[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

Journals

Article Types

Countries / Regions

Search Results (18)

Search Parameters:
Keywords = lithium hexafluorophosphate

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 3404 KiB  
Article
Unravelling Lithium Interactions in Non-Flammable Gel Polymer Electrolytes: A Density Functional Theory and Molecular Dynamics Study
by Nasser AL-Hamdani, Paula V. Saravia, Javier Luque Di Salvo, Sergio A. Paz and Giorgio De Luca
Batteries 2025, 11(1), 27; https://doi.org/10.3390/batteries11010027 - 14 Jan 2025
Viewed by 448
Abstract
Lithium metal batteries (LiMBs) have emerged as extremely viable options for next-generation energy storage owing to their elevated energy density and improved theoretical specific capacity relative to traditional lithium batteries. However, safety concerns, such as the flammability of organic liquid electrolytes, have limited [...] Read more.
Lithium metal batteries (LiMBs) have emerged as extremely viable options for next-generation energy storage owing to their elevated energy density and improved theoretical specific capacity relative to traditional lithium batteries. However, safety concerns, such as the flammability of organic liquid electrolytes, have limited their extensive application. In the present study, we utilize molecular dynamics and Density Functional Theory based simulations to investigate the Li interactions in gel polymer electrolytes (GPEs), composed of a 3D cross-linked polymer matrix combined with two different non-flammable electrolytes: 1 M lithium hexafluorophosphate (LiPF6) in ethylene carbonate (EC)/dimethyl carbonate (DMC) and 1 M lithium bis(fluorosulfonyl)imide (LiFSI) in trimethyl phosphate (TMP) solvents. The findings derived from radial distribution functions, coordination numbers, and interaction energy calculations indicate that Li⁺ exhibits an affinity with solvent molecules and counter-anions over the functional groups on the polymer matrix, highlighting the preeminent influence of electrolyte components in Li⁺ solvation and transport. Furthermore, the second electrolyte demonstrated enhanced binding energies, implying greater ionic stability and conductivity relative to the first system. These findings offer insights into the Li+ transport mechanism at the molecular scale in the GPE by suggesting that lithium-ion transport does not occur by hopping between polymer functional groups but by diffusion into the solvent/counter anion system. The information provided in the work allows for the improvement of the design of electrolytes in LiMBs to augment both safety and efficiency. Full article
(This article belongs to the Special Issue Advances in Lithium-Ion Battery Safety and Fire)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Monomer structures of the 3D-ADCL-PE; (<b>b</b>,<b>c</b>) structures of the solvents and salts used as electrolytes: E<sub>1</sub> and E<sub>2</sub>, respectively.</p>
Full article ">Figure 2
<p>MD equilibrated GPE: (<b>a</b>) 3D cross-linked polymer plus 1 M Li PF<sub>6</sub> in EC/DMC; (<b>b</b>) 3D cross-linked polymer and 1 M LiFSI in TMP. Boxes are cubic and have the same dimensions in all xyz coordinates.</p>
Full article ">Figure 3
<p>(<b>a</b>) Li-X RDFs associated with all interatomic distances among Li ions and atoms of the GPE-E<sub>1</sub> electrolyte components; (<b>b</b>) RDF associated with the atoms of the functional groups of the cross-linked polymer.</p>
Full article ">Figure 4
<p>(<b>a</b>) CN associated with all atoms interacting with Li<sup>+</sup> in GPE E<sub>1</sub> for the electrolyte atoms; (<b>b</b>) CN of the atoms of the functional groups of the cross-linked polymer.</p>
Full article ">Figure 5
<p>(<b>a</b>) Li-X RDFs associated with all inter-atomic distances among Li ions and atoms of the GPE E<sub>2</sub> for electrolyte components; (<b>b</b>) RDF associated with the atoms of the functional groups of the cross-linked polymer.</p>
Full article ">Figure 6
<p>(<b>a</b>) CN associated with all the atoms interacting with Li<sup>+</sup> in GPE E<sub>2</sub>. for the electrolyte atoms and (<b>b</b>) CN of the atoms of the functional groups of the cross-linked polymer.</p>
Full article ">Figure 7
<p>Cluster structures extracted from MD equilibrated systems: (<b>a</b>) central Li<sup>+</sup> interacting at short distance with (EC:DMC) solvent, (<b>b</b>) central Li<sup>+</sup> interacting at short distance with (DMC:EC) solvent and PF<sub>6</sub><sup>−</sup> counter-anions, (<b>c</b>) Li<sup>+</sup> interacting with TMP solvent, and (<b>d</b>) interacting with TMP solvent and FSI<sup>−</sup> counter-anions. All 4 clusters had parts of the polymer equilibrated box as well. Colors: Li<sup>+</sup> (violet), C (gray), H (white), O (red), F (yellow), and P (orange). Target Li ion is represented with its van der Waals radii while the rest are visualized by the ball-and-stick representation.</p>
Full article ">Figure 8
<p>Li<sup>+</sup> interaction energies for E<sub>1</sub> and E<sub>2</sub> electrolyte compositions. The light blue column refers to the C-E<sub>1</sub> and C-E<sub>2</sub> clusters, while the orange column refers to the C<sub>PF6</sub><sup>−</sup>-E<sub>1</sub> and C<sub>FSI</sub><sup>−</sup>-E<sub>2</sub> ones.</p>
Full article ">Scheme 1
<p>A diagram illustrating the process flow and stages involved in the simulation.</p>
Full article ">
16 pages, 6075 KiB  
Article
Developmental Toxicity and Apoptosis in Zebrafish: The Impact of Lithium Hexafluorophosphate (LiPF6) from Lithium-Ion Battery Electrolytes
by Boyu Yang, Luning Sun, Zheng Peng, Qing Zhang, Mei Lin, Zhilin Peng and Lan Zheng
Int. J. Mol. Sci. 2024, 25(17), 9307; https://doi.org/10.3390/ijms25179307 - 28 Aug 2024
Viewed by 1023
Abstract
With the growing dependence on lithium-ion batteries, there is an urgent need to understand the potential developmental toxicity of LiPF6, a key component of these batteries. Although lithium’s toxicity is well-established, the biological toxicity of LiPF6 has been minimally explored. [...] Read more.
With the growing dependence on lithium-ion batteries, there is an urgent need to understand the potential developmental toxicity of LiPF6, a key component of these batteries. Although lithium’s toxicity is well-established, the biological toxicity of LiPF6 has been minimally explored. This study leverages the zebrafish model to investigate the developmental impact of LiPF6 exposure. We observed morphological abnormalities, reduced spontaneous movement, and decreased hatching and swim bladder inflation rates in zebrafish embryos, effects that intensified with higher LiPF6 concentrations. Whole-mount in situ hybridization demonstrated that the specific expression of the swim bladder outer mesothelium marker anxa5b was suppressed in the swim bladder region under LiPF6 exposure. Transcriptomic analysis disclosed an upregulation of apoptosis-related gene sets. Acridine orange staining further supported significant induction of apoptosis. These findings underscore the environmental and health risks of LiPF6 exposure and highlight the necessity for improved waste management strategies for lithium-ion batteries. Full article
(This article belongs to the Collection Feature Papers in Molecular Toxicology)
Show Figures

