[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
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

Journals

Article Types

Countries / Regions

Search Results (198)

Search Parameters:
Keywords = Fano resonance

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 8579 KiB  
Article
Fano and Electromagnetically Induced Transparency Resonances in Dual Side-Coupled Photonic Crystal Nanobeam Cavities
by Yong Zhao, Yuxuan Chen and Lijun Hao
Materials 2024, 17(24), 6213; https://doi.org/10.3390/ma17246213 - 19 Dec 2024
Viewed by 351
Abstract
We propose two types of structures to achieve the control of Fano and electromagnetically induced transparency (EIT) line shapes, in which dual one-dimensional (1D) photonic crystal nanobeam cavities (PCNCs) are side-coupled to a bus waveguide with different gaps. For the proposed type Ⅰ [...] Read more.
We propose two types of structures to achieve the control of Fano and electromagnetically induced transparency (EIT) line shapes, in which dual one-dimensional (1D) photonic crystal nanobeam cavities (PCNCs) are side-coupled to a bus waveguide with different gaps. For the proposed type Ⅰ and type Ⅱ systems, the phase differences between the nanobeam periodic structures of the two cavities are π and 0, respectively. The whole structures are theoretically analyzed via the coupled mode theory and numerically demonstrated using the three-dimensional finite-difference time-domain (3D FDTD) method. The simulation results show that the proposed structure can achieve several kinds of spectra, including Fano, EIT and asymmetric EIT line shapes, which is dependent on the width of the bus waveguide. Compared to the previously proposed Fano resonator with 1D PCNCs, the proposed structures have the advantages of high transmission at the resonant peak, low insertion loss at non-resonant wavelengths, a wide free spectral range (FSR) and a high roll-off rate. Therefore, we believe the proposed structure can find broad applications in optical switches, modulators and sensors. Full article
Show Figures

Figure 1

Figure 1
<p>Simplified model of two side-coupled standing-wave cavities (<span class="html-italic">C</span><sub>1</sub> and <span class="html-italic">C</span><sub>2</sub>).</p>
Full article ">Figure 2
<p>The calculated transmission spectra of the (<b>a</b>) type Ⅰ and (<b>b</b>) type Ⅱ systems based on the temporary CMT. The coupling Q-factors are <span class="html-italic">Q<sub>w</sub></span><sub>1</sub> = 500 and <span class="html-italic">Q<sub>w</sub></span><sub>2</sub> = 2000, and the intrinsic Q-factors are <span class="html-italic">Q<sub>i</sub></span><sub>1</sub> = <span class="html-italic">Q<sub>i</sub></span><sub>2</sub> = 1 × 10<sup>5</sup>. The resonant wavelengths of <span class="html-italic">C</span><sub>1</sub> and <span class="html-italic">C</span><sub>2</sub> are <span class="html-italic">λ</span><sub>1</sub> = <span class="html-italic">λ</span><sub>2</sub> = 1550 nm in (<b>a</b>) and <span class="html-italic">λ</span><sub>1</sub> = 1550 nm and <span class="html-italic">λ</span><sub>2</sub> = 1549 nm in (<b>b</b>).</p>
Full article ">Figure 3
<p>Models of two side-coupled cavities with phase-shifted periodic structure under weak perturbation conditions: (<b>a</b>) Type Ⅰ system with same resonant wavelengths and grating phase difference of <span class="html-italic">π</span> between <span class="html-italic">C</span><sub>1</sub> and <span class="html-italic">C</span><sub>2</sub>. (<b>b</b>) Type Ⅱ system with different resonant wavelengths and same grating phase between <span class="html-italic">C</span><sub>1</sub> and <span class="html-italic">C</span><sub>2</sub>.</p>
Full article ">Figure 4
<p>The calculated transmission spectra of the (<b>a</b>) type Ⅰ and (<b>b</b>) type Ⅱ systems under weak perturbation conditions based on the coupled mode equations of Equations (6)–(8).</p>
Full article ">Figure 5
<p>Schematics of proposed (<b>a</b>) type Ⅰ and (<b>b</b>) type Ⅱ dual-PCNC systems on SOI platform.</p>
Full article ">Figure 6
<p>The reflection spectra of the designed Bragg reflector with 4 tapered holes and 9 uniform holes and the Bragg reflector with 13 uniform holes.</p>
Full article ">Figure 7
<p>A schematic of the calculation of the effective index of the strip waveguide with a hole.</p>
Full article ">Figure 8
<p>Simulated (<b>a</b>) transmission and (<b>b</b>) reflection spectra when only <span class="html-italic">C</span><sub>1</sub> or <span class="html-italic">C</span><sub>2</sub> is side-coupled to the bus waveguide.</p>
Full article ">Figure 9
<p>(<b>a</b>) The transmission and reflection spectra of the type Ⅰ dual-PCNC system with an EIT-like line shape, and the |<span class="html-italic">H<sub>y</sub></span>| profile at the (<b>b</b>) transmission peak point <span class="html-italic">A</span>, (<b>c</b>) transmission dip point <span class="html-italic">B</span> and (<b>d</b>) transmission dip point <span class="html-italic">C</span> calculated by the 3D FDTD method.</p>
Full article ">Figure 10
<p>(<b>a</b>) Transmission spectra and (<b>b</b>) phase of type Ⅰ dual-PCNC system with different bus waveguide widths calculated by 3D FDTD method. Red dot is point of maximum slope change in phase relative to wavelength.</p>
Full article ">Figure 11
<p>(<b>a</b>) The transmission and reflection spectra of the type Ⅰ dual-PCNC system with a Fano line shape (<span class="html-italic">w</span><sub>0</sub> = 435 nm), and the |<span class="html-italic">H<sub>y</sub></span>| profile at (<b>b</b>) point <span class="html-italic">A</span>, (<b>c</b>) point <span class="html-italic">B</span> and (<b>d</b>) point <span class="html-italic">C</span> calculated by the 3D FDTD method.</p>
Full article ">Figure 12
<p>(<b>a</b>) The transmission and reflection spectra of the type Ⅱ dual-PCNC system with an EIT-like line shape, and the |<span class="html-italic">H<sub>y</sub></span>| profile at the (<b>b</b>) transmission peak point <span class="html-italic">A</span>, (<b>c</b>) transmission dip point <span class="html-italic">B</span> and (<b>d</b>) transmission dip point <span class="html-italic">C</span> calculated by the 3D FDTD method.</p>
Full article ">Figure 13
<p>(<b>a</b>) The transmission spectra and (<b>b</b>) phase of the type Ⅱ dual-PCNC system with different bus waveguide widths calculated by the 3D FDTD method. The red dot is the point of the maximum slope change in phase relative to the wavelength.</p>
Full article ">Figure 14
<p>(<b>a</b>) The transmission and reflection spectra of the type Ⅱ dual-PCNC system with an asymmetric EIT line shape (<span class="html-italic">w</span><sub>0</sub> = 380) nm, and the |<span class="html-italic">H<sub>y</sub></span>| profile at (<b>b</b>) point <span class="html-italic">A</span>, (<b>c</b>) point <span class="html-italic">B</span> and (<b>d</b>) point <span class="html-italic">C</span> calculated by the 3D FDTD method.</p>
Full article ">Figure 15
<p>(<b>a</b>) The transmission and reflection spectra of the type Ⅱ dual-PCNC system with a Fano line shape and (<span class="html-italic">w</span><sub>0</sub> = 460), and the |<span class="html-italic">H<sub>y</sub></span>| profile at (<b>b</b>) point <span class="html-italic">A</span>, (<b>c</b>) point <span class="html-italic">B</span> and (<b>d</b>) point <span class="html-italic">C</span> calculated by the 3D FDTD method.</p>
Full article ">Figure 16
<p>The transmission spectra of the (<b>a</b>) type Ⅰ and (<b>b</b>) type Ⅱ dual-PCNC systems with EIT line shapes with different waveguide width deviations (∆<span class="html-italic">w</span>).</p>
Full article ">Figure 17
<p>The transmission spectra of the (<b>a</b>) type Ⅰ and (<b>b</b>) type Ⅱ dual-PCNC systems with EIT line shapes with different hole radius deviations (∆<span class="html-italic">r</span>).</p>
Full article ">
13 pages, 5166 KiB  
Article
A Design of Vanadium Dioxide for Dynamic Color Gamut Modulation Based on Fano Resonance
by Junyang Zhu, Ruimei Zeng, Yiwen Yang, Yiqun Zhou, Zhen Gao, Qi Wang, Ruijin Hong and Dawei Zhang
Crystals 2024, 14(12), 1096; https://doi.org/10.3390/cryst14121096 - 19 Dec 2024
Viewed by 357
Abstract
In this paper, a design of vanadium dioxide for dynamic color gamut modulation based on Fano resonance is proposed. This approach facilitates color modulation by manipulating the phase transition state of vanadium dioxide. The device integrates both broadband and narrowband filters, featuring a [...] Read more.
In this paper, a design of vanadium dioxide for dynamic color gamut modulation based on Fano resonance is proposed. This approach facilitates color modulation by manipulating the phase transition state of vanadium dioxide. The device integrates both broadband and narrowband filters, featuring a structure consisting of a top silver mesh, a layer of vanadium dioxide, and a Fabry–Pérot cavity, which allows for effective modulation of the reflectance spectrum. Simulation results demonstrate that when vanadium dioxide is in its insulating state, the maximum reflectivity observed in the device spectrum, reaching 43.1%, appears at 475 nm. Conversely, when vanadium dioxide transitions to its metallic state, the peak wavelength shifts to 688 nm, accompanied by an increased reflectance peak of 59.3%. Analysis of electric field distributions reveals that the intensity caused by surface plasmonic resonance dominates over the excited Fano resonance while vanadium dioxide is in its insulating state, which is the opposite of when vanadium dioxide transitions to its metallic state. This behavior exhibits an excellent dynamic color-tuning capability. Specifically, the phase transition of vanadium dioxide results in a color difference ∆E2000 of up to 36.7, while maintaining good color saturation. This technique holds significant potential for applications such as dynamic color display and anti-counterfeit labeling. Full article
(This article belongs to the Special Issue Preparation and Characterization of Optoelectronic Functional Films)
Show Figures

