[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

remove_circle_outline

Search Results (203)

Search Parameters:
Keywords = tin dioxide

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
26 pages, 1669 KiB  
Review
Advancements in Surface Modification of NiTi Alloys for Orthopedic Implants: Focus on Low-Temperature Glow Discharge Plasma Oxidation Techniques
by Justyna Witkowska, Jerzy Sobiecki and Tadeusz Wierzchoń
Int. J. Mol. Sci. 2025, 26(3), 1132; https://doi.org/10.3390/ijms26031132 - 28 Jan 2025
Viewed by 337
Abstract
Nickel–titanium (NiTi) shape memory alloys are promising materials for orthopedic implants due to their unique mechanical properties, including superelasticity and shape memory effect. However, the high nickel content in NiTi alloys raises concerns about biocompatibility and potential cytotoxic effects. This review focuses on [...] Read more.
Nickel–titanium (NiTi) shape memory alloys are promising materials for orthopedic implants due to their unique mechanical properties, including superelasticity and shape memory effect. However, the high nickel content in NiTi alloys raises concerns about biocompatibility and potential cytotoxic effects. This review focuses on the recent advancements in surface modification techniques aimed at enhancing the properties of NiTi alloys for biomedical applications, with particular emphasis on low-temperature glow discharge plasma oxidation methods. The review explores various surface engineering strategies, including oxidation, nitriding, ion implantation, laser treatments, and the deposition of protective coatings. Among these, low-temperature plasma oxidation stands out for its ability to produce uniform, nanocrystalline layers of titanium dioxide (TiO2), titanium nitride (TiN), and nitrogen-doped TiO2 layers, significantly enhancing corrosion resistance, reducing nickel ion release, and promoting osseointegration. Plasma-assisted oxynitriding processes enable the creation of multifunctional coatings with improved mechanical and biological properties. The applications of modified NiTi alloys in orthopedic implants, including spinal fixation devices, joint prostheses, and fracture fixation systems, are also discussed. Despite these promising advancements, challenges remain in achieving large-scale reproducibility, controlling process parameters, and reducing production costs. Future research directions include integrating bioactive and antibacterial coatings, enhancing surface structuring for controlled biological responses, and expanding clinical validation. Addressing these challenges can unlock the full potential of surface-modified NiTi alloys in advanced orthopedic applications for safer, longer-lasting, and more effective medical implants. Full article
(This article belongs to the Special Issue Biomaterials for Dental and Orthopedic Applications)
18 pages, 12286 KiB  
Article
Effects of Annealing Conditions on the Catalytic Performance of Anodized Tin Oxide for Electrochemical Carbon Dioxide Reduction
by Nicolò B. D. Monti, Juqin Zeng, Micaela Castellino, Samuele Porro, Mitra Bagheri, Candido F. Pirri, Angelica Chiodoni and Katarzyna Bejtka
Nanomaterials 2025, 15(2), 121; https://doi.org/10.3390/nano15020121 - 16 Jan 2025
Viewed by 537
Abstract
The electrochemical reduction of CO2 (CO2RR) to value-added products has garnered significant interest as a sustainable solution to mitigate CO2 emissions and harness renewable energy sources. Among CO2RR products, formic acid/formate (HCOOH/HCOO) is particularly attractive [...] Read more.
The electrochemical reduction of CO2 (CO2RR) to value-added products has garnered significant interest as a sustainable solution to mitigate CO2 emissions and harness renewable energy sources. Among CO2RR products, formic acid/formate (HCOOH/HCOO) is particularly attractive due to its industrial relevance, high energy density, and potential candidate as a liquid hydrogen carrier. This study investigates the influence of the initial oxidation state of tin on CO2RR performance using nanostructured SnOx catalysts. A simple, quick, scalable, and cost-effective synthesis strategy was employed to fabricate SnOx catalysts with controlled oxidation states while maintaining consistent morphology and particle size. The catalysts were characterized using SEM, TEM, XRD, Raman, and XPS to correlate structure and surface properties with catalytic performance. Electrochemical measurements revealed that SnOx catalysts annealed in air at 525 °C exhibited the highest formate selectivity and current density, attributed to the optimized oxidation state and the presence of oxygen vacancies. Flow cell tests further demonstrated enhanced performance under practical conditions, achieving stable formate production with high faradaic efficiency over prolonged operation. These findings highlight the critical role of tin oxidation states and surface defects in tuning CO2RR performance, offering valuable insights for the design of efficient catalysts for CO2 electroreduction to formate. Full article
(This article belongs to the Section Energy and Catalysis)
Show Figures

Figure 1

Figure 1
<p>TEM characterization of all studied catalysts: top line BF-TEM and bottom line HRTEM for Pristine, Air′370 °C, Air′525 °C and N2′525 °C.</p>
Full article ">Figure 2
<p>XRD patterns (<b>A</b>) and Raman spectra (<b>B</b>) obtained from each sample.</p>
Full article ">Figure 3
<p>XPS HR spectra for (<b>A</b>) C1s, (<b>B</b>) O1s, (<b>C</b>) Sn3d, and (<b>D</b>) Valence Band regions.</p>
Full article ">Figure 4
<p>FE towards CO<sub>2</sub>RR products and the current density dedicated to their production. Each sample is labelled by the above legend, and its three studied potential are grouped and identified by the labelled below.</p>
Full article ">Figure 5
<p>Possible mechanisms for CO<sub>2</sub> ECR to formate/formic acid. (<b>A</b>) monodentate intermediate route; (<b>B</b>,<b>C</b>) CO<sub>2</sub><sup>−</sup> radical intermediate route; and (<b>D</b>) surface-bound carbonate intermediate route. Adapted from [<a href="#B75-nanomaterials-15-00121" class="html-bibr">75</a>].</p>
Full article ">Figure 6
<p>Catalyst Air-525 °C: (<b>A</b>): FEs distribution and current densities at selected potentials; (<b>B</b>) long-term stability was assessed over 30 h at −1.2 V in 2 M KHCO<sub>3</sub>.</p>
Full article ">
16 pages, 4313 KiB  
Article
Mimicking Axon Growth and Pruning by Photocatalytic Growth and Chemical Dissolution of Gold on Titanium Dioxide Patterns
by Fatemeh Abshari, Moritz Paulsen, Salih Veziroglu, Alexander Vahl and Martina Gerken
Molecules 2025, 30(1), 99; https://doi.org/10.3390/molecules30010099 - 30 Dec 2024
Viewed by 513
Abstract
Biological neural circuits are based on the interplay of excitatory and inhibitory events to achieve functionality. Axons form long-range information highways in neural circuits. Axon pruning, i.e., the removal of exuberant axonal connections, is essential in network remodeling. We propose the photocatalytic growth [...] Read more.
Biological neural circuits are based on the interplay of excitatory and inhibitory events to achieve functionality. Axons form long-range information highways in neural circuits. Axon pruning, i.e., the removal of exuberant axonal connections, is essential in network remodeling. We propose the photocatalytic growth and chemical dissolution of gold lines as a building block for neuromorphic computing mimicking axon growth and pruning. We predefine photocatalytic growth areas on a surface by structuring titanium dioxide (TiO2) patterns. Placing the samples in a gold chloride (HAuCl4) precursor solution, we achieve the controlled growth of gold microstructures along the edges of the indium tin oxide (ITO)/TiO2 patterns under ultraviolet (UV) illumination. A potassium iodide (KI) solution is employed to dissolve the gold microstructures. We introduce a real-time monitoring setup based on an optical transmission microscope. We successfully observe both the growth and dissolution processes. Additionally, scanning electron microscopy (SEM) analysis confirms the morphological changes before and after dissolution, with dissolution rates closely aligned to the growth rates. These findings demonstrate the potential of this approach to emulate dynamic biological processes, paving the way for future applications in adaptive neuromorphic systems. Full article
(This article belongs to the Special Issue Photocatalytic Materials and Photocatalytic Reactions)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Simplified schematic representation of a neural network (<b>left</b>) and its abstraction into 2D schematic geometry on a surface (<b>right</b>), retaining the essential connectivity of the network. (<b>b</b>) Process for mimicking axon-like connections: (<b>left</b>) a patterned template defines regions for selective gold growth, (<b>center</b>) UV illumination induces photocatalytic gold growth along predefined pathways, forming axon-like structures, and (<b>right</b>) chemical dissolution selectively prunes the gold connections, mimicking axonal pruning.</p>
Full article ">Figure 2
<p>Real-time monitoring of gold growth (<b>a</b>) Sequential optical microscope images of a selected region around the TiO<sub>2</sub> edge, including the TiO<sub>2</sub> surface on the left side and the glass substrate on the right side, taken at various time points, <span class="html-italic">t</span> = 0, <span class="html-italic">t</span> = 10, <span class="html-italic">t</span> = 20, and <span class="html-italic">t</span> = 30 min, illustrating the gradual growth of gold structures over 30 min. (<b>b</b>) Transmission profiles across the TiO<sub>2</sub> edge at <span class="html-italic">t</span> = 0, <span class="html-italic">t</span> = 10, <span class="html-italic">t</span> = 20, and <span class="html-italic">t</span> = 30 min, normalized to the reference transmission on the glass substrate. (<b>c</b>) Time-dependent transmission curves showing dissolution progress at the TiO<sub>2</sub> edge and surface compared to the glass substrate.</p>
Full article ">Figure 3
<p>Real-time monitoring of gold dissolution (<b>a</b>) Sequential optical microscope images of a selected region around the TiO<sub>2</sub> edge, including the TiO<sub>2</sub> surface on the left side and the glass substrate on the right side, taken at various time points, <span class="html-italic">t</span> = 0, <span class="html-italic">t</span> = 5, <span class="html-italic">t</span> = 10, and <span class="html-italic">t</span> = 30 min, illustrating the gradual dissolution of gold structures over 30 min. (<b>b</b>) Transmission profiles across the TiO<sub>2</sub> edge at <span class="html-italic">t</span> = 0, <span class="html-italic">t</span> = 10, <span class="html-italic">t</span> = 20, and <span class="html-italic">t</span> = 30 min, normalized to the reference transmission on the glass substrate. (<b>c</b>) Time-dependent transmission curves showing dissolution progress at the TiO<sub>2</sub> edge and surface compared to the glass substrate.</p>
Full article ">Figure 4
<p>SEM images of the template post-growth, illustrating the formation of gold particles along the edges of the TiO<sub>2</sub> patterns.</p>
Full article ">Figure 5
<p>SEM images of gold particles on TiO<sub>2</sub>: (<b>a</b>) Before dissolution, with flower-shaped particles on the surface; (<b>b</b>) After dissolution, showing reduced and irregularly shaped particles, though not fully removed.</p>
Full article ">Figure 6
<p>Schematic representation of the substrate preparation process. The glass substrate is negatively patterned with AZ5214E photoresist after lithography. A 6 nm ITO layer and a 70 nm TiO<sub>2</sub> layer are then deposited via sputtering. After the lift-off process, the final patterned TiO<sub>2</sub> structures with the underlying ITO layer are revealed.</p>
Full article ">Figure 7
<p>Schematic of the photocatalytic illumination setup. The sample with TiO<sub>2</sub> structures is submerged in the HAuCl<sub>4</sub> solution and illuminated by two angled UV LEDs (365 nm) under a transmission microscope connected to a camera.</p>
Full article ">Figure 8
<p>Visual representation of the KI solution diluted with DI water at different ratios. From left to right: pure KI solution, 1:200 KI solution to DI water, 1:300 KI solution to DI water, and 1:400 KI solution to DI water.</p>
Full article ">
17 pages, 3272 KiB  
Article
ITO-TiO2 Heterojunctions on Glass Substrates for Photocatalytic Gold Growth Along Pattern Edges
by Fatemeh Abshari, Moritz Paulsen, Salih Veziroglu, Alexander Vahl and Martina Gerken
Catalysts 2024, 14(12), 940; https://doi.org/10.3390/catal14120940 - 19 Dec 2024
Viewed by 540
Abstract
This study investigates the effects of varying indium tin oxide (ITO) layer thicknesses and the patterning of the ITO layer on the growth of metallic gold (Au) nano- and microstructures on titanium dioxide (TiO2) templates. The ITO-TiO2 heterojunction serves to [...] Read more.
This study investigates the effects of varying indium tin oxide (ITO) layer thicknesses and the patterning of the ITO layer on the growth of metallic gold (Au) nano- and microstructures on titanium dioxide (TiO2) templates. The ITO-TiO2 heterojunction serves to collect photogenerated electrons in the ITO sublayer, facilitating their transport to the pattern edges and concentrating photocatalytic activity at these edges. Six template types were fabricated on glass substrates, with systematic variations in ITO thickness (0, 3, 6, 10, and 30 nm) and different ITO patterning methods (either continuous or patterned with the TiO2 layer). Photocatalytic gold growth was carried out on each of the substrates, and morphological analysis was conducted using scanning electron microscopy (SEM). Results showed that a 6 nm ITO layer beneath a 70 nm TiO2 layer yielded the most uniform gold lines, characterized by 3D flower-shaped structures and enhanced edge growth. Conductance measurements indicated a value of 23 mS, suggesting potential applications in bio-inspired electronics. These findings provide insights into optimizing gold structure growth for advanced neuromorphic devices. Full article
(This article belongs to the Special Issue State-of-the-Art of Heterostructured Photocatalysts)
Show Figures

