[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (25)

Search Parameters:
Keywords = Li6PS5Cl

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 10800 KiB  
Article
On the Stability of the Interface between Li2TiS3 Cathode and Li6PS5Cl Solid State Electrolytes for Battery Applications: A DFT Study
by Riccardo Rocca, Naiara Leticia Marana, Fabrizio Silveri, Maddalena D’Amore, Eleonora Ascrizzi, Mauro Francesco Sgroi, Nello Li Pira and Anna Maria Ferrari
Batteries 2024, 10(10), 351; https://doi.org/10.3390/batteries10100351 - 7 Oct 2024
Viewed by 923
Abstract
Lithium-titanium-sulfur cathodes have garnered interest due to their distinctive properties and potential applications in lithium-ion batteries. They present various benefits, including lower cost, enhanced safety, and greater energy density compared to the commonly used transition metal oxides. The current trend in lithium-ion batteries [...] Read more.
Lithium-titanium-sulfur cathodes have garnered interest due to their distinctive properties and potential applications in lithium-ion batteries. They present various benefits, including lower cost, enhanced safety, and greater energy density compared to the commonly used transition metal oxides. The current trend in lithium-ion batteries is to move to all-solid-state chemistries in order to improve safety and energy density. Several chemistries for solid electrolytes have been studied, tested, and characterized to evaluate the applicability in energy storage system. Among those, sulfur-based Argyrodites have been coupled with cubic rock-salt type Li2TiS3 electrodes. In this work, Li2TiS3 surfaces were investigated with DFT methods in different conditions, covering the possible configurations that can occur during the cathode usage: pristine, delithiated, and overlithiated. Interfaces were built by coupling selected Li2TiS3 surfaces with the most stable Argyrodite surface, as derived from a previous study, allowing us to understand the (electro)chemical compatibility between these two sulfur-based materials. Full article
(This article belongs to the Special Issue Recent Process of Solid State Lithium Batteries)
Show Figures

Figure 1

Figure 1
<p>(001) Argyrodite surface. The yellow, gray, green, and magenta spheres represent the sulfur, lithium, chlorine, and phosphorus atoms.</p>
Full article ">Figure 2
<p>Schematic representation of LTS surfaces (<b>a</b>) (100), (<b>b</b>) (110), (<b>c</b>) (111). Yellow, gray, and light blue spheres refer to sulfur, lithium, and titanium atoms, respectively.</p>
Full article ">Figure 3
<p>Comparison between the bottom of the conducting band (BCB) and the top of the valence band (TVB) levels for the (<b>a</b>–<b>d</b>) isolated LTS and Argyrodite surfaces, and (<b>e</b>–<b>i</b>) formed interfaces for different LTS lithium content and the two Argyrodite terminations. The lines in blue, red, cyan, and magenta represent the energetic level of pristine LTS and Argyrodite surfaces, and the same surfaces on the formed interfaces, respectively.</p>
Full article ">Figure 4
<p>Final optimized (100)LTS/(001)Argy interfaces (<b>a</b>) LTS/Argy-Li<sub>2</sub>S, (<b>b</b>) LTS/Argy-LPSC, (<b>c</b>) LTS<sub>over</sub>/Argy-Li<sub>2</sub>S, (<b>d</b>) LTS<sub>even</sub>/Argy-Li<sub>2</sub>S, and (<b>e</b>) LTS<sub>odd</sub>/Argy-Li<sub>2</sub>S. The yellow, gray, light blue, green, and magenta spheres represent the sulfur, lithium, titanium, chlorine, and phosphorus atoms.</p>
Full article ">Figure 5
<p>Charge density difference between interface and the pristine sub-units of (<b>a</b>) LTS/Argy-Li<sub>2</sub>S, (<b>b</b>) LTS/Argy-LPSC, (<b>c</b>) LTS<sub>over</sub>/Argy-Li<sub>2</sub>S, (<b>d</b>) LTS<sub>even</sub>/Argy-Li<sub>2</sub>S, and (<b>e</b>) LTS<sub>odd</sub>/Argy-Li<sub>2</sub>S. The isosurface corresponds to 0.001 e<sup>−</sup>/bhor<sup>3</sup> with the charge accumulation (depletion) plotted in red (blue). The yellow, gray, light blue, green, and magenta spheres represent the sulfur, lithium, titanium, chlorine, and phosphorus atoms.</p>
Full article ">Figure 6
<p>Spin density for the (<b>a</b>) LTS<sub>odd</sub>/Argy-Li<sub>2</sub>S and (<b>b</b>) LTS<sub>over</sub>/Argy-Li<sub>2</sub>S interface. The isosurface corresponds to 0.001 e<sup>−</sup>/bohr<sup>−3</sup>. The yellow, gray, light blue, green, and magenta spheres represent the sulfur, lithium, titanium, chlorine, and phosphorus atoms.</p>
Full article ">
12 pages, 1233 KiB  
Article
Properties of Chitin and Its Regenerated Hydrogels from the Insect Zophobas morio Fed Citrus Biomass or Polystyrene
by Guillermo Ignacio Guangorena Zarzosa and Takaomi Kobayashi
Gels 2024, 10(7), 433; https://doi.org/10.3390/gels10070433 - 29 Jun 2024
Viewed by 1488
Abstract
The potential of insects as a recycling tool has recently attracted attention. In this study, chitin was extracted with 1 M HCl for 24 h at 20 °C, followed by 1 M NaOH for 5 h at 90 °C, and bleached with 2.5% [...] Read more.
The potential of insects as a recycling tool has recently attracted attention. In this study, chitin was extracted with 1 M HCl for 24 h at 20 °C, followed by 1 M NaOH for 5 h at 90 °C, and bleached with 2.5% v/v NaOCl for 2 h at 20 °C from Zophobas morio (ZM) insects fed citrus waste biomass (OP) or polystyrene foam (PS). The highest survival rate was found in the OP group. The properties of the resulting chitin material are reported, as well as the preparation of hydrogels using a DMAc/LiCl solvent. All chitins obtained were α-chitin. The degrees of deacetylation, crystallinity, molecular weight, and solubility in DMAc/LiCl were similar between the PS and biomass feeds, and they showed similar viscosities in the DMAc/LiCl solution. All hydrogels obtained had similar properties and viscoelastic behavior, indicating that the resultant chitins and their hydrogels from ZM were similar between those fed with citrus biomass and those fed with PS. Full article
(This article belongs to the Special Issue Advances in Chemistry and Physics of Hydrogels)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FTIR spectra of extracted chitins from different diet conditions.</p>
Full article ">Figure 2
<p>X-ray diffraction patterns of extracted chitins under different diet conditions.</p>
Full article ">Figure 3
<p>(<b>a</b>) Shear viscosity of 1% <span class="html-italic">w</span>/<span class="html-italic">w</span> DMAc/6% LiCl chitin solutions measured at 20 °C, and (<b>b</b>) molecular weight distributions.</p>
Full article ">Figure 4
<p>Chitin extraction and hydrogel preparation process.</p>
Full article ">Figure 5
<p>Strain sweep measurements of G′ and G″ and tan δ for chitin hydrogels measured from 0.01% strain to 100% strain at 20 °C.</p>
Full article ">
16 pages, 5696 KiB  
Article
Functionalization of Cathode–Electrolyte Interface with Ionic Liquids for High-Performance Quasi-Solid-State Lithium–Sulfur Batteries: A Low-Sulfur Loading Study
by Milinda Kalutara Koralalage, Varun Shreyas, William R. Arnold, Sharmin Akter, Arjun Thapa, Badri Narayanan, Hui Wang, Gamini U. Sumanasekera and Jacek B. Jasinski
Batteries 2024, 10(5), 155; https://doi.org/10.3390/batteries10050155 - 30 Apr 2024
Cited by 1 | Viewed by 1818
Abstract
We introduce a quasi-solid-state electrolyte lithium-sulfur (Li–S) battery (QSSEB) based on a novel Li-argyrodite solid-state electrolyte (SSE), Super P–Sulfur cathode, and Li-anode. The cathode was prepared using a water-based carboxymethyl cellulose (CMC) solution and styrene butadiene rubber (SBR) as the binder while Li [...] Read more.
