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

Time-domain characterization of electric field vector in multi-terahertz pulses using polarization-modulated electro-optic sampling

Open Access Open Access

Abstract

We demonstrated characterizing the electric field waveform of multi-terahertz pulses (10 − 50 THz) as vector quantities in the time domain by applying the polarization modulated electro-optic sampling (POMEOS) method. The problem of an ultrabroadband gate pulse was solved by modifying the fitting function in POMEOS, and its validity was confirmed through numerical simulations. High accuracy and precision of approximately 1 mrad with 3 s accumulation were demonstrated. Our method can be applied not only to multi-terahertz polarization measurements for linear response but also to the evaluation of the driving field of intense pulses for nonlinear response or material control.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Recently, laser-based coherent pulsed light sources with a fixed carrier-envelope phase have been developed in the 10–50 THz range, which lies between the terahertz (THz) and mid-infrared regions and is often called the multi-THz region [17]. The detection of the time-domain waveforms of the electric field is possible using ultrashort optical pulses with a pulse width of approximately 10 fs, in a similar configuration to the conventional THz time-domain spectroscopy (TDS) around the 1 THz region. Such a multi-THz TDS is highly beneficial for measuring the broadband and complex response functions of materials and the ultrafast dynamics in the pump-probe scheme [8,9]. More importantly, the response function of a material is generally a tensor quantity. Multi-THz polarization-resolved measurements are required to obtain the off-diagonal components that reveal vibrational circular dichroism or anomalous Hall conductivity. Moreover, recently the ultrafast control of material properties by utilizing the time-varying polarization degrees of freedom in intense multi-THz light pulses has attracted tremendous attention for controlling topological materials [1016], nonlinear phononics [17,18], and chiral phonons [19,20]. Therefore, there is a high demand for an evaluation technique for the vector field waveforms of multi-THz pulses.

Conventionally, polarization-sensitive measurements have been implemented using polarization-modulated Fourier transform infrared (FTIR) spectroscopy by using a photoelastic modulator (PEM) for the mid-infrared region [21,22]. Despite the advantages of high frequency resolution and a broad spectral range, it is difficult to apply FTIR to time-resolved measurements on an ultrafast femtosecond time scale. It is also impossible to determine the time evolution of the electric field vector because the phase is not detected.

Another challenge for polarization-resolved spectroscopy in the multi-THz region is the absence of necessary optical elements. For approximately 1 THz, polarization-resolved TDS can be performed with free-standing wire-grid polarizers (WGPs) by simple projection in two directions [2325] or by modulating the signal by the high-speed rotation of a polarizer to eliminate artifacts [26]. However, it is difficult to fabricate free-standing WGPs in the multi-THz region because of much shorter wavelengths. Although WGPs are available on substrates such as Thallium Bromo-Iodide (KRS-5), the group velocity dispersion of the substrate severely distorts the waveform. Apart from WGPs, there are other approaches, such as multicontact photoconductive antennas [27,28] and air-biased coherent detection [29,30]. However, these methods also have limitations in terms of the bandwidth or intensity of the laser, as well as difficulty with cross-talk effects.

In this work, we applied the polarization-modulated electro-optic sampling (POMEOS) method, which was developed in THz-TDS around 1 THz [3133], to measure the waveforms of multi-THz pulses as vector values. In POMEOS, modulating the polarization of the optical gate pulse generates the modulated electro-optic (EO) signal, which allows us to directly determine both the amplitude and azimuthal angle of the electric field without any polarizers. However, the extension of POMEOS to the multi-THz region was not obvious because the 10-fs-class ultrashort pulse contains a broadband spectrum, where the retardation by the polarization modulator has a wavelength dependence. To solve this problem, we modified the fitting function of the modulated signals. Owing to this improvement, the time-domain characterization of the multi-THz electric field vector at 10 − 50 THz was successfully demonstrated by modulating the broadband gate pulse. Based on the numerical simulation, we delve into the adjustments made to POMEOS for the 1-THz region. We also address the prerequisites for accurate measurements in the multi-THz range. These results open a new avenue for ultrafast science in the multi-THz region.

2. Experiments and results

A Yb:KGW regenerative amplifier (PH2-2mJ-SP, Light Conversion) with a center wavelength of 1030 nm and a pulse width of 160 fs was used as the light source. A part of the output was compressed to less than 13 fs with a pulse energy of 60 µJ, by using multi-plate broadening and dispersion compensation [7,34,35]. The beam is split into two beams with a controlled time delay for multi-THz TDS [7,8]. A 10-µm-thick GaSe crystal with ab-plane was used for generating broadband multi-THz pulses which cover 10–50 THz by intra-pulse difference frequency generation process. Another 10-µm-thick GaSe crystal with ab-plane was used for the detection of the electric field of multi-THz pulses with the EO sampling (EOS) method [1,2,36].

