Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (175)

Search Parameters:
Keywords = THz time-domain spectroscopy

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 2443 KiB  
Article
Exploring the Impact of 3D Printing Parameters on the THz Optical Characteristics of COC Material
by Mateusz Kaluza, Michal Walczakowski and Agnieszka Siemion
Materials 2024, 17(20), 5104; https://doi.org/10.3390/ma17205104 - 19 Oct 2024
Viewed by 643
Abstract
In terahertz (THz) optical systems, polymer-based manufacturing processes are employed to ensure product quality and the material performance necessary for proper system maintenance. Therefore, the precise manufacturing of system components, such as optical elements, is crucial for the optimal functioning of the systems. [...] Read more.
In terahertz (THz) optical systems, polymer-based manufacturing processes are employed to ensure product quality and the material performance necessary for proper system maintenance. Therefore, the precise manufacturing of system components, such as optical elements, is crucial for the optimal functioning of the systems. In this study, the authors investigated the impact of various 3D printing parameters using fused deposition modeling (FDM) on the optical properties of manufactured structures within the THz radiation range. The measurements were conducted on 3D printed samples using highly transparent and biocompatible cyclic olefin copolymer (COC), which may find applications in THz passive optics for “in vivo” measurements. The results of this study indicate that certain printing parameters significantly affect the optical behavior of the fabricated structures. The improperly configured printing parameters result in the worsening of THz optical properties. This is proved through a significant change in the refractive index value and undesirable increase in the absorption coefficient value. Furthermore, such misconfigurations may lead to the occurrence of defects within the printed structures. Finally, the recommended printing parameters, which improve the optical performance of the manufactured structures are presented. Full article
(This article belongs to the Special Issue Polymers, Processing and Sustainability)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Spatial model of the cylindrical sample sliced using reference printing parameters with various infill patterns. The colors of the sliced objects indicate the predicted printing speeds achieved by the nozzle during the printing process. (<b>a</b>) Three-dimensional model of the sample; (<b>b</b>) sliced model with (aligned) rectilinear infill; (<b>c</b>) sliced model with concentric infill; (<b>d</b>) sliced model with octagram spiral infill.</p>
Full article ">Figure 2
<p>Exemplary 3D printed samples manufactured from COC material using FDM technology, subsequently examined with THz TDS; (<b>a</b>,<b>b</b>) reference samples without visible deformations; (<b>c</b>,<b>d</b>) samples manufactured with too-thick layers that result in deformations (marked with a red outline and red arrows) in the central part of the samples; (<b>e</b>,<b>f</b>) samples printed with an overly high material flow rate that results in overflowing material leading to the deformations (marked with a red outline and red arrows) visible on the top surface and the edges of the samples.</p>
Full article ">Figure 3
<p>Three-dimensional printed samples manufactured from COC material using FDM technology, subsequently examined with THz TDS. The photographs correspond to the samples manufactured with different infill patterns: (<b>a</b>) rectilinear; (<b>b</b>) aligned rectilinear; (<b>c</b>) concentric; (<b>d</b>) octagram spiral; and (<b>e</b>) the cuboid sample printed in the vertical position with rectilinear infill pattern.</p>
Full article ">Figure 4
<p>THz TDS measurement results obtained from two different reference samples, each measured twice at appropriate time intervals. Each measurement was performed with a corresponding reference measurement (signal without the sample) and covered different sample areas; (<b>a</b>) the absorption coefficient <math display="inline"><semantics> <mi>α</mi> </semantics></math> (solid lines) and refractive index <span class="html-italic">n</span> (dashed lines) in the THz frequency domain, displayed over a broader data range common to all measurements presented in this study; (<b>b</b>) the absorption coefficient in the THz frequency domain; (<b>c</b>) the refractive index in the THz frequency domain.</p>
Full article ">Figure 5
<p>THz optical properties of COC samples 3D printed using FDM, illustrating the impact of changes in various printing parameters. The data show the measured absorption coefficient <math display="inline"><semantics> <mi>α</mi> </semantics></math> (solid lines) and refractive index <span class="html-italic">n</span> (dashed lines) in the THz frequency domain, indicating the influence of the following printing parameters: (<b>a</b>) nozzle temperatures; (<b>b</b>) nozzle diameters; (<b>c</b>) printing speed; (<b>d</b>) infill patterns. The change in these parameters does not change the THz optical parameters of the samples.</p>
Full article ">Figure 6
<p>THz optical properties of COC samples 3D printed using FDM, illustrating the impact of changes in various printing parameters. The data show the measured absorption coefficient <math display="inline"><semantics> <mi>α</mi> </semantics></math> (solid lines) and refractive index <span class="html-italic">n</span> (dashed lines) in the THz frequency domain, indicating the influence of the following printing parameters: (<b>a</b>) extrusion width; (<b>b</b>) cooling level; (<b>c</b>) material flow ratio; (<b>d</b>) line thickness. The change in these parameters introduce changes in the THz optical parameters of the samples.</p>
Full article ">
16 pages, 3710 KiB  
Article
Experimental Analysis of Terahertz Wave Scattering Characteristics of Simulated Lunar Regolith Surface
by Suyun Wang and Kazuma Hiramatsu
Remote Sens. 2024, 16(20), 3819; https://doi.org/10.3390/rs16203819 - 14 Oct 2024
Viewed by 605
Abstract
This study investigates terahertz (THz) wave scattering from a simulated lunar regolith surface, with a focus on the Brewster feature, backscattering, and bistatic scattering within the 325 to 500 GHz range. We employed a generalized power-law spectrum to characterize surface roughness and fabricated [...] Read more.
This study investigates terahertz (THz) wave scattering from a simulated lunar regolith surface, with a focus on the Brewster feature, backscattering, and bistatic scattering within the 325 to 500 GHz range. We employed a generalized power-law spectrum to characterize surface roughness and fabricated Gaussian correlated surfaces from Durable Resin V2 using 3D printing technology. The complex dielectric permittivity of these materials was determined through THz time-domain spectroscopy (THz-TDS). Our experimental setup comprised a vector network analyzer (VNA) equipped with dual waveguide frequency extenders for the WR-2.2 band, transmitter and receiver modules, polarizing components, and a scattering chamber. We systematically analyzed the effects of root-mean-square (RMS) height, correlation length, dielectric constant, frequency, polarization, and observation angle on THz scattering. The findings highlight the significant impact of surface roughness on the Brewster angle shift, backscattering, and bistatic scattering. These insights are crucial for refining theoretical models and developing algorithms to retrieve physical parameters for lunar and other celestial explorations. Full article
(This article belongs to the Special Issue Future of Lunar Exploration)
Show Figures