Figure 1

Figure 1
<p>Schematic illustrations of template structures and the illumination setup. (<b>a</b>) Three types of TiO<sub>2</sub> patterns on glass substrates: the top shows a TiO<sub>2</sub> pattern on glass, the middle shows a TiO<sub>2</sub> pattern with a thin ITO layer beneath TiO<sub>2</sub> on glass, and the bottom shows a TiO<sub>2</sub> pattern with a thin ITO layer covering the entire glass substrate. (<b>b</b>) Schematic of the illumination setup, showing the template placed in a beaker with HAuCl<sub>4</sub> solution under UV LED illumination (365 nm).</p>
Full article ">Figure 2
<p>SEM images of Type 1/TiO<sub>2</sub>-70_ITO-0 template showing two types of grown gold particles on the surface of TiO<sub>2</sub>: (<b>a</b>) Sample 1, and (<b>b</b>) Sample 2, highlighting the presence of 2D stacks and smaller spherical particles. A partially formed 2D stack is indicated by a red square in (<b>a</b>).</p>
Full article ">Figure 3
<p>SEM images of Type 2/TiO<sub>2</sub>-70_ITO-3 template showing the morphology of grown gold particles on the surface of TiO<sub>2</sub>: (<b>a</b>) Sample 1 and (<b>b</b>) Sample 2 depicting the distribution of 2D stacks and smaller spherical particles. A partially formed 2D stack is indicated by a red square in (<b>a</b>).</p>
Full article ">Figure 4
<p>SEM images of Type 3/TiO<sub>2</sub>-70_ITO-6 template after gold growth: (<b>a</b>) Sample 1 and (<b>b</b>) Sample 2, highlighting the emergence of 3D flower-like structures along the edges of the TiO<sub>2</sub> patterns, alongside isolated spherical particles and limited 2D stacks.</p>
Full article ">Figure 5
<p>SEM images of the TiO<sub>2</sub> surface with manual scratches of varying widths, showing the growth of gold particles on the TiO<sub>2</sub> surface. The nucleation of gold occurs primarily within the scratches.</p>
Full article ">Figure 6
<p>SEM images of Type 4/TiO<sub>2</sub>-70_ITO-10 template after gold growth: (<b>a</b>) Sample 1 and (<b>b</b>) Sample 2, showing distinct morphologies of gold particles including sparse 2D stacks and unevenly distributed 3D flower-shaped particles on the TiO<sub>2</sub> surface, with visible gaps indicating limited aggregation.</p>
Full article ">Figure 7
<p>SEM images of Type 5/TiO<sub>2</sub>-70_ITO-30 template after gold growth: (<b>a</b>) Sample 1 and (<b>b</b>) Sample 2, illustrating three distinct morphologies of gold particles, including spherical particles, 2D stacks with varied shapes, and hedgehog-like flower structures on the TiO<sub>2</sub> surface, with varying distribution and density.</p>
Full article ">Figure 8
<p>SEM images of Type 6/TiO<sub>2</sub>-70_ITO-6-unstr. template after gold growth: (<b>a</b>) Sample 1 and (<b>b</b>) Sample 2, illustrating primarily 2D stacks in low quantities, 3D flower-shaped particles on the TiO<sub>2</sub> surface, and spherical particles on the ITO layer, with attempts at cluster interaction but no formation of a continuous line along the edges.</p>
Full article ">Figure 9
<p>Schematic illustration of the substrate preparation process. (<b>a</b>) Lithography process steps utilizing AZ5214E photoresist on soda-lime glass substrates. (<b>b</b>) Schematic representation of the deposition processes for ITO and TiO<sub>2</sub> across different template types (Type 1/TiO<sub>2</sub>-70_ITO-0: TiO<sub>2</sub> only, Type 2: TiO<sub>2</sub> with 3 nm ITO, Type 3: TiO<sub>2</sub> with 6 nm ITO, Type 4: TiO<sub>2</sub> with 10 nm ITO, Type 5: TiO<sub>2</sub> with 30 nm ITO, and Type 6: uniform 6 nm ITO layer with TiO<sub>2</sub>.</p>
Full article ">
14 pages, 4160 KiB  
Article
Selective CO2 Detection at Room Temperature with Polyaniline/SnO2 Nanowire Composites
by Gen Li, Muhammad Hilal, Hyojung Kim, Jiyeon Lee, Zhiyong Chen, Bin Li, Yunhao Cui, Jian Hou and Zhicheng Cai
Coatings 2024, 14(12), 1590; https://doi.org/10.3390/coatings14121590 - 19 Dec 2024
Viewed by 523
Abstract
In this study, tin oxide (SnO2)/polyaniline (PANI) composite nanowires (NWs) with varying amounts of PANI were synthesized for carbon dioxide (CO2) gas sensing at room temperature (RT, 25 °C). SnO2 NWs were fabricated via the vapor–liquid–solid (VLS) method, [...] Read more.
In this study, tin oxide (SnO2)/polyaniline (PANI) composite nanowires (NWs) with varying amounts of PANI were synthesized for carbon dioxide (CO2) gas sensing at room temperature (RT, 25 °C). SnO2 NWs were fabricated via the vapor–liquid–solid (VLS) method, followed by coating with PANI. CO2 sensing investigations revealed that the sensor with 186 μL PANI exhibited the highest response to CO2 at RT. Additionally, the optimized sensor demonstrated excellent selectivity for CO2, long-term stability, and reliable performance across different humidity levels. The enhanced sensing performance of the optimized sensor was attributed to the formation of SnO2-PANI heterojunctions and the optimal PANI concentration. This study underscores the potential of SnO2-PANI composites for CO2 detection at RT. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the preparation process for PANI/SnO<sub>2</sub> composites.</p>
Full article ">Figure 2
<p>SEM and TEM characterization of PANI/SnO<sub>2</sub> composites: (<b>a</b>–<b>d</b>) SEM images of SP0, SP1, SP2, and SP3, respectively; (<b>e</b>–<b>h</b>) magnified SEM images showing surface morphology of individual nanowires; (<b>i</b>) TEM image of a single SP2 nanowire; (<b>j</b>) magnified TEM image revealing lattice fringes of the SnO<sub>2</sub> core and PANI coating.</p>
Full article ">Figure 3
<p>XRD patterns of PANI/SnO<sub>2</sub> composites (SP0, SP1, SP2, and SP3) showing characteristic peaks of SnO<sub>2</sub> and PANI with varying polymer content.</p>
Full article ">Figure 4
<p>XPS analysis of PANI/SnO<sub>2</sub> composites (SP0, SP1, SP2, and SP3): (<b>a</b>) Sn 3d peaks, (<b>b</b>–<b>d</b>) C 1s spectrum of PANI, (<b>e</b>–<b>g</b>) N 1s spectrum of PANI, and (<b>h</b>–<b>k</b>) O 1s peaks highlighting oxygen states.</p>
Full article ">Figure 5
<p>Gas sensing performance of PANI/SnO<sub>2</sub> composites: (<b>a</b>) dynamic response transient curves under varying CO<sub>2</sub> concentrations, (<b>b</b>) response–recovery curves, (<b>c</b>) response as a function of CO<sub>2</sub> concentration, (<b>d</b>) response and recovery times, (<b>e</b>) selectivity for various gases, (<b>f</b>) repeatability over multiple cycles, (<b>g</b>) long-term stability, and (<b>h</b>) humidity effects on CO<sub>2</sub> response and baseline resistance.</p>
Full article ">Figure 6
<p>Schematic representation of the sensing mechanism in PANI/SnO<sub>2</sub> composites.</p>
Full article ">
16 pages, 5091 KiB  
Article
Novel Sequential Detection of NO2 and C2H5OH in SnO2 MEMS Arrays for Enhanced Selectivity in E-Nose Applications
by Mahaboobbatcha Aleem, Yilu Zhou, Swati Deswal, Bongmook Lee and Veena Misra
Chemosensors 2024, 12(12), 268; https://doi.org/10.3390/chemosensors12120268 - 19 Dec 2024
Viewed by 3103
Abstract
This study explores the surface chemistry and electrical responses of ultra-high-sensitivity SnO2 MEMS arrays to enable a novel sequential detection methodology for detecting nitrogen dioxide (NO2) and ethanol (C2H5OH) as a route to achieve selective gas [...] Read more.
This study explores the surface chemistry and electrical responses of ultra-high-sensitivity SnO2 MEMS arrays to enable a novel sequential detection methodology for detecting nitrogen dioxide (NO2) and ethanol (C2H5OH) as a route to achieve selective gas sensing in electronic nose (E-nose) applications. Utilizing tin oxide (SnO2) thin films deposited via atomic layer deposition (ALD), the array achieves the lowest reported detection limits of 8 parts per billion (ppb) for NO2. The research delves into the detection mechanisms of NO2 and C2H5OH, both individually and in subsequent exposures, assessing the sensor’s dynamic response across various operating temperatures. It demonstrates rapid response and recovery times, with averages of 48 s and 277 s for NO2 and 40 and 48 for C2H5OH. Understanding the role of individual gases on the SnO2 surface chemistry is paramount in discerning subsequent gas exposure behavior. The oxidizing behavior of C2H5OH following NO2 exposure is attributed to interactions between NO2 and oxygen vacancies on the SnO2 surface, which leads to the formation of nitrate or nitrite species. These species subsequently influence interactions with C2H5OH, inducing oxidizing properties, and need to be carefully considered. Principal component analysis (PCA) was used to further improve the sensor’s capability to precisely identify and quantify gas mixtures, improving its applicability for real-time monitoring in complex scenarios. Full article
