[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
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (173)

Search Parameters:
Keywords = graphene field-effect transistor

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 4348 KiB  
Article
Basic Characteristics of Ionic Liquid-Gated Graphene FET Sensors for Nitrogen Cycle Monitoring in Agricultural Soil
by Naoki Shiraishi, Jian Lu, Fatin Bazilah Fauzi, Ryo Imaizumi, Toyohiro Tsukahara, Satoshi Mogari, Shosuke Iida, Yusuke Matsukura, Satoshi Teramoto, Keisuke Yokoi, Izumi Ichinose and Mutsumi Kimura
Biosensors 2025, 15(1), 55; https://doi.org/10.3390/bios15010055 - 16 Jan 2025
Viewed by 283
Abstract
Nitrogen-based fertilizers are crucial in agriculture for maintaining soil health and increasing crop yields. Soil microorganisms transform nitrogen from fertilizers into NO3–N, which is absorbed by crops. However, some nitrogen is converted to nitrous oxide (N2O), a [...] Read more.
Nitrogen-based fertilizers are crucial in agriculture for maintaining soil health and increasing crop yields. Soil microorganisms transform nitrogen from fertilizers into NO3–N, which is absorbed by crops. However, some nitrogen is converted to nitrous oxide (N2O), a greenhouse gas with a warming potential about 300-times greater than carbon dioxide (CO2). Agricultural activities are the main source of N2O emissions. Monitoring N2O can enhance soil health and optimize nitrogen fertilizer use, thereby supporting precision agriculture. To achieve this, we developed ionic liquid-gated graphene field-effect transistor (FET) sensors to measure N2O concentrations in agricultural soil. We first fabricated and tested the electrical characteristics of the sensors. Then, we analyzed their transfer characteristics in our developed N2O evaluation system using different concentrations of N2O and air. The sensors demonstrated a negative shift in transfer characteristic curves when exposed to N2O, with a Dirac point voltage difference of 0.02 V between 1 and 10 ppm N2O diluted with pure air. These results demonstrate that the ionic liquid-gated graphene FET sensor is a promising device for N2O detection for agricultural soil applications. Full article
(This article belongs to the Special Issue Application of Biosensors in Environmental Monitoring)
Show Figures

Figure 1

Figure 1
<p>Schematic of N<sub>2</sub>O monitoring in agricultural soil using an ionic liquid-gated graphene FET sensor: (<b>a</b>) spatial variation and chemical variation in nitrogen in agricultural soil; (<b>b</b>) schematic illustration and (<b>c</b>–<b>e</b>) working principle of an ionic liquid-gated graphene FET sensor in chamber.</p>
Full article ">Figure 2
<p>Fabrication process of an ionic liquid-gated graphene FET sensor: (<b>a</b>) graphene etching using O<sub>2</sub> plasma; (<b>b</b>) Cr/Au/Cr deposition using vapor deposition and patterning using a lift-off process; (<b>c</b>) 100 nm Au coating on electrodes using a lift-off process; (<b>d</b>) ionic liquid dropped; (<b>e</b>) photograph of graphene pattern formed using O<sub>2</sub> plasma; (<b>f</b>) photograph of a processed wafer; (<b>g</b>) SEM image of a fabricated sensor chip containing 6 graphene FETs with the same graphene channel; (<b>h</b>) enlarged SEM image of channel 1; (<b>i</b>) photograph of an ionic liquid-gated graphene FET sensor; (<b>j</b>) enlarged photograph of channel 1.</p>
Full article ">Figure 3
<p>Fabrication process of an ionic liquid-gated graphene FET sensor with a hydrophobic layer: (<b>a</b>) graphene etching using O<sub>2</sub> plasma; (<b>b</b>) Au/Cr deposition using vapor deposition and patterning using a lift-off process; (<b>c</b>) CYTOP formation; (<b>d</b>) etching of Au/Cr on graphene channel; (<b>e</b>) ionic liquid dropped; (<b>f</b>) photograph of a processed wafer with a hydrophobic layer; (<b>g</b>) SEM image of a fabricated sensor chip containing 6 FETs with different graphene channel sizes; (<b>h</b>) enlarged SEM image of channel 5; (<b>i</b>) photograph of an ionic liquid-gated graphene FET sensor; (<b>j</b>) enlarged photograph of channel 5.</p>
Full article ">Figure 4
<p>Current–voltage characteristics of the source–drain electrodes of the ionic-gated graphene FET sensor without a hydrophobic layer. The Y-axis of I<sub>DS</sub> represents the source–drain current. The X-axis of V<sub>DS</sub> represents the source–drain voltage.</p>
Full article ">Figure 5
<p>(<b>a</b>) Schematic view of our N<sub>2</sub>O evaluation system; (<b>b</b>) photograph of the thermos-hydrostat chamber; (<b>c</b>) evaluation module of the wire-bonded sensor on a ceramic package.</p>
Full article ">Figure 6
<p>Transfer characteristic curve of the 200 nL [PMIM][BF<sub>4</sub>]-ionic liquid-gated graphene FET sensor with a hydrophobic layer exposed to air, followed by 10 ppm N<sub>2</sub>O diluted with N<sub>2</sub> and then by air at a constant 1000 sccm for 10 min. The Y-axis of I<sub>DS</sub> represents the source–drain current. The X-axis of V<sub>DS</sub> represents the source–drain voltage.</p>
Full article ">Figure 7
<p>Transfer characteristic curve of the 100 nL [PMIM][BF<sub>4</sub>]-ionic liquid-gated graphene FET sensor without a hydrophobic layer exposed to air, 1 ppm N<sub>2</sub>O diluted with pure air, air, and lastly 10 ppm N<sub>2</sub>O diluted with pure air at 1000 sccm for 30 min. The Y-axis of I<sub>DS</sub> represents the source–drain current. The X-axis of V<sub>DS</sub> represents the source–drain voltage.</p>
Full article ">Figure 8
<p>(<b>a</b>) Transfer characteristic curve of the 100 nL [PMIM][BF<sub>4</sub>]-ionic liquid-gated graphene FET sensor without a hydrophobic layer exposed to 0.02 ppm N<sub>2</sub>O diluted with pure air and 1 ppm and 10 ppm N<sub>2</sub>O diluted with pure air at 1000 sccm for 5 min. The Y-axis of I<sub>DS</sub> represents the source–drain current. The X-axis of V<sub>DS</sub> represents the source–drain voltage; (<b>b</b>) N<sub>2</sub>O concentration dependence of source–drain current at a gate voltage of 0.3 V. The I<sub>DS</sub> at a gate voltage of 0.3 V of 0.02 ppm N<sub>2</sub>O diluted with pure air and 1 ppm and 10 ppm N<sub>2</sub>O diluted with pure air in <a href="#biosensors-15-00055-f008" class="html-fig">Figure 8</a>a were plotted.</p>
Full article ">Figure 9
<p>Transfer characteristic curve of the 100 nL [PMIM][BF<sub>4</sub>]-ionic liquid-gated graphene FET sensor without a hydrophobic layer exposed to 1 ppm and 10 ppm N<sub>2</sub>O diluted with pure air and 50,000 ppm CO<sub>2</sub> diluted with pure air at 1000 sccm for 5 min. The Y-axis of I<sub>DS</sub> represents the source–drain current. The X-axis of V<sub>DS</sub> represents the source–drain voltage.</p>
Full article ">Figure 10
<p>(<b>a</b>) Photograph of parylene C-coated ionic-gated graphene FET sensor; (<b>b</b>) enlarged micrograph of ch1 of parylene C-coated ionic-gated graphene FET sensor; (<b>c</b>) enlarged micrograph of ch1 before the application of the parylene C coating.</p>
Full article ">Figure 11
<p>Dirac point voltage shift of 100 nL [PMIM][BF4]-ionic liquid-gated graphene FET sensor exposed to pure water vapor: (<b>a</b>) with parylene C coating; (<b>b</b>) without parylene C.</p>
Full article ">
11 pages, 3979 KiB  
Article
An Enhanced Verilog-A Model for Graphene Field-Effect Transistors Using Variable Fermi Velocity
by Shuwei Ji, John Mappes, Peter Koudelka, Maximilian C. Scardelletti, Christian Zorman and Hossein Miri Lavasani
Electronics 2024, 13(24), 5051; https://doi.org/10.3390/electronics13245051 - 23 Dec 2024
Viewed by 372
Abstract
This paper presents a novel Verilog-A model for the Fermi velocity in Graphene Field-Effect Transistors (GFETs). The Fermi velocity is an important parameter associated with the energy spectrum of the delocalized bonds in graphene which impact the performance of a GFET. Unlike existing [...] Read more.
This paper presents a novel Verilog-A model for the Fermi velocity in Graphene Field-Effect Transistors (GFETs). The Fermi velocity is an important parameter associated with the energy spectrum of the delocalized bonds in graphene which impact the performance of a GFET. Unlike existing GFET models where the Fermi velocity is assumed to have a constant value, the proposed model considers carrier concentrations in the channel and gate dielectrics to create a closed-form solution for the Fermi velocity, a parameter previously demonstrated to vary based on these two factors. The proposed mathematical model is then adapted to Verilog-A for interfacing with computer-aided design (CAD) circuit simulators. To demonstrate the accuracy of the proposed model, the simulation results are compared to measured drain–source currents obtained from various GFET devices (including GFETs measured by authors). The measured results show good agreement with the values predicted using the proposed model (<±1%), demonstrating the superior accuracy of the model compared to other published Verilog-A-based models, especially around the Dirac point. Full article
Show Figures