We introduce a quasi-solid-state electrolyte lithium-sulfur (Li–S) battery (QSSEB) based on a novel Li-argyrodite solid-state electrolyte (SSE), Super P–Sulfur cathode, and Li-anode. The cathode was prepared using a water-based carboxymethyl cellulose (CMC) solution and styrene butadiene rubber (SBR) as the binder while Li6PS5F0.5Cl0.5 SSE was synthesized using a solvent-based process, via the introduction of LiF into the argyrodite crystal structure, which enhances both the ionic conductivity and interface-stabilizing properties of the SSE. Ionic liquids (IL) were prepared using lithium bis(trifluoromethyl sulfonyl)imide (LiTFSI) as the salt, with pre-mixed pyrrolidinium bis(trifluoromethyl sulfonyl)imide (PYR) as solvent and 1,3-dioxolane (DOL) as diluent, and they were used to wet the SSE–electrode interfaces. The effect of IL dilution, the co-solvent amount, the LiTFSI concentration, the C rate at which the batteries are tested and the effect of the introduction of SSE in the cathode, were systematically studied and optimized to develop a QSSEB with higher capacity retention and cyclability. Interfacial reactions occurring at the cathode–SSE interface during cycling were also investigated using electrochemical impedance spectroscopy, cyclic voltammetry, and X-ray photoelectron spectroscopy supported by ab initio molecular dynamics simulations. This work offers a new insight into the intimate interfacial contacts between the SSE and carbon–sulfur cathodes, which are critical for improving the electrochemical performance of quasi-solid-state lithium–sulfur batteries. Full article
Show Figures

Figure 1

Figure 1
<p>Cathode characterizations: (<b>a</b>) TGA measurement for quantifying active sulfur loading in cathode, FEI SEM image with (<b>b</b>) 100 and (<b>c</b>) 10,000 magnifications, (<b>d</b>) FEI SEM cross sectional image obtained with 60° tilt angle at 2000 magnification, (<b>e</b>) TESCAN SEM image for EDAX mapping, and (<b>f</b>) Carbon and (<b>g</b>) sulfur EDAX mapping.</p>
Full article ">Figure 2
<p>Electrochemical testing: (<b>a</b>) electrochemical impedance spectra and (<b>b</b>) discharge curves at 0.05 C rate for batteries with and without ionic liquids.</p>
Full article ">Figure 3
<p>Performance of batteries consisting of SP–S/SSE/Li with ILs: (<b>a</b>) 0.6 M LiTFSI dissolved in PYR, (<b>b</b>) 2 M LiTFSI dissolved in PYR/DOL (1:1), (<b>c</b>) 2 M LiTFSI dissolved in PYR/DOL (3:1), and (<b>d</b>) 4 M LiTFSI dissolved in PYR/DOL (1:1).</p>
Full article ">Figure 4
<p>Performance of batteries consisting of SP–S/SSE/Li with 2 M LiTFSI PYR/DOL (1:1) with different volumes: (<b>a</b>) 10 μL, (<b>b</b>) 20 μL, (<b>c</b>) 40 μL, and performance of batteries consisting of SP–S-SSE/SSE/Li (SSE incorporated in the cathode) with 2 M LiTFSI PYR/DOL (1:1) with different volumes (<b>d</b>) 10 μL, (<b>e</b>) 20 μL and (<b>f</b>) 40 μL.</p>
Full article ">Figure 5
<p>Electrochemical impedance spectra of batteries consisting of SP–S cathode, Li anode, and SSE (<b>a</b>) with no ionic liquid, (<b>b</b>) with IL of LiTFSI (1 M) dissolved in PYR, and (<b>c</b>) with IL of LiTFSI (2 M) dissolved in PYR/DOL (1:1) before and after discharging.</p>
Full article ">Figure 6
<p>Cyclic voltammogram of the batteries consist of SP–S cathode, Li anode, and SSE (<b>a</b>) with no ionic liquid (<b>b</b>) with IL LiTFSI (1 M) dissolved in PYR, and (<b>c</b>) with IL LiTFSI (2 M) dissolved in PYR/DOL (1:1).</p>
Full article ">Figure 7
<p>XPS S2p low binding energy peak of the cathode–SSE interface of (<b>a</b>) no ionic liquid, (<b>b</b>) with LiTFSI (1 M) dissolved in PYR, and (<b>c</b>,<b>d</b>) with LiTFSI (2 M) dissolved in PYR/DOL (1:1) after 1 cycle and 100 cycles.</p>
Full article ">Figure 8
<p>Structural evolution of PYR (PYR14TFSI) interface with (<b>a</b>,<b>b</b>) Li<sub>6</sub>PS<sub>5</sub>F<sub>0.5</sub>Cl<sub>0.5</sub>, (<b>c</b>,<b>d</b>) S8, and (<b>e</b>,<b>f</b>) Li2S obtained from ab initio molecular dynamics simulations (20 ps) under ambient conditions. (<b>g</b>) [TFSI]–decomposition products present at the end of simulation time (* represents the complexes).</p>
Full article ">Figure 9
<p>Structural evolution of 2 M LiTFSI in PYR (PYR14TFSI):DOL (1:1) interface with (<b>a</b>,<b>b</b>) Li<sub>6</sub>PS<sub>5</sub>F<sub>0.5</sub>Cl<sub>0.5</sub>, (<b>c</b>,<b>d</b>) S8, and (<b>e</b>,<b>f</b>) Li2S obtained from ab initio molecular dynamics simulations (20 ps) under ambient conditions.</p>
Full article ">
14 pages, 5387 KiB  
Article
Synthesis and Properties of Polyvinylidene Fluoride-Hexafluoropropylene Copolymer/Li6PS5Cl Gel Composite Electrolyte for Lithium Solid-State Batteries
by Xinghua Liang, Xueli Shi, Lingxiao Lan, Yunmei Qing, Bing Zhang, Zhijie Fang and Yujiang Wang
Gels 2024, 10(3), 199; https://doi.org/10.3390/gels10030199 - 14 Mar 2024
Cited by 2 | Viewed by 2131
Abstract
Gel electrolytes for lithium-ion batteries continue to replace the organic liquid electrolytes in conventional batteries due to their advantages of being less prone to leakage and non-explosive and possessing a high modulus of elasticity. However, the development of gel electrolytes has been hindered [...] Read more.
Gel electrolytes for lithium-ion batteries continue to replace the organic liquid electrolytes in conventional batteries due to their advantages of being less prone to leakage and non-explosive and possessing a high modulus of elasticity. However, the development of gel electrolytes has been hindered by their generally low ionic conductivity at room temperature and high interfacial impedance with electrodes. In this paper, a poly (vinylidene fluoride)-hexafluoropropylene copolymer (PVdF-HFP) with a flexible structure, Li6PS5Cl (LPSCl) powder of the sulfur–silver–germanium ore type, and lithium perchlorate salt (LiClO4) were prepared into sulfide gel composite electrolyte films (GCEs) via a thermosetting process. The experimental results showed that the gel composite electrolyte with 1% LPSCl in the PVdF-HFP matrix exhibited an ionic conductivity as high as 1.27 × 10−3 S·cm−1 at 25 °C and a lithium ion transference number of 0.63. The assembled LiFePO4||GCEs||Li batteries have excellent rate (130 mAh·g−1 at 1 C and 54 mAh·g−1 at 5 C) and cycling (capacity retention was 93% after 100 cycles at 0.1 C and 80% after 150 cycles at 0.2 C) performance. This work provides new methods and strategies for the design and fabrication of solid-state batteries with high ionic conductivity and high specific energy. Full article
(This article belongs to the Section Gel Analysis and Characterization)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>,<b>d</b>) Macroscopic picture of PVdF-HFP/LiClO<sub>4</sub> gel electrolyte film. (<b>b</b>) Moreover, 10 μm SEM image of the surface morphology of PVdF-HFP/LiClO<sub>4</sub>. (<b>c</b>) Furthermore, 5 μm SEM image of the surface morphology of PVdF-HFP/LiClO<sub>4</sub>. (<b>e</b>) Moreover, 20 μm SEM image of the cross-section morphology of PVdF-HFP/LiClO<sub>4</sub>. (<b>f</b>) Furthermore, 10 μm SEM image of the cross-section morphology of PVdF-HFP/LiClO<sub>4</sub>. (<b>g</b>) Moreover, 50 μm SEM image and EDS mappings of PVdF-HFP/LiClO<sub>4</sub>.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>d</b>) Macroscopic picture of PVdF-HFP/LPSCl/LiClO<sub>4</sub> gel composite electrolyte film. (<b>b</b>) Moreover, 10 μm SEM images of the surface morphology of PVdF-HFP/LPSCl/LiClO<sub>4</sub> gel composite electrolyte film. (<b>c</b>) Furthermore, 5 μm SEM images of the surface morphology of PVdF-HFP/LPSCl/LiClO<sub>4</sub> gel composite electrolyte film. (<b>e</b>) Moreover, 20 μm SEM images of the cross-sectional morphology of PVdF-HFP/LPSCl/LiClO<sub>4</sub> gel composite electrolyte film. (<b>f</b>) Furthermore, 10 μm SEM images of the cross-sectional morphology of PVdF-HFP/LPSCl/LiClO<sub>4</sub> gel composite electrolyte film. (<b>g</b>) Moreover, 50 μm SEM image and EDS mappings of PVdF-HFP/LiClO<sub>4</sub>/LPSCl distributions in the SCE membrane.</p>
