[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

Journals

Article Types

Countries / Regions

Search Results (126)

Search Parameters:
Keywords = lithium-oxygen battery

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 18755 KiB  
Article
Experimental Study on Thermal Runaway Characteristics of High-Nickel Ternary Lithium-Ion Batteries under Normal and Low Pressures
by Ye Jin, Di Meng, Chen-Xi Zhao, Jia-Ling Yu, Xue-Hui Wang and Jian Wang
Batteries 2024, 10(8), 287; https://doi.org/10.3390/batteries10080287 - 12 Aug 2024
Viewed by 676
Abstract
High-nickel (Ni) ternary lithium-ion batteries (LIBs) are widely used in low-pressure environments such as in the aviation industry, but their attribute of high energy density poses significant fire hazards, especially under low pressure where thermal runaway behavior is complex, thus requiring relevant experiments. [...] Read more.
High-nickel (Ni) ternary lithium-ion batteries (LIBs) are widely used in low-pressure environments such as in the aviation industry, but their attribute of high energy density poses significant fire hazards, especially under low pressure where thermal runaway behavior is complex, thus requiring relevant experiments. This study investigates the thermal runaway characteristics of LiNi0.8Mn0.1Co0.1O2 (NCM811) 18650 LIBs at different states of charge (SOCs) (75%, 100%) under various ambient pressures (101 kPa, 80 kPa, 60 kPa, 40 kPa). The results show that, as the pressure is decreased from 101 kPa to 40 kPa, the onset time of thermal runaway is extended by 28.2 s for 75% SOC and by 40.8 s for 100% SOC; accordingly, the onset temperature of thermal runaway increases by 19.3 °C for 75% SOC and by 33.5 °C for 100% SOC; the maximum surface temperature decreases by 70.8 °C for 75% SOC and by 68.2 °C for 100% SOC. The cell mass loss and loss rate slightly decrease with reduced pressure. However, ambient pressure has little impact on the time and temperature of venting as well as the voltage drop time. SEM/EDS analysis verifies that electrolyte evaporates faster under low pressure. Furthermore, the oxygen concentration is lower under low pressure, which consequently leads to a delay in thermal runaway. This study contributes to understanding thermal runaway characteristics of high-Ni ternary LIBs and provides guidance for their safe application in low-pressure aviation environments. Full article
(This article belongs to the Special Issue Advances in Lithium-Ion Battery Safety and Fire)
Show Figures

Figure 1

Figure 1
<p>Schematic of the experimental setup.</p>
Full article ">Figure 2
<p>Thermal runaway phenomena of cells at (<b>a</b>) 75% SOC and (<b>b</b>) 100% SOC under different pressures.</p>
Full article ">Figure 3
<p>Temperature evolution curves of cells at 75% SOC under (<b>a</b>) 80 kPa and under (<b>b</b>) 40 kPa. Stage I, II, and III represent heating stage, venting stage, and thermal runaway occurrence stage, respectively.</p>
Full article ">Figure 4
<p>Mechanism diagram of the thermal runaway process of NCM811 LIBs.</p>
Full article ">Figure 5
<p>Temperature evolution curves of cells at (<b>a</b>) 75% SOC and (<b>b</b>) 100% SOC under different pressures.</p>
Full article ">Figure 6
<p>(<b>a</b>) Time of venting and (<b>b</b>) temperature of venting under different pressures.</p>
Full article ">Figure 7
<p>(<b>a</b>) Onset time of thermal runaway and (<b>b</b>) onset temperature of thermal runaway of cells at 75% SOC and 100% SOC under different pressures.</p>
Full article ">Figure 8
<p>Maximum temperature of cells at 75% SOC and 100% SOC under different pressures.</p>
Full article ">Figure 9
<p>Voltage and temperature evolution curves during the experiment. Stages I, II, III, and IV represent the voltage stabilization stage, voltage rise stage, voltage decline stage, and voltage collapse stage, respectively.</p>
Full article ">Figure 10
<p>Voltage drop time of cells at 75% SOC and 100% SOC under different pressures.</p>
Full article ">Figure 11
<p>Characteristic times of voltage drop, venting and thermal runaway of cells at (<b>a</b>) 75% SOC and (<b>b</b>) 100% SOC.</p>
Full article ">Figure 12
<p>Mass loss and loss rate of cells after thermal runaway under different pressures.</p>
Full article ">Figure 13
<p>SEM micrographs with a scale bar of 3 micrometers of the raw cell and burnt cells at 100% SOC under different pressures.</p>
Full article ">Figure 14
<p>EDS elemental composition diagrams based on SEM micrographs with a scale bar of 10 micrometers of the raw cell and burnt cells at 100% SOC under different pressures.</p>
Full article ">
11 pages, 2769 KiB  
Article
A Nitrogen/Oxygen Dual-Doped Porous Carbon with High Catalytic Conversion Ability toward Polysulfides for Advanced Lithium–Sulfur Batteries
by Xiaoyan Shu, Yuanjiang Yang, Zhongtang Yang, Honghui Wang and Nengfei Yu
C 2024, 10(3), 67; https://doi.org/10.3390/c10030067 - 30 Jul 2024
Viewed by 669
Abstract
Lithium–sulfur batteries (LSBs) have attracted widespread attention due to their high theoretical energy density and low cost. However, their development has been constrained by the shuttle effect of lithium polysulfides and their slow reaction kinetics. In this work, a nitrogen/oxygen dual-doped porous carbon [...] Read more.
Lithium–sulfur batteries (LSBs) have attracted widespread attention due to their high theoretical energy density and low cost. However, their development has been constrained by the shuttle effect of lithium polysulfides and their slow reaction kinetics. In this work, a nitrogen/oxygen dual-doped porous carbon (N/O-PC) was synthesized by annealing the precursor of zeolitic imidazolate framework-8 grown in situ on MWCNTs (ZIF-8/MWCNTs). Then, the N/O-PC composite served as an efficient host for LSBs through chemical adsorption and providing catalytic conversion sites of polysulfides. Moreover, the interconnected porous carbon-based structure facilitates electron and ion transfer. Thus, the S/N/O-PC cathode exhibits high cycling stability (a stable capacity of 685.9 mA h g−1 at 0.2 C after 100 cycles). It also demonstrates excellent rate performance with discharge capacities of 1018.2, 890.2, 775.1, 722.7, 640.4, and 579.6 mAh g−1 at 0.2, 0.5, 1.0, 2.0, 3.0, and 5.0 C, respectively. This work provides an effective strategy for designing and developing high energy density, long cycle life LSBs. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of fabrication processes of S/N/O–PC.</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM image, (<b>b</b>,<b>c</b>) TEM images of MWCNTs. (<b>d</b>) Low magnification, and (<b>e</b>) high magnification SEM image of N/O–PC, and (<b>f</b>) TEM image of S/N/O–PC.</p>
Full article ">Figure 3
<p>(<b>a</b>) XRD patterns of ZIF-8, ZIF-8/MWCNTs and N/O–PC. (<b>b</b>) Nitrogen adsorption–desorption isotherms for N/O–PC. (<b>c</b>) XPS high resolution of O 1s. (<b>d</b>) XPS high resolution of N 1s.</p>
Full article ">Figure 4
<p>(<b>a</b>) XRD patterns of crystalline S, S/N/O–PC. (<b>b</b>) TGA curves of crystalline S, S/N/O–PC. (<b>c</b>) Nitrogen adsorption–desorption isotherms for the S/N/O–PC. (<b>d</b>) The pore size distribution for S/N/O-PC.</p>
Full article ">Figure 5
<p>(<b>a</b>) CV curve of S/N/O-PC and S/C cathodes at a scan rate of 0.1 mV s<sup>−1</sup>. (<b>b</b>) Galvanostatic charge/discharge profiles of S/N/O−PC and S/C cathodes at 0.2 C. (<b>c</b>) Cycle performances of S/N/O−PC and S/C cathodes at 0.2 C. (<b>d</b>) Discharge and (<b>e</b>) charge profiles of S/N/O–PC and S/C cathodes at a scan rate of 0.1 mV s<sup>−1</sup>, indicating the overpotentials of solid–liquid phase conversion between soluble LiPSs and insoluble Li<sub>2</sub>S<sub>2</sub>/Li<sub>2</sub>S. (<b>f</b>) Schematic illustration of nitrogen and oxygen heteroatoms on N/O−PC catalyzed lithium polysulfides.</p>
Full article ">Figure 6
<p>(<b>a</b>) The rate performance of S/N/O−PC. (<b>b</b>,<b>c</b>) Charge/discharge curves of the S/N/O−PC and S/C cathodes at various rates, respectively.</p>
Full article ">
26 pages, 7864 KiB  
Review
Advancements in Lithium–Oxygen Batteries: A Comprehensive Review of Cathode and Anode Materials
by Jing Guo, Xue Meng, Qing Wang, Yahui Zhang, Shengxue Yan and Shaohua Luo
Batteries 2024, 10(8), 260; https://doi.org/10.3390/batteries10080260 - 23 Jul 2024
Viewed by 1009
Abstract
As modern society continues to advance, the depletion of non-renewable energy sources (such as natural gas and petroleum) exacerbates environmental and energy issues. The development of green, environmentally friendly energy storage and conversion systems is imperative. The energy density of commercial lithium-ion batteries [...] Read more.
As modern society continues to advance, the depletion of non-renewable energy sources (such as natural gas and petroleum) exacerbates environmental and energy issues. The development of green, environmentally friendly energy storage and conversion systems is imperative. The energy density of commercial lithium-ion batteries is approaching its theoretical limit, and even so, it struggles to meet the rapidly growing market demand. Lithium–oxygen batteries have garnered significant attention from researchers due to their exceptionally high theoretical energy density. However, challenges such as poor electrolyte stability, short cycle life, low discharge capacity, and high overpotential arise from the sluggish kinetics of the oxygen reduction reaction (ORR) during discharge and the oxygen evolution reaction (OER) during charging. This article elucidates the fundamental principles of lithium–oxygen batteries, analyzes the primary issues currently faced, and summarizes recent research advancements in air cathodes and anodes. Additionally, it proposes future directions and efforts for the development of lithium–air batteries. Full article
Show Figures

