ACS Publications. Most Trusted. Most Cited. Most Read
Hot-Electron Dynamics in a Semiconductor Nanowire under Intense THz Excitation
My Activity
  • Open Access
Article

Hot-Electron Dynamics in a Semiconductor Nanowire under Intense THz Excitation
Click to copy article linkArticle link copied!

  • Andrei Luferau*
    Andrei Luferau
    Institute of Ion Beam Physics and Materials Research, Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01328, Germany
    Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, Germany
    *Email: [email protected]
  • Maximilian Obst
    Maximilian Obst
    Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, Germany
    Würzburg-Dresden Cluster of Excellence, EXC 2147 (ct.qmat), Dresden 01062, Germany
  • Stephan Winnerl
    Stephan Winnerl
    Institute of Ion Beam Physics and Materials Research, Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01328, Germany
  • Alexej Pashkin*
    Alexej Pashkin
    Institute of Ion Beam Physics and Materials Research, Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01328, Germany
    *Email: [email protected]
  • Susanne C. Kehr
    Susanne C. Kehr
    Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, Germany
    Würzburg-Dresden Cluster of Excellence, EXC 2147 (ct.qmat), Dresden 01062, Germany
  • Emmanouil Dimakis
    Emmanouil Dimakis
    Institute of Ion Beam Physics and Materials Research, Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01328, Germany
  • Felix G. Kaps
    Felix G. Kaps
    Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, Germany
    Würzburg-Dresden Cluster of Excellence, EXC 2147 (ct.qmat), Dresden 01062, Germany
  • Osama Hatem
    Osama Hatem
    Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, Germany
    Würzburg-Dresden Cluster of Excellence, EXC 2147 (ct.qmat), Dresden 01062, Germany
    Department of Engineering Physics and Mathematics, Faculty of Engineering, Tanta University, Tanta 31511, Egypt
    More by Osama Hatem
  • Kalliopi Mavridou
    Kalliopi Mavridou
    Institute of Ion Beam Physics and Materials Research, Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01328, Germany
    Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, Germany
  • Lukas M. Eng
    Lukas M. Eng
    Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, Germany
    Würzburg-Dresden Cluster of Excellence, EXC 2147 (ct.qmat), Dresden 01062, Germany
    More by Lukas M. Eng
  • Manfred Helm
    Manfred Helm
    Institute of Ion Beam Physics and Materials Research, Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01328, Germany
    Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, Germany
    More by Manfred Helm
Open PDFSupporting Information (1)

ACS Photonics

Cite this: ACS Photonics 2024, 11, 8, 3123–3130
Click to copy citationCitation copied!
https://doi.org/10.1021/acsphotonics.4c00433
Published August 9, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

We report terahertz (THz)-pump/mid-infrared probe near-field studies on Si-doped GaAs–InGaAs core–shell nanowires utilizing THz radiation from the free-electron laser FELBE. Upon THz excitation of free carriers, we observe a red shift of the plasma resonance in both amplitude and phase spectra, which we attribute to the heating of electrons in the conduction band. The simulation of heated electron distributions anticipates a significant electron population in both the L- and X-valleys. The two-temperature model is utilized for quantitative analysis of the dynamics of the electron gas temperature under THz pumping at various power levels.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2024 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

High-quality epitaxial nanowires (NWs) based on III–V semiconductors offer the possibility to fabricate ultrafast optical devices. In particular, single NWs can serve as sensitive elements of terahertz (THz) radiation detectors that operate at room temperature. (1,2) Contactless investigation of the average charge carrier concentration and mobility in large ensembles of NWs is possible using THz time-domain spectroscopy. (3) The local investigation of these properties on individual NWs can be carried out by scattering-type scanning near-field optical microscopy (s-SNOM). (4) This technique offers a spatial resolution far beyond the diffraction limit, and in the case of far-infrared studies, it can be extended beyond λ/4600. (5,6) Utilizing near-infrared (NIR) excitation on (typically intrinsic) NWs, charge-carrier lifetimes in NW ensembles become accessible in far-field NIR-pump/THz-probe experiments. (7,8) Several studies demonstrated the feasibility of integrating the NIR-pump/THz-probe technique with near-field microscopy for the examination of III–V semiconductors. (9−12) The NIR pump served as interband excitation and enabled investigating the local recombination dynamics of photogenerated carriers in individual NWs. However, for some important applications of NWs such as field-effect transistors (13−15) or THz detectors, (1,2) it is important to study the response of doped NWs to intense THz radiation that does not change the carrier density but rather heats up the carrier system via intraband absorption.
Here, we report on THz-pump/mid-infrared (MIR) probe s-SNOM studies on highly doped GaAs/InGaAs core–shell NWs utilizing the intense narrowband THz radiation from the free-electron laser (FEL) FELBE (16) at the Helmholtz-Zentrum Dresden-Rossendorf. The analysis of the near-field spectra enables us to obtain the electron temperature as a function of the pump–probe delay time. Our results quantify the dynamics of hot electrons in the conduction band of the InGaAs shell and enable us to quantify the electron–phonon coupling. Furthermore, we gain insights into the electron heating and cooling dynamics by simulating it with the two-temperature model.

Results and Discussion

Click to copy section linkSection link copied!

Near-Field MIR Spectroscopy of the Nanowires

The samples under study are Si-doped GaAs/InGaAs core–shell NWs grown by molecular beam epitaxy. They consist of a 25 nm-thick GaAs core and an 80 nm-thick In0.44Ga0.56As shell that is homogeneously n-doped with a nominal Si-concentration of about 9 × 1018 cm–3. (17,18) For s-SNOM studies, these NWs are transferred onto a (100)Si substrate and dispersed randomly across it.
Our experiments are carried out with an s-SNOM setup from Neaspec GmbH equipped with a nanoscale Fourier transform infrared (nanoFTIR) module, including a broadband MIR difference-frequency generation (DFG) source (5–15 μm; 20–60 THz; Pavg ∼0.2 mW; 78 MHz; t < 100 fs) used as a probe. (19) Radiation is focused by an off-axis parabolic mirror (numerical aperture: 0.46, focal length: 11 mm) onto a metallized atomic force microscopy (AFM) tip that operates in tapping mode with an oscillation amplitude of about 100 nm. Tip–sample near-field interaction modifies the backscattered radiation, which is detected by a photoconductive mercury cadmium telluride (MCT) detector. To separate the near-field signal from the far-field signal, the detector signal is demodulated by a lock-in amplifier at the second harmonic of the AFM cantilever tapping frequency Ω ≈ 250 kHz. (20) The tip and the sample are located in one of the arms of an asymmetric Michelson interferometer of the nanoFTIR module (Figure 1a). Here, the tip-scattered near-field light optically interferes with the split MIR beam from the reference arm, and the detector signal is recorded as a function of the reference mirror position. (21) Finally, the obtained interferograms are Fourier transformed to obtain spectra of the near-field scattering amplitude s(ω) and phase ϕ(ω).

Figure 1

Figure 1. (a) Sketch of a time-resolved near-field spectroscopy setup based on s-SNOM. A Michelson interferometer enables MIR near-field spectroscopy (nanoFTIR) during THz pumping of the sample. (b) Topography of the GaAs/InGaAs core–shell NW recorded by AFM. The tip position chosen for spectrally resolved scans is highlighted with a red triangle. (c) Near-field amplitude s(ω) and phase ϕ(ω) spectra of the doped GaAs/InGaAs core–shell NW normalized to the response of Si along with the results of two-parameter fit based on the point–dipole and the Drude models.

