Open Access Article
Chulgi Nathan
Hong
a,
Mengwen
Yan
b,
Oleg
Borodin
c,
Travis P.
Pollard
c,
Langyuan
Wu
d,
Manuel
Reiter
a,
Dario Gomez
Vazquez
a,
Katharina
Trapp
a,
Ji Mun
Yoo
a,
Netanel
Shpigel
d,
Jeremy I.
Feldblyum
b and
Maria R.
Lukatskaya
*a
aElectrochemical Energy Systems Laboratory, Department of Mechanical and Process Engineering, ETH Zurich, Zürich 8092, Switzerland. E-mail: mlukatskaya@ethz.ch
bDepartment of Chemistry, The University at Albany, State University of New York, Albany, NY 12222, USA
cBattery Science Branch, DEVCOM Army Research Laboratory, Adelphi, MD 20783, USA
dDepartment of Chemical Sciences, Ariel University, Ariel 40700, Israel
First published on 2nd May 2024
Controlling solid electrolyte interphase (SEI) in batteries is crucial for their efficient cycling. Herein, we demonstrate an approach to enable robust battery performance that does not rely on high fractions of fluorinated species in electrolytes, thus substantially decreasing the environmental footprint and cost of high-energy batteries. In this approach, we use very low fractions of readily reducible fluorinated cations in electrolyte (∼0.1 wt%) and employ electrostatic attraction to generate a substantial population of these cations at the anode surface. As a result, we can form a robust fluorine-rich SEI that allows for dendrite-free deposition of dense Li and stable cycling of Li-metal full cells with high-voltage cathodes. Our approach represents a general strategy for delivering desired chemical species to battery anodes through electrostatic attraction while using minute amounts of additive.
Broader contextStable and cost-effective Li-metal batteries (LiMBs) are necessary for non-incremental improvement of energy density in commercial batteries. However, the implementation of Li-metal anodes is impeded by an unacceptably low cycle life and safety concerns when conventional electrolytes are used. In particular, the formation of electronically inactive “dead” lithium and dendrites takes place during cycling. Prior research suggests that fluorine-rich interfacial layer chemistry is important for the stabilization of Li-metal anodes, which can be achieved when electrolytes with a high fraction of fluorinated solvents and/or salts are used. Herein, we introduce an alternative approach that leverages electrostatic attraction between positively charged fluorinated cations and the negatively charged Li-metal anode, generating a significant population of fluorinated species near the electrode surface with a very low fraction of additive in the electrolyte (∼0.1 wt%). As a result, a robust fluorine-rich interfacial layer is formed, enabling dendrite-free deposition of dense Li metal. In general, we present a strategy for delivering desired chemical species to the battery anodes through electrostatic attraction while using minute amounts of additive and therefore can notably reduce costs and environmental footprint of implementing high energy batteries. |
Recent research suggests that fluorine-rich SEI yields superior performance compared to fluorine-free SEI.3,8,9 To achieve fluorine-rich SEI, state-of-the-art approaches rely on large volume fractions of fluorinated species in the electrolyte that have a statistically higher probability of being reduced at the electrode surface and, therefore, generate fluorine-rich SEI.10 For example, solvent-in-salt electrolytes with a high concentration of fluorine-containing anions yield fluorine-rich SEIs and much improved coulombic efficiencies compared to their dilute counterparts.11 Recently, another way to generate a fluorine enriched SEI from fluorine-containing salts at intermediate concentrations was reported by employing a modified solvent with siloxane groups12 or through usage of surface-modified separators.13 Also, fluorinated solvents themselves can be used to form high-quality SEIs.8,14 The implementation of a fluorinated ionic liquid as a battery solvent is another way to create a fluorine-rich SEI that can suppress formation of dead lithium.15,16
Yet, despite enabling promising performance, these approaches have notable drawbacks when it comes to their practical implementation. The high cost of Li salts and increased solution viscosity at high salt concentration makes highly concentrated electrolytes challenging to implement in commercial batteries. Similarly, the replacement of conventional solvents with heavily fluorinated ones can lead to substantial increases in battery costs and environmental footprint.