In conventional EOS, linearly polarized gate pulses with fixed polarization directions enter the EO crystals. Through propagation in the detection crystal, the gate pulse acquires ellipticity proportional to the multi-THz electric field, which is measured by balanced detection using a quarter-wave plate (QWP) and a polarization beam splitter. In contrast, in this study, the azimuthal angle of the linearly polarized gate pulses before the EO crystal was modulated by a PEM (I/FS50, HINDS Instruments) and achromatic QWP, as shown in Fig. 1(a). The PEM induces birefringence in the crystal, which modulates the ellipticity of the gate pulse when the modulation axis of PEM is set to −45° from the incident polarization direction. The ellipticity was subsequently converted to a modulation of the azimuthal angle of the linear polarization using a QWP whose optical axis was set parallel to the initial polarization. The phase retardation of the PEM can be expressed as $\delta (t )= {\delta _0}\sin ({2\pi {f_{\textrm{PEM}}}t} )$, where ${\delta _0}$ is the modulation depth, ${f_{\textrm{PEM}}}$ is the modulation frequency, and t is the laboratory time. After the QWP1, the polarization azimuthal angle $\psi $ becomes half of $\delta $. When the EO crystal has three-fold rotational symmetry and the [100] axis is horizontal, the modulated EO signal can be expressed as ${E_{\textrm{THz}}}\cos ({2\psi + \theta } )$, where ${E_{\textrm{THz}}}$ and $\theta $ are the amplitude and the azimuthal angle of the multi-THz electric field vector, respectively. Note that the EO signal has a maximum value when both gate and multi-THz polarization are also horizontal without modulation ($\theta = \psi = {\delta _0} = 0$). In total, the EO signal as a function of the laboratory time $S(t )$ becomes

$$S(t )= {E_{\textrm{THz}}}\cos ({{\delta_0}\sin ({2\pi {f_{\textrm{PEM}}}t} )+ \theta } ).$$

 figure: Fig. 1.

Fig. 1. (a) Schematic of POMEOS setup. The orientations of the components are shown in the inset. (b) PEM phase and polarization of gate pulse for each pulse. Because of the relationship between modulation frequency and repetition rate, aliasing occurs. (c, d): An example of raw data of BPD signal (c) and PEM phase (d) for each laser pulse. Blue: chopper closed, Red: chopper opened. (e) Corresponding data as functions of PEM phase. Dot: measured data, circle: interpolated data. (f) Subtracted signal where the effect of residual birefringence is eliminated. Red circle: measured, black line: fitted. (g) Experimental result for a monocycle, broadband multi-THz pulse. (h) The power spectra for the waveform in (g). Red and blue show x and y components.

Download Full Size | PDF

In the experiments, the modulation depth ${\delta _0}$ was set to $\pi /2$ at 1030 nm, the center wavelength of gate spectrum.