(This article belongs to the Special Issue Electronic Nose and Electronic Tongue for Substance Analysis)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Integration of a SnO<sub>2</sub> sensor within a PCB board. (<b>b</b>) Visualization of heat distribution across the micro-heater surface under applied voltage, and (<b>c</b>) subsequent temperature variation cycle over time.</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic representation for interaction of oxygen molecules on SnO<sub>2</sub> at different temperatures, (<b>b</b>) sensor response to 1 ppm NO<sub>2</sub> and 10 ppm of C<sub>2</sub>H<sub>5</sub>OH gas as a function of temperature, (<b>c</b>) NO<sub>2</sub> sensing response at various concentration levels (100 ppb to 1000 ppm) (@140 °C), (<b>d</b>) the repeatability sensing graph representing 1 ppm NO<sub>2</sub> across 10 cycles, (<b>e</b>) C<sub>2</sub>H<sub>5</sub>OH sensing response at various concentration levels (1 ppm to 6.5 ppm), (@240 °C), and (<b>f</b>) the repeatability sensing graph representing 1 ppm C<sub>2</sub>H<sub>5</sub>OH across 10 cycles.</p>
Full article ">Figure 3
<p>(<b>a</b>) Gas-sensing mechanism of NO<sub>2</sub> and (<b>b</b>) C<sub>2</sub>H<sub>5</sub>OH.</p>
Full article ">Figure 4
<p>(<b>a</b>) Sequential exposure to 1 ppm NO<sub>2</sub> and 5 ppm C<sub>2</sub>H<sub>5</sub>OH, and (<b>b</b>) sequential exposure to 5 ppm C<sub>2</sub>H<sub>5</sub>OH and 1 ppm NO<sub>2</sub>.</p>
Full article ">Figure 5
<p>(<b>a</b>) Sensor response to sequential exposures of NO<sub>2</sub> at concentrations of 50 and 100 ppb (1 cycle), followed by C<sub>2</sub>H<sub>5</sub>OH at 5 ppm (2 cycles); (<b>b</b>) response to NO<sub>2</sub> at 300 and 500 ppb (1 cycle), followed by C<sub>2</sub>H<sub>5</sub>OH at 5 ppm (2 cycles); and (<b>c</b>) response to NO<sub>2</sub> at 700, 900, and 1100 ppb (1 cycle), followed by C<sub>2</sub>H<sub>5</sub>OH at 5 ppm (2 cycles).</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>f</b>) Gas-sensing mechanism of subsequent NO<sub>2</sub> and C<sub>2</sub>H<sub>5</sub>OH sensing.</p>
Full article ">Figure 7
<p>2-D results of PCA NO<sub>2</sub> and C<sub>2</sub>H<sub>5</sub>OH, and sequential exposure of NO<sub>2</sub> and C2H<sub>5</sub>OH are distinguishable.</p>
Full article ">Scheme 1
<p>Process flow adopted for E-nose array sensor fabrication.</p>
Full article ">
17 pages, 4525 KiB  
Article
Highly Sensitive and Selective SnO2-Gr Sensor Photoactivated for Detection of Low NO2 Concentrations at Room Temperature
by Isabel Sayago, Carlos Sánchez-Vicente and José Pedro Santos
Nanomaterials 2024, 14(24), 1994; https://doi.org/10.3390/nano14241994 - 12 Dec 2024
Viewed by 658
Abstract
Chemical nanosensors based on nanoparticles of tin dioxide and graphene-decorated tin dioxide were developed and characterized to detect low NO2 concentrations. Sensitive layers were prepared by the drop casting method. SEM/EDX analyses have been used to investigate the surface morphology and the [...] Read more.
Chemical nanosensors based on nanoparticles of tin dioxide and graphene-decorated tin dioxide were developed and characterized to detect low NO2 concentrations. Sensitive layers were prepared by the drop casting method. SEM/EDX analyses have been used to investigate the surface morphology and the elemental composition of the sensors. Photoactivation of the sensors allowed for detecting ultra-low NO2 concentrations (100 ppb) at room temperature. The sensors showed very good sensitivity and selectivity to NO2 with low cross-responses to the other pollutant gases tested (CO and CH4). The effect of humidity and the presence of graphene on sensor response were studied. Comparative studies revealed that graphene incorporation improved sensor performance. Detections in complex atmosphere (CO + NO2 or CH4 + NO2, in humid air) confirmed the high selectivity of the graphene sensor in near-real conditions. Thus, the responses were of 600%, 657% and 540% to NO2 (0.5 ppm), NO2 (0.5 ppm) + CO (5 ppm) and NO2 (0.5 ppm) + CH4 (10 ppm), respectively. In addition, the detection mechanisms were discussed and the possible redox equations that can change the sensor conductance were also considered. Full article
(This article belongs to the Special Issue Advanced Nanomaterials in Gas and Humidity Sensors)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) SEM micrographs and (<b>b</b>) EDX elemental mapping images of one sensitive layer (SnO<sub>2</sub>-Gr).</p>
Full article ">Figure 2
<p>TEM images of (<b>a</b>) Gr-SnO<sub>2</sub> on grids (<b>b</b>) pristine SnO<sub>2</sub> nanoparticles, (<b>c</b>) Gr-SnO<sub>2</sub> and (<b>d</b>) HRTEM images of Gr-SnO<sub>2</sub>.</p>
Full article ">Figure 3
<p>Resistance changes in the SnO<sub>2</sub>-NPs sensor tested with and without UV-LED illumination at RT in air.</p>
Full article ">Figure 4
<p>SnO<sub>2</sub> sensor: (<b>a</b>) Dynamic response to NO<sub>2</sub> at RT in air atmosphere with and without UV-LED illumination and (<b>b</b>) responses to 0.5 ppm NO<sub>2</sub> under different conditions (with and without UV-LED, dry and humid air).</p>
Full article ">Figure 5
<p>Dynamic response curves at RT under UV-LED illumination to NO<sub>2</sub> different concentrations of the tested sensors: (<b>a</b>) SnO<sub>2</sub> and (<b>b</b>) SnO<sub>2</sub>-Gr. (<b>c</b>) Response of the SnO<sub>2</sub> and SnO<sub>2</sub>-Gr sensors to 0.1, 0.3 and 0.5 ppm NO<sub>2</sub> at RT under UV-LED illumination in dry and humid air (50% RH). (<b>d</b>) Sensor responses versus NO<sub>2</sub> gas concentration in dry and humid air (50% RH) with UV-LED illumination, where the circles denote experimental results and the dotted lines represent fitting curves.</p>
Full article ">Figure 6
<p>(<b>a</b>) Responses of the sensors to 0.3 ppm NO<sub>2</sub>, 5 ppm CO and 5 ppm CH<sub>4</sub> in dry and wet air. (<b>b</b>) Selectivity of the tested sensors to NO<sub>2</sub> at RT and under UV-LED illumination, in dry and humid air (50%). SnO<sub>2</sub>-Gr sensor dynamic response at RT in humid air (45% RH) and under UV-LED illumination to different gas mixtures: (<b>c</b>) mixture 1 (NO<sub>2</sub> + CO) (<b>d</b>) mixture 2 (NO<sub>2</sub> + CH<sub>4</sub>).</p>
Full article ">Figure 7
<p>Detection mechanism scheme.</p>
Full article ">
17 pages, 2885 KiB  
Article
Advanced SnO2 Thin Films: Stability and Sensitivity in CO Detection
by Nadezhda K. Maksimova, Tatiana D. Malinovskaya, Valentina V. Zhek, Nadezhda V. Sergeychenko, Evgeniy V. Chernikov, Denis V. Sokolov, Aleksandra V. Koroleva, Vitaly S. Sobolev and Petr M. Korusenko
Int. J. Mol. Sci. 2024, 25(23), 12818; https://doi.org/10.3390/ijms252312818 - 28 Nov 2024
Viewed by 462
Abstract
This paper presents the results of a study on the characteristics of semiconductor sensors based on thin SnO2 films modified with antimony, dysprosium, and silver impurities and dispersed double Pt/Pd catalysts deposited on the surface to detect carbon monoxide (CO). An original [...] Read more.
This paper presents the results of a study on the characteristics of semiconductor sensors based on thin SnO2 films modified with antimony, dysprosium, and silver impurities and dispersed double Pt/Pd catalysts deposited on the surface to detect carbon monoxide (CO). An original technology was developed, and ceramic targets were made from powders of Sn-Sb-O, Sn–Sb-Dy–O, and Sn–Sb-Dy-Ag–O systems synthesized by the sol–gel method. Films of complex composition were obtained by RF magnetron sputtering of the corresponding targets, followed by technological annealing at various temperatures. The morphology of the films, the elemental and chemical composition, and the electrical and gas-sensitive properties were studied. Special attention was paid to the effect of the film composition on the stability of sensor parameters during long-term tests under the influence of CO. It was found that different combinations of concentrations of antimony, dysprosium, and silver had a significant effect on the size and distribution of nanocrystallites, the porosity, and the defects of films. The mechanisms of degradation under prolonged exposure to CO were examined. It was established that Pt/Pd/SnO2:0.5 at.% Sb film with optimal crystallite sizes and reduced porosity provided increased stability of carbon monoxide sensor parameters, and the response to the action of 100 ppm carbon monoxide was G1/G0 = 2–2.5. Full article
Show Figures