Figure 1

Figure 1
<p>The equivalent circuit model used to solve for the Fermi velocity and (<b>a</b>,<b>b</b>) to solve for the carrier concentration of the channel at the drain and source, respectively.</p>
Full article ">Figure 2
<p>Schematic diagrams of the voltage bias states in a GFET: (<b>a</b>) electron conduction, (<b>b</b>,<b>c</b>) ambipolar conduction with opposite channel voltage bias, and (<b>d</b>) hole conduction. The images were adapted from [<a href="#B20-electronics-13-05051" class="html-bibr">20</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) A comparison of the proposed effective mobility approximation with prior work and the actual mobility and (<b>b</b>) the relative difference between the actual mobility and the mobility estimated using the two models (new and old) [<a href="#B5-electronics-13-05051" class="html-bibr">5</a>].</p>
Full article ">Figure 4
<p>A circuit diagram of the proposed small-signal model of the GFET using current sources influenced by <span class="html-italic">V<sub>ch</sub></span> and the averaged Fermi velocity <span class="html-italic">v<sub>F</sub></span> varied by top and back gate biasing voltages. The contact resistance and parasitic capacitance are also shown.</p>
Full article ">Figure 5
<p>Simulated drain and source Fermi velocity values for drain source voltage at (<b>a</b>) 0.35 V and (<b>b</b>) 1.1 V using this work’s model with input data from Han Wang et al.’s device [<a href="#B21-electronics-13-05051" class="html-bibr">21</a>].</p>
Full article ">Figure 6
<p>Comparison of simulated I-V characteristics to measurement data from Wang et al. [<a href="#B21-electronics-13-05051" class="html-bibr">21</a>]. (<b>a</b>) I–V characteristics with constant <span class="html-italic">V<sub>GS</sub></span>; (<b>b</b>) I–V characteristics with constant <span class="html-italic">V<sub>BS</sub></span>.</p>
Full article ">Figure 7
<p>A comparison of the <span class="html-italic">I<sub>DS</sub></span> vs. <span class="html-italic">V<sub>BS</sub></span> model with the measured results from Graphenea’s GFET.</p>
Full article ">Figure 8
<p>Simulation of <span class="html-italic">I<sub>DS</sub></span> vs. <span class="html-italic">V<sub>BS</sub></span> when <span class="html-italic">V<sub>DS</sub></span> varies from 0. 49 V to 0.50 V.</p>
Full article ">
14 pages, 2036 KiB  
Article
New Label-Free DNA Nanosensor Based on Top-Gated Metal–Ferroelectric–Metal Graphene Nanoribbon on Insulator Field-Effect Transistor: A Quantum Simulation Study
by Khalil Tamersit, Abdellah Kouzou, José Rodriguez and Mohamed Abdelrahem
Nanomaterials 2024, 14(24), 2038; https://doi.org/10.3390/nano14242038 - 19 Dec 2024
Viewed by 456
Abstract
In this paper, a new label-free DNA nanosensor based on a top-gated (TG) metal–ferroelectric–metal (MFM) graphene nanoribbon field-effect transistor (TG-MFM GNRFET) is proposed through a simulation approach. The DNA sensing principle is founded on the dielectric modulation concept. The computational method employed to [...] Read more.
In this paper, a new label-free DNA nanosensor based on a top-gated (TG) metal–ferroelectric–metal (MFM) graphene nanoribbon field-effect transistor (TG-MFM GNRFET) is proposed through a simulation approach. The DNA sensing principle is founded on the dielectric modulation concept. The computational method employed to evaluate the proposed nanobiosensor relies on the coupled solutions of a rigorous quantum simulation with the Landau–Khalatnikov equation, considering ballistic transport conditions. The investigation analyzes the effects of DNA molecules on nanodevice behavior, encompassing potential distribution, ferroelectric-induced gate voltage amplification, transfer characteristics, subthreshold swing, and current ratio. It has been observed that the feature of ferroelectric-induced gate voltage amplification using the integrated MFM structure can significantly enhance the biosensor’s sensitivity to DNA molecules, whether in terms of threshold voltage shift or drain current variation. Additionally, we propose the current ratio as a sensing metric due to its ability to consider all DNA-induced modulations of electrical parameters, specifically the increase in on-state current and the decrease in off-state current and subthreshold swing. The obtained results indicate that the proposed negative-capacitance GNRFET-based DNA nanosensor could be considered an intriguing option for advanced point-of-care testing. Full article
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Three-dimensional structure of the label-free DNA sensor based on TG-MFM GNRFET. (<b>b</b>) DNA detection based on the dielectric modulation concept. (<b>c</b>) Lengthwise cut view of the proposed nanoscale biosensor.</p>
Full article ">Figure 2
<p>(<b>a</b>) Flowchart of the computational method used. (<b>b</b>) Drain current values from the literature and our simulator and the P–E proprieties from L–K theory and reported experiment data for the ferroelectric hafnium zirconium oxide.</p>
Full article ">Figure 3
<p>Two-dimensional electron potential distribution at V<sub>DS</sub> = 0.3 V and V<sub>GS</sub> = 0.1 V for baseline TG GNRFET-based biosensor (<b>top figures</b>) and TG-MFM GNRFET-based biosensor (<b>bottom figures).</b> (<b>a</b>,<b>c</b>) Empty open cavity. (<b>b</b>,<b>d</b>) Cavity filled with DNA molecules.</p>
Full article ">Figure 4
<p>V<sub>G-INT</sub> versus V<sub>G-EXT</sub> for the proposed biosensor with different ferroelectric thicknesses considering (<b>a</b>) an empty and (<b>b</b>) filled sensing cavity.</p>
Full article ">Figure 5
<p>The I<sub>DS</sub>-V<sub>GS</sub> characteristics for (<b>a</b>) the baseline and (<b>b</b>) the proposed biosensor considering different DNA dielectric constants.</p>
Full article ">Figure 6
<p>Subthreshold swing (<b>a</b>) and current ratio (<b>b</b>) as functions of DNA dielectric constant for the TG-MFM GNRFET-based DNA sensor.</p>
Full article ">
27 pages, 11141 KiB  
Review
A Review of Bandgap Engineering and Prediction in 2D Material Heterostructures: A DFT Perspective
by Yoonju Oh, Seunghyun Song and Joonho Bae
Int. J. Mol. Sci. 2024, 25(23), 13104; https://doi.org/10.3390/ijms252313104 - 6 Dec 2024
Viewed by 1040
Abstract
The advent of two-dimensional (2D) materials and their capacity to form van der Waals (vdW) heterostructures has revolutionized numerous scientific fields, including electronics, optoelectronics, and energy storage. This paper presents a comprehensive investigation of bandgap engineering and band structure prediction in 2D vdW [...] Read more.
The advent of two-dimensional (2D) materials and their capacity to form van der Waals (vdW) heterostructures has revolutionized numerous scientific fields, including electronics, optoelectronics, and energy storage. This paper presents a comprehensive investigation of bandgap engineering and band structure prediction in 2D vdW heterostructures utilizing density functional theory (DFT). By combining various 2D materials, such as graphene, hexagonal boron nitride (h-BN), transition metal dichalcogenides, and blue phosphorus, these heterostructures exhibit tailored properties that surpass those of individual components. Bandgap engineering represents an effective approach to addressing the limitations inherent in material properties, thereby providing enhanced functionalities for a range of applications, including transistors, photodetectors, and solar cells. Furthermore, this study discusses the current limitations and challenges associated with bandgap engineering in 2D heterostructures and highlights future prospects aimed at unlocking their full potential for advanced technological applications. Full article
Show Figures