Figure 1

Figure 1
<p>Developmental toxicity of LiPF<sub>6</sub> in zebrafish embryos. (<b>A</b>) Light microscopy images at 1–5 dpf showing the range of malformations in embryos exposed to different concentrations of LiPF<sub>6</sub>, including spinal curvature (SC), pericardial edema (PE), yolk sac edema (YSE), skin ulceration (SU), delayed yolk absorption (DYA) and uninflated swim bladder (USB). The scale bar in each image represents 1 mm; (<b>B</b>) LC50 curves at 24, 48, 72, and 96 hpf, showing the lethal concentration of LiPF<sub>6</sub> that results in 50% mortality in the exposed embryos; (<b>C</b>) STM frequency per minute for embryos; (<b>D</b>) hatching rates at 48, 60, and 72 hpf; (<b>E</b>) swim bladder-inflation rates at 96 hpf. Statistical significance between the exposure groups and Con is indicated by: * (<span class="html-italic">p</span> &lt; 0.05), ** (<span class="html-italic">p</span> &lt; 0.01), and **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 2
<p>WISH Analysis of anxa5b in 60 hpf zebrafish exposed to LiPF<sub>6</sub>. (<b>A</b>–<b>D</b>) Representative WISH images of anxa5b expression of 60 hpf zebrafish. Panels show embryos exposed to increasing concentrations of LiPF<sub>6</sub>: control (<b>A</b>), 10 µM (<b>B</b>), 20 µM (<b>C</b>), and 30 µM (<b>D</b>). Corresponding dorsal views, denoted as (<b>A′</b>–<b>D′</b>). The scale bar in each image represents 1 mm; (<b>E</b>) quantitative analysis of WISH signal intensity in the swim bladder mesothelium (n = 6 per group). Statistical significance between the exposure groups and Con is indicated by: **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 3
<p>Transcriptomic analysis of zebrafish exposed to LiPF<sub>6</sub> at 48 hpf. (<b>A</b>) Principal component analysis (PCA) plot of the transcriptome sequencing data from control and LiPF<sub>6</sub>-exposed (40 µM) zebrafish at 48 hpf; (<b>B</b>) volcano plot depicting the DEGs identified from the transcriptome sequencing analysis, with the DEGs identified with a <span class="html-italic">p</span>-value &lt; 0.01 and an absolute fold change &gt; 1.5 (log<sub>2</sub> scale). Pro-apoptotic genes are highlighted, all of which are upregulated DEGs; (<b>C</b>,<b>D</b>) combined dual-axis bar-line charts for GO enrichment analysis of both upregulated (<b>C</b>) and downregulated (<b>D</b>) DEGs. These charts display the top-15 enriched terms ranked by ascending <span class="html-italic">p</span>-value. The bars represent enrichment scores on the upper axis, while the line traces gene counts on the lower axis. Bar coloration indicates the <span class="html-italic">p</span>-value of the enrichment; (<b>E</b>,<b>F</b>) KEGG enrichment-analysis bubble charts for upregulated (<b>E</b>) and downregulated (<b>F</b>) DEGs, displaying the top-15 pathways ranked by ascending <span class="html-italic">p</span>-value.</p>
Full article ">Figure 4
<p>GSEA-KEGG enrichment landscape induced by LiPF<sub>6</sub> exposure. This figure employs a Sankey bubble plot to trace the interplay of core genes across the top-ten upregulated (<b>A</b>) and downregulated (<b>B</b>) GSEA-KEGG terms, ranked by their |NES| from the GSEA-KEGG analysis. The chart highlights genes that are common to at least three pathways, signifying their key roles in the cellular response to LiPF<sub>6</sub>. The Sankey diagram delineates these connections, while the juxtaposed bubble chart provides a quantitative assessment of the enrichment, reflecting the FDR, core gene count, and core gene ratio for each pathway.</p>
Full article ">Figure 5
<p>GSEA of apoptosis-related gene sets in zebrafish exposed to LiPF<sub>6</sub>. (<b>A</b>) Enrichment plots for four GSEA-GO terms related to apoptosis; (<b>B</b>) Enrichment plots for two KEGG-GSEA terms associated with apoptosis; (<b>C</b>) A Sankey bubble plot depicting the overlap of genes across the enriched GO terms, highlighting genes retained in two or more sets.</p>
Full article ">Figure 6
<p>AO staining indicates apoptotic induction in 2 dpf zebrafish by LiPF<sub>6</sub> exposure. (<b>A</b>–<b>E</b>) AO-staining images of zebrafish larvae exposed to increasing concentrations of LiPF<sub>6</sub>: (<b>A</b>) control, (<b>B</b>) 10 µM, (<b>C</b>) 20 µM, (<b>D</b>) 30 µM, and (<b>E</b>) 40 µM. Each panel is annotated with “AO” for fluorescence, “White” for brightfield, and “Merge” for the overlay of fluorescence and brightfield images. The scale bar in each image represents 1 mm; (<b>F</b>) quantitative analysis of AO-fluorescence intensity normalized to the control group (n ≥ 5 per group). Statistical significance between the exposure groups and Con is indicated by: **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 7
<p>Mechanism of apoptosis induced by LiPF<sub>6</sub> exposure in zebrafish. (<b>A</b>) Schematic of key elements in the apoptotic pathways, highlighting the intrinsic and extrinsic routes converging at caspase-7. All gene-encoded elements are depicted in red, signifying upregulation; (<b>B</b>) Log<sub>2</sub>(Fold Change) heatmap of the genes encoding the elements in (<a href="#ijms-25-09307-f007" class="html-fig">Figure 7</a>A), illustrating the upregulation spectrum; (<b>C</b>,<b>D</b>) line graphs illustrating the Log<sub>2</sub>(FPKM) values of genes in the intrinsic (<b>C</b>) and extrinsic (<b>D</b>) apoptotic pathways upon LiPF<sub>6</sub> exposure.</p>
Full article ">
10 pages, 2133 KiB  
Article
Stabilization of the Interface between a PEO-Based Lithium Solid Polymer Electrolyte and a 4-Volt Class Cathode, LiCoO2, by the Addition of LiPF6 as a Lithium Salt
by Sou Taminato, Akino Tsuka, Kento Sobue, Daisuke Mori, Yasuo Takeda, Osamu Yamamoto and Nobuyuki Imanishi
Batteries 2024, 10(4), 140; https://doi.org/10.3390/batteries10040140 - 19 Apr 2024
Viewed by 1746
Abstract
Here, the time dependence of the interfacial resistance for Li/polyethylene oxide (PEO)-Li(CF3SO2)2N (LiTFSI)-LiPF6/LiCoO2 cells was measured to investigate the stabilization effect of LiPF6 on the interface between a solid polymer electrolyte (SPE) and [...] Read more.
Here, the time dependence of the interfacial resistance for Li/polyethylene oxide (PEO)-Li(CF3SO2)2N (LiTFSI)-LiPF6/LiCoO2 cells was measured to investigate the stabilization effect of LiPF6 on the interface between a solid polymer electrolyte (SPE) and a 4-volt class cathode, LiCoO2. Impedance measurements under the applied potentials between 4.1 V and 4.4 V vs. Li/Li+ indicated that the addition of LiPF6 to LiTFSI was effective in improving the stability at high potentials such as 4.4 V vs. Li/Li+. In contrast, the resistance of the non-doped PEO-LiTFSI/LiCoO2 interface increased with time under the lower potential of 4.1 V vs. Li/Li+. Fairly good cycle performance was obtained for the LiPF6-doped cell, even at a cut-off voltage of 4.5 V vs. Li/Li+. Full article
Show Figures

Figure 1

Figure 1
<p>Temperature dependence of the electrical conductivity for PEO<sub>18</sub>(100-x)LiTFSI-xLiPF<sub>6</sub> for various x (wt%).</p>
Full article ">Figure 2
<p>CV curves for (<b>a</b>) Li/ PEO-LiTFSI-LiPF<sub>6</sub>/Au and (<b>b</b>) Li/PEO-LiTFSI-LiPF<sub>6</sub>/Al in the second potential sweep from 2.5 V to 5.0 V vs. Li/Li<sup>+</sup> at 80 °C.</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic pattern of the impedance measurement combined with constant current-constant voltage charging. In this example, the cell was charged to 4.2 V vs. Li/Li<sup>+</sup> by constant current, maintained for 2 h at the voltage (constant voltage charge), and open-circuited, and then, the impedance was measured. The process was repeated 20 times. (<b>b</b>) Assumed equivalent circuit. (<b>c</b>) Time dependence of the impedance spectra for the Li/PEO(LiTFSI 100 wt%)/LiCoO<sub>2</sub> cell at 4.2 V vs. Li/Li<sup>+</sup>. (<b>d</b>) Time dependence of the impedance spectra for the Li/PEO(LiTFSI 80 wt%+LiPF<sub>6</sub> 20 wt%)/LiCoO<sub>2</sub> cell at 4.2 V vs. Li/Li<sup>+</sup>.</p>
Full article ">Figure 4
<p>Variation in interfacial resistance, <span class="html-italic">R</span><sub>LiCoO<sub>2</sub></sub>, with time under various applied potentials for SPEs (<b>a</b>) without LiPF<sub>6</sub> and (<b>b</b>) with LiPF<sub>6</sub>.</p>
Full article ">Figure 5
<p>Charge–discharge curves for the Li/PEO-LiTFSI-LiPF<sub>6</sub>/LiCoO<sub>2</sub> cells with the addition of (<b>a</b>) 0 wt%, (<b>b</b>) 25 wt%, and (<b>c</b>) 100 wt% LiPF<sub>6</sub>, and (<b>d</b>) the cycle performance with cut-off voltages of 2.5 V and 4.4 V with 0.1 C rate at 80 °C. The red lines show the first charge–discharge curves. In the case of PEO:LiPF<sub>6</sub> = 18:1 (<b>c</b>), imperfect penetration of the SPE into the electrode caused a decrease in the first charge capacity because of the low fluent SPE characteristic due to the high concentration of LiPF<sub>6</sub>.</p>
Full article ">Figure 6
<p>Charge–discharge curves for (<b>a</b>) Li/PEO-LiTFSI/LiCoO<sub>2</sub> and (<b>b</b>) Li/PEO-LiPF<sub>6</sub>/LiCoO<sub>2</sub> cells, and (<b>c</b>) the cycle performance. The red lines in (<b>a</b>,<b>b</b>) show the first charge-discharge curves. The cut-off voltage was 2.5–4.5 V with a rate of 0.1 C at 80 °C.</p>
Full article ">
22 pages, 1377 KiB  
Article
Structural and Dynamic Characterization of Li–Ionic Liquid Electrolyte Solutions for Application in Li-Ion Batteries: A Molecular Dynamics Approach
by Michele A. Salvador, Rita Maji, Francesco Rossella, Elena Degoli, Alice Ruini and Rita Magri
Batteries 2023, 9(4), 234; https://doi.org/10.3390/batteries9040234 - 19 Apr 2023
Cited by 2 | Viewed by 2719
Abstract
Pyrrolidinium-based (Pyr) ionic liquids (ILs) have been proposed as electrolyte components in lithium-ion batteries (LiBs), mainly due to their higher electrochemical stability and wider electrochemical window. Since they are not naturally electroactive, in order to allow their use in LiBs, it is necessary [...] Read more.
Pyrrolidinium-based (Pyr) ionic liquids (ILs) have been proposed as electrolyte components in lithium-ion batteries (LiBs), mainly due to their higher electrochemical stability and wider electrochemical window. Since they are not naturally electroactive, in order to allow their use in LiBs, it is necessary to mix the ionic liquids with lithium salts (Li). Li–PF6, Li–BF4, and Li–TFSI are among the lithium salts more frequently used in LiBs, and each anion, namely PF6 (hexafluorophosphate), BF4 (tetrafluoroborate), and TFSI (bis(trifluoromethanesulfonyl)azanide), has its own solvation characteristics and interaction profile with the pyrrolidinium ions. The size of Pyr cations, the anion size and symmetry, and cation–anion combinations influence the Li-ion solvation properties. In this work, we used molecular dynamics calculations to achieve a comprehensive view of the role of each cation–anion combination and of different fractions of lithium in the solutions to assess their relative advantage for Li-ion battery applications, by comparing the solvation and structural properties of the systems. This is the most-comprehensive work so far to consider pyrrolidinium-based ILs with different anions and different amounts of Li: from it, we can systematically determine the role of each constituent and its concentration on the structural and dynamic properties of the electrolyte solutions. Full article
(This article belongs to the Section Battery Modelling, Simulation, Management and Application)
Show Figures