Figure 1

Figure 1
<p>Components and technical challenges of rechargeable Li–O<sub>2</sub> battery.</p>
Full article ">Figure 2
<p>(<b>a</b>) SEI film evolution of L–Si and F–L–Si anodes in Li-ion O<sub>2</sub> batteries during the discharge–charge cycle and their resistibility towards the O<sub>2</sub> crossover effect on Si anodes. (<b>b</b>) The specific capacities along with discharge and charge terminal voltages against cycle number of Li-ion O<sub>2</sub> batteries with F–L–Si anodes. (<b>c</b>) The corresponding electrochemical impedance spectrum (EIS) of the discharged Li-ion O<sub>2</sub> batteries with F–L–Si anodes [<a href="#B22-batteries-10-00260" class="html-bibr">22</a>]. Copyright 2016, The Royal Society of Chemistry. (<b>d</b>) Schematics of the electrode preparation and Li-ion oxygen battery assembling [<a href="#B23-batteries-10-00260" class="html-bibr">23</a>]. Copyright 2021, Elsevier. (<b>e</b>) Schematic representation of the reactions occurring at the electrode/electrolyte interphase of a Li<sub>x</sub>Sn–C/Pyr<sub>14</sub>TFSI–LiTFSI/O<sub>2</sub> cell in the OCV condition under an oxygen atmosphere [<a href="#B24-batteries-10-00260" class="html-bibr">24</a>]. Copyright 2015, American Chemical Society.</p>
Full article ">Figure 3
<p>(<b>a</b>) Digital images of Li pellets with and without BA treatment after exposure to air (38 RH%) at different times. (<b>b</b>) Cycling performance of Li|Li symmetric cells with (red line) and without (black line) BA under O<sub>2</sub> atmosphere [<a href="#B33-batteries-10-00260" class="html-bibr">33</a>]. Copyright 2018, Wiley–VCH. (<b>c</b>) A typical fabrication. (<b>d</b>) SEM image of gel electrolyte coated on the Li wire. (<b>e</b>) Photograph of a fiber-shaped Li–air battery [<a href="#B34-batteries-10-00260" class="html-bibr">34</a>]. Copyright 2016, Wiley–VCH. (<b>f</b>) Optimized solvation structures in SE and BSE. (<b>g</b>) ESP comparison of typical solvation structures in SE and BSE. (<b>h</b>) Mechanism of LiTFA on the regulation of solvation structures and interfacial reactions on Li anode [<a href="#B35-batteries-10-00260" class="html-bibr">35</a>]. Copyright 2023, Wiley–VCH.</p>
Full article ">Figure 4
<p>(<b>a</b>) Li addition and removal energies on Li metal and Li<sub>3</sub>P(001) surfaces. Structures of (<b>b</b>) Li(001) and (<b>c</b>) Li<sub>3</sub>P(001) surfaces. (<b>d</b>) Density of states of bcc Li metal and amorphous Li<sub>3</sub>P. EF represents the Fermi level [<a href="#B43-batteries-10-00260" class="html-bibr">43</a>]. Copyright 2018, American Chemical Society. (<b>e</b>) SEM image of the protected anode surface. (<b>f</b>) The discharge–charge voltage profile over 550 cycles. The inset shows the capacity versus the number of cycles. (<b>g</b>) TEM image of a discharged cathode sample. The inset diffraction patterns show crystallinity corresponding to monoclinic lithium peroxide, Li<sub>2</sub>O<sub>2</sub> [<a href="#B44-batteries-10-00260" class="html-bibr">44</a>]. Copyright 2018, Springer Nature. (<b>h</b>) XPS images showing the distribution of Li<sub>2</sub>CO<sub>3</sub> compositions on the surface of Li metal anodes with 5 wt.% of poly(ethylene oxide) (PEO) in gel polymer (PG) from the pretreated LOB cells at 0.2 mA cm<sup>−2</sup> to 5 V of charge cutoff voltage under Ar or O<sub>2</sub> atmosphere. (<b>i</b>) Formation of polymer-supported solid–electrolyte interface (PS–SEI) layer on Li metal surface by electrochemical precharging to 5 V under oxygen atmosphere [<a href="#B45-batteries-10-00260" class="html-bibr">45</a>]. Copyright 2021, American Chemical Society. (<b>j</b>) Schematic Illustration of Suppressing I<sub>3</sub><sup>−</sup> Shutting with the MXene–Modified Separator in Li–O<sub>2</sub> Batteries [<a href="#B46-batteries-10-00260" class="html-bibr">46</a>]. Copyright 2021, American Chemical Society.</p>
Full article ">Figure 5
<p>(<b>a</b>) Schematic of the cathode based on carbon nanosphere clusters in working Li–O<sub>2</sub> batteries [<a href="#B65-batteries-10-00260" class="html-bibr">65</a>]. Copyright 2021, Elsevier. (<b>b</b>–<b>d</b>) The orthogonal channels are large-area uniform distribution (<b>c</b>) with uniform pore size (~5 μm) (<b>d</b>), and there are slight changes in the local mismatch of pore arrangement orientations like grain boundaries in polycrystalline materials (<b>b</b>). (<b>e</b>) Literature surveys of other low-tortuosity cathodes, Ni/Fe-based cathodes, freestanding cathodes, and noble-metal-based cathodes focus on limited areal capacity and long cycling [<a href="#B66-batteries-10-00260" class="html-bibr">66</a>]. Copyright 2023, American Chemical Society. (<b>f</b>,<b>g</b>) New strategy to obtain well-defined RF, carbon, or SiO<sub>2</sub> rings using patchy droplet templates. (<b>h</b>) SEM images of self-standing HCRE. (<b>i</b>) Galvanostatic discharge profiles of LOBs [<a href="#B67-batteries-10-00260" class="html-bibr">67</a>]. Copyright 2023, American Chemical Society.</p>
Full article ">Figure 6
<p>(<b>a</b>) The XRD patterns of Pt nanocrystals and PtIr multipods. (<b>b</b>) The galvanostatic profiles for Pt and PtIr electrodes during the discharging and charging process with a limited capacity of 1000 mAh g<sup>−1</sup> at 0.1 A g<sup>−1</sup> [<a href="#B76-batteries-10-00260" class="html-bibr">76</a>]. Copyright 2021, Wiley–VCH. (<b>c</b>) Raman spectra of discharge product on Ir-rGO for first and second discharges. (<b>d</b>) Schematic showing lattice match between LiO<sub>2</sub> and Ir<sub>3</sub>Li that may be responsible for the LiO<sub>2</sub> discharge product found on the Ir-rGO cathode. (<b>e</b>) DFT electronic band structure (<b>left</b>) and density of states (DOS) plot (<b>right</b>) of ferromagnetic bulk crystalline LiO<sub>2</sub> close to the Fermi level (Ef) based on a spin-polarized calculation with electronic spin-up and spin-down states shown [<a href="#B77-batteries-10-00260" class="html-bibr">77</a>]. Copyright 2016, Springer Nature. (<b>f</b>) HAADF–STEM images of Ru0.3 SAs-NC (Ru single atoms are marked with red circles). (<b>g</b>) Corresponding EDS maps reveal the homogeneous distribution of Ru and N within the carbon support of Ru0.3 SAs-NC. (<b>h</b>) Gibbs free energy diagrams at 2.97 V for the discharge−charge reactions on the active surface of pyrolyzed ZIF-8, Ru0.1 SAs-NC, and Ru0.3 SAs-NC [<a href="#B81-batteries-10-00260" class="html-bibr">81</a>]. Copyright 2020, American Chemical Society.</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic representation of a rechargeable Li/O<sub>2</sub> battery. (<b>b</b>) Variation of discharge capacity with cycle number for several porous Mn-based [<a href="#B87-batteries-10-00260" class="html-bibr">87</a>]. Copyright 2008, Wiley–VCH. (<b>c</b>) The schematic diagram of the formation and decomposition of discharge product Li<sub>2</sub>O<sub>2</sub> on the N–Ti<sub>3</sub>C<sub>2</sub>(H) surface during the discharge and charge process, respectively [<a href="#B101-batteries-10-00260" class="html-bibr">101</a>]. Copyright 2023, Wiley–VCH. Top (<b>d</b>) and side (<b>e</b>) views of the Ti<sub>2</sub>C MXene monolayer. The capitals in the figure are the possible adsorption sites. Green and gray atoms are Ti and C, respectively. (<b>f</b>) Calculated total density of states (<b>left</b>) and projected density of states of Ti 3d (<b>right</b>) for Ti<sub>2</sub>C, Ti<sub>2</sub>CO<sub>2</sub>, Ti<sub>2</sub>CF<sub>2</sub>, and Ti<sub>2</sub>C(OH)<sub>2</sub>. The vertical dashed line at E = 0 eV represents the Fermi energy [<a href="#B102-batteries-10-00260" class="html-bibr">102</a>]. Copyright 2019, American Chemical Society. (<b>g</b>) Schematic diagram of the reaction during the cycling process in LOBs (* represents adsorptive state). [<a href="#B103-batteries-10-00260" class="html-bibr">103</a>]. Copyright 2021, Elsevier.</p>
Full article ">Figure 8
<p>(<b>a</b>) Schematic illustration for preparation of 3DOM-LFO catalyst and structure of the rechargeable Li–O<sub>2</sub> batteries. (<b>b</b>) FESEM images of 3DOM-LFO after calcination at 600 °C for 3 h. Inset in (<b>a</b>): magnified FESEM image [<a href="#B109-batteries-10-00260" class="html-bibr">109</a>]. Copyright 2014, The Royal Society of Chemistry. (<b>c</b>) Low- and (<b>d</b>) High-magnification TEM images of PNT–LSM [<a href="#B110-batteries-10-00260" class="html-bibr">110</a>]. Copyright 2013, Wiley–VCH. (<b>e</b>) XRD patterns of LaCo<sub>x</sub>Mn<sub>1−x</sub>O<sub>3−σ</sub> (x = 1, 0.75, and 0) samples. (<b>f</b>) Li–O<sub>2</sub> batteries with SP and LaCo<sub>x</sub>Mn<sub>1−x</sub>O<sub>3−σ</sub> electrode (x = 1, 0.75, and 0) for (<b>a</b>) galvanostatic discharge/charge curves at a current density of 200 mA g<sup>−1</sup>. (<b>g</b>) Mechanism of oxygen vacancy for ORR/OER processes occurring on the surface of LaCo<sub>0.75</sub>Mn<sub>0.25</sub>O<sub>3−σ</sub> catalyst during charging/discharging in Li–O<sub>2</sub> Battery [<a href="#B111-batteries-10-00260" class="html-bibr">111</a>]. Copyright 2020, American Chemical Society. (<b>h</b>) Full discharge and charge curves of the cell with LNCO cathode (red curves) and with SPC cathode (blue curves). (<b>i</b>) Gibbs free energy diagrams of the LNCO electrode reactions. The insets are the optimized structures of the LNCO (1 1 0) with adsorbates at corresponding discharging steps. (<b>j</b>) Schematic illustration of the discharge pathway on the LNCO cathode surface [<a href="#B112-batteries-10-00260" class="html-bibr">112</a>]. Copyright 2023, Elsevier.</p>
Full article ">Figure 9
<p>(<b>a</b>) HAADF–STEM image and EDS element mapping images of one segment of an HEA–PtPdIrRuAg SNR. Pt (blue), Pd (red), Ag (magenta), Ir (orange), and Ru (green) signals. (<b>b</b>) Possibility of an atomic pair in different HEAs. The color bar represents the possibility of the bonding between different metals, which ranges from 0 to 1. (<b>c</b>) ORR performances of quinary PtPdIrRuAg SNRs [<a href="#B113-batteries-10-00260" class="html-bibr">113</a>]. Copyright 2023, American Chemical Society. (<b>d</b>) Schematic diagram of electron transfer in HEAPtIr. (<b>e</b>) Orbital interactions between LiO<sub>2</sub>/Li<sub>2</sub>O<sub>2</sub> and catalysts with different d-band centers and (<b>f</b>) the corresponding catalytic effects [<a href="#B115-batteries-10-00260" class="html-bibr">115</a>]. Copyright 2023, Wiley–VCH. (<b>g</b>) The single-crystal structure of Na–Pb–MOF. (<b>h</b>) Comparison of overall overpotentials of Na–Pb–MOF and the reported MOF-based electrocatalysts [<a href="#B116-batteries-10-00260" class="html-bibr">116</a>]. Copyright 2023, American Chemical Society. (<b>i</b>) Schematic illustration of the role of the redox mediator (RM) in a Li–O<sub>2</sub> battery system made using a hierarchical CNT fibril electrode [<a href="#B32-batteries-10-00260" class="html-bibr">32</a>]—copyright 2014, Wiley–VCH.</p>
Full article ">
16 pages, 3731 KiB  
Article
Experimental Investigation on Thermal Runaway of Lithium-Ion Batteries under Low Pressure and Low Temperature
by Di Meng, Jingwen Weng and Jian Wang
Batteries 2024, 10(7), 243; https://doi.org/10.3390/batteries10070243 - 6 Jul 2024
Viewed by 866
Abstract
Understanding the thermal runaway mechanism of lithium-ion batteries under low pressure and low temperature is paramount for their application and transportation in the aviation industry. This work investigated the coupling effects of ambient pressure (100 kPa, 70 kPa, 40 kPa) and ambient temperature [...] Read more.
Understanding the thermal runaway mechanism of lithium-ion batteries under low pressure and low temperature is paramount for their application and transportation in the aviation industry. This work investigated the coupling effects of ambient pressure (100 kPa, 70 kPa, 40 kPa) and ambient temperature (−15 °C, 0 °C, 25 °C) on thermal behaviors in an altitude temperature chamber. The experimental results indicate that lowering ambient pressure and temperature could attenuate the thermal runaway intensity, which is mainly attributable to the reduction in oxygen concentration and the increase in heat loss. Such a dual effect leads to the maximum temperature decreasing from 811.9 °C to 667.5 °C, and the maximum temperature rate declines up to 2.6 times. Correspondingly, the whole thermal runaway process is deferred, the total time increases from 370 s to 503 s, and the time interval, Δt, from safety venting gains by 32.3% as the ambient pressure and temperature decrease. This work delivers an in-depth understanding of the thermal characteristics under low pressure and low temperature and provides meritorious guidance for the safety of cell transportation in aviation. Full article
(This article belongs to the Section Battery Performance, Ageing, Reliability and Safety)
Show Figures