Herein, we establish an alternative approach that relies on the electrostatic attraction of fluorinated cations to a negatively charged anode. Through this approach, a significant population of fluorinated species can reach the electrode surface even when the overall additive concentration in the bulk electrolyte is in the millimolar range. Importantly, to ensure the predominant contribution of fluorinated cations to SEI formation, their reduction potential should be significantly higher than that of solvent molecules or anions in electrolyte (as schematically depicted in Fig. 1a). To this end, we selected fluorinated methylpyridinium cations (Fig. 1a) that can offer early decomposition potentials of ∼2 V vs. Li/Li+ (at least ∼1.5 V before solvent decomposition starts). First, we show that the addition of fluorinated methylpyridinium cations even in millimolar amounts (0.08–0.14 wt%) to a conventional electrolyte based on 1,2-dimethoxyethane (DME) enables F-enriched SEI and dense Li plating with an increased coulombic efficiency of 99.6% (compared to 96.4% obtained in additive-free electrolytes). Second, we demonstrate prolonged cycling of a full cell with Li metal anode and Ni-rich high-voltage cathode in ether-based electrolytes with 99.6% average coulombic efficiency, which is achieved by suppressing oxidative decomposition of DME upon addition of small amounts of fluorinated cations. Third, the fluorinated cations also suppress corrosion of the Al current collector caused by chlorine impurities in the electrolyte, thus assisting with prolonged cycling of the full cell.
Cyclic voltammetry (CV) profiles collected in additive-containing electrolytes show a pronounced reduction peak at ∼1.98 V vs. Li/Li+ (Fig. 1b) during the 1st cycle, which is absent in the additive-free electrolyte. The current associated with the reduction peak scales with the concentration of TFP (Fig. S2, ESI†). Furthermore, an identical position of TFP-related reduction peak is observed regardless of whether TFP additive is used in the form of perchlorate or triflate salt (Fig. S3, ESI†). The reduction peak disappears in the 2nd cycle, consistent with passivation of the anode surface, preventing further reduction of fluorinated cations. These experiments confirm that the reduction of fluorinated cations occurs at potentials nearly ∼1.5 V higher than the onset of decomposition of the additive-free electrolyte.20 In contrast, no distinct reduction CV peaks or subsequent passivation was observed upon addition of the same concentrations of the neutral analogue, 2,4,6-trifluoropyridine (TFN), indicating importance of the additive charge in enabling its efficiency (Fig. 1b). These results are also in agreement with density functional theory (DFT) calculations that predict TFN's low reduction potential of 0.23 V vs. Li/Li+ (Fig. S1a, ESI†).
To quantify how the addition of fluorinated cations affects the coulombic efficiency of Li plating/stripping, we used a modified Aurbach protocol21 (Fig. S4a, ESI†). The test revealed a substantial improvement of average coulombic efficiency from 96.4% for the TFP-free electrolyte to 99.6% for the cells containing TFP (∼0.1 wt%, Fig. S4b, ESI†). This efficiency is comparable to those reported for the best-performing electrolytes containing fluorinated solvent8,22 or salts in high concentrations,11,23 demonstrating that large fractions of fluorinated species in electrolytes can be avoided when aiming for high coulombic efficiencies.
Next, we performed long-term galvanostatic cycling at 10 mA cm−2 using a Li0–Li0 symmetric cell configuration (Fig. 1c). A dramatic difference was observed between the cells with and without TFP. For cells with additive-free electrolyte, a continuous increase in overpotential due to the formation of dead lithium was observed with cycling (similar to previous reports24,25), resulting in cell failure after 400 h (Fig. 1d). In contrast, addition of millimolar amounts of fluorinated cations enabled outstanding cycling stability and, after initial stabilization, the overpotential remained almost unchanged for at least 3000 h of cycling (Fig. 1e). Moreover, the cycling stability increases with the concentration of fluorinated cations (Fig. S5, ESI†), showing the greatest stability for the 18 mM concentration (Fig. 1c).