Full article ">Figure 3
<p>(<b>a</b>) Impedance values of the battery electrolyte corresponding to the addition of different proportions of LPSCl for PVdF-HFP/LPSCl/LiClO<sub>4</sub>. (<b>b</b>) Ionic conductivity of the gel composite electrolyte film corresponding to the addition of different proportions of LPSCl. (<b>c</b>) Lithium-ion transference number of PVdF-HFP/LiClO<sub>4</sub> gel composite electrolyte. (<b>d</b>) Lithium-ion transference number of PVdF-HFP/LPSCl/LiClO<sub>4</sub> gel composite electrolyte.</p>
Full article ">Figure 4
<p>(<b>a</b>) DTG results of PVdF-HFP/LiClO<sub>4</sub> gel electrolyte film and PVdF-HFP/LPSCl/LiClO4 gel composite electrolyte film; (<b>b</b>) thermogravimetric analysis curves of PVdF-HFP, PVdF-HFP/LiClO<sub>4</sub>, and PVdF-HFP/LPSCl/LiClO<sub>4</sub> gel composite electrolyte films; (<b>c</b>) X-ray diffraction (XRD) patterns of PVdF-HFP, PVdF-HFP/LiClO<sub>4</sub> gel electrolyte film, and PVdF-HFP/LPSCl/LiClO<sub>4</sub> gel composite electrolyte film; (<b>d</b>) Fourier-transform infrared (FTIR) spectral curves of PVdF-HFP, PVdF-HFP/LiClO<sub>4</sub> gel electrolyte film, and PVdF-HFP/LPSCl/LiClO<sub>4</sub> gel composite electrolyte film.</p>
Full article ">Figure 5
<p>(<b>a</b>) Comparative cycling performance graphs of PVdF-HFP/LiClO<sub>4</sub> and PVdF-HFP/LPSCl/LiClO<sub>4</sub> solid-state batteries at 0.1 C; (<b>b</b>) graphs of current cycling performance of PVdF-HFP/LPSCl/LiClO<sub>4</sub> solid-state batteries at 0.2 C; (<b>c</b>) graphs of PVdF-HFP/LiClO<sub>4</sub> and PVdF-HFP/LPSCl/LiClO<sub>4</sub> solid-state battery performance from 0.1 C to 5 C multiplication rate; (<b>d</b>) charge–discharge curve of PVdF-HFP/LPSCl/LiCLlO<sub>4</sub> solid-state battery at 0.2 C.</p>
Full article ">Figure 6
<p>(<b>a</b>) CV curve of PVdF-HFP/LiClO<sub>4</sub> solid-state battery; (<b>b</b>) CV curve of PVdF-HFP/LPSCl/LiClO<sub>4</sub> solid-state battery.</p>
Full article ">Figure 7
<p>Flowchart of gel composite electrolyte film preparation.</p>
Full article ">
12 pages, 3448 KiB  
Article
Analyzing the Effect of Nano-Sized Conductive Additive Content on Cathode Electrode Performance in Sulfide All-Solid-State Lithium-Ion Batteries
by Jae Hong Choi, Sumyeong Choi, Tom James Embleton, Kyungmok Ko, Kashif Saleem Saqib, Jahanzaib Ali, Mina Jo, Junhyeok Hwang, Sungwoo Park, Minhu Kim, Mingi Hwang, Heesoo Lim and Pilgun Oh
Energies 2024, 17(1), 109; https://doi.org/10.3390/en17010109 - 24 Dec 2023
Cited by 3 | Viewed by 1869
Abstract
All-solid-state lithium-ion batteries (ASSLBs) have recently received significant attention due to their exceptional energy/power densities, inherent safety, and long-term electrochemical stability. However, to achieve energy- and power-dense ASSLBs, the cathode composite electrodes require optimum ionic and electrical pathways and hence the development of [...] Read more.
All-solid-state lithium-ion batteries (ASSLBs) have recently received significant attention due to their exceptional energy/power densities, inherent safety, and long-term electrochemical stability. However, to achieve energy- and power-dense ASSLBs, the cathode composite electrodes require optimum ionic and electrical pathways and hence the development of electrode designs that facilitate such requirements is necessary. Among the various available conductive materials, carbon black (CB) is typically considered as a suitable carbon additive for enhancing electrode conductivity due to its affordable price and electrical-network-enhancing properties. In this study, we examined the effect of different weight percentages (wt%) of nano-sized CB as a conductive additive within a cathode composite made up of Ni-rich cathode material (LiNi0.8Co0.1Mn0.1O2) and solid electrolyte (Li6PS5Cl). Composites including 3 wt%, 5 wt%, and 7 wt% CB were produced, achieving capacity retentions of 66.1%, 65.4%, and 44.6% over 50 cycles at 0.5 C. Despite an increase in electrical conductivity of the 7 wt% CB sample, a significantly lower capacity retention was observed. This was attributed to the increased resistance at the solid electrolyte/cathode material interface, resulting from the presence of excessive CB. This study confirms that an excessive amount of nano-sized conductive material can affect the interfacial resistance between the solid electrolyte and the cathode active material, which is ultimately more important to the electrochemical performance than the electrical pathways. Full article
(This article belongs to the Special Issue Emerging Topics in Future Energy Materials)
Show Figures

Figure 1

Figure 1
<p>Scanning electron microscopy (SEM) images of the surface of (<b>a</b>) 3 wt%, (<b>b</b>) 5 wt%, and (<b>c</b>) 7 wt% carbon black (CB) composites after pelletizing, and the corresponding zoomed-in images of these samples (<b>d</b>–<b>f</b>), respectively.</p>
Full article ">Figure 2
<p>(<b>a</b>) Electrochemical performance of 3 wt%, 5 wt%, and 7 wt% CB composites for the formation (initial charge–discharge) voltage profiles in the voltage range of 2.5–4.3 V at 0.05 C; (<b>b</b>) zoomed-in initial voltage profiles at 0.05 C (<b>c</b>) following cycle performance and coulombic efficiency at 0.5 C.</p>
Full article ">Figure 3
<p>(<b>a</b>) DC polarization (I–V analysis) at 30 °C and (<b>b</b>) calculated electronic conductivity of composite with 3, 5, and 7 wt% of CB.</p>
Full article ">Figure 4
<p>Nyquist plots (0.1 Hz–100 kHz) and their fittings for (<b>a</b>) 3 wt%, (<b>b</b>) 5 wt%, and (<b>c</b>) 7 wt% CB composites comparing results obtained after 50 cycles at 0.5 C.</p>
Full article ">Figure 5
<p>Charge–discharge voltage profiles of composites with (<b>a</b>) 3 wt%, (<b>b</b>) 5 wt%, and (<b>c</b>) 7 wt% CB at 0.5 C cycling in the voltage range of 2.5–4.3 V at 30 °C. The curves were created using the rate data in <a href="#energies-17-00109-f002" class="html-fig">Figure 2</a>c.</p>
Full article ">Figure 6
<p>Scheme of the electrical conductivity of the cathode electrode and ionic resistance between cathode materials and solid electrolyte after fabricating a composite using nano-sized carbon black as a conductive material.</p>
Full article ">
12 pages, 2695 KiB  
Article
Percolation Behavior of a Sulfide Electrolyte–Carbon Additive Matrix for Composite Cathodes in All-Solid-State Batteries
by Elias Reisacher, Pinar Kaya and Volker Knoblauch
Batteries 2023, 9(12), 595; https://doi.org/10.3390/batteries9120595 - 15 Dec 2023
Cited by 1 | Viewed by 2799
Abstract
To achieve high energy densities with sufficient cycling performance in all-solid-state batteries, the fraction of active material has to be maximized while maintaining ionic and electronic conduction throughout the composite cathode. It is well known that low-surface-area carbon additives added to the composite [...] Read more.