Figure 1(b) shows a schematic of the retardation of the PEM and polarization state of the gate pulses in the laboratory time. If the repetition rate of the laser ${f_{\textrm{rep}}}$ is much higher than ${f_{\textrm{PEM}}}$, the modulated EO signal can be easily fitted using Eq. (1). However, ${f_{\textrm{rep}}}$ of a regenerative amplifier is often lower than ${f_{\textrm{PEM}}}$, e.g., ${f_{\textrm{rep}}} = 3\; \textrm{kHz}$ and ${f_{\textrm{PEM}}}\sim 50\; \textrm{kHz}$ in our case. In such a case, the EO signal is not a smooth function of the laboratory time because of aliasing, as shown in Fig. 1(c). In a previous study [31], ${f_{\textrm{PEM}}}$ was locked to $m/n\; \times {f_{\textrm{rep}}}$ to resolve the aliasing, where m and n are small integers. By contrast, we measured the EO signal S and the PEM reference signal (the square wave of ${f_{\textrm{PEM}}}$) for each laser pulse simultaneously for free-running ${f_{\textrm{rep}}}$ and ${f_{\textrm{PEM}}}$. In this free-running scheme, the setup becomes simpler than synchronized scheme because we do not need a frequency divider/multiplier, high voltage oscillator, nor active feedback of PEM frequency. By post data analysis, the PEM phase, which corresponds to $2\pi {f_{\textrm{PEM}}}t$, was determined by the phase of reference signal at trigger timing. The PEM phases fill the whole range of 0–$2\pi $ with several tens of laser pulses (about 40 pulses here) because of a moderate aliasing, as shown in Fig. 1(d). This filling feature is quite different compared with the discrete PEM phases in synchronized scheme [31]. The EO signal can then be plotted as a smooth function of the PEM phase, as shown in Fig. 1(e). An optical chopper with a modulation frequency of ${f_{\textrm{rep}}}/2$ was inserted to eliminate artifacts caused by residual birefringence in the EO crystal. The red and blue points in Figs. 1(c)–(e) correspond to the data with and without the multi-THz field, respectively. The latter shows the effect of the residual birefringence, which must be removed before the fitting analysis. To determine the difference between the data with and without the multi-THz field, a dataset with the same PEM phase was necessary. For this purpose, the PEM phases were split into 60 segments and the pulse-resolved data (dots in Fig. 1(e)) were averaged for each segment. Subsequently, these representative points were interpolated into evenly spaced PEM phases, as indicated by the circles in Fig. 1(e). After averaging and interpolation, the difference signal caused by the multi-THz field was obtained, as shown in Fig. 1(f). The data were well fitted by the function ${a_0} + {E_{\textrm{THz}}}\cos ({{\delta_0}\sin ({\varphi - {\varphi_0}} )+ \theta } )$, which is a modified function of Eq. (1). The offset term ${a_0}$ originates mainly from the wavelength dependence of the modulation depth of the PEM, because the gate pulse has a large bandwidth from 900 to 1100 nm [7], where we expect 22% difference in the modulation depth between the spectral edges. Other origins of ${a_0}$, such as chirp of gate pulse or residual birefringence [37], are also discussed in Section 3 and Supplemental document. From the fitting, ${E_{\textrm{THz}}}$ and $\theta $ were determined at each delay time between the multi-THz and gate pulses. The fitting function includes five parameters: ${a_0},\textrm{}{E_{\textrm{THz}}},\textrm{}{\delta _0},\textrm{}{\varphi _0}$, and $\theta $. Among these, ${\delta _0}$ and ${\varphi _0}$ depend only on the PEM setting. These were fixed to the values obtained by fitting certain data, typically at the peak amplitude. Therefore, there remained only three fitting parameters, ${a_0},\textrm{}{E_{\textrm{THz}}},$ and $\theta $, when the delay time of the gate pulse was scanned. From ${E_{\textrm{THz}}}$ and $\theta $ at each delay time, the waveform of multi-THz pulses was obtained as vector values. The angle of $\theta = 0$ was defined by measuring a field with a known polarization, typically horizontal polarization. Figure 1(g) shows the experimental results for the waveform of a broadband multi-THz electric pulse generated by GaSe. Horizontally polarized monocycle pulses were successfully measured. Figure 1(h) shows the power spectra of x and y components for the same pulse, demonstrating broadband measurements covering 10–50 THz.

To confirm the characterization of the arbitrarily polarized pulses, we measured other polarization directions and ellipticities. First, linear polarization was tested: The GaSe crystal for multi-THz generation was rotated by 10°/step, so as to rotate the polarization plane by 30°/step. To ensure high polarization degree, a WGP on a KRS-5 substrate was inserted and rotated by 30°/step simultaneously. The results are presented in Figs. 2(a)–(f). In all polarization directions, linear polarization was effectively measured using our method. Second, a QWP for 33.3 THz was inserted to generate a circular polarization. Figure 2(g) shows the results when the angle between the incident polarization and the fast axis is 45°. The electric field was rotated clockwise along the $xy$-plane. We also observed a counterclockwise rotation when the QWP was rotated to –45° (Fig. 2(h)). Figure 2(i) shows the power spectra of the right- and left-circular polarization components of the pulse in Fig. 2(g).

 figure: Fig. 2.

Fig. 2. Results of the waveforms of electric field vectors. (a − f) Linearly polarized multi-THz pulses with 30° rotation step. (g, h) Circularly polarized pulse. Red arrows show the direction of the rotation of the field. (i) Power spectra of circular polarization in (g). Red and blue show right- and left-circular components, respectively.

Download Full Size | PDF

A right-handed component that was much larger than the left was successfully identified. The spectral peak at approximately 23 THz originates from a combination of the incident spectrum and the transmittance of the QWP.

Next, the performance of polarization measurements was evaluated using linearly polarized multi-THz pulses. To evaluate precision, we obtained the statistics of 100-times-repeated measurements at a fixed time delay. Figure 3(a) shows a histogram of the measured direction of the electric field. The accumulation time for each measurement was 1 s, which contained 3,000 laser pulses. The data followed a normal distribution, with a standard deviation of 1.7 mrad. The dependence on the accumulation pulse number is shown in Fig. 3(b). The standard deviation decreased in proportion to the inverse square root of the accumulation number, which is consistent with a normal distribution. The precision reached 1 mrad for 9,000 laser pulses (3 s). The accuracy was also evaluated by comparing the angle of the WGP and the detected field direction, as shown in Figs. 3(c)–(e), for different angle ranges. From Fig. 3(c), the angle was accurately measured in a wide range from –90° to +90°, as also confirmed in Figs. 2(a)–(f). As shown in Figs. 3(d) and (e) with enlarged scales, the measurement was also accurate in the small-angle ranges on the order of 1 mrad.

 figure: Fig. 3.

