Abstract
While near-field infrared nanospectroscopy provides a powerful tool for nanoscale material characterization, broadband nanospectroscopy of elementary material excitations in the single-digit terahertz (THz) range remains relatively unexplored. Here, we study liquid-Helium-cooled photoconductive Hg1–XCdXTe (MCT) for use as a fast detector in near-field nanospectroscopy. Compared to the common T = 77 K operation, liquid-Helium cooling reduces the MCT detection threshold to ∼22 meV, improves the noise performance, and yields a response bandwidth exceeding 10 MHz. These improved detector properties have a profound impact on the near-field technique, enabling unprecedented broadband nanospectroscopy across a range of 5 to >50 THz (175 to >1750 cm–1, or <6 to 57 μm), i.e., covering what is commonly known as the “THz gap”. Our approach has been implemented as a user program at the National Synchrotron Light Source II, Upton, USA, where we showcase ultrabroadband synchrotron nanospectroscopy of phonons in ZnSe (∼7.8 THz) and BaF2 (∼6.7 THz), as well as hyperbolic phonon polaritons in GeS (6–8 THz).
Keywords: scattering-type scanning near-field optical microscopy (s-SNOM), synchrotron infrared nanospectroscopy (SINS), terahertz, polariton interferometry, hyperbolic phonon polaritons, van der Waals materials
Introduction
Infrared near-field nanospectroscopy combines the nanometer spatial resolution of atomic force microscopy with the high information density of optical spectroscopy. Due to its unique capabilities, the technique has continuously been expanding and has been applied to material characterization throughout various fields of engineering, chemistry, geology, and physics.1−3 A major and rapidly growing subfield of near-field nanospectroscopy includes polariton interferometry, which provides direct access to the dispersion of deeply subwavelength-confined polariton modes.4−13
Many fundamental material excitations reside in the energy range of several 10 meV, e.g., phenomena related to phonons, excitons, plasmons, Landau level transitions, charge density waves, and magnons. Consequently, there is a strong desire to extend near-field nanospectroscopy toward the far-infrared and terahertz (THz) spectral range.2,3,14 A successful approach to this goal has been combining scattering-type scanning near-field optical microscopy (s-SNOM) with infrared radiation from accelerator-based light sources.1−3,9,10,14−21 Specifically, synchrotron-based infrared near-field spectroscopy has pushed the spectral limit of nanospectroscopy down to ∼10 THz (320 cm–1, 31 μm), thereby enabling many new experiments.3,14−16,19−21 Additionally, coming from lower frequencies (microwave regime), nanospectroscopy and nanoimaging around and below 1 THz is nowadays more frequently achieved, e.g., via THz time-domain spectroscopy and setups based on Schottky diodes or photoconductive antennas.22−29 Despite such advances, the lack of suitable sources and matching detectors in the 1–10 THz range—the so-called “THz gap”2,3,30 —has so far hampered extending broadband near-field nanospectroscopy into the center of this important spectral region. While examples of near-field nanospectroscopy and polariton interferometry in the 1–10 THz spectral range do exist,9,10,18,31−38 such studies usually employ high-intensity narrowband laser sources to compensate for less-than-ideal detectors, thereby inducing limitations for spectroscopy.
The most common method for achieving near-field infrared nanospectroscopy at mid- to far-infrared wavelengths involves the scattering of infrared light from the tip of an atomic force microscope (known as “apertureless scattering”), while the tip tracks the surface of a material in tapping mode.2,15,39,40Figure 1a illustrates the implementation of this scheme at the National Synchrotron Light Source II (NSLS-II), Upton, USA. The light scattered from the oscillating tip is not a simple sinusoid due to the interaction of the tip with a material, and a more localized response is recovered by demodulating the signal at the second or third harmonic of the tapping frequency.39 With typical AFM tapping frequencies in the range of up to 300 kHz,2,15 the detector for sensing the scattered infrared radiation requires a response bandwidth approaching 1 MHz.
Figure 1.
(a) Sketch of the near-field nanospectroscopy setup at the National Synchrotron Light Source II (NSLS-II). Infrared synchrotron radiation is guided via an asymmetric Michelson interferometer and focused onto an atomic force microscopy (AFM) tip. The latter is vibrating at frequency Ω (tapping mode). Light scattered from the tip is guided to a suitable detector as described in the main text. Near-field contributions to the detected signal are separated from the far-field background via lock-in demodulation of the detector signal at harmonics nΩ of the tip frequency with n = 1,2,3,··· The experimental near-field data shown in this paper uses n = 2. (b) Calculated band gap for a Hg1–xCdxTe (MCT) infrared detector element for x = 0.173 as a function of temperature below 100 K; based on ref (41). Cooling MCT to T = 4.2 K shifts its detection threshold to <200 cm–1 (>50 μm, <6 THz).
The commonly employed spectral range covers wavelengths from 3 to 16 μm (625 to 3300 cm–1, 19 THz to 100 THz), for which fast photodiodes made from both III–V and II–VI compound semiconductors are commercially available. Detectors for reaching into the far-infrared include the II–VI alloy mercury–cadmium-telluride (Hg1–xCdxTe or MCT) operating at T = 77 K for wavelengths λ out to λ ∼ 24 μm (417 cm–1, 12 THz; x near 0.174) and doped (i.e., extrinsic) germanium or silicon operated at T = 4.2 K for reaching λ ∼ 30 μm (330 cm–1, 10 THz). These types of detectors have been used successfully for near-field nanospectroscopy in combination with the broadband, high radiance infrared produced as synchrotron radiation.1,3,14,42 Another extrinsic photoconductor, Ge/Ga, operates over a somewhat limited spectral range below λ ∼ 120 μm. Though successfully used with narrowband free-electron laser sources down to λ ∼ 30 μm,9,10,18,34−36 there has not been reported measurement results for the synchrotron source. Extending infrared broadband nanospectroscopy to longer wavelengths has been a challenge as the standard detector for reaching to 50 μm and beyond, the doped silicon composite bolometer operating at T = 4.2 K, has a response time measured in hundreds of microseconds, so it is orders of magnitude too slow for detection at AFM tip modulation frequencies. Faster thermal detectors, such as InSb hot electron bolometers (HEB), can achieve the necessary speed but the HEB effect is limited to very long wavelengths. Superconducting transition edge bolometers can also be fast but tend to have a limited dynamic range as well as challenges in manufacture.
One aspect of the Hg1–xCdxTe system that is less well-known is the dependence of the band gap and detection threshold as a function of temperature.41,43,44 In many semiconductors, lowering the temperature causes the band edge to shift to higher energies. However, for Hg1–xCdxTe with x < 0.5, the band edge shifts to lower energies with decreasing T.41,43−45 This was previously noted by the Soleil synchrotron infrared group for a mid-infrared detector intended for high spectral resolution applications.45 Typical values for x in commercial detectors range from 0.2 to 0.17 for a detection threshold in the far-infrared. When operated at T = 77 K, the threshold for x = 0.173 is near 400 cm–1. However, as illustrated in Figure 1b, by cooling this material to about 4.2 K, its threshold shifts down to below 200 cm–1.41
In this work, we test the operation of a photoconductive
MCT element
having x near 0.173 at its nominal operating temperature
of T = 77 K and when cooled with liquid helium (
) to about T = 4.2 K. Details
of the experimental realization of our detector setup are described
in Methods. In agreement with the known band
gap dependence on temperature (Figure 1b),41 the detection threshold
for T = 4.2 K shifts down to below 200 cm–1. Additionally, similar to the observations of the Soleil group,45 the detector signal-to-noise shows improvement.