Figure 1

Figure 1
<p>Schematic of the experimental setup.</p>
Full article ">Figure 2
<p>Thermal runaway phenomenon of batteries in aviation conditions: (<b>a</b>) 101 kPa, 25 °C, (<b>b</b>) 70 kPa, 0 °C, (<b>c</b>) 40 kPa, −15 °C.</p>
Full article ">Figure 3
<p>The cell surface temperature variations: (<b>a</b>) typical surface temperature at 25 °C under different ambient pressures, (<b>b</b>) typical surface temperature at 70 kPa with different ambient temperatures.</p>
Full article ">Figure 3 Cont.
<p>The cell surface temperature variations: (<b>a</b>) typical surface temperature at 25 °C under different ambient pressures, (<b>b</b>) typical surface temperature at 70 kPa with different ambient temperatures.</p>
Full article ">Figure 4
<p>The critical temperature of voltage drop (<span class="html-italic">T<sub>v-drop</sub></span>), safety venting (<span class="html-italic">T<sub>venting</sub></span>), thermal runaway onset temperature (<span class="html-italic">T<sub>onset</sub></span>), and maximum surface temperature (<span class="html-italic">T<sub>max</sub></span>) under different conditions.</p>
Full article ">Figure 5
<p>Temperature rate evolution under different conditions: (<b>a</b>) temperature rate at 101 kPa, 25 °C; (<b>b</b>) temperature rate at 40 kPa, −15 °C.</p>
Full article ">Figure 6
<p>The axial jet fire temperature distributions during thermal runaway. (<b>a</b>) 101 kPa, 25 °C, (<b>b</b>) 40 kPa, −15 °C.</p>
Full article ">Figure 7
<p>Chemical analysis of the battery before and after thermal runaway. (<b>a</b>) fresh battery; (<b>b</b>) 101 kPa, 25 °C; (<b>c</b>) 40 kPa, −15 °C.</p>
Full article ">Figure 8
<p>Mass loss of cells after thermal runaway.</p>
Full article ">
12 pages, 2693 KiB  
Article
Synthesis of High-Entropy Perovskite Hydroxides as Bifunctional Electrocatalysts for Oxygen Evolution Reaction and Oxygen Reduction Reaction
by Sangwoo Chae, Akihito Shio, Tomoya Kishida, Kosuke Furutono, Yumi Kojima, Gasidit Panomsuwan and Takahiro Ishizaki
Materials 2024, 17(12), 2963; https://doi.org/10.3390/ma17122963 - 17 Jun 2024
Viewed by 559
Abstract
Oxygen reduction reaction (ORR) and oxygen evolutionc reaction (OER) are important chemical reactions for a rechargeable lithium–oxygen battery (LOB). Recently, high-entropy alloys and oxides have attracted much attention because they showed good electrocatalytic performance for oxygen evolution reaction (OER) and/or oxygen reduction reaction [...] Read more.
Oxygen reduction reaction (ORR) and oxygen evolutionc reaction (OER) are important chemical reactions for a rechargeable lithium–oxygen battery (LOB). Recently, high-entropy alloys and oxides have attracted much attention because they showed good electrocatalytic performance for oxygen evolution reaction (OER) and/or oxygen reduction reaction (ORR). In this study, we aimed to synthesize and characterize CoSn(OH)6 and two types of high-entropy perovskite hydroxides, that is, (Co0.2Cu0.2Fe0.2Mn0.2Mg0.2)Sn(OH)6 (CCFMMSOH) and (Co0.2Cu0.2Fe0.2Mn0.2Ni0.2)Sn(OH)6 (CCFMNSOH). TEM observation and XRD measurements revealed that the high-entropy hydroxides CCFMMSOH and CCFMNSOH had cubic crystals with sides of approximately 150–200 nm and crystal structures similar to those of perovskite-type CSOH. LSV measurement results showed that the high-entropy hydroxides CCFMMSOH and CCFMNSOH showed bifunctional catalytic functions for the ORR and OER. CCFMNSOH showed better catalytic performance than CCFMMSOH. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of (<b>a</b>) CSOH, (<b>b</b>) CCFMMSOH, and (<b>c</b>) CCFMNSOH, and (<b>d</b>) enlarged XRD patterns of CSOH, CCFMMSOH, and CCFMNSOH at 2<span class="html-italic">θ</span> = 22.5° to 23.2°.</p>
Full article ">Figure 2
<p>(<b>a</b>) TEM image, (<b>b</b>) STEM image, and elemental mapping images of CCFMMSOH: (<b>c</b>) Co, (<b>d</b>) Cu, (<b>e</b>) Fe, (<b>f</b>) Mn, (<b>g</b>) Mg, (<b>h</b>) Sn, and (<b>i</b>) O.XRD patterns of (<b>a</b>) Co-NC@CNT, (<b>b</b>) Ni-NC@CNT, and (<b>c</b>) CSCNT.</p>
Full article ">Figure 3
<p>(<b>a</b>) TEM image, (<b>b</b>) STEM image, and elemental mapping images of CCFMNSOH: (<b>c</b>) Co, (<b>d</b>) Cu, (<b>e</b>) Fe, (<b>f</b>) Mn, (<b>g</b>) Ni, (<b>h</b>) Sn, and (<b>i</b>) O.</p>
Full article ">Figure 4
<p>High-resolution XPS O1s spectra and deconvolution of (<b>a</b>) CSOH, (<b>b</b>) CCFMMSOH, and (<b>c</b>) CCFMNSOH.</p>
Full article ">Figure 5
<p>(<b>a</b>) N<sub>2</sub> adsorption-desorption plots and (<b>b</b>) pore diameter distribution curves of CSOH, CCFMMSOH, and CCFMNSOH.</p>
Full article ">Figure 6
<p>(<b>a</b>) ORR curves of CSOH, CCFMMSOH, CCFMNSOH, and Pt/C 20 wt% in O<sub>2</sub>-saturated 0.1 M KOH aqueous solution. Change in (<b>b</b>) electron transfer numbers (<span class="html-italic">n</span>), and (<b>c</b>) HO<sub>2</sub><sup>−</sup> yields of CSOH, CCFMMSOH, CCFMNSOH, and Pt/C 20 wt.% as a function of potential.</p>
Full article ">Figure 7
<p>(<b>a</b>) OER curves of CSOH, CCFMMSOH, CCFMNSOH, and RuO<sub>2</sub> in O<sub>2</sub>-saturated 0.1 M KOH aqueous solution. (<b>b</b>) Tafel plots of the composite samples synthesized at CSOH, CCFMMSOH, CCFMNSOH, and commercial RuO<sub>2</sub>.</p>
Full article ">
16 pages, 4656 KiB  
Article
Using Sandwiched Silicon/Reduced Graphene Oxide Composites with Dual Hybridization for Their Stable Lithium Storage Properties
by Yuying Yang, Rui Zhang, Qiang Zhang, Liu Feng, Guangwu Wen, Lu-Chang Qin and Dong Wang
Molecules 2024, 29(10), 2178; https://doi.org/10.3390/molecules29102178 - 7 May 2024
Viewed by 720
Abstract
Using silicon/reduced graphene oxide (Si/rGO) composites as lithium-ion battery (LIB) anodes can effectively buffer the volumetric expansion and shrinkage of Si. Herein, we designed and prepared Si/rGO-b with a sandwiched structure, formed by a duple combination of ammonia-modified silicon (m-Si) nanoparticles (NP) with [...] Read more.
Using silicon/reduced graphene oxide (Si/rGO) composites as lithium-ion battery (LIB) anodes can effectively buffer the volumetric expansion and shrinkage of Si. Herein, we designed and prepared Si/rGO-b with a sandwiched structure, formed by a duple combination of ammonia-modified silicon (m-Si) nanoparticles (NP) with graphene oxide (GO). In the first composite process of m-Si and GO, a core–shell structure of primal Si/rGO-b (p-Si/rGO-b) was formed. The amino groups on the m-Si surface can not only hybridize with the GO surface to fix the Si particles, but also form covalent chemical bonds with the remaining carboxyl groups of rGO to enhance the stability of the composite. During the electrochemical reaction, the oxygen on the m-Si surface reacts with lithium ions (Li+) to form Li2O, which is a component of the solid–electrolyte interphase (SEI) and is beneficial to buffering the volume expansion of Si. Then, the p-Si/rGO-b recombines with GO again to finally form a sandwiched structure of Si/rGO-b. Covalent chemical bonds are formed between the rGO layers to tightly fix the p-Si/rGO-b, and the conductive network formed by the reintroduced rGO improves the conductivity of the Si/rGO-b composite. When used as an electrode, the Si/rGO-b composite exhibits excellent cycling performance (operated stably for more than 800 cycles at a high-capacity retention rate of 82.4%) and a superior rate capability (300 mA h/g at 5 A/g). After cycling, tiny cracks formed in some areas of the electrode surface, with an expansion rate of only 27.4%. The duple combination of rGO and the unique sandwiched structure presented here demonstrate great effectiveness in improving the electrochemical performance of alloy-type anodes. Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Schematic illustration of the preparation of the Si/rGO-b composite. (<b>b</b>,<b>c</b>) SEM images of the Si/rGO-a and Si/rGO-b composites.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>c</b>) TEM images of Si, m-Si, and the Si/rGO-b composite. (<b>d</b>–<b>f</b>) HRTEM images of Si, m-Si, and the Si/rGO-b composite. Insets are magnified HRTEM images of the marked regions. (<b>g</b>) STEM image and (<b>h</b>–<b>j</b>) EDS mapping of the Si/rGO-b composite.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) XPS data of Si/rGO-b Si2p and C1s, respectively. (<b>c</b>) Raman spectra of rGO and Si/rGO-b. (<b>d</b>) XRD of Si, Si/rGO-a, and Si/rGO-b. (<b>e</b>) The pore size distribution of Si, Si/rGO-a and Si/rGO-b. Inset is the nitrogen adsorption/desorption isotherms of Si, Si/rGO-a and Si/rGO-b. (<b>f</b>) Conductivity of Si, Si/rGO-a and Si/rGO-b.</p>
Full article ">Figure 4
<p>Electrochemical characterization of electrodes. (<b>a</b>) CV curves of the Si/rGO-b composite at a scan rate of 0.1 mV/s within the potential range of 0.01–2.0 V. (<b>b</b>) Charge/discharge curves at different current densities of Si/rGO-b (0.1 C–5 C, 1 C = 1 A/g). (<b>c</b>) Rate performance of Si/rGO-b, Si/rGO-a and pure Si. (<b>d</b>) Comparison of lithium storage properties with previous work on Si-based anodes ([<a href="#B5-molecules-29-02178" class="html-bibr">5</a>,<a href="#B25-molecules-29-02178" class="html-bibr">25</a>,<a href="#B53-molecules-29-02178" class="html-bibr">53</a>,<a href="#B54-molecules-29-02178" class="html-bibr">54</a>,<a href="#B55-molecules-29-02178" class="html-bibr">55</a>,<a href="#B56-molecules-29-02178" class="html-bibr">56</a>,<a href="#B57-molecules-29-02178" class="html-bibr">57</a>,<a href="#B58-molecules-29-02178" class="html-bibr">58</a>,<a href="#B59-molecules-29-02178" class="html-bibr">59</a>], our work: red star). (<b>e</b>) Capacity and CE of Si/rGO-b (800 cycles), Si/rGO-a (500 cycles) and Si (200 cycles) at 1 C.</p>
Full article ">Figure 5
<p>(<b>a</b>) Nyquist plots of EIS data for the Si/rGO-b, Si/rGO-a and Si electrodes after 50 discharge/charge cycles. Inset shows the equivalent circuit for the electrodes. (<b>b</b>) <span class="html-italic">R<sub>ct</sub></span> of Si/rGO-b, Si/rGO-a and Si electrodes before and after 50 cycles. (<b>c</b>) Lithium ion diffusion rate of Si/rGO-b, Si/rGO-a and Si. Inset is the corresponding slope <span class="html-italic">σ</span> value of three types of electrodes. (<b>d</b>) CV curves at different scan rates from 0.2 to 1.0 mV/s of the Si/rGO-b electrode. (<b>e</b>) <span class="html-italic">b</span> value of the Si/rGO electrode via the relationship of log(<span class="html-italic">i</span>) vs. log(<span class="html-italic">v</span>). (<b>f</b>) Ratio of the capacitive contribution at different scan rates.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>c</b>) SEM images of the (<b>a</b>) Si/rGO-b, (<b>b</b>) Si/rGO-a and (<b>c</b>) Si electrodes before and after 200 cycles, respectively. (<b>d</b>–<b>i</b>) Cross-sectional view of the Si/rGO-b (<b>d</b>,<b>g</b>), Si/rGO-a (<b>e</b>,<b>h</b>) and (<b>f</b>,<b>i</b>) Si electrodes before and after 200 cycles, respectively. (<b>j</b>) The thickness variation of Si/rGO-b, Si/rGO-a and Si electrode before and after 200 cycles. (<b>k</b>) The corresponding percentage change in the thickness of the three types of electrodes.</p>
Full article ">Figure 7
<p>Morphological changes of the active materials Si, Si/rGO-a and Si/rGO-b during lithiation/delithiation processes.</p>
Full article ">
19 pages, 11634 KiB  
Article
A Study on the Microstructure Regulation Effect of Niobium Doping on LiNi0.88Co0.05Mn0.07O2 and the Electrochemical Performance of the Composite Material under High Voltage
by Xinrui Xu, Junjie Liu, Bo Wang, Jiaqi Wang, Yunchang Wang, Weisong Meng and Feipeng Cai
Materials 2024, 17(9), 2127; https://doi.org/10.3390/ma17092127 - 30 Apr 2024
Viewed by 771
Abstract
High-nickel ternary materials are currently the most promising lithium battery cathode materials due to their development and application potential. Nevertheless, these materials encounter challenges like cation mixing, lattice oxygen loss, interfacial reactions, and microcracks. These issues are exacerbated at high voltages, compromising their [...] Read more.
High-nickel ternary materials are currently the most promising lithium battery cathode materials due to their development and application potential. Nevertheless, these materials encounter challenges like cation mixing, lattice oxygen loss, interfacial reactions, and microcracks. These issues are exacerbated at high voltages, compromising their cyclic stability and safety. In this study, we successfully prepared Nb5+-doped high-nickel ternary cathode materials via a high-temperature solid-phase method. We investigated the impact of Nb5+ doping on the microstructure and electrochemical properties of LiNi0.88Co0.05Mn0.07O2 ternary cathode materials by varying the amount of Nb2O5 added. The experimental results suggest that Nb5+ doping does not alter the crystal structure but modifies the particle morphology, yielding radially distributed, elongated, rod-like structures. This morphology effectively mitigates the anisotropic volume changes during cycling, thereby bolstering the material’s cyclic stability. The material exhibits a discharge capacity of 224.4 mAh g−1 at 0.1C and 200.3 mAh g−1 at 1C, within a voltage range of 2.7 V–4.5 V. Following 100 cycles at 1C, the capacity retention rate maintains a high level of 92.9%, highlighting the material’s remarkable capacity retention and cyclic stability under high-voltage conditions. The enhancement of cyclic stability is primarily due to the synergistic effects caused by Nb5+ doping. Nb5+ modifies the particle morphology, thereby mitigating the formation of microcracks. The formation of high-energy Nb-O bonds prevents oxygen precipitation at high voltages, minimizes the irreversibility of the H2–H3 phase transition, and thereby enhances the stability of the composite material at high voltages. Full article
(This article belongs to the Topic Advanced Nanomaterials for Lithium-Ion Batteries)
Show Figures

