[go: up one dir, main page]

 
 
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

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
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,104)

Search Parameters:
Keywords = carbon storage capacity

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 24084 KiB  
Article
Upcycling Waste Polyethylene Terephthalate to Produce Nitrogen-Doped Porous Carbon for Enhanced Capacitive Deionization
by Hui Yu, Haiyan Duan, Liang Chen, Weihua Zhu, Daria Baranowska, Yumeng Hua, Dengsong Zhang and Xuecheng Chen
Molecules 2024, 29(20), 4934; https://doi.org/10.3390/molecules29204934 (registering DOI) - 18 Oct 2024
Abstract
Porous carbon with a high surface area and controllable pore size is needed for energy storage. It is still a significant challenge to produce porous carbon in an economical way. Nitrogen-doped porous carbon (N-PC) was prepared through carbonization of a mixture of waste [...] Read more.
Porous carbon with a high surface area and controllable pore size is needed for energy storage. It is still a significant challenge to produce porous carbon in an economical way. Nitrogen-doped porous carbon (N-PC) was prepared through carbonization of a mixture of waste PET-derived metal–organic frameworks (MOFs) and ammonium. The obtained N-PC exhibits a large surface area and controlled pore size. When utilized as an electrode material for supercapacitors, the N-PC exhibits a specific capacitance of 224 F g−1, significantly surpassing that of commercial activated carbon (AC), which has a capacitance of 111 F g−1. In the subsequent capacitive deionization (CDI) tests, the N-PC demonstrated a maximum salt adsorption capacity of 19.9 mg g−1 at 1.2 V in a NaCl electrolyte (0.5 g L−1), and the salt adsorption capacity increased to 24.7 mg g−1 at 1.4 V. The N-PC electrode also exhibited superior regeneration. The present work not only presents a potential approach to develop cost-effective electrodes for seawater purification but also paves the way for recycling of waste plastics into high value-added products. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic illustration of the synthesis of N-PC from waste PET and application in desalination. (<b>b</b>,<b>c</b>) TEM and (<b>d</b>) SEM images of waste PET-derived MOF-Al. (<b>e</b>) TEM image of the mixture of PET-derived MOF-Al and urea and (<b>f</b>) N-PC. (<b>g</b>) STEM image and elemental maps for carbon, oxygen, and nitrogen in the N-PC.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD pattern (inset: TGA curve), (<b>b</b>) Raman scattering spectrum, (<b>c</b>) N<sub>2</sub> adsorption/desorption isotherms (inset: pore size distribution), (<b>d</b>) XPS survey spectrum of the N-PC.</p>
Full article ">Figure 3
<p>(<b>a</b>) Cyclic voltammetry (CV) curves of AC and N-PC at a scan rate of 1 mV s<sup>−1</sup>; (<b>b</b>) CV curves of the N-PC electrode over a scanning rate range of 5 to 50 mV/s; (<b>c</b>) specific capacitances of the N-PC determined from CV curves at various scanning rates; (<b>d</b>) electrochemical impedance spectroscopy (EIS) profiles of N-PC and AC, depicted as Nyquist plots, with the inset illustrating the magnified high-frequency region. The data were collected from a 0.5 M aqueous NaCl solution.</p>
Full article ">Figure 4
<p>(<b>a</b>) Charge/discharge curves at a current density of 0.2 A g<sup>−1</sup>; (<b>b</b>) charge/discharge curves across a range of current densities from 0.2 to 1 A g<sup>−1</sup>; (<b>c</b>) cycle performance of the N-PC electrode at a current density of 0.4 A g<sup>−1</sup> (showing the first and last five cycles). All tests were conducted in a 0.5 M NaCl solution.</p>
Full article ">Figure 5
<p>(<b>a</b>) SAC vs. deionization time for the N-PC and AC electrodes, (<b>b</b>) current vs. time and SAC vs. time, and (<b>c</b>) Ragone plots (SAR vs. SAC) for the N-PC electrode in a 500 mg L<sup>−1</sup> NaCl solution at 1.2 V.</p>
Full article ">Figure 6
<p>(<b>a</b>) Graphs depicting the relationship between SAC and deionization time, and (<b>b</b>) Ragone plots illustrating the specific areal rate (SAR) versus SAC for N-PC electrodes at varying NaCl concentrations; (<b>c</b>) Graphs showing the SAC versus deionization time, and (<b>d</b>) Ragone plots depicting the SAR versus SAC for N-PC electrodes at different voltages.</p>
Full article ">Figure 7
<p>Cycling deionization–regeneration curves for PET-based N-PC (NaCl concentration: 100 mg L<sup>−1</sup>, voltage: 1.2 V).</p>
Full article ">
18 pages, 3192 KiB  
Article
The Influence of Thermal Treatment of Activated Carbon on Its Electrochemical, Corrosion, and Adsorption Characteristics
by Andrzej Świątkowski, Elżbieta Kuśmierek, Krzysztof Kuśmierek and Stanisław Błażewicz
Molecules 2024, 29(20), 4930; https://doi.org/10.3390/molecules29204930 (registering DOI) - 18 Oct 2024
Abstract
Activated carbons can be applied in various areas of our daily life depending on their properties. This study was conducted to investigate the effect of thermal treatment of activated carbon on its properties, considering its future use. The characteristics of activated carbon heat-treated [...] Read more.
Activated carbons can be applied in various areas of our daily life depending on their properties. This study was conducted to investigate the effect of thermal treatment of activated carbon on its properties, considering its future use. The characteristics of activated carbon heat-treated at temperatures of 1500, 1800, and 2100 °C based on its future use are presented. The significant effect of the treatment temperature on morphological, adsorption, electrochemical, and corrosion properties was proved. Increasing the temperature above 1800 °C resulted in a significant decrease in the specific surface area (from 969 to 8 m2·g−1) and material porosity—the formation of mesopores (20–100 nm diameter) was observed. Simultaneously, adsorption capability, double layer capacity, and electrochemically active surface area also decreased, which helped to explain the shape of cyclic voltammograms recorded in 2,4-dichlorophenoxyacetic acid and in supporting electrolytes. However, a significant increase in corrosion resistance was found for the carbon material treated at a temperature of 2100 °C (corrosion current decreased by 23 times). Comparison of morphological, adsorption, corrosion, and electrochemical characteristics of the tested activated carbon, its applicability as an electrode material in electrical energy storage devices, and materials for adsorptive removal of organic compounds from wastewater or as a sensor in electrochemical determination of organic compounds was discussed. Full article
(This article belongs to the Special Issue Exclusive Feature Papers in Electrochemistry)
Show Figures