Figure 1

Figure 1
<p>The geometry of wave scattering from rough surface.</p>
Full article ">Figure 2
<p>The rough surface samples are designed with specified RMS heights and correlation lengths.</p>
Full article ">Figure 3
<p>The measured dielectric constant of the material by THz-TDS.</p>
Full article ">Figure 4
<p>The roughness validation of one selected rough surface with an RMS height of 0.8<math display="inline"><semantics> <mi>λ</mi> </semantics></math> and a correlation length of 2<math display="inline"><semantics> <mi>λ</mi> </semantics></math>.</p>
Full article ">Figure 5
<p>The experiment configuration.</p>
Full article ">Figure 6
<p>The polarizer consists of three reflectors.</p>
Full article ">Figure 7
<p>The comparison between the simulated and experimental HH and VV reflections from a flat surface with a dielectric constant of <math display="inline"><semantics> <mrow> <mn>2.597</mn> <mo>+</mo> <mi>j</mi> <mn>0.165</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>The frequency effect on THz scattering from rough surface.</p>
Full article ">Figure 9
<p>The correlation length effect on THz scattering from rough surface. (<b>a</b>) <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 0.8<math display="inline"><semantics> <mi>λ</mi> </semantics></math> (<b>b</b>) <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 0.1<math display="inline"><semantics> <mi>λ</mi> </semantics></math>.</p>
Full article ">Figure 10
<p>The RMS height effect on THz scattering from rough surface. (<b>a</b>) <span class="html-italic">l</span> = 2<math display="inline"><semantics> <mi>λ</mi> </semantics></math> (<b>b</b>) <span class="html-italic">l</span> = 0.4<math display="inline"><semantics> <mi>λ</mi> </semantics></math>.</p>
Full article ">Figure 11
<p>Comparison of bistatic scattering from flat and rough surfaces with RMS heights of 0.5<math display="inline"><semantics> <mi>λ</mi> </semantics></math> and 0.8<math display="inline"><semantics> <mi>λ</mi> </semantics></math> and a fixed correlation length of 2<math display="inline"><semantics> <mi>λ</mi> </semantics></math> at incident angles of 30°, 45°, and 60° for both HH and VV polarizations.</p>
Full article ">Figure 12
<p>Comparison of bistatic scattering from flat and rough surfaces with RMS heights of 0.1<math display="inline"><semantics> <mi>λ</mi> </semantics></math> and 0.08<math display="inline"><semantics> <mi>λ</mi> </semantics></math> and a fixed correlation length of 0.4<math display="inline"><semantics> <mi>λ</mi> </semantics></math> at incident angles of 30°, 45°, and 60° for both HH and VV polarizations.</p>
Full article ">Figure 13
<p>Comparison of bistatic scattering from the flat surface and rough surface with different correlation lengths of 2<math display="inline"><semantics> <mi>λ</mi> </semantics></math>, 4<math display="inline"><semantics> <mi>λ</mi> </semantics></math> and 6<math display="inline"><semantics> <mi>λ</mi> </semantics></math> and a fixed RMS height of 0.8<math display="inline"><semantics> <mi>λ</mi> </semantics></math> at the incident angle of 30°, 45° and 60° for HH and VV polarizations.</p>
Full article ">Figure 14
<p>Comparison of bistatic scattering from rough surfaces with correlation lengths of 1.5<math display="inline"><semantics> <mi>λ</mi> </semantics></math>, 1<math display="inline"><semantics> <mi>λ</mi> </semantics></math>, and 0.4<math display="inline"><semantics> <mi>λ</mi> </semantics></math> and a fixed RMS height of 0.1<math display="inline"><semantics> <mi>λ</mi> </semantics></math> at incident angles of 30°, 45°, and 60° for both HH and VV polarizations.</p>
Full article ">Figure 15
<p>Comparison of three different incident angles 30°, 45°, and 60° for HH and VV polarizations from a Gaussian correlated surface of <span class="html-italic">l</span> = 1.0 <math display="inline"><semantics> <mi>λ</mi> </semantics></math>, <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 0.1 <math display="inline"><semantics> <mi>λ</mi> </semantics></math>. ((<b>left</b>): VV polarization, (<b>right</b>): HH polarization).</p>
Full article ">Figure 16
<p>Comparison of three different incident angles 30°, 45°, and 60° for HH and VV polarizations from a Gaussian correlated surface of <span class="html-italic">l</span> = 2.0 <math display="inline"><semantics> <mi>λ</mi> </semantics></math>, <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 0.5 <math display="inline"><semantics> <mi>λ</mi> </semantics></math>. ((<b>left</b>): VV polarization, (<b>right</b>): HH polarization).</p>
Full article ">
10 pages, 3134 KiB  
Communication
All-Dielectric Metasurface-Based Terahertz Molecular Fingerprint Sensor for Trace Cinnamoylglycine Detection
by Qiyuan Xu, Mingjun Sun, Weijin Wang and Yanpeng Shi
Biosensors 2024, 14(9), 440; https://doi.org/10.3390/bios14090440 - 13 Sep 2024
Viewed by 929
Abstract
Terahertz (THZ) spectroscopy has emerged as a superior label-free sensing technology in the detection, identification, and quantification of biomolecules in various biological samples. However, the limitations in identification and discrimination sensitivity of current methods impede the wider adoption of this technology. In this [...] Read more.
Terahertz (THZ) spectroscopy has emerged as a superior label-free sensing technology in the detection, identification, and quantification of biomolecules in various biological samples. However, the limitations in identification and discrimination sensitivity of current methods impede the wider adoption of this technology. In this article, a meticulously designed metasurface is proposed for molecular fingerprint enhancement, consisting of a periodic array of lithium tantalate triangular prism tetramers arranged in a square quartz lattice. The physical mechanism is explained by the finite-difference time-domain (FDTD) method. The metasurface achieves a high quality factor (Q-factor) of 231 and demonstrates excellent THz sensing capabilities with a figure of merit (FoM) of 609. By varying the incident angle of the THz wave, the molecular fingerprint signal is strengthened, enabling the highly sensitive detection of trace amounts of analyte. Consequently, cinnamoylglycine can be detected with a sensitivity limit as low as 1.23 μg·cm2. This study offers critical insights into the advanced application of THz waves in biomedicine, particularly for the detection of urinary biomarkers in various diseases, including gestational diabetes mellitus (GDM). Full article
(This article belongs to the Special Issue Photonics for Bioapplications: Sensors and Technology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The structural diagram of the all-dielectric metasurface, illustrating the periodic arrangement of the high-index triangular prism tetramer based on the quartz substrate; (<b>b</b>) a unit cell of the periodic structure with a y-polarized source incident downwards in the z direction; (<b>c</b>) the main view of the unit cell (y–z plane) and corresponding parameters.</p>
Full article ">Figure 2
<p>(<b>a</b>) Transmission spectra for x-polarized and y-polarized incident waves at 0°; (<b>b</b>) transmission spectra for x-polarized and y-polarized incident waves at 37°; (<b>c</b>) the electric and magnetic field distribution measured at the surface of the quartz substrate at vertical incidence. The left and right figures correspond to the x-polarized and y-polarized incident wave, respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>) Transmission spectra at different incident angles without any analyte; (<b>b</b>) the experimentally measured refractive index (n) and extinction coefficient (k) of cinnamoylglycine across the relevant frequency range; (<b>c</b>) transmission spectra at different incident angles with a <math display="inline"><semantics> <mrow> <mn>1</mn> <mo> </mo> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> thick layer of analyte; (<b>d</b>) the electric field distribution measured at the substrate surface in the x–y plane at 0.487 THz for specific incident angles, corresponding to the transmission spectra shown in (<b>c</b>), respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>) Comprehensive transmission spectra without any analyte, with the incident angle ranging from 13° to 70°. Specifically, the rightmost line represents the transmission curve for an angle of 13°, while the leftmost line corresponds to 70°; (<b>b</b>) comprehensive transmission spectra with <math display="inline"><semantics> <mrow> <mn>1</mn> <mo> </mo> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> thick cinnamoylglycine, with the incident angle ranging from 13° to 62°. The corresponding envelope curve has been plotted by red line in the figure.</p>
Full article ">Figure 5
<p>(<b>a</b>) Transmission envelope curves for analytes of varying thicknesses; (<b>b</b>) the relationship between the thickness of the analyte and the transmission at <math display="inline"><semantics> <mrow> <mn>0.487</mn> <mo> </mo> <mi mathvariant="normal">T</mi> <mi mathvariant="normal">H</mi> <mi mathvariant="normal">z</mi> </mrow> </semantics></math>.</p>
Full article ">
14 pages, 3670 KiB  
Article
Novel THz Metasurface Biosensor for High-Sensitivity Detection of Vitamin C and Vitamin B9
by Ningyi Wang, Bingwei Liu, Xu Wu and Yan Peng
Photonics 2024, 11(9), 820; https://doi.org/10.3390/photonics11090820 - 30 Aug 2024
Viewed by 806
Abstract
Vitamin C (VC) and Vitamin B9 (VB9) are essential micronutrients integral to numerous biological functions and critical for maintaining human health. The rapid detection of these vitamins is important for verifying nutritional supplements and aiding in clinical diagnoses. This study combined terahertz time-domain [...] Read more.
Vitamin C (VC) and Vitamin B9 (VB9) are essential micronutrients integral to numerous biological functions and critical for maintaining human health. The rapid detection of these vitamins is important for verifying nutritional supplements and aiding in clinical diagnoses. This study combined terahertz time-domain spectroscopy (THz-TDS) with metasurface technology to develop a fast, sensitive, and non-destructive detection method for VC and VB9. Firstly, we determined the characteristic absorption peaks and molecular vibration modes of VC and VB9 within the 0.5–4.0 THz range through quantum chemical calculation and THz-TDS measurement. Then, we designed and fabricated a metasurface biosensor to match its resonance peak with the communal peak of VC and VB9, enhancing the interaction between THz waves and these vitamins. Using this biosensor, we analyzed solutions with different concentrations of VC and VB9. An increase in vitamin concentrations resulted in frequency shifts in the THz resonance peak. Quantifiable relationships between frequency shifts and the vitamin concentrations were established. The detection limits achieved were 158.82 ng/µL for VC and 353.57 ng/µL for VB9, respectively. This method not only demonstrates high sensitivity but also simplifies the operational process, offering an innovative tool for applications in food safety monitoring and clinical diagnostics. Full article
(This article belongs to the Section Optical Interaction Science)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Experimental setup and (<b>b</b>) its operational principle of the THz spectrometer.</p>
Full article ">Figure 2
<p>(<b>a</b>) Molecular structure and (<b>b</b>) theoretical THz spectra of VC; (<b>c</b>) molecular structure and (<b>d</b>) theoretical THz spectra of VB9.</p>
Full article ">Figure 3
<p>THz absorption spectrum of (<b>a</b>) VC and (<b>b</b>) VB9.</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic of the unit structure; (<b>b</b>) periodic structure of the metasurface biosensor; (<b>c</b>) electric field distribution diagram; and (<b>d</b>) theoretical spectrum of the metasurface biosensor simulated using COMSOL Multiphysics 6.1 software.</p>
Full article ">Figure 5
<p>(<b>a</b>) Optical micrograph and actual geometric parameters of the fabricated metasurface biosensor, and (<b>b</b>) measured spectrum of the metasurface biosensor in the 1.5–2.0 THz range.</p>
Full article ">Figure 6
<p>(<b>a</b>) THz absorption spectra for varying VC concentrations, and (<b>b</b>) the relationship between the frequency shift of the metasurface and VC concentrations; (<b>c</b>) THz absorption spectra for varying VB9 concentrations, and (<b>d</b>) the relationship between the frequency shift of the metasurface and VB9 concentrations.</p>
Full article ">
13 pages, 1902 KiB  
Article
Design of an Optimized Terahertz Time-Domain Spectroscopy System Pumped by a 30 W Yb:KGW Source at a 100 kHz Repetition Rate with 245 fs Pulse Duration
by Lennart Hirsch, Dionysis Adamou, Daniele Faccio, Marco Peccianti and Matteo Clerici
Appl. Sci. 2024, 14(15), 6688; https://doi.org/10.3390/app14156688 - 31 Jul 2024
Viewed by 930
Abstract
Ytterbium laser sources are state-of-the-art systems that are increasingly replacing Ti:Sapphire lasers in most applications requiring high repetition rate pulse trains. However, extending these laser sources to THz Time-Domain Spectroscopy (THz-TDS) poses several challenges not encountered in conventional, lower-power systems. These challenges include [...] Read more.
Ytterbium laser sources are state-of-the-art systems that are increasingly replacing Ti:Sapphire lasers in most applications requiring high repetition rate pulse trains. However, extending these laser sources to THz Time-Domain Spectroscopy (THz-TDS) poses several challenges not encountered in conventional, lower-power systems. These challenges include pump rejection, thermal lensing in nonlinear media, and pulse durations exceeding 100 fs, which consequently limit the detection bandwidth in TDS applications. In this article, we describe our design of a THz-TDS beamline that seeks to address these issues. We report on the effectiveness of temperature controlling the Gallium Phosphide (GaP) used to generate the THz radiation and its impact on increasing the generation efficiency and aiding pump rejection while avoiding thermal distortions of the residual pump laser beam. We detail our approach to pump rejection, which can be implemented with off-the-shelf products and minimal customization. Finally, we describe our solution based on a commercial optical parametric amplifier to obtain a temporally compressed probe pulse of 55 fs duration. Our study will prove useful to the increasing number of laboratories seeking to move from the high-energy, low-power THz time-domain spectroscopy systems based on Ti:Sapphire lasers, to medium-energy, high-power systems driven by Yb-doped lasers. Full article
(This article belongs to the Special Issue Applications of Terahertz Sensing and Imaging)
Show Figures

