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

Determining the absolute temporal field of ultra-broadband terahertz-infrared pulses with field-induced second-harmonic spectrograms

Open Access Open Access

Abstract

We demonstrate the use of spectrograms of the field-induced second-harmonic (FISH) signal generated in ambient air, to reconstruct the absolute temporal electric field of ultra-broadband terahertz-infrared (THz-IR) pulses with bandwidths exceeding 100 THz. The approach is applicable even with relatively long (150-femtosecond) optical detection pulses, where the relative intensity and phase can be extracted from the moments of the spectrogram, as demonstrated by transmission spectroscopy of very thin samples. Auxiliary EFISH/ABCD measurements are used to provide the absolute field and phase calibration, respectively. We take into account the beam-shape/propagation effects about the detection focus on the measured FISH signals, which affect the field calibration, and show how an analysis of a set of measurements vs. truncation of the unfocused THz-IR beam can be used to correct for these. This approach could also be applied to the field calibration of ABCD measurements of conventional THz pulses.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The availability of continuum pulses simultaneously covering both the THz and IR spectral ranges opens the way for new spectroscopies with pumping [1,2] and/or probing [3,4] of low-energy excitations, e.g. intra- and inter-band conductivity, phonons and other quanta in complex solids with femtosecond (fs) time resolution. The generation of such pulses can employ nonlinear mixing of the frequency components of fs near-IR (NIR) optical pulses in either ${\chi ^{(2)}}$- or ${\chi ^{(3)}}$-media. For ${\chi ^{(2)}}$-crystals, the use of ultra-broadband NIR pulses from two independent OPAs and mixing in suitable crystals (e.g. GaSe) can yield bandwidths reaching 100 THz [5], and very high peak fields (reaching ${\sim }10~{\,\mathrm {MVcm}^{-1}}$ [6] or even ${\sim }100~{\,\mathrm {MVcm}^{-1}}$ for narrow-band mid-IR pulses [7]), although the requirement of a crystalline medium introduces issues due to phonon absorption, phase-matching, damage threshold, and interference/losses due to surface reflections. In contrast, pulse generation in a gas plasma with an ultra-broadband, two-color (${\omega {-}2\omega }$) NIR-UV pump [812] can reach bandwidths exceeding 100 THz, also with a stable carrier envelope phase, without any inherent damage threshold or low-frequency cut-off in the conventional THz range, and field strengths readily exceeding 1 ${\,\mathrm {MVcm}^{-1}}$ for sufficiently high pump energy [1320].

Simultaneous coherent detection of the full THz-IR bandwidth is non-trivial. Time-domain electro-optic (EO) field sampling in ultra-thin ${\chi ^{(2)}}$-crystals [57] with extremely short optical detection pulses can provide a direct determination of the temporal field (provided the EO response is carefully taken into account [21]), albeit with similar issues for ${\chi ^{(2)}}$-emitters mentioned above. Moreover, sampling the intra-cycle field for bandwidths exceeding 100 THz places severe requirements on the pulse duration – as shown in Fig. 1(a), this essentially requires an optical pulse duration below 10${\,{\mathrm {fs}}}$. One can also use ${\chi ^{(3)}}$-media as the non-linear medium: in particular a combination of terahertz- and electric-field-induced second harmonic generation (TFISH, EFISH) in a gas medium comprises the method of air‐biased coherent detection (ABCD) detection [22,23], now widely employed for conventional THz time-domain detection. While the higher-order nonlinearity requires higher optical intensities, similar detection efficiencies can be achieved with amplifier-laser pulses. The ABCD method was applied to THz-IR transients [20], although with limited sensitivity due to a strong frequency roll-off. One attractive feature of ABCD detection is that the DC electric field $\cal {E}_{\textrm {b}}$ used for EFISH, in principle, provides an absolute field strength with which to calibrate the field of the THz pulse (as opposed to deducing the field from the crystal ${\chi ^{(2)}}$-value and optical retardation in EO detection). As discussed by others [22,2426] (and in this report for the full THz-IR range), one must correctly account for the propagation geometries for both EFISH and TFISH to correctly use the EFISH signal for such calibration (as opposed to directly using the signal ratios [27]).

 figure: Fig. 1.

Fig. 1. (a) Response function for electro-optic time-domain sampling for different optical detection pulse duration, $T_{\mathrm {FWHM}}$ ($S(\nu )=\exp (-\tfrac {\pi ^2}{4\ln {2}}\nu ^2{T_{\mathrm {FWHM}}}^2)$). (b) Schematic of frequency conversion processes for detection of THz-IR pulses, including FW-DFG (equivalent to TFISH for THz frequencies) used here in air. (c) Setup used for spectrogram/ABCD experiments. (d) Example spectrogram using detection pulses with ${T_{\mathrm {FWHM}}}=150{\,{\mathrm {fs}}}$.

Download Full Size | PDF

If one forgoes field detection with sub-cycle resolution, an alternative is to spectrally resolve the signal from mixing the THz-IR and optical-detection pulses. Indeed, such frequency conversion techniques date well back in the history of time-resolved detection of femtosecond pulses. Again, one can employ ${\chi ^{(2)}}$- or ${\chi ^{(3)}}$-media, as depicted in Fig. 1(b). The mixing of the THz-IR and NIR pulses generates sum-frequency (SFG) and difference-frequency (DFG) sidebands which can be detected with an optical spectrometer, and as a function of delay to yield a spectrogram. While the classical method is to employ SFG/DFG in ${\chi ^{(2)}}$-media (which we also employed previously to detect THz-IR pulses [2830]), this has again the issues of using crystalline media, and the SF-/DF-sidebands also overlap with the input optical spectrum for THz frequencies. However, it was realized in recent years [17,19,31,32] that one can extend the TFISH approach to cover the whole THz-IR range – the resulting four-wave (FW) processes between the fundamental optical ($E_1$) and THz-IR (${\mathcal {E}}$) fields also referred to as FW-SFG/DFG – with detection of the signal ($E_2$) in the UV-vis range. As established for TFISH/ABCD [2325,33] and discussed in detail in this paper, the FW-SFG process suffers destructive interference during co-propagation through the focal region [32] due to the relative Gouy-phase evolution of the nonlinear signal polarization $P_{\mathrm{SF}}^{(3)} \propto E_1^2 \mathcal{E}$ and generated field $E_2$ (also well-known for optical third-harmonic generation [34]), although the FW-DFG signal with $P_{\mathrm{DF}}^{(3)} \propto E_1^2 \mathcal{E}^*$ survives.

The resulting FW-DFG spectra can be acquired as a function of delay $\tau$ between $E_1$ and ${\mathcal {E}}$ to build up a spectrogram, which also contains temporal information (relative phase) about ${\mathcal {E}}$, being essentially an XFROG (cross‐correlation frequency resolved optical gating) spectrogram of ${\mathcal {E}}(t)$ with a gate pulse $a(t)=E_1^2(t)$ [19] (similar to that used in the infrared‐UV range [35,36], only here the THz‐IR spectrum extends down to zero‐frequency). Here several modalities have been demonstrated to recover the THz-IR pulse, including the use of chirped optical detection pulses $E_1$ [17,37] (which facilitates single-shot detection and can achieve a finer frequency resolution than the bandwidth of the gate pulse, but requires subsequent phase correction for the retrieved data) or conventional XFROG deconvolution with short detection pulses (e.g. with 30-fs duration [19]). The ability to recover the sub-cycle field profile (i.e. to determine the absolute, or carrier-envelope, phase), as achieved with EO time-domain sampling [57], is highly desirable, especially for non-linear THz-IR spectroscopy. Here an elegant approach was demonstrated [19], whereby an auxiliary ABCD measurement in the same detection scheme delivers the field profile $\mathcal{E}_{\mathrm{THz}}(t)$ of the low-frequency range of ${\mathcal {E}}$, hence providing the absolute phase offset for the latter determined from the FW-DFG spectrogram.

In this paper, we combine and adapt these approaches to estimate the absolute temporal electric field profile ${\mathcal {E}}(t)$ of THz-IR pulses from a two-color air plasma. Here we measure FW-DFG spectrograms with relatively long (150-fs) detection pulses, which we show can still be used to extract the relative spectral intensity and phase of the THz-IR transients (with a duration which can reach below 10 fs) even without explicit XFROG deconvolution, while the absolute phase offset is provided by an auxiliary ABCD measurement of $\cal {E}_{\textrm {THz}}$ as per [19]. We introduce for the calibration of the absolute field a novel approach where we also analyze the relative EFISH and FW-DFG signal intensities and take into account propagation effects (i.e the amplitude- and phase-development of the on-axis input- and signal-fields propagating through the detection focal region, including Gouy- and dispersive phase mismatch) to determine the frequency-dependent FW-DFG response relative to the EFISH response. As this requires determination of the focal behavior of the THz-IR field through the focus [26], we performed additional measurements of the FW-DFG signal vs. radial truncation of the THz-IR beam before focusing, which we show can provide sufficient information to quantify the propagation effects. This method also has potential for absolute field calibration of ABCD measurements of conventional-bandwidth THz pulses.

2. Experimental details

A schematic of the experimental setup is shown in Fig. 1(c), based on a 1-kHz Ti:Al$_2$O$_3$ amplifier laser (Clark-MXR CPA-2101, $\lambda _0=775$ nm, ${T_{\mathrm {FWHM}}}=150{\,{\mathrm {fs}}}$, pulse energy $900{\,\mathrm{\mu}}\textrm {J}$). As in [16,30], we spectrally broaden these pulses in an Ar-gas capillary (2.1 bar) to achieve optical bandwidth corresponding to ${T_{\mathrm {FWHM}}}<20{\,{\mathrm {fs}}}$ and pulse energy ${\sim }400{\,\mathrm{\mu}} \textrm {J} $ with subsequent compression with negative-dispersion mirrors. These pulses are focused in ambient air (${f_{\mathrm {p}}}=200$ mm) via a 150-$\mathrm{\mu}$m-thick $\beta$-BBO crystal ($32^\circ$-cut for type-I second-harmonic generation, SHG) to achieve ${\omega {-}2\omega }$ plasma emission. Note that the orientation angles (and position) of the SHG crystal are carefully optimized to yield the resultant THz-IR bandwidth, which, as we previously reported [16], depends strongly on the resultant frequency detuning of the $\omega$- and $2\omega$-waves. The emitted THz-IR beam passes a small axial stop (to block the $\omega$ and $2\omega$ pump beams), and is collimated by an off-axis paraboloidal mirror (OAPM, protected Al, eff. focal length $76.2$ mm). The beam is focused through a position for various spectroscopy measurements and re-collimated (protected Au, eff. focal lengths 101.6 mm), and then focused for coherent detection (OAPM, $f=76.2$ mm) with a variably delayed co-propagating 150-fs optical pulse ($1/e^2$-beam diameter 3.6 mm, energy $40{\,\mathrm{\mu}} \textrm {J}$, focused by a lens with $f_1=125$ mm through a hole in the last OAPM). Note that the indium-tin-oxide mirror is used to introduce on-axis optical pump beams for other experiments, while a 50-$\mathrm{\mu}$m-thick high-resistivity Si wafer in the collimated beam path suppresses residual $\omega$ and $2\omega$ plasma pump light.