Figure 1

Figure 1
<p>Van der Waals heterostructures. (<b>a</b>) 0D nanoparticles or QDs, (<b>b</b>) 1D nanowires, (<b>c</b>) 2D nanosheets, and (<b>d</b>) 3D bulk materials.</p>
Full article ">Figure 2
<p>Procedure for the bandgap engineering of heterostructures via DFT. Reprinted with permission from [<a href="#B48-ijms-25-13104" class="html-bibr">48</a>], Copyright 2017, Springer Nature. Reprinted with permission from Loan et al. [<a href="#B3-ijms-25-13104" class="html-bibr">3</a>], Copyright 2014, Wiley-CH Verlag GmbH &amp; Co. KGaA. Adapted with permission from [<a href="#B49-ijms-25-13104" class="html-bibr">49</a>], Copyright 2014, American Chemical Society. Adapted from [<a href="#B40-ijms-25-13104" class="html-bibr">40</a>], Copyright 2022, with permission from Elsevier. Adapted from [<a href="#B42-ijms-25-13104" class="html-bibr">42</a>], Copyright 2020, with permission from Elsevier. Adapted from [<a href="#B50-ijms-25-13104" class="html-bibr">50</a>], Copyright 2021, with permission from Elsevier.</p>
Full article ">Figure 3
<p>(<b>a</b>) Band structure of Te−NR on graphene and MoS<sub>2</sub> in both lateral (L) and vertical (V) configurations. Te, graphene, Mo, and S are represented in the fatband as red, green, blue, and yellow, respectively Adapted with permission from [<a href="#B66-ijms-25-13104" class="html-bibr">66</a>]. Copyright 2022 American Chemical Society. (<b>b</b>) Top and side views of the SnSNR/graphene heterostructure, and (<b>c</b>) band structures of F−, Cl−, and Br-passivated heterostructures. Adapted from [<a href="#B67-ijms-25-13104" class="html-bibr">67</a>], Copyright 2018, with permission from Elsevier. (<b>d</b>) Optimized G/MoS<sub>2</sub> heterostructure and the electronic structures of monolayer graphene, monolayer MoS<sub>2</sub>, and the G/MoS<sub>2</sub> heterostructure. The calculated band structures of G/MoS<sub>2</sub> under (<b>e</b>) strain and (<b>f</b>) electric field. Adapted from [<a href="#B69-ijms-25-13104" class="html-bibr">69</a>], Copyright 2018, with permission from Elsevier. Adapted from [<a href="#B70-ijms-25-13104" class="html-bibr">70</a>], Copyright 2018, with permission from Elsevier.</p>
Full article ">Figure 4
<p>(<b>a</b>) Optimized structures of P– and V–MoS<sub>2</sub>/BN/G heterostructures and their electronic structures under (<b>b</b>) various electric field strengths and (<b>c</b>) strain. Used with permission from the Royal Society of Chemistry, from Ref. [<a href="#B71-ijms-25-13104" class="html-bibr">71</a>]; permission conveyed through Copyright Clearance Center, Inc. (<b>d</b>) Designed GaS/graphene heterostructure and its band structures under diverse strain conditions. Reprinted from [<a href="#B73-ijms-25-13104" class="html-bibr">73</a>], with the permission of AIP Publishing. (<b>e</b>) Equilibrium configuration of graphene/GeC and the calculated band structures under varying (<b>f</b>) electric field and (<b>g</b>) strain. Adapted from [<a href="#B74-ijms-25-13104" class="html-bibr">74</a>], Copyright 2019, with permission from Elsevier.</p>
Full article ">Figure 5
<p>(<b>a</b>) Most stable structure of graphene/SiH and its projected electronic structures under diverse conditions of (<b>b</b>) electric field and (<b>c</b>) strain. Used with permission of the Royal Society of Chemistry, from Ref. [<a href="#B75-ijms-25-13104" class="html-bibr">75</a>]; permission conveyed through Copyright Clearance Center, Inc. (<b>d</b>) Optimized structure of graphene/antimonene/graphene and the resulting band structures of monolayer antimonene, monolayer graphene, and the heterostructure. Reprinted from [<a href="#B76-ijms-25-13104" class="html-bibr">76</a>], Copyright 2019, with permission from Elsevier. (<b>e</b>) Relaxed atomic structures of two types of In<sub>2</sub>Se<sub>3</sub>/graphene configurations and their corresponding computed electronic structures. Reprinted from [<a href="#B77-ijms-25-13104" class="html-bibr">77</a>], Copyright 2019, with permission from Elsevier.</p>
Full article ">Figure 6
<p>(<b>a</b>) Computed band structures with and without strain for single−layer MoS<sub>2</sub>, single−layer WS<sub>2</sub>, and WS<sub>2</sub>/MoS<sub>2</sub>, and (<b>b</b>) a plot showing bandgap as a function of strain for these materials. Reprinted from [<a href="#B15-ijms-25-13104" class="html-bibr">15</a>], with the permission of AIP Publishing. (<b>c</b>) Relaxed structures of monolayer MoS<sub>2</sub> and WS<sub>2</sub>, as well as heterostructures of V− and P/MoS<sub>2</sub>/WS<sub>2</sub>, along with the calculated electronic structures of the corresponding designed structures with and without NO<sub>2</sub>. Adapted from [<a href="#B86-ijms-25-13104" class="html-bibr">86</a>], Copyright 2022, with permission from Elsevier. (<b>d</b>) Most stable heterostructure consisting of M<sub>2</sub>CO<sub>2</sub> (M = Ti, Zr, and Hf) and MoS<sub>2</sub> and the corresponding computed band structures. Adapted from [<a href="#B88-ijms-25-13104" class="html-bibr">88</a>], Copyright 2020, with permission from Elsevier. (<b>e</b>) Optimized structure of Zr<sub>2</sub>CO<sub>2</sub>/MSe<sub>2</sub> (M = Mo and W) and the computed projected band structures of both materials. Adapted from [<a href="#B89-ijms-25-13104" class="html-bibr">89</a>], Copyright 2022, with permission from Elsevier.</p>
Full article ">Figure 7
<p>(<b>a</b>) Optimized MoS<sub>2</sub>/β−Ga<sub>2</sub>O<sub>3</sub> heterostructures and their charge density with and without N, and (<b>b</b>) the calculated band structures. Reprinted with permission from [<a href="#B91-ijms-25-13104" class="html-bibr">91</a>], Copyright 2019, Springer Nature. (<b>c</b>) Relaxed atomic structures and obtained electronic structures of WSe<sub>2</sub>/GaN with and without four different types of vacancies. Reprinted from [<a href="#B92-ijms-25-13104" class="html-bibr">92</a>], Copyright 2024, with permission from Elsevier. (<b>d</b>) Most stable structures of two different stacking configurations consisting of WS<sub>2</sub> and C<sub>2</sub>N, and their band structures with and without 1% strain. (<b>e</b>) Graph representing the transition of bandgap and heterojunction types under different strain strengths. Adapted from [<a href="#B93-ijms-25-13104" class="html-bibr">93</a>], Copyright 2022, with permission from Elsevier.</p>
Full article ">Figure 8
<p>(<b>a</b>) Calculated electronic structures of the optimized CN and GQDs/CN structures. Adapted from [<a href="#B103-ijms-25-13104" class="html-bibr">103</a>], Copyright 2022, with permission from Elsevier. (<b>b</b>) Relaxed atomic structures and their electronic band structures of g−C<sub>3</sub>N<sub>4</sub>/BlueP heterostructures, with and without a graphene monolayer between them. The fatband colors for g−C<sub>3</sub>N<sub>4</sub>, graphene, and BlueP are yellow, red, and blue, respectively. Adapted from [<a href="#B104-ijms-25-13104" class="html-bibr">104</a>], Copyright 2024, with permission from Elsevier. (<b>c</b>) Most stable atomic structures of two different configurations consisting of g-C<sub>3</sub>N<sub>4</sub> and B<sub>4</sub>C<sub>3</sub>, and the band structures of the two designed heterostructures. Reprinted from [<a href="#B105-ijms-25-13104" class="html-bibr">105</a>], Copyright 2020, with permission from Elsevier.</p>
Full article ">Figure 9
<p>(<b>a</b>) Optimized atomic structures of monolayers of C<sub>2</sub>N, CdS, and CdSe, along with the heterostructures of CdS/C<sub>2</sub>N and CdSe/C<sub>2</sub>N; (<b>b</b>) computed electronic structures and charge density of the two heterostructures. Adapted with permission from [<a href="#B107-ijms-25-13104" class="html-bibr">107</a>]. Copyright 2020 American Chemical Society. (<b>c</b>) Relaxed structure of g−GaN/C<sub>2</sub>N with and without rotation and the obtained projected band structure of the most stable heterostructure configuration. The fatband colors for C<sub>2</sub>N and g−GaN are red and blue, respectively. Reprinted from [<a href="#B108-ijms-25-13104" class="html-bibr">108</a>], Copyright 2023, with permission from Elsevier.</p>
Full article ">
16 pages, 4124 KiB  
Article
Two-Dimensional Pentagonal Materials with Parabolic Dispersion and High Carrier Mobility
by Xiaofei Shao, Xiaobiao Liu and Xikui Ma
Materials 2024, 17(22), 5543; https://doi.org/10.3390/ma17225543 - 13 Nov 2024
Viewed by 752
Abstract
Materials with high carrier mobility, represented by graphene, have garnered significant interest. However, the zero band gap arising from linear dispersion cannot achieve an ideal on–off ratio in field-effect transistors (FETs), limiting practical applications in certain fields. In contrast, parabolic dispersion usually exhibits [...] Read more.
Materials with high carrier mobility, represented by graphene, have garnered significant interest. However, the zero band gap arising from linear dispersion cannot achieve an ideal on–off ratio in field-effect transistors (FETs), limiting practical applications in certain fields. In contrast, parabolic dispersion usually exhibits extremely high carrier mobility and an appropriate band gap. In this work, we predicted a planar pentagonal lattice composed entirely of pentagons (namely penta-MX2 monolayer), where M = Ni, Pd and Pt, X = group V elements. Using first-principles calculations, we demonstrated a parabolic dispersion within this framework, which results in intriguing phenomena, such as a direct band gap (0.551–1.105 eV) and extraordinary high carrier mobility. For penta-MX2 monolayer, the carrier mobility can attain ~1 × 108 cm2 V−1 s−1 (PBE), surpassing those of black phosphorene, graphene and 2D hexagonal materials. This monolayer also displays anisotropic mechanical properties and significant absorption peaks in the ultraviolet spectrum. Remarkably, 2D penta-MX2 monolayers are promising for successful experimental exfoliation, particularly when X is a nitrogen element, opening up new possibilities for designing two-dimensional semiconductor materials characterized by high carrier mobility. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Primitive cell of (<b>a</b>) monoclinic PdP<sub>2</sub> crystal, (<b>b</b>) orthorhombic PtN<sub>2</sub> crystal, and (<b>c</b>) cubic PtP<sub>2</sub> crystal. The (100) plane, (501) plane, and (001) plane are labelled in yellow and exhibit pentagonal lattice features, respectively. (<b>d</b>) Top view and side view of <span class="html-italic">penta</span>-MX<sub>2</sub> monolayer.</p>
Full article ">Figure 2
<p>Phonon spectrum of (<b>a</b>) <span class="html-italic">penta</span>-NiN<sub>2</sub>, (<b>b</b>) <span class="html-italic">penta</span>-PdP<sub>2</sub>, (<b>c</b>) <span class="html-italic">penta</span>-PtN<sub>2</sub>, and (<b>d</b>) <span class="html-italic">penta</span>-PtP<sub>2</sub> monolayer. Ab initio molecular dynamics simulation (AIMDS) of (<b>e</b>) <span class="html-italic">penta</span>-NiN<sub>2</sub> monolayer and <span class="html-italic">penta</span>-PtN<sub>2</sub> monolayer, (<b>f</b>) <span class="html-italic">penta</span>-PdP<sub>2</sub> monolayer and <span class="html-italic">penta</span>-PtP<sub>2</sub> monolayer.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>d</b>) Electronic band structures of <span class="html-italic">penta</span>-MX<sub>2</sub> monolayer, calculated by using the HSE06 functionals. (<b>e</b>) Band gap of <span class="html-italic">penta</span>-MX<sub>2</sub> monolayer, bilayer and three layers by using the HSE06 functionals. The number of layers is represented by n. (<b>f</b>) Three-dimensional TB band structures of the sixth band (highest VB) and the seventh band (lowest CB) of <span class="html-italic">penta</span>-PtP<sub>2</sub> with TB parameters of <span class="html-italic">t</span><sub>0</sub> = 2.138, <span class="html-italic">γ</span> = 1.717, <span class="html-italic">γ</span>′ = 0.57, and <span class="html-italic">ε</span> = 0.20 by fitting the HSE06 band structures in (<b>d</b>). The 3D energy-momentum dispersion is set in close proximity of the Fermi level calculated in the reciprocal space.</p>
Full article ">Figure 4
<p>(<b>a</b>) Predicted carrier mobilities of typical 2D materials (6, 6, 12-graphyne [<a href="#B53-materials-17-05543" class="html-bibr">53</a>], silicene [<a href="#B52-materials-17-05543" class="html-bibr">52</a>], graphdiyne sheet [<a href="#B18-materials-17-05543" class="html-bibr">18</a>], BC<sub>2</sub>N [<a href="#B54-materials-17-05543" class="html-bibr">54</a>], PdSe<sub>2</sub> [<a href="#B55-materials-17-05543" class="html-bibr">55</a>], SiC<sub>6</sub> [<a href="#B56-materials-17-05543" class="html-bibr">56</a>], GeP<sub>3</sub> (2L) [<a href="#B57-materials-17-05543" class="html-bibr">57</a>], InSe [<a href="#B58-materials-17-05543" class="html-bibr">58</a>], NiP<sub>2</sub> [<a href="#B42-materials-17-05543" class="html-bibr">42</a>], graphene [<a href="#B18-materials-17-05543" class="html-bibr">18</a>,<a href="#B52-materials-17-05543" class="html-bibr">52</a>], <span class="html-italic">α</span>-CP [<a href="#B59-materials-17-05543" class="html-bibr">59</a>], NbS<sub>2</sub> [<a href="#B60-materials-17-05543" class="html-bibr">60</a>], Tl<sub>2</sub>O [<a href="#B51-materials-17-05543" class="html-bibr">51</a>], MoS<sub>2</sub> sheet [<a href="#B61-materials-17-05543" class="html-bibr">61</a>]) by employing the acoustic phonon-limited scattering model based on the PBE band structures. For comparison, only the highest value of carrier (electron or hole) mobility of these 2D materials is listed. The unit of mobilities <span class="html-italic">μ</span> is 10<sup>3</sup> cm<sup>2</sup> V<sup>−</sup><sup>1</sup> s<sup>−</sup><sup>1</sup>. (<b>b</b>) Young’s modulus and (<b>c</b>) Poisson’s ratio of <span class="html-italic">penta</span>-MX<sub>2</sub> in polar coordinates. The unit of Young’s modulus is GPa·nm.</p>
Full article ">Figure 5
<p>Optical absorption of <span class="html-italic">penta</span>-MX<sub>2</sub> (<b>a</b>–<b>d</b>) monolayer and (<b>e</b>–<b>h</b>) bilayer by employing the PBE and HSE06 functionals. The color region represents the visible light region from 1.77 eV (700 nm) to 3.10 eV (400 nm). Scissor operator is applied to align the band gap obtained from the PBE calculations with that from the HSE06 calculations. The multi-peaks observed in the HSE06 data are attributed to the inadequate <span class="html-italic">k</span>-point sampling density.</p>
Full article ">
32 pages, 2634 KiB  
Review
Advances in Graphene Field Effect Transistors (FETs) for Amine Neurotransmitter Sensing
by Elmira Alimohammadzadeh and John Hedley
Appl. Sci. 2024, 14(22), 10109; https://doi.org/10.3390/app142210109 - 5 Nov 2024
Viewed by 1202
Abstract
Amine neurotransmitters (NTs) are crucial in the central nervous system, and dysregulation in their levels is implicated in a spectrum of neurological disorders. Thus, a precise and timely assessment of their concentrations is critical for early diagnosis and treatment efficacy monitoring. Graphene-based field [...] Read more.
Amine neurotransmitters (NTs) are crucial in the central nervous system, and dysregulation in their levels is implicated in a spectrum of neurological disorders. Thus, a precise and timely assessment of their concentrations is critical for early diagnosis and treatment efficacy monitoring. Graphene-based field effect transistors (GFETs) have become a ground-breaking instrument in the detection of these NTs because of their exceptional electrical characteristics and adaptability. This paper summarises the significant advancements in GFET biosensors in amine NT detection and highlights developments in the selectivity, sensitivity, and limit of detection (LOD) attained by selecting various graphene materials and functionalisation approaches. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The schematic GFET biosensor design: showing a graphene sheet functionalised with biorecognition elements (purple) and immersed in a medium containing the target analyte (red) and (<b>b</b>) back-gated, top-gated, coplanar-gated, and liquid-gated GFET configurations.</p>
Full article ">Figure 2
<p>Graphene characterisation: (<b>a</b>) Raman spectroscopy of rGO showing the characteristic D, G, and 2D peaks associated with graphene and its derivatives and (<b>b</b>) XPS spectra of rGO with the fitted peak representing different oxygen functional groups that can be used as active sites for surface functionalisation (Reprinted with permission from [<a href="#B141-applsci-14-10109" class="html-bibr">141</a>]. 2023, MDPI).</p>
Full article ">Figure 3
<p>(<b>a</b>) A schematic of the four-point probe setup for measuring the sheet resistance of the graphene layer. (<b>b</b>) The transfer curve of a GFET shows the difference in the Dirac points in forward and reverse sweeps (as indicated by arrows).</p>
Full article ">Figure 4
<p>A typical calibration curve: Region 1 shows concentrations below the minimum detectable limit (LOD) of the analyte. Region 2 highlights the usable range of the sensor. Region 3 demonstrates that when concentrations are too high, no more binding sites are available, and the sensor is termed saturated.</p>
Full article ">Figure 5
<p>(<b>a</b>) I<sub>D</sub>–V<sub>DS</sub> characteristic of TGN-based electrochemical sensor for different pH values (5–9). (<b>b</b>) Conductance–voltage characteristic of TGN-based electrochemical sensor for different pH values (5–10). (Reprinted with permission from [<a href="#B257-applsci-14-10109" class="html-bibr">257</a>]. 2019, IET).</p>
Full article ">
11 pages, 480 KiB  
Article
High-Data-Rate Modulators Based on Graphene Transistors: Device Circuit Co-Design Proposals
by Anibal Pacheco-Sanchez, J. Noé Ramos-Silva, Nikolaos Mavredakis, Eloy Ramírez-García and David Jiménez
Electronics 2024, 13(20), 4022; https://doi.org/10.3390/electronics13204022 - 12 Oct 2024
Viewed by 1042
Abstract
The multifunctionality feature of graphene field-effect transistors (GFETs) is exploited here to design circuit building blocks of high-data-rate modulators by using a physics-based compact model. Educated device performance projections are obtained with the experimentally calibrated model and used to choose an appropriate improved [...] Read more.
The multifunctionality feature of graphene field-effect transistors (GFETs) is exploited here to design circuit building blocks of high-data-rate modulators by using a physics-based compact model. Educated device performance projections are obtained with the experimentally calibrated model and used to choose an appropriate improved feasible GFET for these applications. Phase-shift and frequency-shift keying (PSK and FSK) modulation schemes are obtained with 0.6 GHz GFET-based multifunctional circuits used alternatively in different operation modes: inverting and in-phase amplification and frequency multiplication. An adequate baseband signal applied to the transistors’ input also serves to enhance the device and circuit performance reproducibility since the impact of traps is diminished. Quadrature PSK is also achieved by combining two GFET-based multifunctional circuits. This device circuit co-design proposal intends to boost the heterogeneous implementation of graphene devices with incumbent technologies into a single chip: the baseband pulses can be generated with CMOS technology as a front end of line and the multifunctional GFET-based circuits as a back end of line. Full article
Show Figures