Figure 1

Figure 1
<p>Refractive index of VO<sub>2</sub>, (<b>a</b>) in the insulating state, (<b>b</b>) in the metallic state.</p>
Full article ">Figure 2
<p>Structure of the designed FROC; (<b>a</b>) top view, (<b>b</b>) front view. (<b>c</b>) Schematic of the FROC array.</p>
Full article ">Figure 3
<p>Reflection spectrum of (<b>a</b>) broadband filter, (<b>b</b>) narrowband filter.</p>
Full article ">Figure 4
<p>(<b>a</b>) Reflection spectrum when incident light vertically irradiates. (<b>b</b>) Chromaticity coordinates marked in the CIE1931 chromaticity diagram according to the spectrum shown in (<b>a</b>).</p>
Full article ">Figure 5
<p>Reflection spectra (<b>a</b>) of the broadband filter (upper part of the FROC), (<b>b</b>) of the FROC, when <span class="html-italic">h</span><sub>2</sub> is varied from 20 nm to 60 nm at a step of 10 nm. (<b>c</b>) Chromaticity coordinates marked in the CIE1931 chromaticity diagram according to the spectrum shown in (<b>b</b>). The black solid lines show the movement trajectory of the chromaticity coordinates when VO<sub>2</sub> is in its insulating state, and the white dashed lines show the movement trajectory of the chromaticity coordinates when VO<sub>2</sub> is in its metallic state.</p>
Full article ">Figure 6
<p>(<b>a</b>) Reflection spectrum of the FROC when <span class="html-italic">h</span><sub>3</sub> is varied from 10 nm to 40 nm at a step of 10 nm. (<b>b</b>) Chromaticity coordinates marked in the CIE1931 chromaticity diagram according to the spectrum shown in (<b>a</b>). The black solid lines show the movement trajectory of the chromaticity coordinates when VO<sub>2</sub> is in its insulating state, and the white dashed lines show the movement trajectory of the chromaticity coordinates when VO<sub>2</sub> is in its metallic state.</p>
Full article ">Figure 7
<p>Reflection spectra (<b>a</b>) of the upper broadband filter, (<b>b</b>) of the FROC, when <span class="html-italic">L</span> is varied from 190 nm to 130 nm at a step of 10 nm. (<b>c</b>) Chromaticity coordinates marked in the CIE1931 chromaticity diagram according to the spectrum shown in (<b>b</b>). The black solid lines show the movement trajectory of the chromaticity coordinates when VO<sub>2</sub> is in its insulating state, and the white dashed lines show the movement trajectory of the chromaticity coordinates when VO<sub>2</sub> is in its metallic state.</p>
Full article ">Figure 8
<p>Reflection spectra (<b>a</b>) of the upper broadband filter, (<b>b</b>) of the FROC, when <span class="html-italic">h</span><sub>4</sub> is varied from 30 nm to 70 nm at a step of 10 nm. (<b>c</b>) Chromaticity coordinates marked in the CIE1931 chromaticity diagram according to the spectrum shown in (<b>b</b>). The black solid lines show the movement trajectory of the chromaticity coordinates when VO<sub>2</sub> is in its insulating state, and the white dashed lines show the movement trajectory of the chromaticity coordinates when VO<sub>2</sub> is in its metallic state.</p>
Full article ">Figure 9
<p>Electric field distributions of the broadband filter in the x-z plane; (<b>a</b>) wavelength of 695 nm light normally illustrated on VO<sub>2</sub> (insulating state); (<b>b</b>) wavelength of 688 nm light vertically irradiates on VO<sub>2</sub> (metallic state). Electric field distributions of the FROC in the x-z plane; (<b>c</b>) wavelength of 695 nm light vertically irradiates on VO<sub>2</sub> (insulating state); (<b>d</b>) wavelength of 688 nm light vertically irradiates on VO<sub>2</sub> (metallic state). The structures of the broadband filter and FROC depicted by the solid black lines are shown in the figure.</p>
Full article ">Figure 10
<p>(<b>a</b>) “USST” pattern arranged by the array of the designed FROCs, which can be reversibly switched between visible and invisible by changing the temperature. The yellowish area is reflected by the structures without the top square grids (<span class="html-italic">L</span> is set as 0). (<b>b</b>) Butterfly patterns when VO<sub>2</sub> is in the insulating or metallic state and the length <span class="html-italic">L</span> is set as 190 nm or 160 nm, respectively.</p>
Full article ">
12 pages, 2570 KiB  
Article
Multifunctional SERS Chip for Biological Application Realized by Double Fano Resonance
by Weile Zhu, Huiyang Wang, Yuheng Wang, Shengde Liu, Jianglei Di and Liyun Zhong
Nanomaterials 2024, 14(24), 2036; https://doi.org/10.3390/nano14242036 - 19 Dec 2024
Viewed by 411
Abstract
The in situ and label-free detection of molecular information in biological cells has always been a challenging problem due to the weak Raman signal of biological molecules. The use of various resonance nanostructures has significantly advanced Surface-enhanced Raman spectroscopy (SERS) in signal enhancement [...] Read more.
The in situ and label-free detection of molecular information in biological cells has always been a challenging problem due to the weak Raman signal of biological molecules. The use of various resonance nanostructures has significantly advanced Surface-enhanced Raman spectroscopy (SERS) in signal enhancement in recent years. However, biological cells are often immersed in different formulations of culture medium with varying refractive indexes and are highly sensitive to the temperature of the microenvironment. This necessitates that SERS meets the requirements of refractive index insensitivity, low thermal damage, broadband enhancement, and other needs in addition to signal enhancement. Here, we propose a SERS chip with integrated dual Fano resonance and the corresponding analytical model. This model can be used to quickly lock the parameters and then analyze the performance of the dual resonance SERS chip. The simulation and experimental characterization results demonstrate that the integrated dual Fano resonances have the ability for independent broadband tuning. This capability enhances both the excitation and radiation processes of Raman signals simultaneously, ensuring that the resonance at the excitation wavelength is not affected by the culture medium (the refractive index) and reduces heat generation. Furthermore, the dual Fano resonance modes can synergize with each other to greatly enhance both the amplitude and enhanced range of the Raman signal, providing a stable, reliable, and comprehensive detection tool and strategy for fingerprint signal detection of bioactive samples. Full article
(This article belongs to the Section Biology and Medicines)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the SERS chip fabrication steps.</p>
Full article ">Figure 2
<p>The schematic of the proposed SERS chip, which consists of a 3-layer structured thin film system constructed on a silica substrate. The thickness of the upper and lower layers containing gold <span class="html-italic">t</span> = 50 nm, the thickness of PMMA film <span class="html-italic">h</span> = 120 nm, the diameter of the nanohole is <span class="html-italic">d</span>, and the period <span class="html-italic">p</span> = 315 nm.</p>
Full article ">Figure 3
<p>Formation mechanism of double Fano resonance in the proposed SERS chip. (<b>a</b>) The SERS chip with a subwavelength period is equivalent to a 3-layer thin film system under equivalent medium theory; (<b>b</b>) F-P resonance-like reflectance spectra in the equivalent 3-layer thin film system; (<b>c</b>) schematic diagram of the three resonance modes (F-P-like, slr, and SPP-Bloch) supported simultaneously in the SERS chip and their interactions; (<b>d</b>) the surface charge distribution of the SPP-Bloch resonance formation calculated by FEM simulation, where k represents the incident light wave vector and E represents the direction of electric field polarization; (<b>e</b>) reflectance spectra of SPP-Bloch mode with different nanopore periods coupling to the F-P mode, where the red curve represents the reflection spectrum with a nanohole period <span class="html-italic">p</span> = 315 nm; and (<b>f</b>) comparison of reflectance spectra in SERS chip when there is interaction between only two resonance modes (SPP-Bloch and F-P) and three resonance modes (SPP-Bloch, SLR, and F-P).</p>
Full article ">Figure 4
<p>Analytical model analysis of the double Fano resonance mechanism in the SERS chip. (<b>a</b>) Reflection mapping using the proposed analytical model; (<b>b</b>) reflection mapping using FEM simulation; (<b>c</b>,<b>d</b>) reflection spectrum (right ordinate) of nanohole diameter d = 220 nm (black dot in (<b>a</b>)) and d = 270 nm (blue dot in (<b>a</b>)) and the corresponding decomposition of resonance modes (left ordinate), respectively; (<b>e</b>,<b>f</b>) phase relationship among the three resonances generating the double Fano resonance corresponding to (<b>c</b>,<b>d</b>), respectively.</p>
Full article ">Figure 5
<p>(<b>a</b>) Resonance spectrum with refractive index changing from 1.33 to 1.5; (<b>b</b>) electric field enhancement spectra calculated using FEM simulation and analytical model (dot curve) while tunning the second Fano resonance.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) Experimental and simulated reflection spectrum of SERS chip with a nanohole diameter <span class="html-italic">d</span> =220 nm and <span class="html-italic">d</span> = 270 nm, respectively; (<b>c</b>) three-dimensional topography of the SERS chip with nanohole diameter <span class="html-italic">d</span> = 220, 250, and 270 nm characterized by atomic force microscopy; (<b>d</b>) comparison of Raman signals enhanced by SERS chips with nanohole diameters <span class="html-italic">d</span> = 220, <span class="html-italic">d</span> = 250, and <span class="html-italic">d</span> = 270 nm, respectively, with the inset showing a comparison of Raman signals from a flat gold film and a SERS chip with a nanohole diameter of <span class="html-italic">d</span> = 220 nm.</p>
Full article ">
10 pages, 2457 KiB  
Article
Angle-Controlled Nanospectrum Switching from Lorentzian to Fano Lineshapes
by Fu Tang, Qinyang Zhong, Xiaoqiuyan Zhang, Yuxuan Zhuang, Tianyu Zhang, Xingxing Xu and Min Hu
Nanomaterials 2024, 14(23), 1932; https://doi.org/10.3390/nano14231932 - 30 Nov 2024
Viewed by 508
Abstract
The tunability of spectral lineshapes, ranging from Lorentzian to Fano profiles, is essential for advancing nanoscale photonic technologies. Conventional far-field techniques are insufficient for studying nanoscale phenomena, particularly within the terahertz (THz) range. In this work, we use a U-shaped resonant ring on [...] Read more.
The tunability of spectral lineshapes, ranging from Lorentzian to Fano profiles, is essential for advancing nanoscale photonic technologies. Conventional far-field techniques are insufficient for studying nanoscale phenomena, particularly within the terahertz (THz) range. In this work, we use a U-shaped resonant ring on a waveguide substrate to achieve precise modulation of Lorentzian, Fano, and antiresonance profiles. THz scattering scanning near-field optical microscopy (s-SNOM) reveals the underlying physical mechanism of these transitions, driven by time-domain phase shifts between the background excitation from the waveguide and the resonance of the U-shaped ring. Our approach reveals a pronounced asymmetry in the near-field response, which remains undetectable in far-field systems. The ability to control spectral lineshapes at the nanoscale presents promising applications in characterizing composite nanoresonators and developing nanoscale phase sensors. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic diagram of the THz s-SNOM system. (<b>b</b>) Optical image and (<b>c</b>) second-order near-field white-light image of the U-shaped ring.</p>
Full article ">Figure 2
<p>(<b>a</b>) Experimental real of the near-field spectra from 0° to 90°. (<b>b</b>) Experimental and Fano formula fitted imaginary near-field spectra from 0° to 90°. (<b>c</b>) Simulated imaginary part of the spectrum as the U-shaped ring rotates from 0° to 90°. (<b>d</b>) Fano parameters <span class="html-italic">q</span> of experimental and simulated imaginary parts of the spectrum.</p>
Full article ">Figure 3
<p>(<b>a</b>) Near-field time-domain signal from 0° to 90°, (<b>b</b>,<b>c</b>) The enlarged view of the highlighted region of (<b>a</b>).</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>b</b>) Near-field spectra with the tip positioned in the left arm of the U-shaped ring, showing the real (<b>a</b>) and imaginary (<b>b</b>) parts from 0° to 180°. (<b>c</b>,<b>d</b>) Near-field spectra with the tip positioned in the right arm of the U-shaped ring, showing the real (<b>c</b>) and imaginary (<b>d</b>) parts from 0° to 180°.</p>
Full article ">Figure 5
<p>Simulated real (<b>a</b>) and imaginary parts (<b>b</b>) of the spectrum for waveguide substrate thicknesses <span class="html-italic">Z</span> from 8 to 28 µm. Simulated real (<b>c</b>) and imaginary parts (<b>d</b>) of the spectrum for <span class="html-italic">L</span> from 20 to 45 µm.</p>
Full article ">
8 pages, 3430 KiB  
Communication
Fano Resonance-Associated Plasmonic Circular Dichroism in a Multiple-Dipole Interaction Born–Kuhn Model
by Wanlu Bian, Guodong Zhu, Fengcai Ma, Tongtong Zhu and Yurui Fang
Sensors 2024, 24(23), 7517; https://doi.org/10.3390/s24237517 - 25 Nov 2024
Viewed by 469
Abstract
Plasmon chirality has garnered significant interest in sensing application due to its strong electromagnetic field localization and highly tunable optical properties. Understanding the effects of mode coupling in chiral structures on chiral optical activity is particularly important for advancing this field. In this [...] Read more.
Plasmon chirality has garnered significant interest in sensing application due to its strong electromagnetic field localization and highly tunable optical properties. Understanding the effects of mode coupling in chiral structures on chiral optical activity is particularly important for advancing this field. In this work, we numerically investigate the circular dichroism (CD) of elliptical nanodisk dimers arranged in an up-and-down configuration with a specific rotation angle. By adjusting the inter-particle distance and geometric parameters, we introduce the coupling between dipole and electric hexapole modes, forming an extended Born–Kuhn model that achieves strong CD. Our findings show that the coupling of dipole modes with electric hexapole modes in elliptical nanodisks can also show obvious Fano resonance and a strong CD effect, and the structure with the largest Fano asymmetry factor shows the highest CD. In addition, CD spectroscopy is highly sensitive to changes in the refractive index of the surrounding medium, especially in the visible and near-infrared regions, highlighting its potential for application in high-sensitivity refractive index sensors. Full article
(This article belongs to the Section Optical Sensors)
Show Figures