Figure 1

Figure 1
<p>Top row: three of the pyrrolidinium cations (Pyr12, Pyr14, and Pyr16) and the Li-ion; bottom row: anions Cl, BF4, PF6, and TFSI.</p>
Full article ">Figure 2
<p>Density of the neat ionic liquids as a function of the size of the pyrrolidinium aliphatic chain, for all anions.</p>
Full article ">Figure 3
<p>Center-of-mass cation–cation, cation–anion, and anion–anion pair correlation functions (<span class="html-italic">g</span>(<span class="html-italic">r</span>)) for all the cation–anion combinations with: (<b>a</b>) Pyr12 as the cation; (<b>b</b>) Pyr16 as the cation.</p>
Full article ">Figure 4
<p>Coordination number (CN) of all neat ionic liquids.</p>
Full article ">Figure 5
<p>Histograms of the number of anions within 5 Å of one cation.</p>
Full article ">Figure 6
<p>Histograms of the number of cations within 5 Å of one anion.</p>
Full article ">Figure 7
<p>Densities for all systems with different amounts of lithium.</p>
Full article ">Figure 8
<p>Percentage of the increase in density with respect to the neat ionic liquids.</p>
Full article ">Figure 9
<p>Center-of-mass cation–anion <span class="html-italic">g</span>(<span class="html-italic">r</span>) for all the systems with different amounts of lithium.</p>
Full article ">Figure 10
<p>Center-of-mass anion–anion <span class="html-italic">g</span>(<span class="html-italic">r</span>) for all the systems with different amounts of lithium.</p>
Full article ">Figure 11
<p>Center-of-mass Li–anion <span class="html-italic">g</span>(<span class="html-italic">r</span>) for all the systems with different amounts of lithium.</p>
Full article ">Figure 12
<p>Histograms of the number of anions around the Li-ions throughout the trajectories, for all systems. In (<b>a</b>) with 0.25 Li and in (<b>b</b>) with 0.50 Li.</p>
Full article ">Figure 13
<p>Self–diffusion of all species in the systems considered.</p>
Full article ">Figure 14
<p>Transference number (<math display="inline"><semantics> <msub> <mi>t</mi> <mrow> <mi>L</mi> <mi>i</mi> </mrow> </msub> </semantics></math>) as a function of Pyr1X, for the systems with different amounts of lithium.</p>
Full article ">
21 pages, 5681 KiB  
Article
Ammonium and Tetraalkylammonium Salts as Additives for Li Metal Electrodes
by Dario Di Cillo, Luca Bargnesi, Giampaolo Lacarbonara and Catia Arbizzani
Batteries 2023, 9(2), 142; https://doi.org/10.3390/batteries9020142 - 20 Feb 2023
Cited by 4 | Viewed by 2862
Abstract
Lithium metal batteries are considered a promising technology to implement high energy density rechargeable systems beyond lithium-ion batteries. However, the development of dendritic morphology is the basis of safety and performance issues and represents the main limiting factor for using lithium anodes in [...] Read more.
Lithium metal batteries are considered a promising technology to implement high energy density rechargeable systems beyond lithium-ion batteries. However, the development of dendritic morphology is the basis of safety and performance issues and represents the main limiting factor for using lithium anodes in commercial rechargeable batteries. In this study, the electrochemical behaviour of Li metal has been investigated in organic carbonate-based electrolytes by electrochemical impedance spectroscopy measurements and deposition/stripping galvanostatic cycling. Low amounts of tetraalkylammonium hexafluorophosphate salts have been added to the electrolytes with the aim of regulating the lithium deposition/stripping process through the electrostatic shielding effect that improves the lithium deposition. The use of NH4PF6 also determined good lithium deposition/stripping performance due to the chemical modification of the native solid electrolyte interphase via direct reaction with lithium. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>CVs at 20 mV s<sup>−1</sup> of (<b>a</b>) LP30 and (<b>b</b>) PC without and with the addition of 50 mM of NH<sub>4</sub>PF<sub>6</sub>, TMAPF<sub>6</sub>, and TEAPF<sub>6</sub>.</p>
Full article ">Figure 2
<p>CVs at 20 mV s<sup>−1</sup> of (<b>a</b>) 1st CV, (<b>b</b>) 10th CV in LP30, and (<b>c</b>) 1st CV, (<b>d</b>) 10th CV in PC without and with the addition of 50 mM of NH<sub>4</sub>PF<sub>6</sub>, TMAPF<sub>6</sub> and TEAPF<sub>6</sub>.</p>
Full article ">Figure 3
<p>Deposition and stripping cycles of Li||Li symmetric cells with (<b>a</b>) LP30, (<b>b</b>) LP30-NH<sub>4</sub>PF<sub>6</sub>, (<b>c</b>) LP30-TMAPF<sub>6</sub>, (<b>d</b>) LP30-TEAPF<sub>6</sub> electrolytes at 0.125 mA cm<sup>−2</sup> (green), 0.250 mA cm<sup>−2</sup> (red) and 0.500 mA cm<sup>−2</sup> (blue).</p>
Full article ">Figure 4
<p>Deposition and stripping cycles of Li||Li symmetric cells with (<b>a</b>) PC, (<b>b</b>) PC-NH<sub>4</sub>PF<sub>6</sub>, (<b>c</b>) PC-TMAPF<sub>6</sub>, (<b>d</b>) PC-TEAPF<sub>6</sub> electrolytes at 0.125 mA cm<sup>−2</sup> (green), 0.250 mA cm<sup>−2</sup> (red) and 0.500 mA cm<sup>−2</sup> (blue).</p>
Full article ">Figure 5
<p>Electrochemical impedance spectra of Li||Li symmetric cells (<b>a</b>) LP30, (<b>b</b>) LP30-NH<sub>4</sub>PF<sub>6</sub>, (<b>c</b>) LP30-TMAPF<sub>6</sub>, (<b>d</b>) LP30-TEAPF<sub>6</sub> after different cycling time. The green, red and cyan lines in the figures are the fitting results for the t<sub>0</sub>, t<sub>1</sub> and t<sub>2</sub> plots, respectively. The enlarged parts at high frequency of <a href="#batteries-09-00142-f005" class="html-fig">Figure 5</a>a,c,d are in <a href="#batteries-09-00142-f0A1" class="html-fig">Figure A1</a>.</p>
Full article ">Figure 6
<p>Electrochemical impedance spectra of Li||Li symmetric cells (<b>a</b>) PC, (<b>b</b>) PC-NH<sub>4</sub>PF<sub>6</sub>, (<b>c</b>) PC-TMAPF<sub>6</sub>, (<b>d</b>) PC-TEAPF<sub>6</sub> after different cycling time. The green, red and cyan lines in the figures are the fitting results for the t<sub>0</sub>, t<sub>1</sub> and t<sub>2</sub> plots, respectively. The plot at t<sub>2</sub> in <a href="#batteries-09-00142-f006" class="html-fig">Figure 6</a>c has been recorded at the end of the test reported in <a href="#batteries-09-00142-f004" class="html-fig">Figure 4</a>c (after 945 total cycles).</p>
Full article ">Figure 7
<p>Selected overvoltage profiles of lithium deposition/stripping of Li||Li symmetric cells with (<b>a</b>) LP30, (<b>b</b>) LP30-NH<sub>4</sub>PF<sub>6</sub>, (<b>c</b>) LP30-TMAPF<sub>6</sub>, (<b>d</b>) LP30-TEAPF<sub>6</sub> electrolytes at 0.125 mA cm<sup>−2</sup> (green), 0.250 mA cm<sup>−2</sup> (red) and 0.500 mA cm<sup>−2</sup> (blue).</p>
Full article ">Figure 8
<p>Selected overvoltage profiles of lithium deposition/stripping of Li||Li symmetric cells with (<b>a</b>) PC, (<b>b</b>) PC-NH<sub>4</sub>PF<sub>6</sub>, (<b>c</b>) PC-TMAPF<sub>6</sub>, (<b>d</b>) PC-TEAPF<sub>6</sub> electrolytes at 0.125 mA cm<sup>−2</sup> (green), 0.250 mA cm<sup>−2</sup> (red) and 0.500 mA cm<sup>−2</sup> (blue).</p>
Full article ">Figure 9
<p>Nyquist plot of cells with (<b>a</b>) PC and (<b>b</b>) PC-NH<sub>4</sub>PF<sub>6</sub> at different times in rest condition; (<b>c</b>) percentage of the interphase impedance increase over time with respect to the impedance at t<sub>0</sub> in PC (black) and in PC-NH<sub>4</sub>PF<sub>6</sub> (red).</p>
Full article ">Figure 10
<p>X-ray diffraction patterns of lithium samples after immersion in solution containing 50 and 250 mM NH<sub>4</sub>PF<sub>6</sub>. (<b>a</b>) Sample immersed in LP30 with added NH<sub>4</sub>PF<sub>6</sub>, (<b>b</b>) enlargement of signal at 22°; (<b>c</b>) sample immersed in PC with added NH<sub>4</sub>PF<sub>6</sub>.</p>
Full article ">Figure 11
<p>SEM images of pristine lithium (<b>a</b>,<b>b</b>) and cycled samples in (<b>c</b>,<b>d</b>) PC, (<b>e</b>,<b>f</b>) PC-NH<sub>4</sub>PF<sub>6</sub>, (<b>g</b>,<b>h</b>) PC-TMAPF<sub>6</sub>, and (<b>i</b>,<b>j</b>) PC-TEAPF<sub>6</sub> at two magnification.</p>
Full article ">Scheme 1
<p>Equivalent circuit.</p>
Full article ">Figure A1
<p>Enlarged view at high frequency of the Nyquist plot of <a href="#batteries-09-00142-f005" class="html-fig">Figure 5</a>a,c,d. The green, red and cyan lines in the figures are the fitting results for the t<sub>0</sub>, t<sub>1</sub> and t<sub>2</sub> plots, respectively.</p>
Full article ">Figure A2
<p>Lithium disk samples immersed for 2 days in (<b>a</b>) LP30 and (<b>b</b>) PC, both with the addition of 250 mM NH<sub>4</sub>PF<sub>6</sub>.</p>
Full article ">Figure A3
<p>Capacity retention (full circles) and coulombic efficiency (plein circles) data of Li||NMC cells with PC and PC-NH<sub>4</sub>PF<sub>6</sub> electrolytes over testing at C/2, after two formation cycles at C/10 C-rate.</p>
Full article ">
14 pages, 2206 KiB  
Article
Strain Compensation Methods for Fiber Bragg Grating Temperature Sensors Suitable for Integration into Lithium-Ion Battery Electrolyte
by Johanna Unterkofler, Gregor Glanz, Markus Koller, Reinhard Klambauer and Alexander Bergmann
Batteries 2023, 9(1), 34; https://doi.org/10.3390/batteries9010034 - 3 Jan 2023
Cited by 8 | Viewed by 3016
Abstract
Temperature is a crucial factor for the safe operation of lithium-ion batteries. During operation, the internal temperature rises above the external temperature due to poor inner thermal conductivity. Various sensors have been proposed to detect the internal temperature, including fiber Bragg grating sensors. [...] Read more.
Temperature is a crucial factor for the safe operation of lithium-ion batteries. During operation, the internal temperature rises above the external temperature due to poor inner thermal conductivity. Various sensors have been proposed to detect the internal temperature, including fiber Bragg grating sensors. However, to the authors’ knowledge, there is no detailed description of the encapsulation of the fiber Bragg grating sensor in the literature to shield it from strain. In this study, different encapsulation methods for strain compensation were compared to find the encapsulation material most compatible with the electrolyte. For this, we stored the proposed sensors with different encapsulation methods in ethylene carbonate:ethyl methyl carbonate (EC:EMC) 3:7 with LiPF6 (lithium hexafluorophosphate) electrolyte and applied temperature changes. After evaluating the sensor encapsulation methods in terms of handling, diameter, uncertainty, usability, and hysteresis behavior, the most suitable sensor encapsulation was found to be a fused silica capillary with polyimide coating. Full article
Show Figures