Overall, we find that the performance of the detector is well-suited
for infrared near-field nanospectroscopy down to around 175 cm–1 (5.2 THz, 57 μm), i.e., well into the single-digit
THz range. We demonstrate this via point nanospectroscopy measurements
on Au, ZnSe, and BaF2 as well as polariton interferometry
of the hyperbolic phonon polaritons in GeS in the 6–8 THz range.
Results
Far-Field Characterization of Detector Performance
Our far-field characterization of the detector performance consisted of measuring the spectral response, the signal-to-noise, and the response speed at both T = 77 K and T = 4.2 K.
Spectral Response and Signal-to-Noise
The spectral range for the photoconductive response was measured by using a Bruker Vertex 80v spectrometer. Due to the limited availability of the synchrotron light source, we used the spectrometer’s internal infrared source for this measurement. While a better signal-to-noise ratio could be obtained by employing the high spectral radiance of the synchrotron light source, here, this is not required for evaluating the detector’s spectral response. Figure 2a shows the resulting spectra for operation of our modified MCT detector at T = 77 K and T = 4.2 K. The lower operating temperature reduces the detection threshold to around 160 cm–1 and significantly increases the overall response. The inset in Figure 2a shows a “100% line” that has been obtained by taking the ratio of two spectra acquired sequentially with identical measurement parameters. As an ideal system should yield the same result for two such measurements, deviations from unity can be attributed to either spectral drift or the noise of the overall system. The 100% lines in Figure 2a illustrate both the redshift of the detection threshold as well as the better signal-to-noise ratio for operation at lower temperature. Note that the aperture for the measurement at T = 77 K was 2 mm, while it was only 1 mm for T = 4.2 K. As the data has not been scaled to compensate for this, this means the T = 4.2 K is for about 4× less intensity than the T = 77 K data but still shows better signal-to-noise ratio, even in the higher-frequency range, e.g., around 600 cm–1. When integrated over the entire spectral range (to several thousand cm–1, refer to Methods), the signal-to-noise ratio can be quantified to be ∼5× better for operation at T = 4.2 K despite the lower photon flux of the measurement.
Figure 2.
Far-field characterization
of the detector performance. (a) Non-normalized
far-field FTIR spectra acquired with our modified MCT detector cooled
to either T = 4.2 K (1 mm aperture, blue line) or T = 77 K (2 mm aperture, orange line). When cooled with
(4.2 K), the detection threshold redshifts
to around 160 cm–1, and the overall detector sensitivity
is significantly increased. The local minima in the 350 to 550 cm–1 range are due to the spectrometer’s multilayer
mylar beamsplitter. Inset: measured “100% line” as described
in the main text. For negligible spectral drift, deviations from unity
provide an estimate of the overall system noise. (b) Measurement of
a short isolated 10 ps light pulse using the MCT detector cooled to T = 4.2 K (blue) and T = 77 K (orange),
respectively. At
temperature, both sensitivity and speed
are significantly increased. For the inset, both curves are scaled
and offset for a better comparison of the response speed.
Response Speed
Since the response bandwidth is important for the apertureless scattering method, we measured the detector rise and fall times using a single isolated light pulse from the NSLS-II synchrotron storage ring. The pulse itself has an RMS length of about 10 ps, i.e., sufficient for measuring response up to tens of GHz. We used a Femto DHPCA-100 preamplifier set to a frequency response of 80 MHz along with a 1 GHz bandwidth oscilloscope and observed an exponential decay of (20 ± 10) ns at T = 77 K, corresponding to an (8 ± 4) MHz bandwidth (Figure 2b). At T = 4.2 K, the fall time was (7 ± 3) ns, corresponding to (21 ± 8) MHz bandwidth. Thus, the response bandwidth increased by a factor of 2–3 going from 77 to 4.2 K. We do not offer a theoretical analysis of the intrinsic relaxation (recombination) time for MCT as a function of T and x, but note that listed response times for commercial detectors show a trend of faster performance as the band gap becomes smaller. That the photoconductive response magnitude increased for lower T suggests an improved carrier mobility, offsetting the reduced photocarrier density that would stem from a faster relaxation (i.e., carrier recombination) time.
Mid-Infrared-to-THz Near-Field Nanospectroscopy
Its
low-energy detection threshold, high sensitivity, good signal-to-noise
ratio, and high detection speed make
He-cooled MCT
a very suitable detector for
ultrabroadband nanospectroscopy down to the single-digit THz range.
As a direct demonstration of our system’s capabilities, we
show different ultrabroadband nanospectroscopy results, namely, point
spectroscopy of a gold reference sample, spectra of the THz phonon-induced
near-field resonances in ZnSe and BaF2 and polariton interferometry
of THz phonon polaritons in GeS. Near-field nanospectroscopy is performed
using a commercial near-field microscope (NeaSnom by Attocube) at
the infrared MET beamline of the NSLS-II synchrotron, Upton, USA (sketch
of the experimental setup in Figure 1a); to allow for ultrabroadband operation, we use a
diamond beamsplitter. The near-field data shown in the paper are demodulated
at the second harmonic of the tip tapping frequency.
Nanospectroscopy of Gold Reference
As gold provides a spectrally flat infrared response, measuring the spectrum of a gold reference sample allows us to determine the overall spectral response. The latter combines spectral contributions from the broadband synchrotron radiation, absorption features of the optical elements and the system atmosphere, the s-SNOM tip, and the detector response. Figure 3a shows a gold reference point spectrum as measured with our nanospectroscopy system at NSLS-II: We obtain a near-field signal well above the noise level down to about 175 cm–1 (5.2 THz, 57 μm). Note that this exceeds the previous spectral limit of synchrotron infrared nanospectroscopy of 320 cm–1 (9.7 THz, 31 μm)14 by almost an octave. Also, this demonstrates the extension of nanospectroscopy well into the so-called THz gap from 1 to 10 THz,30 a spectral region that historically has been very challenging to access with near-field techniques.2,3
Figure 3.

Midinfrared-to-THz near-field nanospectroscopy with the AFM tip at a fixed sample position (point spectroscopy). (a) Near-field amplitude spectrum of a gold reference sample illustrating the broad spectral response of our system. The system covers the spectral range from about 175 to beyond 1750 cm–1, i.e., a full order of magnitude. Inset: detailed view of the low-frequency onset. The dashed horizontal line marks the noise floor. (b,c) Measured near-field amplitude of (b) ZnSe and (c) BaF2 in the spectral range from 175 to 1200 cm–1 (5.2–36.0 THz). Each spectrum has been obtained within 10 min measurement time at a resolution of 3.3 cm–1. The near-field amplitude is normalized to a spectrum taken on a separate gold reference sample. For both materials, phonon-induced resonant tip–sample interaction results in a peak of the normalized amplitude in the spectral range from 200 to 300 cm–1 (6–9 THz, approximate peak position marked by a green dash-dotted line). For comparison, the insets show literature values46,47 for the real part (full line) and imaginary part (dashed line) of the dielectric permittivity in the spectral range of the phonon modes. While the specific position of the phonon-polariton-induced near-field resonance depends on experimental parameters and the complex dielectric material permittivity, as a rule-of-thumb, the resonance is expected in the spectral range where the real part of the permittivity ε′ is in the range −10 < ε′ < −1.18,48−50
THz Phonon-Induced Near-Field Resonances
ZnSe and BaF2 have widespread application as optical materials in the mid-infrared spectral range, e.g., as windows or beam splitters. While transparent in the mid infrared, both materials provide a clear phonon resonance at <300 cm–1.46,47,51 Here, we probe the phonon polariton modes of ZnSe and BaF2 as an example for two materials whose polaritonic responses were previously not accessible via broadband infrared nanospectroscopy.