To achieve high energy densities with sufficient cycling performance in all-solid-state batteries, the fraction of active material has to be maximized while maintaining ionic and electronic conduction throughout the composite cathode. It is well known that low-surface-area carbon additives added to the composite cathode enhance the rate capability; however, at the same time, they can lead to rapid decomposition of the solid electrolyte in thiophosphate-based cells. Thus, the fraction of such conductive additives has to be well balanced. Within this study we determined the electronic percolation threshold of a conducting matrix consisting of Li6PS5Cl and C65. Furthermore, we systematically investigated the microstructure and effective conductivity (σeff) of the conducting matrix. The percolation threshold pc was determined as ~4 wt.-% C65, and it is suggested that below pc, the ionic contribution is dominant, which can be seen in temperature-dependent σeff and blocked charge transport at low frequencies. Above pc, the impedance of the conducting matrix becomes frequency-independent, and the ohmic law applies. Thus, the conducting matrix in ASSB can be regarded as an electronic and ionic conducting phase between active material particles. Additionally, guidelines are provided to enable electronic conduction in the conducting matrix with minimal C65 content. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns of the precursor mixture after ball milling (black) and the corresponding pattern after reaction induced by heat treatment at 550 °C (red). Additionally, the reference pattern (ICSD: 418490) of Li-argyrodite is plotted. (<b>b</b>) SEM image of the synthesized solid electrolyte after heat treatment. (<b>c</b>) Surface of a pelletized separator. (<b>d</b>) Fabricated separator on the bottom current collector of the measurement cell.</p>
Full article ">Figure 2
<p>(<b>a</b>) Nyquist plot of the synthesized SE in blocking condition at temperatures between 65 °C and 5 °C. (<b>b</b>) Arrhenius plot of Li<sub>6</sub>PS<sub>5</sub>Cl to determine the activation energy for Li-ion transport.</p>
Full article ">Figure 3
<p>Nyquist plot of the conducting matrix with (<b>a</b>) pure SE, (<b>b</b>) CM-1: 1 wt.-% C65, (<b>c</b>) CM-3: 3 wt.-% C65 and (<b>d</b>) CM-4: 4 wt.-% C65 measured between 5 and 65 °C with temperatures steps of 10 °C (7.0 MHz–1.0 Hz).</p>
Full article ">Figure 4
<p>Effective conductivities of the conducting matrix as a function of the temperature. Different colors represent the weight fraction of C65 in the powder mixture.</p>
Full article ">Figure 5
<p>Effective conductivities of the conducting matrix as a function of the C65 weight and volume fraction at 25 °C. The approximate range of the percolation threshold <span class="html-italic">p<sub>c</sub></span> is highlighted. The X-axes are not scaled and display tested C65 fractions.</p>
Full article ">Figure 6
<p>(<b>a</b>) Schema of the conducting matrix microstructure with different weight fractions of C65; (<b>b</b>–<b>e</b>) correlated top-view images of the conducting matrix with increasing C65 content (yellow arrows indicate solid electrolyte particles and dark blue arrows indicate C65 nanoparticles): (<b>b</b>) CM-1, (<b>c</b>) CM-2, (<b>d</b>) CM-5 and (<b>e</b>) CM-10.</p>
Full article ">
12 pages, 10472 KiB  
Article
Micro-Sized MoS6@15%Li7P3S11 Composite Enables Stable All-Solid-State Battery with High Capacity
by Mingyuan Chang, Mengli Yang, Wenrui Xie, Fuli Tian, Gaozhan Liu, Ping Cui, Tao Wu and Xiayin Yao
Batteries 2023, 9(11), 560; https://doi.org/10.3390/batteries9110560 - 17 Nov 2023
Cited by 2 | Viewed by 2343
Abstract
All-solid-state lithium batteries without any liquid organic electrolytes can realize high energy density while eliminating flammability issues. Active materials with high specific capacity and favorable interfacial contact within the cathode layer are crucial to the realization of good electrochemical performance. Herein, we report [...] Read more.
All-solid-state lithium batteries without any liquid organic electrolytes can realize high energy density while eliminating flammability issues. Active materials with high specific capacity and favorable interfacial contact within the cathode layer are crucial to the realization of good electrochemical performance. Herein, we report a high-capacity polysulfide cathode material, MoS6@15%Li7P3S11, with a particle size of 1–4 μm. The MoS6 exhibited an impressive initial specific capacity of 913.9 mAh g−1 at 0.1 A g−1. When coupled with the Li7P3S11 electrolyte coating layer, the resultant MoS6@15%Li7P3S11 composite showed improved interfacial contact and an optimized ionic diffusivity range from 10−12–10−11 cm2 s−1 to 10−11–10−10 cm2 s−1. The Li/Li6PS5Cl/MoS6@15%Li7P3S11 all-solid-state lithium battery delivered ultra-high initial and reversible specific capacities of 1083.8 mAh g−1 and 851.5 mAh g−1, respectively, at a current density of 0.1 A g−1 within 1.0–3.0 V. Even under 1 A g−1, the battery maintained a reversible specific capacity of 400 mAh g−1 after 1000 cycles. This work outlines a promising cathode material with intimate interfacial contact and superior ionic transport kinetics within the cathode layer as well as high specific capacity for use in all-solid-state lithium batteries. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of procedure for preparing MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns of MoS<sub>6</sub> and MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite. (<b>b</b>) Raman spectra of MoS<sub>6</sub> and MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite.</p>
Full article ">Figure 3
<p>SEM images of (<b>a</b>) MoS<sub>6</sub> and (<b>b</b>) MoS<sub>6</sub>@15% Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub>, EDS mapping of (<b>c</b>) MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite, TEM images of (<b>d</b>) MoS<sub>6</sub> and (<b>e</b>) MoS<sub>6</sub>@15% Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub>, and HRTEM images of (<b>f</b>) MoS<sub>6</sub>@15% Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub>.</p>
Full article ">Figure 4
<p>CV curves of (<b>a</b>) MoS<sub>6</sub> and (<b>b</b>) MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite; galvanostatic discharge/charge profiles of (<b>c</b>) MoS<sub>6</sub> and (<b>d</b>) MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite cathodes at 0.1 A g<sup>−1</sup>; (<b>e</b>) cyclic performances of MoS<sub>6</sub> and MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite at 0.1 A g<sup>−1</sup> within 1.0–3.0 V (the solid circles represent discharge capacities); (<b>f</b>) Nyquist plots and equivalent circuit diagram of MoS<sub>6</sub> and MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite cathodes after 1st and 20th cycles at 0.1 A g<sup>−1</sup> within 1.0–3.0 V.</p>
Full article ">Figure 5
<p>(<b>a</b>) Rate performances of MoS<sub>6</sub> and MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite under different current densities. (<b>b</b>) Ragone plot deduced from the rate performances shown in (<b>a</b>). (<b>c</b>) Long-term cyclic performance of MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite at 1 A g<sup>−1</sup>.</p>
Full article ">Figure 6
<p>CV curves at different scan rates for (<b>a</b>) MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite and (<b>c</b>) MoS<sub>6</sub>. The log (peak current) vs. log (scan rate) fitted plots at reduction and oxidation peaks of (<b>b</b>) MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite and (<b>d</b>) MoS<sub>6</sub>.</p>
Full article ">Figure 7
<p>GITT plot of Li/Li<sub>6</sub>PS<sub>5</sub>Cl/MoS<sub>6</sub>@15%Li<sub>7</sub>P<sub>3</sub>S<sub>11</sub> composite and Li/Li<sub>6</sub>PS<sub>5</sub>Cl/MoS<sub>6</sub> all-solid-state lithium batteries.</p>
Full article ">
16 pages, 3832 KiB  
Article
New Insights of Infiltration Process of Argyrodite Li6PS5Cl Solid Electrolyte into Conventional Lithium-Ion Electrodes for Solid-State Batteries
by Artur Tron, Andrea Paolella and Alexander Beutl
Batteries 2023, 9(10), 503; https://doi.org/10.3390/batteries9100503 - 4 Oct 2023
Cited by 2 | Viewed by 4851
Abstract
All-solid-state lithium-ion batteries based on solid electrolytes are attractive for electric applications due to their potential high energy density and safety. The sulfide solid electrolyte (e.g., argyrodite) shows a high ionic conductivity (10−3 S cm−1). There is an open question [...] Read more.