Figure 1

Figure 1
<p><b>Top left</b>: ambipolar transfer characteristic of a GFET showing approximate definitions for the drain current in the different operation regimes. <b>Bottom left</b>: input AC signals. <b>Top right</b>: output AC signal. <b>Bottom right</b>: equations of the analog AC+DC signals for the general case (in black) and specific cases. <span class="html-italic">A</span>, <span class="html-italic">B</span>, <span class="html-italic">C</span>, <span class="html-italic">D</span> and <span class="html-italic">E</span> are arbitrary constants and <math display="inline"><semantics> <mrow> <mo>Γ</mo> <mo>=</mo> <mn>0.5</mn> <msub> <mi>R</mi> <mi mathvariant="normal">d</mi> </msub> <mi>C</mi> <msup> <mi>A</mi> <mn>2</mn> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mo>Φ</mo> <mo>=</mo> <msub> <mi>V</mi> <mi>DS</mi> </msub> <mo>−</mo> <mi>B</mi> <msub> <mi>R</mi> <mi mathvariant="normal">d</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <msub> <mi>R</mi> <mi mathvariant="normal">d</mi> </msub> </semantics></math> is the output resistance seen from the drain terminal. <math display="inline"><semantics> <msub> <mi>I</mi> <mi>d</mi> </msub> </semantics></math> is obtained by replacing <math display="inline"><semantics> <msub> <mi>V</mi> <mi>GS</mi> </msub> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mi>GS</mi> </msub> <mo>+</mo> <msub> <mi>v</mi> <mi>in</mi> </msub> </mrow> </semantics></math> in the corresponding <math display="inline"><semantics> <msub> <mi>I</mi> <mi mathvariant="normal">D</mi> </msub> </semantics></math>.</p>
Full article ">Figure 2
<p>Transfer characteristics of a 300 <math display="inline"><semantics> <mi mathvariant="normal">n</mi> </semantics></math><math display="inline"><semantics> <mi mathvariant="normal">m</mi> </semantics></math> long GFET. <b>Left</b>: Trap-reduced data obtained with opposing pulses. Markers are experimental data and lines are modeling results. Inset shows the applied opposing <math display="inline"><semantics> <msub> <mi>V</mi> <mi>GS</mi> </msub> </semantics></math> pulses and constant <math display="inline"><semantics> <msub> <mi>V</mi> <mi>DS</mi> </msub> </semantics></math>. <b>Right panel</b>: optimized device modeling results. <math display="inline"><semantics> <msub> <mi>V</mi> <mi>DS</mi> </msub> </semantics></math> is 0.1 <span class="html-italic">V</span>, 0.2 <span class="html-italic">V</span> and 0.3 <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math> for all cases.</p>
Full article ">Figure 3
<p><b>Top</b>: Schematic of the multifunctional GFET circuit used for data modulation at <math display="inline"><semantics> <mrow> <mn>0.6</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi>GHz</mi> </semantics></math>. In-phase and inverting amplification obtained with <math display="inline"><semantics> <msub> <mi>V</mi> <mi>GS</mi> </msub> </semantics></math> equal to −0.1 V and 0.5 <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math>, respectively, whereas the circuit works as a frequency doubler at <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mi>GS</mi> </msub> <mo>=</mo> <msub> <mi>V</mi> <mi>Dirac</mi> </msub> </mrow> </semantics></math>. Matching (stability) networks are indicated by the dashed (dotted) boxes and are the same regardless of the operation mode. DC and AC filtering between signal sources and circuit elements are not shown. Input AC power is of <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>30</mn> </mrow> </semantics></math> dBm at <math display="inline"><semantics> <mrow> <mn>0.6</mn> <mo> </mo> <mrow> <mi>GHz</mi> </mrow> </mrow> </semantics></math>. Values of circuit elements are <math display="inline"><semantics> <mrow> <msub> <mi>L</mi> <mn>1</mn> </msub> <mspace width="3.33333pt"/> <mo>=</mo> <mspace width="3.33333pt"/> <mn>155</mn> <mo> </mo> <mrow> <mi mathvariant="normal">n</mi> <mi mathvariant="normal">H</mi> </mrow> <mo>,</mo> <msub> <mi>C</mi> <mn>1</mn> </msub> <mspace width="3.33333pt"/> <mo>=</mo> <mspace width="3.33333pt"/> <mn>375</mn> <mo> </mo> <mrow> <mi mathvariant="normal">f</mi> <mi mathvariant="normal">F</mi> </mrow> <mo>,</mo> <msub> <mi>L</mi> <mn>2</mn> </msub> <mspace width="3.33333pt"/> <mo>=</mo> <mspace width="3.33333pt"/> <mn>41</mn> <mo> </mo> <mrow> <mi mathvariant="normal">n</mi> <mi mathvariant="normal">H</mi> </mrow> <mo>,</mo> <msub> <mi>C</mi> <mn>2</mn> </msub> <mspace width="3.33333pt"/> <mo>=</mo> <mspace width="3.33333pt"/> <mn>1.6</mn> <mo> </mo> <mrow> <mi mathvariant="normal">p</mi> <mi mathvariant="normal">F</mi> </mrow> <mo>,</mo> <msub> <mi>R</mi> <mn>1</mn> </msub> <mspace width="3.33333pt"/> <mo>=</mo> <mspace width="3.33333pt"/> <mn>10.9</mn> <mo> </mo> <mrow> <mi mathvariant="normal">k</mi> <mo>Ω</mo> </mrow> </mrow> </semantics></math>. <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mi>DD</mi> </msub> <mo>=</mo> <mn>0.3</mn> <mo> </mo> <mi mathvariant="normal">V</mi> </mrow> </semantics></math>. <b>Bottom</b>: S-parameters of the amplifiers for the PCA design: continuous lines represent results of the in-phase amplifier (<math display="inline"><semantics> <mrow> <mo>@</mo> <msub> <mi>V</mi> <mrow> <mi>GS</mi> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>) and dashed–dotted lines show results of the inverting amplifier (<math display="inline"><semantics> <mrow> <mo>@</mo> <msub> <mi>V</mi> <mrow> <mi>GS</mi> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math>).</p>
Full article ">Figure 4
<p><math display="inline"><semantics> <msub> <mi>v</mi> <mi>in</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>v</mi> <mi>out</mi> </msub> </semantics></math> signals for the PCA design in both operation modes: In-phase amplifier (<b>left</b>) and inverting amplifier (<b>right</b>). Only a 10 <math display="inline"><semantics> <mi mathvariant="normal">n</mi> </semantics></math><math display="inline"><semantics> <mi mathvariant="normal">s</mi> </semantics></math> fram is shown for a better visualization of the signals.</p>
Full article ">Figure 5
<p>Frequency doubler results: <math display="inline"><semantics> <msub> <mi>v</mi> <mi>in</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>v</mi> <mi>out</mi> </msub> </semantics></math> signals over a 10 <math display="inline"><semantics> <mi mathvariant="normal">n</mi> </semantics></math><math display="inline"><semantics> <mi mathvariant="normal">s</mi> </semantics></math> frame (<b>left</b>) and output power spectrum over frequency (<b>right</b>).</p>
Full article ">Figure 6
<p>Baseband, carrier and modulated signals achieved with the GFET-based multifunctional circuits. <b>Left</b>: PSK signals. Phase of output signal is included in the bottom plot. <b>Right</b>: FSK signals.</p>
Full article ">Figure 7
<p>Schematic representation of <math display="inline"><semantics> <mrow> <mn>0.6</mn> <mo> </mo> </mrow> </semantics></math><math display="inline"><semantics> <mi mathvariant="normal">M</mi> </semantics></math><math display="inline"><semantics> <mi>Hz</mi> </semantics></math>-QPSK modulator obtained with two GFET-based PSK circuits. Each PSK block corresponds to the circuit shown in <a href="#electronics-13-04022-f003" class="html-fig">Figure 3</a>. <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>GS</mi> <mn>1</mn> </mrow> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>GS</mi> <mn>2</mn> </mrow> </msub> </semantics></math> correspond to the baseband signal of the PCA design (with values of −<math display="inline"><semantics> <mrow> <mn>0.1</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>0.5</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math>).</p>
Full article ">Figure 8
<p>Signals involved in the GFET-based quadrature PSK modulation.</p>
Full article ">
12 pages, 27277 KiB  
Article
Process Development of a Liquid-Gated Graphene Field-Effect Transistor Gas Sensor for Applications in Smart Agriculture
by Jian Lu, Naoki Shiraishi, Ryo Imaizumi, Lan Zhang and Mutsumi Kimura
Sensors 2024, 24(19), 6376; https://doi.org/10.3390/s24196376 - 1 Oct 2024
Viewed by 1093
Abstract
A compact, multi-channel ionic liquid-gated graphene field-effect transistor (FET) has been proposed and developed in our work for on-field continuous monitoring of nitrate nitrogen and other nitrogen fertilizers to achieve sustainable and efficient farming practices in agriculture. However, fabricating graphene FETs with easy [...] Read more.
A compact, multi-channel ionic liquid-gated graphene field-effect transistor (FET) has been proposed and developed in our work for on-field continuous monitoring of nitrate nitrogen and other nitrogen fertilizers to achieve sustainable and efficient farming practices in agriculture. However, fabricating graphene FETs with easy filling of ionic liquids, minimal graphene defects, and high process yields remains challenging, given the sensitivity of these devices to processing conditions and environmental factors. In this work, two approaches for the fabrication of our graphene FETs were presented, evaluated, and compared for high yields and easy filling of ionic liquids. The process difficulties, major obstacles, and improvements are discussed herein in detail. Both devices, those fabricated using a 3 μm-thick CYTOP® layer for position restriction and volume control of the ionic liquid and those using a ~20 nm-thick photosensitive hydrophobic layer for the same purpose, exhibited typical FET characteristics and were applicable to various application environments. The research findings and experiences presented in this paper will provide important references to related societies for the design, fabrication, and application of liquid-gated graphene FETs. Full article
(This article belongs to the Special Issue Gas Sensors: Progress, Perspectives and Challenges)
Show Figures