In Figure 1b, the AFM topography image of such a NW under study is shown, highlighting the chosen tip position for spectrally resolved scans with a red triangle. The orientation of NWs with respect to the incident angle of the laser shows no significant impact on the spectral response, while positioning the tip at the center of the NW prevents geometry-related artifacts like interference or shadowing.
A point–dipole model incorporating the frequency-dependent infrared permittivity ε(ω) of the NWs is used to estimate the near-field response of the samples. (22,23) In this model, the field Escat scattered by the tip is approximated by a dipolar near-field interaction between the tip apex and the sample surface, resulting in
Escatα0(1+rp)21α0β16π(a+h)3Einc
(1)
where α0 = 4πa3 is the polarizability of the tip with a radius a induced by the incident field Einc, rp is the Fresnel reflection coefficient for p-polarized light, h is the tip apex–sample separation, and β = (ε – 1)/(ε + 1). Consideration of higher harmonic demodulation involves incorporating a sinusoidal tip oscillation h(t). The subsequent Fourier transformation of Escat[h(t)] results in Ensneiϕn, where sn and ϕn represent the amplitude and phase of the scattered electric field at demodulation order n, respectively.
The frequency dependence of the permittivity ε(ω) in a doped semiconductor can be described by the Drude model as follows
εNW(ω)=εopticωpl2εopticω2+iωγel
(2)
where εoptic is the high-frequency permittivity, and the second term in the equation describes the response of free charge carriers with the plasma frequency ωpl2 = ne2/(m0εoptic) and the plasmonic damping γel = e/(μm*), where e is the elementary charge, n is the concentration of the charge carriers, and m* and μ are their effective mass and mobility, respectively. Here, we neglect the contribution from the polar phonon resonance since it is located well below the frequency range addressed in this study. Since the InGaAs shell is much thicker than the GaAs core, we neglect the core–shell structure and estimate the NW response as the plasmonic response of the heavily doped shell alone.
The unpumped near-field MIR response of the NW acquired in the standard nanoFTIR mode at demodulation order n = 2 is shown in Figure 1c. Both the amplitude spectrum s(ω) = sNW(ω)/sSi(ω) and the phase spectrum ϕ(ω) = ϕNW(ω) – ϕSi(ω) are normalized to the response of the Si substrate, assuming its permittivity to be constant for the photon energy range of the MIR probe. (24) Due to the interference between the tip-scattered radiation Escat and the reference beam, which is approximately identical to the incident on the tip field Einc, the MCT detector measures the homodyne detection intensity I|Escat||Einc|ei(ϕscatϕinc) rather than just the intensity of the scattered probe I|Escat|2. (25) Thus, normalizing the NW spectra to the Si reference eliminates the unknown parameters of the amplitude and phase of incident field Einc. In accordance with eq 1, this simplifies modeling the plasmonic response to just two physical parameters. The fitting of the experimental data provides us with the plasma frequency ωpl0 = (1050 ± 10) cm–1 and the damping γel0= (260 ± 20) cm–1. As seen from Figure 1c, the model can reliably reproduce the experimental amplitude and phase spectra.
Samples under study are grown in such a way that Si dopants are incorporated into InGaAs as a donor. (17) Due to the heavy n-doping of the sample, the observed plasmonic resonance is attributed to the response of electrons in the conduction band. The InGaAs Γ-valley conduction band is nonparabolic and can be approximately described by the equation εk (1 + αεk) = ℏ2k2/(2mΓ), (26) with the Γ-point effective mass mΓ and the nonparabolicity parameter α = 1/Eg, (27) where Eg is the band gap. The nonparabolicity leads to a variation of the effective mass with electron density m*(n). By solving the equation ne2/[ε0εopticm*(n)] = ωpl02, we obtain the unpumped electron density n0 = (8.5 ± 0.2) × 1018 cm–3 and the distribution function of electrons with a chemical potential of εF = 265 meV at 300 K. The value of the average effective mass m0* corresponding to the electron density n0 equates to 0.06me. Knowing the effective mass m0* also allows us to convert the fitted value of damping γel0 to the mobility μ0= (600 ± 50) cm2 V–1 s–1.

THz-Pump/MIR-Probe Experiment

For THz-pump/MIR-probe s-SNOM studies, we utilize the intense narrowband THz radiation from FELBE (16) as an intraband pump. Its pulses have a repetition rate of 13 MHz and a duration at full width at half-maximum (fwhm) of 2.85 ps at the chosen wavelength of λ = 25 μm (12 THz). This is the shortest wavelength that can be efficiently suppressed by a ZnSe high-pass filter placed before the detection scheme to avoid saturation of the detector. The temporal profile of the pulse and the duration were obtained from the measured FEL spectrum using the model in ref (28). Unless otherwise mentioned, the data shown in the paper are obtained with an average excitation power of Pavg = 6 mW (fluence of ≈12 μJ/cm2). The FEL pump beam is guided to the parabolic mirror with a focal length of 11 mm that focuses the beam, having a diameter of ≈4 mm, onto the tip apex into a diffraction-limited spot of ≈70 μm, resulting in an applied electric field of ≈80 kV/cm. The DFG probe source is locked to the sixth harmonic of the FEL repetition rate and synchronized with the FEL pulse train. The time delay between pump and probe pulses is varied by an optical delay line.
The 6-fold difference in the repetition rates of the FEL and the DFG source leads to an undesirable situation when the pump pulse excites the sample, and only every sixth probe pulse measures the real response of the sample to this excitation. This can be overcome technically during the measurement process via the sideband demodulation technique by introducing an additional modulation of the pump beam to directly detect the photoinduced change in the interferogram. (29) Since the FEL operates in a pulsed mode, our pump is originally modulated at the FEL repetition rate of 13 MHz, which we use as a carrier frequency, while the second harmonic of the AFM cantilever tapping frequency 2Ω ≈ 500 kHz is utilized as a sideband (see Supporting Information). Nonetheless, the 2 MHz cutoff frequency of the employed MCT detector diminishes the efficiency of the double demodulation technique. Substituting the employed detector with a faster one, however, results in compromised sensitivity within the probed range. To overcome this issue, we further suggest a data processing approach that enables the extraction of the relevant excited probe pulses from data acquired without additional modulation.
As previously mentioned, when the pump pulse excites the sample, only every sixth probe pulse measures the actual response of the sample to this excitation. The five following probe pulses are idle and measured the response of the sample in an unpumped state until the next pump pulse arrives. Assuming that the detector equally averages intensities of all the received probe pulses, we can define an average-measured intensity Imix as follows
Imix=Ipumped+5×Iunpumped6
(3)
where Ipumped and Iunpumped are the intensities of the probe pulses scattered from the sample in the excited and unexcited states, respectively. In the case of interferometric studies, eq 3 is also valid for each point of the interferogram and, therefore, is applicable to describe the mixing of the pumped and unpumped interference patterns as a whole. Thus, given the interferogram of the sample in the unexcited state, we can extract the pumped interferogram from the measured mixed interference pattern Imix. Considering the linearity of the Fourier transform, the same procedure can be applied to extract complex pumped scattering spectra from the measured mixed spectra. The extracted pumped spectra exhibit a superior signal-to-noise ratio (SNR) compared to the results obtained via side-demodulation detection (see Supporting Information) and are used for the presentation of the results in the following.
Figure 2a,b shows the near-field amplitude s(ω) and phase ϕ(ω) spectra of the InGaAs NW obtained with and without THz pumping, extracted according to eq 3, and normalized to the response of Si. Despite the increased noise in the extracted spectra, the implemented fitting model works well and additionally serves as a guide for the eye to observe a red shift of the NW plasma resonance upon the excitation. To trace the full dynamics of the resonance shift, we perform a series of spectroscopic scans with a 2 ps step of the optical delay line. Figure 2c,d depicts the results in the form of a color map, where every line represents extracted and normalized to Si near-field spectra obtained for different time delays between the THz-pump and broadband MIR-probe counted from an arbitrary time zero. We acknowledge that the time zero is set to the maximum of the FEL pulse (Figure 4b) as we do not have a direct method to determine it due to the noninstantaneous nature and relatively long rise time of our pump pulse. One can see a red shift in both the amplitude and phase of the near-field plasmonic response and recovery on the same time scale. To quantify the observed shift of the plasma resonance, all the spectra are fitted with the point–dipole model, and the fitted values of the plasma frequency ωpl for different time delays are plotted in Figure 2e. Since the intraband pump does not excite new electrons into the conduction band, the carrier density n0=(8.5±0.2)×1018cm3 should remain constant. Thus, the time evolution of the plasma frequency ωpl can be converted to the time evolution of the effective mass m* (Figure 2f), which quantitatively illustrates the dynamics of the electron gas heating, reaching values of up to 0.09me for the 6 mW THz pump.