Figure 1

Figure 1
<p>The investigated sensor samples (SP); SP1: optical fiber in PEEK capillary, single-point mounting; SP2: optical fiber in fused silica tube with polyimide coating, single-point mounting; SP3: optical fiber in fused silica tube, removed polyimide coating, single-point mounting; SP4: optical fiber in fused silica tube, removed polyimide coating, double-point mounting; vertical lines indicate the FBGs; gray circles indicate adhesive points; gray circles with an X indicate a connection between encapsulation and the fiber.</p>
Full article ">Figure 2
<p>Setup of the electrolyte measurement in the climatic chamber. Test tubes filled with the electrolyte and the inserted fibers in a test tube holder.</p>
Full article ">Figure 3
<p>Standard calibration in air; FBG3 from each sensor. SP1: optical fiber in PEEK capillary, single-point mounting; SP2: optical fiber in fused silica tube with polyimide coating, single-point mounting; SP3: optical fiber in fused silica tube, removed polyimide coating, single-point mounting; SP4: optical fiber in fused silica tube, removed polyimide coating, double-point mounting.</p>
Full article ">Figure 4
<p>Calibration curve in air for FBG3 from each sensor. SP1: optical fiber in PEEK capillary, single-point mounting; SP2: optical fiber in fused silica tube with polyimide coating, single-point mounting; SP3: optical fiber in fused silica tube, removed polyimide coating, single-point mounting; SP4: optical fiber in fused silica tube, removed polyimide coating, double-point mounting.</p>
Full article ">Figure 5
<p>Standard calibration in air, all sensors. Temperature calculation with the polynomial fit from <a href="#batteries-09-00034-f004" class="html-fig">Figure 4</a>/ <a href="#batteries-09-00034-t0A2" class="html-table">Table A2</a>. SP1: optical fiber in PEEK capillary, single-point mounting; SP2: optical fiber in fused silica tube with polyimide coating, single-point mounting; SP3: optical fiber in fused silica tube, removed polyimide coating, single-point mounting; SP4: optical fiber in fused silica tube, removed polyimide coating, double-point mounting.</p>
Full article ">Figure 6
<p>Representative behaviors of the FBGs of each sensor’s encapsulation in the electrolyte. SP1: optical fiber in PEEK capillary, single-point mounting; SP2: optical fiber in fused silica tube with polyimide coating, single-point mounting; SP3: optical fiber in fused silica tube, removed polyimide coating, single-point mounting; SP4: optical fiber in fused silica tube, removed polyimide coating, double-point mounting.</p>
Full article ">Figure 7
<p>Calibration curve for the measurement in electrolyte for FBG3 for each sensor. Polynomial fit. SP1: optical fiber in PEEK capillary, single-point mounting; SP2: optical fiber in fused silica tube with polyimide coating, single-point mounting; SP3: optical fiber in fused silica tube, removed polyimide coating, single-point mounting; SP4: optical fiber in fused silica tube, removed polyimide coating, double-point mounting.</p>
Full article ">Figure 8
<p>Temperature calculation using the polynomial fit from <a href="#batteries-09-00034-f007" class="html-fig">Figure 7</a>/<a href="#batteries-09-00034-t0A3" class="html-table">Table A3</a> for each FBG of each sensor encapsulation in the electrolyte. SP1: optical fiber in PEEK capillary, single-point mounting; SP2: optical fiber in fused silica tube with polyimide coating, single-point mounting; SP3: optical fiber in fused silica tube, removed polyimide coating, single-point mounting; SP4: optical fiber in fused silica tube, removed polyimide coating, double-point mounting.</p>
Full article ">Figure 9
<p>Sensitivity comparison between the calibration in air and the electrolyte. SP1: optical fiber in PEEK capillary, single-point mounting; SP2: optical fiber in fused silica tube with polyimide coating, single-point mounting; SP3: optical fiber in fused silica tube, removed polyimide coating, single-point mounting; SP4: optical fiber in fused silica tube, removed polyimide coating, double-point mounting.</p>
Full article ">Figure A1
<p>Split fiber (SP4). Coating peeled off the fiber and the adhesive changed color after electrolyte contact (from transparent to amber-colored). Blue and red marks: FBG positions.</p>
Full article ">
14 pages, 2874 KiB  
Article
Determining the Composition of Carbonate Solvent Systems Used in Lithium-Ion Batteries without Salt Removal
by Mohammad Parhizi, Louis Edwards Caceres-Martinez, Brent A. Modereger, Hilkka I. Kenttämaa, Gozdem Kilaz and Jason K. Ostanek
Energies 2022, 15(8), 2805; https://doi.org/10.3390/en15082805 - 12 Apr 2022
Cited by 1 | Viewed by 4250
Abstract
In this work, two methods were investigated for determining the composition of carbonate solvent systems used in lithium-ion (Li-ion) battery electrolytes. One method was based on comprehensive two-dimensional gas chromatography with electron ionization time-of-flight mass spectrometry (GC×GC/EI TOF MS), which often enables unknown [...] Read more.
In this work, two methods were investigated for determining the composition of carbonate solvent systems used in lithium-ion (Li-ion) battery electrolytes. One method was based on comprehensive two-dimensional gas chromatography with electron ionization time-of-flight mass spectrometry (GC×GC/EI TOF MS), which often enables unknown compound identification by their electron ionization (EI) mass spectra. The other method was based on comprehensive two-dimensional gas chromatography with flame ionization detection (GC×GC/FID). Both methods were used to determine the concentrations of six different commonly used carbonates in Li-ion battery electrolytes (i.e., ethylene carbonate (EC), propylene carbonate (PC), dimethyl carbonate (DMC), diethyl carbonate (DEC), ethyl methyl carbonate (EMC), and vinylene carbonate (VC) in model compound mixtures (MCMs), single-blind samples (SBS), and a commercially obtained electrolyte solution (COES). Both methods were found to be precise (uncertainty < 5%), accurate (error < 5%), and sensitive (limit of detection <0.12 ppm for FID and <2.7 ppm for MS). Furthermore, unlike the previously reported methods, these methods do not require removing lithium hexafluorophosphate salt (LiPF6) from the sample prior to analysis. Removal of the lithium salt was avoided by diluting the electrolyte solutions prior to analysis (1000-fold dilution) and using minimal sample volumes (0.1 µL) for analysis. Full article
(This article belongs to the Topic Energy Storage and Conversion Systems)
Show Figures

Figure 1

Figure 1
<p>Calibration curve of vinylene carbonate (VC) with 95% confidence intervals.</p>
Full article ">Figure 2
<p>Three-dimensional chromatogram exhibiting effective separation of the carbonates in the 100% calibration solution when using the GC×GC/FID system.</p>
Full article ">Figure 3
<p>GC×GC/FID chromatogram of 0.2% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) solution of DMC, EMC, DEC, VC, PC, and EC in acetone.</p>
Full article ">Figure 4
<p>70 eV EI mass spectra obtained by using GC×GC/EI TOF MS for DMC, EMC, DEC, VC, PC, and EC.</p>
Full article ">Figure 5
<p>Expected and measured volume percentages of carbonates in MCM #2 (<b>a</b>) and MCM #3 (<b>b</b>) determined with GC×GC/FID.</p>
Full article ">Figure 6
<p>Volume percentages determined for carbonates in SBS #1 (<b>a</b>), SBS #2 (<b>b</b>), and COES (<b>c</b>).</p>
Full article ">Figure 7
<p>GC×GC/FID chromatograms of acetone before (<b>a</b>) and after (<b>b</b>) ten samples containing LiPF<sub>6</sub> were analyzed.</p>
Full article ">Figure 8
<p>GC×GC/FID chromatograms of salt-containing samples: (<b>a</b>) COES and (<b>b</b>) SBS #2.</p>
Full article ">
18 pages, 4779 KiB  
Article
Optimization of LIB Electrolyte and Exploration of Novel Compounds via the Molecular Dynamics Method
by Ken-ichi Saitoh, Yoshihiro Takai, Tomohiro Sato, Masanori Takuma and Yoshimasa Takahashi
Batteries 2022, 8(3), 27; https://doi.org/10.3390/batteries8030027 - 21 Mar 2022
Cited by 7 | Viewed by 4695
Abstract
Due to great interest in the development of electric vehicles and other applications, improving the performances of lithium-ion batteries (LIBs) is crucial. Specifically, components of electrolytes for LIBs should be adequately chosen from hundreds of thousands of candidate compounds. In this study, we [...] Read more.
Due to great interest in the development of electric vehicles and other applications, improving the performances of lithium-ion batteries (LIBs) is crucial. Specifically, components of electrolytes for LIBs should be adequately chosen from hundreds of thousands of candidate compounds. In this study, we aimed to evaluate some physical properties expected for combinations of molecules for electrolytes by microscopic simulations. That is, the viscosity, ionic conductivity, degree of dissociation, diffusion coefficient, and conformation of each molecule were analyzed via molecular dynamics (MD) simulations. We aimed to understand how molecular-sized structures and properties collaboratively affect the behavior of electrolytes. The practical models of molecules we used were ethylene carbonate (EC), fluoroethylene carbonate (FEC), propylene carbonate (PC), butylene carbonate (BC), γ-butyrolactone (GBL), γ-valerolactone (GVL), dimethyl carbonate (DMC), ethyl-methyl carbonate (EMC), diethyl carbonate (DEC), and lithium hexafluorophosphate (LiPF6). Many molecular systems of electrolytes were simulated, in which one molar LiPF6 was mixed into a single or combined solvent. It was found that small solvent molecules diffused with relative ease, and they contributed to the higher ionic conductivity of electrolytes. It was clarified that the diffusion coefficient of lithium (Li) ions is greatly affected by the surrounding solvent molecules. We can conclude that high-permittivity solvents can be selectively coordinated around Li ions, and Li salts are sufficiently dissociated, even when there is only a small content of high-permittivity solvent. Thus, we can confirm solely by MD simulation that one of the better candidates for solvent molecules, formamide (F), will exhibit higher performance than the current solvents. Full article
Show Figures