Figure 1

Figure 1
<p>Conventional THz-TDS beamline. BS: Beamsplitter; OR: Optical rectification via GaP crystal; EO: electro-optic wafer; QWP: quarter-wave plate; PBS: polarising beam splitter (Wollaston prism); BD: balanced detector.</p>
Full article ">Figure 2
<p>Generation efficiency for increasing pump power. The optimal condition in our experiment was reached with a 3.5 mm <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <msup> <mi>e</mi> <mn>2</mn> </msup> </mrow> </semantics></math> diameter pump beam (red curve), for which the conversion efficiency was saturated at 30 W, the maximum laser power available for pumping the rectification process. As a comparison, we show that the conversion efficiency saturates at lower powers for smaller beam sizes (blue and green data).</p>
Full article ">Figure 3
<p>A schematic of our low-loss THz-TDS beamline utilizing a long focal length parabolic and a flat pump rejection mirror.</p>
Full article ">Figure 4
<p>Simulated intensity (<b>a</b>) and phase (<b>b</b>) profiles of the THz beam in the plane of the EO crystal. The beam measures 330 × 316 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m.</p>
Full article ">Figure 5
<p>Efficiency of THz generation for cooled (blue) and uncooled GaP (red).</p>
Full article ">Figure 6
<p>Comparison of cooled vs. uncooled GaP crystal. (<b>a</b>) Thermal image of crystal surface after 7 min of exposure to 25 W average pump power. (<b>b</b>) Thermal image of crystal surface in water-cooled mount after 7 min of exposure to 25 W average pump power. (<b>c</b>) Far-field image of 23.5 W pump beam after transmission through uncooled crystal. Substantial thermal lensing is evident, resulting in beam aberrations and an increase of beam size. (<b>d</b>) Far-field image of 23.5 W pump beam after transmission through water-cooled crystal. Minimal thermal lensing is observed.</p>
Full article ">Figure 7
<p>THz traces taken with cooled (blue) and uncooled (red) GaP at a pump power of 28 W. The scattered light resulting from thermal lensing leads to a reduction in the SNR in THz traces taken under these conditions.</p>
Full article ">Figure 8
<p>FROG scan of the compressed probe pulse. (<b>a</b>) shows the amplitude of the pulse in the time domain. The dashed red line indicates the phase of the pulse. (<b>b</b>) shows the frequency spectrum of the pulse. Again, the dashed blue line indicates the phase of the pulse. The dashed black lines indicate the position of the zero phase (shown on the right ordinate). The lines indicating the phase of the pulse are almost ideally flat across the pulse profile, indicating the pulse is free of chirp.</p>
Full article ">Figure 9
<p>Comparison of THz traces taken with a 55 fs probe and a 245 fs probe. (<b>a</b>) Time-domain traces were taken with compressed (red) vs. uncompressed (blue) probe pulses. The improved resolution of the trace taken with a compressed probe is visible. The initial rise in the region between <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>0.1</mn> </mrow> </semantics></math> ps and <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>0.4</mn> </mrow> </semantics></math> ps shows a sharper definition in the compressed case. (<b>b</b>) Experimental (solid) and simulated (dashed) frequency spectra of the THz field probed with compressed (red) and uncompressed (blue) probe pulses. The THz trace recorded with a compressed probe exhibits ≃0.54 THz greater FWHM bandwidth compared to the uncompressed case. We note that the figure has been zoomed in to highlight the difference in detection bandwidth between the compressed and uncompressed probe pulse; the noise floor is not visible in the figure and lies at ∼−60 dB.</p>
Full article ">
22 pages, 4157 KiB  
Article
Characterization of Indium Tin Oxide (ITO) Thin Films towards Terahertz (THz) Functional Device Applications
by Anup Kumar Sahoo, Wei-Chen Au and Ci-Ling Pan
Coatings 2024, 14(7), 895; https://doi.org/10.3390/coatings14070895 - 17 Jul 2024
Viewed by 1067
Abstract
In this study, we explored the manipulation of optical properties in the terahertz (THz) frequency band of radio-frequency (RF) sputtered indium tin oxide (ITO) thin films on highly resistive silicon substrate by rapid thermal annealing (RTA). The optical constants of as-deposited and RTA-processed [...] Read more.
In this study, we explored the manipulation of optical properties in the terahertz (THz) frequency band of radio-frequency (RF) sputtered indium tin oxide (ITO) thin films on highly resistive silicon substrate by rapid thermal annealing (RTA). The optical constants of as-deposited and RTA-processed ITO films annealed at 400 °C, 600 °C and 800 °C are determined in the frequency range of 0.2 to 1.0 THz. The transmittance can be changed from ~27% for as-deposited to ~10% and ~39% for ITO films heat-treated at different annealing temperatures (Ta’s). Such variations of optical properties in the far infrared for the samples under study are correlated with their mobility and carrier concentration, which are extracted from Drude–Smith modeling of THz conductivity with plasma frequency, scattering time and the c-parameters as fitting parameters. Resistivities of the films are in the range of 10−3 to 10−4 Ω-cm, confirming that annealed ITO films can potentially be used as transparent conducting electrodes for photonic devices operating at THz frequencies. The highest mobility, μ = 47 cm2/V∙s, with carrier concentration, Nc = 1.31 × 1021 cm−3, was observed for ITO films annealed at Ta = 600 °C. The scattering times of the samples were in the range of 8–21 fs, with c-values of −0.63 to −0.87, indicating strong backscattering of the carriers, mainly by grain boundaries in the polycrystalline film. To better understand the nature of these films, we have also characterized the surface morphology, microscopic structural properties and chemical composition of as-deposited and RTA-processed ITO thin films. For comparison, we have summarized the optical properties of ITO films sputtered onto fused silica substrates, as-deposited and RTA-annealed, in the visible transparency window of 400–800 nm. The optical bandgaps of the ITO thin films were evaluated with a Tauc plot from the absorption spectra. Full article
(This article belongs to the Special Issue Thermoelectric Thin Films for Thermal Energy Harvesting)
Show Figures