Figure 2

Figure 2. (a,b) Near-field amplitude s(ω) and phase ϕ(ω) spectra of the doped GaAs/InGaAs core–shell NW obtained with (red) and without (blue) THz pumping (6 mW) and normalized to the response of Si. The spectra are extracted from experimental data according to eq 3 and plotted along with the results of the two-parameter fit based on the point–dipole and the Drude models. (c,d) Color maps illustrating the evolution of the near-field amplitude s(ω) and the phase ϕ(ω) spectra of the doped InGaAs NW upon THz photoexcitation (6 mW), normalized to the response of Si. Every line represents normalized near-field spectra obtained for different time delays between the THz-pump and broadband MIR probe. (e) Fitting parameter of the plasma frequency ωpl as a function of pump–probe delay time. (f) Time evolution of the effective mass m* of electron gas upon intraband THz pumping.

Figure 3

Figure 3. Simulation results: (a) In0.44Ga0.56As band structure scheme depicting the conduction band valleys relevant to the experiment [ (34). The blue dashed line is a parabolic approximation of the nonparabolic Γ-valley. (b,c) Normalized electron distributions for 300 and 1500 K. (d) Calculated fractions of electrons of each conduction band valley versus temperature. (e) Dependence of the total effective mass on temperature. The dashed line represents the simulation neglecting the impact of side valley transfer.

Figure 4

Figure 4. (a) Temporal evolution of the electron-gas temperature upon THz-pumping of various powers. (b) Simulation of the temporal evolution of the electron temperature [the same axis as part (a)] based on the two-temperature model for three peak electric-field amplitudes of the FEL radiation inside the NW. The gray dashed line represents the normalized intensity profile of the FEL pulse.

Dynamics of Electron Temperature

Several theoretical (30,31) and experimental (32,33) studies are dedicated to the heating of free electrons by MIR radiation. They associate the increase in effective mass of highly doped n-GaAs both with the nonparabolicity of the Γ-valley and, at high temperatures, with the transfer of electrons to the heavier mass l-valley, while the small contribution of X-valley electrons is typically neglected. In the case of the In0.44Ga0.56As alloy, a schematic band diagram of which is depicted in Figure 3a, the energy separation between the L- and X-minima almost disappears. (34) Therefore, in addition to the nonparabolicity (the parabolic approximation of the Γ-valley is shown in Figure 3a as a dashed line), we have to consider the redistribution of heated electrons between the three conduction valleys. By taking into account the boundary condition of carrier density conservation, we calculate electron distributions for various temperatures of the heated electron gas. Figure 3b,c depicts the calculated electron distributions for 300 and 1500 K, respectively, showing the change of the valley populations and the Fermi–Dirac distribution.
To quantify the repopulation of hot electrons and its effect on the effective mass, we introduce the value of the fraction of electrons in the ith valley, fi, normalized to the total density of electrons n0. Figure 3d shows the calculated temperature dependence of the fraction of electrons in each valley. Initially, all the electrons stay within the Γ-valley, and in order to transfer just 1% of them to the side valleys, the electron gas must be heated above 500 K. As the heating increases, the tail of the distribution widens enough for faster repopulation of electrons, which are evenly distributed over two side valleys up to 1000 K. At higher temperatures, the fraction of electrons settled in the X-valley becomes greater than that for the l-valley due to the fact that the density-of-states effective mass of X-valley electrons is somewhat larger than that of l-valley electrons. (34) The contribution of electrons from each valley to the total effective mass is estimated as for conductivity problems
1m*=fXmX*+fLmL*+fΓmΓ*(nΓ)
(4)
where fi is the fraction of electrons in the corresponding valley, mΓ*(nΓ) is the effective mass in the Γ-valley considering the nonparabolicity, and mX* and mL* are optical effective masses for anisotropic X- and L-minima, respectively, characterized by longitudinal ml and transverse mt effective masses and defined as (m)−1=(ml–1+2mt–1)/3. As a result, we obtain a simulated dependence of the total effective mass m*(T) on the temperature shown in Figure 3e. To emphasize the influence of side-valley electrons, we additionally simulated the temperature dependence of the total effective mass m* considering that all electrons stay within the nonparabolic Γ-valley (Figure 3e, dashed line). At temperatures below 500 K, both curves are equal, and the effective mass rises because of the Γ-valley nonparabolicity, while further heating leads to a significant increase of the total effective mass due to the transfer of electrons to the side valleys.
Knowing the temperature dependence of the total effective mass m*(T) enables us to extract the time evolution of the electron temperature T(t) from the experimental results on the temporal changes in the effective mass m*(t). Figure 4a presents the dynamics of electron temperature T(t) upon THz-pumping of various powers. The data points related to the excitation power of 6 mW are extracted directly from the data set depicted in Figure 2f, replacing each mass value m* with the associated temperature T according to the simulated m*(T) dependence. The data points corresponding to other power excitations are acquired in the same manner, and detailed outcomes of the related pump–probe experiments are available in the Supporting Information. In Figure 4a, all data points within the temperature range below 500 K exhibit significant error bars, attributed to error propagation resulting from the relatively flat segment of the simulated m*(T) dependence. At higher temperatures, the slope of the m*(T) is larger, which makes it possible to track temperature changes more precisely. However, excessively high temperatures are also challenging to track. Exciting the sample with a power of 27 mW enables us to discern temperatures up to approximately 2000 K as the upper detection limit. Excessive heating causes the plasma resonance to be shifted beyond our MIR probing range, and the exact temperature of the electron gas remains unknown.