Figure 1

Figure 1
<p>Schematic view of the structure and the layout of each ionic liquid-gated graphene FET (left figure), and the cross-section views of the graphene channel in X direction (upper-right figure) and in Y direction (lower-right figure).</p>
Full article ">Figure 2
<p>Fabrication process flow using CYTOP<sup>®</sup>: (<b>a</b>) graphene etching by O<sub>2</sub> plasma; (<b>b</b>) Au/Cr electrode deposition by sputter and patterning by lift-off; (<b>c</b>) CYTOP<sup>®</sup> coating and (<b>d</b>) etching; (<b>e</b>) channel release by Au/Cr wet etching. (<b>f</b>,<b>g</b>) are photos of the wafer after graphene etching in (<b>a</b>) and after channel release in (<b>e</b>), correspondingly.</p>
Full article ">Figure 3
<p>Fabrication process flow using LDW-N010: (<b>a</b>) graphene etching by O<sub>2</sub> plasma; (<b>b</b>) Au/Cr electrode deposition by sputter and patterning by lift-off; (<b>c</b>) LDW-N010 coating, photolithography, and development.</p>
Full article ">Figure 4
<p>(<b>a</b>) SEM image of a fabricated chip using CYTOP<sup>®</sup>; (<b>b</b>) SEM image of an individual FET using CYTOP<sup>®</sup> (length: 20 μm; width: 250 μm); (<b>c</b>) optical image of the graphene channel (length: 20 μm; width: 500 μm) with chemical residuals and Cr undercut by wet etching; (<b>d</b>) optical image of the graphene channel (length: 20 μm; width: 50 μm) with Cr undercut by wet etching; (<b>e</b>) optical image of the graphene channel (length: 20 μm; width: 50 μm) without chemical residuals, the Cr undercut is neglectable.</p>
Full article ">Figure 5
<p>(<b>a</b>) SEM image of a fabricated individual FET (length: 20 μm; width: 500 μm) using LDW-N010; (<b>b</b>) optical image of the chip with well-defined hydrophobic pattern by LDW-N010; (<b>c</b>) optical image of the graphene channel (length: 100 μm; width: 100 μm) after using N-methyl pyrrolidone (NMP) to remove the photoresist during Au/Cr lift-off; (<b>d</b>) optical image of the graphene channel (length: 20 μm; width: 200 μm) after using acetone to remove the photoresist during Au/Cr lift-off; (<b>e</b>) SEM image of a fabricated individual FET (channel length: 100 μm; width: 50 μm) using negative photoresist ZPN1150-90.</p>
Full article ">Figure 6
<p>Measured contact angle on (<b>a</b>) LDW-N010 surface; (<b>b</b>) Au electrode surface without LDW-N010; (<b>c</b>) interface surface between Au and LDW-N010.</p>
Full article ">Figure 7
<p>(<b>a</b>) optical image of the chip using CYTOP<sup>®</sup> after filling with H<sub>2</sub>O at various volumes; (<b>b</b>) optical image of the chip using LDW-N010 after filling with H<sub>2</sub>O at various volumes; (<b>c</b>) optical image of the chip using ZPN1150-90 after filling with H<sub>2</sub>O at various volume; (<b>d</b>) photo of the chip using CYTOP<sup>®</sup> after filling with H<sub>2</sub>O at various volume; (<b>e</b>) 3D microscopic image of the chip using CYTOP<sup>®</sup> and after filling with H<sub>2</sub>O at 600 nL.</p>
Full article ">Figure 8
<p>(<b>a</b>) Photo of our evaluation module after wire bonding of the ionic liquid-gated graphene FET device onto a ceramic package; (<b>b</b>) photo of the chip using CYTOP<sup>®</sup> after filling ionic liquid [PMIM][BF4] with various volumes; (<b>c</b>) current-voltage characteristics of the source-drain electrodes (I<sub>ds</sub>-V<sub>ds</sub>) of the graphene FET before and after filling of the ionic liquid [PMIM][BF4]. No gate voltage was applied during the Ids-Vds test.</p>
Full article ">Figure 9
<p>Measure drain-source current (I<sub>ds</sub>) of the graphene FET at different gate voltages (V<sub>g</sub>) when the FET was exposed to air or sealed inside our self-developed soil simulation chamber. The volume of the filled [PMIM][BF4] was (<b>a</b>) 100 nL and (<b>b</b>) 200 nL, respectively. The measurement was done at 25 °C with a relative humidity of 51.2%.</p>
Full article ">
15 pages, 2388 KiB  
Article
Digitalization of Enzyme-Linked Immunosorbent Assay with Graphene Field-Effect Transistors (G-ELISA) for Portable Ferritin Determination
by Melody L. Candia, Esteban Piccinini, Omar Azzaroni and Waldemar A. Marmisollé
Biosensors 2024, 14(8), 394; https://doi.org/10.3390/bios14080394 - 16 Aug 2024
Viewed by 1490
Abstract
Herein, we present a novel approach to quantify ferritin based on the integration of an Enzyme-Linked Immunosorbent Assay (ELISA) protocol on a Graphene Field-Effect Transistor (gFET) for bioelectronic immunosensing. The G-ELISA strategy takes advantage of the gFET inherent capability of detecting pH changes [...] Read more.
Herein, we present a novel approach to quantify ferritin based on the integration of an Enzyme-Linked Immunosorbent Assay (ELISA) protocol on a Graphene Field-Effect Transistor (gFET) for bioelectronic immunosensing. The G-ELISA strategy takes advantage of the gFET inherent capability of detecting pH changes for the amplification of ferritin detection using urease as a reporter enzyme, which catalyzes the hydrolysis of urea generating a local pH increment. A portable field-effect transistor reader and electrolyte-gated gFET arrangement are employed, enabling their operation in aqueous conditions at low potentials, which is crucial for effective biological sample detection. The graphene surface is functionalized with monoclonal anti-ferritin antibodies, along with an antifouling agent, to enhance the assay specificity and sensitivity. Markedly, G-ELISA exhibits outstanding sensing performance, reaching a lower limit of detection (LOD) and higher sensitivity in ferritin quantification than unamplified gFETs. Additionally, they offer rapid detection, capable of measuring ferritin concentrations in approximately 50 min. Because of the capacity of transistor miniaturization, our innovative G-ELISA approach holds promise for the portable bioelectronic detection of multiple biomarkers using a small amount of the sample, which would be a great advancement in point–of–care testing. Full article
(This article belongs to the Special Issue Current Advance in Transistor-Based Biosensors for Diagnostics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Photograph of the portable measurement station and schematic of a gFET sensor. (<b>B</b>) Characteristic transfer curves for a streptavidin-modified gFET by varying the pH from 3 to 11 obtained at a V<sub>DS</sub> = 0.05 V in a solution of 140 mM NaCl and 1 mM KH<sub>2</sub>PO<sub>4</sub>. (<b>C</b>) Change in Dirac potential as a function of pH for an unmodified gFET (red line) and modified gFET (black line). Slopes’ values from the linear fitting are reported for each gFET.</p>
Full article ">Figure 2
<p>(<b>A</b>) Schematic of the Strept-gFET fabrication process. (<b>B</b>) Changes in I<sub>DS</sub> after the addition of 1 mM urea for a Strept-gFET incubated with different concentrations of b-urease (V<sub>DS</sub> = 50 mV, V<sub>GS</sub> = −250 mV, 0.1 mM HEPES buffer with 10 mM KCl, pH 6). (<b>C</b>) Correlation between the slopes of the I<sub>DS</sub> vs. time curves and the concentration of b-urease used in the modification of Strept-gFET. Bars correspond to the deviation of the slopes of two Strept-gFETs. The red line is the Hill-like model fitting. (<b>D</b>) G-ELISA scheme with capillary pH meter and comparison of changes in I<sub>DS</sub> for a G-ELISA after incubation of 1 nM ferritin (V<sub>DS</sub> = 50 mV, V<sub>GS</sub> = −250 mV, 0.1 mM HEPES buffer with 10 mM KCl, pH 6) and measured pH changes with a capillary pH meter after the addition of 1 mM urea.</p>
Full article ">Figure 3
<p>G-ELISA scheme for the detection of Ferritin.</p>
Full article ">Figure 4
<p>(<b>A</b>) Changes in I<sub>DS</sub> for a G-ELISA after incubation with different concentrations of ferritin and the addition of 1 mM urea (V<sub>DS</sub> = 50 mV, V<sub>GS</sub> = −250 mV, 0.1 mM HEPES buffer with 10 mM KCl, pH 6). (<b>B</b>) I<sub>DS</sub>% as a function of the ferritin concentration calculated at different times, from 550 to 900 s. (<b>C</b>) Correlation between the I<sub>DS</sub> slopes and the ferritin concentration obtained from the G-ELISA measurements and the detection with 1 mM urea. (<b>D</b>) Comparison of conventional I<sub>DS</sub>% gFET (V<sub>DS</sub> = 50 mV, V<sub>GS</sub> = −250 mV, HEPES buffer × 0.1 at pH 7.4) and I<sub>DS</sub>% at 900 s of G-ELISA as a function of ferritin concentration. Data for conventional gFET were reproduced from Piccinini et al., 2022 [<a href="#B52-biosensors-14-00394" class="html-bibr">52</a>].</p>
Full article ">
14 pages, 3929 KiB  
Article
Deep Learning Approach for Modeling the Power Consumption and Delay of Logic Circuits Employing GNRFET Technology
by Recep Emir, Dilek Surekci Yamacli, Serhan Yamacli and Sezai Alper Tekin
Electronics 2024, 13(15), 2993; https://doi.org/10.3390/electronics13152993 - 29 Jul 2024
Viewed by 1015
Abstract
The interest in alternative logic technologies is continuously increasing for short nanometer designs. From this viewpoint, logic gates, full adder and D-latch designs based on graphene nanoribbon field effect transistors (GNRFETs) at 7 nm technology nodes were presented, considering that these structures are [...] Read more.
The interest in alternative logic technologies is continuously increasing for short nanometer designs. From this viewpoint, logic gates, full adder and D-latch designs based on graphene nanoribbon field effect transistors (GNRFETs) at 7 nm technology nodes were presented, considering that these structures are core elements for digital integrated circuits. Firstly, NOT, NOR and NAND gates were implemented using GNRFETs. Then, 28T full adder and 18T D-latch circuits based on CMOS logic were designed using GNRFETs. As the first result of this work, it was shown through HSPICE simulations that the average power consumption of the considered logic circuits employing GNRFETs was 78.6% lower than those built using classical Si-based MOSFETs. Similarly, the delay advantage of the logic circuits employing GNRFETs was calculated to be 53.2% lower than those using Si-based MOSFET counterparts. In addition, a deep learning model was developed to model both the power consumption and the propagation delay of GNRFET-based logic inverters. As the second result, it was demonstrated that the developed deep learning model could accurately represent the power consumption and delay of GNRFET-based logic circuits with the coefficient of determination (R2) values in the range of 0.86 and 0.99. Full article
Show Figures