Figure 1

Figure 1
<p>An example of a snapshot of molecules in a non-equilibrium MD simulation conducted with a constant electric field.</p>
Full article ">Figure 2
<p>Correlations of physical properties between molecular simulations and experiments (dashed lines represent the equivalence between computations and experiments). (<b>a</b>) For density [<a href="#B44-batteries-08-00027" class="html-bibr">44</a>,<a href="#B45-batteries-08-00027" class="html-bibr">45</a>]; (<b>b</b>) for diffusion coefficient [<a href="#B46-batteries-08-00027" class="html-bibr">46</a>,<a href="#B47-batteries-08-00027" class="html-bibr">47</a>]; (<b>c</b>) for viscosity [<a href="#B46-batteries-08-00027" class="html-bibr">46</a>,<a href="#B47-batteries-08-00027" class="html-bibr">47</a>].</p>
Full article ">Figure 3
<p>Comparison of solvation shell. (<b>a</b>) EC + 1M-LiPF<sub>6</sub>; (<b>b</b>) DMC + 1M-LiPF<sub>6</sub>.</p>
Full article ">Figure 4
<p>Correlation between molecular radius and coordination number of Li<sup>+</sup>.</p>
Full article ">Figure 5
<p>Correlation between molecular radius and degree of dissociation.</p>
Full article ">Figure 6
<p>The curve of the Stokes–Einstein relation using the viscosity and the radius of the solvent molecule, which was calculated by MD simulations.</p>
Full article ">Figure 7
<p>Relationship between <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mrow> <mi>s</mi> <mi>o</mi> <mi>l</mi> <mi>v</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mrow> <mi>c</mi> <mi>a</mi> <mi>t</mi> <mi>i</mi> <mi>o</mi> <mi>n</mi> </mrow> </msub> </mrow> </semantics></math> or <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mrow> <mi>a</mi> <mi>n</mi> <mi>i</mi> <mi>o</mi> <mi>n</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>The correlation between the radii of solvent molecules and their ionic conductivity. The ionic conductivity is inversely proportional to the radius of the solvent, as approximated by the dashed curve.</p>
Full article ">Figure 9
<p>Ionic conductivity in mixed-electrolyte models.</p>
Full article ">Figure 10
<p>Estimation of the solvation shell about Li<sup>+</sup> in mixed electrolytes. (<b>a</b>) EC + DMC + 1M-LiPF<sub>6</sub>; (<b>b</b>) EC + EMC + 1M-LiPF<sub>6</sub>; (<b>c</b>) EC + DEC + 1M-LiPF<sub>6</sub>.</p>
Full article ">Figure 10 Cont.
<p>Estimation of the solvation shell about Li<sup>+</sup> in mixed electrolytes. (<b>a</b>) EC + DMC + 1M-LiPF<sub>6</sub>; (<b>b</b>) EC + EMC + 1M-LiPF<sub>6</sub>; (<b>c</b>) EC + DEC + 1M-LiPF<sub>6</sub>.</p>
Full article ">Figure 11
<p>A schematic of a molecular configuration of the selective solvation (LVS means a solvent molecule with low viscosity, HPS means a solvent molecule with high permittivity).</p>
Full article ">Figure 12
<p>Molecular structure of formamide, the smallest one investigated here, compared with other larger molecules of solvents treated above. (<b>a</b>) Formamide (F); (<b>b</b>) ethylene carbonate (EC); (<b>c</b>) γ-valerolactone (GVL); (<b>d</b>) diethyl carbonate (DEC).</p>
Full article ">Figure 13
<p>Comparison of the ionic conductivity of electrolytes against the radii of their solvents, including the result of formamide (except for F, plots are the same as in <a href="#batteries-08-00027-f008" class="html-fig">Figure 8</a>).</p>
Full article ">Figure 14
<p>RDF around Li<sup>+</sup> in the model of 3F + 7DMC + 1M-LiPF<sub>6</sub> (which is the model named 3F + 7DMC).</p>
Full article ">
15 pages, 4118 KiB  
Article
LLCZN/PEO/LiPF6 Composite Solid-State Electrolyte for Safe Energy Storage Application
by Samuel Adjepong Danquah, Jacob Strimaitis, Clifford F. Denize, Sangram K. Pradhan and Messaoud Bahoura
Batteries 2022, 8(1), 3; https://doi.org/10.3390/batteries8010003 - 7 Jan 2022
Cited by 10 | Viewed by 5365
Abstract
All-solid-state batteries (ASSBs) are gaining traction in the arena of energy storage due to their promising results in producing high energy density and long cycle life coupled with their capability of being safe. The key challenges facing ASSBs are low conductivity and slow [...] Read more.
All-solid-state batteries (ASSBs) are gaining traction in the arena of energy storage due to their promising results in producing high energy density and long cycle life coupled with their capability of being safe. The key challenges facing ASSBs are low conductivity and slow charge transfer kinetics at the interface between the electrode and the solid electrolyte. Garnet solid-state electrolyte has shown promising results in improving the ion conductivity but still suffers from poor capacity retention and rate performance due to the interfacial resistance between the electrodes. To improve the interfacial resistance, we prepared a composite consisting of Li7La2.75Ca0.25Zr1.75Nb0.25O12 (LLCZN) garnet material as the ceramic, polyethylene oxide (PEO) as the polymer, and lithium hexafluorophosphate (LiPF6) as the salt. These compounds are mixed in a stoichiometric ratio and developed into a very thin disc-shaped solid electrolyte. The LLCZN provides a lithium-ion transport path to enhance the lithium-ion conduction during charging and discharging cycles, while the LiPF6 contributes more lithium ions via the transport path. The PEO matrix in the composite material aids in bonding the compounds together and creating a large contact area, thereby reducing the issue of large interfacial resistance. FESEM images show the porous nature of the electrolyte which promotes the movement of lithium ions through the electrolyte. The fabricated LLCZN/PEO/LiPF6 solid-state electrolyte shows outstanding electrochemical stability that remains at 130 mAh g−1 up to 150 charging and discharging cycles at 0.05 mA cm−2 current. All the specific capacities were calculated based on the mass of the cathode material (LiCoO2). In addition, the coin cell retains 85% discharge capacity up to 150 cycles with a Coulombic efficiency of approximately 98% and energy efficiency of 90% during the entire cycling process. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation for the synthesis of (<b>a</b>) LLCZN powder via sol–gel method and (<b>b</b>) composite LLCZN/PEO/LiPF<sub>6</sub> in a ratio of 1:1:1 by weight.</p>
Full article ">Figure 2
<p>Powder X-ray diffraction patterns of as-prepared LLCZN via sol–gel technique and LLCZN/PEO/LiPF<sub>6</sub> composite powder samples.</p>
Full article ">Figure 3
<p>FESEM images of surface morphologies for the composite LLCZN/PEO/LiPF<sub>6</sub> powder for solid-state electrolyte before cycling at different magnifications of (<b>a</b>) 10.0 µm, (<b>b</b>) 20.0 µm, (<b>c</b>) 50 µm, and (<b>d</b>) 100 µm.</p>
Full article ">Figure 4
<p>FTIR spectra of composite, LiPF<sub>6,</sub> PEO, and LLCZN.</p>
Full article ">Figure 5
<p>Thermogravimetric analysis (TGA) of LLCZN/PEO/LiPF<sub>6</sub> composite and the insert is pristine LLCZN powder.</p>
Full article ">Figure 6
<p>Brunauer–Emmett–Teller (BET) analysis of LLCZN/PEO/LiPF<sub>6</sub> composite: (<b>a</b>) adsorption–desorption isotherm curve and (<b>b</b>) pore size distribution.</p>
Full article ">Figure 7
<p>The electrochemical impedance spectra of LLCZN/PEO/LiPF<sub>6</sub> composite with and without e-beam.</p>
Full article ">Figure 8
<p>Charge–discharge voltage–time profiles and electrochemical performance of LiCoO<sub>2</sub>|LLCZN/PEO/LiPF<sub>6</sub>|Li cells with voltage window of 3–4.5 V at different currents (<b>a</b>) 0.05 mA cm<sup>−2</sup>, (<b>b</b>) 0.10 mA cm<sup>−2</sup>, (<b>c</b>) 0.24 mA cm<sup>−2</sup>, (<b>d</b>) 0.50 mA cm<sup>−2</sup>, and (<b>e</b>) 0.05 mA cm<sup>−2</sup> (final).</p>
Full article ">Figure 8 Cont.
<p>Charge–discharge voltage–time profiles and electrochemical performance of LiCoO<sub>2</sub>|LLCZN/PEO/LiPF<sub>6</sub>|Li cells with voltage window of 3–4.5 V at different currents (<b>a</b>) 0.05 mA cm<sup>−2</sup>, (<b>b</b>) 0.10 mA cm<sup>−2</sup>, (<b>c</b>) 0.24 mA cm<sup>−2</sup>, (<b>d</b>) 0.50 mA cm<sup>−2</sup>, and (<b>e</b>) 0.05 mA cm<sup>−2</sup> (final).</p>
Full article ">Figure 9
<p>Electrochemical performance of LiCoO<sub>2</sub>|LLCZN/PEO/LiPF<sub>6</sub>|Li cells with voltage window of 3–4.5 V during the first 20 cycles at different currents: (<b>a</b>) specific capacity cycling performance; (<b>b</b>) specific energy cycling performance; (<b>c</b>) charge–discharge curve at 0.05 mA cm<sup>−2</sup> for selected cycles; (<b>d</b>) cycling performance.</p>
Full article ">
7 pages, 2725 KiB  
Article
Preparation of Carbon Nanowall and Carbon Nanotube for Anode Material of Lithium-Ion Battery
by Seokwon Lee, Seokhun Kwon, Kangmin Kim, Hyunil Kang, Jang Myoun Ko and Wonseok Choi
Molecules 2021, 26(22), 6950; https://doi.org/10.3390/molecules26226950 - 17 Nov 2021
Cited by 12 | Viewed by 3188
Abstract
Carbon nanowall (CNW) and carbon nanotube (CNT) were prepared as anode materials of lithium-ion batteries. To fabricate a lithium-ion battery, copper (Cu) foil was cleaned using an ultrasonic cleaner in a solvent such as trichloroethylene (TCE) and used as a substrate. CNW and [...] Read more.
Carbon nanowall (CNW) and carbon nanotube (CNT) were prepared as anode materials of lithium-ion batteries. To fabricate a lithium-ion battery, copper (Cu) foil was cleaned using an ultrasonic cleaner in a solvent such as trichloroethylene (TCE) and used as a substrate. CNW and CNT were synthesized on Cu foil using plasma-enhanced chemical vapor deposition (PECVD) and water dispersion, respectively. CNW and CNT were used as anode materials for the lithium-ion battery, while lithium hexafluorophosphate (LiPF6) was used as an electrolyte to fabricate another lithium-ion battery. For the structural analysis of CNW and CNT, field emission scanning electron microscope (FE-SEM) and Raman spectroscopy analysis were performed. The Raman analysis showed that the carbon nanotube in composite material can compensate for the defects of the carbon nanowall. Cyclic voltammetry (CV) was employed for the electrochemical properties of lithium-ion batteries, fabricated by CNW and CNT, respectively. The specific capacity of CNW and CNT were calculated as 62.4 mAh/g and 49.54 mAh/g. The composite material with CNW and CNT having a specific capacity measured at 64.94 mAh/g, delivered the optimal performance. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The schematic diagram of the Cu foil cleaning process, (<b>b</b>) synthesizing the CNT using water dispersion method and (<b>c</b>) synthesizing the composite material with CNW and CNT.</p>
Full article ">Figure 2
<p>The schematic diagram of PECVD.</p>
Full article ">Figure 3
<p>The FE-SEM surface image of the (<b>a</b>) CNW (scale bar, 3 μm) with enlarged CNW (inset, scale bar, 500 nm) and (<b>b</b>) CNT (scale bar, 3 μm) with enlarged CNT (inset, scale bar, 500 nm). The FE-SEM cross-section image of the (<b>c</b>) CNW (scale bar, 3 μm), (<b>d</b>) CNT (scale bar, 3 μm) and (<b>e</b>) composite material with CNW and CNT (scale bar, 4 μm).</p>
Full article ">Figure 4
<p>(<b>a</b>) The Rama shift of the CNW (black line), CNT (red line), and composite material with CNW and CNT (blue line). (<b>b</b>) The I<sub>D</sub>/I<sub>G</sub> and I<sub>2D</sub>/I<sub>G</sub> ratio of CNW, CNT and composite material from Raman shift.</p>
Full article ">Figure 5
<p>The CV graphs of (<b>a</b>) CNW, (<b>b</b>) CNT and (<b>c</b>) composite material for 1, 5, 10 cycle. (<b>d</b>) The CV graph of the 12th cycle for each of the samples.</p>
Full article ">Figure 6
<p>The Specific capacity of the CNW, CNT, and composite material for 12 cycles.</p>
Full article ">
15 pages, 4626 KiB  
Article
LiCoO2/Graphite Cells with Localized High Concentration Carbonate Electrolytes for Higher Energy Density
by Xin Ma, Peng Zhang, Huajun Zhao, Qingrong Wang, Guangzhao Zhang, Shang-Sen Chi, Zhongbo Liu, Yunxian Qian, Jun Wang, Chaoyang Wang and Yonghong Deng