Figure 1

Figure 1
<p>The SEM images of as-deposited ITO thin film with (<b>a</b>) top view (<b>b</b>) cross-sectional view, ITO thin film with RTA-treated at 400 °C with (<b>c</b>) top view (<b>d</b>) cross-sectional view (N/A), ITO thin film with RTA-treated at 600 °C with (<b>e</b>) top view (<b>f</b>) cross-sectional view and RTA-treated at 800 °C with (<b>g</b>) top and (<b>h</b>) cross-sectional view.</p>
Full article ">Figure 2
<p>The AFM surface topography images of thin films (<b>a</b>) as-deposited, (<b>b</b>) RTA-treated at 400 °C (<b>c</b>) RTA-treated at 800 °C and (<b>d</b>) RTA-treated at 800 °C ITO thin film. The surface with scaling of particle size of (<b>e</b>) as-deposited (<b>f</b>) RTA-treated at 400 °C, (<b>g</b>) RTA-treated at 600 °C and (<b>h</b>) RTA-treated at 800 °C ITO thin film.</p>
Full article ">Figure 3
<p>Element analysis of (<b>a</b>) as-deposited (<b>b</b>) RTA-treated at 600 °C and (<b>c</b>) RTA-treated at 800 °C ITO thin film.</p>
Full article ">Figure 4
<p>THz (<b>a</b>) temporal waveforms and (<b>b</b>) amplitude spectra transmitted through bare HR silicon, as-deposited, RTA-treated 400 °C, RTA-treated 600 °C and RTA-treated 800 °C ITO thin films. Inset in (<b>b</b>) shows phase linearity of the THz signal transmitted through a ITO/HR-Si annealed at 800 °C.</p>
Full article ">Figure 5
<p>(<b>a</b>) The real (<span class="html-italic">n</span>) and imaginary (<span class="html-italic">κ</span>) part of refractive indices of the substrate (HR-Si) (added an error bar for thickness variation of ~350 ± 15 µm). (<b>b</b>) The <span class="html-italic">n</span> and (<b>c</b>) <span class="html-italic">κ</span> ITO films as a function of frequency. Black open square: as-deposited; red circles: RTA-treated at 400 °C; green open triangles: RTA-treated at 600 °C; and blue open triangles: RTA-treated at 800 °C.</p>
Full article ">Figure 6
<p>The real (σ<sub>real</sub>) and imaginary (σ<sub>img</sub>) parts of conductivities as the function of THz frequencies of (<b>a</b>) as-deposited (<b>b</b>) RTA-treated at 400 °C (<b>c</b>) RTA-treated at 600 °C and (<b>d</b>) RTA-treated at 800 °C ITO films. Solid and open circles are experimentally extracted conductivities The dashed curves are fitting lines using both models.</p>
Full article ">Figure 7
<p>THz frequency dependent (<b>a</b>) transmittance (<b>b</b>) absorbance and (<b>c</b>) reflectance of as-deposited, RTA-treated at 400 °C, RTA-treated at 600 °C and RTA-treated at 800 °C ITO thin film.</p>
Full article ">Figure 8
<p>(<b>a</b>) The transmittance (<b>b</b>) reflectance and (<b>c</b>) absorptance of as-deposited, RTA-treated at 400 °C, RTA-treated at 600 °C and RTA-treated at 800 °C ITO thin film coated onto fused silica as well T and R of fused silica (<b>d</b>) in UV-VIS-NIR region.</p>
Full article ">Figure 9
<p>(<b>a</b>) Real (n) and (<b>b</b>) imaginary (κ) refractive indices of as-deposited, RTA-treated at 400 °C, RTA-treated at 600 °C and RTA-treated at 800 °C ITO thin film.</p>
Full article ">Figure 10
<p>The (<b>a</b>) wave-length dependent refractive index and (<b>b</b>) optical bandgap of as-deposited, RTA-treated at 400 °C, RTA-treated at 600 °C and RTA-treated at 800 °C ITO thin film.</p>
Full article ">
14 pages, 5008 KiB  
Article
Terahertz Spectra of Mannitol and Erythritol: A Joint Experimental and Computational Study
by Zeyu Hou, Bingxin Yan, Yuhan Zhao, Bo Peng, Shengbo Zhang, Bo Su, Kai Li and Cunlin Zhang
Molecules 2024, 29(13), 3154; https://doi.org/10.3390/molecules29133154 - 2 Jul 2024
Viewed by 756
Abstract
Sugar substitutes, which generally refer to a class of food additives, mostly have vibration frequencies within the terahertz (THz) band. Therefore, THz technology can be used to analyze their molecular properties. To understand the characteristics of sugar substitutes, this study selected mannitol and [...] Read more.
Sugar substitutes, which generally refer to a class of food additives, mostly have vibration frequencies within the terahertz (THz) band. Therefore, THz technology can be used to analyze their molecular properties. To understand the characteristics of sugar substitutes, this study selected mannitol and erythritol as representatives. Firstly, PXRD and Raman techniques were used to determine the crystal structure and purity of mannitol and erythritol. Then, the THz time-domain spectroscopy (THz-TDS) system was employed to measure the spectral properties of the two sugar substitutes. Additionally, density functional theory (DFT) was utilized to simulate the crystal configurations of mannitol and erythritol. The experimental results showed good agreement with the simulation results. Finally, microfluidic chip technology was used to measure the THz spectroscopic properties of the two sugar substitutes in solution. A comparison was made between their solid state and aqueous solution state, revealing a strong correlation between the THz spectra of the two sugar substitutes in both states. Additionally, it was found that the THz spectrum of a substance in solution is related to its concentration. This study provides a reference for the analysis of sugar substitutes. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of PXRD spectra between experimental results and theoretical calculations. (<b>a</b>) Mannitol and (<b>b</b>) erythritol.</p>
Full article ">Figure 2
<p>Comparison of Raman spectra between experimental results and theoretical calculations. (<b>a</b>) Mannitol and (<b>b</b>) erythritol.</p>
Full article ">Figure 3
<p>THz absorption and refractive index spectra of solid mannitol and erythritol. (<b>a</b>) THz absorption spectrum of mannitol. (<b>b</b>) Refractive index of mannitol. (<b>c</b>) THz absorption spectrum of erythritol. (<b>d</b>) Refractive index of erythritol.</p>
Full article ">Figure 4
<p>Comparison of THz spectroscopy and corresponding molecular vibration modes. (<b>a</b>) Experimental spectrum of mannitol. (<b>b</b>) Simulation spectrum of mannitol. (<b>c</b>–<b>g</b>) Vibration modes of mannitol at 0.95, 1.26, 1.62, 2.03, and 2.35 THz, respectively.</p>
Full article ">Figure 5
<p>Comparison of THz spectroscopy and corresponding molecular vibration modes. (<b>a</b>) Experimental spectrum of erythritol. (<b>b</b>) Simulation spectrum of erythritol. (<b>c</b>–<b>f</b>) Vibration modes of erythritol at 1.79, 1.98, 2.03, and 2.43 THz, respectively.</p>
Full article ">Figure 6
<p>Absorption coefficient spectra of solid mannitol (<b>a</b>) and mannitol in solution state (<b>b</b>).</p>
Full article ">Figure 7
<p>Absorption coefficient spectra of solid erythritol (<b>a</b>) and erythritol in solution state (<b>b</b>).</p>
Full article ">Figure 8
<p>THz spectra of different concentrations of mannitol solutions. (<b>a</b>) Time-domain spectra. (<b>b</b>) Frequency-domain spectra. (<b>c</b>) Signal strength concentration relationship diagram.</p>
Full article ">Figure 9
<p>THz spectra of different concentrations of erythritol solutions. (<b>a</b>) Time-domain spectra. (<b>b</b>) Frequency-domain spectra. (<b>c</b>) Signal strength concentration relationship diagram.</p>
Full article ">Figure 10
<p>THz time-domain spectroscopy system. (<b>a</b>) Schematic diagram and (<b>b</b>) physical diagram.</p>
Full article ">Figure 11
<p>Microfluidic chip diagram. (<b>a</b>) Microfluidic chip preparation process. (<b>b</b>) Schematic diagram of THz detection area composed of COC. (<b>c</b>) Physical diagram of microfluidic chip.</p>
Full article ">Figure 12
<p>Molecular structure diagrams of two samples. (<b>a</b>) Mannitol and (<b>b</b>) erythritol.</p>
Full article ">Figure 13
<p>Crystal cell structure diagrams of mannitol (<b>a</b>) and erythritol (<b>b</b>).</p>
Full article ">
13 pages, 4802 KiB  
Article
Terahertz Sensing of L-Valine and L-Phenylalanine Solutions
by Jingyi Shu, Xinli Zhou, Jixuan Hao, Haochen Zhao, Mingming An, Yichen Zhang and Guozhong Zhao
Sensors 2024, 24(12), 3798; https://doi.org/10.3390/s24123798 - 12 Jun 2024
Cited by 1 | Viewed by 719
Abstract
To detect and differentiate two essential amino acids (L-Valine and L-Phenylalanine) in the human body, a novel asymmetrically folded dual-aperture metal ring terahertz metasurface sensor was designed. A solvent mixture of water and glycerol with a volume ratio of 2:8 was proposed to [...] Read more.
To detect and differentiate two essential amino acids (L-Valine and L-Phenylalanine) in the human body, a novel asymmetrically folded dual-aperture metal ring terahertz metasurface sensor was designed. A solvent mixture of water and glycerol with a volume ratio of 2:8 was proposed to reduce the absorption of terahertz waves by reducing the water content. A sample chamber with a controlled liquid thickness of 15 μm was fabricated. And a terahertz time-domain spectroscopy (THz-TDS) system, which is capable of horizontally positioning the samples, was assembled. The results of the sensing test revealed that as the concentration of valine solution varied from 0 to 20 mmol/L, the sensing resonance peak shifted from 1.39 THz to 1.58 THz with a concentration sensitivity of 9.98 GHz/mmol∗L−1. The resonance peak shift phenomenon in phenylalanine solution was less apparent. It is assumed that the coupling enhancement between the absorption peak position of solutes in the solution and the sensing peak position amplified the terahertz localized electric field resonance, which resulted in the increase in frequency shift. Therefore, it could be shown that the sensor has capabilities in performing the marker sensing detection of L-Valine. Full article
(This article belongs to the Special Issue Terahertz Sensors)
Show Figures