Figure 3b,c shows the measured near-field amplitude point spectra of ZnSe and BaF2 in the broad spectral range from 175 to 1200 cm–1 (5.2 to 36.0 THz). While both materials are spectrally flat at higher frequencies, at <300 cm–1 (<9.0 THz), they show a peak of the near-field amplitude. For ZnSe and BaF2, the maximum of this peak is observed at 222 cm–1 (6.7 THz, 45.0 μm) and 259 cm–1 (7.8 THz, 38.6 μm), respectively. As previously observed for other materials,2,14,17,18,40,48,49,52 this can be attributed to phonon polariton modes inducing a resonant tip–sample near-field interaction. Note that the precise spectral peak position and width as well as the relative peak height are known to depend on experimental parameters such as the tip shape, tip tapping amplitude, average tip–sample distance, and the harmonic order used for lock-in demodulation.3,17,49,52,53 Nevertheless, the spectral positions of the near-field resonances provide an excellent way for nanoscale material characterization3,14,40 and, here, match well to the expectation from literature phonon parameters of ZnSe and BaF2.46,47,51 As a rule-of-thumb, the near-field resonance is expected in the spectral range where the real part of the permittivity ε’ (insets of Figure 3b,c) is in the range −10 < ε′ < −1.18,48−50 This explains why the phonon-induced near-field resonance for BaF2 is observed at higher frequency than the resonance for ZnSe, although the BaF2 transverse optic (TO) phonon mode itself is observed at a lower frequency.46,47,51
Each spectrum in Figure 3 has a spectral resolution of 3.3 cm–1 and has been obtained in a measurement time of 10 min, thereby demonstrating ultrabroadband mid-infrared-to-THz nanospectroscopy at reasonable signal-to-noise ratio and spectral resolution without the need for extreme integration times. Depending on the requirements for a specific experiment, the signal-to-noise ratio can be further increased at the cost of longer integration times.
THz Hyperspectral Polariton Interferometry
GeS is a semiconducting van der Waals material (band gap around 1.6 eV54,55) with many intriguing properties such as an exceptional Seebeck coefficient,56 monolayer ferroelectricity,57 and moiré physics in twisted nanowires.58 Its stacking direction is along the crystalline [001] direction, and exfoliated flakes show highly in-plane anisotropic polariton dispersion,10 which is potentially field-tunable.10,54,55 The combination of these properties makes GeS a promising platform for versatile polariton control.10 Along GeS’s [010] crystalline direction, polariton modes have been observed down to 6.1 THz, thereby extending to lower frequencies than in the other in-plane direction.10 Hence, the hyperbolic polariton modes along the [010] direction of GeS provide a model system to demonstrate polariton interferometry in the <10 THz spectral range.
Figure 4a,b shows the topography and homodyne near-field image of an exfoliated GeS flake (thickness of 149 nm) on SiO2/Si substrate (sample preparation described in Methods). The homodyne near-field signal is acquired while keeping fixed the position of the interferometer reference mirror. A hyperspectral linescan with 3.3 cm–1 spectral resolution (Figure 4c,d) and 50 nm pixel size was measured along the crystalline [010] direction as marked by the blue arrow in Figure 4b. Figure 4c provides an overview of the near-field amplitude (top) and phase (bottom) in the broad spectral range from 175 to 1200 cm–1; the spectral (spatial) information is displayed along the horizontal (vertical) axis. Positive distances refer to a position on the GeS crystal; negative values refer to a position on the SiO2 substrate. SiO2 shows three well-studied phonon-induced near-field resonances at around 500, 800 cm–1 (weak amplitude, refer to phase signal), and 1100 cm–1.14,40 While GeS itself shows no phonon modes at >400 cm–1,59−61 the influence of the SiO2 substrate also modulates the near-field response acquired on the GeS crystal. At lower frequencies, GeS phonon modes induce a strongly anisotropic hyperbolic infrared response.10,59−61 As shown in the zoom-in of the near-field signal in the low-frequency spectral region from 175 to 350 cm–1 in Figure 4d, here, the phonon modes are apparent from a clear modulation of the GeS near-field response. Notably, in the spectral range from about 200 to 260 cm–1 (6.0–7.8 THz, 38–50 μm), a clear fringe pattern is observable both in the near-field amplitude and phase measured on the GeS crystal. This fringe pattern can be attributed to interference patterns caused by hyperbolic phonon polaritons propagating along the [010] direction of GeS.10 The dispersion of these polaritons modes, i.e., their decreasing wavelength with increasing photon energy, can be directly seen in the measurement. For a detailed description of these polariton modes, we refer to ref (10). Compared to previous studies with a high-intensity narrowband laser source,10 we emphasize that the present study demonstrates the first broadband nanospectroscopy of the GeS polariton modes as well as the first interferometric study, resolving both near-field amplitude and phase of the modes.
Figure 4.
(a) Topography of a 149 nm thick GeS flake on a SiO2 substrate. Black arrows mark the crystallographic [100] and [010] directions of GeS. (b) Homodyne (whitelight) near-field image taken simultaneously to (a). The blue arrow marks the position and direction of the hyperspectral linescan shown in (c,d). (c) Near-field amplitude (top) and phase (bottom) of a hyperspectral linescan taken along the crystallographic [010] direction of GeS as marked by the blue arrow in (b). The spatial dimension of the linescan is shown in the vertical direction; positive distances refer to a position on the GeS crystal; negative values refer to a position on the SiO2 substrate. The spectral dimension of the linescan is shown in the horizontal direction; the spectral resolution is 3.3 cm–1. SiO2 shows three well-studied phonon resonances at around 500, 800 cm–1 (weak amplitude, refer to phase signal), and 1100 cm–1.14,40 GeS shows polariton modes only at lower frequencies.10,59−61 (d) Zoom-in of the low-frequency region (175–350 cm–1) of the near-field amplitude and phase spectra as marked by a white-dashed box in (c). A clear fringe pattern can be observed in the spectral range from about 200–260 cm–1 (6.0–7.8 THz), which can be attributed to hyperbolic phonon polariton modes propagating along the [010] direction of GeS.10
Discussion and Conclusions
We have tested liquid-helium-cooled
photoconductive MCT (Hg1–XCdXTe
with x near 0.173) for usage as a fast detector for
far-infrared near-field nanospectroscopy in the single-digit THz frequency
range. Building up on previous literature results,41,43−45 we found that cooling MCT down to 4.2 K lowers its
detection threshold to 22 meV. In addition, compared to operation
at 77 K, we found that cooling MCT to 4.2 K increased both its sensitivity
and speed. This suggests a strongly improved carrier mobility that
overcompensates for the reduced photocarrier density expected from
faster carrier recombination. The combination of increased detector
sensitivity and speed will be beneficial for various mid- to far-infrared
optical techniques that require a fast response time. This potentially
includes any optical setup that currently employs a
-cooled MCT detector, i.e., both
far- and
near-field imaging and spectroscopy techniques. Note that the increased
detector responsivity also covers the mid-infrared spectral range
around 1000 cm–1 (30 THz, 10 μm, refer to Methods and ref (45)), which is a common spectral range for many
table-top near-field imaging and nanospectroscopy setups.2 Going beyond the conventional spectral range
covered by MCT detectors, we demonstrate that the properties of our
specific
He-cooled MCT
detector are well suited for
near-field nanospectroscopy down to 175 cm–1 (5.2
THz, 57 μm). Further extension of this spectral range to even
lower THz frequencies may be achieved via the optimization of the
MCT stoichiometry. Here, near-field point nanospectroscopy down to
the single-digit THz range is demonstrated for different samples (Au
reference, BaF2, ZnSe). For ZnSe and BaF2, we
observe a phonon-induced resonant near-field interaction around 222
cm–1 (6.7 THz, 45.0 μm) and 259 cm–1 (7.8 THz, 38.6 μm), respectively. The good signal-to-noise
ratio of our measurements also enables us to take hyperspectral linescans
with high spectral (3.3 cm–1) and spatial resolution
(50 nm pixel size for a 7.5 μm long linescan). We use this for
polariton interferometry of the hyperbolic phonon polariton modes
of GeS in the 6 to 8 THz frequency range, resolving both the near-field
amplitude and phase of these modes.