Figure 1

Figure 1
<p>Born–Kuhn model composed of two nanodisks. (<b>a</b>) Structure scheme, R<sub>1</sub> = 100 nm, R<sub>2</sub> = 22.5 nm, R<sub>3</sub> = 50 nm, R<sub>4</sub> = 400 nm, the thickness of the two particles H is 50 nm, the gap g between the upper and lower is 50 nm, the angle is 120°, and the light incident is perpendicular to the cross plane of the two nanodisks. (<b>b</b>) The extinction cross-section of individual nanodisks. The insets show the charge distributions of the nanodisks at the resonance wavelength of 800 nm. (<b>c</b>) The extinction cross-section of the structure excited by LCP and RCP. The insets show the surface charge distributions at the corresponding peaks. (<b>d</b>) CD spectrum.</p>
Full article ">Figure 2
<p>(<b>a</b>) Extinction cross-section of a dimer composed of small dimers. (<b>b</b>) The nano-configuration of a dimer composed of two large particles. CPL excites the extinction cross-section of the two large particles. (<b>c</b>) The extinction CD spectrum and the charge distribution of the upper and lower particles after excitation of the corresponding formant. (<b>d</b>) The extinction CD spectrum and the charge distribution of the upper and lower particles after the LCP and RCP formant excitation.</p>
Full article ">Figure 3
<p>(<b>a</b>) The extinction cross-section and the corresponding CD spectrum for the distance <span class="html-italic">g</span> between the two particles at 40nm, and the illustration shows the charge distribution at the corresponding resonance. (<b>b</b>) <span class="html-italic">g</span> = 30 nm. (<b>c</b>) <span class="html-italic">g</span> = 20 nm. (<b>d</b>) <span class="html-italic">g</span> = 10 nm.</p>
Full article ">Figure 4
<p>(<b>a</b>) The central dot of the small nanodisk overlaps with the vertex of the large nanodisk. (<b>b</b>) The vertex of the left end of the small nanodisk overlaps with the vertex of the large nanodisk. (<b>c</b>) The center point of the small nanodisk overlaps with the middle of the center point and the vertex of the large nanodisk. (<b>d</b>) The vertex of the right end of the small nanodisk overlaps with the vertex of the large nanodisk.</p>
Full article ">Figure 5
<p>(<b>a</b>) The CD spectrum after the surrounding medium environment n increased from 1.1 to 1.4. (<b>b</b>) The CD spectrum when the angle between the two particles increased from 10° to 80°.</p>
Full article ">
16 pages, 2769 KiB  
Article
A Reflective Terahertz Point Source Meta-Sensor with Asymmetric Meta-Atoms for High-Sensitivity Bio-Sensing
by Luwei Zheng, Kazuki Hara, Hironaru Murakami, Masayoshi Tonouchi and Kazunori Serita
Biosensors 2024, 14(12), 568; https://doi.org/10.3390/bios14120568 - 23 Nov 2024
Viewed by 528
Abstract
Biosensors operating in the terahertz (THz) region are gaining substantial interest in biomedical analysis due to their significant potential for high-sensitivity trace-amount solution detection. However, progress in compact, high-sensitivity chips and methods for simple, rapid and trace-level measurements is limited by the spatial [...] Read more.
Biosensors operating in the terahertz (THz) region are gaining substantial interest in biomedical analysis due to their significant potential for high-sensitivity trace-amount solution detection. However, progress in compact, high-sensitivity chips and methods for simple, rapid and trace-level measurements is limited by the spatial resolution of THz waves and their strong absorption in polar solvents. In this work, a compact nonlinear optical crystal (NLOC)-based reflective THz biosensor with a few arrays of asymmetrical meta-atoms was developed. A near-field point THz source was locally generated at a femtosecond-laser-irradiation spot via optical rectification, exciting only the single central meta-atom, thereby inducing Fano resonance. The reflective resonance response demonstrated dependence on several aspects, including structure asymmetricity, geometrical size, excitation point position, thickness and array-period arrangement. DNA samples were examined using 1 μL applied to an effective sensing area of 0.234 mm2 (484 μm × 484 μm) for performance evaluation. The developed Fano resonance sensor exhibited nearly double sensitivity compared to that of symmetrical sensors and one-gap split ring resonators. Thus, this study advances liquid-based sensing by enabling easy, rapid and trace-level measurements while also driving the development of compact and highly sensitive THz sensors for biological samples. Full article
(This article belongs to the Section Optical and Photonic Biosensors)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic of the experimental setup around the sample. (<b>b</b>) The geometric parameters and array-period arrangement of the three types of meta-atom structures developed. For the double-gap asymmetrical structure, the overall size (<b><span class="html-italic">L</span></b>) of one single meta-atom is 84 µm × 84 µm with a linewidth (<b><span class="html-italic">W</span></b>) of 10 µm and gaps (<b><span class="html-italic">G</span></b>) of 20 µm. One gap is shifted to one side with offset distance (<b><span class="html-italic">D</span></b>), which determines the asymmetry of the meta-structures. The meta-atom arrays are arranged in 3 × 3 and 5 × 5 patterns with periods of 100 µm and 120 µm, respectively. The gap direction is at an angle of 54 degrees to the [001] crystal orientation of the GaAs substrate. The laser irradiation spot of ~<span class="html-italic">ϕ</span> 20 µm is at the center of the central meta-atom. For comparison, two additional meta-atom structures were developed, as shown at the bottom of the diagram: double-gap and single-gap symmetrical structures. In the double-gap structures, the electric field aligns parallel to the gap direction, while in the single-gap structure, the electric field is oriented perpendicular to the gap direction, as indicated by the black arrows. The gap size (<b><span class="html-italic">G</span><sub>1</sub></b>) of the single-gap structure is 5 µm.</p>
Full article ">Figure 2
<p>Reflective resonance analysis with the meta-structure arrayed in a 5 × 5 pattern with a period of 100 µm and gap offset distance of 20 µm. (<b>a</b>) The calculated and measured reflectance spectra. There main resonances are observed: 1st resonance, 2nd resonance (Fano resonance), and 3rd resonance, which are indicated by black and red arrows for simulation and experiment, respectively. (<b>b</b>) The corresponding electric field distribution of the 2nd resonance (Fano resonance) at 0.425THz. A strong electric filed coupling region was observed around the shorter rod, as the red arrow indicates. (<b>c</b>) The calculated 2nd resonance frequency spectra of the meta-structure immersed in materials with different refractive indices. (<b>d</b>) The calculated resonance frequency versus refractive index with a linear fitting result. The evaluated sensitivity is 43.02 GHz/RIU.</p>
Full article ">Figure 3
<p>Geometric size dependence. The simulated reflectance spectra of the meta-atom structures with different geometric parameters. The overall dimensions (<b><span class="html-italic">L</span></b>) of each meta-atom were set as 82 μm × 82 μm, 84 μm × 84 μm, and 86 μm × 86 μm, respectively. The linewidth (<b><span class="html-italic">W</span></b>) was defined as 8 μm, 10 μm, and 12 μm, and the gap sizes (<b><span class="html-italic">G</span></b>) were set at 18 μm, 20 μm, and 22 μm. These three parameters were varied in combination, resulting in a total of seven different cases.</p>
Full article ">Figure 4
<p>Position dependence of the excitation of point THz sources. (<b>a</b>) Schematic of laser excitation point positions: P<sub>0</sub>(0, 0), P<sub>1</sub>(−12.5, 12.5), P<sub>2</sub>(12.5, 12.5), P<sub>3</sub>(−12.5, −12.5), P<sub>4</sub>(50, 50), P<sub>5</sub>(0, 50), and P<sub>6</sub>(50, 0). (<b>b</b>) The corresponding reflectance spectra at each laser excitation point position.</p>
Full article ">Figure 5
<p>Thickness dependence. The reflectance spectra of the meta-atom structures with different thickness from 10 nm to 160 nm.</p>
Full article ">Figure 6