Figure 1

Figure 1
<p>SEM images of the activated carbon samples (AC1500, AC1800 and AC2100) recorded at magnifications of 500 (<b>top</b>) and 2000 (<b>bottom</b>).</p>
Full article ">Figure 2
<p>Nitrogen adsorption–desorption isotherms (<b>left side</b>) and pore size distribution dV/dD plots (<b>right side</b>) determined for the activated carbon samples.</p>
Full article ">Figure 3
<p>Adsorption isotherms of 2,4-D from 0.1 mol·L<sup>−1</sup> KCl onto activated carbons (line: fitting of Langmuir model). Experimental conditions: 2,4-D initial concentrations = 0.3–1.0 mmol·L<sup>−1</sup>, activated carbon dosage = 0.5 g·L<sup>−1</sup>, temperature = 23 °C, pH = native (original).</p>
Full article ">Figure 4
<p>Potentiodynamic polarization curves recorded for the activated carbon samples in KCl solution (0.1·mol L<sup>−1</sup>).</p>
Full article ">Figure 5
<p>Exemplary potentiodynamic polarization curves recorded for AC1800 material in 0.1 mol·L<sup>−1</sup> KCl in three consecutive measurements.</p>
Full article ">Figure 6
<p>Cyclic voltammograms recorded on activated carbon samples in a 0.001 mol·L<sup>−1</sup> 2,4-D solution (0.1 mol·L<sup>−1</sup> KCl); v = 20 mV·s<sup>−1</sup>, AC1500 and AC1800 (left axis); AC2100 (right axis).</p>
Full article ">Figure 7
<p>Cyclic voltammograms recorded in 0.1 mol·L<sup>−1</sup> KCl for AC1500 (left axis), AC1800 (left axis), and AC2100 (right axis) carbon materials; v = 10 mV·s<sup>−1</sup>.</p>
Full article ">Figure 8
<p>Dependences <span class="html-italic">1/q*</span> vs. <span class="html-italic">v</span><sup>1/2</sup> (<b>A</b>) and <span class="html-italic">q*</span> vs. <span class="html-italic">v</span><sup>−1/2</sup> (<b>B</b>) determined for the activated carbons in KCl solution. AC1500 and AC1800—left axis, AC2100—right axis.</p>
Full article ">
13 pages, 2832 KiB  
Article
High-Performance Dual-Redox-Mediator Supercapacitors Based on Buckypaper Electrodes and Hydrogel Polymer Electrolytes
by Garbas A. Santos Junior, Kélrie H. A. Mendes, Sarah G. G. de Oliveira, Gabriel J. P. Tonon, Neide P. G. Lopes, Thiago H. R. da Cunha, Mario Guimarães Junior, Rodrigo L. Lavall and Paulo F. R. Ortega
Polymers 2024, 16(20), 2903; https://doi.org/10.3390/polym16202903 (registering DOI) - 15 Oct 2024
Viewed by 501
Abstract
In recent years, the demand for solid, thin, and flexible energy storage devices has surged in modern consumer electronics, which require autonomy and long duration. In this context, hybrid supercapacitors have become strategic, and significant efforts are being made to develop cells with [...] Read more.
In recent years, the demand for solid, thin, and flexible energy storage devices has surged in modern consumer electronics, which require autonomy and long duration. In this context, hybrid supercapacitors have become strategic, and significant efforts are being made to develop cells with higher energy densities while preserving the power density of conventional supercapacitors. Motivated by these requirements, we report the development of a new high-performance dual-redox-mediator supercapacitor. In this study, cells were constructed using fully moldable buckypapers (BPs), composed of carbon nanotubes and cellulose nanofibers, as electrodes. We evaluated the compatibility of BPs with hydrogel polymer electrolytes, based on 1 mol L−1 H2SO4 and polyvinyl alcohol (PVA), supplemented with different redox species: methylene blue, indigo carmine, and hydroquinone. Solid cells were constructed containing two active redox species to maximize the specific capacity of each electrode. Considering the main results, the dual-redox-mediator supercapacitor exhibits high energy density of 32.0 Wh kg−1 (at 0.8 kW kg−1) and is capable of delivering 25.9 Wh kg−1 at high power demand (4.0 kW kg−1). Stability studies conducted over 10,000 galvanostatic cycles revealed that the PVA polymer matrix benefits the system by inhibiting the crossover of redox species within the cell. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic representation and (<b>b</b>) image of the obtained CNT/CNF buckypaper. (<b>c</b>) Schematic of the HGPE buckypaper preparation process and (<b>d</b>) photography of the HGPE buckypaper. (<b>e</b>) Schematic representation of the cell configuration and (<b>f</b>) photography of the flexible solid-state supercapacitor device. Schematic representation of the fabricated symmetric SC device assembled using 0.8-HGPE buckypaper and redox mediator: (<b>g</b>) carmine indigo, (<b>h</b>) methylene blue, (<b>i</b>) hydroquinone.</p>
Full article ">Figure 2
<p>(<b>a</b>) Nyquist plot—inset shows the equivalent circuit; (<b>b</b>) GCD curves at J = 1 A g<sup>−1</sup>; (<b>c</b>) specific capacity of SCs based on HGPE electrolyte compared to the liquid electrolyte system at 1 A g<sup>−1</sup> (3.54 mA cm<sup>−2</sup>).</p>
Full article ">Figure 3
<p>Cyclic voltammetry for cells constructed with (<b>a</b>) 0.8-HGPE—the inset shows the magnified cyclic voltammetry—and 0.8-HGPE containing (<b>b</b>) methylene blue, (<b>c</b>) hydroquinone, and (<b>d</b>) indigo carmine. The inset also shows the dependence of peak currents on the square root of the scan rate for both anodic and cathodic potentials. The anodic peak current is represented by empty squares, while the cathodic peak current is represented by filled squares.</p>
Full article ">Figure 4
<p>Dual-redox-mediator solid-state SC: (<b>a</b>) schematic configuration; (<b>b</b>) GCD curves at different current densities, inset compares the system with 0.8-HGPE SC, at 1 A g<sup>−1</sup>; (<b>c</b>) cyclic voltammetry at different scan rates, inset compares the system with 0.8-HGPE SC, at 100 mV s<sup>−1</sup>; (<b>d</b>) specific capacity compared to 0.8-HGPE SC.</p>
Full article ">Figure 5
<p>Dual-redox-mediator solid-state SC: (<b>a</b>) evolution of the potential of the electrodes at different current densities (red line, positive electrode (methylene blue redox mediator); blue line, negative electrode (indigo carmine redox mediator)); (<b>b</b>) cycling stability for 10,000 cycles at 2.5 A g<sup>−1</sup> (8.84 mA cm<sup>−2</sup>) and coulombic efficiency; (<b>c</b>) comparison of the potential evolution of the electrodes at the 1st and the 10,000th cycle. (<b>d</b>) Ragone plot of the dual-redox-mediator solid-state SC, compared with some previously published systems [<a href="#B19-polymers-16-02903" class="html-bibr">19</a>,<a href="#B25-polymers-16-02903" class="html-bibr">25</a>,<a href="#B26-polymers-16-02903" class="html-bibr">26</a>,<a href="#B27-polymers-16-02903" class="html-bibr">27</a>,<a href="#B28-polymers-16-02903" class="html-bibr">28</a>,<a href="#B29-polymers-16-02903" class="html-bibr">29</a>,<a href="#B30-polymers-16-02903" class="html-bibr">30</a>,<a href="#B31-polymers-16-02903" class="html-bibr">31</a>].</p>
Full article ">
15 pages, 5215 KiB  
Article
Molecular Insights into CO2 Diffusion Behavior in Crude Oil
by Chunning Gao, Yongqiang Zhang, Wei Fan, Dezhao Chen, Keqin Wu, Shuai Pan, Yuchuan Guo, Haizhu Wang and Keliu Wu
Processes 2024, 12(10), 2248; https://doi.org/10.3390/pr12102248 (registering DOI) - 15 Oct 2024
Viewed by 413
Abstract
CO2 flooding plays a significant part in enhancing oil recovery and is essential to achieving CCUS (Carbon Capture, Utilization, and Storage). This study aims to understand the fundamental theory of CO2 dissolving and diffusing into crude oil and how these processes [...] Read more.
CO2 flooding plays a significant part in enhancing oil recovery and is essential to achieving CCUS (Carbon Capture, Utilization, and Storage). This study aims to understand the fundamental theory of CO2 dissolving and diffusing into crude oil and how these processes vary under reasonable reservoir conditions. In this paper, we primarily use molecular dynamics simulation to construct a multi-component crude oil model with 17 hydrocarbons, which is on the basis of a component analysis of oil samples through laboratory experiments. Then, the CO2 dissolving capacity of the multi-component crude was quantitatively characterized and the impacts of external conditions—including temperature and pressure—on the motion of the CO2 dissolution and diffusion coefficients were systematically investigated. Finally, the swelling behavior of mixed CO2–crude oil was analyzed and the diffusion coefficients were predicted; furthermore, the levels of CO2 impacting the oil’s mobility were analyzed. Results showed that temperature stimulation intensified molecular thermal motion and increased the voids between the alkane molecules, promoting the rapid dissolution and diffusion of CO2. This caused the crude oil to swell and reduced its viscosity, further improving the mobility of the crude oil. As the pressure increased, the voids between the internal and external potential energy of the crude oil models became wider, facilitating the dissolution of CO2. However, when subjected to external compression, the CO2 molecules’ diffusing progress within the oil samples was significantly limited, even diverging to zero, which inhabited the improvement in oil mobility. This study provides some meaningful insights into the effect of CO2 on improving molecular-scale mobility, providing theoretical guidance for subsequent investigations into CO2–crude oil mixtures’ complicated and detailed behavior. Full article
(This article belongs to the Special Issue Advances in Enhancing Unconventional Oil/Gas Recovery, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Molecular structure of C<sub>8</sub>H<sub>18</sub> and CO<sub>2</sub>. (<b>b</b>) Schematics of molecular simulation of the multi-component crude oil model. The crude oil model density was set as 0.82 g/cm<sup>3</sup>.</p>
Full article ">Figure 2
<p>Comparison of density values of octane in crude oil models under different conditions.</p>
Full article ">Figure 3
<p>Density distribution of CO<sub>2</sub> molecules dissolved in different crude oil systems at various pressures and temperatures: (<b>a</b>) the system at 313 K and 3 MPa; (<b>b</b>) the system at 313 K and 10 MPa; (<b>c</b>) the system at 318 K and 3 MPa; (<b>d</b>) the system at 318 K and 10 MPa; (<b>e</b>) the system at 323 K and 3 MPa; (<b>f</b>) the system at 323 K and 10 MPa. In the figure, blue represents the boundary of the model, gray indicates the crude oil molecules, and red depicts the dissolved CO<sub>2</sub>.</p>
Full article ">Figure 3 Cont.
<p>Density distribution of CO<sub>2</sub> molecules dissolved in different crude oil systems at various pressures and temperatures: (<b>a</b>) the system at 313 K and 3 MPa; (<b>b</b>) the system at 313 K and 10 MPa; (<b>c</b>) the system at 318 K and 3 MPa; (<b>d</b>) the system at 318 K and 10 MPa; (<b>e</b>) the system at 323 K and 3 MPa; (<b>f</b>) the system at 323 K and 10 MPa. In the figure, blue represents the boundary of the model, gray indicates the crude oil molecules, and red depicts the dissolved CO<sub>2</sub>.</p>
Full article ">Figure 4
<p>Curves of model swelling rates of crude oil systems with temperature and pressure before and after dissolving CO<sub>2</sub>: (<b>a</b>) crude oil systems; (<b>b</b>) CO<sub>2</sub>–crude oil systems.</p>
Full article ">Figure 4 Cont.
<p>Curves of model swelling rates of crude oil systems with temperature and pressure before and after dissolving CO<sub>2</sub>: (<b>a</b>) crude oil systems; (<b>b</b>) CO<sub>2</sub>–crude oil systems.</p>
Full article ">Figure 5
<p>Swellling rate curves of crude oil systems before and after dissolving CO<sub>2</sub> molecules at three temperatures: (<b>a</b>) 313 K; (<b>b</b>) 318 K; (<b>c</b>) 323 K.</p>
Full article ">Figure 5 Cont.
<p>Swellling rate curves of crude oil systems before and after dissolving CO<sub>2</sub> molecules at three temperatures: (<b>a</b>) 313 K; (<b>b</b>) 318 K; (<b>c</b>) 323 K.</p>
Full article ">Figure 6
<p>The dissolved CO<sub>2</sub> molecules in crude oil systems at varying temperatures.</p>
Full article ">Figure 7
<p>Diffusion coefficient curves of CO<sub>2</sub> molecules at various pressures in 313 K, 318 K, and 323 K systems.</p>
Full article ">Figure 8
<p>MSD curves of crude oil systems before and after dissolving CO<sub>2</sub> molecules in 323 K system at different pressures: (<b>a</b>) crude oil systems; (<b>b</b>) CO<sub>2</sub>–crude oil systems.</p>
Full article ">Figure 9
<p>Diffusion coefficient curves of crude oil systems before and after dissolving CO<sub>2</sub> molecules in different systems.</p>
Full article ">
25 pages, 10352 KiB  
Article
Sustainable Logistics: Synergizing Passive Design and PV–Battery Systems for Carbon Footprint Reduction
by Kanwal Yasir, Jingchun Shen and Jing Lin
Buildings 2024, 14(10), 3257; https://doi.org/10.3390/buildings14103257 (registering DOI) - 15 Oct 2024
Viewed by 499
Abstract
As more companies strive for net-zero emissions, mitigating indirect greenhouse gas emissions embedded in value chains—especially in logistics activities—has become a critical priority. In the European logistics sector, sustainability and energy efficiency are receiving growing attention, given the sector’s intersectional role in both [...] Read more.
As more companies strive for net-zero emissions, mitigating indirect greenhouse gas emissions embedded in value chains—especially in logistics activities—has become a critical priority. In the European logistics sector, sustainability and energy efficiency are receiving growing attention, given the sector’s intersectional role in both transportation and construction. This transition toward low-carbon logistics design not only reduces carbon emissions but also yields financial benefits, including operational cost savings and new market opportunities. This study examines the impact of passive design strategies and low-carbon technologies in a Swedish logistics center, assessed using the low-carbon design criteria from the BREEAM International standard, version 6. The findings show that passive energy-efficient measures, such as the installation of 47 skylights for natural daylighting, reduced light power density in accordance with AHSHARE 90.1-2019 and the integration of free night flushing, contribute to a 23% reduction in total energy consumption. In addition, the integration of 600 PV panels and 480 batteries with a capacity of 268 ampere-hours and 13.5 kWh storage, operating at 50 volts, delivers a further 56% reduction in carbon emissions. By optimizing the interaction between passive design and active low-carbon technologies, this research presents a comprehensive feasibility analysis that promotes sustainable logistics practices while ensuring a future-proof, low-carbon operational model. Full article
Show Figures