While here we demonstrate synchrotron-based THz nanospectroscopy, we emphasize that our detector system is expected to work equally well for laser-based THz nanoimaging and nanospectroscopy setups, e.g., based on free-electron-lasers,9,10,18,34−36 gas lasers,31,33,62 or THz quantum cascade lasers.32,37,38 The high signal-to-noise ratio of our measurements implies that our technique will also provide great opportunities for studying the THz polariton dispersion in other van der Waals and two-dimensional materials, e.g., in graphene,28,36 black phosphorus,6 transition metal dichalcogenides,29 or topological insulators such as Bi2Se3.6,32,33 Furthermore, access to a plethora of new phenomena is expected when combining the here presented ultrabroadband THz nanospectroscopy with extreme sample environments in cryogenic29,63−68 or magnetic69,70 near-field microscopy. For example, this may enable the nanoscopic exploration of Cooper-pair and Josephson plasmon polaritons in superconductors that exclusively reside in the THz range.6,7,71 Overall, our work demonstrates the extension of broadband infrared near-field nanospectroscopy well into the single-digit THz range, thereby significantly extending the possibilities for nanoscale optical material characterization and probing fundamental collective excitations at their natural length and energy scales.
The near-field nanospectroscopy setup built for the present work is available for general user operation at the 22-IR infrared beamline of the National Synchrotron Light Source II. Proposals can be submitted free of charge via an online proposal system.
Methods
Detector System Details
The photoconductive MCT detector
for this study was a commercial model purchased from Infrared Associates.
Listed as a “D24″, it is intended for reaching to approximately
λ ∼ 24 μm or frequencies near 400 cm–1 (12.5 THz), corresponding to an alloy composition x near 0.173.41 The detector assembly,
including its ceramic chip carrier mount and field-of-view (FOV) limiting
aperture (see Figure 5a), was removed from its
N2 dewar and mounted onto an
oxygen-free high-thermal-conductivity copper bracket attached to the
copper cold plate of an Infrared Laboratories HDL-5 liquid helium
cryostat, as shown in Figure 5b. A Winston cone collector was not used, as the detector
package was not compatible with the nominal integrating cavity used
at the cone’s exit. However, the field-of-view (FOV) limiting
aperture provided by the original manufacturer was retained.
Figure 5.

(a) MCT detector
package from Infrared Associates including a FOV
limiting baffle and electrical connections (only two are needed).
The MCT photoconductive element itself is the small, bright object
inside and behind the baffle’s circular aperture. The ceramic
chip carrier is hidden behind the FOV baffle. (b) MCT detector package
mounted to an oxygen-free high-thermal-conductivity copper bracket
(just right of center) bolted onto the cold plate of the
reservoir. The CsI window is mounted in
the gold-anodized flange at the right edge. A separate Si/B photoconductive
detector assembly is near the top and includes a Winston cone optic
(cylindrical object) and integrating cavity (inside the small square
block with four bolts).
A 2 mm thick CsI disk
was used for the infrared
vacuum window on
HDL-5, providing good transmission to about 160 cm–1 (63 μm, 4.8 THz), as shown in Figure 6. For testing at T = 77
K rather than at T = 4.2 K, the
He reservoir
was filled with
N2.
Figure 6.

Measured far-infrared transmission for the cesium iodide vacuum window used on the HDL-5 detector cryostat. CsI transmits well (∼90%) up through visible light frequencies.
The spectral range for the photoconductive response
was measured
by using a Bruker Vertex 80v spectrometer. Due to limited availability
of the synchrotron light source, we used the spectrometer’s
internal infrared source for this measurement. While a better signal-to-noise
ratio could be obtained by employing the high spectral radiance of
the synchrotron light source, here, this is not required for evaluating
the detector’s spectral response. Figure 7 shows the same measurement as Figure 2a, i.e., a comparison between
the MCT detector response when operated at either T = 77 K or T = 4.2 K. While Figure 2a shows the low-frequency region at <600
cm–1 to better illustrate the redshift of detection
threshold, Figure 7 shows the detector response in the full spectral range up to 7000
cm–1. We chose a multilayer mylar beamsplitter for
this measurement, which is a good choice for the low-frequency spectral
range shown in Figure 2a. However, we note that this beamsplitter is not optimized for the
mid-infrared spectral region where it induces various local minima
in the measured response, most pronounced in a broad frequency region
around ∼1000 cm–1. Note that we employed
a diamond beamsplitter for the near-field measurements, i.e., the
minima apparent in Figure 7 do not impact the near-field response shown in Figures 3 and 4. Despite the beamsplitter being optimized for the far-infrared spectral
range, Figure 7 clearly
demonstrates the detector responsivity enhancement in an ultrabroad
spectral range: we find that cooling the detector to T = 4.2 K enhances the detector response in the full accessible spectral
range up to several thousand wavenumbers, as previously noted for
a MCT detector with slightly different stoichiometric composition.45 This suggests applications of
He-cooled MCT
detectors also in the mid-infrared
spectral range, with signal-to-noise improvement both for far-field45 and near-field techniques. To further test the
responsivity enhancement of the detector, we performed separate measurements
under even more controlled conditions (broadband response measurement
using a thermal source with fixed aperture, optical chopper, and lock-in
detection). One goal of these separate measurements was to minimize
potential impacts on the measured detector responsivity that may be
expected due to the different photon fluxes incident on the detector
(Auger recombination). As such, the photon flux incident on the detector
was kept constant for the separate measurements. Some minor impact
on the “effective” photon flux (as seen by the detector)
may still originate from the shift of the low-frequency cutoff. However,
the photon flux in the spectral range between ∼160 cm–1 (cutoff at 4.2 K) and ∼400 cm–1 (cutoff
at 77 K) is much smaller than the photon flux in the large detected
spectral region >400 cm–1 such that a measurement
at constant incident photon flux remains meaningful. While we believe
that a detailed study of the photon-flux-dependent detector response
is beyond the scope of the current paper (to be published elsewhere),
we note that our broadband study at constant incident photon flux
gave a factor of 10 higher signal for operation at 4.2 K compared
to 77 K, which is not inconsistent with the results shown in Figures 2a and 7.
Figure 7.

Non-normalized far-field FTIR spectra acquired with our modified
MCT detector cooled to either T = 4.2 K (1 mm aperture,
blue line) or T = 77 K (2 mm aperture, orange line).
When cooled with
(4.2 K), the detection threshold redshifts
to around 160 cm–1, and the overall detector sensitivity
is significantly increased. The local minima at various frequencies
throughout the spectrum are due to the spectrometer’s multilayer
mylar beamsplitter. Figure 2 shows a detailed view of the low-frequency region <600
cm–1.