<p>Gap offset distance, array number and period dependences of the developed meta structures. The calculated (<b>a</b>,<b>b</b>) measured reflectance spectra of meta-structures with changing offset distance from 0 µm to 20 µm with a step of 2 µm and 5 µm, respectively. The array number and period are fixed at 5 × 5 and 100 µm, respectively. The calculated (<b>c</b>) and measured (<b>d</b>) reflectance spectra of meta-structures with different arrays and periods.</p>
Full article ">Figure 7
<p>dsDNA and ssDNA measurements. Reflectance spectra (<b>a</b>,<b>c</b>,<b>e</b>) of the meta-sensor with and without dsDNA and ssDNA in a single measurement. Average resonance frequency shifts (<b>b</b>,<b>d</b>,<b>f</b>) of dsDNA and ssDNA of ten measurements. (<b>a</b>,<b>b</b>) are the results of double-gap asymmetrical meta structure. (<b>c</b>,<b>d</b>) are the results of double-gap symmetrical meta structure. (<b>e</b>,<b>f</b>) are the results of single-gap symmetrical meta structure. The concentration of the measured dsDNA and ssDNA are 100 µg/mL. The array number and period of the three meta-structures are set as 5 × 5 and 100 µm, respectively.</p>
Full article ">
10 pages, 2113 KiB  
Article
Kondo Versus Fano in Superconducting Artificial High-Tc Heterostructures
by Gaetano Campi, Gennady Logvenov, Sergio Caprara, Antonio Valletta and Antonio Bianconi
Condens. Matter 2024, 9(4), 43; https://doi.org/10.3390/condmat9040043 - 31 Oct 2024
Viewed by 811
Abstract
Recently, the quest for high-Tc superconductors has evolved from the trial-and-error methodology to the growth of nanostructured artificial high-Tc superlattices (AHTSs) with tailor-made superconducting functional properties by quantum design. Here, we report the growth by molecular beam epitaxy (MBE) of a superlattice of [...] Read more.
Recently, the quest for high-Tc superconductors has evolved from the trial-and-error methodology to the growth of nanostructured artificial high-Tc superlattices (AHTSs) with tailor-made superconducting functional properties by quantum design. Here, we report the growth by molecular beam epitaxy (MBE) of a superlattice of Mott insulator metal interfaces (MIMIs) made of nanoscale superconducting layers of quantum confined-space charge in the Mott insulator La2CuO4 (LCO), with thickness L intercalated by normal metal La1.55Sr0.45CuO4 (LSCO) with period d. The critical temperature shows the superconducting dome with Tc as a function of the geometrical parameter L/d showing the maximum at the magic ratio L/d = 2/3 where the Fano–Feshbach resonance enhances the superconducting critical temperature. The normal state transport data of the samples at the top of the superconducting dome exhibit Planckian T-linear resistivity. For L/d > 2/3 and L/d < 2/3, the heterostructures show a resistance following Kondo universal scaling predicted by the numerical renormalization group theory for MIMI nanoscale heterostructures. We show that the Kondo temperature, TK, and the Kondo scattering amplitude, R0K, vanish at L/d = 2/3, while TK and R0K increase at both sides of the superconducting dome, indicating that the T-linear resistance regime competes with the Kondo proximity effect in the normal phase of MIMIs. Full article
(This article belongs to the Special Issue Superstripes Physics, 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Pictorial view of a practical realization of the nanoscale AHTS superlattice of quantum wells made of four monolayers (L = 2.64 nm) of undoped La<sub>2</sub>CuO<sub>4</sub> (LCO), electronically doped by the interface space charge, which are intercalated by normal metal units made of two monolayers (W = 1.32 nm) of La<sub>1.55</sub>Sr<sub>0.45</sub>CuO<sub>4</sub> (LSCO), forming a superlattice with a period of d = L + W = 3.96 nm and geometric parameter L/d = 2/3, giving the optimum critical temperature. (<b>b</b>) The superconductivity dome of the critical temperature, T<sub>C</sub>, as a function of L/d in the range of 0.25 &lt; L/d &lt; 0.9. The dashed line represents the predictions by the BPV theory at a Fano–Feshbach shape resonance for the superlattice of the quantum wells. We show the Tc determined as maximum of a derivative of the sheet resistance (blue circles) and maximum of the imaginary part of the mutual inductance (red squares). (<b>c</b>) Sheet resistance as a function of temperature of several LSCO/LCO superlattices where it is normalized at R<sub>s 150K</sub>, which is the resistance measured at 150 K. The samples show the maximum Tc ≈ 43 K around L/d = 2/3, corresponding to a critical temperature in the optimum-doped LSCO. We outline a linear behavior (tick black line), observing how the samples with a maximum Tc, with L/d = 2/3, approach this linear resistivity regime. (<b>d</b>) The normalized sheet resistance as a function of temperature of two LSCO/LCO superlattices with L/d = 2/3 and different periods d = 3.96 nm and 2.97 nm, showing T-linear resistivity in the temperature range 50 K &lt; T &lt; 270 K.</p>
Full article ">Figure 2
<p>(<b>a</b>) (Open circles) Sheet resistance as a function of temperature of 18 superlattice samples with different d and L values indicated in the list on the left. The different d-values correspond to the different colors indicated. (<b>b</b>) Normalized sheet resistance as a function of temperature alongside the modeled lines through Equation (1). We show both the (left panel) linear and the (right panel) logarithmic evolution with temperature. (<b>c</b>) Color maps of normalized sheet resistance as a function of temperature, the geometrical parameter L/d. In panel (<b>d</b>), we show the correspondence between L/d and &lt;δ&gt; = 0.45 × (1 − L/d). The white dashed line indicates the L/d = 2/3 value.</p>
Full article ">Figure 3
<p>(<b>a</b>) Sheet resistance measured at 150 K in all 18 measured samples with different L/d values. We observed an exponential increase for L/d &gt; 2/3. Evolution of the fitting parameters (<b>b</b>) A, (<b>c</b>) B, (<b>d</b>) r<sub>0</sub> + R<sub>0k</sub>, and (<b>e</b>) T<sub>k</sub> in Equation (1). (<b>f</b>) Scatter plot of T<sub>k</sub> versus R<sub>0k</sub> showing the positive correlation between these two parameters.</p>
Full article ">Figure 4
<p>T<sub>K</sub> (left y-axis) and T<sub>C</sub> (right y-axis) as a function of L/d. The three red lines represent the theoretical T<sub>C</sub> in the superlattice with period 2.97 (red solid line) and 3:96 nm (dashed red line) as a function of L = d calculated in [<a href="#B8-condensedmatter-09-00043" class="html-bibr">8</a>].</p>
Full article ">
12 pages, 7395 KiB  
Article
Multi-Cavity Nanorefractive Index Sensor Based on MIM Waveguide
by Weijie Yang, Shubin Yan, Ziheng Xu, Changxin Chen, Jin Wang, Xiaoran Yan, Shuwen Chang, Chong Wang and Taiquan Wu
Nanomaterials 2024, 14(21), 1719; https://doi.org/10.3390/nano14211719 - 28 Oct 2024
Viewed by 615
Abstract
Within this manuscript, we provide a novel Fano resonance-driven micro-nanosensor. Its primary structural components are a metal-insulator-metal (MIM) waveguide, a shield with three disks, and a T-shaped cavity (STDTC). The finite element approach was used to study the gadget in theory. It is [...] Read more.
Within this manuscript, we provide a novel Fano resonance-driven micro-nanosensor. Its primary structural components are a metal-insulator-metal (MIM) waveguide, a shield with three disks, and a T-shaped cavity (STDTC). The finite element approach was used to study the gadget in theory. It is found that the adjustment of the structure and the change of the dimensions are closely related to the sensitivity (S) and the quality factor (FOM). Different model structural parameters affect the Fano resonance, which in turn changes the transmission characteristics of the resonator. Through in-depth experimental analysis and selection of appropriate parameters, the sensor sensitivity finally reaches 3020 nm/RIU and the quality factor reaches 51.89. Furthermore, the installation of this microrefractive index sensor allows for the quick and sensitive measurement of glucose levels. It is a positive contribution to the field of optical devices and micro-nano sensors and meets the demand for efficient detection when applied in medical and environmental scenarios. Full article
(This article belongs to the Section Nanoelectronics, Nanosensors and Devices)
Show Figures