Figure 1

Figure 1
<p>SEM images of (<b>a</b>) NCM88-0 (<b>b</b>), NCM88-0.3Nb (<b>c</b>), NCM88-0.5Nb and (<b>d</b>), NCM88-1Nb.</p>
Full article ">Figure 2
<p>EDS-mapping image of Nb-doped samples: (<b>a</b>) NCM88-0.3Nb, (<b>b</b>) NCM88-0.5 Nb, and (<b>c</b>) NCM88-1Nb.</p>
Full article ">Figure 3
<p>SEM images of NCM88 cathode material after ion milling: (<b>a</b>) NCM88-0, (<b>b</b>) NCM88-0.5Nb, and (<b>c</b>) EDS of NCM88-0.5 Nb.</p>
Full article ">Figure 4
<p>XRD pattern ofNCM88-0, NCM88-0.3Nb, NCM88-0.5Nb and NCM88-1Nb.</p>
Full article ">Figure 5
<p>XPS profile of NCM88: (<b>a</b>) NCM88-0 and (<b>b</b>,<b>c</b>) NCM88-0.5Nb.</p>
Full article ">Figure 6
<p>(<b>a</b>) N<sub>2</sub> adsorption–desorption curve, and (<b>b</b>) pore size distribution of the NCM88-0.5Nb sample.</p>
Full article ">Figure 7
<p>The first charge-discharge curve of NCM88 at a high cut-off voltage of 2.7 V–4.5 V.</p>
Full article ">Figure 8
<p>Electrochemical performance of NCM88 at high cut-off voltages of 2.7 V–4.5 V: (<b>a</b>) cycle performance of NCM88, (<b>b</b>) rate performance of NCM88, (<b>c</b>) capacity distribution of NCM88-0 at different magnifications, and (<b>d</b>) capacity distribution of NCM88-0.5Nb at different magnifications.</p>
Full article ">Figure 9
<p>CV curves for the first three turns at a cut-off voltage of 2.7 V–4.5 V: (<b>a</b>,<b>b</b>) NCM88-0 and (<b>c</b>,<b>d</b>) NCM88-0.5Nb.</p>
Full article ">Figure 10
<p>dQ dV<sup>−1</sup>-V curves for the 1st, 50th, and 100th cycles: (<b>a</b>) NCM88-0, and (<b>b</b>) NCM88-0.5Nb.</p>
Full article ">Figure 11
<p>Electrochemical impedance fitting of NCM88 after 1, 50, and 100 cycles at a high cut-off voltage of 2.7 V–4.5 V: (<b>a</b>) NCM88-0, (<b>b</b>) NCM88-0.5Nb, and (<b>c</b>) analog circuit diagram.</p>
Full article ">Figure 12
<p>DSC plots of NCM88-0 and NCM88-0.5Nb at high cut-off voltages of 2.7 V–4.5 V.</p>
Full article ">Figure 13
<p>SEM image after 100 charge–discharge cycles at a cut-off voltage of 2.7 V–4.5 V. (<b>a</b>) NCM88-0 (<b>b</b>) NCM88-0.5Nb.</p>
Full article ">Figure 14
<p>SEM image of NCM88 after 100 cycles at a high cut-off voltage of 2.7 V–4.5 V: (<b>a</b>,<b>b</b>) NCM88-0 and (<b>c</b>,<b>d</b>) NCM88-0.5Nb.</p>
Full article ">Figure 14 Cont.
<p>SEM image of NCM88 after 100 cycles at a high cut-off voltage of 2.7 V–4.5 V: (<b>a</b>,<b>b</b>) NCM88-0 and (<b>c</b>,<b>d</b>) NCM88-0.5Nb.</p>
Full article ">
14 pages, 4061 KiB  
Article
Binder-Free Three-Dimensional Porous Graphene Cathodes via Self-Assembly for High-Capacity Lithium–Oxygen Batteries
by Yanna Liu, Wen Meng, Yuying Gao, Menglong Zhao, Ming Li and Liang Xiao
Nanomaterials 2024, 14(9), 754; https://doi.org/10.3390/nano14090754 - 25 Apr 2024
Viewed by 779
Abstract
The porous architectures of oxygen cathodes are highly desired for high-capacity lithium–oxygen batteries (LOBs) to support cathodic catalysts and provide accommodation for discharge products. However, controllable porosity is still a challenge for laminated cathodes with cathode materials and binders, since polymer binders usually [...] Read more.
The porous architectures of oxygen cathodes are highly desired for high-capacity lithium–oxygen batteries (LOBs) to support cathodic catalysts and provide accommodation for discharge products. However, controllable porosity is still a challenge for laminated cathodes with cathode materials and binders, since polymer binders usually shield the active sites of catalysts and block the pores of cathodes. In addition, polymer binders such as poly(vinylidene fluoride) (PVDF) are not stable under the nucleophilic attack of intermediate product superoxide radicals in the oxygen electrochemical environment. The parasitic reactions and blocking effect of binders deteriorate and then quickly shut down the operation of LOBs. Herein, the present work proposes a binder-free three-dimensional (3D) porous graphene (PG) cathode for LOBs, which is prepared by the self-assembly and the chemical reduction of GO with triblock copolymer soft templates (Pluronic F127). The interconnected mesoporous architecture of resultant 3D PG cathodes achieved an ultrahigh capacity of 10,300 mAh g−1 for LOBs. Further, the cathodic catalysts ruthenium (Ru) and manganese dioxide (MnO2) were, respectively, loaded onto the inner surface of PG cathodes to lower the polarization and enhance the cycling performance of LOBs. This work provides an effective way to fabricate free-standing 3D porous oxygen cathodes for high-performance LOBs. Full article
(This article belongs to the Section Energy and Catalysis)
Show Figures