Figure 1

Figure 1
<p>Illustrated research objectives in relation to the Low-Carbon Design Indicator criteria from BREEAM-Int V.6.</p>
Full article ">Figure 2
<p>Definitions of both passive and active measures described in this article.</p>
Full article ">Figure 3
<p>Project’s site plan and building layout.</p>
Full article ">Figure 4
<p>Monthly diurnal averages both radiation and dry-bulb temperature.</p>
Full article ">Figure 5
<p>Seasonal wind wheels with respect to both temperature and relative humidity.</p>
Full article ">Figure 6
<p>Proposed night ventilation control algorithm using the existing AHU in the IDA ICE 5 schematic interface.</p>
Full article ">Figure 7
<p>Proposed skylight implementation for the logistic center in line with the working principle of utilizing a daylight-controlled lighting solution.</p>
Full article ">Figure 8
<p>Placement of 600 PV panels on the roof, in their position relative to the 47 implemented skylights.</p>
Full article ">Figure 9
<p>Illuminance result comparison in warehouse zone before (<b>left</b>) and after (<b>right</b>) the described optimization.</p>
Full article ">Figure 10
<p>Daylight factor distribution in the studied logistic center.</p>
Full article ">Figure 11
<p>Temperature comparison during summertime, before (<b>left</b>) and after (<b>right</b>) the described optimization.</p>
Full article ">Figure 12
<p>Delivered-energy overview of the case of 600 PV panels.</p>
Full article ">Figure 13
<p>Annual accumulative electrical load curves for the daily storage solution.</p>
Full article ">Figure 14
<p>Annual delivered-energy overview for the daily storage solution.</p>
Full article ">Figure 15
<p>Annual battery-balance overview for the daily storage solution.</p>
Full article ">Figure 16
<p>Annual battery status overview for the daily storage solution.</p>
Full article ">Figure 17
<p>Annual accumulative electrical load curves for the seasonal storage solution.</p>
Full article ">Figure 18
<p>Annual delivered-energy overview for the seasonal storage solution.</p>
Full article ">Figure 19
<p>Annually battery-balance overview for the seasonal storage solution.</p>
Full article ">Figure 20
<p>Annual battery status overview for the seasonal storage solution.</p>
Full article ">Figure 21
<p>Graphical outputs summary from both passive and active efficient measures.</p>
Full article ">Figure A1
<p>Summary of input paraments in the IDA ICE model of the studied logistic center.</p>
Full article ">
12 pages, 4509 KiB  
Article
Effects of Storage Voltage upon Sodium-Ion Batteries
by Tengfei Song, Brij Kishore, Yazid Lakhdar, Lin Chen, Peter R. Slater and Emma Kendrick
Batteries 2024, 10(10), 361; https://doi.org/10.3390/batteries10100361 - 11 Oct 2024
Viewed by 718
Abstract
Sodium-ion batteries (SIBs) are gaining attention as a safer, more cost-effective alternative to lithium-ion batteries (LIBs) due to their use of abundant and non-critical materials. A notable feature of SIBs is their ability to utilize aluminum current collectors, which are resistant to oxidation, [...] Read more.
Sodium-ion batteries (SIBs) are gaining attention as a safer, more cost-effective alternative to lithium-ion batteries (LIBs) due to their use of abundant and non-critical materials. A notable feature of SIBs is their ability to utilize aluminum current collectors, which are resistant to oxidation, allowing for safer storage at 0 V. However, the long-term impacts of such storage on their electrochemical performance remain poorly understood. This study systematically investigates how storage conditions at various states of charge (SOCs) affect open circuit voltage (OCV) decay, internal resistance, and post-storage cycling stability in two different Na-ion chemistries: Prussian white//hard carbon and layered oxide//hard carbon. Electrochemical Impedance Spectroscopy before and after storage shows a pronounced increase in internal resistance and a corresponding decline in cycling performance when SIBs are stored in a fully discharged state (0 V), particularly for layered oxide-based cells, illustrating the sensitivity of different SIB chemistries to storage conditions. Additionally, a novel reformation protocol is proposed that reactivates cell capacity by rebuilding the solid electrolyte interphase (SEI) layer, offering a recovery path after prolonged storage. These insights into the long-term storage effects on SIBs provide new guidelines for optimizing storage and transport conditions to minimize performance degradation, making them more viable for commercial applications. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of storage conditions on the OCV change of PW//HC cells. (<b>a</b>) The voltage profile of PW//HC cells for the formation process in a voltage window of 1.3–3.8 V, and (<b>b</b>) the voltage profiles for charging or discharging the cells to various SOCs. (<b>c</b>) The change in OCV of the cells over 60 days, and (<b>d</b>) the change in the OCV plotted as a percentage change.</p>
Full article ">Figure 2
<p>Effect of storage conditions on the post-cycling performance of PW//HC cells. (<b>a</b>) Cycling data for cells subjected to 60 days of storage. (<b>b</b>) Voltage profiles of the 0 V stored cell when discharged at 0.2 C and 1 C, respectively. (<b>c</b>) Comparison of capacity changes with cycling when discharged at 1 C and 0.2 C, respectively.</p>
Full article ">Figure 3
<p>Nyquist plots of (<b>a</b>) EIS changes after storage of PW//HC cells. (<b>b</b>) Comparison of fitted impedance before and after storage.</p>
Full article ">Figure 4
<p>Effect of storage conditions on NMST//HC cells as shown by (<b>a</b>) the changes in OCV of cells over 60 days of storage and (<b>b</b>) the change in OCV plotted as a percentage change. (<b>c</b>) Post-cycling performance after storage. (<b>d</b>) Voltage profiles of the 0 V stored cell when discharged at 0.2 C and 1 C, respectively. (<b>e</b>) Comparison of capacity changes with cycling when discharged at 1 C and 0.2 C, respectively.</p>
Full article ">Figure 5
<p>Comparison of impedance change and cycling performance with and without reconditioning after storage; (<b>a1</b>,<b>a2</b>) cell storage at 10% SOC; (<b>b1</b>,<b>b2</b>) cell storage at 0 V; (<b>c1</b>,<b>c2</b>) cell storage in a shorted state.</p>
Full article ">
14 pages, 1664 KiB  
Article
Flexible Highly Thermally Conductive PCM Film Prepared by Centrifugal Electrospinning for Wearable Thermal Management
by Jiaxin Qiao, Chonglin He, Zijiao Guo, Fankai Lin, Mingyong Liu, Xianjie Liu, Yifei Liu, Zhaohui Huang, Ruiyu Mi and Xin Min
Materials 2024, 17(20), 4963; https://doi.org/10.3390/ma17204963 - 11 Oct 2024
Viewed by 416
Abstract
Personal thermal management materials integrated with phase-change materials have significant potential to satisfy human thermal comfort needs and save energy through the efficient storage and utilization of thermal energy. However, conventional organic phase-change materials in a solid state suffer from rigidity, low thermal [...] Read more.
Personal thermal management materials integrated with phase-change materials have significant potential to satisfy human thermal comfort needs and save energy through the efficient storage and utilization of thermal energy. However, conventional organic phase-change materials in a solid state suffer from rigidity, low thermal conductivity, and leakage, making their application challenging. In this work, polyethylene glycol (PEG) was chosen as the phase-change material to provide the energy storage density, polyethylene oxide (PEO) was chosen to provide the backbone structure of the three-dimensional polymer network and cross-linked with the PEG to provide flexibility, and carbon nanotubes (CNTs) were used to improve the mechanical and thermal conductivity of the material. The thermal conductivity of the composite fiber membranes was boosted by 77.1% when CNTs were added at 4 wt%. Water-resistant modification of the composite fiber membranes was successfully performed using glutaraldehyde-saturated steam. The resulting composite fiber membranes had a reasonable range of phase transition temperatures, and the CC4PCF-55 membranes had melting and freezing latent heats of 66.71 J/g and 64.74 J/g, respectively. The results of this study prove that the green CC4PCF-55 composite fiber membranes have excellent flexibility, with good thermal energy storage capacity and thermal conductivity and, therefore, high potential in the field of flexible wearable thermal management textiles. Full article
Show Figures