Importantly, the morphology of Li metal correlated with the evolution of overpotential with cycling. Initially, comparable overpotentials and similar Li morphology were observed for both electrolytes (Fig. S6, ESI†). After 371 hours of cell cycling, for the additive-free electrolyte a rough Li surface with multiple cracks can be seen by scanning electron microscopy (SEM; Fig. 1f and Fig. S7a, ESI†), with highly porous Li deposits and in agreement with prior studies.26 In contrast, a smooth surface with large and dense Li grains was observed for Li metal that was cycled in the electrolyte containing TFP (Fig. 1g and Fig. S7b, ESI†).
:
carbon (F
:
C) atomic ratio for the SEI formed in the presence of cationic additives (F
:
C = 3.1–4.5), whereas the F
:
C ratio of 0.17–0.24 was observed for the SEI generated in reference electrolyte (Fig. 2a). This indicates a dramatically decreased relative contribution of the solvent decomposition products. Furthermore, analysis of F 1s and Li 1s spectra revealed LiF as a dominant species for SEI samples obtained in the presence of fluorinated cations (Fig. 2b and Fig. S8 and Table S1, ESI†). Similarly, a more than three-fold enrichment of SEI with fluorine was observed for a Li metal electrode cycled in the presence of TFP (Li0‖Li0 symmetric cell, one week of galvanostatic cycling at 10 mA cm−2, Fig. S9, ESI†). This indicates that even millimolar addition of the fluorinated cations can yield favorable F-rich SEI.
![]() | ||
Fig. 2 (a) F : C atomic ratio in the SEI layer formed on a Cu electrode as a function of depth. (b) XPS F 1s spectrum of a SEI layer formed on a Cu electrode after cycling in DME + 1 M LiFSI + 18 mM TFP with Ar+ sputtered time of 72 s (estimated depth: 2.0 nm). (c) Voltage, charge, frequency, and dissipation change versus time using EQCM-D analysis for DME + 1 M LiFSI and DME + 1 M LiFSI + 18 mM TFP. (d) Products of reduction at the negative electrode obtained from DFT calculations, see Fig. S11 (ESI†) for further details. | ||
Next, EQCM-D measurements show that the SEI formed in electrolytes with cationic additive is highly rigid since the change of the overtone-normalized resonant frequency (Δfn/n) and energy dissipation (ΔDn) are independent27 of overtone order (n) (Fig. 2c). Therefore, we can apply the Sauerbrey equation to obtain gravimetric information for the formed SEI. By comparing the theoretical and experimental frequencies,28 we attribute the main frequency decrease at ∼2 V to the formation of LiF, which is in good agreement with the XPS results (Fig. S10, ESI† and Fig. 2b). Moreover, no further mass gain was observed after the first CV cycle, indicating the highly desirable robust and passivating nature of the formed SEI. In contrast, the SEI formed in the additive-free electrolyte displays viscoelastic behavior: ΔDn depends on overtone order and larger values of dissipation modulus are observed.27–29 Upon cycling, a continuous increase in charge, Δfn/n and ΔDn can be seen, indicative of uncontrolled SEI growth due to the non-compact and porous nature of the formed interfacial layers.27
Since both XPS and EQCM-D suggest that SEI formed in the presence of TFP is mainly LiF, we further investigated the detailed SEI formation pathway using DFT calculations and MD simulations. The DFT calculations suggest that TFP reduction is the dominant process due to the electrostatic attraction of TFP to the negative electrode and its higher reduction potential (2.1–2.5 V vs. Li/Li+) compared to FSI− (Fig. 2d and Fig. S11, ESI†). Meanwhile, FSI− reduction is unlikely here due to the low affinity of FSI− to the negative electrode30 and low fraction of contact ion pairs (CIPs) or aggregates (AGGs) of LiFSI in DME at this salt concentration (<30%) observed in MD simulations (see Supplementary discussion, ESI†) and previous work.31,32 Reduced TFP* radicals at the anode surface undergo a second reduction and release of F− that results in the formation of LiF when TFP* is in close proximity to Li+ (Fig. 2d and Fig. S11, ESI†). A subsequent reduction and defluorination is also likely when the Li+/TFP−(−F) complex is exposed to potentials below 1.9 V vs. Li/Li+. This allows the additive to deliver multiple F− to form LiF-rich SEI. Alternatively, the reduced TFP* radicals can dimerize after their 1st reduction to form a fluorinated viologen species which undergoes spontaneous defluorination in proximity of Li+ across the newly formed C–C bond (left panel of Fig. S12, ESI†). The viologen species can form an anion (2nd reduction) at ∼0.6 V with subsequent release of another LiF (3rd reduction; in the box in Fig. S12, ESI†).