We compared our modified
He-cooled MCT
detector with the response
of a more common
He-cooled
Si/B photoconductor. Note that
the latter has an f/4 Winston cone to limit the FOV to ±7°,
reducing the background infrared reaching the detector. The MCT detector,
with its simpler ±15 °FOV limiting aperture, had at least
4 times more background falling on the detector element. As such,
the comparison is more qualitative than quantitative. For both detectors, Figure 8 shows a measured
100% line as described in the main paper. For the specific detector
setups used here, we find that the Si/B provides a better signal-to-noise
ratio in the spectral range from just above its cutoff frequency at
around 320 cm–1 to about 800 cm–1; both for frequencies below and above this range, the MCT provides
a better signal-to-noise ratio despite the larger background infrared
falling onto the MCT detector.
Figure 8.

Measured “100% line” of
our modified
He-cooled MCT
detector in comparison to
a more common
He-cooled Si/B
photoconductor.
Sample Preparation: GeS Exfoliation
GeS flakes were mechanically exfoliated using an automatic mechanical exfoliation machine, or an exfoliator,72 one of the home-built modules in the Quantum material press (QPress) facility in the Center for functional nanomaterials (CFN) at Brookhaven National Laboratory (BNL). The exfoliator utilizes a roller assembly to control conditions such as the pressure, temperature, rolling and peeling speed, and peeling force. The exfoliation process has two main steps. First, the GeS bulk crystals (2D Semiconductors) were distributed on blue dicing tape and loaded onto the sample stage of the exfoliator. The exfoliator transferred the crystals on the built-in blue tape. Then, we replace the tape with bulk crystals on the sample stage on a diced 285 nm SiO2-coated Si wafer chip (10 mm × 10 mm) attached to a sample holder. The exfoliator conducted the exfoliation automatically. In each step, we used the conditions for the GeS crystal transfer and exfoliation provided by the facility. After the exfoliation, the SiO2 chip with exfoliated flakes was scanned by the optical microscope of the cataloger module (WiTec Alpha 300RA) in the QPress.
Acknowledgments
We are grateful to Attocube for the use of the NeaSNOM near-field microscope.
Author Contributions
L. Wehmeier, M.K. Liu, D.N. Basov, and G.L. Carr conceived the project and designed the experiments. L. Wehmeier, S. Park, and H. Jang prepared the GeS sample. L. Wehmeier and G.L. Carr performed the experimental measurements and analyzed the experimental data with support from M.K. Liu and C.C. Homes. L. Wehmeier and G.L. Carr cowrote the manuscript with input from all coauthors.
Research on near-field studies of low-dimensional materials was supported as part of Programmable Quantum Materials (Pro-QM), an Energy Frontier Research Center (EFRC) funded by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences (BES), under award DE-SC0019443. The construction of the near-field instrument was funded by the U.S. Department of Energy, Office of Science, National Quantum Information Science Research Centers, Co-design Center for Quantum Advantage (C2QA) under contract number DE-SC0012704. For synchrotron-based infrared nanospectroscopy, this research used the MET beamline of the National Synchrotron Light Source II, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Brookhaven National Laboratory under Contract No. DE-SC0012704. For sample fabrication, this research used Quantum Material Press (QPress) of the Center for Functional Nanomaterials (CFN), which is a U.S. Department of Energy Office of Science User Facility, at Brookhaven National Laboratory under Contract No. DE-SC0012704.
The authors declare the following competing financial interest(s): S. Park has a patent pending regarding the exfoliator used for the GeS exfoliation (patent application number PCT/US2022/040777, filing date 08/18/2022; international publication number WO2023/2023260A2, publication date 02/23/2023). All other authors declare no competing interests.
References
- Bechtel H. A.; Muller E. A.; Olmon R. L.; Martin M. C.; Raschke M. B. Ultrabroadband Infrared Nanospectroscopic Imaging. Proc. Natl. Acad. Sci. U.S.A. 2014, 111 (20), 7191–7196. 10.1073/pnas.1400502111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen X.; Hu D.; Mescall R.; You G.; Basov D. N.; Dai Q.; Liu M. Modern Scattering-Type Scanning Near-Field Optical Microscopy for Advanced Material Research. Adv. Mater. 2019, 31, 1804774. 10.1002/adma.201804774. [DOI] [PubMed] [Google Scholar]
- Bechtel H. A.; Johnson S. C.; Khatib O.; Muller E. A.; Raschke M. B. Synchrotron Infrared Nano-Spectroscopy and -Imaging. Surf. Sci. Rep. 2020, 75 (3), 100493. 10.1016/j.surfrep.2020.100493. [DOI] [Google Scholar]
- Fei Z.; Rodin A. S.; Andreev G. O.; Bao W.; McLeod A. S.; Wagner M.; Zhang L. M.; Zhao Z.; Thiemens M.; Dominguez G.; Fogler M. M.; Neto A. H. C.; Lau C. N.; Keilmann F.; Basov D. N. Gate-Tuning of Graphene Plasmons Revealed by Infrared Nano-Imaging. Nature 2012, 487 (7405), 82–85. 10.1038/nature11253. [DOI] [PubMed] [Google Scholar]
- Chen J.; Badioli M.; Alonso-González P.; Thongrattanasiri S.; Huth F.; Osmond J.; Spasenović M.; Centeno A.; Pesquera A.; Godignon P.; Zurutuza Elorza A.; Camara N.; de Abajo F. J. G.; Hillenbrand R.; Koppens F. H. L. Optical Nano-Imaging of Gate-Tunable Graphene Plasmons. Nature 2012, 487 (7405), 77–81. 10.1038/nature11254. [DOI] [PubMed] [Google Scholar]