Figure 1

Figure 1
<p>The typical GNRFET structure.</p>
Full article ">Figure 2
<p>The structure of the armchair GNR employed in the GNRFET as the channel region.</p>
Full article ">Figure 3
<p>NOT, NOR and NAND gates based on CMOS logic.</p>
Full article ">Figure 4
<p>Full adder circuit block.</p>
Full article ">Figure 5
<p>Full adder circuit based on CMOS logic.</p>
Full article ">Figure 6
<p>D-latch closed circuit.</p>
Full article ">Figure 7
<p>Conventional D-latch.</p>
Full article ">Figure 8
<p>Input-output signals of NOT gate using the GNRFET. (<b>a</b>) Voltage domain. (<b>b</b>) Time domain.</p>
Full article ">Figure 9
<p>Input–output signals of NOR and NAND gates using the GNRFET. (<b>a</b>) NOR, (<b>b</b>) NAND.</p>
Full article ">Figure 10
<p>Input–output signals of FA and D-latch circuits using GNRFET. (<b>a</b>) FA, (<b>b</b>) D-latch.</p>
Full article ">Figure 11
<p>The structure of the developed deep learning model.</p>
Full article ">Figure 12
<p>Actual power dissipation data and the deep learning model result for the GNRFET inverter.</p>
Full article ">Figure 13
<p>Actual delay data and the deep learning model result for the GNRFET inverter.</p>
Full article ">
13 pages, 1122 KiB  
Article
Rapid, Selective, and Ultra-Sensitive Field Effect Transistor-Based Detection of Escherichia coli
by Liena Zaidan, Inna Novodchuk, Alexander H.Xu, Alexandru Nica, Saeed Takaloo, Christopher Lloyd, Reza Karimi, Joe Sanderson, Michal Bajcsy and Mustafa Yavuz
Materials 2024, 17(15), 3648; https://doi.org/10.3390/ma17153648 - 24 Jul 2024
Cited by 1 | Viewed by 2886
Abstract
Escherichia coli (E. coli) was among the first organisms to have its complete genome published (Genome Sequence of E. coli 1997 Science). It is used as a model system in microbiology research. E. coli can cause life-threatening illnesses, particularly in children [...] Read more.
Escherichia coli (E. coli) was among the first organisms to have its complete genome published (Genome Sequence of E. coli 1997 Science). It is used as a model system in microbiology research. E. coli can cause life-threatening illnesses, particularly in children and the elderly. Possible contamination by the bacteria also results in product recalls, which, alongside the potential danger posed to individuals, can have significant financial consequences. We report the detection of live Escherichia coli (E. coli) in liquid samples using a biosensor based on a field-effect transistor (FET) biosensor with B/N co-coped reduced graphene oxide (rGO) gel (BN-rGO) as the transducer material. The FET was functionalized with antibodies to detect E. coli K12 O-antigens in phosphate-buffered saline (PBS). The biosensor detected the presence of planktonic E. coli bacterial cells within a mere 2 min. The biosensor exhibited a limit of detection (LOD) of 10 cells per sample, which can be extrapolated to a limit of detection at the level of a single cell per sample and a detection range of at least 10–108 CFU/mL. The selectivity of the biosensor for E. coli was demonstrated using Bacillus thuringiensis (B. thuringiensis) as a sample contaminant. We also present a comparison of our functionalized BN-rGO FET biosensor with established detection methods of E. coli k12 bacteria, as well as with state-of-the-art detection mechanisms. Full article
Show Figures

Figure 1

Figure 1
<p>The schematics of the device configuration and biosensing mechanism of the BN-rGO gel <span class="html-italic">E. coli</span> biosensor. (<b>a</b>) The back-gated FET device configuration features the BN-rGO gel channel between the source and drain electrodes. (<b>b</b>) The circuit connections of the BN-rGO gel FET device. (<b>c</b>) The device configuration following <span class="html-italic">E. coli</span> antibody functionalization and passivation with ethanolamine. (<b>d</b>) The biosensing response is determined by the shift in the Dirac voltage after introducing the <span class="html-italic">E. coli</span>-containing solution.</p>
Full article ">Figure 2
<p><span class="html-italic">E. coli</span> biosensing performance of antibody-functionalized BN-rGO gel FET. Shift in Dirac and increase in ON-current correspond to increasing concentrations of <span class="html-italic">E. coli</span> in solution.</p>
Full article ">Figure 3
<p>The shift in the Dirac voltage in response to different CFU/mL <span class="html-italic">E. coli</span> concentrations.</p>
Full article ">Figure 4
<p>Specificity and selectivity response of BN-rGO gel FET <span class="html-italic">E. coli</span> biosensor towards the blank sample and <span class="html-italic">B. thuringiensis</span>. (<b>a</b>) Gate voltage in the range of −0.15 to 0.025 V, (<b>b</b>) gate voltage in the range of −0.06 to −0.04 V.</p>
Full article ">
12 pages, 3906 KiB  
Article
Analysis of Local Properties and Performance of Bilayer Epitaxial Graphene Field Effect Transistors on SiC
by Dalal Fadil, Wlodek Strupinski, Emiliano Pallecchi and Henri Happy
Materials 2024, 17(14), 3553; https://doi.org/10.3390/ma17143553 - 18 Jul 2024
Cited by 1 | Viewed by 821
Abstract
Epitaxial bilayer graphene, grown by chemical vapor deposition on SiC substrates without silicon sublimation, is crucial material for graphene field effect transistors (GFETs). Rigorous characterization methods, such as atomic force microscopy and Raman spectroscopy, confirm the exceptional quality of this graphene. Post-nanofabrication, extensive [...] Read more.
Epitaxial bilayer graphene, grown by chemical vapor deposition on SiC substrates without silicon sublimation, is crucial material for graphene field effect transistors (GFETs). Rigorous characterization methods, such as atomic force microscopy and Raman spectroscopy, confirm the exceptional quality of this graphene. Post-nanofabrication, extensive evaluation of DC and high-frequency properties enable the extraction of critical parameters such as the current gain (fmax) and cut-off frequency (ft) of hundred transistors. The Raman spectra analysis provides insights into material property, which correlate with Hall mobilities, carrier densities, contact resistance and sheet resistance and highlights graphene’s intrinsic properties. The GFETs’ performance displays dispersion, as confirmed through the characterization of multiple transistors. Since the Raman analysis shows relatively homogeneous surface, the variation in Hall mobility, carrier densities and contact resistance cross the wafer suggest that the dispersion of GFET transistor’s performance could be related to the process of fabrication. Such insights are especially critical in integrated circuits, where consistent transistor performance is vital due to the presence of circuit elements like inductance, capacitance and coplanar waveguides often distributed across the same wafer. Full article
(This article belongs to the Special Issue Silicon Carbide: Material Growth, Device Processing and Applications)
Show Figures