Figure 1

Figure 1
<p>The optical and SEM images of (<b>a</b>,<b>b</b>) LG and (<b>c</b>,<b>d</b>) PG cathodes; the Raman spectra of (<b>e</b>) LG and (<b>f</b>) PG cathodes.</p>
Full article ">Figure 2
<p>(<b>a</b>) The discharge profiles of PG and LG cathodes at a current density of 200 mA g<sup>−1</sup>; (<b>b</b>) the voltage profiles of PG and LG cathodes at 200 mA g<sup>−1</sup> with a curtailed capacity of 1000 mAh g<sup>−1</sup>; the cross-section SEM images of (<b>c</b>) PG and (<b>d</b>) LG.</p>
Full article ">Figure 3
<p>The SEM images of (<b>a</b>,<b>b</b>) MnO<sub>2</sub>@PG and (<b>c</b>,<b>d</b>) Ru@PG.</p>
Full article ">Figure 4
<p>The TEM and HR-TEM images of (<b>a</b>,<b>b</b>) MnO<sub>2</sub>@PG and (<b>c</b>,<b>d</b>) Ru@PG.</p>
Full article ">Figure 5
<p>The XPS spectra of MnO<sub>2</sub>@PG and Ru@PG: (<b>a</b>) survey and (<b>b</b>) Mn 2p spectra of MnO<sub>2</sub>@PG; (<b>c</b>) survey, (<b>d</b>) Ru 3d, and (<b>e</b>) Ru 2p spectra of Ru@PG.</p>
Full article ">Figure 6
<p>The voltage profiles of LOBs with (<b>a</b>) PG, (<b>b</b>) MnO<sub>2</sub>@PG, and (<b>c</b>) Ru@PG cathodes at different current densities; (<b>d</b>) the voltage profiles of PG, MnO<sub>2</sub>@PG, and Ru@PG at 200 mA g<sup>−1</sup> with a curtailed specific capacity of 1000 mAh g<sup>−1</sup>.</p>
Full article ">Figure 7
<p>The SEM images of discharged (<b>a</b>,<b>b</b>) PG, (<b>c</b>,<b>d</b>) MnO<sub>2</sub>@PG, and (<b>e</b>,<b>f</b>) Ru@PG cathodes (partially discharged to 1000 mAh g<sup>−1</sup> and fully discharged to 2.2 V).</p>
Full article ">Figure 8
<p>Li 1s and O 1s XPS spectra of discharged cathodes: (<b>a</b>,<b>b</b>) GMS, (<b>c</b>,<b>d</b>) MnO<sub>2</sub>@PG, and (<b>e</b>,<b>f</b>) Ru@PG.</p>
Full article ">
16 pages, 6005 KiB  
Article
Pseudo-Eutectic of Isodimorphism to Design Biaxially-Oriented Bio-Based PA56/512 with High Strength, Toughness and Barrier Performances
by Diansong Gan, Yuejun Liu, Tianhui Hu, Shuhong Fan, Lingna Cui, Guangkai Liao, Zhenyan Xie, Xiaoyu Zhu and Kejian Yang
Polymers 2024, 16(8), 1176; https://doi.org/10.3390/polym16081176 - 22 Apr 2024
Cited by 1 | Viewed by 888
Abstract
The biaxially-oriented PA56/512 has excellent mechanical strength, extensibility and water–oxygen barrier properties and has broad application prospects in green packaging, lithium battery diaphragm and medical equipment materials. The correlation between the aggregation structure evolution and macroscopic comprehensive properties of copolymer PA56/512 under biaxial [...] Read more.
The biaxially-oriented PA56/512 has excellent mechanical strength, extensibility and water–oxygen barrier properties and has broad application prospects in green packaging, lithium battery diaphragm and medical equipment materials. The correlation between the aggregation structure evolution and macroscopic comprehensive properties of copolymer PA56/512 under biaxial stretching has been demonstrated in this work. The structure of the random copolymerization sequence was characterized by 13C Nuclear magnetic resonance (NMR). The typical isodimorphism behavior of the co-crystallization system of PA56/512 and its BOPA-56/512 films was revealed by differential scanning calorimetry (DSC) and X-ray diffraction (XRD) tests. And the aggregation structure, including the hydrogen bond arrangement, crystal structure and crystal morphology of PA56/512 before and after biaxial stretching, was investigated by XRD, Fourier-transform infrared spectroscopy (FTIR) and polarized optical microscopy (POM) tests. Furthermore, the effect of the biaxially-oriented stretching process on the mechanical properties of PA56/512 has been demonstrated. In addition, a deep insight into the influence of the structure on the crystallization process and physical–mechanical performance has been presented. The lowest melting point at a 512 content of 60 mol% is regarded as a “eutectic” point of the isodimorphism system. Due to the high disorder of the structural units in the polymer chain, the transition degree of the folded chain (gauche conformation) is relatively lowest when it is straightened to form an extended chain (trans conformation) during biaxially-oriented stretching, and part of the folded chain can be retained. This explains why biaxially stretched PA56/512 has high strength, outstanding toughness and excellent barrier properties at the pseudo-eutectic point. In this study, using the unique multi-scale aggregation structure characteristics of a heterohomodymite polyamide at the pseudo-eutectic point, combined with the new material design scheme and the idea of biaxial-stretching processing, a new idea for customized design of high-performance multifunctional polyamide synthetic materials is provided. Full article
(This article belongs to the Special Issue Advances in Interfacial Compatibility of Polymer Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Diagram of the biaxially-oriented stretching process of BOPA-56/512.</p>
Full article ">Figure 2
<p><sup>13</sup>C NMR spectra of PA56/512 with different composition ratios.</p>
Full article ">Figure 3
<p>(<b>a<sub>1</sub></b>) [<a href="#B21-polymers-16-01176" class="html-bibr">21</a>] The second heating DSC curves of PA56/512 resin with different composition ratios, (<b>a<sub>2</sub></b>) the first heating DSC curves of BOPA-56/512 with different composition ratios, the crystallization curves of (<b>b<sub>1</sub></b>) [<a href="#B21-polymers-16-01176" class="html-bibr">21</a>] PA56/512 resin, (<b>b<sub>2</sub></b>) BOPA-56/512 with different composition ratios. All the samples are set at heating/cooling rates of 10 °C/min.</p>
Full article ">Figure 4
<p>(<b>a</b>) [<a href="#B21-polymers-16-01176" class="html-bibr">21</a>] The melting temperature curves and (<b>b</b>) crystallization temperature curves of PA56/512 before and after biaxial stretching with different content of 512.</p>
Full article ">Figure 5
<p>The mechanical properties curve of PA56/512 resin: (<b>a</b>) tensile stress–strain curve, (<b>b</b>) local enlarged tensile stress–strain curve, (<b>c</b>) tensile strength with different contents of 512, (<b>d</b>) tensile elongation yield with different contents of 512.</p>
Full article ">Figure 6
<p>The tensile properties of BOPA-56/512 films: (<b>a</b>) tensile stress–strain curve, (<b>b</b>) breaking strength with different contents of 512, (<b>c</b>) tensile elongation yield with different contents of 512.</p>
Full article ">Figure 7
<p>Barrier properties of BOPA-56/512 films: (<b>a</b>) oxygen permeability curve; (<b>b</b>) water vapor permeability curve.</p>
Full article ">Figure 8
<p>Comparison of comprehensive properties of various barrier materials.</p>
Full article ">Figure 9
<p>XRD spectra of PA56/512 (<b>a</b>) before and (<b>b</b>) after biaxial stretching with different composition ratios, (<b>c</b>) the d-spacing curves of PA56/512 before and after biaxial stretching with different contents of 512.</p>
Full article ">Figure 10
<p>FTIR spectra in the fingerprint region of PA56/512 (<b>a</b>) before and (<b>b</b>) after biaxial stretching with different composition ratios.</p>
Full article ">Figure 11
<p>POM images of PA56/512 after biaxial stretching: (<b>a</b>) #1′, (<b>b</b>) #2′, (<b>c</b>) #3′, (<b>d</b>) #4′, (<b>e</b>) #5′, (<b>f</b>) #6′, (<b>g</b>) #7′, (<b>h</b>) #8′, (<b>i</b>) #9′.</p>
Full article ">Figure 12
<p>Molecular mechanism explanation of high-performance PA56/512.</p>
Full article ">Scheme 1
<p>Synthetic route of PA56/512.</p>
Full article ">
12 pages, 11320 KiB  
Article
The Optimization of Nickel-Rich Cathode-Material Production on a Pilot Plant Scale
by Agus Purwanto, Muhammad Nur Ikhsanudin, Putri Putih Puspa Asri, Afifah Salma Giasari, Miftakhul Hakam, Cornelius Satria Yudha, Hendri Widiyandari, Endah Retno Dyartanti, Arif Jumari and Adrian Nur
Processes 2024, 12(4), 685; https://doi.org/10.3390/pr12040685 - 28 Mar 2024
Viewed by 1386
Abstract
Lithium-ion batteries (LIBs) remain the cornerstone of EV technology due to their exceptional energy density. The selection of cathode materials is a decisive factor in LIB technology, profoundly influencing performance, energy density, and lifespan. Among these materials, nickel-rich NCM cathodes have gained significant [...] Read more.
Lithium-ion batteries (LIBs) remain the cornerstone of EV technology due to their exceptional energy density. The selection of cathode materials is a decisive factor in LIB technology, profoundly influencing performance, energy density, and lifespan. Among these materials, nickel-rich NCM cathodes have gained significant attention due to their high specific capacity and cost-effectiveness, making them a preferred choice for EV energy storage. However, the transition from the laboratory-scale to industrial-scale production of NMC-811 cathode material presents challenges, particularly in optimizing the oxidation process of Ni2+ ions. This paper addresses the challenges of transitioning NMC-811 cathode material production from a lab scale to a pilot scale, with its high nickel content requiring specialized oxidation processes. The important point emphasized in this transition process is how to produce cathode materials on a pilot scale, but show results equivalent to the laboratory scale. Several optimization variations are carried out, namely, the optimization of the heating rate and the calcination and sintering temperatures, as well as oxygen variations. These two aspects are important for large-scale production. This paper discusses strategies for successful pilot-scale production, laying the foundation for industrial-scale manufacturing. Additionally, NMC-811 cathodes are incorporated into 18650 cylindrical cells, advancing the adoption of high-performance cathode materials. Full article
(This article belongs to the Section Chemical Processes and Systems)
Show Figures

Figure 1

Figure 1
<p>Morphology analysis of NMC-811 cathode material. (<b>a</b>–<b>c</b>) SEM images at 500× and 2500× magnification and particle-size distribution of commercial NMC811. (<b>d</b>–<b>f</b>) SEM images at 500× and 2500× and particle-size distribution of NMC–oxalate precursor. (<b>g</b>–<b>i</b>) SEM images at 500× and 2500× and particle-size distribution of NMC-811 produced with small furnace (FK). (<b>j</b>–<b>l</b>) SEM images at 500× and 2500× and particle-size distribution of NMC-811 produced with large furnace with process condition 1 (FB-1). (<b>m</b>–<b>o</b>) SEM images at 500× and 2500× and particle-size distribution of NMC 811 FB with process condition 2 (FB-2). (<b>p</b>–<b>r</b>) SEM images at 500× and 2500× and particle-size distribution of NMC-811 FB with process condition 3 (FB-3).</p>
Full article ">Figure 2
<p>FTIR spectra of NMC-811 materials.</p>
Full article ">Figure 3
<p>XRD analysis of NMC-811 from small furnace (FK) and large furnace (FB).</p>
Full article ">Figure 4
<p>Thermogravimetric analysis (TGA) of NMC-811 cathode material.</p>
Full article ">Figure 5
<p>Electrochemical performance test results: (<b>a</b>) Full-cell analysis charge–discharge test at 0.05 C, (<b>b</b>) rate performance of NMC-811 cathode material, and (<b>c</b>) cycle stability at 0.5 C.</p>
Full article ">
14 pages, 4371 KiB  
Article
Enhancing the Storage Performance and Thermal Stability of Ni-Rich Layered Cathodes by Introducing Li2MnO3
by Jun Yang, Pingping Yang and Hongyu Wang
Energies 2024, 17(4), 810; https://doi.org/10.3390/en17040810 - 8 Feb 2024
Cited by 1 | Viewed by 935
Abstract
Ni-rich layered cathodes are deemed as a potential candidate for high-energy-density lithium-ion batteries, but their high sensitivity to air during storage and poor thermal stability are a vital challenge for large-scale applications. In this paper, distinguished from the conventional surface modification and ion [...] Read more.
Ni-rich layered cathodes are deemed as a potential candidate for high-energy-density lithium-ion batteries, but their high sensitivity to air during storage and poor thermal stability are a vital challenge for large-scale applications. In this paper, distinguished from the conventional surface modification and ion doping, an effective solid-solution strategy was proposed to strengthen the surface and structural stability of Ni-rich layered cathodes by introducing Li2MnO3. The structural analysis results indicate that the formation of Li2CO3 inert layers on Ni-rich layered cathodes during storage in air is responsible for the increased electrode interfacial impedance, thereby leading to the severe deterioration of electrochemical performance. The introduction of Li2MnO3 can reduce the surface reactivity of Ni-rich cathode materials, playing a certain suppression effect on the formation of surface Li2CO3 layer and the deterioration of electrochemical performances. Additionally, the thermal analysis results show that the heat release of Ni-rich cathodes strongly depends on the charge of states, and Li2MnO3 can suppress oxygen release and significantly enhance the thermal stability of Ni-rich layered cathodes. This work provides a method to improving the storage performance and thermal stability of Ni-rich cathode materials. Full article
(This article belongs to the Special Issue Advanced Design Technologies of Lithium Ion Batteries Electrodes)
Show Figures