- Basov D. N.; Fogler M. M.; Garcia de Abajo F. J. Polaritons in van Der Waals Materials. Science 2016, 354 (6309), aag1992. 10.1126/science.aag1992. [DOI] [PubMed] [Google Scholar]
- Basov D. N.; Asenjo-Garcia A.; Schuck P. J.; Zhu X.; Rubio A. Polariton Panorama. Nanophotonics 2020, 10 (1), 549–577. 10.1515/nanoph-2020-0449. [DOI] [Google Scholar]
- Fali A.; White S. T.; Folland T. G.; He M.; Aghamiri N. A.; Liu S.; Edgar J. H.; Caldwell J. D.; Haglund R. F.; Abate Y. Refractive Index-Based Control of Hyperbolic Phonon-Polariton Propagation. Nano Lett. 2019, 19 (11), 7725–7734. 10.1021/acs.nanolett.9b02651. [DOI] [PubMed] [Google Scholar]
- de Oliveira T. V. A. G.; Nörenberg T.; Álvarez-Pérez G.; Wehmeier L.; Taboada-Gutiérrez J.; Obst M.; Hempel F.; Lee E. J. H.; Klopf J. M.; Errea I.; Nikitin A. Y.; Kehr S. C.; Alonso-González P.; Eng L. M. Nanoscale-Confined Terahertz Polaritons in a van Der Waals Crystal. Adv. Mater. 2021, 33 (2), 2005777. 10.1002/adma.202005777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nörenberg T.; Álvarez-Pérez G.; Obst M.; Wehmeier L.; Hempel F.; Klopf J. M.; Nikitin A. Y.; Kehr S. C.; Eng L. M.; Alonso-González P.; de Oliveira T. V. A. G. Germanium Monosulfide as a Natural Platform for Highly Anisotropic THz Polaritons. ACS Nano 2022, 16 (12), 20174–20185. 10.1021/acsnano.2c05376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sternbach A. J.; Moore S. L.; Rikhter A.; Zhang S.; Jing R.; Shao Y.; Kim B. S. Y.; Xu S.; Liu S.; Edgar J. H.; Rubio A.; Dean C.; Hone J.; Fogler M. M.; Basov D. N. Negative Refraction in Hyperbolic Hetero-Bicrystals. Science 2023, 379 (6632), 555–557. 10.1126/science.adf1065. [DOI] [PubMed] [Google Scholar]
- Hu H.; Chen N.; Teng H.; Yu R.; Xue M.; Chen K.; Xiao Y.; Qu Y.; Hu D.; Chen J.; Sun Z.; Li P.; de Abajo F. J. G.; Dai Q. Gate-Tunable Negative Refraction of Mid-Infrared Polaritons. Science 2023, 379 (6632), 558–561. 10.1126/science.adf1251. [DOI] [PubMed] [Google Scholar]
- Zhang Q.; Hu G.; Ma W.; Li P.; Krasnok A.; Hillenbrand R.; Alù A.; Qiu C.-W. Interface Nano-Optics with van Der Waals Polaritons. Nature 2021, 597 (7875), 187–195. 10.1038/s41586-021-03581-5. [DOI] [PubMed] [Google Scholar]
- Khatib O.; Bechtel H. A.; Martin M. C.; Raschke M. B.; Carr G. L. Far Infrared Synchrotron Near-Field Nanoimaging and Nanospectroscopy. ACS Photonics 2018, 5 (7), 2773–2779. 10.1021/acsphotonics.8b00565. [DOI] [Google Scholar]
- Barcelos I. D.; Bechtel H. A.; de Matos C. J. S.; Bahamon D. A.; Kaestner B.; Maia F. C. B.; Freitas R. O. Probing Polaritons in 2D Materials with Synchrotron Infrared Nanospectroscopy. Adv. Opt. Mater. 2020, 8 (5), 1901091. 10.1002/adom.201901091. [DOI] [Google Scholar]
- Feres F. H.; Mayer R. A.; Wehmeier L.; Maia F. C. B.; Viana E. R.; Malachias A.; Bechtel H. A.; Klopf J. M.; Eng L. M.; Kehr S. C.; González J. C.; Freitas R. O.; Barcelos I. D. Sub-Diffractional Cavity Modes of Terahertz Hyperbolic Phonon Polaritons in Tin Oxide. Nat. Commun. 2021, 12 (1), 1995. 10.1038/s41467-021-22209-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wehmeier L.; Lang D.; Liu Y.; Zhang X.; Winnerl S.; Eng L. M.; Kehr S. C. Polarization-Dependent near-Field Phonon Nanoscopy of Oxides: SrTiO3, LiNbO3, and PbZr0.2Ti0.8O3. Phys. Rev. B 2019, 100 (3), 035444. 10.1103/PhysRevB.100.035444. [DOI] [Google Scholar]
- Wehmeier L.; Nörenberg T.; de Oliveira T. V. A. G.; Klopf J. M.; Yang S.-Y.; Martin L. W.; Ramesh R.; Eng L. M.; Kehr S. C. Phonon-Induced near-Field Resonances in Multiferroic BiFeO3 Thin Films at Infrared and THz Wavelengths. Appl. Phys. Lett. 2020, 116 (7), 071103. 10.1063/1.5133116. [DOI] [Google Scholar]
- Yu S.-J.; Jiang Y.; Roberts J. A.; Huber M. A.; Yao H.; Shi X.; Bechtel H. A.; Gilbert Corder S. N.; Heinz T. F.; Zheng X.; Fan J. A. Ultrahigh-Quality Infrared Polaritonic Resonators Based on Bottom-Up-Synthesized van Der Waals Nanoribbons. ACS Nano 2022, 16 (2), 3027–3035. 10.1021/acsnano.1c10489. [DOI] [PubMed] [Google Scholar]
- Yao Z.; Chen X.; Wehmeier L.; Xu S.; Shao Y.; Zeng Z.; Liu F.; Mcleod A. S.; Gilbert Corder S. N.; Tsuneto M.; Shi W.; Wang Z.; Zheng W.; Bechtel H. A.; Carr G. L.; Martin M. C.; Zettl A.; Basov D. N.; Chen X.; Eng L. M.; Kehr S. C.; Liu M. Probing Subwavelength In-Plane Anisotropy with Antenna-Assisted Infrared Nano-Spectroscopy. Nat. Commun. 2021, 12 (1), 2649. 10.1038/s41467-021-22844-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Longuinhos R.; Cadore A. R.; Bechtel H. A.; J S De Matos C.; Freitas R. O.; Ribeiro-Soares J.; Barcelos I. D. Raman and Far-Infrared Synchrotron Nanospectroscopy of Layered Crystalline Talc: Vibrational Properties, Interlayer Coupling, and Symmetry Crossover. J. Phys. Chem. C 2023, 127 (12), 5876–5885. 10.1021/acs.jpcc.3c00017. [DOI] [Google Scholar]
- Liewald C.; Mastel S.; Hesler J.; Huber A. J.; Hillenbrand R.; Keilmann F. All-Electronic Terahertz Nanoscopy. Optica 2018, 5 (2), 159–163. 10.1364/OPTICA.5.000159. [DOI] [Google Scholar]
- Chen X.; Liu X.; Guo X.; Chen S.; Hu H.; Nikulina E.; Ye X.; Yao Z.; Bechtel H. A.; Martin M. C.; Carr G. L.; Dai Q.; Zhuang S.; Hu Q.; Zhu Y.; Hillenbrand R.; Liu M.; You G. THz Near-Field Imaging of Extreme Subwavelength Metal Structures. ACS Photonics 2020, 7 (3), 687–694. 10.1021/acsphotonics.9b01534. [DOI] [Google Scholar]
- Schäffer S.; Ogolla C. O.; Loth Y.; Haeger T.; Kreusel C.; Runkel M.; Riedl T.; Butz B.; Wigger A. K.; Bolívar P. H. Imaging the Terahertz Nanoscale Conductivity of Polycrystalline CsPbBr3 Perovskite Thin Films. Nano Lett. 2023, 23 (6), 2074–2080. 10.1021/acs.nanolett.2c03214. [DOI] [PubMed] [Google Scholar]
