses to ∼5 fs duration in a gas-filled HC-PCF and vary the repetition rate between 1 and 10 MHz.

2. Theoretical discussion

The nonlinear propagation dynamics of ultrashort pulses in gas-filled hollow-core waveguides can be modelled by the generalized nonlinear Schrödinger equation (NLSE). In the case of a pure Kerr nonlinearity (as is the case for Raman-inactive noble gases), and ignoring for the moment the effects of self-focusing and ionization, the dynamics can be well described by three characteristic length-scales: the dispersion length (${L_\textrm{D}}$), the third-order dispersion (TOD) length (${L_{\textrm{TOD}}}$) and the nonlinear length (${L_{\textrm{NL}}}$) [8,21]. For a pulse of peak power ${P_0}$ and central angular frequency ${\omega _0}$, these lengths are given by [21]:

$$\begin{array}{{c}} {{L_\textrm{D}} = \frac{{\tau _0^2}}{{|{{\beta_2}({{\omega_0}} )} |}},\; \; {L_{\textrm{TOD}}} = \frac{{\tau _0^3}}{{|{{\beta_3}({{\omega_0}} )} |}},\; {L_{\textrm{NL}}} = \frac{1}{{\gamma {P_0}}},\; \; \; \textrm{with}\; \; \gamma = \; \frac{{{\omega _0}{n_2}}}{{c{A_{\textrm{eff}}}}},} \end{array}$$
where the parameter ${\tau _0}$ is related to the full-width-half-maximum (FWHM) pulse intensity duration by ${\tau _0} = \; {\tau _{\textrm{FWHM}}}/\ln \left( {3 + \; \sqrt 8 } \right)$ for a sech2-pulse, $\; {n_2}$ is the nonlinear refractive index, ${A_{\textrm{eff}}}$ the effective mode area and c the speed of light in vacuum. In HC-PCF the group velocity dispersion ${\beta _2}$ and TOD ${\beta _3}$ can be estimated using a simple capillary model (see Appendix 1), provided one is operating far from anti-crossings with resonances in the glass walls around the core [8,22]. In contrast to capillaries, which exhibit very high leakage loss (characterized by short loss lengths ${L_{\textrm{loss}}}$ [8]) at small bore diameters, loss in HC-PCF can typically be neglected since ${L_{\textrm{loss}}}$ is much greater than all the other length-scales.

Conveniently, the characteristic length-scales can be reduced to two parameters: the soliton order $N = \sqrt {{L_\textrm{D}}/{L_{\textrm{NL}}}} $ and the zero dispersion (ZD) frequency ${\omega _{\textrm{ZD}}}$ [defined by ${\beta _2}({{\omega_{\textrm{ZD}}}})= 0$]. Once these parameters are fixed, similar (identical for the same ${\tau _0}$) nonlinear propagation dynamics can be reproduced for a wide range of different parameters [8,12]. Moreover, both N and ${\omega _{\textrm{ZD}}}$ can be continuously tuned by changing the gas species and density (both ${\beta _2}$ and the nonlinear coefficient $\gamma $ depend on the gas properties) [8,23], thus allowing precise control of the nonlinear propagation dynamics of ultrashort laser pulses. By selecting parameters such that ${\beta _2}({{\omega_0}} )< 0$ and $N > 1$, a laser pulse propagating in the fiber will undergo soliton self-compression, reaching the shortest duration at the compression length ${L_{\textrm{comp}}}$, which is essentially equal to the fission length ${L_{\textrm{fiss}}} \approx {L_\textrm{D}}/N$, provided ${L_{\textrm{comp}}} < \; {L_{\textrm{loss}}}$ [8,12,21]. Though widely used, this relation becomes increasingly inaccurate at low soliton orders. A more generally applicable expression takes the form [6]:

$$\begin{array}{{c}} {{L_{\textrm{comp}}} = \; \left( {\frac{1}{{2N}} + \frac{{1.7}}{{{N^2}}}} \right){L_\textrm{D}} = \; {A_\textrm{N}}{L_\textrm{D}}} \end{array}$$
where ${A_\textrm{N}}$ is a dimensionless function of N. Beyond this point, the pulse typically undergoes soliton fission caused by perturbations to the soliton dynamics. Nonetheless, even in the case of strong perturbations, ${L_{\textrm{comp}}}$ is only minimally affected. Larger compression factors ${F_\textrm{c}} = {\tau _0}/{\tau _\textrm{c}}$ can be achieved by increasing the soliton order, where ${\tau _\textrm{c}}$ is the duration of the compressed pulse. However, larger compression factors come at the cost of compression quality, quantified by the quality factor ${Q_\textrm{c}}$, which is the ratio of the energy in the FWHM power profile of the compressed pulse and that in the uncompressed pulse. Analysis of large number of NLSE-based numerical simulations suggests that these scaling factors are: ${F_\textrm{c}}\sim 4.4N$ and ${Q_\textrm{c}}\sim 4.3/N$ (for Gaussian-shaped input pulses) [6,8,24]. Although these relations capture the dependence on N, they become increasingly inaccurate (especially for ${F_\textrm{c}})\; $ as the pulse duration decreases, and provide only rough estimates for gas-filled HC fibers.

In Fig.  1 the compression length and compression factor are plotted against soliton order. The discrete datapoints (marked by symbols) were calculated by simulating the propagation of Gaussian pulses with FWHM durations 25 fs (dark blue), 100 fs (orange) and 250 fs (turquoise) along an Ar-filled HC-PCF using the unidirectional full-field nonlinear wave equation [25], including quantum noise [26] and neglecting fiber loss. Values of the nonlinear susceptibility ${\chi ^{(3 )}}\; $ were taken from [27] and photoionization was included via the Perelomov-Popov-Terent’ev (PPT) rate modified with the Ammosov-Delone-Krainov (ADK) coefficients [28]. The compression length predicted by Eq.  (2), plotted by the full lines in Fig.  1(a), agrees remarkably well with numerical simulations (the corresponding datapoints are denoted by the symbols and were obtained by selecting the position of highest peak power). The accuracy is observed to be typically well within a 15% margin even under the most extreme conditions and for low soliton orders. Additionally, we found that the expression ${Q_\textrm{c}} = 3.7/({N + 2.2} )$ gave a more accurate estimate of the quality factor (not shown), especially for low N. On the other hand, ${F_\textrm{c}}$ exhibits a linear dependence on N [dashed red line in Fig.  2(b)] only for small soliton orders. For short pulse durations (∼ 25 fs) the higher-order terms cause the compression factor to saturate. By analysing large sets of numerical data, we find that the compression factor is accurately described by the following fitting function that includes the TOD contribution:

$$ F_{\mathrm{c}}=\frac{\tau_{0}}{\tau_{\mathrm{c}}}=\frac{3}{A_{\mathrm{N}}(1-N \xi)}, \quad \text { with } \quad \xi=\frac{\beta_{3}\left(\omega_{0}\right)}{\tau_{0} \mid \beta_{2}\left(\omega_{0} \right)\mid}=\frac{\partial_{\omega} \delta\left(\omega_{0}, \omega_{\mathrm{ZD}}\right)}{\tau_{0}\left|\delta\left(\omega_{0}, \omega_{\mathrm{ZD}}\right)\right|} $$
where $\delta $ is an analytical function that depends only on ${\omega _0}$ and ${\omega _{\textrm{ZD}}}$ (Appendix 1) and ${\partial _\omega }$ is the partial derivative with respect to $\omega $. This expression, which is plotted with full lines in Fig.  1(b), provides an accurate estimate of the compression factor over a wide range of pulse durations, core radii and gas parameters. However, strong deviations from the numerical data are seen for $N > {N_{\textrm{lim}}}$, where ${N_{\textrm{lim}}}$ depends on both ${\omega _{\textrm{ZD}}}$ and ${\tau _0}$. This indicates that other quantities have a strong impact on soliton self-compression. Identifying these as higher-order dispersion (HOD), gas photoionization (ION) and self-focusing (SF), we can derive phenomenological expressions that set the limits for optimal self-compression in $({N,{\omega_{\textrm{ZD}}},{\tau_0}} )$ space. The HOD limit is obtained by requiring the bandwidth of the self-compressed pulse to be contained within the anomalous dispersion region, which leads to the following condition for pulses with sech2 temporal profiles:
$$\begin{array}{{c}} {\frac{1}{{{\tau _\textrm{c}}}} \le \frac{{|{{\omega_{\textrm{ZD}}} - {\omega_0}} |}}{{0.315\pi }}\ln \left( {3 + \; \sqrt 8 } \right) \equiv {\mathrm{\Delta }_\omega }\sigma ,} \end{array}$$
where ${\mathrm{\Delta }_\omega } = |{{\omega_{ZD}} - {\omega_0}} |$, $\mathrm{\sigma } = \; \textrm{ln}\left( {3 + \sqrt 8 } \right)/({0.315\pi } )$ and 0.315 is the time-bandwidth product of sech2 pulses. The maximum soliton order for which the relations in Eq.  (3) are valid may then be expressed as a function of ${\omega _{\textrm{ZD}}}$:
$$\begin{array}{{c}} {N_{\textrm{ZD}}^{\textrm{max}}({{\omega_0},\; {\omega_{\textrm{ZD}}},\; {\tau_0}} )= {\mathrm{\Delta }_\omega }\mathrm{\sigma }{\tau _0}\frac{{1 - 3.4\xi + \sqrt {{{({1 + 3.4\xi } )}^2} + 81.6/({{\mathrm{\Delta }_\omega }\mathrm{\sigma }{\tau_0}} )} }}{{2({6 + \xi {\mathrm{\Delta }_\omega }\mathrm{\sigma }{\tau_0}} )}}.} \end{array}$$

For soliton orders greater than $N_{\textrm{ZD}}^{\textrm{max}}$, HOD becomes increasingly important and soliton self-compression is impaired through formation of pre- and post-pulses or pulse break-up. In hollow-core waveguides the influence of HOD can be reduced by shifting ${\omega _{\textrm{ZD}}}$ to frequencies much higher than ${\omega _0}$, thus allowing larger values of N to be chosen. Selecting large $\; {\mathrm{\Delta }_\omega }$ (e.g. by using a lighter filling-gas or decreasing the pressure) is, however, not always a good choice and in practice may not always be possible due to limited pulse energy and, more importantly, additional conditions on the maximum value of N set by SF and ION, which become relevant when ${\omega _{\textrm{ZD}}}$ is far away from ${\omega _0}$. Self-focusing causes energy transfer to higher-order modes, which distorts both the spatial and (since dispersion is mode-dependent) temporal profiles of the self-compressing pulse. This effect can be counter-balanced by the defocusing effect of free electrons resulting from gas photoionization [29]. While a certain amount of ionization is tolerable, it becomes detrimental if it distorts the temporal pulse profile. As we shall discuss later, ionization also becomes a constraint if too much energy is deposited in the gas, which can occur at repetition rates higher than a few hundred kHz [13,30,31].

 figure: Fig. 1.

Fig. 1. Scaling behavior of soliton self-compression and optimal parameter region for 1030 nm pulses. (a) Dependence of compression length on soliton order in argon for zero-dispersion wavelength $2\mathrm{\pi }\textrm{c}/{\omega _{\textrm{ZD}}} \approx 450\; \textrm{nm},$ with FWHM input pulse duration 25 fs (dark blue), 100 fs (orange) and 250 fs (turquoise). The symbols mark numerically calculated datapoints (position of highest peak power), and the full lines are solutions of Eq.  (2). For comparison, the fission length ${L_{\textrm{fiss}}} = {L_\textrm{D}}/N$ is also plotted (dashed lines). (b) Dependence of compression factor ${F_\textrm{c}}$ on soliton order for the same case as in (a). The symbols mark numerically calculated datapoints, and the full lines are solutions of Eq.  (3). The red dot-dashed line is a linear fit to ${F_\textrm{c}} = 4.4N - 4$. (c) Analytical plot of soliton order versus ${\tau _0}{\mathrm{\Delta }_\omega }/({2\pi } )$ for 100 fs input pulses with ${\mathrm{\Delta }_\omega } = |{{\omega_{ZD}} - {\omega_0}} |$. MI: modulational instability (turquoise dashed line); SF: self-focusing (dark blue dotted line); ION: photoionization (yellow to red solid lines); HOD: higher-order dispersion (blue dashed line). The size of the gray area, within which optimal compression occurs, is controlled by the ION limit, which depends on the gas species.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. Ratio of peak compressed intensity I to peak launched intensity I0 for 1030 nm pulses in Ar-filled HC-PCF, plotted against soliton order and ${\tau _0}{\mathrm{\Delta }_\omega }/({2\pi } )$, which depends on both core diameter and pressure. Initial pulse durations are (a) 250 fs, (b) 100 fs and (c) 25 fs. The optimal compression region is triangular in shape (yellow-white area) and shrinks with decreasing input pulse duration. At 1030 nm the maximum soliton order is limited mainly by HOD and ION for shorter pulses, whereas for longer pulses MI constrains the maximum soliton order to $N = 16$. The white dot in (a) marks the experimental parameters.

Download Full Size | PDF

Travers et al. [12] defined the SF limit on maximum soliton order by replacing the peak power in the expression for N with the critical power for self-focusing $P_{\textrm{SF}}^{\textrm{cr}} \approx 4\pi {\varepsilon _0}{c^3}/({\omega ^2}{\chi ^{(3 )}}{\rho _\textrm{r}})$ where ${\rho _\textrm{r}}$ is the gas density relative to standard conditions [32,33], and the ION limit by defining the ionization threshold power as ${P_{\textrm{ION}}} = \; {I_\textrm{c}}{A_{\textrm{eff}}}$, where ${I_\textrm{c}}$ is a critical intensity including an arbitrary factor [12]. In the work reported here we define ${I_\textrm{c}}$ as the clamping intensity, i.e., the intensity at which the ionization-induced phase shift equals the Kerr phase shift (see Appendix 1 for details). Expressing ${\beta _2}$ and $\gamma \; $ as functions of ${\omega _{\textrm{ZD}}}$, the following relations can then be obtained for the HE11-like fundamental mode:

$$\begin{array}{{c}} {N_{\textrm{SF}}^{\textrm{max}}({{\omega_0},{\omega_{\textrm{ZD}}},{\tau_0}} )= {\tau _0}\sqrt {\frac{{2\pi c}}{{S{\omega _0}|{\delta ({{\omega_0},{\omega_{\textrm{ZD}}}} )} |}}} ,} \end{array}$$
$$\begin{array}{{c}} {N_{\textrm{ION}}^{\textrm{max}}({{\omega_0},{\omega_{\textrm{ZD}}},{\tau_0}} )= {\tau _0}\sqrt {\frac{{3u_{01}^2{I_\textrm{c}} {\omega _0} {\chi^{(3)}}} }{{4{\varepsilon _0}\omega _{\textrm{ZD}}^3f({{\omega_{\textrm{ZD}}}} )|{\delta ({{\omega_0},{\omega_{\textrm{ZD}}}} )} |}}} \; ,\; } \end{array}$$
where ${u_{01}} = 2.405$, f is a function of the linear susceptibility (Appendix 1), and $S = 10$ is an empirical factor introduced in [12]. It is important to note that, for intense ultrashort pulses, nonlinear intermodal phase-matching can cause spatio-temporal coupling below the critical power threshold for self-focusing [25,29,34,35]. However, for pulses with peak powers ten times lower than $P_{\textrm{SF}}^{\textrm{cr}}$, as is the case in the current work, intermodal energy transfer can be safely neglected.

An additional constraint on self-compression of pulses hundreds of fs long is set by modulational instability (MI) [21,26], which causes the pulse to break up into a train of sub-pulses before the shortest duration is reached. Although under certain circumstances MI can be suppressed in HC-PCFs by the effects of spectral anti-crossings between the fundamental core mode and core-wall resonances [36], here we restrict the analysis to conventional MI, which is typically observed for N > 16 [26]. Using this as a limit, we can identify an optimal parameter range for soliton self-compression by plotting the maximum soliton order as a function of ${\tau _0}{\mathrm{\Delta }_\omega }/({2\pi } )$, a numerical quantity that is related to the fraction of the pulse bandwidth that lies within the anomalous dispersion region.

In Fig.  1(c), we map out the optimal region for a ${\tau _{\textrm{FWHM}}} = \; 100$ fs pulse centered at ${\omega _0} = 2\pi \cdot 0.29\; $PHz (corresponding to a wavelength of 1030 nm). For the selected parameters, the region is mostly delimited by $N_{\textrm{ZD}}^{\textrm{max}}$ and $N_{\textrm{ION}}^{\textrm{max}}$, while self-focusing becomes an effective limit for pulses centered at shorter wavelengths ($N_{\textrm{SF}}^{\textrm{max}}$ depends inversely on $\sqrt {{\omega _0}} $). As seen in Fig.  1(c), a key limitation to optimal compression is ionization, which strongly depends on gas species. For lighter noble gases, the region of optimal compression widens provided $({N,{\tau_0}{\mathrm{\Delta }_\omega }/({2\pi } )} )$ is appropriately chosen. Both $N_{\textrm{ION}}^{\textrm{max}}$ and $N_{\textrm{SF}}^{\textrm{max}}$ are linear with ${\tau _\textrm{0}}\; $ and $N_{\textrm{ZD}}^{\textrm{max}}$ is approximately linear with it. As a result, the extension of the optimal region increases with ${\tau _0}$ and MI introduces an additional border to the region only for long initial pulse durations. For wide-bore capillaries, where the loss can be very significant, it is important to ensure that ${L_{\textrm{comp}}} < {L_{\textrm{loss}}} \approx {\omega ^2}{a^3}/({3{c^2}u_{01}^2} )$ (assuming a fundamental mode) [12]. In this case an additional constraint can be derived by solving the equation ${L_{\textrm{comp}}} = {L_{\textrm{loss}}}$ for N. The solution yields the following lower limit for the soliton order required for self-compression:

$$\begin{array}{{c}} {N_{\textrm{loss}}^{\textrm{min}}\; = \; \frac{3}{{4a|{\delta ({{\omega_0},{\omega_{\textrm{ZD}}}} )} |}}{{\left( {\frac{{c{u_{01}}{\tau_0}}}{{{\omega_0}}}} \right)}^2}\left[ {1 + \sqrt {1 + \frac{{9a|{\delta ({{\omega_0},{\omega_{\textrm{ZD}}}} )} |\omega_0^2}}{{{{({c{u_{01}}{\tau_0}} )}^2}}}} } \right].} \end{array}$$
$N_{\textrm{loss}}^{\textrm{min}}$ scales in a different way to the other limits, as it depends on the radius a and nonlinearly on the duration and central frequency of the pulse.

We verified the validity of these limits by numerically simulating the propagation of 1030 nm pulses in gas-filled HC fibers. Since at this wavelength self-focusing does not play a significant role, and the computational requirements scale quadratically with the number of modes [37], we only considered propagation of the HE11-like mode. We used Gaussian pulses with initial FWHM durations 250 fs, 100 fs and 25 fs, and for each of these we ran a set of 256 simulations for a range of different soliton orders and zero dispersion frequencies. We considered a fiber with 15 µm core radius filled with argon, but we observed similar pulse propagation patterns using other noble gases and core radii (not shown). The results are shown in Fig.  2, where we plot, as function of N and ${\tau _0}{\mathrm{\Delta }_\omega }/({2\pi } ),$ the peak power of the pulses at the compression point normalized to their power at the fiber input. This quantity coincides with the product of the simulated compression and quality factors. As such it is a good measure for the quality of the self-compression. The curves are interpolated from simulated datapoints; since in each case N depends monotonically on ${\mathrm{\Delta }_\omega }$, the interpolations are very accurate.

For the input pulse durations considered, the simulations are in excellent agreement with the maximum soliton order predicted by Eqs.  (56). Self-compression of long pulses (250 fs) can enhance the peak power more than 20 times and is clearly limited by HOD and MI. In the vicinity of the latter limit, $I/{I_0}$ varies greatly with N and ${\tau _0}{\mathrm{\Delta }_\omega }$ thus indicating a pulse break up (we further verified this by analysing individual numerical simulations). For shorter durations, HOD and ION become dominant. The optimal compression region has a characteristic triangular shape, which shrinks as the initial pulse duration is decreased, while the peak power enhancement falls. The greatest enhancement is always located for $({N,{\tau_0}{\mathrm{\Delta }_\omega }/({2\pi } )} )$ values near the center of gravity of the triangle.

It is important to note that the limit set by ION appears to be more accurate for 25 fs input pulses. For much shorter pulse durations (e.g. < 10 fs) the limit is somewhat too strict, while for 250 fs pulses, $N_{\textrm{ION}}^{\textrm{max}}$/1.7 delimits the optimal region more precisely. This is because the definition of $N_{\textrm{ION}}^{\textrm{max}}$ neglects propagation effects and we calculate the clamping intensity in the multiphoton limit of the PPT rate (Appendix 2). When tunneling ionization becomes dominant, ${I_c}$ should be estimated in a different way, for example by fitting to the ADK rate [38].