Liquids 2021, 1(1), 60-74; https://doi.org/10.3390/liquids1010005 - 10 Nov 2021
Cited by 7 | Viewed by 5689
Abstract
Widening the working voltage of lithium-ion batteries is considered as an effective strategy to improve their energy density. However, the decomposition of conventional aprotic electrolytes at high voltage greatly impedes the success until the presence of high concentration electrolytes (HCEs) and the resultant [...] Read more.
Widening the working voltage of lithium-ion batteries is considered as an effective strategy to improve their energy density. However, the decomposition of conventional aprotic electrolytes at high voltage greatly impedes the success until the presence of high concentration electrolytes (HCEs) and the resultant localized HCEs (LHCEs). The unique solvated structure of HCEs/LHCEs endows the involved solvent with enhanced endurance toward high voltage while the LHCEs can simultaneously possess the decent viscosity for sufficient wettability to porous electrodes and separator. Nowadays, most LHCEs use LiFSI/LiTFSI as the salts and β-hydrofluoroethers as the counter solvents due to their good compatibility, yet the LHCE formula of cheap LiPF6 and high antioxidant α-hydrofluoroethers is seldom investigated. Here, we report a unique formula with 3 mol L−1 LiPF6 in mixed carbonate solvents and a counter solvent α-substituted fluorine compound (1,1,2,2-tetrafluoroethyl-2,2,3,3-tetrafluoropropylether). Compared to a conventional electrolyte, this formula enables dramatic improvement in the cycling performance of LiCoO2//graphite cells from approximately 150 cycles to 1000 cycles within the range of 2.9 to 4.5 V at 0.5 C. This work provides a new choice and scope to design functional LHCEs for high voltage systems. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Viscosity of LiPF<sub>6</sub>/EC/DMC electrolytes (blank line) and LiPF<sub>6</sub>/EC/DMC/TTE electrolytes; (<b>b</b>) Conductivity of LiPF<sub>6</sub>/EC/DMC electrolytes (blank line) and LiPF<sub>6</sub>/EC/DMC/TTE electrolytes; (<b>c</b>) LSV curves of the blank electrolyte, 5 M HCE, 3 M LHCE and 1.5 M LHCE.</p>
Full article ">Figure 2
<p>(<b>a</b>) Initial charge and discharge curves of the LCO//graphite cells with the blank, 3 M LHCE and 1.5 M LHCE electrolytes; (<b>b</b>) Cycling performance of the LCO//graphite cells using the blank, 3 M LHCE and 1.5 M LHCE electrolytes at 0.5 C and 30 °C in the voltage range of 2.9–4.5 V; (<b>c</b>) DCIR results of the LCO//graphite cells with the blank, 3 M LHCE and 1.5 M LHCE electrolytes upon cycling; (<b>d</b>) Rate capabilities of the LCO//graphite cells with the blank, 3 M LHCE and 1.5 M LHCE electrolytes at 30 °C; (<b>e</b>) Chronoamperometry plots of the LCO//graphite cells with the blank, 3 M LHCE and 1.5 M LHCE electrolytes at a charged state of 4.5 V.</p>
Full article ">Figure 3
<p>(<b>a</b>) Initial charge and discharge curves of the LCO//graphite cells with the blank, 3 M LHCE and 1.5 M LHCE electrolytes at 0.5 C in the voltage range of 2.9–4.5 V; (<b>b</b>) The 100th charge and discharge curves of the LCO//graphite cells with different electrolytes; (<b>c</b>) <span class="html-italic">Nyquist</span> diagrams of the LCO//graphite cells with different electrolytes after formation cycles; (<b>d</b>) <span class="html-italic">Nyquist</span> diagrams of these cells after 150 cycles (blank) and 1000 cycles (LHCEs).</p>
Full article ">Figure 4
<p>(<b>a</b>) NMR spectra of electrolytes with different amounts of LiPF<sub>6</sub> in EC/DMC; (<b>b</b>) NMR spectra of pure TTE, and different LHCEs diluted from 5 M HCE of LiPF<sub>6</sub>/EC/DMC.</p>
Full article ">Figure 5
<p>FTIR spectra of blank electrolyte, 5 M HCE, 3 M LHCE and 1.5 M LHCE.</p>
Full article ">Figure 6
<p>DSC curves of blank electrolyte, 5 M HCE, 3 M LHCE and 1.5 M LHCE.</p>
Full article ">Figure 7
<p>(<b>a</b>) Illustration of the solvation structure for the blank electrolyte; (<b>b</b>) Illustration of the solvation structure for the 5 M HCE; (<b>c</b>) Illustration of the solvation structure for the 3 M LHCE.</p>
Full article ">Figure 8
<p>SEM images of the LCO cathodes after cycling with (<b>a</b>) the blank electrolyte, (<b>b</b>) 3 M LHCE and (<b>c</b>) 1.5 M LHCE; SEM images of the graphite anodes after cycling with (<b>d</b>) the blank electrolyte and (<b>e</b>) 3 M LHCE and (<b>f</b>) 1.5 M LHCE.</p>
Full article ">Figure 9
<p>XPS spectra of C 1s, F 1s and O 1s of the LCO cathodes after cycling with different electrolytes.</p>
Full article ">Figure 10
<p>XPS spectra of C 1s, F 1s and O 1s of the graphite anodes after cycling with different electrolytes.</p>
Full article ">
10 pages, 2111 KiB  
Article
Investigation of a Novel Ecofriendly Electrolyte-Solvent for Lithium-Ion Batteries with Increased Thermal Stability
by Marco Ströbel, Larissa Kiefer and Kai Peter Birke
Batteries 2021, 7(4), 72; https://doi.org/10.3390/batteries7040072 - 28 Oct 2021
Cited by 2 | Viewed by 3678
Abstract
This study presents tributyl acetylcitrate (TBAC) as a novel ecofriendly high flash point and high boiling point solvent for electrolytes in lithium-ion batteries. The flash point (TFP=217C) and the boiling point (TBP=331 [...] Read more.
This study presents tributyl acetylcitrate (TBAC) as a novel ecofriendly high flash point and high boiling point solvent for electrolytes in lithium-ion batteries. The flash point (TFP=217C) and the boiling point (TBP=331C) of TBAC are approximately 200 K greater than that of conventional linear carbonate components, such as ethyl methyl carbonate (EMC) or diethyl carbonate (DEC). The melting point (TMP=80C) is more than 100 K lower than that of ethylene carbonate (EC). Furthermore, TBAC is known as an ecofriendly solvent from other industrial sectors. A life cycle test of a graphite/NCM cell with 1 M lithium hexafluorophosphate (LiPF6) in TBAC:EC:EMC:DEC (60:15:5:20 wt) achieved a coulombic efficiency of above 99% and the remaining capacity resulted in 90 percent after 100 cycles (C/4) of testing. As a result, TBAC is considered a viable option for improving the thermal stability of lithium-ion batteries. Full article
(This article belongs to the Special Issue Battery Systems and Energy Storage beyond 2020)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structure of tributyl acetylcitrate [<a href="#B24-batteries-07-00072" class="html-bibr">24</a>].</p>
Full article ">Figure 2
<p>Cycling performance of a graphite/NCM cell with 1 M LiTFSI in TBAC:EC:EMC:DEC (60:15:5:20 wt) electrolyte at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>25</mn> <msup> <mspace width="3.33333pt"/> <mo>∘</mo> </msup> </mrow> </semantics></math>C. The cell was charged and discharged at C—rate <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>/</mo> <mn>4</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>C-rate performance of a graphite/NCM cell with 1 M LiTFSI in TBAC:EC:EMC:DEC (60:15:5:20 wt) electrolyte at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>25</mn> <msup> <mspace width="3.33333pt"/> <mo>∘</mo> </msup> </mrow> </semantics></math>C in the potential range of <math display="inline"><semantics> <mrow> <mn>2.5</mn> <mo>−</mo> <mn>4.2</mn> </mrow> </semantics></math> V.</p>
Full article ">Figure 4
<p>Voltage profiles at different temperatures of a graphite/NCM cell with 1 M LiTFSI in TBAC:EC:EMC:DEC (60:15:5:20 wt) electrolyte.</p>
Full article ">Figure 5
<p>Cycling performance of a graphite/NCM cell with 1 M LiFSI in TBAC:EC:EMC (80:15:5 wt) electrolyte at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>25</mn> <msup> <mspace width="3.33333pt"/> <mo>∘</mo> </msup> </mrow> </semantics></math>C. The cell was charged and discharged at C—rate <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>/</mo> <mn>4</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Cycling performance of a graphite/NCM cell with 1 M LiPF<math display="inline"><semantics> <msub> <mrow/> <mn>6</mn> </msub> </semantics></math> in TBAC:EC:EMC:DEC (60:15:5:20 wt) electrolyte at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>25</mn> <msup> <mspace width="3.33333pt"/> <mo>∘</mo> </msup> </mrow> </semantics></math>C. The cell was charged and discharged at C—rate <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>/</mo> <mn>4</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Electrochemical stability window of (<b>a</b>) LP40, (<b>b</b>) 1 M LiPF<math display="inline"><semantics> <msub> <mrow/> <mn>6</mn> </msub> </semantics></math> dissolved in TBAC:EC:EMC:DEC (60:15:5:20 wt), (<b>c</b>) 1 M LiTFSI in TBAC:EC:EMC:DEC (60:15:5:20 wt) (all with stainless steel as working and counter electrodes), and (<b>d</b>) 1 M LiTFSI in TBAC:EC:EMC:DEC (60:15:5:20 wt) (with aluminum as working electrode). Scan rate 1 mVs<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>. The electrochemical stability of LiPF<math display="inline"><semantics> <msub> <mrow/> <mn>6</mn> </msub> </semantics></math> in TBAC seems to be improved compared to LP40. LiTFSI—TBAC based electrolytes in contact with aluminum show higher decomposition rates, which is assumed to be linked to aluminum corrosion by LiTFSI.</p>
Full article ">
20 pages, 4375 KiB  
Article
An In-Depth Life Cycle Assessment (LCA) of Lithium-Ion Battery for Climate Impact Mitigation Strategies
by Jhuma Sadhukhan and Mark Christensen
Energies 2021, 14(17), 5555; https://doi.org/10.3390/en14175555 - 6 Sep 2021
Cited by 26 | Viewed by 15425
Abstract
Battery energy storage systems (BESS) are an essential component of renewable electricity infrastructure to resolve the intermittency in the availability of renewable resources. To keep the global temperature rise below 1.5 °C, renewable electricity and electrification of the majority of the sectors are [...] Read more.
Battery energy storage systems (BESS) are an essential component of renewable electricity infrastructure to resolve the intermittency in the availability of renewable resources. To keep the global temperature rise below 1.5 °C, renewable electricity and electrification of the majority of the sectors are a key proposition of the national and international policies and strategies. Thus, the role of BESS in achieving the climate impact mitigation target is significant. There is an unmet need for a detailed life cycle assessment (LCA) of BESS with lithium-ion batteries being the most promising one. This study conducts a rigorous and comprehensive LCA of lithium-ion batteries to demonstrate the life cycle environmental impact hotspots and ways to improve the hotspots for the sustainable development of BESS and thus, renewable electricity infrastructure. The whole system LCA of lithium-ion batteries shows a global warming potential (GWP) of 1.7, 6.7 and 8.1 kg CO2 eq kg−1 in change-oriented (consequential) and present with and without recycling credit consideration, scenarios. The GWP hotspot is the lithium-ion cathode, which is due to lithium hexafluorophosphate that is ultimately due to the resource-intensive production system of phosphorous, white, liquid. To compete against the fossil economy, the GWP of BESS must be curbed by 13 folds. To be comparable with renewable energy systems, hydroelectric, wind, biomass, geothermal and solar (4–76 g CO2 eq kWh−1), 300 folds reduction in the GWP of BESS will be necessary. The areas of improvement to lower the GWP of BESS are as follows: reducing scopes 2–3 emissions from fossil resource use in the material production processes by phosphorous recycling, increasing energy density, increasing lifespan by effective services, increasing recyclability and number of lives, waste resource acquisition for the battery components and deploying multi-faceted integrated roles of BESS. Achieving the above can be translated into an overall avoided GWP of up to 82% by 2040. Full article
(This article belongs to the Special Issue Energy Storage for Grid Integration of Renewable Energy)
Show Figures