Figure 1

Figure 1
<p>AFM images of samples: series (I)-693K (<b>a</b>), series (II)-693K (<b>b</b>), (III)-723K (<b>c</b>), series (IV)-693K (<b>d</b>), and (V)-723K (<b>e</b>).</p>
Full article ">Figure 2
<p>Survey PE spectra of samples: (1)—(I)-693K, (2)—(II)-693K, (3)—(III)-723K, (4)—(IV)-693K, (5)—(V)-723K, and reference SnO<sub>2</sub>.</p>
Full article ">Figure 3
<p>Sn 3<span class="html-italic">d</span> (<b>a</b>), O 1<span class="html-italic">s</span> with Sb 3<span class="html-italic">d</span><sub>3/2</sub> (<b>b</b>), Dy 3<span class="html-italic">d</span> (<b>c</b>), and Ag 3<span class="html-italic">d</span> (<b>d</b>) PE spectra of samples: (1)—(I)-693K, (2)—(II)-693K, (3)—(III)-723K, (4)—(IV)-693K, (5)—(V)-723K, and reference compounds (SnO<sub>2</sub>, Sb<sub>2</sub>O<sub>5</sub>, and Ag<sup>0</sup>).</p>
Full article ">Figure 4
<p>Raman spectra of SnO<sub>2</sub> powder as well as samples (I)-693K, (II)-693K, and (IV)-693K before and after long-term (90 days) testing under CO exposure (designated by the number 1 superscript in the sample name).</p>
Full article ">Figure 5
<p>Graphs of conductivity versus (<b>a</b>) CO concentration and (<b>b</b>) response of freshly prepared sensors of series: (1)—(I)-693K, (2)—(II)-693K, (4)—(IV)-693K, and (5)—(V)-723K.</p>
Full article ">Figure 6
<p>Concentration dependences of the response of freshly prepared sensors (curves 1) and sensors after long-term (90 days) testing (curves 2). Films from different series are presented: (<b>a</b>)—(I)-693K, (<b>b</b>)—(II)-693K, (<b>c</b>)—(IV)-693K, and (<b>d</b>)—(V)-723K.</p>
Full article ">Figure 7
<p>SEM images obtained in the back-scattering (BSE) mode: the sensitive element from the side of (<b>a</b>) semiconductor SnO<sub>2</sub> layer and (<b>b</b>) heater; (<b>c</b>) sensors assembled into TO-8 case: 1—sensitive element; 2—Pt electrodes; 3—sapphire substrate; 4—Pt heater.</p>
Full article ">Figure 8
<p>Schematic diagram of the measuring stand.</p>
Full article ">
24 pages, 5398 KiB  
Article
A Nitrogen- and Carbon-Present Tin Dioxide-Supported Palladium Composite Catalyst (Pd/N-C-SnO2)
by Keqiang Ding, Weijia Li, Mengjiao Li, Mengyao Di, Ying Bai, Xiaoxuan Liang and Hui Wang
Electrochem 2024, 5(4), 482-505; https://doi.org/10.3390/electrochem5040032 - 13 Nov 2024
Viewed by 913
Abstract
For the first time, nitrogen- and carbon-present tin dioxide-supported palladium composite catalysts (denoted as Pd/N-C-SnO2) were prepared via an HCH method (HCH is the abbreviation for the hydrothermal process–calcination–hydrothermal process preparation process). In this work, firstly, three catalyst carriers (denoted as [...] Read more.
For the first time, nitrogen- and carbon-present tin dioxide-supported palladium composite catalysts (denoted as Pd/N-C-SnO2) were prepared via an HCH method (HCH is the abbreviation for the hydrothermal process–calcination–hydrothermal process preparation process). In this work, firstly, three catalyst carriers (denoted as cc) were prepared using a hydrothermal-process-aided calcination method, and catalyst carriers prepared using ammonia monohydrate (NH3∙H2O), N,N-dimethylformamide (C3H7NO) and triethanolamine (C6H15NO3) as the nitrogen sources were nominated as cc1, cc2 and cc3, respectively. Secondly, these catalyst carriers were reacted with palladium oxide monohydrate (PdO·H2O) hydrothermally to generate catalysts c1, c2 and c3. As testified by XRD and XPS, besides carbon materials and the N-containing substances, the main substances of all prepared catalysts were SnO2 and metallic palladium (Pd). Above all things, all resultant catalysts, especially c2, showed a prominent electrocatalytic activity towards the ethanol oxidation reaction (EOR). As indicated by the CV (cyclic voltammetry) results, all fabricated catalysts presented a clear electrocatalytic activity towards the EOR. In the CA (chronoamperometry) measurement, the faradaic current density of EOR measured on c2 at −0.27 V vs. an SCE (saturated calomel electrode) after 7200 s was still maintained at about 5.6 mA cm−2. Preparing a novel catalyst carrier, N-C-SnO2, and preparing a new EOR catalyst, Pd/N-C-SnO2, were the principal dedications of this preliminary work, which was very beneficial to the development of Pd-based EOR catalysts. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns of all catalyst carriers in which the standard XRD patterns for both SnO<sub>2</sub> and metallic Sn are also presented. Patterns cc<sub>1</sub>, cc<sub>2</sub> and cc<sub>3</sub> correspond to catalyst carriers cc<sub>1</sub>, cc<sub>2</sub> and cc<sub>3</sub>, respectively. (<b>b</b>) XRD patterns for three prepared catalysts, where the standard XRD patterns for SnO<sub>2</sub> and metallic Pd are also illustrated. Patterns c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> correspond to catalysts c<sub>1</sub>, c<sub>2</sub> and c<sub>33</sub>, respectively.</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>) XRD patterns of all catalyst carriers in which the standard XRD patterns for both SnO<sub>2</sub> and metallic Sn are also presented. Patterns cc<sub>1</sub>, cc<sub>2</sub> and cc<sub>3</sub> correspond to catalyst carriers cc<sub>1</sub>, cc<sub>2</sub> and cc<sub>3</sub>, respectively. (<b>b</b>) XRD patterns for three prepared catalysts, where the standard XRD patterns for SnO<sub>2</sub> and metallic Pd are also illustrated. Patterns c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> correspond to catalysts c<sub>1</sub>, c<sub>2</sub> and c<sub>33</sub>, respectively.</p>
Full article ">Figure 2
<p>(<b>a</b>) EDS patterns of all prepared catalysts. Patterns c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> correspond to catalyst c<sub>1</sub>, c<sub>2</sub> and c<sub>33</sub>, respectively. (<b>b</b>) EDS mappings of C, N, O, Pd and Sn for catalyst c<sub>2</sub>.</p>
Full article ">Figure 3
<p>(<b>a</b>) FTIR spectra of all resultant catalysts. (<b>b</b>) Raman spectra for all prepared catalysts. In (<b>a</b>,<b>b</b>), patterns c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> correspond to catalysts c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>, respectively.</p>
Full article ">Figure 4
<p>XPS spectra for the resultant catalysts. (<b>a</b>) Wide scan XPS survey spectra; (<b>b</b>) high resolution XPS spectra of C 1s; (<b>c</b>) high resolution XPS spectra of O 1s; (<b>d</b>) high resolution XPS spectra of Pd 3d; (<b>e</b>) high resolution XPS spectra of Sn 3d; (<b>f</b>) high resolution XPS spectra of N 1s. In above figures, pattern c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> were the patterns of catalyst c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>, respectively.</p>
Full article ">Figure 4 Cont.
<p>XPS spectra for the resultant catalysts. (<b>a</b>) Wide scan XPS survey spectra; (<b>b</b>) high resolution XPS spectra of C 1s; (<b>c</b>) high resolution XPS spectra of O 1s; (<b>d</b>) high resolution XPS spectra of Pd 3d; (<b>e</b>) high resolution XPS spectra of Sn 3d; (<b>f</b>) high resolution XPS spectra of N 1s. In above figures, pattern c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> were the patterns of catalyst c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>, respectively.</p>
Full article ">Figure 4 Cont.
<p>XPS spectra for the resultant catalysts. (<b>a</b>) Wide scan XPS survey spectra; (<b>b</b>) high resolution XPS spectra of C 1s; (<b>c</b>) high resolution XPS spectra of O 1s; (<b>d</b>) high resolution XPS spectra of Pd 3d; (<b>e</b>) high resolution XPS spectra of Sn 3d; (<b>f</b>) high resolution XPS spectra of N 1s. In above figures, pattern c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> were the patterns of catalyst c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>, respectively.</p>
Full article ">Figure 5
<p>SEM images for three prepared catalysts. The scale bars of SEM images were respectively 2 μm and 200 nm. Image c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> were the images of catalyst c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>.</p>
Full article ">Figure 6
<p>CV (cyclic voltammetry) plots recorded on all prepared electrodes at 50 mV s<sup>−1</sup>, in which the black and red curve were respectively recorded in 1M KOH and a solution of 1M KOH + 1M ethanol. (<b>a</b>–<b>c</b>) were respectively measured using electrode c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>. Above CV curves measured on various electrodes at 50 mV s<sup>−1</sup> in a solution of 1M KOH + 1M ethanol are collected in (<b>d</b>).</p>
Full article ">Figure 6 Cont.