Fig. 3. Precision and accuracy. (a) Histogram of the measured azimuthal angles. The distribution is well fitted by Gaussian function (blue dashed line). (b) Accumulation pulse number dependence of azimuthal angle precision. (c − e) The relationship between measured azimuthal angles and known WGP angles in different angle ranges.

Download Full Size | PDF

3. Discussion

By applying POMEOS to an ultrashort gate pulse with a duration of approximately 10 fs, we successfully demonstrated the measurements of multi-THz time-domain waveforms as vector values. However, the fitting function required an additional offset term ${a_0}$ compared to a previous THz study [31]. Figure 4(a) shows an example of the obtained fitting parameters, which depend on the delay time. As indicated by the green curve, a finite ${a_0}$ depending on the delay time is necessary for better fitting. To explain this, we must go beyond the ideal situation, in which only the azimuthal angle is modulated without ellipticity. This is only the case when the retardation of the QWP and PEM is uniform over the entire gate pulse bandwidth. However, their wavelength dependence may affect the measurements in actual cases because our gate pulse has a broad bandwidth.

 figure: Fig. 4.

Fig. 4. (a) An experimental result for a linearly polarized pulse. Red: ${E_x}$, blue: ${E_y}$, green: offset term, black shaded: residue in fitting (right axis). (b − f) Results of numerical simulations. (b) Ideal case with achromatic PEM and QWP, (c) frequency-dependent PEM and achromatic QWP, and (d) frequency-dependent PEM and QWP. (e) and (f) are calculated for a chirped gate pulse, with the other parameters the same as (c) and (d), respectively.

Download Full Size | PDF

To evaluate this hypothesis, numerical simulations were conducted. Instead of the Pockels effect formalism often adopted for the EOS in the 1-THz region, we describe the interaction between the gate and multi-THz pulses as sum and difference frequency generation (SFG and DFG) processes so that the finite frequencies of the multi-THz pulses are taken into account. The second-order nonlinear susceptibility in a three-fold symmetric system has only 4 non-zero components with one independent parameter; ${\chi _{yyy}} ={-} {\chi _{yxx}} ={-} {\chi _{xxy}} ={-} {\chi _{xyx}}$. The transmitted gate pulse after the EO crystal ${{\boldsymbol E}^{g,\textrm{out}}}(\omega )$ is expressed as

$$\begin{aligned} \left[ {\begin{array}{{cc}} {E_x^{g,\textrm{out}}({\omega ,\textrm{}\tau } )}\\ {E_y^{g,\textrm{out}}({\omega ,\textrm{}\tau } )} \end{array}} \right] &= \left[ {\begin{array}{{c}} {E_x^g(\omega )}\\ {E_y^g(\omega )} \end{array}} \right] + i\omega b\int d\mathrm{\Omega \;\ }\left( {{e^{i\mathrm{\Omega }\tau }}\left[ {\begin{array}{{cc}} { - {E_y}(\mathrm{\Omega } )}&{ - {E_x}(\mathrm{\Omega } )}\\ { - {E_x}(\mathrm{\Omega } )}&{{E_y}(\mathrm{\Omega } )} \end{array}} \right]\left[ {\begin{array}{{c}} {E_x^g({\omega - \mathrm{\Omega }} )}\\ {E_y^g({\omega - \mathrm{\Omega }} )} \end{array}} \right]} \right.\\& + \left. {{e^{ - i\mathrm{\Omega }\tau }}\left[ {\begin{array}{{cc}} { - E_y^\mathrm{\ast }(\mathrm{\Omega } )}&{ - E_x^\mathrm{\ast }(\mathrm{\Omega } )}\\ { - E_x^\mathrm{\ast }(\mathrm{\Omega } )}&{E_y^\mathrm{\ast }(\mathrm{\Omega } )} \end{array}} \right]\left[ {\begin{array}{{c}} {E_x^g({\omega + \mathrm{\Omega }} )}\\ {E_y^g({\omega + \mathrm{\Omega }} )} \end{array}} \right]} \right) \end{aligned}, $$
where ${\boldsymbol E}(\mathrm{\Omega } )$ and ${{\boldsymbol E}^g}(\omega )$ are complex amplitude spectra of multi-THz and gate pulses, respectively, b is a coefficient proportional to the nonlinear susceptibility, and $\tau $ is the time delay between the gate and multi-THz pulses. SFG and DFG have been assumed to be much smaller than the incident light.