All-solid-state lithium-ion batteries based on solid electrolytes are attractive for electric applications due to their potential high energy density and safety. The sulfide solid electrolyte (e.g., argyrodite) shows a high ionic conductivity (10−3 S cm−1). There is an open question related to the sulfide electrode’s fabrication by simply infiltrating methods applied for conventional lithium-ion battery electrodes via homogeneous solid electrolyte solutions, the structure of electrolytes after drying, chemical stability of binders and electrolyte, the surface morphology of electrolyte, and the deepening of the infiltrated electrolyte into the active materials to provide better contact between the active material and electrolyte and favorable lithium ionic conduction. However, due to the high reactivity of sulfide-based solid electrolytes, unwanted side reactions between sulfide electrolytes and polar solvents may occur. In this work, we explore the chemical and electrochemical properties of the argyrodite-based film produced by infiltration mode by combining electrochemical and structural characterizations. Full article
(This article belongs to the Special Issue Recent Progress in All-Solid-State Lithium Batteries)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic illustration of the infiltration process of sulfide-based solid-electrolyte (Li<sub>6</sub>PS<sub>5</sub>Cl in ethanol solvent) into conventional NCM811-based cathode.</p>
Full article ">Figure 2
<p>Solution of solid electrolyte of Li<sub>6</sub>PS<sub>5</sub>Cl into ethanol (<b>a</b>) before and after shaking: (<b>b</b>) Solid electrolyte treated with ethanol and depicted after drying at RT and after 180 °C for 2 h, (<b>c</b>) XRD patterns of the pristine Li<sub>6</sub>PS<sub>5</sub>Cl materials and after treatment with ethanol and subsequent drying at 80 °C and 180 °C.</p>
Full article ">Figure 3
<p>(<b>a</b>) Ionic conductivities and (<b>b</b>) density of powder pellets (densified at 300 MPa) used for EIS measurements of Li<sub>6</sub>PS<sub>5</sub>Cl samples as well as after treatment with ethanol and subsequent drying at 80 °C and 180 °C for 2 h, and without treatment.</p>
Full article ">Figure 4
<p>SEM images and EDX analysis of (<b>a</b>) pristine Li<sub>6</sub>PS<sub>5</sub>Cl, (<b>b</b>) 11 wt%, (<b>c</b>) 5 wt%, and (<b>d</b>) 0.5 wt% of solid electrolyte after ethanol treatment and drying at 180 °C.</p>
Full article ">Figure 5
<p>XRD patterns of NMC811 electrodes infiltrated with Li<sub>6</sub>PS<sub>5</sub>Cl after drying at (<b>a</b>) room temperature and (<b>b</b>) 120 °C for 2 h in a vacuum.</p>
Full article ">Figure 6
<p>SEM images and cross-section of the NCM electrode materials with immersed solid electrolyte (NCM/solid electrolyte later) and their compatibility via ethanol solution before and after applied pressing.</p>
Full article ">Figure 7
<p>(<b>a</b>) EDX results for the infiltrated NCM811 electrode showing 70% porosity after densification. (<b>b</b>) The elemental distribution for 2 points located close to the electrode surface and close to the current collector is given. (<b>c</b>) Schematic illustration of the infiltration process of Li<sub>6</sub>PS<sub>5</sub>Cl solution into the conventional lithium-ion battery electrodes.</p>
Full article ">Figure 8
<p>(<b>a</b>) Electrochemical impedance spectrum with the equivalent circuit model and (<b>b</b>) Nyquist plots of the NMC811 electrodes infiltrated with Li<sub>6</sub>PS<sub>5</sub>Cl solid electrolyte.</p>
Full article ">Figure 9
<p>First charge and discharge curves for (<b>a</b>) all-solid-state battery, (<b>b</b>) non-infiltrated, and (<b>c</b>) infiltrated cathodes.</p>
Full article ">
19 pages, 4853 KiB  
Article
Li4Ln[PS4]2Cl: Chloride-Containing Lithium Thiophosphates with Lanthanoid Participation (Ln = Pr, Nd and Sm)
by Pia L. Lange, Sebastian Bette, Sabine Strobel, Robert E. Dinnebier and Thomas Schleid
Crystals 2023, 13(10), 1408; https://doi.org/10.3390/cryst13101408 - 22 Sep 2023
Viewed by 1000
Abstract
The synthesis and structural analysis of three new chloride-containing lithium thiophosphates(V) Li4Ln[PS4]2Cl with trivalent lanthanoids (Ln = Pr, Nd and Sm) are presented and discussed. Single crystals of Li4Sm[PS4]2Cl [...] Read more.
The synthesis and structural analysis of three new chloride-containing lithium thiophosphates(V) Li4Ln[PS4]2Cl with trivalent lanthanoids (Ln = Pr, Nd and Sm) are presented and discussed. Single crystals of Li4Sm[PS4]2Cl were obtained and used for crystal structure determination by applying X-ray diffraction. The other compounds were found to crystallize isotypically in the monoclinic space group C2/c. Thus, Li4Sm[PS4]2Cl (a = 2089.31(12) pm, b = 1579.69(9) pm, c = 1309.04(8) pm, β = 109.978(3)°, Z = 12) was used as a representative model to further describe the crystal structure in detail since Li4Pr[PS4]2Cl and Li4Nd[PS4]2Cl were confirmed to be isotypic using powder X-ray diffraction measurements (PXRD). In all cases, a trigonal structure in the space group R3¯ (e.g., a = 1579.67(9) pm, c = 2818.36(16) pm, c/a = 1.784, Z = 18, for Li4Sm[PS4]2Cl) displaying almost identical building units worked initially misleadingly. The structure refinement of Li4Sm[PS4]2Cl revealed bicapped trigonal prisms of sulfur atoms coordinating the two crystallographically distinct (Sm1)3+ and (Sm2)3+ cations, which are further coordinated by four anionic [PS4]3− tetrahedra. The compounds also contain chloride anions residing within channel-like pores made of [PS4]3− units. Eight different sites for Li+ cations were identified with various coordination environments (C.N. = 4–6) with respect to chlorine and sulfur. EDXS measurements supported the stoichiometric formula of Li4Ln[PS4]2Cl, and diffuse reflectance spectroscopy revealed optical band gaps of 2.69 eV, 3.52 eV, and 3.49 eV for Li4Sm[PS4]2Cl, Li4Nd[PS4]2Cl, and Li4Pr[PS4]2Cl, respectively. The activation energy for Li+-cation mobility in Li4Sm[PS4]2Cl was calculated as Ea(Li+) = 0.88 eV using BVEL, which indicates potential as a Li+-cation conductor. Full article
Show Figures

Figure 1

Figure 1
<p>SEM picture of the Li<sub>4</sub>Sm[PS<sub>4</sub>]<sub>2</sub>Cl crystal after the X-ray diffraction experiment. The crystal was exposed for a short time to ambient air.</p>
Full article ">Figure 2
<p>Extended unit-cell content of Li<sub>4</sub>Sm[PS<sub>4</sub>]<sub>2</sub>Cl with highlighted [PS<sub>4</sub>]<sup>3−</sup> tetrahedra as viewed along the <span class="html-italic">c</span>-axis.</p>
Full article ">Figure 3
<p>Extended unit-cell content of Li<sub>4</sub>Sm[PS<sub>4</sub>]<sub>2</sub>Cl with highlighted [PS<sub>4</sub>]<sup>3−</sup> tetrahedra as viewed along [1 0 <math display="inline"><semantics> <mrow> <mover> <mn>1</mn> <mo>¯</mo> </mover> </mrow> </semantics></math>] to visualize the putative trigonal symmetry leading to twinning.</p>
Full article ">Figure 4
<p>Graphical result of the final Rietveld refinement of Li<sub>4</sub>Nd[PS<sub>4</sub>]<sub>2</sub>Cl using the trigonal structure model with space group <span class="html-italic">R</span><math display="inline"><semantics> <mrow> <mover> <mn>3</mn> <mo>¯</mo> </mover> </mrow> </semantics></math> and <span class="html-italic">a</span> = 1584.09(8) pm, <span class="html-italic">c</span> = 2834.49(15) pm for <span class="html-italic">Z</span> = 18. The high-angle part starting at 2<span class="html-italic">θ</span> = 20° is enlarged for clarity.</p>
Full article ">Figure 5
<p>Graphical result of the final Rietveld refinement of Li<sub>4</sub>Nd[PS<sub>4</sub>]<sub>2</sub>Cl using the monoclinic structure model with space group <span class="html-italic">C</span>2/<span class="html-italic">c</span> and a = 2098.54(11) pm, <span class="html-italic">b</span> = 1580.60(8) pm, <span class="html-italic">c</span> = 1318.02(9) pm, <span class="html-italic">β</span> = 110.018(4)° for <span class="html-italic">Z</span> = 12. The high-angle part starting at 2<span class="html-italic">θ</span> = 20° is enlarged for clarity.</p>
Full article ">Figure 6
<p>Graphical result of the final Le Bail fit [<a href="#B27-crystals-13-01408" class="html-bibr">27</a>] for Li<sub>4</sub>Pr[PS<sub>4</sub>]<sub>2</sub>Cl.</p>
Full article ">Figure 7
<p>Coordination environments for the Sm<sup>3+</sup>-cation sites in the monoclinic structure refinement of Li<sub>4</sub>Sm[PS<sub>4</sub>]<sub>2</sub>Cl (Sm1, (<b>a</b>); Sm2, (<b>b</b>) and for the trigonal refinement for the single Sm<sup>3+</sup> site, (<b>c</b>).</p>
Full article ">Figure 8
<p>(<b>a</b>) [PS<sub>4</sub>]<sup>3−</sup>-connected network of [SmS<sub>8</sub>]<sup>13−</sup> polyhedra in both structure descriptions of Li<sub>4</sub>Sm[PS<sub>4</sub>]<sub>2</sub>Cl and (<b>b</b>) dimers of (Cl2)<sup>−</sup>-centered (Li<sup>+</sup>)<sub>5</sub> square pyramids sharing an edge.</p>
Full article ">Figure 9
<p><span class="html-italic">Kubelka</span>-<span class="html-italic">Munk</span> plot from DRS data to determine the optical band gap to be 3.49 eV with assigned <span class="html-italic">f → f</span> transitions for the bulk praseodymium material Li<sub>4</sub>Pr[PS<sub>4</sub>]<sub>2</sub>Cl.</p>
Full article ">Figure 10
<p><span class="html-italic">Kubelka</span>-<span class="html-italic">Munk</span> plot from DRS data to determine the optical band gap to be 3.52 eV with assigned <span class="html-italic">f → f</span> transitions for the bulk neodymium material Li<sub>4</sub>Nd[PS<sub>4</sub>]<sub>2</sub>Cl.</p>
Full article ">Figure 11
<p><span class="html-italic">Kubelka</span>-<span class="html-italic">Munk</span> plot from DRS data to determine the optical band gap to be 2.69 eV with assigned <span class="html-italic">f → f</span> transitions for the bulk samarium material Li<sub>4</sub>Sm[PS<sub>4</sub>]<sub>2</sub>Cl.</p>
Full article ">Figure 12
<p>Bond-valence energy-landscape isosurfaces with different depth levels for Li<sub>4</sub>Sm[PS<sub>4</sub>]<sub>2</sub>Cl (monoclinic, <span class="html-italic">C</span>2/<span class="html-italic">c</span>) with levels 0.3 (<b>a</b>), 1.9 (<b>b</b>) and 2.3 (<b>c</b>).</p>
Full article ">
10 pages, 4208 KiB  
Article
Optimization of Annealing Process of Li6PS5Cl for All-Solid-State Lithium Batteries by Box–Behnken Design
by Zhihua Zhang, Yan Chai, De Ning, Jun Wang, Dong Zhou and Yongli Li
Batteries 2023, 9(9), 480; https://doi.org/10.3390/batteries9090480 - 21 Sep 2023
Cited by 2 | Viewed by 2413
Abstract
Li6PS5Cl possesses high ionic conductivity and excellent interfacial stability to electrodes and is known as a promising solid-state electrolyte material for all-solid-state batteries (ASSBs). However, the optimal annealing process of Li6PS5Cl has not been studied [...] Read more.