- Aghamiri N. A.; Huth F.; Huber A. J.; Fali A.; Abate Y.; Hillenbrand R.; Abate Y. Hyperspectral Time-Domain Terahertz Nano-Imaging. Opt. Express 2019, 27 (17), 24231–24242. 10.1364/OE.27.024231. [DOI] [PubMed] [Google Scholar]
- Moon K.; Park H.; Kim J.; Do Y.; Lee S.; Lee G.; Kang H.; Han H. Subsurface Nanoimaging by Broadband Terahertz Pulse Near-Field Microscopy. Nano Lett. 2015, 15 (1), 549–552. 10.1021/nl503998v. [DOI] [PubMed] [Google Scholar]
- Stinson H. T.; Sternbach A.; Najera O.; Jing R.; Mcleod A. S.; Slusar T. V.; Mueller A.; Anderegg L.; Kim H. T.; Rozenberg M.; Basov D. N. Imaging the Nanoscale Phase Separation in Vanadium Dioxide Thin Films at Terahertz Frequencies. Nat. Commun. 2018, 9 (1), 3604. 10.1038/s41467-018-05998-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang J.; Chen X.; Mills S.; Ciavatti T.; Yao Z.; Mescall R.; Hu H.; Semenenko V.; Fei Z.; Li H.; Perebeinos V.; Tao H.; Dai Q.; Du X.; Liu M. Terahertz Nanoimaging of Graphene. ACS Photonics 2018, 5 (7), 2645–2651. 10.1021/acsphotonics.8b00190. [DOI] [Google Scholar]
- Jing R.; Shao Y.; Fei Z.; Lo C. F. B.; Vitalone R. A.; Ruta F. L.; Staunton J.; Zheng W. J.-C.; Mcleod A. S.; Sun Z.; Jiang B.; Chen X.; Fogler M. M.; Millis A. J.; Liu M.; Cobden D. H.; Xu X.; Basov D. N. Terahertz Response of Monolayer and Few-Layer WTe2 at the Nanoscale. Nat. Commun. 2021, 12 (1), 5594. 10.1038/s41467-021-23933-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Borak A. Toward Bridging the Terahertz Gap with Silicon-Based Lasers. Science 2005, 308 (5722), 638–639. 10.1126/science.1109831. [DOI] [PubMed] [Google Scholar]
- Huber A. J.; Keilmann F.; Wittborn J.; Aizpurua J.; Hillenbrand R. Terahertz Near-Field Nanoscopy of Mobile Carriers in Single Semiconductor Nanodevices. Nano Lett. 2008, 8 (11), 3766–3770. 10.1021/nl802086x. [DOI] [PubMed] [Google Scholar]
- Pogna E. A. A.; Viti L.; Politano A.; Brambilla M.; Scamarcio G.; Vitiello M. S. Mapping Propagation of Collective Modes in Bi2Se3 and Bi2Te2.2Se0.8 Topological Insulators by near-Field Terahertz Nanoscopy. Nat. Commun. 2021, 12 (1), 6672. 10.1038/s41467-021-26831-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen S.; Bylinkin A.; Wang Z.; Schnell M.; Chandan G.; Li P.; Nikitin A. Y.; Law S.; Hillenbrand R. Real-Space Nanoimaging of THz Polaritons in the Topological Insulator Bi2Se3. Nat. Commun. 2022, 13 (1), 1374. 10.1038/s41467-022-28791-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barnett J.; Wehmeier L.; Heßler A.; Lewin M.; Pries J.; Wuttig M.; Klopf J. M.; Kehr S. C.; Eng L. M.; Taubner T. Far-Infrared Near-Field Optical Imaging and Kelvin Probe Force Microscopy of Laser-Crystallized and -Amorphized Phase Change Material Ge3Sb2Te6. Nano Lett. 2021, 21 (21), 9012–9020. 10.1021/acs.nanolett.1c02353. [DOI] [PubMed] [Google Scholar]
- 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), 113102. 10.1063/1.4943793. [DOI] [Google Scholar]
- Feres F. H.; Barcelos I. D.; Cadore A. R.; Wehmeier L.; Nörenberg T.; Mayer R. A.; Freitas R. O.; Eng L. M.; Kehr S. C.; Maia F. C. B. Graphene Nano-Optics in the Terahertz Gap. Nano Lett. 2023, 23 (9), 3913–3920. 10.1021/acs.nanolett.3c00578. [DOI] [PubMed] [Google Scholar]
- Giordano M. C.; Mastel S.; Liewald C.; Columbo L. L.; Brambilla M.; Viti L.; Politano A.; Zhang K.; Li L.; Davies A. G.; Linfield E. H.; Hillenbrand R.; Keilmann F.; Scamarcio G.; Vitiello M. S. Phase-Resolved Terahertz Self-Detection near-Field Microscopy. Opt. Express 2018, 26 (14), 18423. 10.1364/OE.26.018423. [DOI] [PubMed] [Google Scholar]
- Degl’Innocenti R.; Wallis R.; Wei B.; Xiao L.; Kindness S. J.; Mitrofanov O.; Braeuninger-Weimer P.; Hofmann S.; Beere H. E.; Ritchie D. A. Terahertz Nanoscopy of Plasmonic Resonances with a Quantum Cascade Laser. ACS Photonics 2017, 4 (9), 2150–2157. 10.1021/acsphotonics.7b00687. [DOI] [Google Scholar]
- Knoll B.; Keilmann F. Enhanced Dielectric Contrast in Scattering-Type Scanning near-Field Optical Microscopy. Opt. Commun. 2000, 182 (4–6), 321–328. 10.1016/S0030-4018(00)00826-9. [DOI] [Google Scholar]
- Amarie S.; Keilmann F. Broadband-Infrared Assessment of Phonon Resonance in Scattering-Type near-Field Microscopy. Phys. Rev. B 2011, 83 (4), 045404. 10.1103/PhysRevB.83.045404. [DOI] [Google Scholar]
- Chu J.; Xu S.; Tang D. Energy Gap versus Alloy Composition and Temperature in Hg1-xCdxTe. Appl. Phys. Lett. 1983, 43 (11), 1064–1066. 10.1063/1.94237. [DOI] [Google Scholar]
- Dumas P.; Carr G. L.. Infrared Spectroscopy and Spectro-Microscopy with Synchrotron Radiation; Brookhaven National Lab: Upton, NY, United States, 2020. [Google Scholar]
- Schmit J. L.; Stelzer E. L. Temperature and Alloy Compositional Dependences of the Energy Gap of Hg1-xCdxTe. J. Appl. Phys. 1969, 40 (12), 4865–4869. 10.1063/1.1657304. [DOI] [Google Scholar]
- Scott M. W. Energy Gap in Hg1-xCdxTe by Optical Absorption. J. Appl. Phys. 1969, 40 (10), 4077–4081. 10.1063/1.1657147. [DOI] [Google Scholar]
- Faye M.; Bordessoule M.; Kanouté B.; Brubach J.-B.; Roy P.; Manceron L. Improved Mid Infrared Detector for High Spectral or Spatial Resolution and Synchrotron Radiation Use. Rev. Sci. Instrum. 2016, 87 (6), 063119. 10.1063/1.4954405. [DOI] [PubMed] [Google Scholar]
- Querry M. R.. Optical Constants of Minerals and Other Materials from the Millimeter to the Ultraviolet;, Contractor Report CRDEC-CR-88009; 1987. https://apps.dtic.mil/sti/citations/ADA192210 (accessed 2023–06–25).