The gate pulse ${{\boldsymbol E}^g}(\omega )$ just before the EO crystal is described by using PEM and QWP Jones matrices,

$$\left[ {\begin{array}{{c}} {E_x^g({\omega ;\varphi } )}\\ {E_y^g({\omega ;\varphi } )} \end{array}} \right] = \left[ {\begin{array}{{cc}} 1&0\\ 0&{{e^{i\Delta (\omega )}}} \end{array}} \right]R\left( {\frac{\pi }{4}} \right)\textrm{}\left[ {\begin{array}{{cc}} 1&0\\ 0&{{e^{i{\delta_0}(\omega )\sin \varphi }}} \end{array}} \right]R\left( { - \frac{\pi }{4}} \right)\left[ {\begin{array}{{c}} 1\\ 0 \end{array}} \right], $$
where $\varphi$ is the PEM phase, ${\delta _0}(\omega )$ is the modulation depth of PEM considering frequency dependence, $\Delta (\omega )$ is phase retardation of QWP, and $R({ \pm \pi /4} )$ is the rotation matrix for angle of ${\pm} \pi /4$. We assume $\Delta (\omega )= \pi /2$ for an ideal QWP and $\Delta (\omega )= ({\pi /2} )\times \omega /{\omega _0}$ for a zero-order QWP, where ${\omega _0}$ is the center frequency of the gate that corresponds to the wavelength of 1030 nm. The modulation depth of PEM ${\delta _0}(\omega )$ is also assumed to have a similar frequency dependence for ideal (${\delta _0}(\omega )= {\delta _0}$) and actual (${\delta _0}(\omega )= {\delta _0} \times \omega /{\omega _0}$) situations. The gate pulse $E_x^{g,\; \textrm{WP}}({\omega ,\tau ;\varphi } )$ before the Wollaston prism in balance detection is described as
$$\left[ {\begin{array}{{c}} {E_x^{g,\textrm{WP}}({\omega ,\tau ;\varphi } )}\\ {E_y^{g,\textrm{WP}}({\omega ,\tau ;\varphi } )} \end{array}} \right] = R\left( {\frac{\pi }{4}} \right)\textrm{}\left[ {\begin{array}{{cc}} 1&0\\ 0&{{e^{i\Delta (\omega )}}} \end{array}} \right]R\left( { - \frac{\pi }{4}} \right)\left[ {\begin{array}{{c}} {E_x^{g,\textrm{out}}({\omega ,\textrm{}\tau ;\varphi } )}\\ {E_y^{g,\textrm{out}}({\omega ,\textrm{}\tau ;\varphi } )} \end{array}} \right].$$
Then, the balanced signal $S({\tau ;\varphi } )$ becomes
$$S({\tau ;\varphi } )= \int d\omega ({{{|{E_x^{g,\textrm{WP}}({\omega ,\tau ;\varphi } )} |}^2} - {{|{E_y^{g,\textrm{WP}}({\omega ,\tau ;\varphi } )} |}^2}} ).$$
Similar with the experiment, the signals $S({\tau ;\varphi } )$ with and without multi-THz pulse are calculated, and the difference between them are fitted with the model function for each delay time $\tau $. The gate spectrum ${{\boldsymbol E}^g}(\omega )$ was assumed to be a Gaussian function with transform-limited pulse duration of 13 fs. The multi-THz field was also set as Gaussian with a center frequency of 30 THz and an FWHM of 20 THz. The effects of the imperfections in the PEM or QWP and the chirp in the gate pulse were evaluated in the simulations.

Figure 4(b) shows the results for the ideal case, where the PEM and QWP are perfectly achromatic, and the gate pulse is transform-limited. The retrieved ${E_x}$ almost coincides with the input waveform, and ${E_y}$ and ${a_0}$ are zero. The residue of the fitting, defined as the sum of the squared fitting errors and plotted on the right axis, was also quite small in this ideal case. Figure 4(c) shows the result when the frequency dependence of the PEM retardation is considered. Although ${a_0}$ has a finite value compared with that in Fig. 4(b), ${E_x}$ and ${E_y}$ are not affected. The zero order QWP is shown in Fig. 4(d). ${E_x}$ does not largely change, whereas ${E_y}$ shows a larger artifact than that of Fig. 4(c). We also calculated for zero order QWP and achromatic PEM, and confirmed that the result is quite similar with Fig. 4(c). This means that the measurement is valid as long as either PEM or QWP is achromatic.