The focal region is equipped with cylindrical electrodes (diameter 2.5 mm with flat end surfaces, spacing 0.45 mm) to apply a 1-kHz/500-Hz sinusoidal electric field with peak amplitude $\cal {E}_{\textrm {b}}$ = 25.8 kV cm-1 used for the 1-kHz EFISH calibration measurements (as confirmed with high-voltage probes directly on the electrodes, remaining below air breakdown) to allow the generation of an additional ABCD/EFISH detection signal. Following the detection focus, the signal beam passes dichroic filters to suppress the input optical beam $E_1$, is recollimated and coupled with variable split ratio to a compact UV-vis spectrometer (Avantes Avaspec-HSC1024x58TEC-EVO, 335-510 nm, resolution <0.4 nm) for FW-DFG/EFISH spectral measurements, and a photomultiplier tube (Hamamatsu H9656-04) for ABCD measurements. The spectral response of the spectrometer is flat over the FW-DFG spectral range, and the spectra are used without further wavelength correction. The complete setup is purged with dry air to minimize THz/IR absorption from water vapor.

3. Results

3.1 Extraction of THz-IR spectrum and chirp

For our approach, we employ a relatively long (150-fs), transform-limited optical detection pulse to measure the FW-DFG spectrogram $S_T(\nu,\tau )$, in contrast to previous reports (e.g. 30 fs in [19]). There are several motivations for this: (i) the resultant spectral convolution of the THz-IR spectrum $I(\nu )$ with that of the detection gate pulse still yields a reasonable spectral resolution without deconvolution; (ii) for pump-probe experiments, one can use a constant detection delay $\tau =\tau _0$ (as the rather long gate pulse can maintain temporal overlap with the THz-IR pulse) and only scan the pump-probe delay ${{{\tau }_{\mathrm {ex}}}}$ to acquire $\Delta I(\nu ;{{{\tau }_{\mathrm {ex}}}})$ via $S_T(\nu,\tau _0;{{{\tau }_{\mathrm {ex}}}})$ [30]; (iii) one can more readily treat propagation effects in the FW-DFG detection process (Sec. 3.3) when the optical gate is reasonably narrow-band. Nevertheless, given that the THz-IR pulses have a duration on the order of 10 fs, one could question whether any robust retrieval of temporal information is possible with such relatively long detection pulses. We demonstrate that this is the case in this section.

A measured reference FW-DFG spectrogram $S_T(\nu ',\tau )$ is shown in Fig. 1(d), where $\nu '=2\nu _0-\nu _S$ is the DF offset (i.e. THz-IR) frequency, $\nu _S=c/\lambda$ the measured signal frequency, and $\nu _0=c/\lambda _0$ ($\lambda _0$ the center wavelength of the Ti:Al$_2$O$_3$ amplifier laser). A small portion of the FW-SFG range ($\nu ' < 0$) is also shown, where one sees that the signal is strongly suppressed (as mentioned above and discussed in detail below). Hereon in this section, we will drop the dash and write the DF frequency as $\nu$ to improve readability, although in later sections (beginning in Sec. 3.3 to the end) $\nu =\nu _S-2\nu _0$ will carry a sign to describe both FW-SFG ($\nu >0$) and FW-DFG ($\nu =-\nu '<0$). In order to extract spectral and temporal information from the spectrogram, we employ standard spectrogram analysis methods [30], where we first neglect propagation effects (treated below) and assume an ideal spectrogram signal (Appendix 5.1). To estimate the THz-IR intensity spectrum $I(\nu )$, we calculate the spectral marginal $M(\nu )=\smallint \mathrm {d}\tau \, S_T(\nu,\tau )$ (integration over $(-\infty,\infty )$ implicit), which as derived in Eq. (9) of Appendix 5.1 yields $M(\nu )\propto I(\nu )\ast G(\nu )$, i.e. the spectrum convolved with the detection gate spectrum $G(\nu )=|a(\nu )|^2$ where $a=E_1^2$ (here $G(\nu )$ has a FWHM of 2.6 ${\mathrm {THz}}$, as determined from the EFISH spectrum in Sec. 3.2 below). The spectral marginal $M(\nu )$ is shown in Fig. 1(d) (to the left of the spectrogram), and demonstrates a reasonably continuous THz-IR spectrum with a 10%-bandwidth of ${\sim }100\;{\mathrm {THz}}$, reaching the noise floor near 150 ${\mathrm {THz}}$. The corresponding temporal marginal $M(\tau )=\smallint \mathrm {d}\nu \, S_T(\nu,\tau )$ is also shown below the spectrogram, and as expected reflects only the temporal profile of the optical detection pulse $E_1$. However, as employed previously [30], we can also extract the temporal first moment ${\bar \tau }(\nu )=\smallint \mathrm {d}\tau \, \tau \cdot S_T(\nu,\tau )$ from the spectrogram, which one can show (Eq. (11)) yields ${\bar \tau }(\nu ) = \langle {T_{\mathrm {g}}}(\nu )\rangle$, i.e. the spectral group delay ${T_{\mathrm {g}}}(\Omega )=-\partial _\Omega \varphi (\Omega )$ ($\Omega =2\pi \nu$) convolved with $G$. In Fig. 1(d), we plot ${\bar \tau }$ superposed on the spectrogram, which is seen to vary by less than 20 ${\mathrm {fs}}$ across the full bandwidth, with small dispersive deviations due to residual H$_2$O/CO$_2$ absorption lines, mainly the H$_2$O bending mode at 45 THz (1500 cm$^{-1}$) and CO$_2$ asymmetric stretch mode at 70 THz (2300 cm$^{-1}$), also seen in $M(\nu )$. This weak dispersion is mainly due to the thin (50-$\mathrm{\mu}$m) Si wafer in the beam (Sec. 2), where one can calculate a group dispersive delay of 180 fs mm$^{-1}$ [38] over the range 0-100 ${\mathrm {THz}}$ (despite the absence of dispersion from allowed-phonon contributions – note that Ge has an even higher dispersion).

In order to demonstrate the sensitivity of ${\bar \tau }$, we performed transmission measurements on a thin (15-$\mathrm{\mu}$m) polypropylene (PP) membrane at the sample focus: the reference and sample spectrograms are shown in Fig. 2(a) and (b), respectively. The relative intensity transmission $T(\nu )=M(\nu )/{M_{\mathrm {ref}}}(\nu )$ and group delay $\Delta {T_{\mathrm {g}}}(\nu )={\bar \tau }(\nu )-{{\bar \tau }_{\mathrm {ref}}}(\nu )$ extracted from the spectrograms are shown in Fig. 2(c), along with model results calculated using literature data for the PP vibrational absorption lines [39] (specifically, CH$_{2,3}$ bending modes at ${\sim }40$-$45\;{\mathrm {THz}}$ and stretching modes at ${\sim }85$-$90{\,{\mathrm {THz}}}$) and accounting for the 2.6-THz frequency resolution. As can be seen, the spectra estimated from the spectrogram analysis are quite faithful to the model data. In particular, the expected etalon interference effects are clearly resolved in both $T(\nu )$ and $\Delta {T_{\mathrm {g}}}(\nu )$, the latter demonstrating a relative sensitivity below 10 fs. Note that the model calculations are based on a simple Lorentzian-band model of the main absorption lines with a constant background refractive index $n_{\infty }=1.45$ of PP, such that some of the residual deviations seen in Fig. 2(c) are due to neglecting weaker bands (moreover, the truncation of $\Delta {T_{\mathrm {g}}}$ about the stretching modes arises due to the finite $t$-scan range of the spectrograms).

 figure: Fig. 2.

Fig. 2. (a) Reference spectrogram, and (b) spectrogram with 15-$\mu$m-thick polypropylene film in sample position. Additional chirp for reference/sample (compared to Fig. 1(d)) due to the inclusion of an additional 525-$\mu$m-thick high-resistivity silicon wafer in THz-IR beam path for these measurements. (c) Corresponding intensity transmission and group delay of polypropylene sample extracted from $M(\nu )$ and ${\bar \tau }(\nu )$ of the reference/sample spectrograms (solid curves), and model data using literature IR absorption line data [39].

Download Full Size | PDF

Having established that ${\bar \tau }$ provides a reliable estimator of ${T_{\mathrm {g}}}$, and hence also the relative spectral phase via $\varphi (\Omega )=-\smallint ^{\Omega }\mathrm {d}\Omega '\,{T_{\mathrm {g}}}(\Omega ')$, a determination of the low-frequency absolute phase is sufficient to the absolute phase $\Phi (\nu )$ for the full THz-IR range [19].

3.2 Field and phase calibration using EFISH/ABCD

We now proceed to use auxiliary EFISH/ABCD measurements to determine the absolute phase and temporal field of the THz-IR pulses, at first without the correction for propagation effects which will be treated in the next section. While the EFISH signals are typically only measured in a spectrally integrated way via their interferometric contribution to an ABCD signal in conventional THz time-domain detection (with alternating bias field polarity $\pm \cal {E}_{\textrm {b}}$), one can also couple the EFISH signal to the spectrometer as per the FW-DFG signal. This is demonstrated in Fig. 3, which shows spectrograms for (a) FW-DFG alone ($S_T(\nu,\tau )$) and (b) with the addition of the EFISH signal ($S_{TE}(\nu,\tau )$). For the latter, we use a 1-kHz frequency for the sinusoidal bias field, such that $\cal {E}_{\textrm {b}}$ is unipolar for each laser shot, and the optical phase of the EFISH signal is constant (and not alternating, as for ABCD). As expected from the principle of ABCD, given that the two TFISH/EFISH signal fields, $b_T$ and $b_E$ respectively, are coherent (see Appendix 5.1 for their mathematical form), they exhibit interference depending on their relative phase, as can be scrutinized by forming the residual $\Delta S_{TE}=S_{TE} - S_{T} - S_{E}$ ($S_{E}$ clearly being independent of $\tau$). The resulting spectrogram $\Delta S_{TE}(\nu,\tau )$ is shown in Fig. 3(d) (with corresponding marginals included in Fig. 3(c), where we scale $M_E = N_\tau \cdot S_{E}$ (where $N_\tau$ is the number of delays in each spectrogram), which shows that the interference is constructive for the lowest THz frequencies (for the laboratory direction of EFISH bias chosen). In principle, this measurement is sufficient to deduce the absolute phase $\Phi$ of the low-frequency range of the THz-IR pulses, seeing as (neglecting the spectral convolution with the optical detection gate)

$$S_{TE}(\nu,\tau)\propto|b_T(\nu,\tau)+b_E(\nu)|^2 = |b_T(\nu,\tau)|^2 + |b_E(\nu)|^2 + 2 |b_T(\nu,\tau)b_E(\nu)|\cos\Phi(\nu).$$

However, to circumvent some of the issues with the spectral convolution, we adopt here the same approach as in [19], and use an auxiliary ABCD measurement of $\cal {E}_{\textrm {THz}}$, as shown in Fig. 3(e), with the corresponding spectral intensity and phase in (f), which remain above the noise floor out to ${\sim }5{\,{\mathrm {THz}}}$. Here we employ the extreme frequency range of the ABCD spectral phase for matching to that extracted from the FW-DFG spectrogram (as shown in Fig. 3(f)), to minimize the impact of spectral convolution for the latter, where satisfactory phase overlap is achieved.

 figure: Fig. 3.