Figure 1

Figure 1
<p>Metasurface sensor’s (<b>a</b>) unit structure schematic; (<b>b</b>) working principle diagram; and (<b>c</b>) Lithography process.</p>
Full article ">Figure 2
<p>Vertical incident THz-TDS: (<b>a</b>) optical path and (<b>b</b>) system diagram.</p>
Full article ">Figure 3
<p>Experimental preprocessing diagram.</p>
Full article ">Figure 4
<p>Asymmetric folded double open ring’s (<b>a</b>) experimental characterization and simulation comparison; (<b>b</b>) surface electric field distribution; (<b>c</b>) surface current distribution; (<b>d</b>) the influence of the thickness of the cover on the frequency shift of the resonant peak; (<b>e</b>) the influence of the change in refractive index on the resonant peak frequency shift; and (<b>f</b>) the effect of L2 size on sensor performance.</p>
Full article ">Figure 5
<p>Terahertz spectrum of glycerol: (<b>a</b>) time-domain spectroscopy; (<b>b</b>) frequency domain spectrum; (<b>c</b>) transmission spectrum; (<b>d</b>) refractive index spectrum.</p>
Full article ">Figure 6
<p>Terahertz spectrum of solutions of water mixed with glycerol in different proportions (<b>a</b>) frequency domain spectrum; (<b>b</b>) transmission spectrum; (<b>c</b>) refractive index spectrum; (<b>d</b>) absorption spectrum; (<b>e</b>) transmission spectrum of the sensor with mixed solutions.</p>
Full article ">Figure 7
<p>(<b>a</b>) Transmission spectrum of L-Valine solution; (<b>b</b>) the optimized transmission spectrum of L-Valine solution; (<b>c</b>) fitting analysis of L-Phenylalanine sensing frequency shift and solution concentration; (<b>d</b>) transmission spectrum of L-Phenylalanine solution; (<b>e</b>) the optimized transmission spectrum of L-Valine solution; (<b>f</b>) fitting analysis of L-Phenylalanine sensing frequency shift and solution concentration.</p>
Full article ">Figure 8
<p>Two amino acid sensing frequency shifts and solution concentration fitting comparison.</p>
Full article ">
13 pages, 6163 KiB  
Article
Qualitative and Quantitative Detection of Typical Reproductive Hormones in Dairy Cows Based on Terahertz Spectroscopy and Metamaterial Technology
by Shuang Liang, Jingbo Zhao, Wenwen Zhao, Nan Jia, Zhiyong Zhang and Bin Li
Molecules 2024, 29(10), 2366; https://doi.org/10.3390/molecules29102366 - 17 May 2024
Viewed by 1137
Abstract
Progesterone (PROG) and estrone (E1) are typical reproductive hormones in dairy cows. Assessing the levels of these hormones in vivo can aid in estrus identification. In the present work, the feasibility of the qualitative and quantitative detection of PROG and E [...] Read more.
Progesterone (PROG) and estrone (E1) are typical reproductive hormones in dairy cows. Assessing the levels of these hormones in vivo can aid in estrus identification. In the present work, the feasibility of the qualitative and quantitative detection of PROG and E1 using terahertz time-domain spectroscopy (THz-TDS) and metamaterial technology was preliminarily investigated. First, the time domain spectra, frequency domain spectra, and absorption coefficients of PROG and E1 samples were collected and analyzed. A vibration analysis was conducted using density functional theory (DFT). Subsequently, a double-ring (DR) metamaterial structure was designed and simulated using the frequency domain solution algorithm in CST Studio Suite (CST) software. This aimed to ensure that the double resonance peaks of DR were similar to the absorption peaks of PROG and E1. Finally, the response of DR to different concentrations of PROG/E1 was analyzed and quantitatively modeled. The results show that a qualitative analysis can be conducted by comparing the corresponding DR resonance peak changes in PROG and E1 samples at various concentrations. The best R2 for the PROG quantitative model was 0.9872, while for E1, it was 0.9828. This indicates that terahertz spectral–metamaterial technology for the qualitative and quantitative detection of the typical reproductive hormones PROG and E1 in dairy cows is feasible and worthy of in-depth exploration. This study provides a reference for the identification of dairy cow estrus. Full article
Show Figures

Figure 1

Figure 1
<p>Changes in reproductive hormone content during the estrus cycle in cows.</p>
Full article ">Figure 2
<p>Metamaterial structure diagram: (<b>a</b>) double-ring structure and (<b>b</b>) thickness.</p>
Full article ">Figure 3
<p>Diagram of the THz-TDS system.</p>
Full article ">Figure 4
<p>Flow chart of research.</p>
Full article ">Figure 5
<p>Terahertz spectral information: (<b>a</b>) time domain, (<b>b</b>) frequency domain, and (<b>c</b>) absorption coefficient.</p>
Full article ">Figure 6
<p>Comparison of experimental and calculated results: (<b>a</b>) PROG and (<b>b</b>) E<sub>1.</sub></p>
Full article ">Figure 7
<p>Vibration identification of (<b>a</b>) PROG at 1.02 THz and (<b>b</b>) 1.24 THz; (<b>c</b>) E<sub>1</sub> at 0.98 THz.</p>
Full article ">Figure 8
<p>Metamaterial design and simulation; (<b>a</b>) simulation and test results; (<b>b</b>) responses to different refractive indices; (<b>c</b>) relationship between resonant peak frequency shifts, intensity changes, and the refractive index; (<b>d</b>) electric field distribution at 0.819 THz; (<b>e</b>) electric field distribution at 1.394 THz; (<b>f</b>) surface current distribution at 0.819 THz; (<b>g</b>) surface current distribution at 1.394 THz.</p>
Full article ">Figure 9
<p>Transmission spectra of different concentrations of PROG and E<sub>1</sub>; (<b>a</b>) PROG transmission spectra; (<b>b</b>,<b>c</b>) detailed spectrum of transmission peak at 0.819 THz and 1.394 THz for PROG; (<b>d</b>) E<sub>1</sub> transmission spectra; (<b>e</b>,<b>f</b>) detailed spectrum of transmission peak at 0.819 THz and 1.394 THz for E<sub>1.</sub></p>
Full article ">Figure 10
<p>Qualitative analysis: (<b>a</b>) matching the degree of the PROG/E<sub>1</sub> absorption peak with the metamaterial resonance peak, (<b>b</b>) frequency change om the 0.819 THz resonance peak, and (<b>c</b>) intensity change.</p>
Full article ">Figure 11
<p>Resonance peak change: (<b>a</b>) 0.819 THz frequency shift, (<b>b</b>) 0.819 THz intensity change, and (<b>c</b>) 1.394 THz intensity change and PROG concentration linear fitting results.</p>
Full article ">Figure 12
<p>Resonance peak change: (<b>a</b>) 0.819 THz intensity change and (<b>b</b>) 1.394 THz intensity change and E<sub>1</sub> concentration linear fitting results.</p>
Full article ">
12 pages, 1040 KiB  
Article
Gamma Irradiation Effect on Polymeric Chains of Epoxy Adhesive
by Carino Ferrante, Leonardo Lucchesi, Alessia Cemmi, Ilaria Di Sarcina, Jessica Scifo, Adriano Verna, Andrea Taschin, Luca Senni, Marco Beghini, Bernardo Disma Monelli and Fabrizio Raffaelli
Polymers 2024, 16(9), 1202; https://doi.org/10.3390/polym16091202 - 25 Apr 2024
Viewed by 998
Abstract
The study of materials for space exploration is one of the most interesting targets of international space agencies. An essential tool for realizing light junctions is epoxy adhesive (EA), which provides an elastic and robust material with a complex mesh of polymeric chains [...] Read more.
The study of materials for space exploration is one of the most interesting targets of international space agencies. An essential tool for realizing light junctions is epoxy adhesive (EA), which provides an elastic and robust material with a complex mesh of polymeric chains and crosslinks. In this work, a study of the structural and chemical modification of a commercial two-part flexible EA (3M™ Scotch-Weld™ EC-2216 B/A Gray), induced by 60Co gamma radiation, is presented. Combining different spectroscopic techniques, such as the spectroscopic Fourier transform infrared spectroscopy (FTIR), the THz time-domain spectroscopy (TDS), and the electron paramagnetic resonance (EPR), a characterization of the EA response in different regions of the electromagnetic spectrum is performed, providing valuable information about the structural and chemical properties of the polymers before and after irradiation. A simultaneous dissociation of polymeric chain and crosslinking formation is observed.The polymer is not subject to structural modification at an absorbed dose of 10 kGy, in which only transient free radicals are observed. Differently, between 100 and 500 kGy, a gradual chemical degradation of the samples is observed together with a broad and long-living EPR signal appearance. This study also provides a microscopic characterization of the material useful for the mechanism evaluation of system degradation. Full article
(This article belongs to the Section Polymer Chemistry)
Show Figures