3. Experimental results

The limits discussed above are particularly useful for identifying the range of parameters for optimal soliton self-compression. Experimental implementation requires, however, consideration of additional constraints, such as core diameter, fiber length, and gas pressure handling capabilities. Furthermore, at high repetition rates, even weak gas photoionization can cause buildup of post-recombination refractive index changes, an effect which can occur already at hundreds of kHz repetition rates and impair the compression dynamics [30,31]. To minimize these effects, the interval between subsequent pulses should be longer than the time needed for the photoionization-induced heat to dissipate radially to the core-walls. This is described by the diffusion time, estimated by ${t_{\textrm{diff}}} = {a^2}/\alpha $, where a is the core radius and $\alpha = \; \kappa /({\varrho \; {C_\textrm{p}}} )$ is the thermal diffusivity, where $\varrho $ is the gas density, ${C_\textrm{p}}$ the heat capacity at constant pressure, and κ the thermal conductivity, which is larger for lighter gases. Matching the nonlinearity and ${\omega _{\textrm{ZD}}}$ of a heavier gas with a lighter gas requires much higher pressures, resulting in higher values of ${t_{\textrm{diff}}}$. To visualise this scaling, it is convenient to express the diffusion time as a function of the zero-dispersion frequency, derived from the ideal gas law using the relations in Appendix 1:

$$\begin{array}{{c}} {{t_{\textrm{diff}}} = \frac{{{C_p}}}{\kappa }\frac{M}{{{k_\textrm{B}}{N_\textrm{A}}}}\frac{{{p_0}}}{{{T_0}}}\frac{{u_{01}^2{c^2}}}{{f({{\omega_{\textrm{ZD}}}} )\; \omega _{\textrm{ZD}}^3}}} \end{array}$$
where M is the molar mass of the gas, ${k_\textrm{B}}$ Boltzmann's constant, ${N_\textrm{A}}$ Avogadro's number, ${p_0} = 1.01325\; \textrm{bar},\; $ ${T_0} = 273.15\; \textrm{K}$ and $f({{\omega_{\textrm{ZD}}}} )$ is defined in Appendix 1. Figure  3(a) plots ${t_{\textrm{diff}}}$ as a function of ${\omega _{\textrm{ZD}}}$ for different noble gases. At fixed zero dispersion frequency, the nonlinear coefficient is slightly larger for the heavier gases, whereas the diffusion time decreases with the molar mass. The only exception is neon, which does not follow the general trend and yields the longest diffusion time. Although lighter noble gases provide a wider optimal compression region, the long diffusion time introduces a new constraint. This can be viewed as a distortion of the $N_{\textrm{ION}}^{\textrm{max}}$ curve at high repetition rates, which results in a lower limit for the soliton order set by ionization. In light of this, we expect neon to be the worst noble gas (despite its high price) for compressing MHz trains of pulses. Counterintuitively, in some cases heavier gases are a better choice for scaling to higher repetition rates, provided the single-shot ION limit [Fig.  1(c)] does not already impair pulse compression for the chosen parameters.

 figure: Fig. 3.

Fig. 3. Single-stage pulse compression to few-cycle durations. (a) Diffusion time as a function of the zero-dispersion frequency for different noble gases. The dashed gray line marks the experimental setpoint at $2\mathrm{\pi }\textrm{c}/{\omega _{\textrm{ZD}}} \approx 523\; \textrm{nm}$. The upper horizontal axis allows comparison with Fig.  1(c); however, note that the diffusion time is independent of the input pulse duration. (b) Scaling of core radius, gas pressure and compression length with pulse energy for 1030 nm, 250 fs pulses in argon at a fixed soliton order of 13.6. The dashed gray line marks the experimental parameters, where a core radius of 15 µm, a pressure of 5.6 bar and a fiber length of 0.8 m are required to compress pulses with an energy of 4 µJ. (c) Sketch of the experimental setup: λ/2: half-wave plate, TFP: thin-film polarizer, OAPM: off-axis parabolic mirror, W: wedge, CM: chirped-mirror. (d) Scanning electron micrograph of the cross-section (zoom) of the 80-cm-long, 8-capillary single-ring PCF with core diameter of 30 µm.

Download Full Size | PDF

We experimentally verified this prediction by compressing 250 fs pulses at ${\omega _0}\sim 2\pi \cdot 0.29\; \textrm{PHz}$ (1030 nm) to nearly single-cycle duration in a fiber with 15 µm core radius filled with either Ar or Ne. To achieve this ∼55-fold temporal compression with the highest efficiency, given the laser and fiber parameters we have access to, we used the numerical results in Fig.  2(a) and selected $N = 13.6$ and ${\tau _0}{\mathrm{\Delta }_\omega }/({2\pi } )\; \sim \; 40$ (yielding ${\lambda _{ZD}} = 2\pi c/{\omega _{\textrm{ZD}}} \approx $ 523 nm), values that are expected to yield the desired ${F_\textrm{c}}$ and compression quality.

To relate the optimal $({N,{\tau_0}{\mathrm{\Delta }_\omega }/({2\pi } )} )$-region with the physical parameters of gas-filled HC fibers, we express the core radius a and the gas pressure p as functions of the input pulse parameters and the zero-dispersion frequency. The core radius, given as:

$$\begin{array}{{c}} {a = \; \sqrt {\frac{{{\chi ^{(3)}{u_{01}^2}}}}{{4{\varepsilon _0}}}\frac{{{\omega _0}}}{{\omega _{\textrm{ZD}}^3\; |{\delta ({{\omega_0},{\omega_{\textrm{ZD}}}} )} |\; f({{\omega_{\textrm{ZD}}}} )}}\frac{{{\tau _0}E}}{{{N^2}}}} \; ,} \end{array}$$
where E is the energy of the input pulse, is used to calculate the required gas pressure:
$$\begin{array}{{c}} {p = \; \frac{{u_{01}^2{c^2}}}{{{a^2}\omega _{\textrm{ZD}}^3}}\frac{1}{{f({{\omega_{\textrm{ZD}}}} )}}\frac{T}{{{T_0}}}{p_0}\; ,} \end{array}$$
where T is the ambient temperature and ${T_0}\; = \; 273.15\; K$. Figure  3(b) shows the core radius, argon pressure and compression length for $N = 13.6$, ${\tau _0}{\mathrm{\Delta }_\omega }/({2\pi } )= 40$, and input energies between 1 and 20 µJ. The core radius [Eq.  (9)] and gas pressure [Eq.  (10)] show the same dependence on pulse duration as on energy, only the compression length scales differently. From the plot we find that for a pulse energy of 4 µJ a core radius of 15 µm is suitable. Following the dashed vertical line at this energy, we obtain the required gas pressure (5.6 bar) and the compression length (0.8 m), which consequently was taken to be the length of the fiber. Figure  3(c) shows the experimental setup, which is similar to the soliton compression stage reported in [13].

Pump pulses (FWHM duration ∼250 fs) were generated by a commercial 1030 nm ytterbium fiber laser at repetition rates between 50 kHz and 19.2 MHz and up to 100 W average power. A half-wave plate and a thin-film polarizer were used for power control, and a second half-wave plate corrected for potential nonlinear polarization rotation along the fiber, to ensure horizontal polarization after the fiber, to meet the requirements of subsequent optics and diagnostics. The pulses were compressed in an 8-capillary single-ring PCF [Fig.  3(d)] with a length of 80 cm, which was mounted in a gas-filled cell. The core-wall thickness of ∼190 nm was chosen to ensure that the anti-crossings lie far from the pump wavelength, so as not to impair the compression [22]. Light leaving the cell through the uncoated magnesium fluoride window was collimated using a silver-coated off-axis parabolic mirror. Reflecting the beam at two glass wedges lowered the intensity to prevent damage to the diagnostics. The compressed pulses were characterized with a custom-built, all-reflective, dispersion-free, second-harmonic generation frequency-resolved optical gating device (SHG FROG). To achieve transform limited pulses in spite of the dispersion introduced by the output window of the gas cell and propagation in air, we used double-angle chirped mirrors (group delay dispersion of −80 fs2) and thin fused silica wedges, which were only adjusted once at the beginning to exactly compensate the beam path dispersion.