Figure 1

Figure 1
<p>Centrifugal electrostatic spinning for the preparation of flexible and highly thermally conductive phase-change thermal storage membranes.</p>
Full article ">Figure 2
<p>Images of (<b>a</b>) PEO, (<b>b</b>) PCF-40, (<b>c</b>) PCF-45, (<b>d</b>) PCF-50, (<b>e</b>) PCF-55, (<b>f</b>) PCF-60, (<b>g</b>) C<sub>2</sub>PCF-55, (<b>h</b>) C<sub>4</sub>PCF-55, (<b>i</b>) C<sub>6</sub>PCF-55, (<b>j</b>) C<sub>8</sub>PCF-55, (<b>k</b>) C<sub>10</sub>PCF-55, and (<b>l</b>) CC<sub>4</sub>PCF-55 fibrous membranes and the corresponding single fiber.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>c</b>) Folding, curling, and stretching images, respectively, of C<sub>4</sub>PCF-55 and CC<sub>4</sub>PCF-55 fibrous membranes. (<b>d</b>–<b>f</b>) Optical images of a CC<sub>4</sub>PCF-55 membrane immersed in water for 0 min, 5 min, and 10 min, respectively. (<b>g</b>) FT-IR spectra of C<sub>4</sub>PCF-55 and CC<sub>4</sub>PCF-55 fibrous membranes. (<b>h</b>,<b>i</b>) Water contact angle images of C<sub>4</sub>PCF-55 and CC<sub>4</sub>PCF-55 fibrous membranes, respectively. (<b>j</b>) UV—visible spectrogram of C<sub>4</sub>PCF-55 and CC<sub>4</sub>PCF-55 fibrous membranes.</p>
Full article ">Figure 4
<p>(<b>a</b>) Thermal conductivity of the composite fibrous membranes. (<b>b</b>) XRD patterns of PEG1000, C<sub>4</sub>PCF-55, and CC<sub>4</sub>PCF-55. (<b>c</b>) Shape stability of PEG1000 and CC<sub>4</sub>PCF-55.</p>
Full article ">Figure 5
<p>Thermal properties of the prepared composite fiber membranes: DSC curves of PEG and the composite fiber membranes for (<b>a</b>) the endothermic process and (<b>b</b>) the exothermic process. (<b>c</b>) Melting/freezing enthalpies (ΔH<sub>m</sub>/ΔH<sub>f</sub>) and melting/freezing phase transition temperatures (T<sub>m</sub>/T<sub>f</sub>) of the composite fiber membranes. DSC curves of the CC<sub>4</sub>PCF-55 fiber membrane after 2, 20, and 50 cycles in the (<b>d</b>) endothermic process and (<b>e</b>) exothermic process.</p>
Full article ">Figure 6
<p>Representative thermal images of the PEO and CC<sub>4</sub>PCF55 membranes recorded by an IR camera in the (<b>a</b>) heating and (<b>b</b>) cooling processes. (<b>c</b>,<b>d</b>) Temperature evolution graphs of the PEO and CC<sub>4</sub>PCF55 fiber membranes.</p>
Full article ">
35 pages, 2134 KiB  
Review
Geochemistry in Geological CO2 Sequestration: A Comprehensive Review
by Jemal Worku Fentaw, Hossein Emadi, Athar Hussain, Diana Maury Fernandez and Sugan Raj Thiyagarajan
Energies 2024, 17(19), 5000; https://doi.org/10.3390/en17195000 - 8 Oct 2024
Viewed by 705
Abstract
The increasing level of anthropogenic CO2 in the atmosphere has made it imperative to investigate an efficient method for carbon sequestration. Geological carbon sequestration presents a viable path to mitigate greenhouse gas emissions by sequestering the captured CO2 deep underground in [...] Read more.
The increasing level of anthropogenic CO2 in the atmosphere has made it imperative to investigate an efficient method for carbon sequestration. Geological carbon sequestration presents a viable path to mitigate greenhouse gas emissions by sequestering the captured CO2 deep underground in rock formations to store it permanently. Geochemistry, as the cornerstone of geological CO2 sequestration (GCS), plays an indispensable role. Therefore, it is not just timely but also urgent to undertake a comprehensive review of studies conducted in this area, articulate gaps and findings, and give directions for future research areas. This paper reviews geochemistry in terms of the sequestration of CO2 in geological formations, addressing mechanisms of trapping, challenges, and ways of mitigating challenges in trapping mechanisms; mineralization and methods of accelerating mineralization; and the interaction between rock, brine, and CO2 for the long-term containment and storage of CO2. Mixing CO2 with brine before or during injection, using microbes, selecting sedimentary reservoirs with reactive minerals, co-injection of carbonate anhydrase, and enhancing the surface area of reactive minerals are some of the mechanisms used to enhance mineral trapping in GCS applications. This review also addresses the potential challenges and opportunities associated with geological CO2 storage. Challenges include caprock integrity, understanding the lasting effects of storing CO2 on geological formations, developing reliable models for monitoring CO2–brine–rock interactions, CO2 impurities, and addressing public concerns about safety and environmental impacts. Conversely, opportunities in the sequestration of CO2 lie in the vast potential for storing CO2 in geological formations like depleted oil and gas reservoirs, saline aquifers, coal seams, and enhanced oil recovery (EOR) sites. Opportunities include improved geochemical trapping of CO2, optimized storage capacity, improved sealing integrity, managed wellbore leakage risk, and use of sealant materials to reduce leakage risk. Furthermore, the potential impact of advancements in geochemical research, understanding geochemical reactions, addressing the challenges, and leveraging the opportunities in GCS are crucial for achieving sustainable carbon mitigation and combating global warming effectively. Full article
(This article belongs to the Collection Feature Papers in Carbon Capture, Utilization, and Storage)
Show Figures

Figure 1

Figure 1
<p>Reduction in volume of CO<sub>2</sub> with increasing depth [<a href="#B35-energies-17-05000" class="html-bibr">35</a>].</p>
Full article ">Figure 2
<p>Pressure–temperature phase diagram for CO<sub>2</sub> [modified from [<a href="#B45-energies-17-05000" class="html-bibr">45</a>]].</p>
Full article ">Figure 3
<p>Security of CO<sub>2</sub> trapping mechanisms over time [<a href="#B66-energies-17-05000" class="html-bibr">66</a>].</p>
Full article ">Figure 4
<p>Solubility and mineral trapping in basalt formation [<a href="#B66-energies-17-05000" class="html-bibr">66</a>].</p>
Full article ">Figure 5
<p>Schematic of the trail of residual CO<sub>2</sub> that is left behind due to snap-off as the plume migrates upward during the post-injection period [<a href="#B79-energies-17-05000" class="html-bibr">79</a>].</p>
Full article ">Figure 6
<p>Main mechanisms for loss of wellbore integrity [modified from [<a href="#B159-energies-17-05000" class="html-bibr">159</a>]].</p>
Full article ">
11 pages, 2290 KiB  
Article
Enhancing Electrochemical Performance of Si@CNT Anode by Integrating SrTiO3 Material for High-Capacity Lithium-Ion Batteries
by Nischal Oli, Diana C. Liza Castillo, Brad R. Weiner, Gerardo Morell and Ram S. Katiyar
Molecules 2024, 29(19), 4750; https://doi.org/10.3390/molecules29194750 - 8 Oct 2024
Viewed by 712
Abstract
Silicon (Si) has attracted worldwide attention for its ultrahigh theoretical storage capacity (4200 mA h g−1), low mass density (2.33 g cm−3), low operating potential (0.4 V vs. Li/Li+), abundant reserves, environmentally benign nature, and low cost. [...] Read more.
Silicon (Si) has attracted worldwide attention for its ultrahigh theoretical storage capacity (4200 mA h g−1), low mass density (2.33 g cm−3), low operating potential (0.4 V vs. Li/Li+), abundant reserves, environmentally benign nature, and low cost. It is a promising high-energy-density anode material for next-generation lithium-ion batteries (LIBs), offering a replacement for graphite anodes owing to the escalating energy demands in booming automobile and energy storage applications. Unfortunately, the commercialization of silicon anodes is stringently hindered by large volume expansion during lithiation–delithiation, the unstable and detrimental growth of electrode/electrolyte interface layers, sluggish Li-ion diffusion, poor rate performance, and inherently low ion/electron conductivity. These present major safety challenges lead to quick capacity degradation in LIBs. Herein, we present the synergistic effects of nanostructured silicon and SrTiO3 (STO) for use as anodes in Li-ion batteries. Si and STO nanoparticles were incorporated into a multiwalled carbon nanotube (CNT) matrix using a planetary ball-milling process. The mechanical stress resulting from the expansion of Si was transferred via the CNT matrix to the STO. We discovered that the introduction of STO can improve the electrochemical performance of Si/CNT nanocomposite anodes. Experimental measurements and electrochemical impedance spectroscopy provide evidence for the enhanced mobility of Li-ions facilitated by STO. Hence, incorporating STO into the Si@CNT anode yields promising results, exhibiting a high initial Coulombic efficiency of approximately 85%, a reversible specific capacity of ~800 mA h g−1 after 100 cycles at 100 mA g−1, and a high-rate capability of 1400 mA g−1 with a capacity of 800 mA h g−1. Interestingly, it exhibits a capacity of 350 mAh g−1 after 1000 lithiation and delithiation cycles at a high rate of 600 mA hg−1. This result unveils and sheds light on the design of a scalable method for manufacturing Si anodes for next-generation LIBs. Full article
(This article belongs to the Special Issue Advanced Nanomaterials for Energy Storage Devices)
Show Figures