<p>CV (cyclic voltammetry) plots recorded on all prepared electrodes at 50 mV s<sup>−1</sup>, in which the black and red curve were respectively recorded in 1M KOH and a solution of 1M KOH + 1M ethanol. (<b>a</b>–<b>c</b>) were respectively measured using electrode c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>. Above CV curves measured on various electrodes at 50 mV s<sup>−1</sup> in a solution of 1M KOH + 1M ethanol are collected in (<b>d</b>).</p>
Full article ">Figure 7
<p>Successive CV curves measured at a scan rate of 50 mV s<sup>−1</sup> in a solution of 1M KOH + 1M ethanol, in which (<b>a</b>–<b>c</b>) were respectively plotted employing electrode c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>, and in each figure, CV curves were successively recorded at the 1st, 5th, 10th, 15th, 20th and 25th cycles. For all above successive CV curves, the relationship between the peak f current density and the cycling number was again summarized in (<b>d</b>), in which curve c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> were respectively recorded using electrode c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>.</p>
Full article ">Figure 7 Cont.
<p>Successive CV curves measured at a scan rate of 50 mV s<sup>−1</sup> in a solution of 1M KOH + 1M ethanol, in which (<b>a</b>–<b>c</b>) were respectively plotted employing electrode c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>, and in each figure, CV curves were successively recorded at the 1st, 5th, 10th, 15th, 20th and 25th cycles. For all above successive CV curves, the relationship between the peak f current density and the cycling number was again summarized in (<b>d</b>), in which curve c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> were respectively recorded using electrode c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>.</p>
Full article ">Figure 8
<p>The CA (Chronoamperometry) curves recorded in a solution of 1M KOH + 1M ethanol, the working potential was fixed at −0.27 V vs. SCE. Curve c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> were respectively recorded using electrode c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>.</p>
Full article ">Figure 9
<p>CV curves of various catalyst-coated GC electrodes measured in 1M KOH at 50 mV s<sup>−1</sup>. Curves c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> correspond to the cases of using electrodes c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>, respectively.</p>
Full article ">Figure 10
<p>EIS (electrochemical impedance spectroscopy) measurement results for various catalyst-coated GC electrodes measured in an electrolyte solution containing 1M KOH and 1M ethanol. In each figure, curves c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> correspond to the cases of using electrodes c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>, respectively. (<b>a</b>) Nyquist plots; (<b>b</b>) Bode plots; (<b>c</b>) curves showing the relationship between the total impedance (Z) and the applied frequency.</p>
Full article ">Figure 10 Cont.
<p>EIS (electrochemical impedance spectroscopy) measurement results for various catalyst-coated GC electrodes measured in an electrolyte solution containing 1M KOH and 1M ethanol. In each figure, curves c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub> correspond to the cases of using electrodes c<sub>1</sub>, c<sub>2</sub> and c<sub>3</sub>, respectively. (<b>a</b>) Nyquist plots; (<b>b</b>) Bode plots; (<b>c</b>) curves showing the relationship between the total impedance (Z) and the applied frequency.</p>
Full article ">
9 pages, 3435 KiB  
Article
The Synthesis of Materials with a Hierarchical Structure Based on Tin Dioxide
by Ekaterina Bondar, Elena Dmitriyeva, Igor Lebedev, Anastasiya Fedosimova, Aigul Shongalova, Sayora Ibraimova, Ainagul Kemelbekova, Ulzhalgas Issayeva, Bagdat Rakymetov and Bedelbek Nurbaev
Nanomaterials 2024, 14(22), 1813; https://doi.org/10.3390/nano14221813 - 13 Nov 2024
Viewed by 876
Abstract
This article presents the results of the formation of hierarchical micro–nano structures in nanostructured tin dioxide films obtained from the lyophilic film-forming system SnCl4/EtOH/NH4OH. The classification of the shape and size of the synthesized structures, in relation to the [...] Read more.
This article presents the results of the formation of hierarchical micro–nano structures in nanostructured tin dioxide films obtained from the lyophilic film-forming system SnCl4/EtOH/NH4OH. The classification of the shape and size of the synthesized structures, in relation to the pH of the solution, is presented. Measurements were carried out on an X-ray diffractometer to study the crystal structure of the samples analyzed. It was found that SnO2 and NH4Cl crystallites participate in the formation of the synthesized hierarchical structures. It is shown that the mechanism of the formation of hierarchical structures depends on the amount of ammonium hydroxide added. This makes it possible to control the shape and size of the synthesized structures by changing the ratio of precursors. Full article
(This article belongs to the Topic Advances in Functional Thin Films)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The structure of the film obtained from the film-forming system SnCl<sub>4</sub>/EtOH/NH<sub>4</sub>OH via the sol–gel method: (<b>a</b>) without adding NH<sub>4</sub>OH; (<b>b</b>) 0.4 mL NH<sub>4</sub>OH per 100 mL of solution; (<b>c</b>) 0.8 mL NH<sub>4</sub>OH per 100 mL of solution; (<b>d</b>) 1.6 mL NH<sub>4</sub>OH per 100 mL of solution.</p>
Full article ">Figure 2
<p>Raman spectra of the sample: (<b>a</b>) without adding NH<sub>4</sub>OH; (<b>b</b>) 1.6 mL NH<sub>4</sub>OH per 100 mL of solution.</p>
Full article ">Figure 3
<p>X-ray diffraction pattern of the crystal structure of the film obtained from the film-forming system SnCl<sub>4</sub>/EtOH/NH<sub>4</sub>OH by the sol–gel method, measured on a DRON-6 X-ray diffractometer: (<b>a</b>) sample without NH<sub>4</sub>OH; (<b>b</b>) sample with NH<sub>4</sub>OH.</p>
Full article ">Figure 4
<p>Elementary cells of (<b>a</b>) SnO<sub>2</sub> and (<b>b</b>) NH<sub>4</sub>Cl.</p>
Full article ">Figure 5
<p>(<b>a</b>) Julien fractal aggregate; (<b>b</b>) two-dimensional disordered aggregate.</p>
Full article ">Figure 6
<p>The mapping of samples obtained from the film-forming system SnCl<sub>4</sub>/EtOH/NH<sub>4</sub>OH with different additive contents: (<b>a</b>) without additive; (<b>b</b>) 0.4 mL NH<sub>4</sub>OH per 100 mL of solution; (<b>c</b>) 0.8 mL NH<sub>4</sub>OH per 100 mL of solution; (<b>d</b>) 1.6 mL NH<sub>4</sub>OH per 100 mL of solution.</p>
Full article ">
29 pages, 8267 KiB  
Review
A Comparative Review of Graphene and MXene-Based Composites towards Gas Sensing
by Pushpalatha Vijayakumar Vaishag and Jin-Seo Noh
Molecules 2024, 29(19), 4558; https://doi.org/10.3390/molecules29194558 - 25 Sep 2024
Cited by 1 | Viewed by 2319
Abstract
Graphene and MXenes have emerged as promising materials for gas sensing applications due to their unique properties and superior performance. This review focuses on the fabrication techniques, applications, and sensing mechanisms of graphene and MXene-based composites in gas sensing. Gas sensors are crucial [...] Read more.
Graphene and MXenes have emerged as promising materials for gas sensing applications due to their unique properties and superior performance. This review focuses on the fabrication techniques, applications, and sensing mechanisms of graphene and MXene-based composites in gas sensing. Gas sensors are crucial in various fields, including healthcare, environmental monitoring, and industrial safety, for detecting and monitoring gases such as hydrogen sulfide (H2S), nitrogen dioxide (NO2), and ammonia (NH3). Conventional metal oxides like tin oxide (SnO2) and zinc oxide (ZnO) have been widely used, but graphene and MXenes offer enhanced sensitivity, selectivity, and response times. Graphene-based sensors can detect low concentrations of gases like H2S and NH3, while functionalization can improve their gas-specific selectivity. MXenes, a new class of two-dimensional materials, exhibit high electrical conductivity and tunable surface chemistry, making them suitable for selective and sensitive detection of various gases, including VOCs and humidity. Other materials, such as metal-organic frameworks (MOFs) and conducting polymers, have also shown potential in gas sensing applications, which may be doped into graphene and MXene layers to improve the sensitivity of the sensors. Full article
(This article belongs to the Special Issue The Way Forward in MXenes Materials)
Show Figures