Figure 1

Figure 1
<p>XRD patterns and magnifications of cathodes before and after storage: (<b>a</b>,<b>c</b>) NCM-811. (<b>b</b>,<b>d</b>) LNCMO-1090.</p>
Full article ">Figure 2
<p>The C 1s XPS spectra of cathodes before and after storage: (<b>a</b>) NCM-811 and (<b>b</b>) LNCMO-1090.</p>
Full article ">Figure 3
<p>SEM images of cathodes before and after storage: (<b>a</b>,<b>b</b>) NCM-811 and (<b>c</b>,<b>d</b>) LNCMO-1090.</p>
Full article ">Figure 4
<p>TEM images of cathodes before and after storage: (<b>a</b>,<b>c</b>) NCM-811 and (<b>b</b>,<b>d</b>) LNCMO-1090.</p>
Full article ">Figure 5
<p>The Nyquist plots and the corresponding equivalent circuit model (inset) of the cathodes charged to 4.3 V before and after storage: (<b>a</b>) NCM-811 and (<b>b</b>) LNCMO-1090.</p>
Full article ">Figure 6
<p>Comparison of electrochemical performance of cathode materials before and after storage. (<b>a</b>,<b>b</b>) The first charge/discharge profiles at 0.1 C. (<b>c</b>,<b>d</b>) Cycling performance at 0.1 C. (<b>e</b>,<b>f</b>) Rate capability.</p>
Full article ">Figure 7
<p>(<b>a</b>) The DSC curves of the NCM-811 cathode with different charge states. (<b>b</b>) The DSC curves of the NCM-811 and LNCMO-1090 cathodes with charge state of 4.5 V vs. Li/Li<sup>+</sup>.</p>
Full article ">
15 pages, 5015 KiB  
Article
Investigating the Influence of Three Different Atmospheric Conditions during the Synthesis Process of NMC811 Cathode Material
by Arianna Tiozzo, Keyhan Ghaseminezhad, Asya Mazzucco, Mattia Giuliano, Riccardo Rocca, Matteo Dotoli, Giovanna Nicol, Carlo Nervi, Marcello Baricco and Mauro Francesco Sgroi
Crystals 2024, 14(2), 137; https://doi.org/10.3390/cryst14020137 - 29 Jan 2024
Cited by 1 | Viewed by 1905
Abstract
Lithium-ion batteries (LIBs) are fundamental for the energetic transition necessary to contrast climate change. The characteristics of cathode active materials (CAMs) strongly influence the cell performance, so improved CAMs need to be developed. Currently, Li(Ni0.8Mn0.1Co0.1)O2 (NMC811) [...] Read more.
Lithium-ion batteries (LIBs) are fundamental for the energetic transition necessary to contrast climate change. The characteristics of cathode active materials (CAMs) strongly influence the cell performance, so improved CAMs need to be developed. Currently, Li(Ni0.8Mn0.1Co0.1)O2 (NMC811) is state-of-the-art among the cathodic active materials. The aim of this work is the optimization of the procedure to produce NMC811: two different syntheses were investigated, the co-precipitation and the self-combustion methods. For a better understanding of the synthesis conditions, three different types of atmospheres were tested during the calcination phase: air (partially oxidizing), oxygen (totally oxidizing), and nitrogen (non-oxidizing). The synthesized oxides were characterized by X-ray Powder Diffraction (XRPD), Scanning Electron Microscopy (SEM), Energy Dispersive X-ray (EDX), Inductively Coupled Plasma (ICP), and Particle Size Distribution (PSD). The most promising materials were tested in a half-cell set up to verify the electrochemical performances. The procedure followed during this study is depicted in the graphical abstract. The oxidizing atmospheric conditions turned out to be the most appropriate to produce NMC811 with good electrochemical properties. Full article
Show Figures

Figure 1

Figure 1
<p>Workflow and processes adopted in the present work.</p>
Full article ">Figure 2
<p>Components of a coin cell with NMC811 cathode and lithium metal anode. The cell is composed of a metallic lithium anode and the cathode materials deposited on an aluminum current collector. The two electrodes are divided using the polymeric separator impregnated by the electrolyte. The spacers and the spring are used to apply pressure and to keep the cell components in good contact. The cell is enclosed into a metallic case composed of two parts (negative and positive) that are not electrically connected.</p>
Full article ">Figure 3
<p>XRD of the commercial NMC811.</p>
Full article ">Figure 4
<p>XRD results of CPT-O<sub>2</sub> and CPT-Air.</p>
Full article ">Figure 5
<p>XRD results of SCS-O<sub>2</sub> and SCS-Air.</p>
Full article ">Figure 6
<p>XRD results of CPT-N<sub>2</sub>.</p>
Full article ">Figure 7
<p>XRD results of SCS-N<sub>2</sub>.</p>
Full article ">Figure 8
<p>SEM images of (<b>A</b>) SCS-O<sub>2</sub>, (<b>B</b>) SCS-Air, (<b>C</b>) SCS-N<sub>2</sub>, (<b>D</b>) CPT-O<sub>2</sub>, (<b>E</b>) CPT-Air, and (<b>F</b>) CPT-N<sub>2</sub>.</p>
Full article ">Figure 9
<p>EDS map of CPT-O<sub>2</sub>.</p>
Full article ">Figure 10
<p>Cycling performance of prepared electrodes.</p>
Full article ">Figure 11
<p>C-rate performance of the electrodes.</p>
Full article ">
17 pages, 6132 KiB  
Article
Preparation and Lithium-Ion Capacitance Performance of Nitrogen and Sulfur Co-Doped Carbon Nanosheets with Limited Space via the Vermiculite Template Method
by Fang Yang, Pingzheng Jiang, Qiqi Wu, Wei Dong, Minghu Xue and Qiao Zhang
Molecules 2024, 29(2), 536; https://doi.org/10.3390/molecules29020536 - 22 Jan 2024
Viewed by 964
Abstract
Nitrogen and sulfur co-doped graphene-like carbon nanosheets (CNSs) with a two-dimensional structure are prepared by using methylene blue as a carbon source and expanded vermiculite as a template. After static negative pressure adsorption, high-temperature calcination, and etching in a vacuum oven, they are [...] Read more.
Nitrogen and sulfur co-doped graphene-like carbon nanosheets (CNSs) with a two-dimensional structure are prepared by using methylene blue as a carbon source and expanded vermiculite as a template. After static negative pressure adsorption, high-temperature calcination, and etching in a vacuum oven, they are embedded in the limited space of the vermiculite template. The addition of an appropriate number of mixed elements can improve the performance of a battery. Via scanning electron microscopy, it is found that the prepared nitrogen–sulfur-co-doped carbon nanosheets exhibit a thin yarn shape. The XPS results show that there are four elements of C, N, O, and S in the carbon materials (CNS-600, CNS-700, CNS-800, CNS-900) prepared at different temperatures, and the N atom content shows a gradually decreasing trend. It is mainly doped into a graphene-like network in four ways (graphite nitrogen, pyridine nitrogen, pyrrole nitrogen, and pyridine nitrogen oxide), while the S element shows an increasing trend, mainly in the form of thiophene S and sulfur, which is covalently linked to oxygen. The results show that CNS-700 has a discharge-specific capacity of 460 mAh/g at a current density of 0.1 A/g, and it can still maintain a specific capacity of 200 mAh/g at a current density of 2 A/g. The assembled lithium-ion capacitor has excellent energy density and power density, with a maximum power density of 20,000 W/kg. Full article
(This article belongs to the Special Issue Advanced Functional Nanomaterials for Energy Conversion and Storage)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of MB-700, CNS-600, CNS-700, CNS-800, and CNS-900 (<b>a</b>); Raman spectra of MB-700, CNS-600, CNS-700, CNS-800, and CNS-900 (<b>b</b>); the nitrogen adsorption and desorption curves of vermiculite-based template carbon materials calcined at four different temperatures (<b>c</b>); and the pore size distribution curves of vermiculite-based template carbon materials calcined at four different temperatures (<b>d</b>).</p>
Full article ">Figure 2
<p>The XPS full spectrum (<b>a</b>) and the atomic percentage of N and S (<b>b</b>) in the five samples.</p>
Full article ">Figure 3