Figure 1

Figure 1
<p>Structural analysis: (<b>a</b>) XRD pattern of the Si@STO@CNT composite; (<b>b</b>) XRD pattern of CNTs; (<b>c</b>) XRD pattern of STO; and (<b>d</b>) crystal structure of Si and STO, respectively.</p>
Full article ">Figure 2
<p>SEM analysis: (<b>a</b>) 100 nm; (<b>b</b>) 500 nm; and (<b>c</b>) SEM mapping of Si; (<b>d</b>) Sr; (<b>e</b>) Ti; and (<b>f</b>) O.</p>
Full article ">Figure 3
<p>Electrochemical performance of composite Si/STO/CNTs. (<b>a</b>) Galvanostatic charge–discharge (GCD) curve at a low current density of 100 mA g<sup>−1</sup>; (<b>b</b>) cycling performance at a current density of 100 mA g<sup>−1</sup>; (<b>c</b>) rate curve at a range of current densities from 100 to 1400 mA g<sup>−1</sup>; (<b>d</b>) rate performance at a range of current densities from 100 to 1400 mA g<sup>−1</sup>; (<b>e</b>) cyclic voltammetry at a scan rate of 0.1 mV s<sup>−1</sup>.</p>
Full article ">Figure 4
<p>Cyclic performance at current densities of (<b>a</b>) 400 mA g<sup>−1</sup> and (<b>b</b>) 600 mA g<sup>−1</sup>.</p>
Full article ">Figure 5
<p>Electrochemical Impedance Spectroscopy analysis of composite Si/STO/CNTs. (<b>a</b>) Before lithiation–delithiation and (<b>b</b>) after 150 lithiation–delithiation cycles.</p>
Full article ">
16 pages, 3527 KiB  
Article
Organic Carbon Storage in Waterlogging Soils in Ávila, Spain: A Traditional Agrosilvopastoral Region
by María P. Alvarez-Castellanos, Laura Escudero-Campos, Jorge Mongil-Manso, Francisco J. San Jose, Adrián Jiménez-Sánchez and Raimundo Jiménez-Ballesta
Land 2024, 13(10), 1630; https://doi.org/10.3390/land13101630 - 8 Oct 2024
Viewed by 432
Abstract
Soils play a crucial role in the protection, management, and ecological understanding of the La Moraña region, located in Ávila province, Central Spain, which has a moderate population, traditional agriculture, livestock farming, and low industrial activity, resulting in relatively low environmental degradation. The [...] Read more.
Soils play a crucial role in the protection, management, and ecological understanding of the La Moraña region, located in Ávila province, Central Spain, which has a moderate population, traditional agriculture, livestock farming, and low industrial activity, resulting in relatively low environmental degradation. The region’s soils often experience prolonged water stagnation, influencing its agronomy, ecology, and economy. This study aimed to estimate and understand the soil’s role in the C sequestration of an agrosilvopastoral system under conditions of temporary water stagnation and different land uses. The results showed that ryegrass-magaza and Pinus pinaster show more content in soil carbon sequestration storage (98.7 and 92.4 Mg per hectare) compared to the adjacent degraded rangeland (75.8 and 63.9 Mg ha−1). Arenosols exhibited a higher total amount of SOC stocks. The soil profile with ryegrass sequestered more nitrogen (9.7 Mg ha−1) than other land uses; moreover, Arenosols have a lower nitrogen sequestration capacity even in low-forest conditions. The study highlights significant differences in carbon accumulation due to the management practices, temporary water layers, and parent material. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area. UTM coordinates, ETRS89 datum, Zone 30T: P1, P2, P3 and P4 = soil profiles; S1, S2….S8 = surface soil samples.</p>
Full article ">Figure 2
<p>Distribution of soil profiles in the landscape.</p>
Full article ">Figure 3
<p>Tendency for SOC content to decrease in the analyzed soil profiles.</p>
Full article ">Figure 4
<p>The excess of water induces reduction processes, resulting in generally low soil chroma and value (&lt;2).</p>
Full article ">
13 pages, 2699 KiB  
Article
Insight into the Reversible Hydrogen Storage of Titanium-Decorated Boron-Doped C20 Fullerene: A Theoretical Prediction
by Zhiliang Chai, Lili Liu, Congcong Liang, Yan Liu and Qiang Wang
Molecules 2024, 29(19), 4728; https://doi.org/10.3390/molecules29194728 - 6 Oct 2024
Viewed by 592
Abstract
Hydrogen storage has been a bottleneck factor for the application of hydrogen energy. Hydrogen storage capacity for titanium-decorated boron-doped C20 fullerenes has been investigated using the density functional theory. Different boron-doped C20 fullerene absorbents are examined to avoid titanium atom clustering. [...] Read more.
Hydrogen storage has been a bottleneck factor for the application of hydrogen energy. Hydrogen storage capacity for titanium-decorated boron-doped C20 fullerenes has been investigated using the density functional theory. Different boron-doped C20 fullerene absorbents are examined to avoid titanium atom clustering. According to our research, with three carbon atoms in the pentagonal ring replaced by boron atoms, the binding interaction between the Ti atom and C20 fullerene is stronger than the cohesive energy of titanium. The calculated results revealed that one Ti atom can reversibly adsorb four H2 molecules with an average adsorption energy of −1.52 eV and an average desorption temperature of 522.5 K. The stability of the best absorbent structure with a gravimetric density of 4.68 wt% has been confirmed by ab initio molecular dynamics simulations. These findings suggest that titanium-decorated boron-doped C20 fullerenes could be considered as a potential candidate for hydrogen storage devices. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Top and side views of C<sub>20</sub>. (<b>b</b>) Top and side views of the adsorbent B123 model. (<b>c</b>) Binding energy of different amounts of boron doping, where brown, green and blue represent carbon, boron and titanium atoms respectively. (<b>d</b>) Changes in energy and bond length in the ab initio molecular dynamics simulation (300 K, 5 ps) of the B123 model.</p>
Full article ">Figure 2
<p>Hydrogen adsorption on the B123 model, where brown, green, blue and white represent carbon, boron, titanium and hydrogen atoms respectively.</p>
Full article ">Figure 3
<p>Desorption temperature as a function of pressure in the B123 model.</p>
Full article ">Figure 4
<p>(<b>a</b>) Relative energy as a function of temperature under a given pressure in the B123 model. (<b>b</b>) Relative energy as a function of pressure under a given temperature in the B123 model.</p>
Full article ">Figure 5
<p>Total density of states of C<sub>20</sub> fullerene and B123 model.</p>
Full article ">Figure 6
<p>Partial density of states for (<b>a</b>) C 2p orbital of C<sub>20</sub>; (<b>b</b>) C 2p orbital of B123; (<b>c</b>) B 2p orbital of isolated B; (<b>d</b>) Ti 3d orbital of isolated Ti; (<b>e</b>) B 2p orbital of B123; and (<b>f</b>) Ti 3d orbital of B123.</p>
Full article ">Figure 7
<p>Partial density of states for the (<b>a</b>) H 1s orbital of isolated H<sub>2</sub>; (<b>b</b>) H 1s orbital of C<sub>20</sub> fullerene + H<sub>2</sub>; (<b>c</b>) H 1s orbital of B123 + H<sub>2</sub>; (<b>d</b>) Ti 3d orbital of B123; and (<b>e</b>) Ti 3d orbital of B123 + H<sub>2</sub>. Fermi level is set to zero energy.</p>
Full article ">Figure 8
<p>Charge density difference for the (<b>a</b>) B123 system; (<b>b</b>) B123 + H<sub>2</sub> system. Yellow and blue colors represent charge-gained and charge-lost regions, respectively.</p>
Full article ">
31 pages, 6153 KiB  
Review
Sodium-Ion Battery at Low Temperature: Challenges and Strategies
by Yan Zhao, Zhen Zhang, Yalong Zheng, Yichao Luo, Xinyu Jiang, Yaru Wang, Zhoulu Wang, Yutong Wu, Yi Zhang, Xiang Liu and Baizeng Fang
Nanomaterials 2024, 14(19), 1604; https://doi.org/10.3390/nano14191604 - 4 Oct 2024
Viewed by 529
Abstract
Sodium-ion batteries (SIBs) have garnered significant interest due to their potential as viable alternatives to conventional lithium-ion batteries (LIBs), particularly in environments where low-temperature (LT) performance is crucial. This paper provides a comprehensive review of current research on LT SIBs, focusing on electrode [...] Read more.
Sodium-ion batteries (SIBs) have garnered significant interest due to their potential as viable alternatives to conventional lithium-ion batteries (LIBs), particularly in environments where low-temperature (LT) performance is crucial. This paper provides a comprehensive review of current research on LT SIBs, focusing on electrode materials, electrolytes, and operational challenges specific to sub-zero conditions. Recent advancements in electrode materials, such as carbon-based materials and titanium-based materials, are discussed for their ability to enhance ion diffusion kinetics and overall battery performance at colder temperatures. The critical role of electrolyte formulation in maintaining battery efficiency and stability under extreme cold is highlighted, alongside strategies to mitigate capacity loss and cycle degradation. Future research directions underscore the need for further improvements in energy density and durability and scalable manufacturing processes to facilitate commercial adoption. Overall, LT SIBs represent a promising frontier in energy storage technology, with ongoing efforts aimed at overcoming technical barriers to enable widespread deployment in cold-climate applications and beyond. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) A schematic illustration of the synthesis for Zn-HC. (<b>b</b>) A schematic diagram of the enhanced rate capability mechanism. (<b>c</b>) The low-temperature performance (−40 °C) of the GCD curves (the second cycle) and cycling stabilities. Reproduced with permission from ref. [<a href="#B40-nanomaterials-14-01604" class="html-bibr">40</a>]. Copyright 2023 Wiley-VCH GmbH. HRTEM image of (<b>d</b>) CTS 1300, (<b>e</b>) CTSFS 1300. (<b>f</b>) Cycling performances of CTS 1300 and CTSFS 1300 at −20 °C. Reproduced with permission from ref. [<a href="#B41-nanomaterials-14-01604" class="html-bibr">41</a>]. Copyright 2024 Elsevier. (<b>g</b>) A schematic diagram illustrating the synthesis route of the samples. (<b>h</b>) The GCD curves at −40 °C (the second cycle). Reproduced with permission from ref. [<a href="#B42-nanomaterials-14-01604" class="html-bibr">42</a>]. Copyright 2023 Wiley-VCH GmbH.</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic diagram of synthetic route for H-NTO microspheres. (<b>b</b>) Cycle performance at −40 °C of H-NTO electrode. (<b>c</b>) Atomic ratio of C, F, and Na in SEI formed for cycled H-NTO electrode. Reproduced with permission from ref. [<a href="#B47-nanomaterials-14-01604" class="html-bibr">47</a>]. Copyright 2023 Wiley-VCH GmbH. (<b>d</b>) Capacitive contributions to charge storage at 1.0 mV s<sup>−1</sup> of a-KTiO<sub>x</sub>/Ti<sub>2</sub>CT<sub>x</sub>. (<b>e</b>) Schematic illustration for preparation of a-KTiO<sub>x</sub>/Ti<sub>2</sub>CT<sub>x</sub>. FESEM images of (<b>f</b>) pristine Ti<sub>2</sub>AlC MAX, (<b>g</b>) multi-layered Ti<sub>2</sub>CT<sub>x</sub>, and (<b>h</b>) a-KTiO<sub>x</sub>/Ti<sub>2</sub>CT<sub>x</sub>. (<b>i</b>) Cycling performances at 0.1 A g<sup>−1</sup> of a-KTiO<sub>x</sub>Ti<sub>2</sub>CT<sub>x</sub> at −25 °C. Reproduced with permission from ref. [<a href="#B49-nanomaterials-14-01604" class="html-bibr">49</a>]. Copyright 2022 Elsevier Inc.</p>
Full article ">Figure 3
<p>(<b>a</b>) Illustration of fabrication of WS<sub>2</sub>/Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> heterojunction. (<b>b</b>) Illustration of electron redistribution at interface between layered WS<sub>2</sub> and Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> MXene. (<b>c</b>) Sodium storage performance at −20 °C: cycling at 0.1 A g<sup>−1</sup> of WS<sub>2</sub>/Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> heterojunction, pure WS<sub>2</sub>, and Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> MXene. (<b>d</b>) Rate capabilities of WS<sub>2</sub>/Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>, pure WS<sub>2</sub>, and Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub> MXene. Reproduced with permission from ref. [<a href="#B50-nanomaterials-14-01604" class="html-bibr">50</a>]. Copyright 2023 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published by ELSEVIER B.V. and Science Press. (<b>e</b>,<b>f</b>) SEM images of in situ TiO<sub>2</sub>@rGO. (<b>g</b>) Contribution ratio of capacitive and diffusion-controlled charges in TiO<sub>2</sub>@rGO at different scan rates. (<b>h</b>) Separation of capacitive (red region) and diffusion currents in TiO<sub>2</sub>@rGO at a scan rate of 10 mV S<sup>−1</sup>. (<b>i</b>) Charge and discharge capacity of TiO<sub>2</sub>@rGO versus cycle number at current densities of 5C in temperature of −40 °C. Reproduced with permission from ref. [<a href="#B51-nanomaterials-14-01604" class="html-bibr">51</a>]. Copyright 2020 Elsevier B.V.</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic diagram of the synthesis procedure of ZnSe@NCNFs. (<b>b</b>,<b>c</b>) Na<sup>+</sup> diffusion coefficient during cycle process. (<b>d</b>) Electrochemical performance of ZnSe@NCNFs in low temperatures for SIBs on 0.5 and 1 A g<sup>−1</sup>. Reproduced with permission from ref. [<a href="#B54-nanomaterials-14-01604" class="html-bibr">54</a>]. Copyright 2021 American Chemical Society. (<b>e</b>) Synthesis diagram of FPS/GPNC. (<b>f</b>) Long cycling stability of FPS/GPNC at −20 °C. Reproduced with permission from ref. [<a href="#B55-nanomaterials-14-01604" class="html-bibr">55</a>]. Copyright 2022, American Chemical Society.</p>
Full article ">Figure 5
<p>(<b>a</b>) The schematic of crystal structure for NbSSe. (<b>b</b>) Cycle performance for NbSSe at different temperatures (−20 °C, 0 °C, 20 °C, 40 °C). Reproduced with permission from ref. [<a href="#B60-nanomaterials-14-01604" class="html-bibr">60</a>]. copyright 2022 Elsevier B.V. (<b>c</b>) Cycle stability of the Bi//NFPP@C batteries at −40 °C. (<b>d</b>) CV curve in current versus time form and corresponding 2D color-filled contour plot in wavenumber versus time form of the in situ FTIR/ATR test of the Na//Bi half battery. (<b>e</b>) The galvanostatic charge/discharge curves of Na//Bi half battery during the in situ XRD test and corresponding 2D color-filled contour plot. Reproduced with permission from ref. [<a href="#B67-nanomaterials-14-01604" class="html-bibr">67</a>]. Copyright 2022 Wiley-VCH GmbH. (<b>f</b>) Schematic illustration of the synthesis process of Bi@3DCF. (<b>g</b>) cycling stability at 1 A g<sup>−1</sup> for EMP-Bi@3DCF and Bi@3DCF at −20 °C. Reproduced with permission from ref. [<a href="#B65-nanomaterials-14-01604" class="html-bibr">65</a>]. Copyright 2020 Elsevier Ltd.</p>
Full article ">Figure 6
<p>(<b>a</b>) A schematic illustration of the structural stabilization of NVP electrode materials by doping K ions into the Na sites. (<b>b</b>) Cycle performance at 1 C of NVP, NVP-K<sub>0.05</sub>, and NVP-K<sub>0.10</sub> samples under the low temperature of −25 °C. Reproduced with permission from ref. [<a href="#B87-nanomaterials-14-01604" class="html-bibr">87</a>]. Copyright 2023 Elsevier Ltd. (<b>c</b>) Crystal structure of the as-constructed Na<sub>4</sub>VMn<sub>0.7</sub>Ni<sub>0.3</sub>(PO<sub>4</sub>)<sub>3</sub>. (<b>d</b>) SEM of the NVMNP@C material. (<b>e</b>) Cycle performance (−40 °C) of NVMNP@C and NVMP@C cathodes at 20 mA g<sup>−1</sup>. Reproduced with permission from ref. [<a href="#B88-nanomaterials-14-01604" class="html-bibr">88</a>]. Copyright 2024 The Authors. (<b>f</b>) Schematic illustration (fi) of the synthesis of NFPP–N and (fii) the electronic and Na<sup>+</sup> diffusion paths on one-dimensional NFPP nanoribbons. (<b>g</b>) Rate performances of NFPP–N and NFPP–P at −15 °C. (<b>h</b>) The cycling performances of NFPP–N and NFPP–P at 0.05 C at −15, 0, and 25 °C. Reproduced with permission from ref. [<a href="#B89-nanomaterials-14-01604" class="html-bibr">89</a>]. Copyright 2021 Elsevier B.V.</p>
Full article ">Figure 7
<p>(<b>a</b>) A schematic illustration of the electronic coupling between NTP and in situ generated carbonaceous skeleton. Reproduced with permission from ref. [<a href="#B90-nanomaterials-14-01604" class="html-bibr">90</a>]. Copyright 2022 Elsevier Inc. (<b>b</b>) Schematic illustration of the crystal structure of Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>2</sub>F<sub>3</sub>. Reproduced with permission from ref. [<a href="#B93-nanomaterials-14-01604" class="html-bibr">93</a>]. Copyright 2020 Elsevier Ltd. (<b>c</b>) Diagram of the V−F−C bonding between NVPOF and graphene. (<b>d</b>) Cycling performance at 2 C at −40 °C of the pure NVPOF and CB-NVPOF at −40 °C. Reproduced with permission from ref. [<a href="#B94-nanomaterials-14-01604" class="html-bibr">94</a>]. Copyright 2023, American Chemical Society. (<b>e</b>) Schematic illustration of the formation mechanism of the NVPF formed by HTS and tube furnace (TF). (<b>f</b>) A schematic illustration of the crystal structure of NVPF. Reproduced with permission from ref. [<a href="#B95-nanomaterials-14-01604" class="html-bibr">95</a>]. Copyright 2024 Wiley-VCH GmbH.</p>
Full article ">Figure 8
<p>(<b>a</b>) Schematic illustration showing the mixed conducting network in the PB/CNT. (<b>b</b>) GDC profiles at 0.1 C of the PB/CNT cathode at −25 °C for 100 cycles. (<b>c</b>) Rate capabilities of the PB/CNT cathode at different temperatures. Reproduced with permission from ref. [<a href="#B112-nanomaterials-14-01604" class="html-bibr">112</a>]. Copyright 2016 WILEY-VCH Verlag GmbH &amp; Co. KGaA, Weinheim. (<b>d</b>) Galvanostatic cycling test of Na symmetric cells at a current density of 0.5 mA cm<sup>−2</sup>. (<b>e</b>) Low-temperature cycling test of SSIBs. (<b>f</b>) Corresponding charge–discharge curves of SSIBs. Reproduced with permission from ref. [<a href="#B113-nanomaterials-14-01604" class="html-bibr">113</a>]. Copyright 2019 WILEY-VCH Verlag GmbH &amp; Co. KGaA, Weinheim. (<b>g</b>) Schematic illustrations of the structures of the monoclinic MnHCF-S-170. (<b>h</b>) Space utilization and water consumption comparison of the as-prepared samples. (<b>i</b>) Long-term cycling performances at −10 °C and 50 °C at the current density of 100 mA g<sup>−1</sup>. Reproduced with permission from ref. [<a href="#B114-nanomaterials-14-01604" class="html-bibr">114</a>]. Copyright 2022 Wiley-VCH GmbH.</p>
Full article ">Figure 9
<p>(<b>a</b>) Cycling performance of P<sub>2</sub>/O<sub>3</sub>-NaMnNiCuFeTiOF at different temperatures. (<b>b</b>) Schematic diagram of structural evolution of P<sub>2</sub>/O<sub>3</sub>-NaMnNiCuFeTiOF. Reproduced with permission from ref. [<a href="#B117-nanomaterials-14-01604" class="html-bibr">117</a>]. Copyright 2023 Elsevier B.V. (<b>c</b>) Schematic illustration of effect of Co<sup>3+</sup> doping on electrode structure. (<b>d</b>) GITT curves and the corresponding D<sub>Na</sub><sup>+</sup> values for the NNCM cathode material at −40 °C. (<b>e</b>) Rate capabilities at different rates from 0.2 to 5 C at −40 °C. Reproduced with permission from ref. [<a href="#B118-nanomaterials-14-01604" class="html-bibr">118</a>]. Copyright 2021 Elsevier B.V. (<b>f</b>) P<sub>2</sub>-type crystal structure viewed along a axis (left) and c axis (right). (<b>g</b>) Typical charge/discharge profiles of NM-2 at 60, 45, 25, 5, and −30 °C. (<b>h</b>) Cycle performance at −30 °C of NM-2. Reproduced with permission from ref. [<a href="#B119-nanomaterials-14-01604" class="html-bibr">119</a>]. Copyright 2023 Wiley-VCH GmbH.</p>