Figure 1

Figure 1
<p>Sketch of typical reaction for the polymeric crosslinking [<a href="#B9-polymers-16-01202" class="html-bibr">9</a>,<a href="#B10-polymers-16-01202" class="html-bibr">10</a>]. Specifically, we report the bond formation between a diamine compound and 4 epoxy group of polymeric chain. Depending on how many amino groups there are in the crosslinker, the crosslink structure changes.</p>
Full article ">Figure 2
<p>Epoxy adhesive FTIR and THz-TDS measurements at different total absorbed doses are shown. Each FTIR line is the average of more than 10 FTIR measurements normalized to the 2923 cm<sup>−1</sup> peak. Different FTIR spectral regions of interest are highlighted in (<b>a</b>–<b>c</b>). The black arrows of ∼1180 cm<sup>−1</sup>, ∼1725 cm<sup>−1</sup>, and 827 cm<sup>−1</sup> indicate the <math display="inline"><semantics> <mi>γ</mi> </semantics></math>ray-induced broadening, intensity increasing, and frequency shift of FTIR peak, respectively. Differently, the refractive index and absorption coefficient obtained by the THz-TDS setup are reported in (<b>d</b>). Part of the figure was reproduced/adapted with permission from [<a href="#B20-polymers-16-01202" class="html-bibr">20</a>], Elsevier, 2021.</p>
Full article ">Figure 3
<p>Fit of two FTIR spectral regions with a sum of Lorentzian convoluted with the spectral resolution of the instrument (Gaussian profile with variance 4.7 cm<sup>−1</sup>). (<b>a</b>) The best fit (black lines) of the nonirradiated FTIR spectrum (blue dots in the upper panel) and FTIR spectrum at a dose of 500 kGy (red dots in the lower panel) with the convoluted Lorentzian contribution with colored solid lines. The bandwidth of Lorentzian peaks is reported with the same colors in (<b>b</b>) for different doses. The profiles highlight a collective peak broadening related to a more heterogeneous chemical structure of polymeric chains and crosslinks.</p>
Full article ">Figure 4
<p>Distribution of the exact maximum position for the FTIR peak around 827 cm<sup>−1</sup> for the different samples before irradiation (lowest panel) and after the irradiation at the different indicated doses (higher panels). The spectra reported in <a href="#polymers-16-01202-f002" class="html-fig">Figure 2</a> are, instead, the average of all the spectra acquired before the irradiation and after. The position is obtained by fitting with a Gaussian spectral range of 6 cm<sup>−1</sup> around the peak maximum. The red dashed line is a visual guide to highlight the dose-dependent blueshift.</p>
Full article ">Figure 5
<p>PCA of FTIR spectra with two components. (<b>a</b>) The spectra associated with the PC1 and PC2. The spectrum of PC1 is divided by a factor of 10 for graphical reasons. A combination of two spectra is also reported, with the yellow line discerning the role of PC2 addition in the spectral profile. (<b>b</b>) Histogram of the ratio of coefficient PC2 and PC1 for different doses. (<b>c</b>) The scatter plot of coefficient PC2 and PC1. The colors of bars and squares in (<b>b</b>,<b>c</b>) indicate the <math display="inline"><semantics> <mi>γ</mi> </semantics></math>ray dose of the samples in agreement with <a href="#polymers-16-01202-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 6
<p>Temporal evolution of EPR spectra of EA. In (<b>a</b>), the raw spectral profiles are reported at different doses. The color of each thin line indicates the time delay with respect to the end of gamma irradiation (see the color bar). The EPR spectra before irradiation are reported with black lines. The difference of the EPR spectrum with respect to the nonirradiated sample is reported in (<b>b</b>). The green and red vertical dashed lines point to the magnetic field value in correspondence with <span class="html-italic">g</span> contributions observed in nonirradiated and irradiated samples, respectively. The samples are sliced from the bulk specimens with a size &gt; 2 mm.</p>
Full article ">Figure 7
<p>Temporaldynamic of EPR signal in EA. (<b>a</b>) The EPR spectra of the EA samples with a thickness of 150 µm are reported. The color of each thin line indicates the time delay with respect to the end of gamma irradiation at 50 kGy (see the color bar). (<b>b</b>) The transient EPR intensity at 3462 G (<a href="#polymers-16-01202-f006" class="html-fig">Figure 6</a>b), normalized to 1 for the first measurement after irradiation for each absorbed dose, is reported as a function of the recovery time through colored circles. The black circles represent the time profile for the sample with a thickness of 150 µm, reported in (<b>a</b>). The solid lines represent the exponential best fit of the experimental data with time scales of 66, 100, and 64 h for 10, 100, and 500 kGy, respectively.</p>
Full article ">Figure 8
<p>Dose dependence of EPR signal difference (see <a href="#polymers-16-01202-f006" class="html-fig">Figure 6</a>b) at 3411 G and area of C=O peak in FTIR measurement, in the left and right axis, respectively. The area of the C=O peak is calculated between 1700 and 1800 cm<sup>−1</sup> [<a href="#B20-polymers-16-01202" class="html-bibr">20</a>]. The two profiles are compatible considering the error of measurement. Part of the figure was reproduced/adapted with permission from [<a href="#B20-polymers-16-01202" class="html-bibr">20</a>], Elsevier, 2021.</p>
Full article ">
21 pages, 4330 KiB  
Article
Terahertz Time-Domain Spectroscopy of Blood Serum for Differentiation of Glioblastoma and Traumatic Brain Injury
by Denis A. Vrazhnov, Daria A. Ovchinnikova, Tatiana V. Kabanova, Andrey G. Paulish, Yury V. Kistenev, Nazar A. Nikolaev and Olga P. Cherkasova
Appl. Sci. 2024, 14(7), 2872; https://doi.org/10.3390/app14072872 - 28 Mar 2024
Cited by 1 | Viewed by 1086
Abstract
The possibility of the differentiation of glioblastoma from traumatic brain injury through blood serum analysis by terahertz time-domain spectroscopy and machine learning was studied using a small animal model. Samples of a culture medium and a U87 human glioblastoma cell suspension in the [...] Read more.
The possibility of the differentiation of glioblastoma from traumatic brain injury through blood serum analysis by terahertz time-domain spectroscopy and machine learning was studied using a small animal model. Samples of a culture medium and a U87 human glioblastoma cell suspension in the culture medium were injected into the subcortical brain structures of groups of mice referred to as the culture medium injection groups and glioblastoma groups, accordingly. Blood serum samples were collected in the first, second, and third weeks after the injection, and their terahertz transmission spectra were measured. The injection caused acute inflammation in the brain during the first week, so the culture medium injection group in the first week of the experiment corresponded to a traumatic brain injury state. In the third week of the experiment, acute inflammation practically disappeared in the culture medium injection groups. At the same time, the glioblastoma group subjected to a U87 human glioblastoma cell injection had the largest tumor size. The THz spectra were analyzed using two dimensionality reduction algorithms (principal component analysis and t-distributed Stochastic Neighbor Embedding) and three classification algorithms (Support Vector Machine, Random Forest, and Extreme Gradient Boosting Machine). Constructed prediction data models were verified using 10-fold cross-validation, the receiver operational characteristic curve, and a corresponding area under the curve analysis. The proposed machine learning pipeline allowed for distinguishing the traumatic brain injury group from the glioblastoma group with 95% sensitivity, 100% specificity, and 97% accuracy with the Extreme Gradient Boosting Machine. The most informative features for these groups’ differentiation were 0.37, 0.40, 0.55, 0.60, 0.70, and 0.90 THz. Thus, an analysis of mouse blood serum using terahertz time-domain spectroscopy and machine learning makes it possible to differentiate glioblastoma from traumatic brain injury. Full article
(This article belongs to the Section Optics and Lasers)
Show Figures