Two-Temperature Modeling

For interpretation of the temporal response shown in Figure 4a, we performed a modeling of the electron temperature dynamics using the two-temperature model (35)
Ce(Te)dTedt=PFEL(t)g(TeTl)
(5)
CldTldt=g(TeTl)
(6)
where Te and Tl are the temperatures of the electrons and the lattice, respectively; Ce and Cl are the specific heats of the electron and the lattice subsystems, respectively; g is the electron–phonon coupling constant; and PFEL(t) is the FEL power density absorbed by the electrons. The latter function can be written as
PFEL(t)=σ1(ωFEL,Te(t))EFEL2(t)
(7)
where σ1FEL,Te) = ωFELIm[εNWFEL,Te)]ε0 is the optical conductivity of the NW at the FEL frequency ωFEL estimated from the Drude model parameters in eq 2, and EFEL is the local electric field in the NW under the AFM tip induced by the FEL pumping. Since the effective mass depends on the transient electronic temperature Te(t), as depicted in Figure 3e, the plasma frequency ωpl and the optical conductivity in the Drude model also vary with time in our simulation. In contrast to metals, where the Fermi energy is large, the condition kBT ≪ εF is not fulfilled in our case of a doped semiconductor. Therefore, the usual assumption that the electronic specific heat Ce is proportional to Te does not hold. In our modeling, we used the Ce(Te) dependence calculated for the actual band structure of In0.44Ga0.56As (see Supporting Information for details).
Figure 4b shows the simulation results for three peak electric-field amplitudes of the FEL radiation inside the NW that properly scale with the FEL powers used in the experiment. We obtain reasonably good agreement with the experimentally measured temperature dynamics by varying only two parameters: the electron–phonon coupling g and the conversion factor between the FEL power and the local electric field EFEL2. The simulation confirms the quicker rise and slower decay of the electron temperature and accurately reproduces the substantial power-dependent extension in the duration the system stays in a heated state compared with the FEL pulse duration. This observation allows us to exclude the effect of band dispersion modification, which would cause a nearly instantaneous response lasting for just over the duration of the FEL pulse.
The two-temperature model captures the peak temperature levels for two lower excitation fields, but the simulated temperature for the highest field of 75 kV/cm seems to be overestimated. Although the experimental data also reveal that the electron temperature exceeds 2000 K around the peak for an FEL power of 27 mW, the simulation shows even more extreme heating with a maximal Te of 3700 K for a maximum field strength of 170 kV/cm that is reachable in our experiments. According to our calculations, the heating should transfer up to 87% of electrons to the side valleys.
First, let us discuss electron–phonon coupling. The coupling coefficient used in the simulation is independent of Te and equals g ≈ 1014 J/(m3 K s). It results in temperature relaxation times of 2–5 ps for the electronic temperature between 500 and 2000 K. This agrees with theoretical results predicting a nearly temperature-independent electron cooling time around 1 ps for electronic temperatures up to 500 K. (36) We would like to point out that our assumption of the temperature-independent electron–phonon coupling does not take into account that the LO phonon emission rate for the valley electrons is much higher than that for the Γ-valley. (37) This would lead to an increase in the electron–phonon coupling g at high electronic temperatures, consequently accelerating the cooling dynamics. However, this effect should not drastically affect the temperature dynamics in the range of electronic temperatures detected in our experiment.
Second, we compare the estimated peak FEL field in the focus of the parabolic mirror with the simulation results. Knowing the FEL pulse parameters and considering the Fresnel transmission coefficient of ≈25% for the p-polarized pump, we can estimate the peak FEL field inside the NW samples to be ≈15 kV/cm for an FEL power of 3 mW. Taking into account uncertainties in the experimental and modeling parameters, this value nearly matches the local electric field EFEL used in the simulations. This fact demonstrates that the pump-induced THz near-field under the SNOM tip does not have a superior effect on the sample heating compared to the focused far-field. At first glance, this result contradicts the near-field modeling predicting a few orders of magnitude enhancement of the near-field at the sample surface. (38,39) However, the normal electric field component inside the sample is reduced by a factor |ε| with respect to the field outside. In our case, the permittivity of the doped NW at the FEL frequency can be estimated from eq 2, giving |εNW| ≈ 60. Thus, it is possible that the tip enhancement of the near-field is compensated by the plasma screening inside the NW, resulting in a comparable effect on the sample heating from both near- and far-field.

Conclusions

Click to copy section linkSection link copied!

Our investigation of heavily doped GaAs/InGaAs core–shell NWs via THz-pump/MIR-probe s-SNOM provides valuable insights into the heating and cooling dynamics of electron gas in an individual NW. The localized plasmon resonance experiences a strong red shift under THz-FEL excitation. We have demonstrated that this behavior can be explained by the increase in the averaged effective electron mass caused by the hot-electron distribution that populates side valleys of the conduction band in InGaAs. Additionally, the simulation of heated electron distributions has allowed estimation of the temporal evolution of electron-gas temperatures under THz pumping. The observed features of the temporal response were accurately reproduced by the two-temperature model, yielding estimated values for both electron–phonon coupling and the local pump field. The results provide relaxation times that are consistent with theoretical predictions and suggest that the pump-induced THz near-field under the SNOM tip has a minimal impact on sample heating when compared to that of the focused far-field. Importantly, the THz-pump/MIR-probe s-SNOM has proven itself as a powerful tool for quantitatively exploring the nonlinear interaction of nanostructures with THz radiation, further broadening its applicability in the field of nanoscale science and technology.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsphotonics.4c00433.

  • Results of the sideband modulation technique along with the point–dipole model-based simulations, extended power-dependent results, and calculation of the specific heat of the electron gas (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
    • Maximilian Obst - Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, GermanyWürzburg-Dresden Cluster of Excellence, EXC 2147 (ct.qmat), Dresden 01062, GermanyOrcidhttps://orcid.org/0000-0003-0370-425X
    • Stephan Winnerl - Institute of Ion Beam Physics and Materials Research, Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01328, Germany
    • Susanne C. Kehr - Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, GermanyWürzburg-Dresden Cluster of Excellence, EXC 2147 (ct.qmat), Dresden 01062, GermanyOrcidhttps://orcid.org/0000-0002-3857-673X
    • Emmanouil Dimakis - Institute of Ion Beam Physics and Materials Research, Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01328, GermanyOrcidhttps://orcid.org/0000-0002-7546-0621
    • Felix G. Kaps - Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, GermanyWürzburg-Dresden Cluster of Excellence, EXC 2147 (ct.qmat), Dresden 01062, Germany
    • Osama Hatem - Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, GermanyWürzburg-Dresden Cluster of Excellence, EXC 2147 (ct.qmat), Dresden 01062, GermanyDepartment of Engineering Physics and Mathematics, Faculty of Engineering, Tanta University, Tanta 31511, EgyptOrcidhttps://orcid.org/0000-0003-4746-0963
    • Kalliopi Mavridou - Institute of Ion Beam Physics and Materials Research, Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01328, GermanyInstitut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, Germany
    • Lukas M. Eng - Institut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, GermanyWürzburg-Dresden Cluster of Excellence, EXC 2147 (ct.qmat), Dresden 01062, GermanyOrcidhttps://orcid.org/0000-0002-2484-4158
    • Manfred Helm - Institute of Ion Beam Physics and Materials Research, Helmholtz-Zentrum Dresden-Rossendorf, Dresden 01328, GermanyInstitut für Angewandte Physik, Technische Universität Dresden, Dresden 01062, Germany
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

The authors are grateful to J. Michael Klopf and the ELBE team for the operation of the FEL FELBE and for their dedicated support and to Thales de Oliveira and Xiaoxiao Sun for their experimental assistance at the Helmholtz-Zentrum Dresden-Rossendorf. A.L. expresses gratitude to Markus B. Raschke for insightful discussions on sideband demodulation technique conveyed through private communications. M.O., S.C.K., F.G.K., O.H., and L.M.E. acknowledge funding by the Bundesministerium für Bildung und Forschung (BMBF, Federal Ministry of Education and Research, Germany) grant nos. 05K19ODA, 05K19ODB, and 05K22ODA as well as the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) through project CRC1415 (ID: 417590517) and the Würzburg-Dresden Cluster of Excellence “ct.qmat” (EXC 2147, ID: 390858490).

References

Click to copy section linkSection link copied!