Figure 1

Figure 1
<p>Service provisions of energy storage systems.</p>
Full article ">Figure 2
<p>Activities and materials involved in the life cycle of BESS.</p>
Full article ">Figure 3
<p>Percentage life cycle energy storage and saving by BESS in various services in a whole distributed energy system.</p>
Full article ">Figure 4
<p>Results of GWP hotspot analysis of BESS.</p>
Full article ">Figure 5
<p>ReCiPe (M) (H) life cycle impact characterisations of 1 kg of each key material shown in <a href="#energies-14-05555-t001" class="html-table">Table 1</a>. 1,4 DCB: 1,4-dichlorobenzene. CFC11: chlorofluorocarbon. GLO: global. RoW, rest of the world and CH: Switzerland in Ecoinvent.</p>
Full article ">Figure 6
<p>CML life cycle impact characterisations of 1 kg of each key material shown in <a href="#energies-14-05555-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 7
<p>ILCD life cycle impact characterisations of 1 kg of each key material shown in <a href="#energies-14-05555-t001" class="html-table">Table 1</a>. CTU: comparative toxic units. NMVOC: non-methane volatile organic compound.</p>
Full article ">Figure 8
<p>GWP of material productions in the entire system to show the contributing materials.</p>
Full article ">Figure 9
<p>The UK electricity generation by alternative resource (both TWh (<b>top</b>) and percentage (<b>bottom</b>) per annum) from 2016 through to 2020 [<a href="#B2-energies-14-05555" class="html-bibr">2</a>].</p>
Full article ">Figure 9 Cont.
<p>The UK electricity generation by alternative resource (both TWh (<b>top</b>) and percentage (<b>bottom</b>) per annum) from 2016 through to 2020 [<a href="#B2-energies-14-05555" class="html-bibr">2</a>].</p>
Full article ">
20 pages, 3219 KiB  
Article
A Genome-Wide Screen in Saccharomyces cerevisiae Reveals a Critical Role for Oxidative Phosphorylation in Cellular Tolerance to Lithium Hexafluorophosphate
by Xuejiao Jin, Jie Zhang, Tingting An, Huihui Zhao, Wenhao Fu, Danqi Li, Shenkui Liu, Xiuling Cao and Beidong Liu
Cells 2021, 10(4), 888; https://doi.org/10.3390/cells10040888 - 13 Apr 2021
Cited by 10 | Viewed by 2881
Abstract
Lithium hexafluorophosphate (LiPF6) is one of the leading electrolytes in lithium-ion batteries, and its usage has increased tremendously in the past few years. Little is known, however, about its potential environmental and biological impacts. In order to improve our understanding of [...] Read more.
Lithium hexafluorophosphate (LiPF6) is one of the leading electrolytes in lithium-ion batteries, and its usage has increased tremendously in the past few years. Little is known, however, about its potential environmental and biological impacts. In order to improve our understanding of the cytotoxicity of LiPF6 and the specific cellular response mechanisms to it, we performed a genome-wide screen using a yeast (Saccharomyces cerevisiae) deletion mutant collection and identified 75 gene deletion mutants that showed LiPF6 sensitivity. Among these, genes associated with mitochondria showed the most enrichment. We also found that LiPF6 is more toxic to yeast than lithium chloride (LiCl) or sodium hexafluorophosphate (NaPF6). Physiological analysis showed that a high concentration of LiPF6 caused mitochondrial damage, reactive oxygen species (ROS) accumulation, and ATP content changes. Compared with the results of previous genome-wide screening for LiCl-sensitive mutants, we found that oxidative phosphorylation-related mutants were specifically hypersensitive to LiPF6. In these deletion mutants, LiPF6 treatment resulted in higher ROS production and reduced ATP levels, suggesting that oxidative phosphorylation-related genes were important for counteracting LiPF6-induced toxicity. Taken together, our results identified genes specifically involved in LiPF6-modulated toxicity, and demonstrated that oxidative stress and ATP imbalance maybe the driving factors in governing LiPF6-induced toxicity. Full article
(This article belongs to the Section Plant, Algae and Fungi Cell Biology)
Show Figures