Fig. 3. Measured spectrograms with (a) FW-DFG alone ($S_T(\nu,\tau )$), and (b) both FW-DFG and EFISH $S_{TE}(\nu,\tau )$ (the latter with uni-polar bias $\cal {E}_{\textrm {b}}$), focusing on low-frequency range (see Fig. 1(d) for full spectral range for reference pulse). (c) Corresponding frequency marginals and that from (d) interferometric residual $\Delta S=S_{TE}-S_T-S_E$. (e) Low-frequency ABCD time-domain field for the same pulses and (f) corresponding spectral intensity and phase. Also included in (f) is the phase obtained from integrating the group delay ${\bar \tau }(\nu )$ extracted from the spectrogram (see Sec. 5.1) after adjusting to the absolute ABCD phase in the interval shown.

Download Full Size | PDF

Having fixed the absolute phase of the THz-IR pulses, it only remains to calibrate the absolute field strength. As detailed in Appendix 5.1, if one simply considers only the on-axis fields on the focal plane (e.g. assumed in [27]), one arrives at the intuitive normalization condition for the recovered field (Eq. (8)):

$$ F_T=\int \mathrm{d} t \mathcal{E}^2(t)=\frac{J_T}{P_E} \mathcal{E}_{\mathrm{b}}^2 $$
where $J_T$ is obtained by integrating the experimental FW-DFG spectrogram $S_T(\nu,\tau )$ over $(\nu,\tau )$ and $P_E$ by integrating the EFISH spectrum $S_E(\nu )$ over $\nu$, and accounting correctly for the absence the FW-SFG spectrum in $S_T$, which, neglecting propagation effects in the derivation of Eq. (1), is a mirror image of the FW-DFG spectrum. Note that the arbitrary, but common scaling of $S_T$ and $S_E$ cancels in Eq. (1). The resultant temporal field ${\mathcal {E}}(t)$ (with a peak field of ${\mathcal {E}}_0 \sim 2.1{\,\mathrm {MVcm}^{-1}}$) is presented later (Fig. 5(d)) – however, as discussed in the Introduction, one must account for the spatial propagation in order to obtain a reliable quantitative estimate for the field, applying the normalization in Eq. (1) with the addition of correction factors for both the FW-DFG and EFISH signals.

3.3 Accounting for beam profile and propagation effects

The role of spatial effects for the co-propagating THz and optical fields on the resultant TFISH and EFISH signals has been considered in detail in previous reports, e.g. [2426,33] for conventional THz bandwidths. Notably, the effect of the Gouy phase on the axial evolution of the nonlinear signal source ${P^{(3)}}(\nu )\propto a(\nu -\nu '){\mathcal {E}}(\nu ')$ ($a=E_1$) and FISH signal $E_2(\nu )$ leads to a suppression of the FW-SFG signal ($\nu >0$) and saturation of the FW-DFG signal ($\nu <0$). While the Gouy-phase effect is not explicitly considered in many studies with ABCD detection (as it is assumed that the relative time-domain signal is not affected), one sees this effect immediately when spectrally resolving the TFISH signal. We build here on those previous treatments, extended to apply to the present case with an ultra-broadband THz-IR pulse (where quasi-DC TFISH treatments no longer hold). However, the propagation effects are in general more complex than the basic result (integrating the axial fields along $z$ with the Gouy phase variation [33]) due to (i) a deviation of the THz-IR beam from a Gaussian spatial form and its frequency dependence [29], and (ii) effects of dispersive phase mismatch, which were considered when assessing different gas media and pressures [24,33] and were also more recently shown to be significant for low-frequency TFISH detection [40].

The theoretical basis is covered in detail in Appendix 5.2, which shows that the “raw” THz-IR field determined from the spectrogram and applying the EFISH calibration (Eq. (1)) must still be corrected for both the EFISH electrode geometry [25,33] (Eq. (18)) and the evolution of the on-axis optical and THz-IR fields throughout the focal region [26], via the relative responsivity factor $H(\nu )$ (Eq. (20)). The latter would appear to necessitate measurements of the THz-IR beam profile throughout the focal region [41], although to perform this with spectral resolution over the full THz-IR bandwidth would indeed be a demanding task.

However, here we adopt a novel, alternate approach, by measuring the FW-DFG spectra $S_T(\nu,\tau =0, R)$ with a variable aperture (radius $R$) in the collimated beam before focusing (see scheme in Fig. 4). As derived in Appendix 5.2 (Eq. (23)), under the assumption of a radially symmetric beam intensity and linear polarization (reasonable assumptions for our THz-IR source [29]), one can deduce the responsivity correction $P(\nu )$ for the FW-DFG signal due to propagation. This highly useful result is facilitated by the Fourier optics relation between the collimated beam field ${\mathcal {E}}_C(r,\nu )$ (a distance $f$ before the focusing OAPM whose focal length is also $f$), the focal-plane field ${\mathcal {E}}(r,z=0,\nu )$, and the axial field ${\mathcal {E}}(r=0,z,\nu )$ in the relevant vicinity of the focal plane. That such an analysis allows one to connect the FW-DFG signal to the collimated beam field in closed-form is not immediately obvious, given that the variable, radial beam truncation simultaneously changes the numerical aperture and circular portion of the transmitted THz-IR beam being focused, but one can show that these two effects essentially factor into two terms in the integral transformation (Eq. (23)).

 figure: Fig. 4.

Fig. 4. Beam profile results from measurements with variable aperture in collimated THz-IR beam path. (a) Experimental and (b) fitted FW-DFG spectra $S_T(\nu,\tau =0,R)$ vs. iris radius $R$, the latter based on the collimated radial beam profile $I_C(r,\nu )$ in (c) (plotted here normalized for each frequency) – see Eq. (2) for model function for ${\mathcal {E}}_C(\nu,R)$ and Appendix 5.2 (Eqs. (23) and (24)) for the integral transformation between ${\mathcal {E}}_C(\nu,R)$ and $S_T(\nu,R)$. Schematic of experimental geometry shown left (${\mathcal {E}}(z,\nu )$ denotes the axial THz-IR field about the focal point, $b_T(z,\nu )$ the FISH signal field, and ${\mathcal {E}}_C(r,\nu )$ the field of the collimated THz/IR beam).

Download Full Size | PDF

The experimental results are shown in Fig. 4(a). While the approach described in Appendix 5.2 can in principle be applied to an arbitrary beam profile ${\mathcal {E}}_C(r,\nu )$, we find the data can be modeled well assuming a conical beam

$${\mathcal{E}}_C(r,\nu) ={\mathcal{E}}_{C0}(\nu) \exp\left( -\frac{r^2}{w_C^2(\nu)}\right) \left[ 1-\exp\left( -\frac{r^2}{h^2(\nu)}\right) \right],$$
(i.e. a Gaussian beam with radius $w_C$ modified by a central Gaussian-shaped depression with width $h$) which yields the $R$-dependence given in Eq. (24) for the relative FW-DFG responsivity $P(R)$. The optical detection beam focal parameter $z_1=1.2$ mm (also required for the correction factors, Eq. (12)) is calculated from the rms radius of the measured intensity profile of the collimated beam ($w_{1C} \rightarrow \sqrt {2\langle r^2 \rangle }=1.8$ mm). The results of fitting the experimental data with this model for each frequency $\nu$ are shown in Fig. 4(b), and are seen to be in reasonable agreement. The corresponding collimated THz-IR radial intensity profile $I_C(r,\nu )=|{\mathcal {E}}_C(r,\nu )|^2$ determined from the fitting procedure for each frequency is shown in Fig. 4(c) (normalized to the peak intensity for each frequency). While one must exercise some caution using such an indirect approach to determine the THz-IR beam profile, the form and frequency-dependence of the beam shape are reasonably consistent with our previous measurement [29] (which used ${\chi ^{(2)}}$-SF-detection in an EO crystal, where the FISH propagation effects are absent). Note that the collimated THz-IR beam shape is affected by both the inherent conical plasma emission [42] as well as the axial stops used in the setup (Fig. 1(c)). In any case, the propagation correction factors $P(\nu )$ are relatively robust to the parameters respectively the beam fitting analysis.

Besides the beam-shape propagation effects treated above (which include the influence of the Gouy phase), as apparent in Eqs. (18,19,23), the dispersive wavevector mismatch ${\Delta k}$ (Eq. (15)) in air also adds an additional factor to the FISH responsivity functions [25,33,40]. While often this is neglected, due to the notion that at ambient pressure the weak dispersion of air leads to NIR-UV phase walk-off lengths on the few-cm scale (compared to mm-scale Rayleigh ranges), we will see that at least for the THz range, it is not negligible for our THz-IR beam focal parameters. That such effects must be considered was again emphasized in a recent report with a weakly focused THz beam, which showed severe effects due to ${\Delta k}$, even well below ambient pressure [40]. The calculated values ${\Delta k}$ (for $\lambda _0=775$ nm here) are shown in Fig. 5(a) for ambient air, and both dry and humid air (using the full set of THz-IR absorption bands for H$_2$O and CO$_2$ [43] and electronic response of N$_2$, O$_2$, Ar and H$_2$O from [44]). One sees that the mismatch is dominated by the residual electronic dispersion, and is asymmetric between FW-DFG and FW-SFG. The latter is as expected, as the DFG process involves wavelengths on the red side of $\lambda _2=\lambda _0/2$ which are closer to $\lambda _0$ such that ${\Delta k}$ is progressively smaller for increasing $\nu '=-\nu$ (one notes that this is a fortuitous situation for FISH detection, given that only FW-DFG survives the Gouy-phase effects). To consider the magnitudes more readily, the right scale in Fig. 5(a) shows the corresponding coherence length $L_c=1/{\Delta k}$ (note this is defined differently than the typical expression $L_c=\pi /{\Delta k}$ for plane-wave nonlinear generation, as its combination here with the beam $z$-dependence yields an exponential (and not sinc-function) dependence, see e.g. Eq. (17)). One sees that $L_c \approx 7$ mm for THz frequencies, growing to $L_c \approx 13$ mm at $\nu '=100{\,{\mathrm {THz}}}$. We can compare these values with the predicted axial-field $z$-dependence obtained from the THz-IR beam analysis above, as shown in Fig. 5(b). Here one sees that especially in the range below $10{\,{\mathrm {THz}}}$, the tails of the axial fields do persist on the many-mm scale and one should evaluate phase-matching factors to inspect how significantly they affect the FW-DFG signal evolution. Note that the non-monotonic development of the curves in Fig. 5(b) around 10 THz is linked to the characteristic shape of $I_C(r,\nu )$ shown in Fig. 4(c), where the intensity of the collimated radiation is found to be concentrated closer to the beam axis than at lower and higher frequencies. We also note that axial fields actually decay here more rapidly with $|z|$ than for a Gaussian beam due to the conical beam shape, which actually reduces the impact of phase mismatch.

 figure: Fig. 5.