This article references 39 other publications.

  1. 1
    Peng, K.; Parkinson, P.; Fu, L.; Gao, Q.; Jiang, N.; Guo, Y.-N.; Wang, F.; Joyce, H. J.; Boland, J. L.; Tan, H. H.; Jagadish, C.; Johnston, M. B. Single nanowire photoconductive terahertz detectors. Nano Lett. 2015, 15 (1), 206210,  DOI: 10.1021/nl5033843
  2. 2
    Peng, K.; Jevtics, D.; Zhang, F.; Sterzl, S.; Damry, D. A.; Rothmann, M. U.; Guilhabert, B.; Strain, M. J.; Tan, H. H.; Herz, L. M.; Fu, L.; Dawson, M. D.; Hurtado, A.; Jagadish, C.; Johnston, M. B. Three-dimensional cross-nanowire networks recover full terahertz state. Science 2020, 368 (6490), 510513,  DOI: 10.1126/science.abb0924
  3. 3
    Parkinson, P.; Lloyd-Hughes, J.; Gao, Q.; Tan, H. H.; Jagadish, C.; Johnston, M. B.; Herz, L. M. Transient terahertz conductivity of GaAs nanowires. Nano Lett. 2007, 7 (7), 21622165,  DOI: 10.1021/nl071162x
  4. 4
    Stiegler, J. M.; Huber, A. J.; Diedenhofen, S. L.; Gomez Rivas, J.; Algra, R. E.; Bakkers, E. P. A. M.; Hillenbrand, R. Nanoscale free-carrier profiling of individual semiconductor nanowires by infrared near-field nanoscopy. Nano Lett. 2010, 10 (4), 13871392,  DOI: 10.1021/nl100145d
  5. 5
    Kuschewski, F.; von Ribbeck, H.-G.; Döring, J.; Winnerl, S.; Eng, L. M.; Kehr, S. C. Narrow-band near-field nanoscopy in the spectral range from 1.3 to 8.5 THz. Appl. Phys. Lett. 2016, 108, 11,  DOI: 10.1063/1.4943793
  6. 6
    Lang, D.; Balaghi, L.; Winnerl, S.; Schneider, H.; Hübner, R.; Kehr, S. C.; Eng, L. M.; Helm, M.; Dimakis, E.; Pashkin, A. Nonlinear plasmonic response of doped nanowires observed by infrared nanospectroscopy. Nanotechnology 2019, 30 (8), 084003,  DOI: 10.1088/1361-6528/aaf5a7
  7. 7
    Joyce, H. J.; Docherty, C. J.; Gao, Q.; Tan, H. H.; Jagadish, C.; Lloyd-Hughes, J.; Herz, L. M.; Johnston, M. B. Electronic properties of GaAs, InAs and InP nanowires studied by terahertz spectroscopy. Nanotechnology 2013, 24 (21), 214006,  DOI: 10.1088/0957-4484/24/21/214006
  8. 8
    Fotev, I.; Balaghi, L.; Schmidt, J.; Schneider, H.; Helm, M.; Dimakis, E.; Pashkin, A. Electron dynamics in In x Ga1–x As shells around GaAs nanowires probed by terahertz spectroscopy. Nanotechnology 2019, 30 (24), 244004,  DOI: 10.1088/1361-6528/ab0913
  9. 9
    Eisele, M.; Cocker, T. L.; Huber, M. A.; Plankl, M.; Viti, L.; Ercolani, D.; Sorba, L.; Vitiello, M. S.; Huber, R. Ultrafast multi-terahertz nano-spectroscopy with sub-cycle temporal resolution. Nat. Photonics 2014, 8 (11), 841845,  DOI: 10.1038/nphoton.2014.225
  10. 10
    Wagner, M.; McLeod, A. S.; Maddox, S. J.; Fei, Z.; Liu, M.; Averitt, R. D.; Fogler, M. M.; Bank, S. R.; Keilmann, F.; Basov, D. N. Ultrafast dynamics of surface plasmons in InAs by time-resolved infrared nanospectroscopy. Nano Lett. 2014, 14 (8), 45294534,  DOI: 10.1021/nl501558t
  11. 11
    Pizzuto, A.; Castro-Camus, E.; Wilson, W.; Choi, W.; Li, X.; Mittleman, D. M. Nonlocal time-resolved terahertz spectroscopy in the near field. ACS Photonics 2021, 8 (10), 29042911,  DOI: 10.1021/acsphotonics.1c01367
  12. 12
    Pushkarev, V.; Němec, H.; Paingad, V. C.; Maňák, J.; Jurka, V.; Novák, V.; Ostatnický, T.; Kužel, P. Charge Transport in Single-Crystalline GaAs Nanobars: Impact of Band Bending Revealed by Terahertz Spectroscopy. Adv. Funct. Mater. 2022, 32 (5), 2107403,  DOI: 10.1002/adfm.202107403
  13. 13
    Egard, M.; Johansson, S.; Johansson, A.-C.; Persson, K.-M.; Dey, A. W.; Borg, B. M.; Thelander, C.; Wernersson, L.-E.; Lind, E. Vertical InAs nanowire wrap gate transistors with ft > 7 GHz and fmax > 20 GHz. Nano Lett. 2010, 10 (3), 809812,  DOI: 10.1021/nl903125m
  14. 14
    Colinge, J.-P.; Lee, C.-W.; Afzalian, A.; Akhavan, N. D.; Yan, R.; Ferain, I.; Razavi, P.; O’Neill, B.; Blake, A.; White, M.; Kelleher, A.-M.; McCarthy, B.; Murphy, R. Nanowire transistors without junctions. Nat. Nanotechnol. 2010, 5 (3), 225229,  DOI: 10.1038/nnano.2010.15
  15. 15
    Morkötter, S.; Jeon, N.; Rudolph, D.; Loitsch, B.; Spirkoska, D.; Hoffmann, E.; Döblinger, M.; Matich, S.; Finley, J. J.; Lauhon, L. J.; Abstreiter, G.; Koblmüller, G. Demonstration of confined electron gas and steep-slope behavior in delta-doped GaAs-AlGaAs core--shell nanowire transistors. Nano Lett. 2015, 15 (5), 32953302,  DOI: 10.1021/acs.nanolett.5b00518
  16. 16
    Helm, M.; Winnerl, S.; Pashkin, A.; Klopf, J. M.; Deinert, J.-C.; Kovalev, S.; Evtushenko, P.; Lehnert, U.; Xiang, R.; Arnold, A.; Wagner, A.; Schmidt, S. M.; Schramm, U.; Cowan, T.; Michel, P. The ELBE infrared and THz facility at Helmholtz-Zentrum Dresden-Rossendorf. Eur. Phys. J. Plus 2023, 138 (2), 158,  DOI: 10.1140/epjp/s13360-023-03720-z
  17. 17
    Dimakis, E.; Ramsteiner, M.; Tahraoui, A.; Riechert, H.; Geelhaar, L. Shell-doping of GaAs nanowires with Si for n-type conductivity. Nano Res. 2012, 5, 796804,  DOI: 10.1007/s12274-012-0263-9
  18. 18
    Balaghi, L.; Bussone, G.; Grifone, R.; Hübner, R.; Grenzer, J.; Ghorbani-Asl, M.; Krasheninnikov, A. V.; Schneider, H.; Helm, M.; Dimakis, E. Widely tunable GaAs bandgap via strain engineering in core/shell nanowires with large lattice mismatch. Nat. Commun. 2019, 10 (1), 2793,  DOI: 10.1038/s41467-019-10654-7
  19. 19
    Huth, F.; Govyadinov, A.; Amarie, S.; Nuansing, W.; Keilmann, F.; Hillenbrand, R. Nano-FTIR absorption spectroscopy of molecular fingerprints at 20 nm spatial resolution. Nano Lett. 2012, 12 (8), 39733978,  DOI: 10.1021/nl301159v
  20. 20
    Hillenbrand, R.; Knoll, B.; Keilmann, F. Pure optical contrast in scattering-type scanning near-field microscopy. J. Microsc. 2001, 202 (1), 7783,  DOI: 10.1046/j.1365-2818.2001.00794.x
  21. 21
    Huth, F.; Schnell, M.; Wittborn, J.; Ocelic, N.; Hillenbrand, R. Infrared-spectroscopic nanoimaging with a thermal source. Nat. Mater. 2011, 10 (5), 352356,  DOI: 10.1038/nmat3006
  22. 22
    Knoll, B.; Keilmann, F. Enhanced dielectric contrast in scattering-type scanning near-field optical microscopy. Opt. Commun. 2000, 182 (4–6), 321328,  DOI: 10.1016/S0030-4018(00)00826-9
  23. 23
    Cvitkovic, A.; Ocelic, N.; Hillenbrand, R. Analytical model for quantitative prediction of material contrasts in scattering-type near-field optical microscopy. Opt. Express 2007, 15 (14), 85508565,  DOI: 10.1364/OE.15.008550
  24. 24
    Chandler-Horowitz, D.; Amirtharaj, P. M. High-accuracy, midinfrared (450 cm–1⩽ ω⩽ 4000 cm–1) refractive index values of silicon. J. Appl. Phys. 2005, 97 (12), 123526,  DOI: 10.1063/1.1923612
  25. 25
    Eisele, M. Ultrafast Multi-Terahertz Nano-Spectroscopy with Sub-cycle Temporal Resolution, Doctoral Dissertation, University of Regensburg, 2015. .
  26. 26
    Lundstrom, M. Fundamentals of Carrier Transport; Cambridge University Press: Cambridge, 2000.
  27. 27
    Nag, B. R. Electron Transport in Compound Semiconductors; Springer-Verlag: Berlin, 1980.
  28. 28
    Regensburger, S.; Winnerl, S.; Klopf, J. M.; Lu, H.; Gossard, A. C.; Preu, S. Picosecond-scale terahertz pulse characterization with field-effect transistors. IEEE Trans. Terahertz Sci. Technol. 2019, 9 (3), 262271,  DOI: 10.1109/TTHZ.2019.2903630
  29. 29
    Nishida, J.; Johnson, S. C.; Chang, P. T. S.; Wharton, D. M.; Dönges, S. A.; Khatib, O.; Raschke, M. B. Ultrafast infrared nano-imaging of far-from-equilibrium carrier and vibrational dynamics. Nat. Commun. 2022, 13 (1), 1083,  DOI: 10.1038/s41467-022-28224-9
  30. 30
    Shkerdin, G.; Stiens, J.; Vounckx, R. Comparative study of the intra- and intervalley contributions to the free-carrier induced optical nonlinearity in n-GaAs. J. Appl. Phys. 1999, 85 (7), 38073818,  DOI: 10.1063/1.369751
  31. 31
    Shkerdin, G.; Stiens, J.; Vounckx, R. A multi-valley model for hot free-electron nonlinearities at 10.6 μm in highly doped n-GaAs. Eur. Phys. J. Appl. Phys. 2000, 12 (3), 169180,  DOI: 10.1051/epjap:2000185
  32. 32
    Kash, K.; Wolff, P. A.; Bonner, W. A. Nonlinear optical studies of picosecond relaxation times of electrons in n-GaAs and n-GaSb. Appl. Phys. Lett. 1983, 42 (2), 173175,  DOI: 10.1063/1.93864
  33. 33
    Auyang, S. Y.; Wolff, P. A. Free-carrier-induced third-order optical nonlinearities in semiconductors. JOSA B 1989, 6 (4), 595605,  DOI: 10.1364/JOSAB.6.000595
  34. 34
    Vurgaftman, I.; Meyer, J. R.; Ram-Mohan, L. R. Band parameters for III--V compound semiconductors and their alloys. J. Appl. Phys. 2001, 89 (11), 58155875,  DOI: 10.1063/1.1368156
  35. 35
    Anisimov, S. I.; Kapeliovich, B. L.; Perelman, T. L. Electron emission from metal surfaces exposed to ultrashort laser pulses. Zh. Eksp. Teor. Fiz. 1974, 66 (2), 375377
  36. 36
    Lee, S.-C.; Galbraith, I.; Pidgeon, C. R. Influence of electron temperature and carrier concentration on electron-LO-phonon intersubband scattering in wide GaAs/AlxGa1-xAs quantum wells. Phys. Rev. B 1995, 52 (3), 18741881,  DOI: 10.1103/PhysRevB.52.1874
  37. 37
    Zhou, J.-J.; Bernardi, M. Ab initio electron mobility and polar phonon scattering in GaAs. Phys. Rev. B 2016, 94, 201201,  DOI: 10.1103/PhysRevB.94.201201
  38. 38
    Mooshammer, F.; Huber, M. A.; Sandner, F.; Plankl, M.; Zizlsperger, M.; Huber, R. Quantifying nanoscale electromagnetic fields in near-field microscopy by Fourier demodulation analysis. ACS Photonics 2020, 7 (2), 344351,  DOI: 10.1021/acsphotonics.9b01533
  39. 39
    Huth, F.; Chuvilin, A.; Schnell, M.; Amenabar, I.; Krutokhvostov, R.; Lopatin, S.; Hillenbrand, R. Resonant antenna probes for tip-enhanced infrared near-field microscopy. Nano Lett. 2013, 13 (3), 10651072,  DOI: 10.1021/nl304289g