<p>N1s and S2p fitting spectra of MB-700 (<b>a</b>,<b>b</b>), N1s and S2p fitting spectra of CNS-600 (<b>c</b>,<b>d</b>), N1s and S2p fitting spectra of CNS-700 (<b>e</b>,<b>f</b>), N1s and S2p fitting spectra of CNS-800 (<b>g</b>,<b>h</b>), N1s and S2p fitting spectra of CNS-900 (<b>i</b>,<b>j</b>).</p>
Full article ">Figure 4
<p>SEM map of MB-700 (<b>a</b>), CNS-600 (<b>b</b>), CNS-800 (<b>c</b>), CNS-900 (<b>d</b>), and CNS-700 for CNS (<b>e</b>,<b>f</b>); TEM image of CNS-700 (<b>g</b>); and HRTEM image of CNS-700 (<b>h</b>,<b>i</b>) (the small yellow frames in each graph refer to the locally enlarged regions).</p>
Full article ">Figure 5
<p>Charge–discharge curves of samples at different carbonization temperatures: CNS-600 (<b>a</b>), CNS-700 (<b>b</b>), CNS-800 (<b>c</b>), CNS-900 (<b>d</b>), MB-700 (<b>e</b>), and AC impedance spectra of samples at different carbonization temperatures (<b>f</b>).</p>
Full article ">Figure 6
<p>Cycle performance diagram at 0.1 A/g (<b>a</b>); cycle performance diagram at 1 A/g (<b>b</b>); cycle performance diagram at 2 A/g (<b>c</b>); rate performance diagrams of 0.05, 0.1, 0.2, 0.5, 1, 2, 4 A/g (<b>d</b>).</p>
Full article ">Figure 7
<p>Cyclic voltammograms of four templated carbon materials. CNS-600 (<b>a</b>), CNS-700 (<b>b</b>), CNS-800 (<b>c</b>), CNS-900 (<b>d</b>), and carbon material MB-700 prepared at different calcination temperatures (<b>e</b>); GITT test curve of four template carbon materials in the charging and discharging process (<b>f</b>); lithium-ion diffusion coefficient of four template carbon materials in the discharging process (<b>g</b>); lithium-ion diffusion coefficient curve of four template carbon materials in the charging process (<b>h</b>).</p>
Full article ">Figure 8
<p>CNS-700//AC lithium-ion capacitor cyclic volt–ampere curve at different scanning rates (<b>a</b>); constant current charge–discharge curve at different current densities (<b>b</b>); Thouragongtu (<b>c</b>); long cycle capacity retention rate and coulomb efficiency diagram (<b>d</b>).</p>
Full article ">
12 pages, 3079 KiB  
Article
Solution-Plasma Synthesis and Characterization of Transition Metals and N-Containing Carbon–Carbon Nanotube Composites
by Kodai Sasaki, Kaiki Yamamoto, Masaki Narahara, Yushi Takabe, Sangwoo Chae, Gasidit Panomsuwan and Takahiro Ishizaki
Materials 2024, 17(2), 320; https://doi.org/10.3390/ma17020320 - 8 Jan 2024
Cited by 1 | Viewed by 1284
Abstract
Lithium–air batteries (LABs) have a theoretically high energy density. However, LABs have some issues, such as low energy efficiency, short life cycle, and high overpotential in charge–discharge cycles. To solve these issues electrocatalytic materials were developed for oxygen reduction reaction (ORR) and oxygen [...] Read more.
Lithium–air batteries (LABs) have a theoretically high energy density. However, LABs have some issues, such as low energy efficiency, short life cycle, and high overpotential in charge–discharge cycles. To solve these issues electrocatalytic materials were developed for oxygen reduction reaction (ORR) and oxygen evolution reaction (OER), which significantly affect battery performance. In this study, we aimed to synthesize electrocatalytic N-doped carbon-based composite materials with solution plasma (SP) using Co or Ni as electrodes from organic solvents containing cup-stacked carbon nanotubes (CSCNTs), iron (II) phthalocyanine (FePc), and N-nethyl-2-pyrrolidinone (NMP). The synthesized N-doped carbon-based composite materials were characterized by transmission electron microscopy (TEM), X-ray diffraction (XRD), Raman spectroscopy, and X-ray photoelectron spectroscopy (XPS). TEM observation and XPS measurements revealed that the synthesized carbon materials contained elemental N, Fe, and electrode-derived Co or Ni, leading to the successful synthesis of N-doped carbon-based composite materials. The electrocatalytic activity for ORR of the synthesized carbon-based composite materials was also evaluated using electrochemical measurements. The electrochemical measurements demonstrated that the electrocatalytic performance for ORR of N-doped carbon-based composite material including Fe and Co showed superiority to that of N-doped carbon-based composite material including Fe and Ni. The difference in the electrocatalytic performance for ORR is discussed regarding the difference in the specific surface area and the presence ratio of chemical bonding species. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of solution plasma equipment.</p>
Full article ">Figure 2
<p>TEM and elemental mapping images of (<b>a</b>) Co-NC@CNT, (<b>b</b>) Ni-NC@CNT, and (<b>c</b>) pristine CNT.</p>
Full article ">Figure 3
<p>XRD patterns of (a) Co-NC@CNT, (b) Ni-NC@CNT, and (c) CSCNT.</p>
Full article ">Figure 4
<p>Raman spectra of (a) Co−NC@CNT, (b) Ni−NC@CNT, and (c) CSCNT.</p>
Full article ">Figure 5
<p>XPS N1<span class="html-italic">s</span> spectra of (<b>a</b>) Co-NC@CNT and (<b>b</b>) Ni-NC@CNT.</p>
Full article ">Figure 6
<p>Linear sweep voltammograms (LSVs) of (a) Co-NC@CNT, (b) Ni-NC@CNT, (c) CSCNT, and (d) commercial Pt/C at 1600 rpm in the O<sub>2</sub>-saturated 0.1 M KOH solution.</p>
Full article ">Figure 7
<p>(<b>a</b>) Change in electron transfer numbers of Co-NC@CNT, Ni-NC@CNT, and CSCNT as a function of potentials. (<b>b</b>) Change in H<sub>2</sub>O<sub>2</sub> yield of Co-NC@CNT, Ni-NC@CNT, and CSCNT as a function of potentials.</p>
Full article ">Figure 8
<p>Chronoamperometric responses of Co-NC@CNT and Pt/C in O<sub>2</sub>-saturated 0.1 M KOH.</p>
Full article ">
26 pages, 4507 KiB  
Review
A Brief Review of MoO3 and MoO3-Based Materials and Recent Technological Applications in Gas Sensors, Lithium-Ion Batteries, Adsorption, and Photocatalysis
by Mário Gomes da Silva Júnior, Luis Carlos Costa Arzuza, Herbet Bezerra Sales, Rosiane Maria da Costa Farias, Gelmires de Araújo Neves, Hélio de Lucena Lira and Romualdo Rodrigues Menezes
Materials 2023, 16(24), 7657; https://doi.org/10.3390/ma16247657 - 15 Dec 2023
Cited by 2 | Viewed by 1897
Abstract
Molybdenum trioxide is an abundant natural, low-cost, and environmentally friendly material that has gained considerable attention from many researchers in a variety of high-impact applications. It is an attractive inorganic oxide that has been widely studied because of its layered structure, which results [...] Read more.
Molybdenum trioxide is an abundant natural, low-cost, and environmentally friendly material that has gained considerable attention from many researchers in a variety of high-impact applications. It is an attractive inorganic oxide that has been widely studied because of its layered structure, which results in intercalation ability through tetrahedral/octahedral holes and extension channels and leads to superior charge transfer. Shape-related properties such as high specific capacities, the presence of exposed active sites on the oxygen-rich structure, and its natural tendency to oxygen vacancy that leads to a high ionic conductivity are also attractive to technological applications. Due to its chemistry with multiple valence states, high thermal and chemical stability, high reduction potential, and electrochemical activity, many studies have focused on the development of molybdenum oxide-based systems in the last few years. Thus, this article aims to briefly review the latest advances in technological applications of MoO3 and MoO3-based materials in gas sensors, lithium-ion batteries, and water pollution treatment using adsorption and photocatalysis techniques, presenting the most relevant and new information on heterostructures, metal doping, and non-stoichiometric MoO3−x. Full article
Show Figures