Fig. 5. (a) Calculated phase mismatch ${\Delta k}$ vs. frequency, due to dispersion of air (for two values of relative humidity ${c_{\mathrm {H_2O}}}$); corresponding $1/e$-coherence length scale at right. (b) Normalized on-axis field amplitude vs distance about focal plane based on the fitted THz-IR beam parameters from Fig. 4. (c) Corresponding field responsivity factor $H(\nu ')$ for FW-DFG (blue solid curve), relative to EFISH for geometry used here. Dashed curves: responsivity factors due to dispersive phase mismatch alone. (d) Recovered temporal electric field in focal plane: raw result using simple focal-plane treatment (dashed) and after correcting for relative FW-DFG responsivity in (c) due to propagation effects.

Download Full Size | PDF

In Fig. 5(c), we plot the total FW-DFG field responsivity spectrum $H(\nu ')$ (including both beam propagation and mismatch), relative to the EFISH responsivity (from Eq. (18), plotted as a horizontal line). As can be seen, this exceeds unity, reflecting the fact that the effective interaction length for FW-DFG is larger than for EFISH with our experimental beam parameters. Also plotted are the relative contributions due to phase mismatch alone (dashed curves), calculated by taking the ratio of $H(\nu )$ with ${\Delta k}\neq 0$ and ${\Delta k}=0$, respectively. One sees here that ${\Delta k}$ does significantly degrade the responsivity below ${\sim }10{\,{\mathrm {THz}}}$, falling to ${\sim }0.5$ for $\nu '\rightarrow 0$ (while for our electrode geometry and optical focusing, ${\Delta k}$ only reduces the relative EFISH field responsivity by a small factor 0.05).

We can now apply this relative responsivity function to the THz-IR spectral field determined from the spectrogram/ABCD analysis in the last section. The corresponding temporal electric fields ${\mathcal {E}}(t)$ (from both the focal-plane calibration and after responsivity correction) are shown in Fig. 5(d), where one sees an essentially single-cycle pulse (blue curve) with a small residual positive chirp. While the duration of the main half-cycle is $9.5{\,{\mathrm {fs}}}$ (reflecting the mean frequency of ${\sim }50{\,{\mathrm {THz}}}$), a more objective measure of the pulse duration is the FWHM of the field envelope ($\hat {{\mathcal {E}}}(t)$, also included in the figure) which is $17.0{\,{\mathrm {fs}}}$, while the corresponding intensity FWHM based on $I(t)=|\hat {{\mathcal {E}}}(t)|^2$ is $10.1{\,{\mathrm {fs}}}$. Importantly, as expected from the responsivity spectrum, the corrected peak field is ${\mathcal {E}}_0 = 1.2{\,\mathrm {MVcm}^{-1}}$, compared to $2.1{\,\mathrm {MVcm}^{-1}}$ from the simple focal-plane calibration. Hence for our present geometry, one would significantly overestimate the field using the simplified approach. Note that ${\mathcal {E}}_0$ is not corrected for the loss/dispersion of the thin Si wafer in the beam (field transmission $0.7$), such that higher fields are potentially available at the sample location. Moreover, from the THz-IR beam results in Fig. 4(c), we see that the numerical aperture (and hence ${\mathcal {E}}_0$) could be increased significantly, by (i) increasing the collimated beam diameter (which is well below 10 mm across most of the THz-IR spectrum of Fig. 4(c)) using a longer focal length for the initial collimation-OAPM after the plasma, and (ii) using a tighter re-focusing than used here ($f=76.2$ mm). Hence, for this THz-IR source, peak fields approaching 10${\,\mathrm {MVcm}^{-1}}$ could be achieved.

To further elucidate the above results, it is instructive to inspect how the propagation effects are predicted to develop vs. distance through the focal region. In Fig. 6(a), we plot the FW-DFG/SFG intensity spectrum for a set of $z$-positions, using the experimental spectrum and beam parameters from above, i.e. based on Eq. (14), where we also include the results for ${\Delta k}=0$. One sees that for ${\Delta k}=0$ (dashed curves), while FW-DFG and SFG initially grow symmetrically before reaching the focal plane ($z<0$), for the FW-SFG process the destructive Gouy-phase effects already set in upon reaching the focal plane ($z=0$), with the signal vanishing with increasing $z>0$. The phase mismatch for FW-SFG exacerbates these effects, explaining why the FW-SFG signal is so effectively suppressed experimentally, even if one considers a loss of focal symmetry in practice, e.g. due to non-ideal beam shape/wavefront and collinearity. For the FW-DFG process, one sees again that the phase mismatch is only significant at lower DF frequencies $\nu '=-\nu$, which further emphasizes how the situation here differs from conventional plane-wave THz phase-matching (where the THz dispersion plays the dominant role in ${\Delta k}$ and typically increases for higher frequencies due to the THz-wavelength scaling, ${\Delta k}\propto \nu$, and typical THz dispersion).

 figure: Fig. 6.

Fig. 6. (a) Simulated evolution of the FW-DFG/SFG signal intensity spectrum for selected positions $z$ relative to focal plane. Results are based on the experimental THz-IR spectrum and beam parameters determined above, and include both results with (solid curves, ${\Delta k}\neq 0$) and without (dotted curves, ${\Delta k}=0$) dispersive phase mismatch. (b,c) Focal spot diameter (top), Rayleigh range (middle) and relative field responsivity (bottom, as per Fig. 5(c)) for two experimental scenarios with a Gaussian THz-IR beam and frequency-independent collimated beam diameter $2w_C$: (b) $2w_C=25$ mm, $f=76.2$ mm, $f_1=125$ mm; (c) $2w_C=15$ mm, $f=152.4$ mm, $f_1=250$ mm.

Download Full Size | PDF

Finally, we consider some auxiliary model predictions for the propagation effects for other beam/focal scenarios in Fig. 6(b,c). For simplicity, we assume a Gaussian (not conical) collimated/focal beam profile, with a frequency-independent collimated beam diameter $2w_C$ (as one would obtain e.g. with a large-area THz emitter [45]). In each case, we plot the focal spot diameters $2w_{0,1}$ for the THz-IR and optical beams, respectively, the corresponding Rayleigh ranges $z_{0,1}$ and resultant field responsivity (as per Fig. 5(c), normalized to the EFISH signal, and including the relative contribution due to dispersive mismatch alone). We assume the same geometry for the EFISH electrodes as above. The first scenario in Fig. 6(b) employs the same focal lengths $f,f_1$ and input optical beam diameter as the experiments presented above, but with a larger THz-IR beam with $2w_C=25$ mm. One sees that while the FW-DFG responsivity below 10 THz rises above unity, for higher frequencies it falls significantly below this. One can see why this is the case by inspecting the Rayleigh range $z_0(\nu ')$, which becomes increasingly shorter for larger $\nu '$, such that the effective interaction length becomes much less than the EFISH electrode length. Clearly, in such a scenario, the focal-plane calibration would seriously underestimate the IR spectral range. For our experimental results above, such a trend is countered by the fact that the collimated beam diameter decreases with increasing frequency (Fig. 4(c)), due to the physical emission mechanisms in the plasma [42]. In Fig. 6(c), we inspect a scenario with much weaker focusing, in order to be close to the experimental conditions in [40], i.e. both $f,f_1$ a factor 2 higher and a moderate collimated THz beam size $2w_C=15$ mm. We also adopt the larger input optical beam used in [40] with a 10-mm diameter, such that $z_1$ is actually slightly smaller than the last scenario. We see that the THz responsivity is now below unity, while the IR responsivity is above for the whole range shown. However, our main interest in this scenario is due to the result in [40] that dispersive mismatch seriously degrades the TFISH conversion efficiency (for a low-bandwidth LiNbO$_3$ THz source). As seen, the responsivity in this range (<3 THz) falls to very low values, due to the long Rayleigh range of the THz beam whereby the phase mismatch effects become highly destructive (which was studied and circumvented in [40] by terminating the THz co-propagation with an iris in the focal region). As demonstrated above, the dispersive mismatch effects are much less significant for our THz-IR source and focal parameters (Fig. 5(c)), although clearly one should bear dispersive mismatch effects in mind when designing the focal geometry.

4. Conclusion

We have demonstrated an approach to reconstruct the calibrated, temporal electric field of ultra-broadband THz-IR pulses using FISH in ambient air. Here we have used longer, transform-limited optical detection pulses than in previous reports [19] and extracted the intensity spectrum and relative phase directly from the FW-DFG spectrogram (without deconvolution), which has some practical advantages in certain contexts, e.g. allowing simple single-shot detection of the intensity spectrum via $S_T(\nu ',\tau =0)$ with the same configuration, albeit with a reduced spectral resolution (although in many cases this resolution is sufficient). While not employed here, deconvolution of the full spectrograms would still be possible to recover superior resolution [28]. Extending on the THz-range studies of propagation effects for the relative FW-DFG and EFISH signals, we have shown that one must account for these (in our case, to avoid overestimating the field strength), and we have presented a formalism and technique for estimating these with auxiliary measurements with variable radial truncation of the collimated beam. This technique could also be adapted to account for propagation effects in ABCD measurements on conventional THz sources, in order to correctly calibrate the peak temporal fields.

The approach here again emphasizes how FISH detection in ambient gases is admittedly prone to an “uncontrolled” effective interaction length, and in this sense detection in thin sensor materials (such as EO time-domain field sampling with careful calibration of the EO response [21]) allows a more direct determination of the focal electric field. However, as discussed in the Introduction, for such high THz-IR bandwidths this requires very short optical detection pulses, which may not be available. Nevertheless, it is evident that the field calibration would be simplified if the ${\chi ^{(3)}}$-interaction would be confined to a thin sensor medium. One could consider a narrow gas jet [24], although as shown here, any vacuum cell windows would have to be chosen carefully not to cause severe dispersion/absorption effects. One very interesting approach would be to use FISH in solid-state ${\chi ^{(3)}}$-media [46,47], although again, even for very thin media, dispersion/absorption effects will need to be addressed. Hence, the use of a truly surface-based ${\chi ^{(3)}}$-effect (integrating an optional bias field for the EFISH signal, and using tight focusing) may have the most potential in future approaches.

5. Appendices

While there are several theoretical treatments in the literature for the nonlinear signals generated by EFISH, ABCD and FW-SFG/DFG, we present a relatively self-contained development here, in particular to define notation for the novel results (e.g. mapping of collimated beam to the axial integral for the FW-DFG signal through the focus), as well as to avoid quasi-DC approximations which are commonly used for EFISH/TFISH with low-frequency THz pulses (based on the assumption that the THz bandwidth is much smaller than the optical bandwidth, which is clearly untenable in the present case). Moreover, we employ the carrier phasor convention $e^{i(\omega t - k z)}$ throughout, which leads to opposite signs in various expressions for the beam propagation and Gouy phases.