Figure 1

Figure 1
<p>The idea of the study (<b>a</b>): red area illustrates the CMI state, blue, the GBM state, and white area, the healthy state. Arrow means comparison of groups for TBI and GBM differentiation. List of ML binary models under study (<b>b</b>).</p>
Full article ">Figure 2
<p>THz time-domain spectrometer used in the transmission mode.</p>
Full article ">Figure 3
<p>The averaged THz spectra in the time domain (<b>a</b>) and the frequency domain (<b>b</b>) for all studied groups.</p>
Full article ">Figure 4
<p>Proposed ML pipeline steps.</p>
Full article ">Figure 5
<p>Illustration of application of unsupervised ML methods (t-SNE and PCA) for the separability analysis of the CMI groups.</p>
Full article ">Figure 6
<p>t-SNE (<b>a</b>) and PCA (<b>b</b>) visualization of the CMI groups.</p>
Full article ">Figure 7
<p>ROC-AUC analysis of CMI groups: 1st vs. 2nd weeks using SVM (<b>a</b>), 1st vs. 3rd weeks using Catboost (<b>b</b>), and 2nd vs. 3rd weeks using Catboost (<b>c</b>).</p>
Full article ">Figure 8
<p>Informative feature analysis of CMI groups: 1st vs. 2nd weeks using Catboost (<b>a</b>), 1st vs. 3rd weeks using Catboost (<b>b</b>), 1st vs. 2nd weeks using SVM (<b>c</b>), and 1st vs. 3rd weeks using SVM (<b>d</b>).</p>
Full article ">Figure 9
<p>ROC-AUC analysis of SVM classifier (<b>a</b>) of the TBI group vs. the GBM group and the corresponding informative features (<b>b</b>).</p>
Full article ">
13 pages, 3573 KiB  
Article
High-Density Polyethylene Custom Focusing Lenses for High-Resolution Transient Terahertz Biomedical Imaging Sensors
by Debamitra Chakraborty, Robert Boni, Bradley N. Mills, Jing Cheng, Ivan Komissarov, Scott A. Gerber and Roman Sobolewski
Sensors 2024, 24(7), 2066; https://doi.org/10.3390/s24072066 - 24 Mar 2024
Cited by 3 | Viewed by 1364
Abstract
Transient terahertz time-domain spectroscopy (THz-TDS) imaging has emerged as a novel non-ionizing and noninvasive biomedical imaging modality, designed for the detection and characterization of a variety of tissue malignancies due to their high signal-to-noise ratio and submillimeter resolution. We report our design of [...] Read more.
Transient terahertz time-domain spectroscopy (THz-TDS) imaging has emerged as a novel non-ionizing and noninvasive biomedical imaging modality, designed for the detection and characterization of a variety of tissue malignancies due to their high signal-to-noise ratio and submillimeter resolution. We report our design of a pair of aspheric focusing lenses using a commercially available lens-design software that resulted in about 200 × 200-μm2 focal spot size corresponding to the 1-THz frequency. The lenses are made of high-density polyethylene (HDPE) obtained using a lathe fabrication and are integrated into a THz-TDS system that includes low-temperature GaAs photoconductive antennae as both a THz emitter and detector. The system is used to generate high-resolution, two-dimensional (2D) images of formalin-fixed, paraffin-embedded murine pancreas tissue blocks. The performance of these focusing lenses is compared to the older system based on a pair of short-focal-length, hemispherical polytetrafluoroethylene (TeflonTM) lenses and is characterized using THz-domain measurements, resulting in 2D maps of the tissue refractive index and absorption coefficient as imaging markers. For a quantitative evaluation of the lens effect on the image resolution, we formulated a lateral resolution parameter, R2080, defined as the distance required for a 20–80% transition of the imaging marker from the bare paraffin region to the tissue region in the same image frame. The R2080 parameter clearly demonstrates the advantage of the HDPE lenses over TeflonTM lenses. The lens-design approach presented here can be successfully implemented in other THz-TDS setups with known THz emitter and detector specifications. Full article
(This article belongs to the Special Issue Research Development in Terahertz and Infrared Sensing Technology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Transmittance (including surface reflections) with respect to THz frequency for an HDPE material; (<b>b</b>) THz transmittance (including surface reflections) with respect to HDPE material thickness. The data for these plots are adopted from [<a href="#B17-sensors-24-02066" class="html-bibr">17</a>,<a href="#B19-sensors-24-02066" class="html-bibr">19</a>].</p>
Full article ">Figure 2
<p>Schematic of a plano-aspheric lens. All dimensions are in mm.</p>
Full article ">Figure 3
<p>(<b>a</b>) Surface profile map of the aspheric <span class="html-italic">x</span>–<span class="html-italic">y</span> surface of one of the fabricated lenses with reference to the ideal designed aspheric surface. (<b>b</b>) Line plot of the surface error along the lens radius.</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic diagram of the THz-TDS setup using the pair of hemispherical Teflon<sup>TM</sup> lenses. Optical pulses of 100 fs wide, with an 800 nm wavelength were generated by a commercial Ti:Sapphire laser and were split using a 50:50 polarizing beam-splitter. The pump beam was used to excite a low-temperature grown GaAs (LT-GaAs) THz emitter, whereas the probe beam was used to excite an LT-GaAs detector antenna. The emitted THz beam is passed through an HH-Si lens and a hemispherical Teflon lens to focus the signal on a sample plane. Another Teflon<sup>TM</sup> lens and an HH-Si lens attached to the detector antenna focus the beam onto the detector. (<b>b</b>) Schematic diagram of the optical system using the pair of aspheric HDPE lenses instead of Teflon<sup>TM</sup> lenses. A custom-made plano-aspheric HDPE lens tightly focuses the THz beam from the HH-Si lens attached to the THz emitter. After that, the diverging beam from the focal point is collimated/focused by the second HDPE lens. The rays are then passed through the HH-Si lens attached to the detector to focus it on the LT-GaAs photoconductive antenna detector (the rightmost element). (<b>c</b>) &lt;1% energy coupling from the 50 µm diameter emitter to the 50 µm diameter detector using a hemispherical Teflon<sup>TM</sup> lens. The irradiance at the center of the detector is 74%. The cross-section line scans at a normalized energy scale are provided along both the <span class="html-italic">x</span> and the <span class="html-italic">y</span> axes. The ray-trace results were generated by launching 1.00 × 10<sup>6</sup> rays, of which 1.25 × 10<sup>3</sup> passed and 0.125% power was transmitted. (<b>d</b>) An energy coupling of &gt;90% energy was passed from the 50 µm diameter emitter to the 50 µm diameter detector using an aspheric HDPE lens. The irradiance occurs at the center if the detector is at 74%. The cross-section line scans on a normalized energy scale are provided along both the <span class="html-italic">x</span> and the <span class="html-italic">y</span> axes. The ray-trace results were generated by launching 1.00 × 10<sup>5</sup> rays, of which 9.08 × 10<sup>4</sup> passed and 90.8% power was transmitted. (<b>e</b>) A 600 µm diameter spot at the focal point (sample plane) using a hemispherical Teflon<sup>TM</sup> lens. The inset shows a spot diagram, resulting in about 300 × 300 μm<sup>2</sup> focal spot size corresponding to 1 THz central frequency. A very small back circle at the center indicates the diffraction limit (also known as the ‘Airy disk’) of 76 µm. The green dots represent rays originating on the axis location of the emitter, traced geometrically by refraction at the lens surfaces. The geometrical rms radius is found to be 534 µm. Hence the system is not diffraction-limited. (<b>f</b>) A 200 µm diameter spot at the focal point (sample plane) aspheric HDPE lens. The inset shows a spot diagram at the focal point using the aspheric lenses, resulting in about a 100 × 100-μm<sup>2</sup> focal spot size corresponding to 1 THz central frequency. The back circle at the figure center indicates the diffraction limit (also known as the ‘Airy disk’) of 565 µm. The geometrical rms radius is found to be 225 µm, which is the full-width-at-half-maximum (FWHM) of relative energy vs. position diagram. Hence, the HDPE system is diffraction-limited.</p>
Full article ">Figure 5
<p>(<b>a</b>) THz transients using Teflon<sup>TM</sup> lenses (blue solid line is multiplied 10× for better visibility) and HDPE lenses (red dashed line), measured under identical conditions without purging the system with dry nitrogen [hence, water signatures are visible in the FFTs shown in (<b>b</b>)]. (<b>b</b>) Fast Fourier transforms of time-domain waveforms shown in (<b>a</b>) with 10 dB bandwidths indicated.</p>
Full article ">Figure 6
<p>(<b>a</b>) An optical image of a 4 mm FFPE healthy murine tissue used in this study. High-resolution, 2D THz-imaging maps of the same tissue, using (<b>b</b>) the refractive index and (<b>c</b>) the absorption coefficient as imaging markers.</p>
Full article ">Figure 7
<p>(<b>a</b>) The double-sigmoid fit function was applied to the line scans at <span class="html-italic">y</span> = 3 mm for both the <span class="html-italic">n</span> and <span class="html-italic">α</span> maps shown earlier in <a href="#sensors-24-02066-f006" class="html-fig">Figure 6</a>. (<b>b</b>) Double-sigmoid fit for <span class="html-italic">n</span>. (<b>c</b>) Double-sigmoid fit for <span class="html-italic">α</span>.</p>
Full article ">
19 pages, 18294 KiB  
Article
Non-Destructive Testing of a Fiber-Web-Reinforced Polymethacrylimide Foam Sandwich Panel with Terahertz Time-Domain Spectroscopy
by Yu Liu, Yefa Hu, Jinguang Zhang, Haixin Liu and Meng Wan
Sensors 2024, 24(6), 1715; https://doi.org/10.3390/s24061715 - 7 Mar 2024
Cited by 1 | Viewed by 1051
Abstract
Terahertz (THz) non-destructive testing can detect internal defects in dielectric materials. However, this technology is mainly used for detecting thin and simple structures at present, lacking validations for the detection effectiveness of internal defects in thicker and more complex structures, such as fiber-web-reinforced [...] Read more.
Terahertz (THz) non-destructive testing can detect internal defects in dielectric materials. However, this technology is mainly used for detecting thin and simple structures at present, lacking validations for the detection effectiveness of internal defects in thicker and more complex structures, such as fiber-web-reinforced composite sandwich panels. In this study, samples of fiber-web-reinforced polymethacrylimide foam sandwich panels, which are, respectively, 20 mm and 30 mm thick, were made to detect the internal debonding, inclusion, pore, and crack defects by the THz time-domain spectroscopy system (THz-TDS). The peak-to-peak-imaging algorithm, maximum-amplitude-imaging algorithm, minimum-amplitude-imaging algorithm, pulse-width-imaging algorithm, and time-of-flight-imaging algorithm were used to process and image the collected THz signals. The results showed that the peak-to-peak-imaging algorithm had the best performance. To address the low imaging resolution of THz-TDS, a block-based super-resolution reconstruction method—SSSRGAN—is proposed, which can improve image resolution while maintaining the clear edge contours of defects. The defect-detection results of the samples showed that THz-TDS could detect all pore, debonding, and crack defects, with a minimum size of 3 mm for pores and debonding and a minimum thickness of 1 mm for cracks. The method showed poor detection performance for inclusions with a thickness of 0.053 mm, but could still extract the defect features. Based on the THz-TDS reflection mode measurement principle, the thickness information of the panel, foam core, and web of the samples was calculated: the measurement error was no more than 0.870 mm for Sample #1 and no more than 0.270 mm for Sample #2, demonstrating the accuracy of THz-TDS in measuring the dimensions of sandwich panel structures. In general, THz technology shows potential for detecting internal defects and performing dimensional measurements in complex structures. With the advancement of portable devices and enhancements in detection speed, real-time on-site detection is anticipated in the future. Full article
(This article belongs to the Section Optical Sensors)
Show Figures