Figure 1

Figure 1
<p>Published scientific papers directly related to MoO<sub>3</sub> and MoO<sub>3</sub>-based systems in the last 20 years (Web of Science search with MoO<sub>3</sub> keyword, 2023 ongoing, Access on 9 October 2023).</p>
Full article ">Figure 2
<p>MoO<sub>3</sub> crystalline structure illustrations.</p>
Full article ">Figure 3
<p><b>Right</b> panel: Crystal structure of MoO<sub>3</sub> with H+ intercalation and hydrogen molybdenum bronze. <b>Left</b> panel: Electronic band structure manipulation by H+ intercalation. Figure reproduced with permission of [<a href="#B24-materials-16-07657" class="html-bibr">24</a>] Copyright 2017, Nature Publishing Group.</p>
Full article ">Figure 4
<p>Scheme of the reaction mechanism of MoO<sub>3</sub> materials exposed to hydrogen gas. Based on Ref. [<a href="#B36-materials-16-07657" class="html-bibr">36</a>].</p>
Full article ">Figure 5
<p>Published scientific papers and citations directly related to MoO<sub>3</sub> and MoO<sub>3</sub>-based systems applied to gas sensors in the last 20 years (Web of Science search with MoO<sub>3</sub> and gas sensors keywords, 2023 ongoing, Access on 9 October 2023).</p>
Full article ">Figure 6
<p>Responses of MoO<sub>3</sub> nanorods to NO<sub>2</sub>, CO, and CH<sub>4</sub> gases at 40 ppm and different temperatures. Figure reproduced with permission of [<a href="#B41-materials-16-07657" class="html-bibr">41</a>] Copyright 2012, Elsevier.</p>
Full article ">Figure 7
<p>Schematic diagram illustrating the 1-butylamine sensing mechanism on the MoO<sub>3</sub> surface showing the dehydrogenation pathway of (<b>a</b>) 1-butylamine and (<b>b</b>) 1-butylimine. Figure reproduced with permission of [<a href="#B37-materials-16-07657" class="html-bibr">37</a>] Copyright 2021, Elsevier.</p>
Full article ">Figure 8
<p>A schematic diagram illustrates an external load applied to both electrodes. This is the discharging process of a lithium-ion battery. Figure reproduced with permission of [<a href="#B68-materials-16-07657" class="html-bibr">68</a>] Copyright 2017, Wiley.</p>
Full article ">Figure 9
<p>Published scientific papers and citations directly related to MoO<sub>3</sub> and MoO<sub>3</sub>-based systems applied to LIBs in the last 20 years (Web of Science search with MoO<sub>3</sub> and lithium-ion batteries keywords, 2023 ongoing, Access on 9 October 2023).</p>
Full article ">Figure 10
<p>Published scientific papers and citations directly related to MoO<sub>3</sub> and MoO<sub>3</sub>-based systems applied to adsorption in the last 20 years (Web of Science search with MoO<sub>3</sub> and adsorption keywords, 2023 ongoing, Access on 9 October 2023).</p>
Full article ">Figure 11
<p>Removal efficiency of metal ions Cu<sup>2+</sup>, Pb<sup>2+</sup>, Zn<sup>2+</sup>, Cr<sup>3+</sup>, and Cd<sup>2+</sup> removed by the α-MoO<sub>3</sub> nanosheet array system. Figure reproduced with permission of [<a href="#B94-materials-16-07657" class="html-bibr">94</a>] Copyright 2017, Elsevier.</p>
Full article ">Figure 12
<p>Published scientific papers and citations directly related to MoO<sub>3</sub> and MoO<sub>3</sub>-based systems applied to photocatalysis in the last 20 years (Web of Science search with MoO<sub>3</sub> and photocatalysis keywords, 2023 ongoing, Access on 9 October 2023).</p>
Full article ">Figure 13
<p>Schematic representation of visible light photocatalytic degradation of MB dye by MoO<sub>3</sub>.</p>
Full article ">
Back to TopTop