Figure 1

Figure 1
<p>(<b>1</b>) A schematic illustrating the fabrication process of an LIG sensor. (<b>a</b>) The mechanism involved in the formation of LIG on the PI substrate. (<b>b</b>) A schematic demonstrating the realization of the LIG sensor using laser patterning followed by an IPL-sintered copper electrical pad on the PI substrate. (<b>2</b>) Morphological analysis of the LIG sensor. (<b>a</b>) Photograph of the fabricated LIG sensor. (<b>b</b>,<b>c</b>) FESEM images of IPL and laser-sintered CuNPs, respectively. (<b>d</b>) FESEM micrograph depicting the interface between IPL-sintered CuNPs and laser-sintered CuNPs. FESEM micrographs of the LIG (<b>e</b>) at lower magnification and (<b>f</b>,<b>g</b>) at higher magnification. (<b>h</b>,<b>i</b>) HR-TEM images of the LIG. (<b>j</b>) SAED pattern were obtained from the LIG. FESEM micrographs illustrating the interface between LIG and copper electrode (<b>k</b>) at lower magnification and (<b>l</b>) at higher magnification. Reproduced from Paeng, C.; Shanmugasundaram, A.; We, G.; Kim, T.; Park, J.; Lee, D.W.; Yim, C. Rapid and Flexible Humidity Sensor Based on Laser-Induced Graphene for Monitoring Human Respiration. <span class="html-italic">ACS Appl. Nano Mater.</span> <b>2024</b>, <span class="html-italic">7</span>, 4772–4783 [<a href="#B48-molecules-29-04558" class="html-bibr">48</a>].</p>
Full article ">Figure 2
<p>(<b>1</b>) Left panel: Schematic representation of the steps undertaken for the sensors’ preparation and treatment. Right panel: (<b>a</b>) Scheme of the system employed for NH sensing measurements; (<b>b</b>) a drawing that symbolizes the sensor’s connection to the Keithley multimeter; (<b>c</b>) digital image of the sensor with magnified sections featuring SEM images of the drop-cast GO suspension, providing visual evidence of the even distribution of GO suspension and electrode coverage. (<b>2</b>) SEM micrographs of sensor surface after plasma treatment: (<b>a</b>) 0 s (GO), (<b>b</b>) 10 s (rGO), and (<b>c</b>) 240 s (rGO), pinpointing no apparent damages on the surface. Reproduced from Kurtishaj Hamzaj, A.; Donà, E.; M Santhosh, N.; Shvalya, V.; Košiček, M.; Cvelbar, U. Plasma-Modification of Graphene Oxide for Advanced Ammonia Sensing. <span class="html-italic">Appl. Surf. Sci.</span> <b>2024</b>, <span class="html-italic">660</span>, 1–11 [<a href="#B50-molecules-29-04558" class="html-bibr">50</a>].</p>
Full article ">Figure 3
<p>(<b>1</b>) Schematic procedures for alkalization of Ti<sub>3</sub>C<sub>2</sub>T<span class="html-italic"><sub>x</sub></span> MXene. (<b>2</b>) SEM images of Ti<sub>3</sub>AlC<sub>2</sub> MAX phase, control Ti<sub>3</sub>C<sub>2</sub>T<span class="html-italic"><sub>x</sub></span> MXene, 0.05 M NaOH-treated MXene, and 0.2 M NaOH-treated MXene. Reproduced from Bae, Y.H.; Park, S.; Noh, J.S. Control of Electrical Properties of Ti<sub>3</sub>C<sub>2</sub>T<span class="html-italic"><sub>x</sub></span> Mediated by Facile Alkalization. <span class="html-italic">Surfaces and Interfaces</span> <b>2023</b>, <span class="html-italic">41</span>, 103,258 [<a href="#B55-molecules-29-04558" class="html-bibr">55</a>].</p>
Full article ">Figure 4
<p>Diagrammatic representation of (<b>1</b>) material preparation and (<b>2</b>) zeta potential results of Nb<sub>2</sub>CT<span class="html-italic"><sub>x</sub></span> and Nb<sub>2</sub>CT<span class="html-italic"><sub>x</sub></span>-CTAB together with stability test results of prepared samples. Reproduced from Rathi, K.; Arkoti, N.K.; Pal, K. Fabrication of Delaminated 2D Metal Carbide MXenes (Nb<sub>2</sub>CT<span class="html-italic"><sub>x</sub></span>) by CTAB-Based NO<sub>2</sub> Gas Sensor with Enhanced Stability. <span class="html-italic">Adv. Mater. Interfaces</span> <b>2022</b>, <span class="html-italic">9</span>, 1–10 [<a href="#B56-molecules-29-04558" class="html-bibr">56</a>].</p>
Full article ">Figure 5
<p>(<b>1</b>) Schematic processes for fabricating graphene/ZnO composite sensor on a flexible printed circuit board. (<b>2</b>) Picture of a wearable NFC tag system. Reproduced from Santos-Betancourt, A.; Santos-Ceballos, J.C.; Alouani, M.A.; Malik, S.B.; Romero, A.; Ramírez, J.L.; Vilanova, X.; Llobet, E. ZnO Decorated Graphene-Based NFC Tag for Personal NO<sub>2</sub> Exposure Monitoring during a Workday. <span class="html-italic">Sensors</span> <b>2024</b>, <span class="html-italic">24</span>, 1431 [<a href="#B71-molecules-29-04558" class="html-bibr">71</a>].</p>
Full article ">Figure 6
<p>(<b>1</b>) Transient response curves of (<b>a</b>) pristine ZnO, (<b>b</b>) unirradiated ZnO/(2 wt%) Ti<sub>3</sub>C<sub>2</sub>T<span class="html-italic"><sub>x</sub></span>, and (<b>c</b>) MW-irradiated (5 min) ZnO/(2 wt%) Ti<sub>3</sub>C<sub>2</sub>T<span class="html-italic"><sub>x</sub></span> nanocomposite sensors to various concentration NO<sub>2</sub> gas at 300 °C. (<b>d</b>) Corresponding calibration curves. (<b>2</b>) Dynamic resistance responses of MW-irradiated (5 min) ZnO/(2 wt.%) Ti<sub>3</sub>C<sub>2</sub>T<span class="html-italic"><sub>x</sub></span> nanocomposite gas sensor to NO<sub>2</sub> gas at various temperatures: (<b>a</b>) 150 °C, (<b>b</b>)200 °C, (<b>c</b>) 250 °C, (<b>d</b>) 300 °C, and (<b>e</b>) 350 °C. (<b>f</b>) Corresponding calibration curves at these temperatures. Reproduced from Shin, K.Y.; Mirzaei, A.; Oum, W.; Kim, E.B.; Kim, H.M.; Moon, S.; Kim, S.S.; Kim, H.W. Enhanced NO<sub>2</sub> Gas Response of ZnO/Ti<sub>3</sub>C<sub>2</sub>T<span class="html-italic"><sub>x</sub></span> MXene Nanocomposites by Microwave Irradiation. <span class="html-italic">Sensors Actuators B Chem</span>. <b>2024</b>, <span class="html-italic">409</span>, 135605 [<a href="#B73-molecules-29-04558" class="html-bibr">73</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Optical image showing three 3D MXene/gelatin origami animal models, including rose (<b>a1</b>), crane (<b>a2</b>), and squirrel (<b>a3</b>), respectively. Scale bar: 10 mm. (<b>b</b>) Optical image showing three types of 3D MXene/gelatin origamis prepared by mechanically guided compressive buckling, including an openwork octagonal structure (<b>b1</b>), a hexagonal umbrella (<b>b2</b>), and a six-legged ant (<b>b3</b>), respectively. Scale bar: 10 mm. (<b>c</b>) Detachment test of the 3D origami and the planar pattern on the surface of the shrinking balloons. Scale bar: 10 mm. (<b>d</b>) Sensing response of the proposed 3D MXene/gelatin origami sensor to 50 ppm of NH<sub>3</sub>. Reproduced from Wang, Z.; Yan, F.; Yu, Z.; Cao, H.; Ma, Z.; YeErKenTai, Z.N.; Li, Z.; Han, Y.; Zhu, Z. Fully Transient 3D Origami Paper-Based Ammonia Gas Sensor Obtained by Facile MXene Spray Coating. <span class="html-italic">ACS Sensors</span> <b>2024</b>, <span class="html-italic">9</span>, 1447–1457 [<a href="#B81-molecules-29-04558" class="html-bibr">81</a>].</p>
Full article ">Figure 8
<p>(<b>1</b>) Humidity sensing performance of LIG sensor. (<b>a</b>) Schematic representation of the experimental setup employed for gas sensing analysis. (<b>b</b>,<b>c</b>) Dynamic sensing response of the LIG sensor showcasing the variation in sensor resistance with respect to different RH levels. (<b>d</b>) Sensing response of the LIG sensor in the presence of other interfering gases such as NH<sub>3</sub>, EtOH, CO, SO<sub>2</sub>, and NO<sub>2</sub>. (<b>e</b>) Long-term stability of the LIG sensor (working temperature 25 °C, 67% RH). (<b>2</b>) (<b>a</b>) A schematic of breath sensing analysis. (<b>b</b>) Current versus voltage characteristics of the LIG sensor at three different temperatures. (<b>c</b>–<b>f</b>) Investigation of the LIG sensor’s response behavior under varying respiratory patterns, including slow, normal, fast, and alternating periods of respiration and apnea. Reproduced from Paeng, C.; Shanmugasundaram, A.; We, G.; Kim, T.; Park, J.; Lee, D.W.; Yim, C. Rapid and Flexible Humidity Sensor Based on Laser-Induced Graphene for Monitoring Human Respiration. <span class="html-italic">ACS Appl. Nano Mater.</span> <b>2024</b>, <span class="html-italic">7</span>, 4772–4783 [<a href="#B48-molecules-29-04558" class="html-bibr">48</a>].</p>
Full article ">Figure 9