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 2 publications.

  1. Rainer Hillenbrand, Yohannes Abate, Mengkun Liu, Xinzhong Chen, D. N. Basov. Visible-to-THz near-field nanoscopy. Nature Reviews Materials 2025, 10 (4) , 285-310. https://doi.org/10.1038/s41578-024-00761-3
  2. Andrei Luferau, Alexej Pashkin, Stephan Winnerl, Maximilian Obst, Susanne C. Kehr, Emmanouil Dimakis, Thales V. A. G. de Oliveira, Lukas M. Eng, Manfred Helm. Time-resolved nanospectroscopy of III–V semiconductor nanowires. Nanoscale Advances 2025, 7 https://doi.org/10.1039/D5NA00307E

ACS Photonics

Cite this: ACS Photonics 2024, 11, 8, 3123–3130
Click to copy citationCitation copied!
https://doi.org/10.1021/acsphotonics.4c00433
Published August 9, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

950

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. (a) Sketch of a time-resolved near-field spectroscopy setup based on s-SNOM. A Michelson interferometer enables MIR near-field spectroscopy (nanoFTIR) during THz pumping of the sample. (b) Topography of the GaAs/InGaAs core–shell NW recorded by AFM. The tip position chosen for spectrally resolved scans is highlighted with a red triangle. (c) Near-field amplitude s(ω) and phase ϕ(ω) spectra of the doped GaAs/InGaAs core–shell NW normalized to the response of Si along with the results of two-parameter fit based on the point–dipole and the Drude models.