- Kaiser W.; Spitzer W. G.; Kaiser R. H.; Howarth L. E. Infrared Properties of CaF2, SrF2, and BaF2. Phys. Rev. 1962, 127 (6), 1950–1954. 10.1103/PhysRev.127.1950. [DOI] [Google Scholar]
- Hillenbrand R.; Taubner T.; Keilmann F. Phonon-Enhanced Light-Matter Interaction at the Nanometre Scale. Nature 2002, 418 (6894), 159–162. 10.1038/nature00899. [DOI] [PubMed] [Google Scholar]
- Taubner T.; Hillenbrand R.; Keilmann F. Performance of Visible and Mid-Infrared Scattering-Type near-Field Optical Microscopes. J. Microsc. 2003, 210 (3), 311–314. 10.1046/j.1365-2818.2003.01164.x. [DOI] [PubMed] [Google Scholar]
- Keilmann F.; Hillenbrand R. Near-Field Microscopy by Elastic Light Scattering from a Tip. Philos. Trans. R. Soc., A 2004, 362 (1817), 787–805. 10.1098/rsta.2003.1347. [DOI] [PubMed] [Google Scholar]
- Deneuville A.; Tanner D.; Holloway P. H. Optical Constants of ZnSe in the Far Infrared. Phys. Rev. B 1991, 43 (8), 6544–6550. 10.1103/PhysRevB.43.6544. [DOI] [PubMed] [Google Scholar]
- McArdle P.; Lahneman D. J.; Biswas A.; Keilmann F.; Qazilbash M. M. Near-Field Infrared Nanospectroscopy of Surface Phonon-Polariton Resonances. Phys. Rev. Res. 2020, 2 (2), 023272. 10.1103/PhysRevResearch.2.023272. [DOI] [Google Scholar]
- Wang H.; Wang L.; Jakob D. S.; Xu X. G. Tomographic and Multimodal Scattering-Type Scanning near-Field Optical Microscopy with Peak Force Tapping Mode. Nat. Commun. 2018, 9 (1), 2005. 10.1038/s41467-018-04403-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ulaganathan R. K.; Lu Y.-Y.; Kuo C.-J.; Tamalampudi S. R.; Sankar R.; Boopathi K. M.; Anand A.; Yadav K.; Mathew R. J.; Liu C.-R.; Chou F. C.; Chen Y.-T. High Photosensitivity and Broad Spectral Response of Multi-Layered Germanium Sulfide Transistors. Nanoscale 2016, 8 (4), 2284–2292. 10.1039/C5NR05988G. [DOI] [PubMed] [Google Scholar]
- Zhang S.; Wang N.; Liu S.; Huang S.; Zhou W.; Cai B.; Xie M.; Yang Q.; Chen X.; Zeng H. Two-Dimensional GeS with Tunable Electronic Properties via External Electric Field and Strain. Nanotechnology 2016, 27 (27), 274001. 10.1088/0957-4484/27/27/274001. [DOI] [PubMed] [Google Scholar]
- Ding G.; Gao G.; Yao K. High-Efficient Thermoelectric Materials: The Case of Orthorhombic IV-VI Compounds. Sci. Rep. 2015, 5 (1), 9567. 10.1038/srep09567. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fei R.; Kang W.; Yang L. Ferroelectricity and Phase Transitions in Monolayer Group-IV Monochalcogenides. Phys. Rev. Lett. 2016, 117 (9), 097601. 10.1103/PhysRevLett.117.097601. [DOI] [PubMed] [Google Scholar]
- Sutter P.; Wimer S.; Sutter E. Chiral Twisted van Der Waals Nanowires. Nature 2019, 570 (7761), 354–357. 10.1038/s41586-019-1147-x. [DOI] [PubMed] [Google Scholar]
- Wiley J. D.; Buckel W. J.; Schmidt R. L. Infrared Reflectivity and Raman Scattering in GeS. Phys. Rev. B 1976, 13 (6), 2489–2496. 10.1103/PhysRevB.13.2489. [DOI] [Google Scholar]
- Yu L.-M.; Degiovanni A.; Thiry P. A.; Ghijsen J.; Caudano R.; Lambin Ph. Infrared Optical Constants of Orthorhombic IV-VI Lamellar Semiconductors Refined by a Combined Study Using Optical and Electronic Spectroscopies. Phys. Rev. B 1993, 47 (24), 16222–16228. 10.1103/PhysRevB.47.16222. [DOI] [PubMed] [Google Scholar]
- Mihajlovic P.; Nikolic P. M.; Hughes O. H. Infrared Reflectivity of GeS. J. Phys. C: Solid State Phys. 1976, 9 (21), L599–L602. 10.1088/0022-3719/9/21/002. [DOI] [Google Scholar]
- Alonso-González P.; Nikitin A. Y.; Gao Y.; Woessner A.; Lundeberg M. B.; Principi A.; Forcellini N.; Yan W.; Vélez S.; Huber A. J.; Watanabe K.; Taniguchi T.; Casanova F.; Hueso L. E.; Polini M.; Hone J.; Koppens F. H. L.; Hillenbrand R. Acoustic Terahertz Graphene Plasmons Revealed by Photocurrent Nanoscopy. Nat. Nanotechnol. 2017, 12 (1), 31–35. 10.1038/nnano.2016.185. [DOI] [PubMed] [Google Scholar]
- Döring J.; Lang D.; Wehmeier L.; Kuschewski F.; Nörenberg T.; Kehr S. C.; Eng L. M. Low-Temperature Nanospectroscopy of the Structural Ferroelectric Phases in Single-Crystalline Barium Titanate. Nanoscale 2018, 10 (37), 18074–18079. 10.1039/C8NR04081H. [DOI] [PubMed] [Google Scholar]
- Yang H. U.; Hebestreit E.; Josberger E. E.; Raschke M. B. A Cryogenic Scattering-Type Scanning near-Field Optical Microscope. Rev. Sci. Instrum. 2013, 84 (2), 023701. 10.1063/1.4789428. [DOI] [PubMed] [Google Scholar]
- McLeod A. S.; van Heumen E.; Ramirez J. G.; Wang S.; Saerbeck T.; Guenon S.; Goldflam M.; Anderegg L.; Kelly P.; Mueller A.; Liu M. K.; Schuller I. K.; Basov D. N. Nanotextured Phase Coexistence in the Correlated Insulator V2O3. Nat. Phys. 2017, 13 (1), 80–86. 10.1038/nphys3882. [DOI] [Google Scholar]
- Luo W.; Boselli M.; Poumirol J.-M.; Ardizzone I.; Teyssier J.; van der Marel D.; Gariglio S.; Triscone J.-M.; Kuzmenko A. B. High Sensitivity Variable-Temperature Infrared Nanoscopy of Conducting Oxide Interfaces. Nat. Commun. 2019, 10 (1), 2774. 10.1038/s41467-019-10672-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dong Y.; Xiong L.; Phinney I. Y.; Sun Z.; Jing R.; McLeod A. S.; Zhang S.; Liu S.; Ruta F. L.; Gao H.; Dong Z.; Pan R.; Edgar J. H.; Jarillo-Herrero P.; Levitov L. S.; Millis A. J.; Fogler M. M.; Bandurin D. A.; Basov D. N. Fizeau Drag in Graphene Plasmonics. Nature 2021, 594 (7864), 513–516. 10.1038/s41586-021-03640-x. [DOI] [PubMed] [Google Scholar]
- Zhao W.; Zhao S.; Li H.; Wang S.; Wang S.; Utama M. I. B.; Kahn S.; Jiang Y.; Xiao X.; Yoo S.; Watanabe K.; Taniguchi T.; Zettl A.; Wang F. Efficient Fizeau Drag from Dirac Electrons in Monolayer Graphene. Nature 2021, 594 (7864), 517–521. 10.1038/s41586-021-03574-4. [DOI] [PubMed] [Google Scholar]
- Dapolito M.; Tsuneto M.; Zheng W.; Wehmeier L.; Xu S.; Chen X.; Sun J.; Du Z.; Shao Y.; Jing R.; Zhang S.; Bercher A.; Dong Y.; Halbertal D.; Ravindran V.; Zhou Z.; Petrovic M.; Gozar A.; Carr G. L.; Li Q.; Kuzmenko A. B.; Fogler M. M.; Basov D. N.; Du X.; Liu M. Infrared Nano-Imaging of Dirac Magnetoexcitons in Graphene. Nat. Nanotechnol. 2023, 1–7. 10.1038/s41565-023-01488-y. [DOI] [PubMed] [Google Scholar]
- Kim R. H. J.; Park J.-M.; Haeuser S. J.; Luo L.; Wang J. A Sub-2 Kelvin Cryogenic Magneto-Terahertz Scattering-Type Scanning near-Field Optical Microscope (Cm-THz-sSNOM). Review of Scientific Instruments 2023, 94 (4), 043702. 10.1063/5.0130680. [DOI] [PubMed] [Google Scholar]
- Sun Z.; Fogler M. M.; Basov D. N.; Millis A. J. Collective Modes and Terahertz Near-Field Response of Superconductors. Phys. Rev. Res. 2020, 2 (2), 023413. 10.1103/PhysRevResearch.2.023413. [DOI] [Google Scholar]
- Park S.; Yager K.. Roll-to-Roll Mechanized Exfoliator and Automatic 2D Materials Transfer and Layering System. Patent application number PCT/US2022/040777, WO 2023023260 A2, August 18, 2022. https://patentimages.storage.googleapis.com/f2/f1/5c/46f830a8af203b/WO2023023260A2.pdf (accessed 2023–06–26).