We found that the calculated core radius, gas pressure and fiber length are exceptionally accurate for predicting the real experimental behavior, despite the high soliton order of 13.6, which produces extreme nonlinear propagation dynamics, with strong spectral broadening and fast temporal compression. Using 5.6 bar of argon, we could compress the pulses to a duration of 4.6 fs at 1 MHz repetition rate with a pulse energy of 4.3 µJ. As we always optimized the input pulse energy to yield the broadest output spectrum, we found that the maximum spectral width and the required pulse energy decreased with increasing repetition rate: 3.8 µJ at 5 MHz and 3.6 µJ at 10 MHz. As the maximum spectral width decreased, the pulse duration increased and we obtained at 5 MHz trains of pulses with duration 4.9 fs, and 10 MHz trains with duration 5.4 fs. The measured spectra and retrieved temporal pulse shapes for different repetition rates are shown in Figs.  4(a) and 4(c), the corresponding measured and retrieved FROG traces at 1 MHz in Fig.  4(d). We attribute the decreasing spectral width to changes in the transverse refractive index distribution caused by cumulative post-recombination heating, which significantly alters the dispersion and nonlinearity, thus limiting the compression [30,39]. The power delivered through the gas-cell during pulse compression, including losses at the fiber and input and output windows, varied depending on the repetition rate, having similar values at 1 MHz (86.8%), 5 MHz (84.5%), and 10 MHz (84.7%). The maximum output power was ∼30 W at 10 MHz. In the experiment a quality factor of ${Q_{\textrm{exp}}}\sim 0.24$ was achieved, very close to the expected value of ${\sim} 0.23$ for $N = 13.6$ obtained from the numerical fit.

 figure: Fig. 4.

Fig. 4. Experimental results. Spectra measured at different repetition rates at the output of the 80-cm-long single-ring PCF for (a) 5.6 bar of argon and (b) 52 bar of neon. (c) Retrieved temporal shapes of the compressed pulses at different repetition rates in argon and (d) corresponding measured and retrieved SHG FROG traces at a repetition rate of 1 MHz.

Download Full Size | PDF

We next investigated the dynamics in the lighter noble gas, neon. For this, we selected a similar zero-dispersion frequency, corresponding to ${\lambda _{\textrm{ZD}}} = 2\pi c/{\omega _{\textrm{ZD}}} \approx $ 509 nm, by filling the fiber with 52 bar (the maximum pressure the output window of the gas cell could safely sustain). Since the resulting nonlinearity was marginally lower than for the Ar-filled fiber, we used a slightly higher pulse energy to match the soliton order. When launching 5.2 µJ pulses at 1 MHz into the fiber, we measured output spectra slightly broader than for Ar at the same repetition rate, and slightly shorter pulse durations (4.5 fs). At higher repetition rate we found that, in contrast to argon, compression in neon was much more unstable. Already at 5 MHz the output was too unstable for full characterization by SHG FROG and at 10 MHz no stable output could be achieved. This behavior is in accord with our predictions. The weaker ionization of neon compared to argon results in a slightly higher temporal compression at low repetition rates. As the repetition rate increases, however, the longer diffusion time seems to lead to a higher post-recombination heating in neon, affecting the propagation dynamics more strongly than in argon. We attribute this to a stronger non-uniformity in transverse refractive index profile, which affects the nonlinearity and dispersion during pulse compression. A full understanding of these phenomena will require, however, further careful investigations.

4. Conclusion and outlook

In conclusion, an optimal region for soliton self-compression in hollow-core fibers can be identified in $({N,{\omega_{\textrm{ZD}}},{\tau_0}} )$ space. When the optical loss is negligible, e.g., in anti-resonant guiding PCF far from spectral anti-crossings with resonances in the core walls, the optimal region is enclosed within limits set by modulational instability, higher-order dispersion, self-focusing and photoionization. Of these, only photoionization depends on the choice of filling gas, becoming especially important when scaling to high repetition rates. The analytically derived limits are in excellent agreement with numerical simulations and reveal a characteristic triangular shape for the optimal compression region that shrinks with decreasing input pulse duration. Furthermore, we find that increasing the photoionization limit by employing lighter gases broadens the optimal region. However, matching ${\omega _{\textrm{ZD}}}$ and N of a heavier gas (e.g. argon) using a lighter gas (e.g. neon) results in longer diffusion times, leading to the thermal buildup of refractive index changes and impairing pulse compression.

When the core radius, gas pressure and fiber length are matched to the calculated values, good agreement is obtained between experiment and theory. Stable compression of 250 fs pulses at 1030 nm to almost single-cycle durations at repetition rates up to 10 MHz can be achieved experimentally in argon. As expected, the performance in neon is worse, stable compression being possible only at 1 MHz repetition rate.

We expect these results to be helpful in the design of high-performance soliton-compression systems. Proper selection of experimental parameters makes soliton self-compression highly convenient for compressing even long (∼250 fs) few-µJ pulses in a single compression stage down to few-cycle duration. The compression efficiency is similar to that achieved in two-stage systems based on gas-filled capillaries, though these require much higher pulse energies. The much lower pulse energy requirements allow operation at MHz instead of kHz repetition rates, allowing mode-locked pump lasers to be used.

Appendix 1

According to the capillary model [8], the propagation constant $\beta (\omega )$ of the HEnm modes in a gas-filled HC waveguide of radius a is given by:

$$\begin{array}{{c}} {\beta (\omega )= \frac{\omega }{c}{n_{\textrm{eff}}} \approx \frac{\omega }{c}\left[ {1 + \frac{{{\rho_r}{\chi^{(1 )}}}}{2} - \; \frac{1}{2}{{\left( {\frac{{{u_{n - 1,m}}c}}{{a\omega }}} \right)}^2}} \right],} \end{array}\; $$
where ${\rho _\textrm{r}}$ is the gas density relative to standard conditions, ${\chi ^{(1 )}}(\omega )$ the linear susceptibility of the gas, and ${u_{\textrm{qm}}}$ is the m-th zero of the q-th order Bessel function of the first kind. The group velocity dispersion and TOD follow from this expression as:
$$\begin{array}{{c}} {{\beta _2}(\omega )\approx \; \frac{{{\rho _\textrm{r}}}}{c}f(\omega )- \frac{{u_{n - 1,m}^2c}}{{{a^2}{\omega ^3}}},\; \; \; \; {\beta _3} \approx \frac{{{\rho _\textrm{r}}}}{c}{\partial _\omega }f(\omega )+ \frac{{3u_{n - 1,m}^2c}}{{{a^2}{\omega ^4}}},} \end{array}$$
$$\textrm{with}\; f(\omega )= {\partial _\omega }{\chi ^{(1 )}} + \; \frac{\omega }{2}\partial _\omega ^2{\chi ^{(1 )}}$$
where ${\partial _\omega }$ the derivative with respect to the frequency $\omega .$ If we consider only the HE11-like fundamental mode, we can relate the relative density to the zero dispersion frequency via ${\rho _\textrm{r}} = \; u_{01}^2{c^2}/[{{a^2}\omega_{\textrm{ZD}}^3f({{\omega_{\textrm{ZD}}}} )} ]$ with a obtained from Eq.  (9) and thus express ${\beta _2}$ and ${\beta _3}$ as functions of ${\omega _{\textrm{ZD}}}$:
$$\begin{array}{{c}} {{\beta _2}({\omega ,{\omega_{\textrm{ZD}}},\; a} )= \frac{{u_{01}^2c}}{{{a^2}}}\; \left( {\frac{{f(\omega )}}{{f({{\omega_{\textrm{ZD}}}} )}}\frac{1}{{\omega_{\textrm{ZD}}^3}} - \frac{1}{{{\omega^3}}}} \right) = \frac{{\delta ({\omega ,{\omega_{\textrm{ZD}}}} )}}{{{a^2}}}\; ,} \end{array}$$
$$\; {\beta _3}({\omega ,{\omega_{\textrm{ZD}}},\; a} )= \frac{{\; {\partial _\omega }\; \delta ({\omega ,{\omega_{\textrm{ZD}}}} )}}{{{a^2}}}.\; $$

Similarly, the nonlinear coefficient can also be expressed as function of ${\omega _{ZD}}$:

$$\begin{array}{{c}} {\gamma = \; \frac{{3{\omega _0}{\chi ^{(3 )}}{\rho _\textrm{r}}}}{{4{\varepsilon _0}{c^2}n_{\textrm{eff}}^2{A_{\textrm{eff}}}}} \approx \frac{{{\omega _0}{\chi ^{(3 )}}}}{{2{\varepsilon _0}{a^4}}}\frac{{u_{01}^2}}{{\omega _{\textrm{ZD}}^3f({{\omega_{\textrm{ZD}}}} )}},\; } \end{array}\; $$
where ${\varepsilon _0}$ is the vacuum permittivity, ${\chi ^{(3 )}}$ the nonlinear susceptibility, and we have used the approximation $n_{\textrm{eff}}^2 \approx 1$ and ${A_{\textrm{eff}}} \approx 1.5{a^2}$ [8].

Appendix 2

The clamping intensity is defined as the intensity ${I_\textrm{c}}$ at which the nonlinear phase shift induced by ionization equals the Kerr phase shift. This can be expressed as [28]:

$$\begin{array}{{c}} {{n_2}{I_\textrm{c}} = \; \frac{{{N_\textrm{e}}({{I_\textrm{c}}} )}}{{2{N_\textrm{c}}}},} \end{array}$$
where ${N_\textrm{c}}\; $ is the critical density and ${N_\textrm{e}}$ is the free-electron density, which can be calculated from the ionization rate. To obtain an analytical expression for ${I_\textrm{c}}$, we consider only multiphoton ionization (MPI), thus ${N_\textrm{e}}({{I_\textrm{c}}} )\approx {\sigma _\textrm{K}}I_\textrm{c}^K{\tau _0}{N_0}$, where ${N_0}$ is the density of neutral atoms, and ${\sigma _\textrm{K}}\; is\; $ the MPI cross-section for multi-photon absorption by K photons. The last two parameters can be derived from the PPT rate in the multi-photon limit [28]. Table  1 lists the MPI coefficients calculated at a wavelength of 1030 nm for the noble gases.

Tables Icon

Table 1. MPI cross-sections for the noble gases and for a wavelength of 1030 nm

By solving Eq.  (15), we obtain the following expression for the clamping intensity:

$$\begin{array}{{c}} {{I_c} = \; {{\left( {\frac{{2{n_2}{N_\textrm{c}}}}{{{\sigma_\textrm{K}}{\tau_0}{N_0}}}} \right)}^{{\raise0.7ex\hbox{$1$} \!\mathord{\left/ {\vphantom {1 {K - 1}}} \right.}\!\lower0.7ex\hbox{${K - 1}$}}}} = \; {{\left( {\frac{{3{\chi^{(3 )}}{N_\textrm{c}}{\rho_r}}}{{2{\varepsilon_0}cn_{\textrm{eff}}^2{\sigma_\textrm{K}}{\tau_0}{N_0}}}} \right)}^{{\raise0.7ex\hbox{$1$} \!\mathord{\left/ {\vphantom {1 {K - 1}}} \right.}\!\lower0.7ex\hbox{${K - 1}$}}}}.} \end{array}$$

Noting that ${N_0} \propto {\rho _\textrm{r}},\; $ this expression is only weakly pressure-dependent, via neff.

Funding

Max-Planck-Gesellschaft.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. M. Müller, A. Klenke, A. Steinkopff, H. Stark, A. Tünnermann, and J. Limpert, “3.5 kW coherently combined ultrafast fiber laser,” Opt. Lett. 43(24), 6037–6040 (2018). [CrossRef]  

2. M. Müller, C. Aleshire, A. Klenke, E. Haddad, F. Légaré, A. Tünnermann, and J. Limpert, “10.4 kW coherently combined ultrafast fiber laser,” Opt. Lett. 45(11), 3083–3086 (2020). [CrossRef]  

3. F. Calegari, G. Sansone, S. Stagira, C. Vozzi, and M. Nisoli, “Advances in attosecond science,” J. Phys. B: At. Mol. Opt. Phys. 49(6), 062001 (2016). [CrossRef]  

4. P. B. Corkum and F. Krausz, “Attosecond science,” Nat. Phys. 3(6), 381–387 (2007). [CrossRef]  

5. R. R. Gattass and E. Mazur, “Femtosecond laser micromachining in transparent materials,” Nat. Photonics 2(4), 219–225 (2008). [CrossRef]  

6. E. M. Dianov, P. V. Mamyshev, and A. M. Prokhorov, “Nonlinear fiber optics,” Quantum Electron. 18, 1–15 (1988). [CrossRef]  

7. L. F. Mollenauer, R. H. Stolen, J. P. Gordon, and W. J. Tomlinson, “Extreme picosecond pulse narrowing by means of soliton effect in single-mode optical fibers,” Opt. Lett. 8(5), 289–291 (1983). [CrossRef]  

8. 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,” J. Opt. Soc. Am. B 28(12), A11–A26 (2011). [CrossRef]  

9. N. Lilienfein, C. Hofer, M. Högner, T. Saule, M. Trubetskov, V. Pervak, E. Fill, C. Riek, A. Leitenstorfer, J. Limpert, F. Krausz, and I. Pupeza, “Temporal solitons in free-space femtosecond enhancement cavities,” Nat. Photonics 13(3), 214–218 (2019). [CrossRef]  

10. A. V. Mitrofanov, A. A. Voronin, M. V. Rozhko, D. A. Sidorov-Biryukov, A. B. Fedotov, A. Pugžlys, V. Shumakova, S. Ališauskas, A. Baltuška, and A. M. Zheltikov, “Self-compression of high-peak-power mid-infrared pulses in anomalously dispersive air,” Optica 4(11), 1405–1408 (2017). [CrossRef]  