5.1. Ideal spectrogram signal and spectral intensity/chirp estimation

Substituting the real fundamental optical field $\mathcal {E}_{1j}(t)=\tfrac {1}{2}A_j(t)e^{i\omega _0 t}+\mathrm {c.c.}$ and THz-IR field ${\mathcal {E}}_k(t)$ ($j,k=x,y$) into the third-order nonlinear polarisation (assumed instantaneous) yields [48] $\mathcal{P}^{(3)}_i=\tfrac {1}{2} {P^{(3)}_i} e^{2i\omega _0 t}+\mathrm {c.c.}$ with

$${P}^{(3)}_i=\tfrac{1}{2}\varepsilon_0{\chi^{(3)}_{ijkl}} (A_j A_k{\mathcal{E}}_l + A_j {\mathcal{E}}_k A_k + {\mathcal{E}}_j A_k A_l),$$
where we retain only the terms about the second harmonic $2\omega _0$. Taking the optical field polarized along the $x$-axis ($A_j=A\cdot \delta _{jx}$) and invoking the symmetry properties of ${\chi ^{(3)}}$ in isotropic media [48], one has ${P}^{(3)}_x=\tfrac {3}{2}{\chi ^{(3)}}\;A^2{\mathcal {E}}_x$ and ${P^{(3)}_{y}}=\tfrac {1}{2}{\chi ^{(3)}}\;A^2{\mathcal {E}}_y$ where ${\chi ^{(3)}}\;\equiv {\chi ^{(3)}_{iiii}}$. Hence the FISH intensity generated for the THz-IR (or DC) field polarization orthogonal to the optical one is $3^2$ weaker than for the parallel case, and we neglect the former in the following, with all scalar fields now referring to the $x$-components. Defining $a(t)=A^2(t)$ and $\gamma =\tfrac {3}{2}\varepsilon _0{\chi ^{(3)}}$ we have
$${P^{(3)}}(\Omega)=\mathcal{F}\{\gamma a(t){\mathcal{E}}(t)\}=\frac{\gamma}{2\pi}\!\!\int\!\!{\mathrm{d}\Omega'\,a(\Omega-\Omega'){\mathcal{E}}(\Omega')},$$
(where integration over $(-\infty,\infty )$ is assumed unless otherwise specified), which is simply a restatement of ${P^{(3)}}(t)=\gamma a(t){\mathcal {E}}(t)$ in the frequency domain, but will be required below when propagation is taken into account. In terms of the absolute FISH signal frequency $\omega _s$, $\Omega =\omega _s-2\omega _0$, i.e. $\Omega >0$ corresponds to FW-SFG and $\Omega <0$ to FW-DFG.

Defining $\mathcal {E}_2=\tfrac {1}{2}Be^{2i\omega _0 t}+\mathrm {c.c.}$, the wave equation for $B(\mathbf {r},\Omega )$ (applying the slowly-varying amplitude approximation) is [24,25,48]:

$$~-2ik_2\partial_zB ={-}\nabla^2_\perp B - \mu_0\omega_2^2 {P^{(3)}}e^{ik_2z}$$
with $k_2(\Omega )=k(\omega _2 + \Omega )$ and $\omega _2=2\omega _0$. Neglecting the spatial dependence and propagation (treated below), Eq. (4) reduces to
$$\partial_zB ={-}i\frac{\mu_0\omega_2}{2k_2} {P^{(3)}}\; \qquad \rightarrow \qquad B(\Omega) ={-}i\frac{\mu_0\omega_2}{2k_2} {P^{(3)}}(\Omega)\cdot\delta z.$$
for a short propagation distance $\delta z$.

For the FW-SFG/DFG spectrogram field (with $a$ shifted by a variable time $\tau$, or equivalently ${\mathcal {E}}$ shifted by $-\tau$) we have ${\mathcal {E}}(\Omega )\rightarrow {\mathcal {E}}(\Omega )e^{i\Omega \tau }$

$$B_T(\Omega,\tau)= \frac{\beta}{2\pi}\delta z \int{ \mathrm{d}\Omega'\, e^{i\Omega' \tau}a(\Omega-\Omega'){\mathcal{E}}(\Omega')}, \qquad \beta={-}i\frac{3{\chi^{(3)}}\;\omega_2^2}{4k_2 c^2}$$
whereas for the EFISH field, with ${\mathcal {E}}(\Omega )\rightarrow \mathcal {E}_{\textrm {b}}\cdot 2\pi \delta (\Omega)$ yields $B_E(\Omega )= \beta \delta z a(\Omega ) \cal {E}_{\textrm {b}}$.

In order to use $B_E$ to calibrate the THz-IR field, we can calculate the spectrogram “energy”, which is given by:

$$J_T = \frac{1}{2\pi}\!\!\int\!\!\mathrm{d}\tau\,\!\!\int\!\! \mathrm{d}\Omega\, S_T(\Omega,\tau) =\frac{1}{2\pi}\!\!\int\!\!\mathrm{d}\tau\,\!\!\int\!\! \mathrm{d}\Omega\, |B(\Omega,\tau)|^2 = \gamma^2 F_a \frac{1}{2\pi}\!\!\int\!\! \mathrm{d}\Omega\, |{\mathcal{E}}(\Omega)|^2,$$
where $F_a=\tfrac {1}{2\pi }\smallint \mathrm {d}\Omega \, |a(\Omega )|^2$. The corresponding EFISH “power” is given by $P_E=\tfrac {1}{2\pi }\smallint \mathrm {d}\Omega \,|B_E(\Omega )|^2=\gamma ^2 F_a \mathcal {E}_{\textrm {b}}^2$. Hence the terms $\gamma ^2 F_a$ cancel in the ratio $J_T/P_E$ and the correctly scaled THz-IR field must satisfy:
$$F_T = \frac{1}{2\pi}\!\!\int\!\! \mathrm{d}\Omega\, |{\mathcal{E}}(\Omega)|^2 =\!\!\int\!\! \mathrm{d}t\, {\mathcal{E}}^2(t) = \frac{J_T}{P_E}\mathcal{E}_{\textrm{b}}^2,$$
which can be used to find the correct field normalization for ${\mathcal {E}}$, after calculating $J_T$ and $P_E$ from the experimental data (with arbitrary, but common scaling).

For the moments extracted from the FW-DFG spectrogram, we have for the spectral marginal:

$$M(\Omega) = \!\!\int\!\! \mathrm{d}\tau\, S_T(\Omega,\tau) = \!\!\int\!\! \mathrm{d}\tau\, |B_T(\Omega,\tau)|^2 = \frac{\beta^2}{2\pi} \!\!\int\!\! \mathrm{d}\Omega'\, |a(\Omega-\Omega')|^2 |{\mathcal{E}}(\Omega')|^2,$$
where the final result is gained by substituting Eq. (6) for $B_T$ and resolving the resultant triple integral. Hence $M(\Omega )$ yields the spectral intensity $I(\Omega )=|{\mathcal {E}}(\Omega )|^2$ convolved with the spectral response function $G(\Omega )=|a(\Omega )|^2$ (the latter being relatively narrow-band for our optical pulses). For the temporal first moment,
$${\bar \tau}(\Omega) = \frac{1}{M(\Omega)}\!\!\int\!\! \mathrm{d}\Omega'\, \tau \cdot S(\Omega,\tau) \rightarrow -\frac{2\pi i}{M(\Omega)}\!\!\int\!\! \mathrm{d}\Omega'\, {\mathcal{E}}(\Omega') \partial_\Omega'{\mathcal{E}}^*(\Omega')|a(\Omega-\Omega')|^2$$
$$\rightarrow \frac{2\pi}{M(\Omega)}\!\!\int\!\!\mathrm{d}\Omega'\,|{\mathcal{E}}(\Omega')|^2 |a(\Omega-\Omega')|^2 {T_{\mathrm{g}}}(\Omega'),$$
where we invoke that ${\mathcal {E}}$ is smoothly varying over the bandwidth of $G=|a|^2$ and used the result ${\mathcal {E}}\partial {\mathcal {E}}^* = |{\mathcal {E}}|\partial {\mathcal {E}}+ i|{\mathcal {E}}|^2 {T_{\mathrm {g}}}$. Observing Eq. (9) for $M(\Omega )$, we see that ${\bar \tau }$ yields the intensity-weighted local group delay $\langle {T_{\mathrm {g}}}\rangle$ convolved with the gate $G=|a|^2$.

5.2. FISH signals accounting for spatial propagation

The expressions for the broadband FISH signals begin with Eq. (4) above, which are solved analogously to [2426]. We assume a Gaussian beam for the input optical detection field,

$$A(\mathbf{r},\Omega)=A_0(\Omega)Q_1(z)R_1(r,z), \quad Q_1(z)=\frac{1}{1-iz/z_1}, \quad R_1(r,z)=\exp\left(-\frac{Q_1(z)r^2}{w_1^2}\right)$$
where $z_1=k_1w_1^2/2$, and $w_1$ is the $1/e^2$-intensity beam radius at the focal plane. Assuming the THz-IR beam is sufficiently larger than the optical beams, from Eq. (3) we then have
$${P^{(3)}}(\mathbf{r},\Omega)=\frac{\gamma}{2\pi}Q_1^2 R_1^2(r,z) \!\!\int\!\! \mathrm{d}\Omega'\,a_0(\Omega-\Omega'){\mathcal{E}}(\omega')$$
where $a_0(t)=A_0^2(t)$. Note that $\Omega\;>\;0$ corresponds to FW-SFG and $\Omega \;<\;0$ to FW-DFG (for the real temporal field ${\mathcal {E}}(-\Omega )={\mathcal {E}}^*(\Omega )$). Then the ansatz $B(\mathbf {r},\Omega )=b(\Omega,z)Q_2(z)R_2(r,z)$ is a solution to Eq. (4) where $w_2=w_1/\sqrt {2}$ ($z_2=z_1$), $Q_2=Q_1$, $R_2=R_1^2$, and with Eq. (13) the on-axis SH field is governed by:
$$\partial_z b(z,\Omega)={-}i\frac{\beta}{2\pi} Q_1(z)\!\!\int\!\! \mathrm{d}\Omega'\, a_0(\Omega-\Omega'){\mathcal{E}}(z,\Omega')e^{i{\Delta k}(\Omega,\Omega')z}$$
where we factored out the propagation phase shifts (but not Gouy phase), i.e. ${\mathcal {E}}(z,\Omega )\rightarrow {\mathcal {E}}(z,\Omega )e^{-ik(\Omega )z}$ and $a_0(z,\Omega )\rightarrow a_0(z,\Omega )e^{-2ik_{10}z}e^{-i\Omega k'_{10}z}$ (the latter valid due to the narrow-band nature of $a_0$), where the dispersive mismatch is given by
$${\Delta k}(\Omega,\Omega')=k_2(\Omega)-k(\Omega')-[2k_{10}+(\Omega-\Omega')k'_{10}]$$
with $k_{10}=k_1(\Omega =0)$, $k'_{10}=\partial _\Omega k_1(\Omega =0)=1/{v_{\mathrm {g1}}}$. The solution of Eq. (14) for a Gaussian THz-IR beam, with
$${\mathcal{E}}(z,\Omega)={\mathcal{E}}_0(\Omega)Q_0(z,\Omega), \qquad Q_0(z,\Omega)= \frac{1}{1-iz/z_0(\Omega)}$$
has been expressed in the literature [24,25]. Note that here with $z_0(\Omega )=k_0(\Omega )w_0^2/2$, $\mathrm {sgn}(z_0)=\mathrm {sgn}(\Omega )$, i.e. $k_0$ and $z_0$ are algebraically negative for FW-DFG. As shown in Fig. 5(a), ${\Delta k}\;>\;0$ for both FW-SFG and DFG, which yields
$$b_T(z\rightarrow\infty,\Omega)={-}i\beta\!\!\int\!\! \mathrm{d}\Omega'\, a_0(\Omega-\Omega'){\mathcal{E}}_0(\Omega')\frac{z_1\cdot z_0(\Omega')}{z_1 -z_0(\Omega')}e^{-{\Delta k}(\Omega,\Omega')z_1}\Theta(-\Omega')$$
i.e. the output FISH signal vanishes for FW-SFG (the Heaviside function $\Theta (-\Omega )=0$), and the response for FW-DFG is affected by the relative Rayleigh ranges of the input optical ($z_1$) and THz-IR ($z_0(\Omega <0)<0$) beams and degraded by any dispersive phase mismatch ${\Delta k}$ (note the exponential form of the latter, as opposed to the sinc-function dependence in other plane-wave nonlinear conversion processes).