Such an increase in the oxidation stability of the DME-based electrolyte with TFP cations implies that these electrolytes can be used with high voltage cathodes. To study this, we first performed galvanostatic cycling using a LiNi0.8Co0.1Mn0.1O2 (NCM811) cathode and the Li metal anode (NCM811‖Li metal) in a coin cell configuration with different concentrations of TFP additive to find the optimal one (Fig. S14a, ESI†). Without TFP, the cells show severe capacity fading and a decrease in coulombic efficiency around the 30th cycle, in agreement with previous reports (Fig. S14 and S15, ESI†).8 In contrast, cells with TFP demonstrate a dramatic improvement in cycling stability even for the TFP concentrations as low as 4 mM (∼0.03 wt%; 250 μm Li full cells; Fig. S14a, ESI†), with the optimal concentration being 12 mM (0.1 wt%). The NCM811‖50 μm Li metal full cells with the electrolyte containing 12 mM TFP maintained 94% of its discharge capacity even after 275 cycles (as measured at 0.1C) and an average coulombic efficiency of 99.6% (1C, Fig. 3a). In comparison, for the full cells containing 12 mM LiClO4 (Fig. S15, ESI†), we observed a much faster capacity loss than for 12 mM TFP-ClO4.
The charge/discharge voltage profiles (Fig. 3b) show that the cells containing fluorinated cations can be successfully and repetitively charged up to 4.2 V. The differential capacity profiles (dQ/dV, Fig. S16, ESI†) for the cells with TFP showed both phase transitions expected for the NCM811 cathode during charging/discharging [from hexagonal (H1) to monoclinic (M) occurring between 3.6 V and 3.8 V and from monoclinic to hexagonal (H2) at ∼4.0 V].35,36 This is in contrast to the cells without additive that showed severe discharge capacity fading when charging up to 4.2 V due to electrolyte decomposition (Fig. S14, ESI†).8 The electrochemical impedance spectroscopy data also show a minimal increase in impedance with cycling for the full cells with TFP (Fig. S17, ESI†), in good agreement with the cells’ dQ/dV profiles.
In addition, we found that the presence of TFP in the electrolyte helps to suppress current collector corrosion. SEM of the aluminum current collectors that supported the NCM811 cathode showed that after 160 cycles in the electrolyte with TFP+, the current collector had no signs of corrosion, while in the additive-free electrolyte a significant roughening of the Al current collector was observed already after 20 cycles (Fig. S18, ESI†). Based on the DFT calculations, we suggest that TFP+ may act as a scavenger of chloride ions and chlorine radicals that are often present in commercial LiFSI, resulting in decreased corrosion of aluminum and cathode surfaces (Fig. S19, ESI†).