Figure 1

Figure 1
<p>Two-dimensional schematic diagram of a shield-shaped structure with three discs and a T-shaped cavity.</p>
Full article ">Figure 2
<p>(<b>a</b>) Transmission spectra of different structures; (<b>b</b>) Shield structure with T-cavity.</p>
Full article ">Figure 3
<p>Magnetic field strength distribution (<b>a</b>) Single shield cavity; (<b>b</b>) Shield with T-cavity; (<b>c</b>) STDTC.</p>
Full article ">Figure 4
<p>(<b>a</b>) Transmission curves with varying R<sub>1</sub> lengths; (<b>b</b>) Sensitivity fit lines with varying R<sub>1</sub> lengths; (<b>c</b>) Comparing FWHM with FOM for various R<sub>1</sub> lengths.</p>
Full article ">Figure 5
<p>(<b>a</b>) Transmission curves with varying L lengths; (<b>b</b>) Sensitivity fit lines with varying L lengths; (<b>c</b>) Comparing FWHM with FOM for various L lengths.</p>
Full article ">Figure 6
<p>(<b>a</b>) Transmission curves with varying r<sub>1</sub> lengths; (<b>b</b>) Sensitivity fit lines with varying r<sub>1</sub> lengths.</p>
Full article ">Figure 7
<p>(<b>a</b>) Transmission curves with varying g lengths; (<b>b</b>) Comparing FWHM with FOM for various g lengths.</p>
Full article ">Figure 8
<p>(<b>a</b>) Transmission spectra for various indexes of refraction; (<b>b</b>) Refractive index sensitivity fit lines.</p>
Full article ">Figure 9
<p>(<b>a</b>) Transmission spectrum of glucose concentration; (<b>b</b>) Fitted line of glucose concentration.</p>
Full article ">
11 pages, 508 KiB  
Article
Quantum Interference Effects on Josephson Current through Quadruple-Quantum-Dot Molecular Inserted between Superconductors
by Yumei Gao, Yaohong Shen, Feng Chi, Zichuan Yi and Liming Liu
Micromachines 2024, 15(10), 1225; https://doi.org/10.3390/mi15101225 - 30 Sep 2024
Viewed by 773
Abstract
We study theoretically the Josephson current through a junction composed of quadruple quantum dots (QDs), of which only one is coupled directly to the left and right superconductor leads (denoted by QD1). The other three QDs are side-coupled to QD1 and free from [...] Read more.
We study theoretically the Josephson current through a junction composed of quadruple quantum dots (QDs), of which only one is coupled directly to the left and right superconductor leads (denoted by QD1). The other three QDs are side-coupled to QD1 and free from coupling to the leads. It is found that when the energy levels of all the four QDs are identical, the Josephson current varying with energy level of QD1 develops three peaks with two narrow and one wide, showing the typical Dicke lineshape. With increasing inter-dot coupling strength, the triple-peak configuration is well retained and accompanied by an obviously increased current amplitude. The critical current as a function of the energy level of QD1 shows a single resonance peak whose position and height depend on the energy levels of the side-coupled QDs and the inter-dot coupling strengths. We also find that the curve of the critical current versus energy levels of the side-coupled QDs shows a pair of Fano resonances and the same number Fano antiresonances (valleys). When the energy levels of the side-coupled QDs are different from each other, another Fano resonance and antiresonance are induced due to the quantum interference effect. The present results are compared with those in double and triple QDs systems, and may serve as unique means, such as the combination of quantum Dicke and Fano effects, to manipulate the Josehpson currents. Full article
(This article belongs to the Special Issue Quantum Tunneling Devices and Sensors)
Show Figures