For the EFISH signal, we can use Eq. (14), with ${\mathcal {E}}(z,\Omega )\rightarrow 2\pi \mathcal {E}_{\textrm {b}} \Pi (2z/L_b)\cdot \delta (\Omega )$, where we assume a parallel-plate electrode geometry here (as opposed to the more involved field distribution of thin wires [24]), i.e. $\Pi (2z/L_b)$ is the unit rectangular function with width $L_b$ about $z=0$. Substitution yields for the output field

$$b_E(\Omega) ={-}i\beta Z_E a_0(\Omega) \cal{E}_{\textrm{b}},$$
where $Z_E\rightarrow Z_{E0} = 2 z_1 \tan ^{-1}\left (\tfrac {1}{2}L_b/z_1\right )$ for ${\Delta k}_0 = {\Delta k}(\Omega,0) = 0$, and the less transparent expression $Z_E = i z_1 e^{+\Delta k_0 z_1}[{\mathrm {E}_1}(\xi )-{\mathrm {E}_1}(\xi ^*)]$ for $\Delta k_0 \neq 0$, where ${\mathrm {E}_1}$ is the exponential integral function and $\xi = \Delta k_0(z_1+i\tfrac {1}{2}L_b)$ (note that the factor $e^{+\Delta k_0 z_1}$ is compensated by a term in the ${\mathrm {E}_1}$ functions which decays more rapidly with $\Delta k_0$, yielding $Z_E \;<\; Z_{E0}$ as expected).

As presented in the main text, our THz-IR beam does not correspond to a Gaussian beam as per Eqs. (16) and (17). One can write the more general solution to Eq. (14) as:

$$b_T(z\rightarrow \infty,\Omega)={-}i\frac{\beta}{2\pi} \!\!\int\!\! \mathrm{d}\Omega'\, a_0(\Omega-\Omega') \!\!\int_{-\infty}^{\infty}\!\! \mathrm{d}z\, Q_1(z)e^{i\Delta k(\Omega')z} {\mathcal{E}}(z,\Omega')$$
where, as per the solution Eq. (17), the $z$-integrand decays sufficiently rapidly for large $|z|$ for our focal conditions to extend the finite focal axis range to $z=(-\infty,\infty )$. Note also that we invoked the narrow-band approximation for $a(\Omega )$ to take $\Omega \rightarrow \Omega '$ for $\Delta k$.

We are now in the position to define a relative FISH field responsivity (i.e. field correction factor), $H(\Omega )$ between the FISH signals with propagation effects, and the focal-plane expressions in Sec. 5.1, i.e. using Eqs. (6), (19) and (18):

$$\frac{b_T(\Omega)}{b_E}= H(\Omega) \frac{B_T(\Omega)}{B_E}, \qquad H(\Omega) = \frac{1}{Z_E} \frac{ \!\!\int_{-\infty}^{\infty}\!\! \mathrm{d}z\, Q_1(z)e^{i\Delta k(\Omega)z} {\mathcal{E}}(z,\Omega) }{{\mathcal{E}}(0,\Omega)}$$
where we invoked the narrow-band limit $a(\Omega )\rightarrow 2\pi a_0 \delta (\Omega )$ for the input optical field and consider $\tau =0$.

While one can proceed to substitute different beam models for ${\mathcal {E}}(z,\Omega )$ into Eq. (19) or (20) [26], we derive a transformation here to rewrite this in terms of the collimated beam ${\mathcal {E}}_C$ before the focusing mirror (focal length $f$). If this is referenced to the plane a distance $f$ before the mirror, we can apply the Fourier optics relation between ${\mathcal {E}}_C(X,Y)$ and the focal plane field ${\mathcal {E}}(x,y,0)$ [49]. For a given THz-IR frequency component (with wavevector $k_0$), one can derive

$$\tilde{\mathcal{E}}(u,v,0)=i\frac{2\pi f}{k_0} {\mathcal{E}}_C(\frac{f}{k_0}u,\frac{f}{k_0}v)$$
where $\tilde {\mathcal {E}}$ is the 2D spatial Fourier spectrum of ${\mathcal {E}}$. Helmholtz propagation about the focal plane leads to $\tilde {\mathcal {E}}(u,v,z)=\tilde {\mathcal {E}}(u,v,0)\exp [{\tfrac {i}{2k_0}(u^2+v^2)z}]$, and the on-axis field can be expressed as ${\mathcal {E}}(0,0,z)=\tfrac {1}{(2\pi )^2}\smallint \!\smallint \mathrm {d}u\,\mathrm {d}v\,\tilde {\mathcal {E}}(u,v,z)$.

We now invoke the assumption of radial symmetry for the collimated/focal beams, $\tilde {\mathcal {E}}(u,v,z)\rightarrow \tilde {\mathcal {E}}(\rho,z)$ ($\rho =\sqrt {u^2+v^2}$), such that ${\mathcal {E}}(r=0,z)=\tfrac {1}{2\pi }\smallint \mathrm {d}\rho \, \rho \tilde {\mathcal {E}}(\rho )$. Substituting $r=\tfrac {f}{k_0}\rho$ yields

$${\mathcal{E}}(0,z)=i\frac{k_0}{f}\!\!\int_{0}^{R}\!\!\mathrm{d}r\,r \exp\left( i\frac{k_0}{2f^2}r^2\,z \right) {\mathcal{E}}_C(r).$$

While the result for the full collimated beam corresponds to $R\rightarrow \infty$, Eq. (22) also holds for finite $R$ corresponding to radial truncation of the beam with an iris aperture. One can substitute Eq. (22) into Eq. (19) to obtain the output FISH field:

$$\begin{aligned}b_T(\Omega;R)&=\frac{\beta z_1}{f} \!\!\int\!\! \mathrm{d}\Omega'\, k_0(\Omega') a_0(\Omega-\Omega') \Theta(-\Omega') e^{-\Delta k(\Omega')z_1} P(\Omega',R), \\ P(\Omega',R)&=\!\!\int_{0}^{R}\!\!\mathrm{d}r\,r {\mathcal{E}}_C(r,\Omega') \exp(\zeta(\Omega') r^2) \end{aligned}$$
where $\zeta (\Omega )=\tfrac {1}{2}k_0(\Omega )z_1/f^2$. The form of Eq. (23) allows us to analyze the experimental THz-IR beam shape which is contained completely in $P(\Omega,R)$ which, for a plane wavefront, is a real quantity. Hence one can fit the relative form of $S_T(\Omega,\tau =0;R)\propto |b_T(\Omega ;R)|^2\propto P^2(\Omega ;R)$ (where we assume the beam properties are smoothly varying over the finite width of $a(\Omega )$) from a set of measurements with varying $R$, as performed in Sec. 3.3. Note that in principle one could apply this to an arbitrary beam shape, noting that $\partial _R P \propto {\mathcal {E}}_C(R)$, although here we employ an ansatz for the beam profile and fit the parameters for each THz-IR frequency.

One can show that substituting a collimated Gaussian beam, ${\mathcal {E}}_C(r,\Omega )={\mathcal {E}}_{C0}e^{-r^2/w_C^2}$, with $w_C=2f/k_0 w_0$ and ${\mathcal {E}}_{C0}=-i(w_0/w_C){\mathcal {E}}_0$, into Eq. (23) yields the same solution Eq. (17). However, for the conical experimental beam shape, we take the model function in Eq. (2), which when substituted into Eq. (23) yields:

$$P(R) = \frac{1}{2}\left( \frac{1-e^{{-}d_1 R^2}}{d_1} - \frac{1-e^{{-}d_2 R^2}}{d_2} \right)$$
where $d_1(\Omega )=1/w_C^2(\Omega ) + \zeta (\Omega )$ and $d_2(\Omega )=d_1(\Omega )+1/h^2(\Omega )$. As described in the main text, Eq. (24) is used to fit $S_T(\Omega,\tau =0;R)$ as shown in Fig. 4, to allow one to evaluate the propagation factors in Eq. (23) and correct the THz-IR spectra relative to the EFISH signal.

Funding

Deutsche Forschungsgemeinschaft (422213477).

Acknowledgments

We gratefully acknowledge funding by the German Research Foundation (DFG) via the Collaborative Research Center TRR 288 (422213477, project B08).

Disclosures

The authors declare no conflicts of interest.

Data availability

The raw experimental, analyzed, and fitted data, as well as simulation results have been placed on the Zenodo data repository [50]. Details of the analysis/simulation codes may be obtained from the authors upon reasonable request.