Next, we performed molecular dynamics simulations (MD) to provide insights into electrolyte structure and the solvation of the TFP+ additive (Fig. 4a and Fig. S21, ESI†). Higher TFP+ concentrations were used in MD simulations to ensure that at least 12 TFP+ cations are present in a simulation cell. Two TFP+ concentrations (0.27 M and 0.14 M TFP) were examined to understand concentration dependence of additive solvation (Fig. S22, ESI†). A snapshot of the MD simulation cell is shown in Fig. 4 together with the representative Li+ solvates (Fig. S23, ESI†). At room temperature, the Li+(DME)3 solvates are prevalent: 75% of Li+ are not coordinated by either ClO4− or FSI− anions. The rest of Li+ are part of the Li+/FSI− and Li+/ClO4− contact ion pairs (CIPs) shown in Fig. 4c and Fig. S23a–d (ESI†).
Fig. 4b and Fig. S24 (ESI†) show the radial distribution functions (RDFs) for the respective electrolytes. We observe similar magnitudes of the first peak for Li+ with ether oxygen atoms of DME [EO(DME)] and O(ClO4−), while Li–O(FSI) peak is significantly smaller indicating strong affinity of Li+ to DME and ClO4− anions compared to FSI− (Fig. S24, ESI†). The TFP+–ClO4− RDF also shows a much higher first peak than that for TFP+–FSI− (Fig. 4b). Interestingly, both magnitude and widths of the first TFP+/ClO4− peak (Fig. 4b) are larger than the first peak in the Li+/ClO4− RDF (Fig. S24, ESI†), indicating significantly stronger predisposition of the TFP+ cation to form CIPs and aggregates than the corresponding lithium salts. We observe that while most of Li+ (75%) exist as free ions in agreement with the previous reports for 1 M LiFSI in DME,40 a significant fraction of TFP+ cations (∼80%) participate in CIPs and aggregates. Negatively charged aggregates such as TFP+(ClO4−)2 are expected to be found at the positive electrode surface, while free TFP+ and positive aggregates such as (TFP+)2(Anion−) are expected to be found at the negative electrode surface.
We observe the following concentration dependence of the TFP+ environments that is summarized in Table S2 (ESI†) for both [TFP+][ClO4−] concentrations. MD simulations also reveal an increase in “free” (not coordinated by either anion) TFP+ cations from 21% to 28% as the concentration of TFP+ decreases from 0.27 M to 0.14 M. This indicates a greater fraction of free TFP+ at lower concentrations. It is important to emphasize that our electrochemical experiments have TFP+ concentrations that are ∼10–25 times lower than those in the simulated electrolytes. Therefore, we expect an even higher fraction of free TFP+ in the 10–20 mM TFP solutions yielding a sufficient amount of free TFP+ to be electrostatically attracted to anode surface to participate in formation of LiF-rich SEI (in agreement with our EQCM-D results, Fig. 2c).
:
dichloroethane 1
:
3 (v
:
v) mixture at 60 °C to obtain the final product (1.4 g, 4.7 mmol, 94%). Approximately 0.5 g of pure N-methyl-2,4,6-trifluoropyridinium triflate was then dissolved in acetonitrile (2 mL) and passed through the CIO4− ion exchange column with a constant flow of acetonitrile. The eluted solution (ca. 75 mL) was subject to rotary evaporation to afford the crude N-methyl-2,4,6-trifluoropyridinium perchlorate solid salt. The compound was then recrystallized by allowing an acetonitrile:dichloroethane 1
:
3 (v
:
v) solution to cool from 40 °C to 0 °C to obtain the final product (0.24 g, 0.97 mmol, 48%).
IR: νmax 3070 cm−1, 2161 cm−1, 2034 cm−1, 1675 cm−1, 1597 cm−1, 1496 cm−1, 1167 cm−1, 1077 cm−1, 877 cm−1, 622 cm−1.
1H NMR (500 MHz, acetonitrile-d3) δ 1.94 (p, J = 2.5 Hz, 1H), 4.02 (s, 1H), 7.64 (dd, J = 7.0, 2.4 Hz, 1H). (Fig. S25a, ESI.†)
13C NMR (126 MHz, acetonitrile-d3) δ 1.31, 35.56, 100.99, 118.26, 159.01, 161.09, 176.58, 178.69. (Fig. S25b, ESI.†)
19F NMR (471 MHz, acetonitrile-d3) δ −75.32 (t, J = 28.6 Hz), −69.33 (s). (Fig. S25c, ESI.†)
HRMS (m/z) N-methyl-2,4,6-trifluoropyridinium cation [M + H]+: Calcd: 148.0368, found: 148.0370; perchlorate anion [M − H]−: Calcd: 98.9410, found: 98.8446.