Figure 1

Figure 1
<p>Identification of LiPF<sub>6</sub>-sensitive mutants by genome-wide screening. (<b>A</b>) Sensitivity of <span class="html-italic">S. cerevisiae</span> to LiPF<sub>6</sub>, LiCl, NaPF<sub>6</sub>, and both LiCl and NaPF<sub>6</sub>. Cell growth of BY4741 treated with different concentrations of LiPF<sub>6</sub>, LiCl, and NaPF<sub>6</sub>, respectively, was measured by reading absorbance at 600 nm (OD<sub>600</sub>) at the indicated time points. Growth curves were performed in triplicate. Growth was represented by mean OD<sub>600</sub> values and error bars indicate SE. (<b>B</b>) GO term analysis for the 75 genes in <a href="#cells-10-00888-t001" class="html-table">Table 1</a>, the deletion of which resulted in LiPF<sub>6</sub> sensitivity. (<b>C</b>) KEGG analysis for the 75 genes in <a href="#cells-10-00888-t001" class="html-table">Table 1</a>. The <span class="html-italic">p</span>-value was corrected using the Benjamini and Hochberg (1995) correction method [<a href="#B45-cells-10-00888" class="html-bibr">45</a>]. (<b>D</b>) Genetic interactions, physical interactions, and co-localization of the 75 LiPF<sub>6</sub>-sensitive genes. Grey, red, and blue edges indicate genetic interactions, physical interactions, and co-localization, respectively. The node colors indicate different functions. SE, Standard Error; GO, Gene Ontology; KEGG, Kyoto Encyclopedia of Genes and Genomes.</p>
Full article ">Figure 2
<p>Oxidative phosphorylation-related gene deletion strains show increased sensitivity to LiPF<sub>6</sub>. (<b>A</b>) Phenotypes of the SGA control strain (<span class="html-italic">MATa his3Δ::kanMX4</span>) and deletion mutants in our screen. Each strain was arranged in quadruplicate. (<b>B</b>) Spot test to verify the screening results. The control strain and <span class="html-italic">cox5a</span><span class="html-italic">∆</span>, <span class="html-italic">cox12</span><span class="html-italic">∆</span>, <span class="html-italic">cox14</span><span class="html-italic">∆</span>, <span class="html-italic">qcr2</span><span class="html-italic">∆</span>, and <span class="html-italic">qcr6</span><span class="html-italic">∆</span> strains were grown to mid-log phase in YPD medium and then diluted to an OD<sub>600</sub> of 0.5. Cells were serially diluted onto YPD agar plates either containing 3 mM LiPF<sub>6</sub> or no LiPF<sub>6</sub>. Plates were photographed after 48 h of incubation at 30 °C. Images shown are representative of triplicates. (<b>C</b>) Spot test of the complementation strains. Deletion mutant strains transformed with empty vector or plasmid expressing the corresponding genes were grown to mid-log phase in YPD medium before diluting to an OD<sub>600</sub> of 0.5 and cells were serially diluted onto YPD plates with 3 mM LiPF<sub>6</sub> or without LiPF<sub>6</sub>. The control strain transformed with empty vector served as a control. Images were representative of triplicates. SGA, Synthetic Genetic Array; YPD, Yeast Peptone Dextrose.</p>
Full article ">Figure 3
<p>Effects of LiPF<sub>6</sub> on mitochondrial morphology, ROS, and ATP synthesis. (<b>A</b>) Mitochondrial morphology of BY4741 was observed with or without LiPF<sub>6</sub>. A total of 10 μM CCCP-treated cells served as the positive control. Mitochondria stained with MitoTracker Red CMXRos are shown on the left; 3D volume images of mitochondria are shown in the second column; bright-field micrographs are shown in the third column; and merged images are shown on the right. “−”: without LiPF<sub>6</sub> and CCCP; “+ LiPF<sub>6</sub>”: with 4 mM LiPF<sub>6</sub>; “+ CCCP”: with 10 μM CCCP. Scale bar represents 5 μm. The percentage of cells exhibiting tubular or fragmented mitochondria was calculated. At least 200 cells of each sample were used for quantitation. Error values indicate the SE from three independent experiments. ***, <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) Intracellular ROS were detected in BY4741 in the absence or presence of LiPF<sub>6</sub>. ROS signal stained with H<sub>2</sub>DCFDA was shown on the left; bright-field micrographs are shown in the middle; and merged images are shown on the right. “−”: without LiPF<sub>6</sub>; “+”: with 4 mM LiPF<sub>6</sub>. Scale bar represents 5 μm. (<b>C</b>) Mitochondrial ATP synthesis was measured after incubation with 0 or 4 mM LiPF<sub>6</sub>. The vertical axis represents the ATP content per mg protein. Error bars indicate the SE from three independent experiments. **, <span class="html-italic">p</span> &lt; 0.01, Student’s <span class="html-italic">t</span>-test. CCCP, Carbonyl cyanide m-chlorophenylhydrazone; ROS, Reactive Oxygen Species; SE, Standard Error.</p>
Full article ">Figure 4
<p>Oxidative phosphorylation-related genes are required for the regulation of LiPF<sub>6</sub>-induced ROS. (<b>A</b>) The presence of ROS in BY4741 and the deletion mutant strains in medium with or without LiPF<sub>6</sub> was determined. The H<sub>2</sub>O<sub>2</sub>-treated SGA control strain (<span class="html-italic">his3Δ</span>) served as a positive control. Left, ROS signal; middle, bright-field micrographs; right, merged images. Scale bar represents 5 μm. (<b>B</b>) Quantifications of H<sub>2</sub>DCFDA-positive cells. Three independent biological experiments were carried out, and for each replicate, a minimum of 200 cells were counted. The vertical axis represents the percentage of cells with a ROS signal, and the horizontal axis represents the different strains. Error bars indicate SE. ***, <span class="html-italic">p</span> &lt; 0.001, **, <span class="html-italic">p</span> &lt; 0.01, *, <span class="html-italic">p</span> &lt; 0.05, Student’s <span class="html-italic">t</span>-test. (<b>C</b>) Quantifications of fluorescence intensity per cell. Three independent biological experiments were carried out and for each replicate, and a minimum of 200 cells were counted. The vertical axis represents the fluorescence intensity per cell, and the horizontal axis represents the different strains. Error bars indicate SE. ***, <span class="html-italic">p</span> &lt; 0.001, *, <span class="html-italic">p</span> &lt; 0.05, Student’s <span class="html-italic">t</span>-test. (<b>D</b>) Western bolt analysis of Cox5a before and after LiPF<sub>6</sub> treatment. Pgk1 served as a loading control. Accumulation levels of Cox5a protein were quantified using ImageJ software. Error bars indicate SE. **, <span class="html-italic">p</span> &lt; 0.01, Student’s <span class="html-italic">t</span>-test. ROS, Reactive Oxygen Species; SE, Standard Error.</p>
Full article ">Figure 5
<p>Deletion of the oxidative phosphorylation-related genes aggravated the decrease in ATP production under LiPF<sub>6</sub> treatment. ATP synthesis abilities were compared between the control strain (<span class="html-italic">his3Δ</span>) and the deletion mutants. The vertical axis represents the ATP content per mg protein. The data shown represent averages of three experiments, and error bars indicate SE. ***, <span class="html-italic">p</span> &lt; 0.001, **, <span class="html-italic">p</span> &lt; 0.01, *, <span class="html-italic">p</span> &lt; 0.05, Student’s <span class="html-italic">t</span>-test. YPD, Yeast Peptone Dextrose; SE, Standard Error.</p>
Full article ">Figure 6
<p>Oxidative phosphorylation-related genes are not hypersensitive to LiCl or NaPF<sub>6</sub>. The SGA control strain and deletion strains were grown to mid-log phase in YPD medium before diluting to an OD<sub>600</sub> of 0.5. Cultures were serially diluted onto YPD plates containing different concentrations of LiCl or NaPF<sub>6</sub>. Plates were incubated at 30 °C and photographed. SGA, Synthetic Genetic Array; YPD, Yeast Peptone Dextrose.</p>
Full article ">
13135 KiB  
Article
Capillary Electrophoresis as Analysis Technique for Battery Electrolytes: (i) Monitoring Stability of Anions in Ionic Liquids and (ii) Determination of Organophosphate-Based Decomposition Products in LiPF6-Based Lithium Ion Battery Electrolytes
by Marcelina Pyschik, Martin Winter and Sascha Nowak
Separations 2017, 4(3), 26; https://doi.org/10.3390/separations4030026 - 5 Sep 2017
Cited by 5 | Viewed by 5500
Abstract
In this work, a method for capillary electrophoresis (CE) hyphenated to a high-resolution mass spectrometer was presented for monitoring the stability of anions in ionic liquids (ILs) and in commonly used lithium ion battery (LIB) electrolytes. The investigated ILs were 1-methyl-1-propylpyrrolidinium bis(trifluoromethanesulfonyl)imide (PYR13TFSI) [...] Read more.
In this work, a method for capillary electrophoresis (CE) hyphenated to a high-resolution mass spectrometer was presented for monitoring the stability of anions in ionic liquids (ILs) and in commonly used lithium ion battery (LIB) electrolytes. The investigated ILs were 1-methyl-1-propylpyrrolidinium bis(trifluoromethanesulfonyl)imide (PYR13TFSI) and 1-methyl-1-propylpyrrolidinium bis(fluorosulfonyl)imide (PYR13FSI). The method development was conducted by adjusting the following parameters: buffer compositions, buffer concentrations, and the pH value. Also the temperature and the voltage applied on the capillary were optimized. The ILs were aged at room temperature and at 60 °C for 16 months each. At both temperatures, no anionic decomposition products of the FSI− and TFSI− anions were detected. Accordingly, the FSI− and TFSI− anions were thermally stable at these conditions. This method was also applied for the investigation of LIB electrolyte samples, which were aged at 60 °C for one month. The LP30 (50/50 wt. % dimethyl carbonate/ethylene carbonate and 1 M lithium hexafluorophosphate) electrolyte was mixed with the additive 1,3-propane sultone (PS) and with one of the following organophosphates (OP): dimethyl phosphate (DMP), diethyl phosphate (DEP), and triethyl phosphate (TEP), to investigate the influence of these compounds on the formation of OPs. Full article
(This article belongs to the Special Issue Ionic Liquid for Separations)
Show Figures

Figure 1

Figure 1
<p>Electropherogram of bis(trifluoromethanesulfonyl)imide (TFSI<sup>−</sup>) with the buffers A (40.0 mM ammonium formate buffer) and B (50.0 mM ammonium acetate buffer).</p>
Full article ">Figure 2
<p>Electropherograms showing the TFSI<sup>−</sup> peak using an acetate buffer during method development: (<b>A</b>) measured at different pH values; (<b>B</b>) measured with different concentrations of the acetate buffer; (<b>C</b>) measured at different temperatures and (<b>D</b>) measured under different applied voltages.</p>
Full article ">Figure 3
<p>Electropherograms of the detected <span class="html-italic">m/z</span> ratio 279.9173 of the fresh and aged TFSI<sup>−</sup> sample (<b>A</b>) and of the detected <span class="html-italic">m/z</span> ratio 179.9238 of the fresh and aged FSI<sup>−</sup> sample (<b>B</b>).</p>
Full article ">Figure 4
<p>Electropherogram of the aged reference electrolyte sample.</p>
Full article ">Figure 5
<p>Electropherogram of the organophosphates (OPs) in the aged electrolyte sample containing dimethyl phosphate (DMP) and 1,3-propane sultone (PS).</p>
Full article ">Figure 6
<p>Electropherogram of the OPs in the aged electrolyte sample containing DEP (9) and PS.</p>
Full article ">Figure 7
<p>Electropherogram of the OPs in the aged electrolyte sample containing TEP and PS.</p>
Full article ">
Back to TopTop