References

1. A. Pashkin, A. Sell, T. Kampfrath, and R. Huber, “Electric and magnetic terahertz nonlinearities resolved on the sub-cycle scale,” New J. Phys. 15(6), 065003 (2013). [CrossRef]  

2. B. Mayer, C. Schmidt, A. Grupp, J. Bühler, J. Oelmann, R. E. Marvel, R. F. Haglund, T. Oka, D. Brida, A. Leitenstorfer, and A. Pashkin, “Tunneling breakdown of a strongly correlated insulating state in VO2 induced by intense multiterahertz excitation,” Phys. Rev. B 91(23), 235113 (2015). [CrossRef]  

3. H. Shirai, T.-T. Yeh, Y. Nomura, C.-W. Luo, and T. Fuji, “Ultrabroadband midinfrared pump-probe spectroscopy using chirped-pulse up-conversion in gases,” Phys. Rev. Appl. 3(5), 051002 (2015). [CrossRef]  

4. T.-T. Yeh, C.-M. Tu, W.-H. Lin, C.-M. Cheng, W.-Y. Tzeng, C.-Y. Chang, H. Shirai, T. Fuji, R. Sankar, F.-C. Chou, M. M. Gospodinov, T. Kobayashi, and C.-W. Luo, “Femtosecond time-evolution of mid-infrared spectral line shapes of Dirac fermions in topological insulators,” Sci. Rep. 10(1), 9803 (2020). [CrossRef]  

5. C. Riek, D. V. Seletskiy, and A. Leitenstorfer, “Femtosecond measurements of electric fields: From classical amplitudes to quantum fluctuations,” Eur. J. Phys. 38(2), 024003 (2017). [CrossRef]  

6. F. Junginger, A. Sell, O. Schubert, B. Mayer, D. Brida, M. Marangoni, G. Cerullo, A. Leitenstorfer, and R. Huber, “Single-cycle multiterahertz transients with peak fields above 10 MV/cm,” Opt. Lett. 35(15), 2645 (2010). [CrossRef]  

7. A. Sell, A. Leitenstorfer, and R. Huber, “Phase-locked generation and field-resolved detection of widely tunable terahertz pulses with amplitudes exceeding 100 MV/cm,” Opt. Lett. 33(23), 2767 (2008). [CrossRef]  

8. D. J. Cook and R. M. Hochstrasser, “Intense terahertz pulses by four-wave rectification in air,” Opt. Lett. 25(16), 1210 (2000). [CrossRef]  

9. M. Kreß, T. Löffler, S. Eden, M. Thomson, and H. G. Roskos, “Terahertz-pulse generation by photoionization of air with laser pulses composed of both fundamental and second-harmonic waves,” Opt. Lett. 29(10), 1120–1122 (2004). [CrossRef]  

10. T. Bartel, P. Gaal, K. Reimann, M. Woerner, and T. Elsaesser, “Generation of single-cycle THz transients with high electric-field amplitudes,” Opt. Lett. 30(20), 2805 (2005). [CrossRef]  

11. H. Roskos, M. Thomson, M. Kreß, and T. Löffler, “Broadband THz emission from gas plasmas induced by femtosecond optical pulses: From fundamentals to applications,” Laser Photonics Rev. 1(4), 349–368 (2007). [CrossRef]  

12. J. Buldt, H. Stark, M. Müller, C. Grebing, C. Jauregui, and J. Limpert, “Gas-plasma-based generation of broadband terahertz radiation with 640 mW average power,” Opt. Lett. 46(20), 5256 (2021). [CrossRef]  

13. M. Kreß, T. Löffler, M. D. Thomson, R. Dörner, H. Gimpel, K. Zrost, T. Ergler, R. Moshammer, U. Morgner, J. Ullrich, and H. G. Roskos, “Determination of the carrier-envelope phase of few-cycle laser pulses with terahertz-emission spectroscopy,” Nat. Phys. 2(5), 327–331 (2006). [CrossRef]  

14. T. Fuji and T. Suzuki, “Generation of sub-two-cycle mid-infrared pulses by four-wave mixing through filamentation in air,” Opt. Lett. 32(22), 3330 (2007). [CrossRef]  

15. K. Y. Kim, A. J. Taylor, J. H. Glownia, and G. Rodriguez, “Coherent control of terahertz supercontinuum generation in ultrafast laser–gas interactions,” Nat. Photonics 2(10), 605–609 (2008). [CrossRef]  

16. M. D. Thomson, V. Blank, and H. G. Roskos, “Terahertz white-light pulses from an air plasma photo-induced by incommensurate two-color optical fields,” Opt. Express 18(22), 23173 (2010). [CrossRef]  

17. C. R. Baiz and K. J. Kubarych, “Ultrabroadband detection of a mid-IR continuum by chirped-pulse upconversion,” Opt. Lett. 36(2), 187 (2011). [CrossRef]  

18. C. Calabrese, A. M. Stingel, L. Shen, and P. B. Petersen, “Ultrafast continuum mid-infrared spectroscopy: Probing the entire vibrational spectrum in a single laser shot with femtosecond time resolution,” Opt. Lett. 37(12), 2265 (2012). [CrossRef]  

19. Y. Nomura, H. Shirai, and T. Fuji, “Frequency-resolved optical gating capable of carrier-envelope phase determination,” Nat. Commun. 4(1), 2820 (2013). [CrossRef]  

20. E. Matsubara, M. Nagai, and M. Ashida, “Coherent infrared spectroscopy system from terahertz to near infrared using air plasma produced by 10-fs pulses,” J. Opt. Soc. Am. B 30(6), 1627 (2013). [CrossRef]  

21. C. Kübler, R. Huber, S. Tübel, and A. Leitenstorfer, “Ultrabroadband detection of multi-terahertz field transients with GaSe electro-optic sensors: Approaching the near infrared,” Appl. Phys. Lett. 85(16), 3360–3362 (2004). [CrossRef]  

22. N. Karpowicz, J. Dai, X. Lu, Y. Chen, M. Yamaguchi, H. Zhao, X.-C. Zhang, L. Zhang, C. Zhang, M. Price-Gallagher, C. Fletcher, O. Mamer, A. Lesimple, and K. Johnson, “Coherent heterodyne time-domain spectrometry covering the entire “terahertz gap”,” Appl. Phys. Lett. 92(1), 011131 (2008). [CrossRef]  

23. X. Lu, N. Karpowicz, Y. Chen, and X.-C. Zhang, “Systematic study of broadband terahertz gas sensor,” Appl. Phys. Lett. 93(26), 261106 (2008). [CrossRef]  

24. X. Lu, N. Karpowicz, and X.-C. Zhang, “Broadband terahertz detection with selected gases,” Journal of the Optical Society of America B 26(9), A66 (2009). [CrossRef]  

25. X. Lu and X.-C. Zhang, “Investigation of ultra-broadband terahertz time-domain spectroscopy with terahertz wave gas photonics,” Frontiers of Optoelectronics 7(2), 121–155 (2014). [CrossRef]  

26. A. V. Borodin, M. N. Esaulkov, A. A. Frolov, A. P. Shkurinov, and V. Y. Panchenko, “Possibility of direct estimation of terahertz pulse electric field,” Opt. Lett. 39(14), 4092 (2014). [CrossRef]  

27. H.-W. Du, F. Tang, D.-Y. Zhang, W. Sheng, and J.-Y. Mao, “Calibration of the field strength of broadband terahertz radiation in air coherent detection technique,” J. Appl. Phys. 124(14), 143101 (2018). [CrossRef]  

28. M. D. Thomson, V. Blank, and H. G. Roskos, “Recovery of ultra-broadband terahertz pulses from sum-frequency spectrograms using a generalized deconvolution method,” EPJ Web Conf. 41, 09011 (2013). [CrossRef]  

29. V. Blank, M. D. Thomson, and H. G. Roskos, “Spatio-spectral characteristics of ultra-broadband THz emission from two-colour photoexcited gas plasmas and their impact for nonlinear spectroscopy,” New J. Phys. 15(7), 075023 (2013). [CrossRef]  

30. F. Meng, M. D. Thomson, B. E. Sernelius, M. Jörger, and H. G. Roskos, “Ultrafast dynamic conductivity and scattering rate saturation of photoexcited charge carriers in silicon investigated with a midinfrared continuum probe,” Phys. Rev. B 91(7), 075201 (2015). [CrossRef]  

31. M. Clerici, D. Faccio, L. Caspani, M. Peccianti, O. Yaakobi, B. E. Schmidt, M. Shalaby, F. Vidal, F. Légaré, T. Ozaki, and R. Morandotti, “Spectrally resolved wave-mixing between near- and far-infrared pulses in gas,” New J. Phys. 15(12), 125011 (2013). [CrossRef]  

32. Y. Nomura, Y.-T. Wang, T. Kozai, H. Shirai, A. Yabushita, C.-W. Luo, S. Nakanishi, and T. Fuji, “Single-shot detection of mid-infrared spectra by chirped-pulse upconversion with four-wave difference frequency generation in gases,” Opt. Express 21(15), 18249 (2013). [CrossRef]  

33. H. He and X.-C. Zhang, “Analysis of Gouy phase shift for optimizing terahertz air-biased-coherent-detection,” Appl. Phys. Lett. 100(6), 061105 (2012). [CrossRef]  

34. G. Bjorklund, “Effects of focusing on third-order nonlinear processes in isotropic media,” IEEE J. Quantum Electron. 11(6), 287–296 (1975). [CrossRef]  

35. S. Linden, J. Kuhl, and H. Giessen, XFROG—Cross-correlation Frequency-resolved Optical Gating (Springer US, Boston, MA, 2000), 313–322.

36. R. Trebino, R. Jafari, S.A. Akturk, P. Bowlan, P. Guang, P. Zhu, E. Escoto, and G. Steinmeyer, “Highly reliable measurement of ultrashort laser pulse,” J. Appl. Phys. 128, 171103 (2020). [CrossRef]  

37. W.-H. Huang, Y. Zhao, S. Kusama, F. Kumaki, C.-W. Luo, and T. Fuji, “Generation of sub-half-cycle 10 µm pulses through filamentation at kilohertz repetition rates,” Opt. Express 28 (24), 36527 (2020). [CrossRef]  

38. E. D. Palik, ed., Handbook of Optical Constants of Solids. 1 (Acad. Press, Boston, 2004), 6th ed.

39. M. C. Tobin, “The infrared spectra of polymers. III. The infrared and Raman spectra of isotactic polypropylene,” J. Phys. Chem. 64(2), 216–219 (1960). [CrossRef]  

40. A. Beer, D. Hershkovitz, and S. Fleischer, “Iris-assisted terahertz field-induced second-harmonic generation in air,” Opt. Lett. 44(21), 5190 (2019). [CrossRef]  

41. P. Klarskov, A. C. Strikwerda, K. Iwaszczuk, and P. U. Jepsen, “Experimental three-dimensional beam profiling and modeling of a terahertz beam generated from a two-color air plasma,” New J. Phys. 15(7), 075012 (2013). [CrossRef]  

42. V. A. Andreeva, O. G. Kosareva, N. A. Panov, D. E. Shipilo, P. M. Solyankin, M. N. Esaulkov, P. González de Alaiza Martínez, A. P. Shkurinov, V. A. Makarov, L. Bergé, and S. L. Chin, “Ultrabroad Terahertz Spectrum Generation from an Air-Based Filament Plasma,” Phys. Rev. Lett. 116(6), 063902 (2016). [CrossRef]  

43. L. Rothman, I. Gordon, A. Barbe, et al., “The HITRAN 2008 molecular spectroscopic database,” J. Quant. Spectrosc. Radiat. Transf. 110(9-10), 533–572 (2009). [CrossRef]  

44. A. A. Voronin and A. M. Zheltikov, “The generalized Sellmeier equation for air,” Sci. Rep. 7(1), 46111 (2017). [CrossRef]  