Figure 1

Figure 1
<p>The schematic diagram of the quadruple quantum dots sandwiched between the left and right superconductors with energy gap <math display="inline"><semantics> <mo>Δ</mo> </semantics></math>. Only quantum dot 1 with energy level <math display="inline"><semantics> <msub> <mi>ε</mi> <mn>1</mn> </msub> </semantics></math> couples directly to the superconductors with a strength of <math display="inline"><semantics> <mo>Γ</mo> </semantics></math>, and interacts simultaneously with quantum dot 2 with energy level <math display="inline"><semantics> <msub> <mi>ε</mi> <mn>2</mn> </msub> </semantics></math> by coupling strength <math display="inline"><semantics> <msub> <mi>t</mi> <mn>1</mn> </msub> </semantics></math>. Quantum dot 2 is further connected to quantum dots 3 and 4 individually having energy levels <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>2</mn> </msub> <mo>+</mo> <mi>δ</mi> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>2</mn> </msub> <mo>−</mo> <mi>δ</mi> </mrow> </semantics></math> by the same coupling strength <math display="inline"><semantics> <msub> <mi>t</mi> <mn>2</mn> </msub> </semantics></math>.</p>
Full article ">Figure 2
<p>Contour plots of the Josephson currents <span class="html-italic">J</span> as a function of (<math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>0</mn> </msub> <mo>,</mo> <mi>ϕ</mi> </mrow> </semantics></math>) in (<b>a</b>), and (<math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>0</mn> </msub> <mo>,</mo> <msub> <mi>t</mi> <mn>0</mn> </msub> </mrow> </semantics></math>) in (<b>b</b>). Other parameters are indicated in the figures.</p>
Full article ">Figure 3
<p>Josephson current <span class="html-italic">J</span> as a function of phase bias <math display="inline"><semantics> <mi>ϕ</mi> </semantics></math> for various <math display="inline"><semantics> <msub> <mi>t</mi> <mn>2</mn> </msub> </semantics></math> in (<b>a</b>), and different <math display="inline"><semantics> <mi>δ</mi> </semantics></math> in (<b>b</b>) for the indicated parameters.</p>
Full article ">Figure 4
<p>Critical Josephson current <math display="inline"><semantics> <msub> <mi>J</mi> <mi>c</mi> </msub> </semantics></math> as a function of energy level of QD1 for positive <math display="inline"><semantics> <msub> <mi>ε</mi> <mn>2</mn> </msub> </semantics></math> in (<b>a</b>), and negative <math display="inline"><semantics> <msub> <mi>ε</mi> <mn>2</mn> </msub> </semantics></math> in (<b>b</b>) for the indicated parameters.</p>
Full article ">Figure 5
<p>Josephson current as a function of <math display="inline"><semantics> <msub> <mi>ε</mi> <mn>1</mn> </msub> </semantics></math> for fixed <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>2</mn> </msub> <mo>=</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.3</mn> <mo>Δ</mo> </mrow> </semantics></math> and different <math display="inline"><semantics> <msub> <mi>t</mi> <mn>2</mn> </msub> </semantics></math> in (<b>a</b>), different <math display="inline"><semantics> <mi>δ</mi> </semantics></math> in (<b>b</b>) for the indicated parameters.</p>
Full article ">Figure 6
<p>Josephson current as a function of <math display="inline"><semantics> <msub> <mi>ε</mi> <mn>2</mn> </msub> </semantics></math> for varying <math display="inline"><semantics> <msub> <mi>ε</mi> <mn>1</mn> </msub> </semantics></math>. Other parameters are <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>t</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.3</mn> <mo>Δ</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>δ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Josephson current as a function of <math display="inline"><semantics> <msub> <mi>ε</mi> <mn>2</mn> </msub> </semantics></math> for fixed <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.3</mn> <mo>Δ</mo> </mrow> </semantics></math> and different <math display="inline"><semantics> <msub> <mi>t</mi> <mn>2</mn> </msub> </semantics></math> in (<b>a</b>), different <math display="inline"><semantics> <mi>δ</mi> </semantics></math> in (<b>b</b>) for the indicated parameters.</p>
Full article ">
13 pages, 4279 KiB  
Article
Highly Sensitive Plasmon Refractive Index Sensor Based on MIM Waveguide
by Wen Jiang, Shubin Yan, Xiaoran Yan, Aiwei Xu, Guang Liu, Chong Wang, Lei Li, Xiangyang Mu and Guowang Gao
Micromachines 2024, 15(8), 987; https://doi.org/10.3390/mi15080987 - 30 Jul 2024
Viewed by 947
Abstract
This paper introduces a novel plasmon refractive index nanosensor structure based on Fano resonance. The structure comprises a metal–insulator–metal (MIM) waveguide with an inverted rectangular cavity and a circle minus a small internal circle plus a rectangular cavity (CMSICPRC). This study employs the [...] Read more.
This paper introduces a novel plasmon refractive index nanosensor structure based on Fano resonance. The structure comprises a metal–insulator–metal (MIM) waveguide with an inverted rectangular cavity and a circle minus a small internal circle plus a rectangular cavity (CMSICPRC). This study employs the finite element method (FEM) to analyze the sensing characteristics of the structure. The results demonstrate that the geometrical parameters of specific structures exert a considerable influence on the sensing characteristics. Simulated experimental data show that the maximum sensitivity of this structure is 3240 nm/RIU, with a figure of merit (FOM) of 52.25. Additionally, the sensor can be used in biology, for example, to detect the concentration of hemoglobin in blood. The sensitivity of the sensor in this application, according to our calculations, can be 0.82 nm∙g/L. Full article
(This article belongs to the Special Issue Advances in Photodetecting Materials, Devices and Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic two-dimensional (2D) layout of the designed sensor fabric.</p>
Full article ">Figure 2
<p>Transmission spectra of single ring, single stub circular cavity, single CMSICPRC, and all systems.</p>
Full article ">Figure 3
<p>Sensitivity fit lines for different structures.</p>
Full article ">Figure 4
<p>Transmission spectra of a single short rod in a circle at different angles.</p>
Full article ">Figure 5
<p>(<b>a</b>) Transmission spectra of different radii a; (<b>b</b>) sensitivity fitted lines of different radii a; (<b>c</b>) variation in FWHM values of different radii a.</p>
Full article ">Figure 6
<p>(<b>a</b>) Transmission spectra for different lengths h; (<b>b</b>) detail view at (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math>+ <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> </mrow> </semantics></math>)/2.</p>
Full article ">Figure 7
<p>(<b>a</b>) Transmission spectra at different distances g; (<b>b</b>) sensitivity fit lines at different distances g; (<b>c</b>) variation in FWHM values at different distances g.</p>
Full article ">Figure 8
<p>(<b>a</b>) Transmission spectra for different lengths of L; (<b>b</b>) sensitivity fit lines for different lengths of L.</p>
Full article ">Figure 9
<p>(<b>a</b>) Transmission spectra of different radii R; (<b>b</b>) sensitivity fitted lines of different radii R; (<b>c</b>) variation in FWHM and FOM values of different radii R.</p>
Full article ">Figure 10
<p>(<b>a</b>) Transmission spectra of different concentrations of H; (<b>b</b>) sensitivity fit curves for different hemoglobin concentrations.</p>
Full article ">
13 pages, 6217 KiB  
Article
High-Performance Refractive Index and Temperature Sensing Based on Toroidal Dipole in All-Dielectric Metasurface
by Jingjing Zhao, Xinye Fan, Wenjing Fang, Wenxing Xiao, Fangxin Sun, Chuanchuan Li, Xin Wei, Jifang Tao, Yanling Wang and Santosh Kumar
Sensors 2024, 24(12), 3943; https://doi.org/10.3390/s24123943 - 18 Jun 2024
Cited by 2 | Viewed by 1253
Abstract
This article shows an all-dielectric metasurface consisting of “H”-shaped silicon disks with tilted splitting gaps, which can detect the temperature and refractive index (RI). By introducing asymmetry parameters that excite the quasi-BIC, there are three distinct Fano resonances with nearly 100% modulation depth, [...] Read more.
This article shows an all-dielectric metasurface consisting of “H”-shaped silicon disks with tilted splitting gaps, which can detect the temperature and refractive index (RI). By introducing asymmetry parameters that excite the quasi-BIC, there are three distinct Fano resonances with nearly 100% modulation depth, and the maximal quality factor (Q-factor) is over 104. The predominant roles of different electromagnetic excitations in three distinct modes are demonstrated through near-field analysis and multipole decomposition. A numerical analysis of resonance response based on different refractive indices reveals a RI sensitivity of 262 nm/RIU and figure of merit (FOM) of 2183 RIU−1. This sensor can detect temperature fluctuations with a temperature sensitivity of 59.5 pm/k. The proposed metasurface provides a novel method to induce powerful TD resonances and offers possibilities for the design of high-performance sensors. Full article
(This article belongs to the Special Issue Novel Optical Biosensing Technology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Diagram of the metasurface array. (<b>b</b>) Diagram of the unit metasurface. (<b>c</b>) The <span class="html-italic">x</span>–<span class="html-italic">z</span> side view of the unit metasurface.</p>
Full article ">Figure 2
<p>(<b>a</b>) The transmission spectra curves of metasurface structure at symmetry and symmetry breaking. (<b>b</b>) The blue solid line shows the simulation curve for the resonant mode FR3, and the red dashed line indicates the fitted curve.</p>
Full article ">Figure 3
<p>(<b>a</b>) Transmission spectra of metasurface at different values of tilt angle <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>.</mo> </mrow> </semantics></math> (<b>b</b>) The correlation of the Q-factor and <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>α</mi> </mrow> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math> of the modes FR1 and FR3.</p>
Full article ">Figure 4
<p>(<b>a</b>) The variation of the transmission spectrum with the splitting gap <span class="html-italic">g</span> when other parameters are held constant. (<b>b</b>) The variation of the transmission spectrum with the silicon disc width <span class="html-italic">w</span> when other parameters are held constant.</p>
Full article ">Figure 5
<p>Multipole decomposition of three resonance modes.</p>
Full article ">Figure 6
<p>The near-field distributions of the normalized electric and magnetic fields of the nanostructure in the three resonance modes. The vectorial distribution of the electric field in the resonance modes is indicated with white arrows. The vector distribution of the magnetic field in the resonance modes is indicated with black arrows.</p>
Full article ">Figure 7
<p>Transmission spectra of asymmetric structures at <span class="html-italic">θ</span> = 5<math display="inline"><semantics> <mrow> <mo>°</mo> </mrow> </semantics></math> for different loss levels k.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>b</b>) Transmission spectra of asymmetric structures at different polarization angles <math display="inline"><semantics> <mrow> <mi>φ</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>(<b>a</b>) Transmission spectra of resonant modes at various temperatures. (<b>b</b>) The relationship between wavelength shift and temperature is analyzed in three resonance modes.</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>c</b>) Transmission spectra of the proposed metasurface structure with refractive indices of 1, 1.01, 1.02, 1.03, and 1.04, respectively. (<b>d</b>) Correlation between wavelength shift and refractive index.</p>
Full article ">Figure 11
<p>Manufacturing process for metasurface structures.</p>
Full article ">
14 pages, 3246 KiB  
Article
Nanosensor Based on the Circular Ring with External Rectangular Ring Structure
by Shuwen Chang, Shubin Yan, Yiru Su, Jin Wang, Yuhao Cao, Yi Zhang, Taiquan Wu and Yifeng Ren
Photonics 2024, 11(6), 568; https://doi.org/10.3390/photonics11060568 - 17 Jun 2024
Cited by 1 | Viewed by 783
Abstract
This paper presents a novel nanoscale refractive index sensor, which is produced by using a metal–insulator–metal (MIM) waveguide structure coupled with the circular ring with an external rectangular ring (CRERR) structure with the Fano resonance phenomenon. In this study, COMSOL software was used [...] Read more.
This paper presents a novel nanoscale refractive index sensor, which is produced by using a metal–insulator–metal (MIM) waveguide structure coupled with the circular ring with an external rectangular ring (CRERR) structure with the Fano resonance phenomenon. In this study, COMSOL software was used to model and simulate the structure, paired with an analysis of the output spectra to detail the effect of constructional factors on the output Fano curve as measured from a finite element method. After a series of studies, it was shown that an external rectangular ring is the linchpin of the unsymmetrical Fano resonance, while the circular ring’s radius strongly influences the transducer’s capability to achieve a maximum for 3180 nm/RIU sensitivity and a FOM of 54.8. The sensor is capable of achieving sensitivities of 0.495 nm/mgdL−1 and 0.6375 nm/mgdL−1 when detecting the concentration of the electrolyte sodium and potassium ions in human blood and is expected to play an important role in human health monitoring. Full article
(This article belongs to the Special Issue New Perspectives in Optical Design)
Show Figures