Figures 4(e) and 4(f) show the results for a chirped gate pulse with a duration of 24 fs; the other conditions were the same as those in Figs. 4(c) and 4(d), respectively. In both cases, ${E_x}$ is slightly deformed because of the poor time resolution. When the QWP was achromatic (Fig. 4(e)), ${E_y}$ was almost zero although ${a_0}$ and residue has finite values especially at the zero-crossing delay of ${E_x}$ waveform. In contrast, ${E_y}$ shows a large artifact when the QWP is of the zero order (Fig. 4(f)). From these simulations, we confirmed that the POMEOS fitting function, including the offset term ${a_0}$ is also valid for a 10-fs-class broadband gate pulse as long as achromatic QWP is used. We consider that our experimental results correspond to a mixture of slight chirp in the gate pulse and small imperfections in the achromatic QWP (∼0.004). We also conducted numerical simulations with residual birefringence in EO crystal (See Supplemental document). We confirmed the residual birefringence effect is well eliminated with subtraction between with and without multi-THz pulse if the QWP is achromatic.

4. Conclusions

We demonstrated the characterization of time-domain waveform of a multi-THz pulse as vector quantities. The POMEOS technique was successfully extended to the multi-THz region with a broadband gate pulse shorter than 15 fs. The fitting function in POMEOS was modified with an offset term and its validity was confirmed through numerical simulations. Owing to the fast polarization modulation, we demonstrated high accuracy and precision of approximately 1 mrad with only 3 s of data accumulation. Our method evaluates a multi-THz electric field vector, which can be applied to ultrafast time-resolved measurements of the chirality of molecules, metamaterials, or off-diagonal response functions in magnetic systems away from equilibrium. Our method can also be important in Floquet engineering using a designed complex vector field, such as counter-rotating bicircular polarized light [1416] by utilizing the vector field-shaping technique [38].

Funding

Precursory Research for Embryonic Science and Technology (JPMJPR2006); Ministry of Education, Culture, Sports, Science and Technology (JPMXS0118068681).

Disclosures

The authors declare no conflicts of interest.

Data availability

The data for this study are available from the corresponding authors upon request.

Supplemental document

See Supplement 1 for supporting content.

References

1. R. Huber, A. Brodschelm, F. Tauser, et al., “Generation and field-resolved detection of femtosecond electromagnetic pulses tunable up to 41 THz,” Appl. Phys. Lett. 76(22), 3191–3193 (2000). [CrossRef]  

2. C. Kübler, R. Huber, S. Tübel, et al., “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]  

3. M. Ashida, “Ultra-broadband terahertz wave detection using photoconductive antenna,” Jpn. J. Appl. Phys. 47(10), 8221–8225 (2008). [CrossRef]  

4. 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), 2767–2769 (2008). [CrossRef]  

5. M. Knorr, J. Raab, M. Tauer, et al., “Phase-locked multi-terahertz electric fields exceeding 13 MV/cm at a 190 kHz repetition rate,” Opt. Lett. 42(21), 4367–4370 (2017). [CrossRef]  

6. T. Yamakawa, N. Sono, T. Kitao, et al., “Long-term stabilization of carrier envelope phases of mid-infrared pulses for the precise detection of phase-sensitive responses to electromagnetic waves,” AIP Adv. 10(2), 025311 (2020). [CrossRef]  

7. N. Kanda, N. Ishii, J. Itatani, et al., “Optical parametric amplification of phase-stable terahertz-to-mid-infrared pulses studied in the time domain,” Opt. Express 29(3), 3479 (2021). [CrossRef]  

8. N. Kanda, Y. Murotani, T. Matsuda, et al., “Tracking Ultrafast Change of Multiterahertz Broadband Response Functions in a Photoexcited Dirac Semimetal Cd3As2 Thin Film,” Nano Lett. 22(6), 2358–2364 (2022). [CrossRef]  

9. Y. Murotani, N. Kanda, T. N. Ikeda, et al., “Stimulated Rayleigh Scattering Enhanced by a Longitudinal Plasma Mode in a Periodically Driven Dirac Semimetal Cd3As2,” Phys. Rev. Lett. 129(20), 207402 (2022). [CrossRef]  

10. T. Oka and H. Aoki, “Floquet theory of photo-induced topological phase transitions: application to graphene,” Phys. Rev. B 79(8), 081406 (2009). [CrossRef]  

11. J. W. McIver, B. Schulte, F. U. Stein, et al., “Light-induced anomalous Hall effect in graphene,” Nat. Phys. 16(1), 38–41 (2020). [CrossRef]  

12. R. Wang, B. Wang, R. Shen, et al., “Floquet Weyl semimetal induced by off-resonant light,” Europhys. Lett. 105(1), 17004 (2014). [CrossRef]  