45. T. Löffler, M. Kreß, M. D. Thomson, T. Hahn, N. Hasegawa, and H. G. Roskos, “Comparative performance of terahertz emitters in amplifier-laser-based systems,” Semicond. Sci. Technol. 20(7), S134–S141 (2005). [CrossRef]  

46. M. Clerici, L. Caspani, E. Rubino, M. Peccianti, M. Cassataro, A. Busacca, T. Ozaki, D. Faccio, and R. Morandotti, “Counterpropagating frequency mixing with terahertz waves in diamond,” Opt. Lett. 38(2), 178 (2013). [CrossRef]  

47. A. Tomasino, R. Piccoli, Y. Jestin, S. Delprat, M. Chaker, M. Peccianti, M. Clerici, A. Busacca, L. Razzari, and R. Morandotti, “Invited Article: Ultra-broadband terahertz coherent detection via a silicon nitride-based deep sub-wavelength metallic slit,” APL Photonics 3(11), 110805 (2018). [CrossRef]  

48. R. W. Boyd, Nonlinear Optics (Academic Press, Amsterdam ; Boston, 2008), 3rd ed.

49. J. W. Goodman, Introduction to Fourier Optics (Roberts & Co., Englewood, Colo, 2005), 3rd ed.

50. M. D. Thomson, K. Warawa, F. Meng, and H. Roskos, "Determining the absolute temporal field of ultra-broadband terahertz-infrared pulses with field-induced second-harmonic spectrograms - Data," Zenodo (2023), https://doi.org/10.5281/zenodo.7671804

Data availability

The raw experimental, analyzed, and fitted data, as well as simulation results have been placed on the Zenodo data repository [50]. Details of the analysis/simulation codes may be obtained from the authors upon reasonable request.

50. M. D. Thomson, K. Warawa, F. Meng, and H. Roskos, "Determining the absolute temporal field of ultra-broadband terahertz-infrared pulses with field-induced second-harmonic spectrograms - Data," Zenodo (2023), https://doi.org/10.5281/zenodo.7671804

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 (6)

Fig. 1.
Fig. 1. (a) Response function for electro-optic time-domain sampling for different optical detection pulse duration, $T_{\mathrm {FWHM}}$ ($S(\nu )=\exp (-\tfrac {\pi ^2}{4\ln {2}}\nu ^2{T_{\mathrm {FWHM}}}^2)$). (b) Schematic of frequency conversion processes for detection of THz-IR pulses, including FW-DFG (equivalent to TFISH for THz frequencies) used here in air. (c) Setup used for spectrogram/ABCD experiments. (d) Example spectrogram using detection pulses with ${T_{\mathrm {FWHM}}}=150{\,{\mathrm {fs}}}$.
Fig. 2.
Fig. 2. (a) Reference spectrogram, and (b) spectrogram with 15-$\mu$m-thick polypropylene film in sample position. Additional chirp for reference/sample (compared to Fig. 1(d)) due to the inclusion of an additional 525-$\mu$m-thick high-resistivity silicon wafer in THz-IR beam path for these measurements. (c) Corresponding intensity transmission and group delay of polypropylene sample extracted from $M(\nu )$ and ${\bar \tau }(\nu )$ of the reference/sample spectrograms (solid curves), and model data using literature IR absorption line data [39].
Fig. 3.
Fig. 3. Measured spectrograms with (a) FW-DFG alone ($S_T(\nu,\tau )$), and (b) both FW-DFG and EFISH $S_{TE}(\nu,\tau )$ (the latter with uni-polar bias $\cal {E}_{\textrm {b}}$), focusing on low-frequency range (see Fig. 1(d) for full spectral range for reference pulse). (c) Corresponding frequency marginals and that from (d) interferometric residual $\Delta S=S_{TE}-S_T-S_E$. (e) Low-frequency ABCD time-domain field for the same pulses and (f) corresponding spectral intensity and phase. Also included in (f) is the phase obtained from integrating the group delay ${\bar \tau }(\nu )$ extracted from the spectrogram (see Sec. 5.1) after adjusting to the absolute ABCD phase in the interval shown.
Fig. 4.
Fig. 4. Beam profile results from measurements with variable aperture in collimated THz-IR beam path. (a) Experimental and (b) fitted FW-DFG spectra $S_T(\nu,\tau =0,R)$ vs. iris radius $R$, the latter based on the collimated radial beam profile $I_C(r,\nu )$ in (c) (plotted here normalized for each frequency) – see Eq. (2) for model function for ${\mathcal {E}}_C(\nu,R)$ and Appendix 5.2 (Eqs. (23) and (24)) for the integral transformation between ${\mathcal {E}}_C(\nu,R)$ and $S_T(\nu,R)$. Schematic of experimental geometry shown left (${\mathcal {E}}(z,\nu )$ denotes the axial THz-IR field about the focal point, $b_T(z,\nu )$ the FISH signal field, and ${\mathcal {E}}_C(r,\nu )$ the field of the collimated THz/IR beam).
Fig. 5.
Fig. 5. (a) Calculated phase mismatch ${\Delta k}$ vs. frequency, due to dispersion of air (for two values of relative humidity ${c_{\mathrm {H_2O}}}$); corresponding $1/e$-coherence length scale at right. (b) Normalized on-axis field amplitude vs distance about focal plane based on the fitted THz-IR beam parameters from Fig. 4. (c) Corresponding field responsivity factor $H(\nu ')$ for FW-DFG (blue solid curve), relative to EFISH for geometry used here. Dashed curves: responsivity factors due to dispersive phase mismatch alone. (d) Recovered temporal electric field in focal plane: raw result using simple focal-plane treatment (dashed) and after correcting for relative FW-DFG responsivity in (c) due to propagation effects.
Fig. 6.
Fig. 6. (a) Simulated evolution of the FW-DFG/SFG signal intensity spectrum for selected positions $z$ relative to focal plane. Results are based on the experimental THz-IR spectrum and beam parameters determined above, and include both results with (solid curves, ${\Delta k}\neq 0$) and without (dotted curves, ${\Delta k}=0$) dispersive phase mismatch. (b,c) Focal spot diameter (top), Rayleigh range (middle) and relative field responsivity (bottom, as per Fig. 5(c)) for two experimental scenarios with a Gaussian THz-IR beam and frequency-independent collimated beam diameter $2w_C$: (b) $2w_C=25$ mm, $f=76.2$ mm, $f_1=125$ mm; (c) $2w_C=15$ mm, $f=152.4$ mm, $f_1=250$ mm.

Equations (26)

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

S T E ( ν , τ ) | b T ( ν , τ ) + b E ( ν ) | 2 = | b T ( ν , τ ) | 2 + | b E ( ν ) | 2 + 2 | b T ( ν , τ ) b E ( ν ) | cos Φ ( ν ) .
F T = d t E 2 ( t ) = J T P E E b 2
E C ( r , ν ) = E C 0 ( ν ) exp ( r 2 w C 2 ( ν ) ) [ 1 exp ( r 2 h 2 ( ν ) ) ] ,
P i ( 3 ) = 1 2 ε 0 χ i j k l ( 3 ) ( A j A k E l + A j E k A k + E j A k A l ) ,
P ( 3 ) ( Ω ) = F { γ a ( t ) E ( t ) } = γ 2 π d Ω a ( Ω Ω ) E ( Ω ) ,
  2 i k 2 z B = 2 B μ 0 ω 2 2 P ( 3 ) e i k 2 z
z B = i μ 0 ω 2 2 k 2 P ( 3 ) B ( Ω ) = i μ 0 ω 2 2 k 2 P ( 3 ) ( Ω ) δ z .
B T ( Ω , τ ) = β 2 π δ z d Ω e i Ω τ a ( Ω Ω ) E ( Ω ) , β = i 3 χ ( 3 ) ω 2 2 4 k 2 c 2
J T = 1 2 π d τ d Ω S T ( Ω , τ ) = 1 2 π d τ d Ω | B ( Ω , τ ) | 2 = γ 2 F a 1 2 π d Ω | E ( Ω ) | 2 ,
F T = 1 2 π d Ω | E ( Ω ) | 2 = d t E 2 ( t ) = J T P E E b 2 ,
M ( Ω ) = d τ S T ( Ω , τ ) = d τ | B T ( Ω , τ ) | 2 = β 2 2 π d Ω | a ( Ω Ω ) | 2 | E ( Ω ) | 2 ,
τ ¯ ( Ω ) = 1 M ( Ω ) d Ω τ S ( Ω , τ ) 2 π i M ( Ω ) d Ω E ( Ω ) Ω E ( Ω ) | a ( Ω Ω ) | 2
2 π M ( Ω ) d Ω | E ( Ω ) | 2 | a ( Ω Ω ) | 2 T g ( Ω ) ,
A ( r , Ω ) = A 0 ( Ω ) Q 1 ( z ) R 1 ( r , z ) , Q 1 ( z ) = 1 1 i z / z 1 , R 1 ( r , z ) = exp ( Q 1 ( z ) r 2 w 1 2 )
P ( 3 ) ( r , Ω ) = γ 2 π Q 1 2 R 1 2 ( r , z ) d Ω a 0 ( Ω Ω ) E ( ω )
z b ( z , Ω ) = i β 2 π Q 1 ( z ) d Ω a 0 ( Ω Ω ) E ( z , Ω ) e i Δ k ( Ω , Ω ) z
Δ k ( Ω , Ω ) = k 2 ( Ω ) k ( Ω ) [ 2 k 10 + ( Ω Ω ) k 10 ]
E ( z , Ω ) = E 0 ( Ω ) Q 0 ( z , Ω ) , Q 0 ( z , Ω ) = 1 1 i z / z 0 ( Ω )
b T ( z , Ω ) = i β d Ω a 0 ( Ω Ω ) E 0 ( Ω ) z 1 z 0 ( Ω ) z 1 z 0 ( Ω ) e Δ k ( Ω , Ω ) z 1 Θ ( Ω )
b E ( Ω ) = i β Z E a 0 ( Ω ) E b ,
b T ( z , Ω ) = i β 2 π d Ω a 0 ( Ω Ω ) d z Q 1 ( z ) e i Δ k ( Ω ) z E ( z , Ω )
b T ( Ω ) b E = H ( Ω ) B T ( Ω ) B E , H ( Ω ) = 1 Z E d z Q 1 ( z ) e i Δ k ( Ω ) z E ( z , Ω ) E ( 0 , Ω )
E ~ ( u , v , 0 ) = i 2 π f k 0 E C ( f k 0 u , f k 0 v )
E ( 0 , z ) = i k 0 f 0 R d r r exp ( i k 0 2 f 2 r 2 z ) E C ( r ) .
b T ( Ω ; R ) = β z 1 f d Ω k 0 ( Ω ) a 0 ( Ω Ω ) Θ ( Ω ) e Δ k ( Ω ) z 1 P ( Ω , R ) , P ( Ω , R ) = 0 R d r r E C ( r , Ω ) exp ( ζ ( Ω ) r 2 )
P ( R ) = 1 2 ( 1 e d 1 R 2 d 1 1 e d 2 R 2 d 2 )
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.