11. G. Jargot, N. Daher, L. Lavenu, X. Delen, N. Forget, M. Hanna, and P. Georges, “Self-compression in a multipass cell,” Opt. Lett. 43(22), 5643–5646 (2018). [CrossRef]  

12. 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(8), 547–554 (2019). [CrossRef]  

13. 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(7), 9099–9110 (2020). [CrossRef]  

14. T. Nagy, S. Hädrich, P. Simon, A. Blumenstein, N. Walther, R. Klas, J. Buldt, H. Stark, S. Breitkopf, P. Jójárt, I. Seres, Z. Várallyay, T. Eidam, and J. Limpert, “Generation of three-cycle multi-millijoule laser pulses at 318 W average power,” Optica 6(11), 1423–1424 (2019). [CrossRef]  

15. M. Gebhardt, C. Gaida, S. Hädrich, F. Stutzki, C. Jauregui, J. Limpert, and A. Tünnermann, “Nonlinear compression of an ultrashort-pulse thulium-based fiber laser to sub-70 fs in Kagome photonic crystal fiber,” Opt. Lett. 40(12), 2770–2773 (2015). [CrossRef]  

16. 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(6), 2426–2430 (2018). [CrossRef]  

17. T. Balciunas, C. Fourcade-Dutin, G. Fan, T. Witting, A. A. Voronin, A. M. Zheltikov, F. Gerome, G. G. Paulus, A. Baltuska, and F. Benabid, “A strong-field driver in the single-cycle regime based on self-compression in a kagome fibre,” Nat. Commun. 6(1), 6117 (2015). [CrossRef]  

18. U. Elu, M. Baudisch, H. Pires, F. Tani, M. H. Frosz, F. Köttig, A. Ermolov, P. St.J. Russell, and J. Biegert, “High average power and single-cycle pulses from a mid-IR optical parametric chirped pulse amplifier,” Optica 4(9), 1024 (2017). [CrossRef]  

19. C. Brahms, F. Belli, and J. C. Travers, “Infrared attosecond field transients and UV to IR few-femtosecond pulses generated by high-energy soliton self-compression,” Phys. Rev. Res. 2(4), 043037 (2020). [CrossRef]  

20. C. M. Heyl, H. Coudert-Alteirac, M. Miranda, M. Louisy, K. Kovacs, V. Tosa, E. Balogh, K. Varjú, A. L’Huillier, A. Couairon, and C. L. Arnold, “Scale-invariant nonlinear optics in gases,” Optica 3(1), 75–81 (2016). [CrossRef]  

21. G. P. Agrawal, Nonlinear Fiber Optics (Academic Press, 2007).

22. F. Tani, F. Köttig, D. Novoa, R. Keding, and P. St.J. Russell, “Effect of anti-crossings with cladding resonances on ultrafast nonlinear dynamics in gas-filled photonic crystal fibers,” Photonics Res. 6(2), 84–88 (2018). [CrossRef]  

23. P. St.J. Russell, P. Hölzer, W. Chang, A. Abdolvand, and J. C. Travers, “Hollow-core photonic crystal fibres for gas-based nonlinear optics,” Nat. Photonics 8(4), 278–286 (2014). [CrossRef]  

24. C.-M. Chen and P. L. Kelley, “Nonlinear pulse compression in optical fibers: scaling laws and numerical analysis,” J. Opt. Soc. Am. B 19(9), 1961–1967 (2002). [CrossRef]  

25. F. Tani, J. C. Travers, and P. St.J. Russell, “Multimode ultrafast nonlinear optics in optical waveguides: numerical modeling and experiments in kagomé photonic-crystal fiber,” J. Opt. Soc. Am. B 31(2), 311 (2014). [CrossRef]  

26. J. M. Dudley, G. Genty, and S. Coen, “Supercontinuum generation in photonic crystal fiber,” Rev. Mod. Phys. 78(4), 1135–1184 (2006). [CrossRef]  

27. H. J. Lehmeier, W. Leupacher, and A. Penzkofer, “Nonresonant third order hyperpolarizability of rare gases and N2 determined by third harmonic generation,” Opt. Commun. 56(1), 67–72 (1985). [CrossRef]  

28. A. Couairon and A. Mysyrowicz, “Femtosecond filamentation in transparent media,” Phys. Rep. 441(2-4), 47–189 (2007). [CrossRef]  

29. A. Crego, E. Conejero Jarque, and J. San Roman, “Influence of the spatial confinement on the self-focusing of ultrashort pulses in hollow-core fibers,” Sci. Rep. 9(1), 9546 (2019). [CrossRef]  

30. J. R. Koehler, F. Köttig, D. Schade, P. St.J. Russell, and F. Tani, “Post-recombination effects in confined gases photoionized at megahertz repetition rates,” Opt. Express 29(4), 4842–4857 (2021). [CrossRef]  

31. B. M. Trabold, M. I. Suresh, J. R. Koehler, M. H. Frosz, F. Tani, and P. St.J. Russell, “Spatio-temporal measurement of ionization-induced modal index changes in gas-filled PCF by prism-assisted side-coupling,” Opt. Express 27(10), 14392–14399 (2019). [CrossRef]  

32. G. Tempea and T. Brabec, “Theory of self-focusing in a hollow waveguide,” Opt. Lett. 23(10), 762–764 (1998). [CrossRef]  

33. G. Fibich and A. L. Gaeta, “Critical power for self-focusing in bulk media and in hollow waveguides,” Opt. Lett. 25(5), 335–337 (2000). [CrossRef]  

34. R. T. Chapman, T. J. Butcher, P. Horak, F. Poletti, J. G. Frey, and W. S. Brocklesby, “Modal effects on pump-pulse propagation in an Ar-filled capillary,” Opt. Express 18(12), 13279–13284 (2010). [CrossRef]  

35. F. Köttig, F. Tani, J. C. Travers, and P. St.J. Russell, “Self-focusing below the critical power in gas-filled hollow-core PCF,” in 2017 Conference on Lasers and Electro-Optics Europe European Quantum Electronics Conference (CLEO/Europe-EQEC) (2017), p. 1.

36. F. Köttig, F. Tani, and P. St.J. Russell, “Modulational-instability-free pulse compression in anti-resonant hollow-core photonic crystal fiber,” Opt. Lett. 45(14), 4044–4047 (2020). [CrossRef]  

37. J. Lægsgaard, “Efficient simulation of multimodal nonlinear propagation in step-index fibers,” J. Opt. Soc. Am. B 34(10), 2266–2273 (2017). [CrossRef]  

38. C. Vozzi, M. Nisoli, G. Sansone, S. Stagira, and S. De Silvestri, “Optimal spectral broadening in hollow-fiber compressor systems,” Appl. Phys. B 80(3), 285–289 (2005). [CrossRef]  

39. F. Köttig, J. R. Koehler, F. Tani, and P. St.J. Russell, “Inter-Pulse Dynamics at High Repetition Rates by Femtosecond Ionization in Gas-Filled Fibres,” in 2019 Conference on Lasers and Electro-Optics Europe European Quantum Electronics Conference (CLEO/Europe-EQEC) (2019), p. 1.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