Full article ">Figure 10
<p>(<b>a</b>) Long-term cycling performance for NVPOF cathodes at −40 and −50 °C. (<b>b</b>) Rate capability for NVPOF cathode at −40 °C. (<b>c</b>) Schematic illustration of the formed CEI in different electrolytes. Reproduced with permission from ref. [<a href="#B127-nanomaterials-14-01604" class="html-bibr">127</a>]. Copyright 2024 Published by Elsevier Ltd. on behalf of the editorial office of the <span class="html-italic">Journal of Materials Science &amp; Technology</span>. (<b>d</b>) Schematic illustration of Na growth behavior and temperature-responsive SEI in the WT electrolyte at subzero and high temperature. (<b>e</b>) Representative solvation sheath determined by MD simulations. The Na, C, O, H, F, S, and N are marked with purple, gray, red, white, light blue, yellow, and navy blue, respectively. (<b>f</b>) Long-term cycling performance of Na/NVP cells operated in the corresponding electrolytes at 0.5 C, −20 °C. Reproduced with permission from ref. [<a href="#B128-nanomaterials-14-01604" class="html-bibr">128</a>]. Copyright 2022 Elsevier Ltd.</p>
Full article ">Figure 11
<p>(<b>a</b>) The designed new low-temperature electrolyte system of ES<sub>6</sub>-BLTE. (<b>b</b>) Cyclability of the Na||NVP cells with CSE, BLTE, and ES<sub>6</sub>-BLTE electrolytes at 0.1 C at −40 °C. (<b>c</b>) Rate performance of Na||NVP cells at −40 °C. Reproduced with permission from ref. [<a href="#B129-nanomaterials-14-01604" class="html-bibr">129</a>]. Copyright 2023 Wiley-VCH GmbH. (<b>d</b>) Ionic conductivity of the electrolyte blends: control (1 M NaPF<sub>6</sub> in EC-PC-DMC 1:1:2 by vol%), 20%MA + control (1 M NaPF<sub>6</sub> in EC-PC-DMC-MA 1:1:2:1 by vol%), and 20%EA + control (1 M NaPF<sub>6</sub> in EC-PC-DMC-EA 1:1:2:1 by vol%) measured at different temperatures. Reproduced with permission from ref. [<a href="#B130-nanomaterials-14-01604" class="html-bibr">130</a>]. Copyright 2022 Elsevier Ltd. (<b>e</b>) The capacity retentions of Sb<sub>2</sub>Se<sub>3</sub>/rGO along with temperature variations between 25 and −15 °C at various current densities of 0.2, 0.5, 1, and 2 A g−1. Reproduced with permission from ref. [<a href="#B57-nanomaterials-14-01604" class="html-bibr">57</a>]. Copyright 2022 Royal Society of Chemistry. (<b>f</b>) Charge–discharge voltage profiles at various current densities of MoS<sub>2</sub>@MXene@D-TiO<sub>2</sub> electrodes. (<b>g</b>) Low-temperature cycling performance of MoS<sub>2</sub>@MXene@D-TiO<sub>2</sub> electrodes. Reproduced with permission from ref. [<a href="#B59-nanomaterials-14-01604" class="html-bibr">59</a>]. Copyright 2022 American Chemical Society. (<b>h</b>) The cycling stability during 500 cycles at different operation temperatures of the 500 mAh soft-packed Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>3</sub>//Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>3</sub> symmetrical battery. Reproduced with permission from ref. [<a href="#B131-nanomaterials-14-01604" class="html-bibr">131</a>]. Copyright 2021 The Chemical Industry and Engineering Society of China, and Chemical Industry Press Co., Ltd.</p>
Full article ">Figure 12
<p>(<b>a</b>) LUMO and the highest occupied molecular orbital (HOMO) level of Na<sup>+</sup>-EC, Na<sup>+</sup>-PC, Na<sup>+</sup>-EMC, Na<sup>+</sup>-DME, and Na<sup>+</sup>-DEGDME complexes. Reproduced with permission from ref. [<a href="#B133-nanomaterials-14-01604" class="html-bibr">133</a>]. Copyright 2021 Wiley-VCH GmbH. (<b>b</b>) Cycling performance of the cells equipped with NDT electrolyte at 25 and −40 °C, at a current density of 100 mA g<sup>−1</sup>. (<b>c</b>) Schematic illustration showing the low-temperature electrolyte design strategy. The solvation structure variations with temperature are driven by entropy change: non-adaptive electrolyte (left) and temperature-adaptive electrolyte (right). Reproduced with permission from ref. [<a href="#B135-nanomaterials-14-01604" class="html-bibr">135</a>]. Copyright 2023 Wiley-VCH GmbH. (<b>d</b>) Schematic illustration of SIBs with conventional electrolytes (left) and the designed electrolyte (right) operating at low temperatures. Characterizations of the SEI film for the cycled HC anode with 1 m NaPF<sub>6</sub>-G<sub>2</sub>/DME electrolyte at −40 °C. (<b>e</b>) HRTEM image of the SEI film on HC anode. STEM images of (<b>f</b>) C, O, F elements mapping and (<b>g</b>) line scan of the SEI film. (<b>h</b>) Cycling stability of HC||Na half-cell with 1 m NaPF<sub>6</sub>-G<sub>2</sub>/DME electrolyte at 0.30 A g<sup>−1</sup>, −40 °C. Reproduced with permission from ref. [<a href="#B136-nanomaterials-14-01604" class="html-bibr">136</a>]. Copyright 2024 Wiley-VCH GmbH.</p>
Full article ">Figure 13
<p>(<b>a</b>) Schematic illustration of the mechanism of improved LT Na reversibility by the DOL-diluted electrolyte. (<b>b</b>) Ionic conductivity of 0.4 m NaPF<sub>6</sub>-G<sub>2</sub> and 0.4 m NaPF<sub>6</sub>-G<sub>2</sub>/DOL electrolytes. (<b>c</b>) Long-term galvanostatic cycling performance of AFSMBs at 1 C under −25 °C and −40 °C. Reproduced with permission from ref. [<a href="#B137-nanomaterials-14-01604" class="html-bibr">137</a>]. Copyright 2024 Wiley-VCH GmbH. (<b>d</b>) A schematic illustration of the working mechanism of the selected electrolyte in the MN cathode. (<b>e</b>) GCD curves of MN cathode with TPP electrolyte for different cycles at −40 °C. (<b>f</b>) Cycling performance of MN cathode in TPP and THF electrolytes at −40 °C. Reproduced with permission from ref. [<a href="#B138-nanomaterials-14-01604" class="html-bibr">138</a>]. Copyright 2023 American Chemical Society.</p>
Full article ">Figure 14
<p>(<b>a</b>) Long-term cycles at −20 °C in 1 C and 5 C of SMBs. Reproduced with permission from ref. [<a href="#B139-nanomaterials-14-01604" class="html-bibr">139</a>]. Copyright 2022 Elsevier B.V. Schematic diagrams of Na<sup>+</sup> migration in (<b>b</b>) THF and (<b>c</b>) PC-based electrolytes and SEI at −20 °C. (<b>d</b>) Long-term cycling performance in 1 M NaPF<sub>6</sub>/THF. Reproduced with permission from ref. [<a href="#B140-nanomaterials-14-01604" class="html-bibr">140</a>]. Copyright 2022 Wiley-VCH GmbH. (<b>e</b>) The cycle performance of defective HT-NW at 1 A g<sup>−1</sup> and 3 A g<sup>−1</sup> in 1.0 M NaCF<sub>3</sub>SO<sub>3</sub> diglyme at −25 °C. Reproduced with permission from ref. [<a href="#B141-nanomaterials-14-01604" class="html-bibr">141</a>]. Copyright 2021 Elsevier B.V. (<b>f</b>) Cycling performance at a rate of 920 mA g<sup>−1</sup> at 25 °C and 92 mA g<sup>−1</sup> at −40 °C. Reproduced with permission from ref. [<a href="#B143-nanomaterials-14-01604" class="html-bibr">143</a>]. Copyright 2022 The Author(s). (<b>g</b>) Temperature-dependent performance of the Na/NVPOF cells in the various systems ranging from 25 °C to −30 °C at 0.2 C. Reproduced with permission from ref. [<a href="#B11-nanomaterials-14-01604" class="html-bibr">11</a>]. Copyright 2021 Royal Society of Chemistry.</p>
Full article ">
16 pages, 4071 KiB  
Article
Improving the Performance of LiFePO4 Cathodes with a Sulfur-Modified Carbon Layer
by Su-hyun Kwak and Yong Joon Park
Batteries 2024, 10(10), 348; https://doi.org/10.3390/batteries10100348 - 1 Oct 2024
Viewed by 532
Abstract
LiFePO₄ (LFP) cathodes are popular due to their safety and cyclic performance, despite limitations in lithium-ion diffusion and conductivity. These can be improved with carbon coating, but further advancements are possible despite commercial success. In this study, we modified the carbon coating layer [...] Read more.
LiFePO₄ (LFP) cathodes are popular due to their safety and cyclic performance, despite limitations in lithium-ion diffusion and conductivity. These can be improved with carbon coating, but further advancements are possible despite commercial success. In this study, we modified the carbon coating layer using sulfur to enhance the electronic conductivity and stabilize the carbon surface layer via two methods: 1-step and 2-step processes. In the 1-step process, sulfur powder was mixed with cellulose followed by heat treatment to form a coating layer; in the 2-step process, an additional coating layer was applied on top of the carbon coating layer. Electrochemical measurements demonstrated that the 1-step sulfur-modified LFP significantly improved the discharge capacity (~152 mAh·g−1 at 0.5 C rate) and rate capability compared to pristine LFP. Raman analyses indicated that sulfur mixed with a carbon source increases the graphitization of the carbon layer. Although the 2-step sulfur modification did not exceed the 1-step process in enhancing rate capability, it improved the storage characteristics of LFP at high temperatures. The residual sulfur elements apparently protected the surface. These findings confirm that sulfur modification of the carbon layer is effective for improving LFP cathode properties, offering a promising approach to enhance the performance and stability of LFP-based lithium-ion batteries. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Initial charge–discharge profiles and (<b>b</b>) discharge capacities at 0.05, 0.1, 0.3, 0.5, and 1 C rates for the pristine LFP and 1-step sulfur-modified LFPs. (<b>c</b>) Initial charge–discharge profiles and (<b>d</b>) discharge capacities at 0.05, 0.1, 0.3, 0.5, and 1 C rates for the pristine LFP and 2-step sulfur-modified LFPs.</p>