Figure 1

Figure 1
<p>AFM images of bilayer graphene surface on 6H-SiC substrate [<a href="#B40-materials-17-03553" class="html-bibr">40</a>]. (<b>a</b>) 60 × 60 µm<sup>2</sup> image and the blue and red dash square 10 × 10 µm<sup>2</sup> is represented in (<b>b</b>). (<b>c</b>) SEM image with 10 µm scale bare. (<b>d</b>) Raman spectroscopy of graphene on SiC substrate. The black trace is the spectrum of graphene and SiC, the red trace is the graphene spectrum once SiC Raman peaks are subtracted, the blue trace is the smoothed spectrum. (<b>e</b>) Raman spectra at different locations (blue, red, and black) on the SiC wafer.</p>
Full article ">Figure 1 Cont.
<p>AFM images of bilayer graphene surface on 6H-SiC substrate [<a href="#B40-materials-17-03553" class="html-bibr">40</a>]. (<b>a</b>) 60 × 60 µm<sup>2</sup> image and the blue and red dash square 10 × 10 µm<sup>2</sup> is represented in (<b>b</b>). (<b>c</b>) SEM image with 10 µm scale bare. (<b>d</b>) Raman spectroscopy of graphene on SiC substrate. The black trace is the spectrum of graphene and SiC, the red trace is the graphene spectrum once SiC Raman peaks are subtracted, the blue trace is the smoothed spectrum. (<b>e</b>) Raman spectra at different locations (blue, red, and black) on the SiC wafer.</p>
Full article ">Figure 2
<p>(<b>a</b>) A picture of the layout of the device. “+” is source-drain contact layer. (<b>b</b>) A schematic image representing the step of the realization of the T gate process with the three layers’ resists. (<b>c</b>) A schematic of the side view of the T gate transistor. (<b>d</b>) (<b>d</b>, <b>left</b>) An SEM image of the transistor in the end of the process. (<b>d</b>, <b>right</b>) An FIB image of the transistor showing the T-gate structure, <span class="html-italic">L<sub>g</sub></span> = 200 nm. (<b>e</b>) A photograph of the final devices fabricated on the 15 mm × 15 mm SiC substrate, including 458 transistors and height cells numbered from 1 to 8.</p>
Full article ">Figure 3
<p>(<b>a</b>) An optical image of the Hall and TLM pattern. (<b>b</b>) The transmission line measurement (TLM) and the linear fit to extract the correspondent contact resistance and sheet resistance. (<b>c</b>) A recapitulation of the averaged Raman spectra of graphene on SiC measured in the eight cells of the wafer. (<b>d</b>) A table summarizing the values of the Hall mobility, carrier densities and Raman peaks’ positions and the full width at the half maximum of the G and 2D peaks. Missing values are due to faulty components.</p>
Full article ">Figure 4
<p>Example of DC and RF characteristics of dual T-gate graphene transistor in cell 7 having gate channel length (<span class="html-italic">L<sub>g</sub></span>) equal to 200 nm and dual-gate width (<span class="html-italic">w</span>) 2 × 30 µm. (<b>a</b>) DC measurement <span class="html-italic">I<sub>DS</sub></span>–<span class="html-italic">V<sub>GS</sub></span> and <span class="html-italic">g<sub>m</sub></span>–<span class="html-italic">V<sub>GS</sub></span>. (<b>b</b>) The voltage transfer characteristics as a function of <span class="html-italic">V<sub>GS</sub></span> varying from −2 V to +2 V by 0.5 V steps. (<b>c</b>) RF characteristic includes the as-measured values of the current gain <span class="html-italic">H<sub>21</sub></span> and the unilateral power gain <span class="html-italic">U</span> as a function of the frequency at <span class="html-italic">V<sub>DS</sub></span> = 1.5 V and <span class="html-italic">V<sub>GS</sub></span> = 1.3 V. The cut off frequency <span class="html-italic">f<sub>t</sub></span> and the maximum oscillation frequency <span class="html-italic">f<sub>max</sub></span> have been extracted [<a href="#B40-materials-17-03553" class="html-bibr">40</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Photography of the 15 mm × 15 mm graphene wafer presented previously in <a href="#materials-17-03553-f002" class="html-fig">Figure 2</a>e. (<b>b</b>) <span class="html-italic">V<sub>GS</sub></span> is the bias gate voltage related the maximum the transconductance and where the best performance of the transistor is expected. &lt;<span class="html-italic">V<sub>GS</sub></span>&gt; is the average value of all the gate voltage measured in each cell. (<b>c</b>,<b>d</b>) Representation of the average values of the cut-off frequency &lt;<span class="html-italic">f<sub>t</sub></span>&gt; and the maximum oscillation frequency &lt;<span class="html-italic">f<sub>max</sub></span>&gt; computed for each cell. (<b>e</b>) Graph summarizing the evolution of the on-probe values of <span class="html-italic">f<sub>t</sub></span> and <span class="html-italic">f<sub>max</sub></span> as function of the biased gate voltage <span class="html-italic">V<sub>GS</sub></span> of the transistors in each cell.</p>
Full article ">
22 pages, 7975 KiB  
Article
Low-Cost Source Measure Unit (SMU) to Characterize Sensors Built on Graphene-Channel Field-Effect Transistors
by Ashley Morgan Galanti and Mark A. Haidekker
Sensors 2024, 24(12), 3841; https://doi.org/10.3390/s24123841 - 14 Jun 2024
Viewed by 1270
Abstract
This study introduces a flexible and low-cost solution for a source measure unit (SMU), which is presented as an alternative to conventional source meter units and a blueprint for sensor FET drivers. An SMU collects current–voltage (I-V) curves with an additional variable voltage [...] Read more.
This study introduces a flexible and low-cost solution for a source measure unit (SMU), which is presented as an alternative to conventional source meter units and a blueprint for sensor FET drivers. An SMU collects current–voltage (I-V) curves with an additional variable voltage or current and is commonly used to characterize semiconductors. We present the hardware design, interfacing, and test results of our SMU. Specifically, we present representative I-V curve measurements for graphene-channel FETs to demonstrate the SMU’s capability to efficiently characterize these devices with minimal noise and sufficient accuracy. This cost-effective solution presents a promising avenue for researchers and developers seeking reliable tools for sensor development and characterization. We demonstrate, with the example of surface illumination, how the sensing behavior of graphene-channel FETs can be characterized without the need for expensive equipment. Additionally, the SMU was validated with known passive and active components, along with probe station integration for semiconductor die-scale connection. The SMU’s focus on collecting I-V curves, coupled with its ability to identify device defects, such as parasitic Schottky junctions or a failed oxide, contributes to its utility in quality testing for semiconductor devices. Its low-cost nature makes it accessible for various research endeavors, enabling efficient data collection and analysis for graphene-based and other nanomaterial-based sensor applications. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic cross-section of a graphene-based sensor FET. The drain-source channel is formed of a graphene nanolayer, whose conductance is influenced by the gate electric field. The entire bulk silicon forms the gate, and a back-gate connection allows the gate potential to be applied. When the graphene is functionalized and binds to the target analyte, the conductivity changes (specifically, the Dirac point shifts), which, in turn, allows the detection of the analyte.</p>
Full article ">Figure 2
<p>Photo of the semiconductor probe station connected to the source measure unit (SMU) hardware (<b>A</b>) presented in this paper. The probe station offers BNC terminals (<b>B</b>) for all three contact micro-positioners (<b>C</b>). A red laser (<b>D</b>) is mounted on a post and allows us to illuminate the device under test with varying light flux levels.</p>
Full article ">Figure 3
<p>Overview of the SMU functional groups. Two identical active channels (A and B) serve either as a controlled-voltage or -current source with respect to the reference ground and Probe C. Each channel’s actual voltage and current are sensed and provided to the datalogger.</p>
Full article ">Figure 4
<p>Detailed overview schematic of one of the two identical active channels. At the center is the output amplifier, which is embedded in a feedback loop to either provide a controlled current (I-mode) or a controlled voltage (V-mode). V- or I-mode is selected with an electronic switch. The current is sensed with the resistor R<sub>s</sub>, and the proportional voltage is presented at the interface, as is the actual probe voltage. Additional digital signals allow the selection of the current or voltage range, and a physical relay is provided to separate the DUT from the circuit until it has been fully configured.</p>
Full article ">Figure 5
<p>Photos of the SMU and datalogger component. (<b>A</b>) Exterior view showing the front panel, which is cut from a copper-clad printed circuit board (PCB) and grounded at the point labeled (4). The BNC connections for the active probes (1) and (2) and two identical ground probes (3) are visible; the ground probe is connected to its own analog ground. The front panel also includes an LCD (5) and additional functions (6) such as a trigger/abort switch, power output, and temperature sensor jacks. (<b>B</b>) Interior view of the entire device, which shows the SMU board, datalogger/CPU board, and the linear power supply (PSU) in context. The SMU board is shielded with copper-clad PCB. To make the SMU board visible, the upper shield, connected to (4) similar to the lower shield, was removed. (<b>C</b>) View of the SMU analog board: (1) and (2) are the active probe outputs, and (3) is the analog ground-referenced header for the passive probe connectors. The sensitive section of the amplifier circuits is limited to the ICs labeled U5 though U8 with the connectors (1) and (2). Also visible are the current gain resistors (6) held in sockets for easy replacement. A flat ribbon cable, connected at (7), carries the digital and low-impedance analog signals, and the header (8) can be used for an optional 16-bit port expander with a serial peripheral interface (SPI) with U1 and U2. The interface elements (7) and (8) allow the SMU to be connected to a wide range of data acquisition hardware with a minimum of four digital outputs, two analog outputs, and two analog inputs.</p>
Full article ">Figure 6
<p>Software interface for source and measurement control of the SMU data collection: (<b>a</b>) flowchart; (<b>b</b>) GUI during measurement. The controls shown on the right side of the GUI act as spin-dials to trigger the flowchart from start to end. Current ranges from –100 μA to +100 μA and voltage ranges from –10 V to +10 V.</p>
Full article ">Figure 7
<p>Results of 1 MΩ resistor test. The I-V curve shows the expected linear behavior of the resistor through the origin with a slope of 0.99 (spanning across ±5 μA—Current Mode 5) based on an input of ±5 V (Voltage Mode 8). The response validated the calibration and SMU functionality to allow for additional testing.</p>
Full article ">Figure 8
<p>Test of temperature stability of the SMU. The SMU was connected to a 1 MOhm resistor and placed inside a cell culture incubator. For the experiment, CO<sub>2</sub> was disabled. V-I measurement with the ±5 μA sensitivity range was performed at room temperature (22 °C) and repeated after one-hour equilibration at 25 °C, 27 °C, and 30 °C. The inset shows a magnified section (gray rectangle) to highlight the datapoints acquired at each temperature. Linear regression gives the slope (i.e., resistance in MΩ) as 0.9867 for 22 °C, 0.9852 for 25 °C, 0.9853 for 27 °C, and 0.9854 for 30 °C.</p>
Full article ">Figure 9
<p>TP2450 p-channel MOSFET datasheet for output and transfer characteristics [<a href="#B20-sensors-24-03841" class="html-bibr">20</a>]: (<b>a</b>) output characteristics; (<b>b</b>) transfer characteristics. The I-V curves serve as the gold standard for the measurements collected in <a href="#sensors-24-03841-f008" class="html-fig">Figure 8</a> and <a href="#sensors-24-03841-f009" class="html-fig">Figure 9</a>.</p>
Full article ">Figure 10
<p>TP2540 P-channel MOSFET using SMU. (<b>a</b>) Schematic for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>D</mi> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math> v <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> curve. The source is connected to the SMU ground, Probe B is connected to the MOSFET gate, and Probe A is connected to the MOSFET drain. A 1 MΩ resistor was attached across the drain and Probe A to limit the current to the SMU max range. (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>D</mi> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math> v <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> curve. The collected curve can be compared to the datasheet in <a href="#sensors-24-03841-f009" class="html-fig">Figure 9</a>.</p>
Full article ">Figure 11
<p>TP2540 P-channel MOSFET arrangement using SMU. (<b>a</b>) The source is connected to the SMU ground, Probe B is connected to the MOSFET drain, and Probe A is connected to the MOSFET gate. A 1 MΩ resistor was attached across the drain and Probe A to limit the current to the SMU max range. (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>G</mi> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math> v <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> curve. The collected curve was compared to the datasheet in <a href="#sensors-24-03841-f007" class="html-fig">Figure 7</a> to confirm device accuracy.</p>
Full article ">Figure 12
<p>Acquisition of the electrical characteristics of an NPN transistor 2N3019. The measurement was performed to test the integration of the SMU with the probe station. For this purpose, the top of the TO-5 can was carefully removed to expose the transistor die. The base region (left probe needle) and emitter region (right probe needle) were directly contacted. The collector was connected through the conductive chuck.</p>
Full article ">Figure 13
<p>Family of <span class="html-italic">I</span><sub>C</sub> over <span class="html-italic">V</span><sub>CE</sub> curves for a medium-power NPN transistor (2N3019), probed at the die level. The curves were acquired at different base currents, and the linear region (<span class="html-italic">V</span><sub>CE</sub> &lt; 0.2 V) can easily be distinguished from the forward active region (<span class="html-italic">V</span><sub>CE</sub> &gt; 0.2 V). In the latter, <span class="html-italic">I</span><sub>C</sub> almost exclusively depends on <span class="html-italic">I</span><sub>B</sub> with the ratio <span class="html-italic">I</span><sub>C</sub>/<span class="html-italic">I</span><sub>B</sub> ranging from approximately 100 to 160 in the covered range for <span class="html-italic">I</span><sub>B</sub>.</p>
Full article ">Figure 14
<p>Schottky junction testing assembly. (<b>a</b>) Testing schematic: A Schottky diode was connected to Probe A in reverse polarity to mimic the metal-to-silicon junction that appears when the oxide breaks down, and Probe B was connected to the other side of the diode and a 10 MΩ resistor, which was also connected to GND. (<b>b</b>) Schematic of an actual graphene FET with the SiO<sub>2</sub> layer damaged by excessive wire bonding force. The graphene-channel FET with a breached oxide layer on the drain electrode was compared to the testing setup for quality testing. Probe A was connected to the gate, Probe B to the drain, and GND to the source.</p>
Full article ">Figure 15
<p>Schottky junction quality test results. (<b>a</b>) I-V curve of the model Schottky junction. The curve shows a steep increase in <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> to reach the current level that the gate is driving at the negative <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>G</mi> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math> range on account of the Schottky diode. (<b>b</b>) I-V curve for the graphene-channel FET with breached oxide layer (Schottky junction). This behavior allows rapid assessment of faulty sensor chips.</p>
Full article ">Figure 16
<p>Graphene-channel FET schematic. As seen in <a href="#sensors-24-03841-f001" class="html-fig">Figure 1</a>, the graphene is located across the source and drain electrodes for analyte detection. For the <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>G</mi> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math> v <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> curve, the gate is connected to Probe A, the drain is connected to Probe B, and the source is connected to GND.</p>
Full article ">Figure 17
<p>Graphene-channel FET: <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math> as a function of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>G</mi> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math>. Graphene allows conduction either through holes or through electrons. A negative gate voltage inhibits conduction through electrons, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math> is carried by the holes. Conversely, a positive gate voltage inhibits conduction through holes, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math> is carried by electrons. The point where the two carrier mechanisms balance out is known as the Dirac point and can be seen in this curve as the minimum <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math> of 30 μA when <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>G</mi> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math> = 0.25 V.</p>
Full article ">Figure 18
<p>Graphene-channel FET <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>G</mi> <mi>S</mi> <mo> </mo> </mrow> </msub> </mrow> </semantics></math> v <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> control curve to juxtapose against laser response. The curve shown in <a href="#sensors-24-03841-f017" class="html-fig">Figure 17</a> was recollected using a different device in order to confirm functionality and prepare for laser illumination testing. The Dirac point shifts to the right and drops much lower compared to the first device tested (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>G</mi> <mi>S</mi> <mo> </mo> </mrow> </msub> <mo>=</mo> </mrow> </semantics></math> ~2.8 at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> <mo>=</mo> </mrow> </semantics></math> ~2.3 µA).</p>
Full article ">Figure 19
<p>G-FET <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>G</mi> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math> v <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> curve laser response. The same sensor used in <a href="#sensors-24-03841-f018" class="html-fig">Figure 18</a> was used with a laser light of wavelength 632 nm illuminating the graphene channel. The excitation resulted in the response shifting upwards by ~48 µA.</p>
Full article ">
16 pages, 3407 KiB  
Article
Performance Projection of Vacuum Gate Dielectric Doping-Free Carbon Nanoribbon/Nanotube Field-Effect Transistors for Radiation-Immune Nanoelectronics
by Khalil Tamersit, Abdellah Kouzou, José Rodriguez and Mohamed Abdelrahem
Nanomaterials 2024, 14(11), 962; https://doi.org/10.3390/nano14110962 - 1 Jun 2024
Cited by 2 | Viewed by 1176
Abstract
This paper investigates the performance of vacuum gate dielectric doping-free carbon nanotube/nanoribbon field-effect transistors (VGD-DL CNT/GNRFETs) via computational analysis employing a quantum simulation approach. The methodology integrates the self-consistent solution of the Poisson solver with the mode space non-equilibrium Green’s function (NEGF) in [...] Read more.
This paper investigates the performance of vacuum gate dielectric doping-free carbon nanotube/nanoribbon field-effect transistors (VGD-DL CNT/GNRFETs) via computational analysis employing a quantum simulation approach. The methodology integrates the self-consistent solution of the Poisson solver with the mode space non-equilibrium Green’s function (NEGF) in the ballistic limit. Adopting the vacuum gate dielectric (VGD) paradigm ensures radiation-hardened functionality while avoiding radiation-induced trapped charge mechanisms, while the doping-free paradigm facilitates fabrication flexibility by avoiding the realization of a sharp doping gradient in the nanoscale regime. Electrostatic doping of the nanodevices is achieved via source and drain doping gates. The simulations encompass MOSFET and tunnel FET (TFET) modes. The numerical investigation comprehensively examines potential distribution, transfer characteristics, subthreshold swing, leakage current, on-state current, current ratio, and scaling capability. Results demonstrate the robustness of vacuum nanodevices for high-performance, radiation-hardened switching applications. Furthermore, a proposal for extrinsic enhancement via doping gate voltage adjustment to optimize band diagrams and improve switching performance at ultra-scaled regimes is successfully presented. These findings underscore the potential of vacuum gate dielectric carbon-based nanotransistors for ultrascaled, high-performance, energy-efficient, and radiation-immune nanoelectronics. Full article
(This article belongs to the Section Nanoelectronics, Nanosensors and Devices)
Show Figures