Figure 1

Figure 1
<p>Schematic two-dimensional (2D) layout of the designed sensor fabric.</p>
Full article ">Figure 2
<p>Transmission spectra of the entire structure (black line), a single CRERR structure (red line) and only two baffle structures (blue line).</p>
Full article ">Figure 3
<p>(<b>a</b>) Normalized distribution of magnetic field at λ = 2232 nm for the CRERR structure; (<b>b</b>) normalized distribution of magnetic field at λ = 2243 nm for the whole system.</p>
Full article ">Figure 4
<p>(<b>a</b>) Transmission spectra at variable refractive indices; (<b>b</b>) sensitivity fitted lines at variable refractive indices.</p>
Full article ">Figure 5
<p>Transmission spectra of rectangular annular cavity rotated at different angles.</p>
Full article ">Figure 6
<p>(<b>a</b>) Transmission spectra at different heights of the rectangular annular cavity; (<b>b</b>) sensitivity fitted lines at different heights of the rectangular annular cavity; (<b>c</b>) variation in FWHM values at different heights of the rectangular annular cavity.</p>
Full article ">Figure 7
<p>(<b>a</b>) Transmission spectra of different Rs; (<b>b</b>) sensitivity fitted lines of different Rs.</p>
Full article ">Figure 8
<p>(<b>a</b>) Transmission spectra of a rectangular baffle at variable heights; (<b>b</b>) variation in FWHM values at different heights of a rectangular baffle.</p>
Full article ">Figure 9
<p>(<b>a</b>) Transmission spectra obtained with various coupling gaps; (<b>b</b>) sensitivity fitted lines with various coupling gaps; (<b>c</b>) variation in FWHM values with various coupling gaps.</p>
Full article ">Figure 10
<p>(<b>a</b>) Transmission spectra of different concentrations of <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>N</mi> <mi>a</mi> </mrow> <mrow> <mo>+</mo> </mrow> </msup> </mrow> </semantics></math>; (<b>b</b>) transmission spectra of different concentrations of <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>K</mi> </mrow> <mrow> <mo>+</mo> </mrow> </msup> </mrow> </semantics></math>; (<b>c</b>) sensitivity fit lines for different concentrations of <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>N</mi> <mi>a</mi> </mrow> <mrow> <mo>+</mo> </mrow> </msup> </mrow> </semantics></math>; (<b>d</b>) sensitivity fit lines for different concentrations of <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>K</mi> </mrow> <mrow> <mo>+</mo> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">
11 pages, 2797 KiB  
Communication
Sensing Characteristic Analysis of All-Dielectric Metasurfaces Based on Fano Resonance in Near-Infrared Regime
by Yongpeng Zhao, Qingfubo Geng, Jian Liu and Zhaoxin Geng
Photonics 2024, 11(5), 482; https://doi.org/10.3390/photonics11050482 - 20 May 2024
Viewed by 1204
Abstract
A novel, all-dielectric metasurface, featuring a missing wedge-shaped nanodisk, is proposed to investigate optical characteristics. By introducing symmetry-breaking to induce Fano resonance, the metasurface achieves an impressive Q-factor of 1202 in the near-infrared spectrum, with a remarkably narrow full width at half maximum [...] Read more.
A novel, all-dielectric metasurface, featuring a missing wedge-shaped nanodisk, is proposed to investigate optical characteristics. By introducing symmetry-breaking to induce Fano resonance, the metasurface achieves an impressive Q-factor of 1202 in the near-infrared spectrum, with a remarkably narrow full width at half maximum (FWHM) of less than 1 nm. The ability to adjust the wavelength resonance by manipulating the structure of the wedge-shaped nanodisk offers a simple and efficient approach for metasurface design. This breakthrough holds great potential for various applications in sensing and optical filtering, marking a significant advancement in the field of nanophotonics. Full article
(This article belongs to the Special Issue Advanced Photonic Sensing and Measurement II)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the all-dielectric metasurfaces and the working principle. (<b>a</b>) Array structure consists of the missing wedge-shaped nanodisk. (<b>b</b>) Top view and cross-section of a cell structure. (<b>c</b>) Top row shows the αe electric dipole mode and αm magnetic dipole mode of the incident x-polarized light excited nanodisk. (<b>d</b>) Bottom row shows symmetry-breaking and induced coupling of the αm magnetic dipole to the αe electric dipole.</p>
Full article ">Figure 2
<p>Reflection spectra and electric field distribution (<span class="html-italic">Ey</span>) of a metasurfaces cell. (<b>a</b>) Reflection spectra of the whole nanodisk. (<b>b</b>) Reflection spectra of the missing wedge-shaped nanodisk. (<b>c</b>) Electric field distribution (<span class="html-italic">Ey</span>) at different wavelength positions (800 nm (M1), 950 nm (M2), 1000 nm (M3), 800 nm (M4), 938 nm (M5), and 1000 nm (M6)).</p>
Full article ">Figure 3
<p>Reflection spectra at different slice angles and the radii of the nanodisk. (<b>a</b>) Reflection spectra of the missing wedge-shaped nanodisk with different <span class="html-italic">θ</span> (130–180°) at R = 160 nm. (<b>b</b>) Reflection spectra of the missing wedge-shaped nanodisk with different <span class="html-italic">R</span> (130–180 nm) at <span class="html-italic">θ</span> = 160°.</p>
Full article ">Figure 4
<p>Electric field distribution of the missing wedge-shaped nanodisk (<span class="html-italic">θ</span> = 170° and <span class="html-italic">R</span> = 180 nm) at the resonance wavelength (<span class="html-italic">λ</span> = 938 nm). (<b>a</b>) Electric field distribution in the <span class="html-italic">x</span>–<span class="html-italic">y</span> plane. (<b>b</b>) Linear variation of the electric field distribution (normalized) in the <span class="html-italic">x</span>–<span class="html-italic">y</span> plane. (<b>c</b>) Electric field distribution in the <span class="html-italic">x–z</span> plane. (<b>d</b>) Linear variation of the electric field distribution (normalized) in the <span class="html-italic">x</span>–<span class="html-italic">z</span> plane.</p>
Full article ">Figure 5
<p>Reflection spectra and <math display="inline"><semantics> <mrow> <mi>F</mi> <mi>W</mi> <mi>H</mi> <mi>M</mi> </mrow> </semantics></math> of the missing wedge-shaped silicon nanodisk at different parameters. (<b>a</b>) Reflection spectra changes with different refractive indices (<span class="html-italic">n</span> = 1.33–1.39) at <span class="html-italic">R</span> = 180 nm, <span class="html-italic">θ</span> = 170°. (<b>b</b>) Reflection spectra at different periods (<span class="html-italic">P</span> = 470–510 nm, at <span class="html-italic">n</span> = 1.33, <span class="html-italic">R</span> = 180 nm, <span class="html-italic">θ</span> = 170°). (<b>c</b>) Reflection spectra of the missing wedge-shaped nanodisk at different θ (130–180°) with <span class="html-italic">R</span> = 160 nm. (<b>d</b>) Reflectance spectra of the missing wedge-shaped nanodisk at different <span class="html-italic">R</span> (150–200 nm) with <span class="html-italic">θ</span> = 160°. (<b>e</b>) Evolution of the resonance wavelength and <math display="inline"><semantics> <mrow> <mi>F</mi> <mi>W</mi> <mi>H</mi> <mi>M</mi> </mrow> </semantics></math> at <span class="html-italic">θ</span> of 130–180°. (<b>f</b>) Evolution of the resonance wavelength and <span class="html-italic">FWHM</span> at <span class="html-italic">R</span> of 150–200 nm and <span class="html-italic">θ</span> of 180°.</p>
Full article ">
14 pages, 3319 KiB  
Article
The Application of Optical Sensors with Built-in Anchor-like Cavities in the Detection of Hemoglobin Concentration
by Wen Jiang, Shubin Yan, Yiru Su, Chong Wang, Taiquan Wu, Yang Cui, Chuanhui Zhu, Yi Zhang, Xiangyang Mu and Guowang Gao
Photonics 2024, 11(5), 402; https://doi.org/10.3390/photonics11050402 - 26 Apr 2024
Viewed by 1226
Abstract
This paper introduces a refractive index sensor based on Fano resonance, utilizing a metal–insulator–metal (MIM) waveguide structure with an Anchor-like cavity. This study utilizes the finite element method (FEM) for analyzing the propagation characteristics of the structure. The evaluation concentrated on assessing how [...] Read more.
This paper introduces a refractive index sensor based on Fano resonance, utilizing a metal–insulator–metal (MIM) waveguide structure with an Anchor-like cavity. This study utilizes the finite element method (FEM) for analyzing the propagation characteristics of the structure. The evaluation concentrated on assessing how the refractive index and the structure’s geometric parameters affect its sensing characteristics. The designed structure demonstrates optimum performance, achieving a maximum sensitivity of 2440 nm/RIU and an FOM of 63. Given its high sensitivity, this nanoscale refractive index sensor is ideal for detecting hemoglobin concentrations in blood, and the sensor’s sensitivity is 0.6 nm·g/L, aiding in clinical prevention and treatment. Full article
(This article belongs to the Special Issue New Perspectives in Optical Design)
Show Figures