    Figure 2

    Figure 2. (a,b) Near-field amplitude s(ω) and phase ϕ(ω) spectra of the doped GaAs/InGaAs core–shell NW obtained with (red) and without (blue) THz pumping (6 mW) and normalized to the response of Si. The spectra are extracted from experimental data according to eq 3 and plotted along with the results of the two-parameter fit based on the point–dipole and the Drude models. (c,d) Color maps illustrating the evolution of the near-field amplitude s(ω) and the phase ϕ(ω) spectra of the doped InGaAs NW upon THz photoexcitation (6 mW), normalized to the response of Si. Every line represents normalized near-field spectra obtained for different time delays between the THz-pump and broadband MIR probe. (e) Fitting parameter of the plasma frequency ωpl as a function of pump–probe delay time. (f) Time evolution of the effective mass m* of electron gas upon intraband THz pumping.

    Figure 3

    Figure 3. Simulation results: (a) In0.44Ga0.56As band structure scheme depicting the conduction band valleys relevant to the experiment [ (34). The blue dashed line is a parabolic approximation of the nonparabolic Γ-valley. (b,c) Normalized electron distributions for 300 and 1500 K. (d) Calculated fractions of electrons of each conduction band valley versus temperature. (e) Dependence of the total effective mass on temperature. The dashed line represents the simulation neglecting the impact of side valley transfer.

    Figure 4

    Figure 4. (a) Temporal evolution of the electron-gas temperature upon THz-pumping of various powers. (b) Simulation of the temporal evolution of the electron temperature [the same axis as part (a)] based on the two-temperature model for three peak electric-field amplitudes of the FEL radiation inside the NW. The gray dashed line represents the normalized intensity profile of the FEL pulse.

  • References


    This article references 39 other publications.