<p>Schematic illustration of the gas sensing process of rGO/SnO<sub>2</sub> nanocomposites. Reproduced from An, D.; Dai, J.; Zhang, Z.; Wang, Y.; Liu, N.; Zou, Y. rGO/SnO<sub>2</sub> Nanocomposite Based Sensor for Ethanol Detection under Low Temperature. <span class="html-italic">Ceram. Int.</span> <b>2024</b>, <span class="html-italic">50</span>, 16272–16283 [<a href="#B59-molecules-29-04558" class="html-bibr">59</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) Schematic illustration of the sensing mechanism for the WO<sub>3</sub>/Nb<sub>2</sub>CT<span class="html-italic"><sub>x</sub></span> heterojunction sensor. (<b>b</b>) The energy band configuration of WO<sub>3</sub>/Nb<sub>2</sub>CT<span class="html-italic"><sub>x</sub></span> heterojunction (E<sub>F</sub> denotes the Fermi level, and E<sub>C</sub> and E<sub>V</sub> are the conduction band edge and valence band edge, respectively). (<b>c</b>) XPS spectra of O 1s for WO<sub>3</sub> and WO<sub>3</sub>/Nb<sub>2</sub>CT<span class="html-italic"><sub>x</sub></span> heterojunction. (<b>d</b>) EPR spectra of WO<sub>3</sub> and WO<sub>3</sub>/Nb<sub>2</sub>CT<span class="html-italic"><sub>x</sub></span> heterojunction. (<b>e</b>) EIS spectra of WO<sub>3</sub> and WO<sub>3</sub>/Nb<sub>2</sub>CT<span class="html-italic"><sub>x</sub></span> heterojunction. Reproduced from Wang, P.; Guo, S.; Zhao, Y.; Hu, Z.; Tang, Y.; Zhou, L.; Li, T.; Li, H.Y.; Liu, H. WO<sub>3</sub> Nanoparticles Supported by Nb<sub>2</sub>CT<span class="html-italic"><sub>x</sub></span> MXene for Superior Acetone Detection under High Humidity. <span class="html-italic">Sensors Actuators B Chem.</span> <b>2024</b>, <span class="html-italic">398</span>, 134710 [<a href="#B64-molecules-29-04558" class="html-bibr">64</a>].</p>
Full article ">
13 pages, 14616 KiB  
Article
Impedance Spectroscopy Study of Charge Transfer in the Bulk and Across the Interface in Networked SnO2/Ga2O3 Core–Shell Nanobelts in Ambient Air
by Maciej Krawczyk, Ryszard Korbutowicz and Patrycja Suchorska-Woźniak
Sensors 2024, 24(19), 6173; https://doi.org/10.3390/s24196173 - 24 Sep 2024
Viewed by 873
Abstract
Metal oxide core–shell fibrous nanostructures are promising gas-sensitive materials for the detection of a wide variety of both reducing and oxidizing gases. In these structures, two dissimilar materials with different work functions are brought into contact to form a coaxial heterojunction. The influence [...] Read more.
Metal oxide core–shell fibrous nanostructures are promising gas-sensitive materials for the detection of a wide variety of both reducing and oxidizing gases. In these structures, two dissimilar materials with different work functions are brought into contact to form a coaxial heterojunction. The influence of the shell material on the transportation of the electric charge carriers along these structures is still not very well understood. This is due to homo-, hetero- and metal/semiconductor junctions, which make it difficult to investigate the electric charge transfer using direct current methods. However, in order to improve the gas-sensing properties of these complex structures, it is necessary to first establish a good understanding of the electric charge transfer in ambient air. In this article, we present an impedance spectroscopy study of networked SnO2/Ga2O3 core–shell nanobelts in ambient air. Tin dioxide nanobelts were grown directly on interdigitated gold electrodes, using the thermal sublimation method, via the vapor–liquid–solid (VLS) mechanism. Two forms of a gallium oxide shell of varying thickness were prepared via halide vapor-phase epitaxy (HVPE), and the impedance spectra were measured at 189–768 °C. The bulk resistance of the core–shell nanobelts was found to be reduced due to the formation of an electron accumulation layer in the SnO2 core. At temperatures above 530 °C, the thermal reduction of SnO2 and the associated decrease in its work function caused electrons to flow from the accumulation layer into the Ga2O3 shell, which resulted in an increase in bulk resistance. The junction resistance of said core–shell nanostructures was comparable to that of SnO2 nanobelts, as both structures are likely connected through existing SnO2/SnO2 homojunctions comprising thin amorphous layers. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Scanning electron microscope image of SnO<sub>2</sub> nanobelts grown on interdigitated electrodes; and (<b>b</b>,<b>c</b>) enlarged image of the microstructure of SnO<sub>2</sub> nanobelts. The inset shows the edge of the Au electrode before synthesis; and (<b>d</b>) Transmission electron microscope image of SnO<sub>2</sub> nanobelts. The inset in the upper corner shows an enlarged view of the amorphous layer on the surface of the nanobelt. The inset in the lower corner shows the selective area diffraction pattern of the imaged nanobelt.</p>
Full article ">Figure 2
<p>(<b>a</b>) X-ray diffractogram of SnO<sub>2</sub> nanobelts; and (<b>b</b>) chemical composition examined along the length and at the end of the nanobelt shown in <a href="#sensors-24-06173-f001" class="html-fig">Figure 1</a>d.</p>
Full article ">Figure 3
<p>SEM images of the microstructure of SnO<sub>2</sub>/Ga<sub>2</sub>O<sub>3</sub> core–shell nanobelts: (<b>a</b>) CS840; (<b>b</b>) enlarged image of CS840; (<b>c</b>) CS1000; and (<b>d</b>) enlarged image of CS1000.</p>
Full article ">Figure 4
<p>(<b>a</b>) SEM image of the cross-section of two SnO<sub>2</sub>/Ga<sub>2</sub>O<sub>3</sub> core–shell fibrous structures with shells synthesized at 840 °C. In the upper part of the image, the sputtered Pt layer before ion beam etching is visible (for more detail, see <a href="#app1-sensors-24-06173" class="html-app">Figure S1 in the Supplementary Materials</a>), and, in the lower part, the alumina substrate can be seen. (<b>b</b>) Corresponding map of the atomic composition of the cross-section; the edges of the SEM image are overlaid to aid visualization.</p>
Full article ">Figure 5
<p>Complex impedance and Bode plots of the tested structures, as measured at (<b>a</b>,<b>b</b>) 626 °C and (<b>c</b>,<b>d</b>) 768 °C.</p>
Full article ">Figure 6
<p>The electrical equivalent circuit fitted to the impedance spectra. <span class="html-italic">R</span><sub>b</sub> is the bulk resistance, <span class="html-italic">R</span><sub>j</sub> is the resistance of junctions between the nanobelts, and <span class="html-italic">CPE</span><sub>j</sub> is the constant-phase element.</p>
Full article ">Figure 7
<p>Arrhenius plots of (<b>a</b>) the resistance of junctions between the structures; (<b>b</b>) the bulk resistance of SnO<sub>2</sub> nanobelts; (<b>c</b>) the bulk resistance of CS840; and (<b>d</b>) the bulk resistance of CS1000.</p>
Full article ">Figure 8
<p>A schematic depiction of electric charge being transported through adjoining SnO<sub>2</sub> nanobelts. The charge depletion layer is shaded in blue. <span class="html-italic">R</span><sub>e</sub> and <span class="html-italic">C</span><sub>e</sub> are the resistance and capacitance of the electrode/semiconductor junction; <span class="html-italic">R</span><sub>b</sub> is the bulk resistance of the SnO<sub>2</sub> nanobelt; and <span class="html-italic">R</span><sub>j</sub> and <span class="html-italic">C</span><sub>j</sub> are the resistance and capacitance of the junction between the nanobelts.</p>
Full article ">Figure 9
<p>Current and voltage measured at 768 °C.</p>
Full article ">
8 pages, 3228 KiB  
Article
Enhancing Tin Dioxide Anode Performance by Narrowing the Potential Range and Optimizing Electrolytes
by Jose Fernando Florez Gomez, Fernando Camacho Domenech, Songyang Chang, Valerio Dorvilien, Nischal Oli, Brad R. Weiner, Gerardo Morell and Xianyong Wu
Batteries 2024, 10(9), 334; https://doi.org/10.3390/batteries10090334 - 21 Sep 2024
Viewed by 1073
Abstract
Tin dioxide (SnO2) is a low-cost and high-capacity anode material for lithium-ion batteries, but the fast capacity fading significantly limits its practical applications. Current research efforts have focused on preparing sophisticated composite structures or optimizing functional binders, both of which increase [...] Read more.
Tin dioxide (SnO2) is a low-cost and high-capacity anode material for lithium-ion batteries, but the fast capacity fading significantly limits its practical applications. Current research efforts have focused on preparing sophisticated composite structures or optimizing functional binders, both of which increase material manufacturing costs. Herein, we utilize pristine and commercially available SnO2 nanopowders and enhance their cycling performance by simply narrowing the potential range and optimizing electrolytes. Specifically, a narrower potential range (0–1 V) mitigates the capacity fading associated with the conversion reaction, whereas an ether-based electrolyte further suppresses the volume expansion related to the alloy reaction. Consequently, this SnO2 anode delivers a promising battery performance, with a high capacity of ~650 mAhg−1 and stable cycling for 100 cycles. Our work provides an alternative approach to developing high-capacity and long-cycling metal oxide anode materials. Full article
Show Figures