Full article ">Figure 2
<p>TEM images of the (<b>a</b>) carbon-uncoated LFP, (<b>b</b>) pristine LFP (carbon-coated), (<b>c</b>) 1-step sulfur-modified SLFP, and (<b>d</b>) 2-step sulfur-modified SLFP.</p>
Full article ">Figure 3
<p>Raman spectra of the (<b>a</b>) pristine LFP, (<b>b</b>) 1-step sulfur-modified SLFP, and (<b>c</b>) 2-step sulfur-modified SLFP.</p>
Full article ">Figure 4
<p>TOF-SIMS spectra of the surface of (<b>a</b>) pristine LFP, (<b>b</b>) 1-step sulfur-modified SLFP, (<b>c</b>) 2-step sulfur-modified SLFP, and (<b>d</b>) comparison of their S<sup>−</sup> intensity.</p>
Full article ">Figure 5
<p>Nyquist plots of the cells containing the pristine LFP, 1-step sulfur-modified SLFP, and 2-step sulfur-modified SLFP after (<b>a</b>) 1 cycle and (<b>b</b>) 100 cycles.</p>
Full article ">Figure 6
<p>Li-diffusion coefficient (D<sub>Li+</sub>) values measured using the GITT method for (<b>a</b>) pristine LFP, (<b>b</b>) 1-step sulfur-modified SLFP, and (<b>c</b>) 2-step sulfur-modified SLFP.</p>
Full article ">Figure 7
<p>Charge–discharge profiles of the pristine LFP, 1-step sulfur-modified SLFP, and 2-step sulfur-modified SLFP after storage for (<b>a</b>) 7 days, (<b>b</b>) 10 days, and (<b>c</b>) 14 days. Discharge capacities at 0.05, 0.1, 0.3, 0.5, and 1 C rates for the pristine LFP, 1-step sulfur-modified SLFP, and 2-step sulfur-modified SLFP after storage for (<b>d</b>) 7 days, (<b>e</b>) 10 days, and (<b>f</b>) 14 days.</p>
Full article ">Figure 8
<p>Schematic summarization of the 1-step and 2-step sulfur modification effects.</p>
Full article ">
11 pages, 3439 KiB  
Article
Binary Biomass-Based Electrolyte Films for High-Performance All-Solid-State Supercapacitor
by Rui Lou, Guocheng Zhang, Taoyuan Niu, Long He, Ying Su and Guodong Wei
Polymers 2024, 16(19), 2772; https://doi.org/10.3390/polym16192772 - 30 Sep 2024
Viewed by 452
Abstract
Solid-state electrolytes have received widespread attention for solving the problem of the leakage of liquid electrolytes and effectively improving the overall performance of supercapacitors. However, the electrochemical performance and environmental friendliness of solid-state electrolytes still need to be further improved. Here, a binary [...] Read more.
Solid-state electrolytes have received widespread attention for solving the problem of the leakage of liquid electrolytes and effectively improving the overall performance of supercapacitors. However, the electrochemical performance and environmental friendliness of solid-state electrolytes still need to be further improved. Here, a binary biomass-based solid electrolyte film (LSE) was successfully synthesized through the incorporation of lignin nanoparticles (LNPs) with sodium alginate (SA). The impact of the mass ratio of SA to LNPs on the microstructure, porosity, electrolyte absorption capacity, ionic conductivity, and electrochemical properties of the LSE was thoroughly investigated. The results indicated that as the proportion of SA increased from 5% to 15% of LNPs, the pore structure of the LSE became increasingly uniform and abundant. Consequently, enhancements were observed in porosity, liquid absorption capacity, ionic conductivity, and overall electrochemical performance. Notably, at an SA amount of 15% of LNPs, the ionic conductivity of the resultant LSE-15 was recorded at 14.10 mS cm−1, with the porosity and liquid absorption capacity reaching 58.4% and 308%, respectively. LSE-15 was employed as a solid electrolyte, while LNP-based carbon aerogel (LCA) served as the two electrodes in the construction of a symmetric all-solid-state supercapacitor (SSC). The SSC device demonstrated exceptional electrochemical storage capacity, achieving a specific capacitance of 197 F g−1 at 0.5 A g−1, along with a maximum energy and power density of 27.33 W h kg−1 and 4998 W kg−1, respectively. Furthermore, the SSC device exhibited highly stable electrochemical performance under extreme conditions, including compression, bending, and both series and parallel connections. Therefore, the development and application of binary biomass-based solid electrolyte films in supercapacitors represent a promising strategy for harnessing high-value biomass resources in the field of energy storage. Full article
(This article belongs to the Section Polymer Membranes and Films)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the LSE preparation process.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>–<b>c</b>) the LSE-5, LSE-10, LSE-15, and (<b>d</b>) LSE-20.</p>
Full article ">Figure 3
<p>(<b>a</b>) Graphical comparisons of the LSE before and after immersion in a 3.3 M KOH solution for 4 h; (<b>b</b>) porosity and electrolyte absorption capacity (<span class="html-italic">η<sub>a</sub></span>); (<b>c</b>) C 1s XPS spectra of LSE-15; (<b>d</b>) EIS curves and the ionic conductivity (<span class="html-italic">δ</span>) of the LSE.</p>
Full article ">Figure 4
<p>CV curves of (<b>a</b>) LSE-5, (<b>b</b>) LSE-10, (<b>c</b>) LSE-15, (<b>d</b>) LSE-20 at different scan rates; GCD curves of (<b>e</b>) LSE-5, (<b>f</b>) LSE-10, (<b>g</b>) LSE-15, (<b>h</b>) LSE-20 at different current densities.</p>
Full article ">Figure 5
<p>(<b>a</b>) CV curves of the SSC at 20 mV s<sup>−1</sup>; (<b>b</b>) GCD curves of the SSC at 0.5 A g<sup>−1</sup>; (<b>c</b>) specific capacitance; (<b>d</b>) Ragone plots [<a href="#B11-polymers-16-02772" class="html-bibr">11</a>,<a href="#B28-polymers-16-02772" class="html-bibr">28</a>,<a href="#B29-polymers-16-02772" class="html-bibr">29</a>]; (<b>e</b>) EIS curves; (<b>f</b>) long-term cycling stability and coulombic efficiency.</p>
Full article ">Figure 6
<p>CV curves of the SSC under compression (<b>a</b>) and bending (<b>b</b>) conditions; the CV (<b>c</b>) and GCD (<b>d</b>) curves of two SSCs in series; the CV (<b>e</b>) and GCD (<b>f</b>) curves of two SSCs in parallel.</p>
Full article ">
14 pages, 3322 KiB  
Article
Pomegranate Peel-Derived Hard Carbons as Anode Materials for Sodium-Ion Batteries
by Qijie Wu, Kewei Shu, Long Zhao and Jianming Zhang
Molecules 2024, 29(19), 4639; https://doi.org/10.3390/molecules29194639 - 29 Sep 2024
Viewed by 627
Abstract
Exploring high-performance carbon anodes that are low-cost and easily accessible is the key to the commercialization of sodium-ion batteries. Producing carbon materials from bio by-products is an intriguing strategy for sodium-ion battery anode manufacture and for high-value utilization of biomass. Herein, a novel [...] Read more.
Exploring high-performance carbon anodes that are low-cost and easily accessible is the key to the commercialization of sodium-ion batteries. Producing carbon materials from bio by-products is an intriguing strategy for sodium-ion battery anode manufacture and for high-value utilization of biomass. Herein, a novel hard carbon (PPHC) was prepared via a facile pyrolysis process followed by acid treatment using biowaste pomegranate peel as the precursor. The morphology and structure of the PPHC were influenced by the carbonization temperature, as evidenced by physicochemical characterization. The PPHC pyrolyzed at 1100 °C showed expanded interlayer spacing and appropriate oxygen group content. When used as a sodium ion battery anode, the PPHC-1100 demonstrated a reversible capacity of up to 330 mAh g−1, maintaining 174 mAh g−1 at an increased current rate of 1 C. After 200 cycles at 0.5 C, the capacity delivered by PPHC-1100 was 175 mAh g−1. The electrochemical behavior of PPHC electrodes was investigated, revealing that the PPHC-1100 possessed increased capacitive-controlled energy storage and improved ion transport properties, which explained its excellent electrochemical performance. This work underscores the feasibility of high-performance sodium-ion battery anodes derived from biowaste and provides insights into the sodium storage process in biomass-derived hard carbon. Full article
(This article belongs to the Section Photochemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) XRD patterns; (<b>b</b>) Raman spectra of PPHC obtained at different temperatures.</p>
Full article ">Figure 2
<p>The XPS C1s spectra (<b>a</b>) and O1s spectra (<b>b</b>) of PPHC-1100.</p>
Full article ">Figure 3
<p>SEM images of PPHC-900 (<b>a</b>), PPHC-1100 (<b>c</b>), and PPHC-1300 (<b>e</b>); and TEM images of PPHC-900 (<b>b</b>), PPHC-1100 (<b>d</b>), and PPHC-1300 (<b>f</b>) (inset: SAED pattern).</p>
Full article ">Figure 4
<p>The first three charge–discharge cycles of (<b>a</b>) PPHC-900, (<b>b</b>) PPHC-1100, (<b>c</b>) PPHC-1300 obtained at 0.1 C. (<b>d</b>) Specific capacity distribution of PPHC electrodes below and above 0.1 V.</p>
Full article ">Figure 5
<p>Rate performance (<b>a</b>) and cycling stability (<b>b</b>) of PPHC electrodes.</p>
Full article ">Figure 6
<p>(<b>a</b>) CV curves of PPHC-1100 at 0.1 mV s<sup>−1</sup>; (<b>b</b>) relationship between the peak current and scan rate in logarithmic format; (<b>c</b>) capacitive contribution to charge storage at a scan rate of 1 mV s<sup>−1</sup>; (<b>d</b>) the contribution ratio of the capacitive and intercalated charge to capacity at different scan rates.</p>
Full article ">Figure 7
<p>(<b>a</b>) The first three cycles of charge–discharge curves of NVP//PPHC-1100 at 0.1 C in the voltage range of 1–4 V; (<b>b</b>) rate performance of NVP//PPHC-1100 at rates from 0.1 to 5 C; (<b>c</b>) cycling performance at a current density of 0.5 C for NVP//PPHC-1100.</p>
Full article ">
Back to TopTop