Reduction potential for the complexes containing an additive, solvent, Li+ denoted as a complex A was calculated as the negative of the free energy of formation of the reduced species A− in solution [ΔGS298(A → A−) = GS298(A−) − GS298(A)] divided by Faraday's constant as given by:
The difference between the Li/Li+ and absolute reduction potential of 1.4 V was subtracted to convert results to Li/Li+ scale as discussed extensively elsewhere.43 Oxidation potential for a complex A was calculated as the free energy of formation of the oxidized specie A+ in solution [ΔGS298(A → A+) = GS298(A+) − GS298(A)] divided by Faraday's constant as given by:
The H-transfer reaction from DME to LiNiO2 cathode surface was adapted from previous work35 and is shown in Fig. 3d. Details and discussion of the molecular dynamics simulations is expanded upon in the ESI.†
Li‖Li symmetric cell tests were performed using CR2032 coin cell parts (Hohsen Corp), 250 μm Li chips (∅ = 11 mm, Xiamen Tmax Battery Equipments Limited), and 8 μm Cu current collectors (∅ = 13 mm, Xiamen Tmax Battery Equipment Limited). Two layers of Celgard 2500 were used as a separator (∅ = 19 mm). The amount of electrolyte used for every coin cell was ∼40–75 μl. The cells were aged at room temperature for at least 5 hours before operation. Galvanostatic cycling was performed for the symmetric cell tests with a current density of 10 mA cm−2 (6 min charge and 6 min discharge, specific areal capacity of 1 mA h cm−2).44 Biologic MPG-200 potentiostat and Arbin battery cycler (LBT21084UC) were used for the data collection.
The coulombic efficiency was obtained by averaging results of at least 5 independent measurements that were generated using a modified Aurbach protocol (described below). The modified Aurbach protocol was performed using asymmetric Cu‖Li coin cells, assembled with 75 μl of electrolyte. The detailed experimental sequence was the following: (1) 10 CV cycles were performed to pre-form SEI on a Cu electrode with a scan rate of 5 mV s−1 within the voltage range of 0 V to 2.5 V; (2) a deposition, stripping and re-deposition of excess amount of Li was performed (25 mA h cm−2 with a current density of 2.5 mA cm−2); (3) 100 galvanostatic Li plating and stripping cycles were performed at current density of 10 mA cm−2 and a specific areal capacity of 1 mA h cm−2; (4) final Li stripping was carried out with current density of 2.5 mA cm−2 with a cut off voltage of 1 V.
For full cell tests, a single-sided NCM811 electrodes (2 mA h cm−2 areal loading on 16 μm thick Al current collector, 99.6%, NEI Corp) were used as a cathode and Li metal chips (50 μm and 250 μm, Xiamen Tmax Battery Equipments Limited) were used as anodes. 50 μm thick Li metal anodes were prepared by thinning 250 μm thick Li foil using a roll press. 20 μm thick Li metal anodes were purchased from China Energy Lithium Co., Ltd. For full cell cycling tests a constant current (CC) step was initially used for all charging steps with a cut-off voltage of 4.2 V followed by a constant current constant voltage (CCCV) step with the terminating conditions of either 5% of the 1C-rate current or 30 minutes of the constant voltage step. A CC step was applied for all discharging steps with a cut-off voltage of 3 V. A sequence of three cycles at 0.1C-rate and 50 cycles at 1C-rate was repeated during full cell testing. EIS data were acquired using a VSP-300 (Biologic) in a frequency range from 200 kHz to 100 mHz, with a sinus amplitude of 10.0 mV.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ee00296b |
| This journal is © The Royal Society of Chemistry 2024 |