13. Y. Murotani, N. Kanda, T. Fujimoto, et al., “Disentangling the Competing Mechanisms of Light-Induced Anomalous Hall Conductivity in Three-Dimensional Dirac Semimetal,” Phys. Rev. Lett. 131(9), 096901 (2023). [CrossRef]  

14. T. V. Trevisan, P. V. Arribi, O. Heinonen, et al., “Bicircular Light Floquet Engineering of Magnetic Symmetry and Topology and Its Application to the Dirac Semimetal Cd3As2,” Phys. Rev. Lett. 128(6), 066602 (2022). [CrossRef]  

15. Y. Ikeda, S. Kitamura, and T. Morimoto, “Floquet engineering of electric polarization with two-frequency drive,” Prog. Theor. Exp. Phys. 2022(4), 04A101 (2022). [CrossRef]  

16. Y. Ikeda, S. Kitamura, and T. Morimoto, “Photocurrent induced by a bicircular light drive in centrosymmetric systems,” Phys. Rev. Lett. 131(9), 096301 (2023). [CrossRef]  

17. D. M. Juraschek, M. Fechner, A. V. Balatsky, et al., “Dynamical multiferroicity,” Phys. Rev. Mater. 1(1), 014401 (2017). [CrossRef]  

18. D. Shin, H. Hübener, U. D. Giovannini, et al., “Phonon-driven spin-Floquet magneto-valleytronics in MoS2,” Nat. Commun. 9, 638 (2018).

19. H. Zhu, J. Yi, M.-Y. Li, et al., “Observation of chiral phonons,” Science 359(6375), 579–582 (2018). [CrossRef]  

20. K. Ishito, H. Mao, Y. Kousaka, et al., “Truly chiral phonons in α-HgS,” Nat. Phys. 19(1), 35–39 (2023). [CrossRef]  

21. L. A. Nafie, M. Diem, and D. W. Vidrine, “Fourier transform infrared vibrational circular dichroism,” J. Am. Chem. Soc. 101(2), 496–498 (1979). [CrossRef]  

22. K. Ohgushi, Y. Okamoto, T. Ogasawara, et al., “Magnetic, Optical, and Magnetooptical Properties of Spinel-Type ACr2X4 (A = Mn, Fe, Co, Cu, Zn, Cd; X = O, S, Se),” J. Phys. Soc. Jpn. 77(3), 034713 (2008). [CrossRef]  

23. F. Miyamaru, T. Kondo, T. Nagashima, et al., “Large polarization change in two-dimensional metallic photonic crystals in subterahertz region,” Appl. Phys. Lett. 82(16), 2568–2570 (2003). [CrossRef]  

24. N. Kanda, K. Konishi, and M. Kuwata-Gonokami, “Terahertz wave polarization rotation with double layered metal grating of complimentary chiral patterns,” Opt. Express 15(18), 11117 (2007). [CrossRef]  

25. T. Matsuda, N. Kanda, T. Higo, et al., “Room-Temperature Terahertz Anomalous Hall Effect in Weyl Antiferromagnet Mn3Sn Thin Films,” Nat. Commun. 11(1), 909 (2020). [CrossRef]  

26. C. M. Morris, R. Valdés Aguilar, A. V. Stier, et al., “Polarization modulation time-domain terahertz polarimetry,” Opt. Express 20(11), 12303–12317 (2012). [CrossRef]  

27. E. Castro-Camus, J. Lloyd-Hughes, M. B. Johnston, et al., “Polarization-sensitive terahertz detection by multicontact photoconductive receivers,” Appl. Phys. Lett. 86(25), 254102 (2005). [CrossRef]  

28. A. Hussain and R. Andrews, “Ultrabroadband polarization analysis of terahertz pulses,” Opt. Express 16(10), 7251–7257 (2008). [CrossRef]  

29. Z. Lü, D. Zhang, C. Meng, et al., “Polarization-sensitive air-biased-coherent-detection for terahertz wave,” Appl. Phys. Lett. 101(8), 081119 (2012). [CrossRef]  

30. J. Zhang, “Polarization-dependent study of THz air-biased coherent detection,” Opt. Lett. 39(14), 4096–4099 (2014). [CrossRef]  

31. N. Nemoto, T. Higuchi, N. Kanda, et al., “Highly precise and accurate terahertz polarization measurements based on electro-optic sampling with polarization modulation of probe pulses,” Opt. Express 22(15), 17915–17929 (2014). [CrossRef]  

32. N. Yasumatsu, A. Kasatani, K. Oguchi, et al., “High-speed terahertz time-domain polarimeter based on an electro-optic modulation technique,” Appl. Phys. Express 7(9), 092401 (2014). [CrossRef]  