Figure 1

Figure 1
<p>The structure of fiber-web-reinforced polymethacrylimide foam sandwich panels: (<b>a</b>) Axis side view. (<b>b</b>) Exploded view.</p>
Full article ">Figure 2
<p>The defect size information (unit: mm).</p>
Full article ">Figure 3
<p>Optical images of the sandwich panel samples (Sample #1 with a thickness of 30 mm; Sample #2 with a thickness of 20 mm), GFRP laminate (#3), and PMI foam (#4).</p>
Full article ">Figure 4
<p>The THz-TDS reflection mode measurement principle.</p>
Full article ">Figure 5
<p>The THz-TDS system equipment.</p>
Full article ">Figure 6
<p>The comparative analysis (time-domain spectrogram and frequency-domain spectrogram) between the reference signal and sample signal.</p>
Full article ">Figure 7
<p>Three sampling points are randomly selected for the samples.</p>
Full article ">Figure 8
<p>The optical performance parameters of the GFRP panel and PMI foam: (<b>a</b>) Refractive index of GFRP panel. (<b>b</b>) Refractive index of PMI foam. (<b>c</b>) Extinction coefficient. (<b>d</b>) Absorption coefficient.</p>
Full article ">Figure 9
<p>Reflected THz signals at defect and non-defect positions of Sample #1 (electric-field waveform related to time [<a href="#B13-sensors-24-01715" class="html-bibr">13</a>]).</p>
Full article ">Figure 10
<p>The principle of THz-TDS non-destructive testing.</p>
Full article ">Figure 11
<p>The results generated by by different imaging algorithms: (<b>a</b>) Peak-to-peak imaging. (<b>b</b>) Maximum-amplitude imaging. (<b>c</b>) Minimum-amplitude imaging. (<b>d</b>) Pulse-width imaging. (<b>e</b>) Time-of-flight imaging.</p>
Full article ">Figure 12
<p>The complete time-domain signal at a certain point of Sample #1: (a) Signal of upper panel. (b) Signal of upper PMI foam core. (c) Signal of lower PMI foam core.</p>
Full article ">Figure 13
<p>The corresponding relationship between the imaging results and the physical images: (<b>a</b>) The upper panel. (<b>b</b>) The upper PMI foam core. (<b>c</b>) The lower PMI foam core.</p>
Full article ">Figure 14
<p>The image slice-segmentation-reconstruction algorithm based on SRGAN.</p>
Full article ">Figure 15
<p>The comparison of the images before and after super-resolution reconstruction: (<b>a</b>) Low-resolution original image. (<b>b</b>) Locally enlarged image of the low-resolution image. (<b>c</b>) Super-resolution reconstructed image. (<b>d</b>) Locally enlarged image of the super-resolution reconstructed image.</p>
Full article ">Figure 16
<p>The time-domain signals and relative imaging results of pore defects: (<b>a</b>) The detection image of pore defects in the upper PMI foam core. (<b>b</b>) The time-domain signal at the position of the pore defect in the upper PMI foam core. (<b>c</b>) The detection image of pore defects in the lower PMI foam core. (<b>d</b>) The time-domain signal at the position of the pore defect in the lower PMI foam core.</p>
Full article ">Figure 17
<p>The time-domain signals and relative imaging results of debonding defects: (<b>a</b>) The detection image of debonding defects between the upper PMI foam core and the web. (<b>b</b>) The time-domain signal at the position of the debonding defect between the upper PMI foam core and the web. (<b>c</b>) The detection image of debonding defects between the lower PMI foam core and the web. (<b>d</b>) The time-domain signal at the position of the debonding defect between the lower PMI foam core and the web.</p>
Full article ">Figure 18
<p>The time-domain signals and relative imaging results of inclusion defects: (<b>a</b>) The detection image of inclusion defects between the upper PMI foam core and the web. (<b>b</b>) The time-domain signal at the position of the inclusion defect between the upper PMI foam core and the web. (<b>c</b>) The detection image of inclusion defects between the lower PMI foam core and the web. (<b>d</b>) The time-domain signal at the position of the inclusion defect between the lower PMI foam core and the web.</p>
Full article ">Figure 19
<p>The time-domain signals and relative imaging results of crack defects: (<b>a</b>) The detection image of crack defects in the upper PMI foam core. (<b>b</b>) The time-domain signal at the defect position in the upper PMI foam core. (<b>c</b>) The detection image of crack defects in the lower PMI foam core. (<b>d</b>) The time-domain signal at the defect position in the lower PMI foam core.</p>
Full article ">
9 pages, 6754 KiB  
Article
Dielectric Terahertz Characterization of Microwave Substrates and Dry Resist
by Silvia Tofani, Tiziana Ritacco, Luca Maiolo, Francesco Maita, Romeo Beccherelli, Walter Fuscaldo and Dimitrios C. Zografopoulos
Crystals 2024, 14(3), 205; https://doi.org/10.3390/cryst14030205 - 21 Feb 2024
Viewed by 1329
Abstract
Microwave fabrication and design techniques are commonly employed in the terahertz (THz) domain. However, a characterization of commercially available microwave dielectric materials is usually lacking at sub-THz and THz frequencies. In this work, we characterized four substrates by Rogers and an Ordyl dry [...] Read more.
Microwave fabrication and design techniques are commonly employed in the terahertz (THz) domain. However, a characterization of commercially available microwave dielectric materials is usually lacking at sub-THz and THz frequencies. In this work, we characterized four substrates by Rogers and an Ordyl dry resist between 0.2 and 2 THz, in terms of relative permittivity and loss tangent. The reflectance spectra of the investigated materials were retrieved by means of THz time-domain spectroscopy in reflection mode and post-processed according to a transmission-line model in which the materials’ parameters are fit by means of the Havriliak–Negami variation of the Debye model. The relative permittivity of the investigated materials showed negligible frequency dispersion in the sub-THz and in the THz range. In terms of the loss tangent, the Rogers substrates revealed a more pronounced frequency-dispersive behavior among different materials, as dictated by the Havriliak–Negami model. The Ordyl resist was dispersive in the 0.2–1.2 THz range and presented a nearly constant loss tangent value between 1.2 and 2 THz. These results may represent a reference for the development of innovative components for THz and sub-THz emerging applications. Full article
(This article belongs to the Special Issue Feature Papers in Crystals 2023)
Show Figures

Figure 1

Figure 1
<p>Investigated material samples: (<b>a</b>) Rogers CLTE-MW; (<b>b</b>) Rogers Ultralam 3850HT; (<b>c</b>) Rogers Duroid 6035HTC; (<b>d</b>) Rogers RO3006; (<b>e</b>) Ordyl FP 415.</p>
Full article ">Figure 2
<p>Schematic of the THz-TDS setup in reflection mode. It consists of a near-infrared laser, photoconductive transmit (Tx) and receive (Rx) antennas, an optical delay line, a set of parabolic mirrors, and a metallic baseplate for the placement of a sample. The red arrows represent the guided signal path, while the red-shaded elements mark the THz beam propagating in the free space.</p>
Full article ">Figure 3
<p>Sample cross-section and its equivalent circuit model.</p>
Full article ">Figure 4
<p>Spectra of the (<b>a</b>) relative dielectric permittivity and (<b>b</b>) loss tangent fit by means of the Havriliak–Negami variation of the Debye model.</p>
Full article ">Figure 5
<p>Measured (black line) and theoretical (green dashed line) spectra based on the Havriliak–Negami variation of the Debye model for the five investigated materials: (<b>a</b>) Rogers CLTE-MW, (<b>b</b>) Rogers Ultralam 3850HT, (<b>c</b>) Rogers Duroid 6035HTC, (<b>d</b>) Rogers RO3006, and (<b>e</b>) Ordyl FP 415.</p>
Full article ">Figure 5 Cont.
<p>Measured (black line) and theoretical (green dashed line) spectra based on the Havriliak–Negami variation of the Debye model for the five investigated materials: (<b>a</b>) Rogers CLTE-MW, (<b>b</b>) Rogers Ultralam 3850HT, (<b>c</b>) Rogers Duroid 6035HTC, (<b>d</b>) Rogers RO3006, and (<b>e</b>) Ordyl FP 415.</p>
Full article ">
13 pages, 3517 KiB  
Article
Study on Bulk-Surface Transport Separation and Dielectric Polarization of Topological Insulator Bi1.2Sb0.8Te0.4Se2.6
by Yueqian Zheng, Tao Xu, Xuan Wang, Zhi Sun and Bai Han
Molecules 2024, 29(4), 859; https://doi.org/10.3390/molecules29040859 - 15 Feb 2024
Cited by 1 | Viewed by 973
Abstract
This study successfully fabricated the quaternary topological insulator thin films of Bi1.2Sb0.8Te0.4Se2.6 (BSTS) with a thickness of 25 nm, improving the intrinsic defects in binary topological materials through doping methods and achieving the separation of transport [...] Read more.
This study successfully fabricated the quaternary topological insulator thin films of Bi1.2Sb0.8Te0.4Se2.6 (BSTS) with a thickness of 25 nm, improving the intrinsic defects in binary topological materials through doping methods and achieving the separation of transport characteristics between the bulk and surface of topological insulator materials by utilizing a comprehensive Physical Properties Measurement System (PPMS) and Terahertz Time-Domain Spectroscopy (THz-TDS) to extract electronic transport information for both bulk and surface states. Additionally, the dielectric polarization behavior of BSTS in the low-frequency (10–107 Hz) and high-frequency (0.5–2.0 THz) ranges was investigated. These research findings provide crucial experimental groundwork and theoretical guidance for the development of novel low-energy electronic devices, spintronic devices, and quantum computing technology based on topological insulators. Full article
(This article belongs to the Special Issue Physicochemical Research on Material Surfaces)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Schematic diagram of Laser Molecular Beam Epitaxy (LMBE) equipment preparation. (<b>b</b>) Schematic diagram of mixed-target assembly. (<b>c</b>) XRD diffraction pattern of the BSTS sample. (<b>d</b>) Schematic illustration of the molecular structure of the BSTS sample.</p>
Full article ">Figure 2
<p>(<b>a</b>) Surface morphology of the BSTS sample at 40.00 kx magnification in SEM characterization. (<b>b</b>) Surface morphology of the BSTS sample at 100.00 kx magnification in SEM characterization. (<b>c</b>) Cross-sectional morphology of the BSTS sample in AFM characterization. (<b>d</b>) Cross-sectional height profile of the BSTS sample, demonstrating a film thickness of 25 nm.</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic illustration of the dual-channel, bulk-surface transport in the BSTS and Bi<sub>2</sub>Se<sub>3</sub> sample and the principle of PPMS measurement. (<b>b</b>) Current–voltage (I-V) relationship of the BSTS and Bi<sub>2</sub>Se<sub>3</sub> sample at room temperature. (<b>c</b>) Relationship between bulk conductivity and temperature in the BSTS and Bi<sub>2</sub>Se<sub>3</sub> sample. (<b>d</b>) Relationship between Hall resistance and magnetic field variation in the BSTS and Bi<sub>2</sub>Se<sub>3</sub> sample. (<b>e</b>) Magnetoresistance variation of the BSTS and Bi<sub>2</sub>Se<sub>3</sub> sample with magnetic field changes in the range of −9T to 9T at 2K temperature.</p>
Full article ">Figure 4
<p>(<b>a</b>) Fundamental working principle of Terahertz Time-Domain Spectroscopy (THz-TDS). (<b>b</b>) Time-domain distribution of the Terahertz pulses for the BSTS sample and reference sample. (<b>c</b>) Terahertz transmission spectrum of the BSTS sample. (<b>d</b>) Phase variation of the BSTS sample in the frequency domain. (<b>e</b>) Real and imaginary parts of surface conductivity of the BSTS sample.</p>
Full article ">Figure 5
<p>(<b>a</b>) Curve showing the variation in dielectric constant of BSTS sample within the frequency range of 10–10<sup>7</sup> Hz. (<b>b</b>) Curve depicting the change in loss factor of the BSTS sample within the frequency range of 10–10<sup>7</sup> Hz. (<b>c</b>) Curve illustrating the fluctuation in dielectric constant of the BSTS sample within the frequency range of 0.5–2.5THz. (<b>d</b>) Curve displaying the alteration in loss factor of the BSTS sample within the frequency range of 0.5–2.5 THz.</p>
Full article ">
Back to TopTop