    1. 1
      Peng, K.; Parkinson, P.; Fu, L.; Gao, Q.; Jiang, N.; Guo, Y.-N.; Wang, F.; Joyce, H. J.; Boland, J. L.; Tan, H. H.; Jagadish, C.; Johnston, M. B. Single nanowire photoconductive terahertz detectors. Nano Lett. 2015, 15 (1), 206210,  DOI: 10.1021/nl5033843
    2. 2
      Peng, K.; Jevtics, D.; Zhang, F.; Sterzl, S.; Damry, D. A.; Rothmann, M. U.; Guilhabert, B.; Strain, M. J.; Tan, H. H.; Herz, L. M.; Fu, L.; Dawson, M. D.; Hurtado, A.; Jagadish, C.; Johnston, M. B. Three-dimensional cross-nanowire networks recover full terahertz state. Science 2020, 368 (6490), 510513,  DOI: 10.1126/science.abb0924
    3. 3
      Parkinson, P.; Lloyd-Hughes, J.; Gao, Q.; Tan, H. H.; Jagadish, C.; Johnston, M. B.; Herz, L. M. Transient terahertz conductivity of GaAs nanowires. Nano Lett. 2007, 7 (7), 21622165,  DOI: 10.1021/nl071162x
    4. 4
      Stiegler, J. M.; Huber, A. J.; Diedenhofen, S. L.; Gomez Rivas, J.; Algra, R. E.; Bakkers, E. P. A. M.; Hillenbrand, R. Nanoscale free-carrier profiling of individual semiconductor nanowires by infrared near-field nanoscopy. Nano Lett. 2010, 10 (4), 13871392,  DOI: 10.1021/nl100145d
    5. 5
      Kuschewski, F.; von Ribbeck, H.-G.; Döring, J.; Winnerl, S.; Eng, L. M.; Kehr, S. C. Narrow-band near-field nanoscopy in the spectral range from 1.3 to 8.5 THz. Appl. Phys. Lett. 2016, 108, 11,  DOI: 10.1063/1.4943793
    6. 6
      Lang, D.; Balaghi, L.; Winnerl, S.; Schneider, H.; Hübner, R.; Kehr, S. C.; Eng, L. M.; Helm, M.; Dimakis, E.; Pashkin, A. Nonlinear plasmonic response of doped nanowires observed by infrared nanospectroscopy. Nanotechnology 2019, 30 (8), 084003,  DOI: 10.1088/1361-6528/aaf5a7
    7. 7
      Joyce, H. J.; Docherty, C. J.; Gao, Q.; Tan, H. H.; Jagadish, C.; Lloyd-Hughes, J.; Herz, L. M.; Johnston, M. B. Electronic properties of GaAs, InAs and InP nanowires studied by terahertz spectroscopy. Nanotechnology 2013, 24 (21), 214006,  DOI: 10.1088/0957-4484/24/21/214006
    8. 8
      Fotev, I.; Balaghi, L.; Schmidt, J.; Schneider, H.; Helm, M.; Dimakis, E.; Pashkin, A. Electron dynamics in In x Ga1–x As shells around GaAs nanowires probed by terahertz spectroscopy. Nanotechnology 2019, 30 (24), 244004,  DOI: 10.1088/1361-6528/ab0913
    9. 9
      Eisele, M.; Cocker, T. L.; Huber, M. A.; Plankl, M.; Viti, L.; Ercolani, D.; Sorba, L.; Vitiello, M. S.; Huber, R. Ultrafast multi-terahertz nano-spectroscopy with sub-cycle temporal resolution. Nat. Photonics 2014, 8 (11), 841845,  DOI: 10.1038/nphoton.2014.225
    10. 10
      Wagner, M.; McLeod, A. S.; Maddox, S. J.; Fei, Z.; Liu, M.; Averitt, R. D.; Fogler, M. M.; Bank, S. R.; Keilmann, F.; Basov, D. N. Ultrafast dynamics of surface plasmons in InAs by time-resolved infrared nanospectroscopy. Nano Lett. 2014, 14 (8), 45294534,  DOI: 10.1021/nl501558t
    11. 11
      Pizzuto, A.; Castro-Camus, E.; Wilson, W.; Choi, W.; Li, X.; Mittleman, D. M. Nonlocal time-resolved terahertz spectroscopy in the near field. ACS Photonics 2021, 8 (10), 29042911,  DOI: 10.1021/acsphotonics.1c01367
    12. 12
      Pushkarev, V.; Němec, H.; Paingad, V. C.; Maňák, J.; Jurka, V.; Novák, V.; Ostatnický, T.; Kužel, P. Charge Transport in Single-Crystalline GaAs Nanobars: Impact of Band Bending Revealed by Terahertz Spectroscopy. Adv. Funct. Mater. 2022, 32 (5), 2107403,  DOI: 10.1002/adfm.202107403
    13. 13
      Egard, M.; Johansson, S.; Johansson, A.-C.; Persson, K.-M.; Dey, A. W.; Borg, B. M.; Thelander, C.; Wernersson, L.-E.; Lind, E. Vertical InAs nanowire wrap gate transistors with ft > 7 GHz and fmax > 20 GHz. Nano Lett. 2010, 10 (3), 809812,  DOI: 10.1021/nl903125m
    14. 14
      Colinge, J.-P.; Lee, C.-W.; Afzalian, A.; Akhavan, N. D.; Yan, R.; Ferain, I.; Razavi, P.; O’Neill, B.; Blake, A.; White, M.; Kelleher, A.-M.; McCarthy, B.; Murphy, R. Nanowire transistors without junctions. Nat. Nanotechnol. 2010, 5 (3), 225229,  DOI: 10.1038/nnano.2010.15
    15. 15
      Morkötter, S.; Jeon, N.; Rudolph, D.; Loitsch, B.; Spirkoska, D.; Hoffmann, E.; Döblinger, M.; Matich, S.; Finley, J. J.; Lauhon, L. J.; Abstreiter, G.; Koblmüller, G. Demonstration of confined electron gas and steep-slope behavior in delta-doped GaAs-AlGaAs core--shell nanowire transistors. Nano Lett. 2015, 15 (5), 32953302,  DOI: 10.1021/acs.nanolett.5b00518
    16. 16
      Helm, M.; Winnerl, S.; Pashkin, A.; Klopf, J. M.; Deinert, J.-C.; Kovalev, S.; Evtushenko, P.; Lehnert, U.; Xiang, R.; Arnold, A.; Wagner, A.; Schmidt, S. M.; Schramm, U.; Cowan, T.; Michel, P. The ELBE infrared and THz facility at Helmholtz-Zentrum Dresden-Rossendorf. Eur. Phys. J. Plus 2023, 138 (2), 158,  DOI: 10.1140/epjp/s13360-023-03720-z
    17. 17
      Dimakis, E.; Ramsteiner, M.; Tahraoui, A.; Riechert, H.; Geelhaar, L. Shell-doping of GaAs nanowires with Si for n-type conductivity. Nano Res. 2012, 5, 796804,  DOI: 10.1007/s12274-012-0263-9
    18. 18
      Balaghi, L.; Bussone, G.; Grifone, R.; Hübner, R.; Grenzer, J.; Ghorbani-Asl, M.; Krasheninnikov, A. V.; Schneider, H.; Helm, M.; Dimakis, E. Widely tunable GaAs bandgap via strain engineering in core/shell nanowires with large lattice mismatch. Nat. Commun. 2019, 10 (1), 2793,  DOI: 10.1038/s41467-019-10654-7
    19. 19
      Huth, F.; Govyadinov, A.; Amarie, S.; Nuansing, W.; Keilmann, F.; Hillenbrand, R. Nano-FTIR absorption spectroscopy of molecular fingerprints at 20 nm spatial resolution. Nano Lett. 2012, 12 (8), 39733978,  DOI: 10.1021/nl301159v
    20. 20
      Hillenbrand, R.; Knoll, B.; Keilmann, F. Pure optical contrast in scattering-type scanning near-field microscopy. J. Microsc. 2001, 202 (1), 7783,  DOI: 10.1046/j.1365-2818.2001.00794.x
    21. 21
      Huth, F.; Schnell, M.; Wittborn, J.; Ocelic, N.; Hillenbrand, R. Infrared-spectroscopic nanoimaging with a thermal source. Nat. Mater. 2011, 10 (5), 352356,  DOI: 10.1038/nmat3006
    22. 22
      Knoll, B.; Keilmann, F. Enhanced dielectric contrast in scattering-type scanning near-field optical microscopy. Opt. Commun. 2000, 182 (4–6), 321328,  DOI: 10.1016/S0030-4018(00)00826-9
    23. 23
      Cvitkovic, A.; Ocelic, N.; Hillenbrand, R. Analytical model for quantitative prediction of material contrasts in scattering-type near-field optical microscopy. Opt. Express 2007, 15 (14), 85508565,  DOI: 10.1364/OE.15.008550
    24. 24
      Chandler-Horowitz, D.; Amirtharaj, P. M. High-accuracy, midinfrared (450 cm–1⩽ ω⩽ 4000 cm–1) refractive index values of silicon. J. Appl. Phys. 2005, 97 (12), 123526,  DOI: 10.1063/1.1923612
    25. 25
      Eisele, M. Ultrafast Multi-Terahertz Nano-Spectroscopy with Sub-cycle Temporal Resolution, Doctoral Dissertation, University of Regensburg, 2015. .
    26. 26
      Lundstrom, M. Fundamentals of Carrier Transport; Cambridge University Press: Cambridge, 2000.
    27. 27
      Nag, B. R. Electron Transport in Compound Semiconductors; Springer-Verlag: Berlin, 1980.
    28. 28
      Regensburger, S.; Winnerl, S.; Klopf, J. M.; Lu, H.; Gossard, A. C.; Preu, S. Picosecond-scale terahertz pulse characterization with field-effect transistors. IEEE Trans. Terahertz Sci. Technol. 2019, 9 (3), 262271,  DOI: 10.1109/TTHZ.2019.2903630
    29. 29
      Nishida, J.; Johnson, S. C.; Chang, P. T. S.; Wharton, D. M.; Dönges, S. A.; Khatib, O.; Raschke, M. B. Ultrafast infrared nano-imaging of far-from-equilibrium carrier and vibrational dynamics. Nat. Commun. 2022, 13 (1), 1083,  DOI: 10.1038/s41467-022-28224-9
    30. 30
      Shkerdin, G.; Stiens, J.; Vounckx, R. Comparative study of the intra- and intervalley contributions to the free-carrier induced optical nonlinearity in n-GaAs. J. Appl. Phys. 1999, 85 (7), 38073818,  DOI: 10.1063/1.369751
    31. 31
      Shkerdin, G.; Stiens, J.; Vounckx, R. A multi-valley model for hot free-electron nonlinearities at 10.6 μm in highly doped n-GaAs. Eur. Phys. J. Appl. Phys. 2000, 12 (3), 169180,  DOI: 10.1051/epjap:2000185
    32. 32
      Kash, K.; Wolff, P. A.; Bonner, W. A. Nonlinear optical studies of picosecond relaxation times of electrons in n-GaAs and n-GaSb. Appl. Phys. Lett. 1983, 42 (2), 173175,  DOI: 10.1063/1.93864
    33. 33
      Auyang, S. Y.; Wolff, P. A. Free-carrier-induced third-order optical nonlinearities in semiconductors. JOSA B 1989, 6 (4), 595605,  DOI: 10.1364/JOSAB.6.000595
    34. 34
      Vurgaftman, I.; Meyer, J. R.; Ram-Mohan, L. R. Band parameters for III--V compound semiconductors and their alloys. J. Appl. Phys. 2001, 89 (11), 58155875,  DOI: 10.1063/1.1368156
    35. 35
      Anisimov, S. I.; Kapeliovich, B. L.; Perelman, T. L. Electron emission from metal surfaces exposed to ultrashort laser pulses. Zh. Eksp. Teor. Fiz. 1974, 66 (2), 375377
    36. 36
      Lee, S.-C.; Galbraith, I.; Pidgeon, C. R. Influence of electron temperature and carrier concentration on electron-LO-phonon intersubband scattering in wide GaAs/AlxGa1-xAs quantum wells. Phys. Rev. B 1995, 52 (3), 18741881,  DOI: 10.1103/PhysRevB.52.1874
    37. 37
      Zhou, J.-J.; Bernardi, M. Ab initio electron mobility and polar phonon scattering in GaAs. Phys. Rev. B 2016, 94, 201201,  DOI: 10.1103/PhysRevB.94.201201
    38. 38
      Mooshammer, F.; Huber, M. A.; Sandner, F.; Plankl, M.; Zizlsperger, M.; Huber, R. Quantifying nanoscale electromagnetic fields in near-field microscopy by Fourier demodulation analysis. ACS Photonics 2020, 7 (2), 344351,  DOI: 10.1021/acsphotonics.9b01533
    39. 39
      Huth, F.; Chuvilin, A.; Schnell, M.; Amenabar, I.; Krutokhvostov, R.; Lopatin, S.; Hillenbrand, R. Resonant antenna probes for tip-enhanced infrared near-field microscopy. Nano Lett. 2013, 13 (3), 10651072,  DOI: 10.1021/nl304289g
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsphotonics.4c00433.

    • Results of the sideband modulation technique along with the point–dipole model-based simulations, extended power-dependent results, and calculation of the specific heat of the electron gas (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.