33. M. Nakagawa, M. Okano, and S. Watanabe, “Polarization-sensitive terahertz time-domain spectroscopy system without mechanical moving parts,” Opt. Express 30(16), 29421–29434 (2022). [CrossRef]  

34. C.-H. Lu, Y.-J. Tsou, H.-Y. Chen, et al., “Generation of intense supercontinuum in condensed media,” Optica 1(6), 400–406 (2014). [CrossRef]  

35. N. Ishii, P. Xia, T. Kanai, et al., “Optical parametric amplification of carrier-envelope phase-stabilized mid-infrared pulses generated by intra-pulse difference frequency generation,” Opt. Express 27(8), 11447–11454 (2019). [CrossRef]  

36. K. Oguchi, H. Iwasaki, M. Okano, et al., “Polarization-sensitive electro-optic detection of terahertz wave using three different types of crystal symmetry: Toward broadband polarization spectroscopy,” Appl. Phys. Lett. 108(1), 011105 (2016). [CrossRef]  

37. N. Yasumatsu and S. Watanabe, “Robustness of electric field vector sensing of electromagnetic waves by analyzing crystal angle dependence of the electro-optic effect,” J. Opt. Soc. Am. B 30(11), 2940–2951 (2013). [CrossRef]  

38. M. Sato, T. Higuchi, N. Kanda, et al., “Terahertz polarization pulse shaping with arbitrary field control,” Nat. Photonics 7(9), 724–731 (2013). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental Document

Data availability

The data for this study are available from the corresponding authors upon 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. (a) Schematic of POMEOS setup. The orientations of the components are shown in the inset. (b) PEM phase and polarization of gate pulse for each pulse. Because of the relationship between modulation frequency and repetition rate, aliasing occurs. (c, d): An example of raw data of BPD signal (c) and PEM phase (d) for each laser pulse. Blue: chopper closed, Red: chopper opened. (e) Corresponding data as functions of PEM phase. Dot: measured data, circle: interpolated data. (f) Subtracted signal where the effect of residual birefringence is eliminated. Red circle: measured, black line: fitted. (g) Experimental result for a monocycle, broadband multi-THz pulse. (h) The power spectra for the waveform in (g). Red and blue show x and y components.
Fig. 2.
Fig. 2. Results of the waveforms of electric field vectors. (a − f) Linearly polarized multi-THz pulses with 30° rotation step. (g, h) Circularly polarized pulse. Red arrows show the direction of the rotation of the field. (i) Power spectra of circular polarization in (g). Red and blue show right- and left-circular components, respectively.
Fig. 3.
Fig. 3. Precision and accuracy. (a) Histogram of the measured azimuthal angles. The distribution is well fitted by Gaussian function (blue dashed line). (b) Accumulation pulse number dependence of azimuthal angle precision. (c − e) The relationship between measured azimuthal angles and known WGP angles in different angle ranges.
Fig. 4.
Fig. 4. (a) An experimental result for a linearly polarized pulse. Red: ${E_x}$, blue: ${E_y}$, green: offset term, black shaded: residue in fitting (right axis). (b − f) Results of numerical simulations. (b) Ideal case with achromatic PEM and QWP, (c) frequency-dependent PEM and achromatic QWP, and (d) frequency-dependent PEM and QWP. (e) and (f) are calculated for a chirped gate pulse, with the other parameters the same as (c) and (d), respectively.

Equations (5)

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

S ( t ) = E THz cos ( δ 0 sin ( 2 π f PEM t ) + θ ) .
[ E x g , out ( ω , τ ) E y g , out ( ω , τ ) ] = [ E x g ( ω ) E y g ( ω ) ] + i ω b d Ω   ( e i Ω τ [ E y ( Ω ) E x ( Ω ) E x ( Ω ) E y ( Ω ) ] [ E x g ( ω Ω ) E y g ( ω Ω ) ] + e i Ω τ [ E y ( Ω ) E x ( Ω ) E x ( Ω ) E y ( Ω ) ] [ E x g ( ω + Ω ) E y g ( ω + Ω ) ] ) ,
[ E x g ( ω ; φ ) E y g ( ω ; φ ) ] = [ 1 0 0 e i Δ ( ω ) ] R ( π 4 ) [ 1 0 0 e i δ 0 ( ω ) sin φ ] R ( π 4 ) [ 1 0 ] ,
[ E x g , WP ( ω , τ ; φ ) E y g , WP ( ω , τ ; φ ) ] = R ( π 4 ) [ 1 0 0 e i Δ ( ω ) ] R ( π 4 ) [ E x g , out ( ω , τ ; φ ) E y g , out ( ω , τ ; φ ) ] .
S ( τ ; φ ) = d ω ( | E x g , WP ( ω , τ ; φ ) | 2 | E y g , WP ( ω , τ ; φ ) | 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.