Figure 1

Figure 1
<p>Lengthwise cut views of (<b>a</b>) the conventional GAA CNT(T)FET and (<b>b</b>) DG GNR(T)FET. Three-dimensional structures of (<b>c</b>) the proposed VGD-DLCNT(T)FET and (<b>d</b>) VGD-DLGNR(T)FET. Lengthwise cut views of (<b>e</b>) the proposed VGD-DLCNT(T)FET and (<b>f</b>) VGD-DLGNR(T)FET. Cross-sectional views of (<b>g</b>) the proposed VGD-DLCNT(T)FET and (<b>h</b>) VGD-DLGNR(T)FET.</p>
Full article ">Figure 2
<p>Illustration of the self-consistent procedure in NEGF-based quantum simulation.</p>
Full article ">Figure 3
<p>Two-dimensional potential distributions of (<b>a</b>) VGD-DLCNTFET, (<b>b</b>) VGD-DLCNTTFET, (<b>c</b>) VGD-DLGNRFET, and (<b>d</b>) VGD-DLGNRTFET.</p>
Full article ">Figure 4
<p>I<sub>DS</sub>-V<sub>GS</sub> transfer characteristics of (<b>a</b>) the VGD-DLCNT(T)FET and (<b>b</b>) the VGD-DLGNR(T)FET.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>) Subthreshold swing versus the gate-to-source voltage, (<b>c</b>,<b>d</b>) I<sub>ON</sub>/I<sub>OFF</sub> current ratio versus I<sub>ON</sub> for VGD-DLCNT(T)FET (<b>left</b> figures) and VGD-DLGNR(T)FET (<b>right</b> figures).</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) Subthreshold swing; (<b>c</b>,<b>d</b>) MRCR versus the control gate length for VGD-DLCNT(T)FETs (<b>left</b> figures) and VGD-DLGNR(T)FETs (<b>right</b> figures).</p>
Full article ">Figure 7
<p>I<sub>DS</sub>-V<sub>GS</sub> transfer characteristics of (<b>a</b>,<b>b</b>) the VGD-DLCNT(T)FETs, (<b>c</b>,<b>d</b>) VGD-DLGNR(T)FETs before and after irradiation, with conventional CNT/GNR-based (T)FETs included for comparison.</p>
Full article ">Figure 8
<p>The I<sub>DS</sub>-V<sub>GS</sub> transfer characteristics of (<b>a</b>) MOSFET-like vacuum nanodevices and (<b>b</b>) TFET vacuum nanodevices for different electrical doping voltages.</p>
Full article ">
12 pages, 1316 KiB  
Article
Comparative Study of Field-Effect Transistors Based on Graphene Oxide and CVD Graphene in Highly Sensitive NT-proBNP Aptasensors
by Anastasiia Kudriavtseva, Stefan Jarić, Nikita Nekrasov, Alexey V. Orlov, Ivana Gadjanski, Ivan Bobrinetskiy, Petr I. Nikitin and Nikola Knežević
Biosensors 2024, 14(5), 215; https://doi.org/10.3390/bios14050215 - 26 Apr 2024
Cited by 1 | Viewed by 2249
Abstract
Graphene-based materials are actively being investigated as sensing elements for the detection of different analytes. Both graphene grown by chemical vapor deposition (CVD) and graphene oxide (GO) produced by the modified Hummers’ method are actively used in the development of biosensors. The production [...] Read more.
Graphene-based materials are actively being investigated as sensing elements for the detection of different analytes. Both graphene grown by chemical vapor deposition (CVD) and graphene oxide (GO) produced by the modified Hummers’ method are actively used in the development of biosensors. The production costs of CVD graphene- and GO-based sensors are similar; however, the question remains regarding the most efficient graphene-based material for the construction of point-of-care diagnostic devices. To this end, in this work, we compare CVD graphene aptasensors with the aptasensors based on reduced GO (rGO) for their capabilities in the detection of NT-proBNP, which serves as the gold standard biomarker for heart failure. Both types of aptasensors were developed using commercial gold interdigitated electrodes (IDEs) with either CVD graphene or GO formed on top as a channel of liquid-gated field-effect transistor (FET), yielding GFET and rGO-FET sensors, respectively. The functional properties of the two types of aptasensors were compared. Both demonstrate good dynamic range from 10 fg/mL to 100 pg/mL. The limit of detection for NT-proBNP in artificial saliva was 100 fg/mL and 1 pg/mL for rGO-FET- and GFET-based aptasensors, respectively. While CVD GFET demonstrates less variations in parameters, higher sensitivity was demonstrated by the rGO-FET due to its higher roughness and larger bandgap. The demonstrated low cost and scalability of technology for both types of graphene-based aptasensors may be applicable for the development of different graphene-based biosensors for rapid, stable, on-site, and highly sensitive detection of diverse biochemical markers. Full article
(This article belongs to the Special Issue Nanotechnology-Enabled Biosensors)
Show Figures

Figure 1

Figure 1
<p>Characterisation of CVD graphene and GO films transferred on IDE electrodes. AFM-image of graphene (<b>a</b>) and rGO film (<b>b</b>) on glass surface (between electrodes). Height profiles of cross-sections of AFM images, shown on (<b>a</b>) and (<b>b</b>) for graphene and rGO, respectively (<b>c</b>). Raman spectra for graphene and reduced GO films on glass surface (<b>d</b>).</p>
Full article ">Figure 2
<p>Immobilization of aptamer on the channel of FET. Change in I<sub>D</sub> – V<sub>G</sub> curves for GFETs (in 0.1× and 0.01×PBS) (<b>a</b>) and rGO-FET (in 1×PBS) (<b>b</b>) after assembling to aptasensor (V<sub>DS</sub> = 100 mV). Dirac point shift change for each step of sensor assembly for GFET and rGO-FET (<b>c</b>). Schematic illustrations of electrical response mechanism in GFET and rGO-FET (<b>d</b>).</p>
Full article ">Figure 3
<p>The GFET and rGO-FET aptasensors’ response to NT-proBNP in 0.01×PBS (pH 7.4). (<b>a</b>) Transfer curves after stabilization for varied NT-proBNP concentration with no washing steps for GFET (solid lines) and rGO-FET (dashed lines), with applied V<sub>DS</sub> = 100 mV and 10 mV for rGO-FET and GFET, respectively. (<b>b</b>) Dirac point shift for each concentration for GFET (●) and rGO-FET (■, data is reproduced with permission from [<a href="#B10-biosensors-14-00215" class="html-bibr">10</a>]). (<b>c</b>) Transfer curves for three NT-proBNP concentration in AS, dissolved in 0.01×PBS for GFET (solid lines) and rGO-FET (dashed lines), with applied V<sub>DS</sub> = 50 mV. (<b>d</b>) Dirac point shift for each concentration of NT-proBNP in AS, diluted in 0.01×PBS for GFET (■) and rGO-FET (●).</p>
Full article ">
Back to TopTop