Figure 1

Figure 1
<p>Physical characterizations of the SnO<sub>2</sub> material. (<b>a</b>) XRD pattern; (<b>b</b>–<b>e</b>) SEM images at different magnifications; (<b>f</b>) EDS elemental and mapping analysis.</p>
Full article ">Figure 2
<p>Electrochemical characterizations of SnO<sub>2</sub> in different voltage ranges. (<b>a</b>) CGD curves in the 0–2 V range; (<b>b</b>) GCD curves in the 0–1 V range; (<b>c</b>) the capacity and Coulombic efficiency comparison during cycling; (<b>d</b>) the scheme of the conversion–alloy reaction mechanism. The electrolyte is 1 M LiPF<sub>6</sub>/EC-DEC (1:1 volume ratio, 10% FEC additives).</p>
Full article ">Figure 3
<p>Electrochemical performance comparison of SnO<sub>2</sub> in carbonate and ether electrolytes. (<b>a</b>) CGD curves in the 0–2 V range using the LiPF<sub>6</sub>/THF+m-THF electrolyte; (<b>b</b>) rate performance comparison; (<b>c</b>) the cycling performance comparison; (<b>d</b>) the impedance comparison; (<b>e</b>) the cross-sectional SEM image of the cycled electrode in the ether electrolyte; (<b>f</b>) the cross-sectional SEM image of the cycled electrode in the conventional carbonate electrolyte.</p>
Full article ">
14 pages, 3909 KiB  
Article
Impact of Annealing in Various Atmospheres on Characteristics of Tin-Doped Indium Oxide Layers towards Thermoelectric Applications
by Anna Kaźmierczak-Bałata, Jerzy Bodzenta, Piotr Szperlich, Marcin Jesionek, Anna Michalewicz, Alina Domanowska, Jeyanthinath Mayandi, Vishnukanthan Venkatachalapathy and Andrej Kuznetsov
Materials 2024, 17(18), 4606; https://doi.org/10.3390/ma17184606 - 20 Sep 2024
Viewed by 976
Abstract
The aim of this work was to investigate the possibility of modifying the physical properties of indium tin oxide (ITO) layers by annealing them in different atmospheres and temperatures. Samples were annealed in vacuum, air, oxygen, nitrogen, carbon dioxide and a mixture of [...] Read more.
The aim of this work was to investigate the possibility of modifying the physical properties of indium tin oxide (ITO) layers by annealing them in different atmospheres and temperatures. Samples were annealed in vacuum, air, oxygen, nitrogen, carbon dioxide and a mixture of nitrogen with hydrogen (NHM) at temperatures from 200 °C to 400 °C. Annealing impact on the crystal structure, optical, electrical, thermal and thermoelectric properties was examined. It has been found from XRD measurements that for samples annealed in air, nitrogen and NHM at 400 °C, the In2O3/In4Sn3O12 share ratio decreased, resulting in a significant increase of the In4Sn3O12 phase. The annealing at the highest temperature in air and nitrogen resulted in larger grains and the mean grain size increase, while vacuum, NHM and carbon dioxide atmospheres caused the decrease in the mean grain size. The post-processing in vacuum and oxidizing atmospheres effected in a drop in optical bandgap and poor electrical properties. The carbon dioxide seems to be an optimal atmosphere to obtain good TE generator parameters—high ZT. The general conclusion is that annealing in different atmospheres allows for controlled changes in the structure and physical properties of ITO layers. Full article
Show Figures

Figure 1

Figure 1
<p>XRD spectra (<b>a</b>) recorded for ITO thin films annealed in different atmospheres at 400 °C and (<b>b</b>) deconvoluted (222) peak for reference sample.</p>
Full article ">Figure 2
<p>(<b>a</b>) The main diffraction peak (222) decomposed into two In<sub>4</sub>Sn<sub>3</sub>O<sub>12</sub> to In<sub>2</sub>O<sub>3</sub> components, (<b>b</b>) two components area ratio for ITO layers annealed from RT to 400 °C in various atmospheres, (<b>c</b>) FWHM factor of (222) reflection recorded as a function of annealing temperature: the In<sub>4</sub>Sn<sub>3</sub>O<sub>12</sub> component and (<b>d</b>) the In<sub>2</sub>O<sub>3</sub> component.</p>
Full article ">Figure 3
<p>1 × 1 mm<sup>2</sup> AFM topographic images with 3D view of ITO thin films annealed at 400 °C in various atmospheres and histogram of surface mean grain size analysis.</p>
Full article ">Figure 4
<p>Absorption spectra (<b>a</b>), energy band gap (<b>b</b>) of ITO thin films annealed in different atmospheres at 400 °C.</p>
Full article ">Figure 5
<p>Calibration curve for thermal measurements.</p>
Full article ">Figure 6
<p>Electrical conductivity (<b>a</b>), carrier concentration (<b>b</b>), mobility (<b>c</b>) as a function of the temperature of ITO thin films annealed in different atmospheres.</p>
Full article ">Figure 7
<p>Seebeck coefficient and thermal conductivity (<b>a</b>), Seebeck coefficient and electrical conductivity (<b>b</b>) versus type of annealing atmosphere.</p>
Full article ">Figure 8
<p>Comparison of figure of merit ZT as a function of electrical conductivity in this work with results collected by Kim, S, et al. [<a href="#B13-materials-17-04606" class="html-bibr">13</a>].</p>
Full article ">
15 pages, 3949 KiB  
Article
Highly Efficient Production of Furfural from Corncob by Barley Hull Biochar-Based Solid Acid in Cyclopentyl Methyl Ether–Water System
by Bo Fan, Linghui Kong and Yucai He
Catalysts 2024, 14(9), 583; https://doi.org/10.3390/catal14090583 - 1 Sep 2024
Cited by 1 | Viewed by 1256
Abstract
Furfural, an important biobased compound, can be synthesized through the chemocatalytic conversion of D-xylose and hemicelluloses from lignocellulose. It has widespread applications in the production of valuable furans, additives, resins, rubbers, synthetic fibers, polymers, plastics, biofuels, and pharmaceuticals. By using barley hulls [...] Read more.
Furfural, an important biobased compound, can be synthesized through the chemocatalytic conversion of D-xylose and hemicelluloses from lignocellulose. It has widespread applications in the production of valuable furans, additives, resins, rubbers, synthetic fibers, polymers, plastics, biofuels, and pharmaceuticals. By using barley hulls (BHs) as biobased support, a heterogeneous biochar Sn-NUS-BH catalyst was created to transform corncob into furfural in cyclopentyl methyl ether–H2O. Sn-NUS-BH had a fibrous structure with voids, a large comparative area, and a large pore volume, which resulted in more catalytic active sites. Through the characterization of the physical and chemical properties of Sn-NUS-BH, it was observed that the Sn-NUS-BH had tin dioxide (Lewis acid sites) and a sulfonic acid group (Brønsted acid sites). This chemocatalyst had good thermostability. At 170 °C for 20 min, Sn-NUS-BH (3.6 wt%) was applied to transform 75 g/L of corncob with ZnCl2 (50 mM) to generate furfural (80.5% yield) in cyclopentyl methyl ether–H2O (2:1, v/v). This sustainable catalytic process shows great promise in the transformation of lignocellulose to furfural using biochar-based chemical catalysts. Full article
Show Figures

Figure 1

Figure 1
<p>SEM characterization of NUS-BH (<b>a</b>) and Sn-NUS-BH solid acid catalyst (<b>b</b>).</p>
Full article ">Figure 2
<p>XRD images of NUS-BH and solid acid Sn-NUS-BH.</p>
Full article ">Figure 3
<p>Infrared characterization of NUS-BH and solid acid Sn-NUS-BH.</p>
Full article ">Figure 4
<p>XPS characterization of NUS-BH and solid acid Sn-NUS-BH. [Gray line, sum of all deconvolution curves; light purple lines, background curve].</p>
Full article ">Figure 5
<p>NH<sub>3</sub>-TPD characterization of NUS-BH and Sn-NUS-BH solid acid catalysts.</p>
Full article ">Figure 6
<p>The effect of five organic solvents (CPME, MIBK, THF, ACN, DMF, and DMSO) on furfural yield [corncob 3 g (75 g/L), organic solvent/water = 1:1 (<span class="html-italic">v</span>/<span class="html-italic">v</span>), Sn-NUS-BH 3.6 wt%, 170 °C, 20 min] (<b>a</b>); The relationship between furfural yield and organic solvent log <span class="html-italic">P</span> [corncob 3 g (75 g/L), organic solvent/water = 1:1 (<span class="html-italic">v</span>/<span class="html-italic">v</span>), Sn-NUS-BH 3.6 wt%, 170 °C, 20 min] (<b>b</b>).</p>
Full article ">Figure 7
<p>The effect of the volumetric ratio of CPME-to-water (1:3–3:1, <span class="html-italic">v</span>/<span class="html-italic">v</span>) on the formation of furfural [corncob 3 g (75 g/L), Sn-NUS-BH 3.6 wt%, 170 °C, 20 min].</p>
Full article ">Figure 8
<p>Effect of Sn-NUS-BH (0–6 wt%) on the catalytic conversion of corncob to furfural under 170 °C for 20 min in CPME–water (2:1, <span class="html-italic">v</span>/<span class="html-italic">v</span>) system.</p>
Full article ">Figure 9
<p>Effects of reaction time (5–60 min) on the chemocatalysis of corncob into furfural with Sn-NUS-BH (3.6 wt%) at 170 °C in CPME–water (2:1, <span class="html-italic">v</span>/<span class="html-italic">v</span>) (<b>a</b>); effects of temperature (150–190 °C) on the chemocatalysis of corncob into furfural with Sn-NUS-BH (3.6 wt%) at 170 °C for 20 min in CPME–water (2:1, <span class="html-italic">v</span>/<span class="html-italic">v</span>) (<b>b</b>).</p>
Full article ">Figure 10
<p>Effects of chloride salts (50 mM) on the chemocatalysis of corncob into furfural with Sn-NUS-BH (3.6 wt%) at 170 °C for 20 min in CPME–water (2:1, <span class="html-italic">v</span>/<span class="html-italic">v</span>) (<b>a</b>); effects of ZnCl<sub>2</sub> dosage (0–500 mM) on the chemocatalysis of corncob into furfural with Sn-NUS-BH (3.6 wt%) at 170 °C for 20 min in CPME–water (2:1, <span class="html-italic">v</span>/<span class="html-italic">v</span>) (<b>b</b>).</p>
Full article ">Figure 10 Cont.
<p>Effects of chloride salts (50 mM) on the chemocatalysis of corncob into furfural with Sn-NUS-BH (3.6 wt%) at 170 °C for 20 min in CPME–water (2:1, <span class="html-italic">v</span>/<span class="html-italic">v</span>) (<b>a</b>); effects of ZnCl<sub>2</sub> dosage (0–500 mM) on the chemocatalysis of corncob into furfural with Sn-NUS-BH (3.6 wt%) at 170 °C for 20 min in CPME–water (2:1, <span class="html-italic">v</span>/<span class="html-italic">v</span>) (<b>b</b>).</p>
Full article ">Figure 11
<p>Recycling of Sn-NUS-BH for converting corncob into furfural at 170 °C for 20 min in CPME–water (2:1, <span class="html-italic">v</span>/<span class="html-italic">v</span>) containing ZnCl<sub>2</sub> (50 mM).</p>
Full article ">
Back to TopTop