Li6PS5Cl possesses high ionic conductivity and excellent interfacial stability to electrodes and is known as a promising solid-state electrolyte material for all-solid-state batteries (ASSBs). However, the optimal annealing process of Li6PS5Cl has not been studied systematically. Here, a Box–Behnken design is used to investigate the interactions of the heating rate, annealing temperature, and duration of annealing process for Li6PS5Cl to optimize the ionic conductivity. The response surface methodology with regression analysis is employed for simulating the data obtained, and the optimized parameters are verified in practice. As a consequence, Li6PS5Cl delivers a rather high conductivity of 4.45 mS/cm at 25 °C, and ASSB consisting of a LiNi0.6Co0.2Mn0.2O2 cathode and lithium anode shows a high initial discharge capacity of 151.7 mAh/g as well as excellent cycling performances for more than 350 cycles, highlighting the importance of the design of experiments. Full article
(This article belongs to the Special Issue Electrolytes for Solid State Batteries)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) The synthesis processes of Li<sub>6</sub>PS<sub>5</sub>Cl. (<b>b</b>) Schematic diagram of the BBD. (<b>c</b>) Schematic diagram of the full factorial design.</p>
Full article ">Figure 2
<p>(<b>a</b>) The comparison between the actual and predicted ionic conductivity values of Li<sub>6</sub>PS<sub>5</sub>Cl. (<b>b</b>) Box–Cox plot for power transforms.</p>
Full article ">Figure 3
<p>Contour plot and 3D response surface for effect of (<b>a</b>,<b>b</b>) the heating rate and temperature, (<b>c</b>,<b>d</b>) the heating rate and duration, and (<b>e</b>,<b>f</b>) the heating rate and temperature on the ionic conductivity of Li<sub>6</sub>PS<sub>5</sub>Cl.</p>
Full article ">Figure 4
<p>(<b>a</b>) Nyquist plots and electronic conductivity test, and (<b>b</b>) SEM of the Li<sub>6</sub>PS<sub>5</sub>Cl synthesized under the optimum annealing process. Rietveld refinement of XRD data for the pristine precursor mixture annealed under (<b>c</b>) condition 1 in <a href="#batteries-09-00480-t002" class="html-table">Table 2</a> and (<b>d</b>) the optimum condition. (<b>e</b>) Charge–discharge curves and (<b>f</b>) cyclic performance of the NCM/Li ASSB using the optimized Li<sub>6</sub>PS<sub>5</sub>Cl.</p>
Full article ">
15 pages, 3995 KiB  
Article
Ionic Conductivity of the Li6PS5Cl0.5Br0.5 Argyrodite Electrolyte at Different Operating and Pelletizing Pressures and Temperatures
by Joshua Dunham, Joshua Carfang, Chan-Yeop Yu, Raziyeh Ghahremani, Rashid Farahati and Siamak Farhad
Energies 2023, 16(13), 5100; https://doi.org/10.3390/en16135100 - 1 Jul 2023
Viewed by 2131
Abstract
All-solid-state lithium batteries (ASSLBs) using argyrodite electrolyte materials have shown promise for applications in electric vehicles (EVs). However, understanding the effects of processing parameters on the ionic conductivity of these electrolytes is crucial for optimizing battery performance and manufacturing methods. This study investigates [...] Read more.
All-solid-state lithium batteries (ASSLBs) using argyrodite electrolyte materials have shown promise for applications in electric vehicles (EVs). However, understanding the effects of processing parameters on the ionic conductivity of these electrolytes is crucial for optimizing battery performance and manufacturing methods. This study investigates the influence of electrolyte operating temperature, electrolyte operating pressure, electrolyte pelletization pressure, and electrolyte pelletizing temperature on the ionic conductivity of the Li6PS5Cl0.5Br0.5 argyrodite electrolyte (AmpceraTM, D50 = 10 µm). A specially designed test cell is employed for the experimental measurements, allowing for controlled pelletization and testing within the same tooling. The results demonstrate the significant impact of the four parameters on the ionic conductivity of the argyrodite electrolyte. The electrolyte operating temperature has a more pronounced effect than operating pressure, and pelletizing temperature exerts a greater influence than pelletizing pressure. This study provides graphs that aid in understanding the interplay between these parameters and achieving desired conductivity values. It also establishes a baseline for the maximum pelletizing temperature before undesirable degradation of the electrolyte occurs. By manipulating the pelletizing pressure, operating pressure, and pelletizing temperature, battery engineers can achieve the desired conductivity for specific applications. The findings emphasize the need to consider operating conditions to ensure satisfactory low-temperature performance, particularly for EVs. Overall, this study provides valuable insights into processing and operating conditions for ASSLBs utilizing the Li6PS5Cl0.5Br0.5 argyrodite electrolyte. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental setup.</p>
Full article ">Figure 2
<p>Nyquist plot for Ampcera Li6PS<sub>5</sub>Cl<sub>0.5</sub>Br<sub>0.5</sub> at −20 °C, 540 MPa pelletizing and 250 MPa operating pressure. Results indicate that there is a grain boundary resistance of 90 Ω and a bulk resistance of 105 Ω based on the Zview fit, which means that the total resistance is 195 Ω. This is the value that was used to calculate the ionic conductivity of the electrolyte. This figure is for visual explanation only; the accuracy of the bulk and grain boundary measurements have not been studied or verified.</p>
Full article ">Figure 3
<p>Electrolyte thickness versus the pelletization pressure for 100 mg electrolyte powder and the die with a circular diameter of 10 mm.</p>
Full article ">Figure 4
<p>(<b>a</b>) Ionic conductivity of the Li<sub>6</sub>PS<sub>5</sub>Cl<sub>0.5</sub>Br<sub>0.5</sub> electrolyte versus the cell operating pressure at different pelletization pressures and at room temperature (25 °C). (<b>b</b>) Ionic conductivity of the Li6PS5Cl0.5Br0.5 electrolyte versus the cell pelletizing pressure at different operating pressures and at room temperature (25 °C).</p>
Full article ">Figure 5
<p>Ionic conductivity of the Li6PS5Cl0.5Br0.5 electrolyte versus the cell operating pressure at different battery operating temperatures and pelletization pressures of (<b>a</b>) 180 MPa (<b>b</b>) 540 MPa and (<b>c</b>) 900 MPa. Error bars are presented to show one standard deviation range for each result.</p>
Full article ">Figure 5 Cont.