Figure 1

Figure 1
<p>Two-dimensional plan of the overall structure.</p>
Full article ">Figure 2
<p>(<b>a</b>) Transmission spectra of single rectangular stub, all systems, and a single Anchor-like structure. (<b>b</b>) Magnetic field distribution for different structures.</p>
Full article ">Figure 3
<p>(<b>a</b>) Transmittance spectrum at different refractive indices and (<b>b</b>) sensitivity fitting curve.</p>
Full article ">Figure 4
<p>(<b>a</b>) The transmittance spectra for radii ranging from 75 to 95 nm. (<b>b</b>) The sensitivity fitting curves for different radii a.</p>
Full article ">Figure 5
<p>(<b>a</b>) The transmittance spectra for l ranging from195 to 235 nm. (<b>b</b>) The sensitivity fitting curves for different l.</p>
Full article ">Figure 6
<p>(<b>a</b>) The transmittance spectra for g ranging from10 to 50 nm, (<b>b</b>) the sensitivity fitting curves for different g, and (<b>c</b>) the trend of FWHM change.</p>
Full article ">Figure 7
<p>The transmittance spectra for h ranging from 45 to 85 nm.</p>
Full article ">Figure 8
<p>(<b>a</b>) The transmittance spectra for 2<span class="html-italic">φ</span> ranging from 120° to 160°, (<b>b</b>) the sensitivity fitting curves for different 2<span class="html-italic">φ</span>.</p>
Full article ">Figure 9
<p>(<b>a</b>) The transmission spectra for l being the same, with 2φ values of 120°and 160°. (<b>b</b>) The distribution of the magnetic field at a wavelength of 2082 nm when 2φ is 120°. (<b>c</b>) The distribution of the magnetic field at a wavelength of 2082 nm when 2φ is 160°.</p>
Full article ">Figure 10
<p>(<b>a</b>) The transmittance spectra for H ranging from 0 to 300 g/L. (<b>b</b>) The sensitivity fitting curve of hemoglobin concentration changes.</p>
Full article ">
10 pages, 3124 KiB  
Communication
Multipolar Analysis in Symmetrical Meta-Atoms Sustaining Fano Resonances
by Vittorio Bonino and Angelo Angelini
Optics 2024, 5(2), 238-247; https://doi.org/10.3390/opt5020017 - 15 Apr 2024
Viewed by 1174
Abstract
We present an optical metasurface with symmetrical individual elements sustaining Fano resonances with high Q-factors. This study combines plane-wave illumination and modal analysis to investigate the resonant behavior that results in a suppression of the forward scattering, and we investigate the role of [...] Read more.
We present an optical metasurface with symmetrical individual elements sustaining Fano resonances with high Q-factors. This study combines plane-wave illumination and modal analysis to investigate the resonant behavior that results in a suppression of the forward scattering, and we investigate the role of the lattice constant on the excited multipoles and on the spectral position and Q-factor of the Fano resonances, revealing the nonlocal nature of the resonances. The results show that the intrinsic losses play a crucial role in modulating the resonance amplitude in specific conditions and that the optical behavior of the device is extremely sensitive to the pitch of the metasurface. The findings highlight the importance of near-neighbor interactions to achieve high Q resonances and offer an important tool for the design of spectrally tunable metasurfaces using simple geometries. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Three-dimensional schematic of the simulated metasurface. (<b>b</b>) Real (blue) and imaginary (red) part of the refractive index used for the nanopillars. (<b>c</b>–<b>h</b>) Graphs showing transmission (blue line), reflection (green line), and absorption (red line) as functions of wavelength for the metasurface, with pitches ranging from 330 nm to 380 nm.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>f</b>) Relative intensity of contributions from different multipolar resonances, including the electric dipole (ED), magnetic dipole (MD), electric quadrupole (EQ), and magnetic quadrupole (MQ), as well as the total sum of all contributions (TOT), assessed over a range of pitch values from 330 nm to 380 nm. On the left axis, the transmittance profile (T) is reported in red for the reader’s convenience.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>c</b>) Norm of the electric field intensity on the XZ plane, passing through the center of a pillar and one of its near neighbors. The black arrows indicate the orientation of the electric field component on the XZ plane. The electric field was evaluated for pillars with three different pitches <span class="html-italic">P</span>, (<b>a</b>) 330 nm, (<b>b</b>) 350 nm, and (<b>c</b>) 380 nm, and at their corresponding resonant wavelengths <math display="inline"><semantics> <mi>λ</mi> </semantics></math>, (<b>a</b>) 656 nm, (<b>b</b>) 669 nm, and (<b>c</b>) 692 nm. (<b>d</b>–<b>f</b>) Schematic representation of the multipole displacement within the pillars. Dipoles with more intense coloring indicate a higher contribution for that specific combination of pitch and wavelength (<math display="inline"><semantics> <mi>λ</mi> </semantics></math>).</p>
Full article ">Figure 4
<p>(<b>a</b>) Quality factors of the resonances for pillars with pitches ranging from 300 nm to 380 nm. (<b>b</b>) Difference in the scattering cross sections of the electric quadrupole (EQ) and magnetic dipole (MD) for the corresponding pitches (blue line). The red dashed line highlights the zero value.</p>
Full article ">Figure 5
<p>Resonance wavelength variation with the pitch of the metasurface (blue dashed line). The error bars indicate the full width at half maximum (FWHM) of the resonances (blue solid lines). Figure of Merit (red solid circles) evaluated for varying pitches.</p>
Full article ">
Back to TopTop