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

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. Scaling behavior of soliton self-compression and optimal parameter region for 1030 nm pulses. (a) Dependence of compression length on soliton order in argon for zero-dispersion wavelength $2\mathrm{\pi }\textrm{c}/{\omega _{\textrm{ZD}}} \approx 450\; \textrm{nm},$ with FWHM input pulse duration 25 fs (dark blue), 100 fs (orange) and 250 fs (turquoise). The symbols mark numerically calculated datapoints (position of highest peak power), and the full lines are solutions of Eq.  (2). For comparison, the fission length ${L_{\textrm{fiss}}} = {L_\textrm{D}}/N$ is also plotted (dashed lines). (b) Dependence of compression factor ${F_\textrm{c}}$ on soliton order for the same case as in (a). The symbols mark numerically calculated datapoints, and the full lines are solutions of Eq.  (3). The red dot-dashed line is a linear fit to ${F_\textrm{c}} = 4.4N - 4$ . (c) Analytical plot of soliton order versus ${\tau _0}{\mathrm{\Delta }_\omega }/({2\pi } )$ for 100 fs input pulses with ${\mathrm{\Delta }_\omega } = |{{\omega_{ZD}} - {\omega_0}} |$ . MI: modulational instability (turquoise dashed line); SF: self-focusing (dark blue dotted line); ION: photoionization (yellow to red solid lines); HOD: higher-order dispersion (blue dashed line). The size of the gray area, within which optimal compression occurs, is controlled by the ION limit, which depends on the gas species.
Fig. 2.
Fig. 2. Ratio of peak compressed intensity I to peak launched intensity I0 for 1030 nm pulses in Ar-filled HC-PCF, plotted against soliton order and ${\tau _0}{\mathrm{\Delta }_\omega }/({2\pi } )$ , which depends on both core diameter and pressure. Initial pulse durations are (a) 250 fs, (b) 100 fs and (c) 25 fs. The optimal compression region is triangular in shape (yellow-white area) and shrinks with decreasing input pulse duration. At 1030 nm the maximum soliton order is limited mainly by HOD and ION for shorter pulses, whereas for longer pulses MI constrains the maximum soliton order to $N = 16$ . The white dot in (a) marks the experimental parameters.
Fig. 3.
Fig. 3. Single-stage pulse compression to few-cycle durations. (a) Diffusion time as a function of the zero-dispersion frequency for different noble gases. The dashed gray line marks the experimental setpoint at $2\mathrm{\pi }\textrm{c}/{\omega _{\textrm{ZD}}} \approx 523\; \textrm{nm}$ . The upper horizontal axis allows comparison with Fig.  1(c); however, note that the diffusion time is independent of the input pulse duration. (b) Scaling of core radius, gas pressure and compression length with pulse energy for 1030 nm, 250 fs pulses in argon at a fixed soliton order of 13.6. The dashed gray line marks the experimental parameters, where a core radius of 15 µm, a pressure of 5.6 bar and a fiber length of 0.8 m are required to compress pulses with an energy of 4 µJ. (c) Sketch of the experimental setup: λ/2: half-wave plate, TFP: thin-film polarizer, OAPM: off-axis parabolic mirror, W: wedge, CM: chirped-mirror. (d) Scanning electron micrograph of the cross-section (zoom) of the 80-cm-long, 8-capillary single-ring PCF with core diameter of 30 µm.
Fig. 4.
Fig. 4. Experimental results. Spectra measured at different repetition rates at the output of the 80-cm-long single-ring PCF for (a) 5.6 bar of argon and (b) 52 bar of neon. (c) Retrieved temporal shapes of the compressed pulses at different repetition rates in argon and (d) corresponding measured and retrieved SHG FROG traces at a repetition rate of 1 MHz.

Tables (1)

Tables Icon

Table 1. MPI cross-sections for the noble gases and for a wavelength of 1030 nm

Equations (19)

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

L D = τ 0 2 | β 2 ( ω 0 ) | , L TOD = τ 0 3 | β 3 ( ω 0 ) | , L NL = 1 γ P 0 , with γ = ω 0 n 2 c A eff ,
L comp = ( 1 2 N + 1.7 N 2 ) L D = A N L D
F c = τ 0 τ c = 3 A N ( 1 N ξ ) ,  with  ξ = β 3 ( ω 0 ) τ 0 β 2 ( ω 0 ) = ω δ ( ω 0 , ω Z D ) τ 0 | δ ( ω 0 , ω Z D ) |
1 τ c | ω ZD ω 0 | 0.315 π ln ( 3 + 8 ) Δ ω σ ,
N ZD max ( ω 0 , ω ZD , τ 0 ) = Δ ω σ τ 0 1 3.4 ξ + ( 1 + 3.4 ξ ) 2 + 81.6 / ( Δ ω σ τ 0 ) 2 ( 6 + ξ Δ ω σ τ 0 ) .
N SF max ( ω 0 , ω ZD , τ 0 ) = τ 0 2 π c S ω 0 | δ ( ω 0 , ω ZD ) | ,
N ION max ( ω 0 , ω ZD , τ 0 ) = τ 0 3 u 01 2 I c ω 0 χ ( 3 ) 4 ε 0 ω ZD 3 f ( ω ZD ) | δ ( ω 0 , ω ZD ) | ,
N loss min = 3 4 a | δ ( ω 0 , ω ZD ) | ( c u 01 τ 0 ω 0 ) 2 [ 1 + 1 + 9 a | δ ( ω 0 , ω ZD ) | ω 0 2 ( c u 01 τ 0 ) 2 ] .
t diff = C p κ M k B N A p 0 T 0 u 01 2 c 2 f ( ω ZD ) ω ZD 3
a = χ ( 3 ) u 01 2 4 ε 0 ω 0 ω ZD 3 | δ ( ω 0 , ω ZD ) | f ( ω ZD ) τ 0 E N 2 ,
p = u 01 2 c 2 a 2 ω ZD 3 1 f ( ω ZD ) T T 0 p 0 ,
β ( ω ) = ω c n eff ω c [ 1 + ρ r χ ( 1 ) 2 1 2 ( u n 1 , m c a ω ) 2 ] ,
β 2 ( ω ) ρ r c f ( ω ) u n 1 , m 2 c a 2 ω 3 , β 3 ρ r c ω f ( ω ) + 3 u n 1 , m 2 c a 2 ω 4 ,
with f ( ω ) = ω χ ( 1 ) + ω 2 ω 2 χ ( 1 )
β 2 ( ω , ω ZD , a ) = u 01 2 c a 2 ( f ( ω ) f ( ω ZD ) 1 ω ZD 3 1 ω 3 ) = δ ( ω , ω ZD ) a 2 ,
β 3 ( ω , ω ZD , a ) = ω δ ( ω , ω ZD ) a 2 .
γ = 3 ω 0 χ ( 3 ) ρ r 4 ε 0 c 2 n eff 2 A eff ω 0 χ ( 3 ) 2 ε 0 a 4 u 01 2 ω ZD 3 f ( ω ZD ) ,
n 2 I c = N e ( I c ) 2 N c ,
I c = ( 2 n 2 N c σ K τ 0 N 0 ) 1 / 1 K 1 K 1 = ( 3 χ ( 3 ) N c ρ r 2 ε 0 c n eff 2 σ K τ 0 N 0 ) 1 / 1 K 1 K 1 .
Select as filters


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