<p>Ionic conductivity of the Li6PS5Cl0.5Br0.5 electrolyte versus the cell operating pressure at different battery operating temperatures and pelletization pressures of (<b>a</b>) 180 MPa (<b>b</b>) 540 MPa and (<b>c</b>) 900 MPa. Error bars are presented to show one standard deviation range for each result.</p>
Full article ">Figure 6
<p>Arrhenius ionic conductivity plot of Ampcera<sup>TM</sup> Li<sub>6</sub>PS<sub>5</sub>Cl<sub>0.5</sub>Br<sub>0.5</sub> at 250 MPa from −20 to 75 °C. The activation energy was found to be 0.275 eV.</p>
Full article ">Figure 7
<p>Ionic conductivity of the Li<sub>6</sub>PS<sub>5</sub>Cl<sub>0.5</sub>Br<sub>0.5</sub> electrolyte versus the cell operating pressure at different battery pelletizing temperatures and operating pressures at a constant pelletizing pressure of 540 MPa.</p>
Full article ">Figure 8
<p>Ionic conductivity of the Li<sub>6</sub>PS<sub>5</sub>Cl<sub>0.5</sub>Br<sub>0.5</sub> electrolyte versus the cell operating pressure at different battery pelletizing pressures and 100 °C pelletizing temperature compared with the best room-temperature pelletizing temperature result.</p>
Full article ">
10 pages, 2854 KiB  
Article
LiNi0.6Co0.2Mn0.2O2 Cathode-Solid Electrolyte Interfacial Behavior Characterization Using Novel Method Adopting Microcavity Electrode
by Rahul S. Ingole, Rajesh Rajagopal, Orynbassar Mukhan, Sung-Soo Kim and Kwang-Sun Ryu
Molecules 2023, 28(8), 3537; https://doi.org/10.3390/molecules28083537 - 17 Apr 2023
Cited by 1 | Viewed by 1981
Abstract
Due to the limitations of organic liquid electrolytes, current development is towards high performance all-solid-state lithium batteries (ASSLBs). For high performance ASSLBs, the most crucial is the high ion-conducting solid electrolyte (SE), with a focus on interface analysis between SE and active materials. [...] Read more.
Due to the limitations of organic liquid electrolytes, current development is towards high performance all-solid-state lithium batteries (ASSLBs). For high performance ASSLBs, the most crucial is the high ion-conducting solid electrolyte (SE), with a focus on interface analysis between SE and active materials. In the current study, we successfully synthesized the high ion-conductive argyrodite-type (Li6PS5Cl) solid electrolyte, which has 4.8 mS cm−1 conductivity at room temperature. Additionally, the present study suggests the quantitative analysis of interfaces in ASSLBs. The measured initial discharge capacity of a single particle confined in a microcavity electrode was 1.05 nAh for LiNi0.6Co0.2Mn0.2O2 (NCM622)-Li6PS5Cl solid electrolyte materials. The initial cycle result shows the irreversible nature of active material due to the formation of the solid electrolyte interphase (SEI) layer on the surface of the active particle; further second and third cycles demonstrate high reversibility and good stability. Furthermore, the electrochemical kinetic parameters were calculated through the Tafel plot analysis. From the Tafel plot, it is seen that asymmetry increases gradually at high discharge currents and depths, which rise asymmetricity due to the increasing of the conduction barrier. However, the electrochemical parameters confirm the increasing conduction barrier with increased charge transfer resistance. Full article
(This article belongs to the Special Issue Recent Progress in Nanomaterials in Electrochemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The crystal structure, (<b>b</b>) powder X-ray diffraction pattern, (<b>c</b>) laser Raman spectrum, and (<b>d</b>) FE-SEM image of the Li<sub>6</sub>PS<sub>5</sub>Cl solid electrolyte.</p>
Full article ">Figure 2
<p>(<b>a</b>) Nyquist plot and (<b>b</b>) Impedance plots for conductivity data between 30 and 100 °C (insets) and the Arrhenius plot of the prepared Li<sub>6</sub>PS<sub>5</sub>Cl solid electrolyte.</p>
Full article ">Figure 3
<p>(<b>a</b>) Initial three cycles of NCM622-Li<sub>6</sub>PS<sub>5</sub>Cl system at 0.2 C current density and (<b>b</b>) differential analysis of initial cycles of NCM622-Li<sub>6</sub>PS<sub>5</sub>Cl system at 0.2 C.</p>
Full article ">Figure 4
<p>Tafel plots of the NCM622-Li<sub>6</sub>PS<sub>5</sub>Cl system at 10% DOD and 50% DOD (red- and royal-colored dashed lines are Tafel slops upon the charging (blue dots) and discharging (red dots), respectively; violet dashed lines refer to the exchange current densities).</p>
Full article ">Figure 5
<p>(<b>a</b>) The variation in electrochemical parameters (i) charge transfer resistance, (ii) exchange current, and (<b>b</b>) diffusion coefficient with the Li ion transformation during the discharge process.</p>
Full article ">Scheme 1
<p>Schematic illustration of preparation of the Li<sub>6</sub>PS<sub>5</sub>Cl solid electrolyte.</p>
Full article ">
9 pages, 3341 KiB  
Article
Sn-Substituted Argyrodite Li6PS5Cl Solid Electrolyte for Improving Interfacial and Atmospheric Stability
by Seul-Gi Kang, Dae-Hyun Kim, Bo-Joong Kim and Chang-Bun Yoon
Materials 2023, 16(7), 2751; https://doi.org/10.3390/ma16072751 - 29 Mar 2023
Cited by 3 | Viewed by 3363
Abstract
Sulfide-based solid electrolytes exhibit good formability and superior ionic conductivity. However, these electrolytes can react with atmospheric moisture to generate H2S gas, resulting in performance degradation. In this study, we attempted to improve the stability of the interface between Li metal [...] Read more.
Sulfide-based solid electrolytes exhibit good formability and superior ionic conductivity. However, these electrolytes can react with atmospheric moisture to generate H2S gas, resulting in performance degradation. In this study, we attempted to improve the stability of the interface between Li metal and an argyrodite Li6Ps5Cl solid electrolyte by partially substituting P with Sn to form an Sn–S bond. The solid electrolyte was synthesized via liquid synthesis instead of the conventional mechanical milling method. X-ray diffraction analyses confirmed that solid electrolytes have an argyrodite structure and peak shift occurs as substitution increases. Scanning electron microscopy and energy-dispersive X-ray spectroscopy analyses confirmed that the particle size gradually increased, and the components were evenly distributed. Moreover, electrochemical impedance spectroscopy and DC cycling confirmed that the ionic conductivity decreased slightly but that the cycling behavior was stable for about 500 h at X = 0.05. The amount of H2S gas generated when the solid electrolyte is exposed to moisture was measured using a gas sensor. Stability against atmospheric moisture was improved. In conclusion, liquid-phase synthesis could be applied for the large-scale production of argyrodite-based Li6PS5Cl solid electrolytes. Moreover, Sn substitution improved the electrochemical stability of the solid electrolyte. Full article
(This article belongs to the Special Issue Applications of High-Performance Electrolyte Materials in Batteries)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the liquid-phase synthesis process of solid electrolyte.</p>
Full article ">Figure 2
<p>(<b>a</b>) X-ray diffraction (XRD) patterns of LPSCl-xSn with different amounts of Sn substitution (x = 0, 0.025, 0.05, 0.075, 0.1) (<b>b</b>) Shift in the XRD patterns (23<math display="inline"><semantics> <mo>°</mo> </semantics></math> &lt; 2<math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> </mrow> </semantics></math>&lt; <math display="inline"><semantics> <mrow> <mn>34</mn> <mo>°</mo> </mrow> </semantics></math>).</p>
Full article ">Figure 3
<p>Scanning electron microscopy images of LPSCl-xSn sulfide-based electrolytes with different amounts of Sn substitution (<b>a</b>) x = 0, (<b>b</b>) x = 0.025, (<b>c</b>) x = 0.05, (<b>d</b>) x = 0.075, (<b>e</b>) x = 0.1.</p>
Full article ">Figure 4
<p>Energy-dispersive X-ray spectroscopy images of LPSCl-xSn sulfide-based electrolytes (<b>a</b>) x = 0 (<b>b</b>) x = 0.025.</p>
Full article ">Figure 5
<p>Ionic conductivity of LPSCl-xSn sulfide-based electrolytes with different amounts of Sn substitution (x = 0, 0.025, 0.05, 0.075, 0.1).</p>
Full article ">Figure 6
<p>(<b>a</b>) Arrhenius plots of the LPSCl-xSn sulfide-based electrolytes with different amounts of Sn substitution; (<b>b</b>) Change trend of the activation energy against different substitution percentages.</p>
Full article ">Figure 7
<p>Li–Li symmetric cells performance of LPSCl-xSn sulfide-based electrolytes with different amounts of Sn substitution (x = 0, x = 0.05).</p>
Full article ">Figure 8
<p><math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mn>2</mn> </msub> <mi mathvariant="normal">S</mi> </mrow> </semantics></math> gas emission amounts when LPSCl-xSn sulfide electrolytes with different Sn substitution amounts (x = 0, 0.025, x = 0.05, 0.075, 0.1) are exposed to moisture in the atmosphere.</p>
Full article ">
12 pages, 3714 KiB  
Article
Li7La3Zr2O12-co-LiNbO3 Surface Modification Improves the Interface Stability between Cathode and Sulfide Solid-State Electrolyte in All-Solid-State Batteries
by Shishuo Liang, Dong Yang, Jianhua Hu, Shusen Kang, Xue Zhang and Yanchen Fan
Membranes 2023, 13(2), 216; https://doi.org/10.3390/membranes13020216 - 9 Feb 2023
Cited by 2 | Viewed by 2633
Abstract
With the rapid development of energy storage and electric vehicles, thiophosphate-based all-solid-state batteries (ASSBs) are considered the most promising power source. In order to commercialize ASSBs, the interfacial problem between high-voltage cathode active materials and thiophosphate-based solid-state electrolytes needs to be solved in [...] Read more.
With the rapid development of energy storage and electric vehicles, thiophosphate-based all-solid-state batteries (ASSBs) are considered the most promising power source. In order to commercialize ASSBs, the interfacial problem between high-voltage cathode active materials and thiophosphate-based solid-state electrolytes needs to be solved in a simple, effective way. Surface coatings are considered the most promising approach to solving the interfacial problem because surface coatings could prevent direct physical contact between cathode active materials and thiophosphate-based solid-state electrolytes. In this work, Li7La3Zr2O12 (LLZO) and LiNbO3 (LNO) coatings for LiCoO2 (LCO) were fabricated by in-situ interfacial growth of two high-Li+ conductive oxide electrolytes on the LCO surface and tested for thiophosphate-based ASSBs. The coatings were obtained from a two-step traditional sol–gel coatings process, the inner coatings were LNO, and the surface coatings were LLZO. Electrochemical evaluations confirmed that the two-layer coatings are beneficial for ASSBs. ASSBs containing LLZO-co-LNO coatings LiCoO2 (LLZO&LNO@LCO) significantly improved long-term cycling performance and discharge capacity compared with those assembled from uncoated LCO. LLZO&LNO@LCO||Li6PS5Cl (LPSC)||Li-In delivered discharge capacities of 138.8 mAh/g, 101.8 mAh/g, 60.2 mAh/g, and 40.2 mAh/g at 0.05 C, 0.1 C, 0.2 C, and 0.5 C under room temperature, respectively, and better capacity retentions of 98% after 300 cycles at 0.05 C. The results highlighted promising low-cost and scalable cathode material coatings for ASSBs. Full article
Show Figures

Figure 1

Figure 1
<p>Illustration of the synthesis of LLZO&amp;LNO@LCO.</p>
Full article ">Figure 2
<p>SEM image of LLZO&amp;LNO@LCO with different magnification, 1000 (<b>A</b>), 3000 (<b>B</b>), 5000 (<b>C</b>), 10,000 (<b>D</b>).</p>
Full article ">Figure 3
<p>(<b>A</b>) selected particles, and elemental O (<b>B</b>), Co (<b>C</b>), Zr (<b>D</b>), Nb (<b>E</b>), and La (<b>F</b>) mapping of LLZO&amp;LNO@LCO.</p>
Full article ">Figure 4
<p>XRD pattern (<b>A</b>) and Raman spectra (<b>B</b>) of LLZO&amp;LNO@LCO and LCO.</p>
Full article ">Figure 5
<p>Volume resistance and density of LLZO&amp;LNO@LCO (<b>A</b>) and LCO (<b>B</b>) under different pressures.</p>
Full article ">Figure 6
<p>Charging and discharging curve of LCO||LPSC||Li-In (<b>A</b>) and LLZO&amp;LNO@LCO||LPSC||Li-In (<b>B</b>,<b>C</b>) discharging capacity of LLZO&amp;LNO@LCO at a different rate, (<b>D</b>) lifecycles of LLZO&amp;LNO@LCO||LPSC||Li-In at 0.05 C.</p>
Full article ">Figure 7
<p>EIS curve of LCO||LPSC||Li-In and LLZO&amp;LNO@LCO||LPSC||Li-In before (<b>A</b>) and after 20 cycles (<b>B</b>).</p>
Full article ">Figure 8
<p>Charging GITT profiles (<b>A</b>) and discharging GITT profiles (<b>B</b>) of LLZO&amp;LNO@LCO and LCO samples measured at the first cycle. Charging polarization voltage profiles (<b>C</b>) and discharging polarization GITT profiles (<b>D</b>) of LLZO&amp;LNO@LCO and LCO samples measured at the first cycle.</p>
Full article ">
13 pages, 2584 KiB  
Article
Rational Optimization of Cathode Composites for Sulfide-Based All-Solid-State Batteries
by Artur Tron, Raad Hamid, Ningxin Zhang and Alexander Beutl
Nanomaterials 2023, 13(2), 327; https://doi.org/10.3390/nano13020327 - 12 Jan 2023
Cited by 10 | Viewed by 4988
Abstract
All-solid-state lithium-ion batteries with argyrodite solid electrolytes have been developed to attain high conductivities of 10−3 S cm−1 in studies aiming at fast ionic conductivity of electrolytes. However, no matter how high the ionic conductivity of the electrolyte, the design of [...] Read more.
All-solid-state lithium-ion batteries with argyrodite solid electrolytes have been developed to attain high conductivities of 10−3 S cm−1 in studies aiming at fast ionic conductivity of electrolytes. However, no matter how high the ionic conductivity of the electrolyte, the design of the cathode composite is often the bottleneck for high performance. Thus, optimization of the composite cathode formulation is of utmost importance. Unfortunately, many reports limit their studies to only a few parameters of the whole electrode formulation. In addition, different measurement setups and testing conditions employed for all-solid-state batteries make a comparison of results from mutually independent studies quite difficult. Therefore, a detailed investigation on different key parameters for preparation of cathodes employed in all-solid-state batteries is presented here. Employing a rational approach for optimization of composite cathodes using solid sulfide electrolytes elucidated the influence of different parameters on the cycling performance. First, powder electrodes made without binders are investigated to optimize several parameters, including the active materials’ particle morphology, the nature and amount of the conductive additive, the particle size of the solid electrolyte, as well as the active material-to-solid electrolyte ratio. Finally, cast electrodes are examined to determine the influence of a binder on cycling performance. Full article
(This article belongs to the Special Issue Sulfur Based Nanomaterials for Secondary Batteries)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic outline of the composite cathode optimization by dry- and wet-chemical processing. In addition, the testing conditions are listed.</p>
Full article ">Figure 2
<p>(<b>a</b>) Potential profiles of 1st charge/discharge cycle at C/20 using different NMC811 active materials; (<b>b</b>) EIS spectra before cycling (inset shows full range).</p>
Full article ">Figure 3
<p>(<b>a</b>) Potential profiles of 1st charge/discharge cycle at C/20 using different conducting additives (C65, VGCF); (<b>b</b>) EIS spectra after first discharge step.</p>
Full article ">Figure 4
<p>Specific electronic conductivity vs. the amount of conductive additive plots for composite cathodes using (<b>a</b>) VGCF and (<b>b</b>) C65. The dashed red lines are sigmoidal fits of the data and indicate the percolation threshold.</p>
Full article ">Figure 5
<p>Potential profiles of 1st charge/discharge cycle at C/20 using different Li<sub>6</sub>PS<sub>5</sub>Cl materials for the composite NMC811 cathode (AM:SE:CA = 67:30:3 wt.%; 40:55:5 vol.%).</p>
Full article ">Figure 6
<p>Potential profiles of composite cathodes using different AM:SE ratios (vol./vol.): (<b>a</b>) Specific charge values are given with respect to the active material and (<b>b</b>) with respect to the electrode material (AM+SE+CA).</p>
Full article ">Figure 7
<p>Potential profile of a cast composite cathode is compared with the profiles of two powder cathodes.</p>
Full article ">Figure 8
<p>Specific charge values with respect to (<b>a</b>) the active material, (<b>b</b>) the whole electrode composite of cathodes using conventional (AM+CA+SE), and sequential mixing AM+(CA+SE) for different AM:SE ratios.</p>
Full article ">
Back to TopTop