Skip to main content
ACS Materials Au logoLink to ACS Materials Au
. 2022 Nov 10;3(2):88–101. doi: 10.1021/acsmaterialsau.2c00060

Transition Metal Dissolution Mechanisms and Impacts on Electronic Conductivity in Composite LiNi0.5Mn1.5O4 Cathode Films

Julia C Hestenes , Jerzy T Sadowski , Richard May §, Lauren E Marbella §,*
PMCID: PMC9999480  PMID: 38089724

Abstract

graphic file with name mg2c00060_0007.jpg

The high-voltage LiNi0.5Mn1.5O4 (LNMO) spinel cathode material offers high energy density storage capabilities without the use of costly Co that is prevalent in other Li-ion battery chemistries (e.g., LiNixMnyCozO2 (NMC)). Unfortunately, LNMO-containing batteries suffer from poor cycling performance because of the intrinsically coupled processes of electrolyte oxidation and transition metal dissolution that occurs at high voltage. In this work, we use operando electron paramagnetic resonance (EPR) and nuclear magnetic resonance (NMR) spectroscopies to demonstrate that transition metal dissolution in LNMO is tightly coupled to HF formation (and thus, electrolyte oxidation reactions as detected with operando and in situ solution NMR), indicative of an acid-driven disproportionation reaction that occurs during delithiation (i.e., battery charging). Leveraging the temporal resolution (s-min) of magnetic resonance, we find that the LNMO particles accelerate the rate of LiPF6 decomposition and subsequent Mn2+ dissolution, possibly due to the acidic nature of terminal Mn-OH groups. X-ray photoemission electron microscopy (XPEEM) provides surface-sensitive and localized X-ray absorption spectroscopy (XAS) measurements, in addition to X-ray photoelectron spectroscopy (XPS), that indicate disproportionation is enabled by surface reconstruction upon charging, which leads to surface Mn3+ sites on the LNMO particle surface that can disproportionate into Mn2+(dissolved) and Mn4+(s). During discharge of the battery, we observe high quantities of metal fluorides (in particular, MnF2) in the cathode electrolyte interphase (CEI) on LNMO as well as the conductive carbon additives in the composite. Electronic conductivity measurements indicate that the MnF2 decreases film conductivity by threefold compared to LiF, suggesting that this CEI component may impede both the ionic and electronic properties of the cathode. Ultimately, to prevent transition metal dissolution and the associated side reactions in spinel-type cathodes (particularly those that operate at high voltages like LNMO), the use of electrolytes that offer improved anodic stability and prevent acid byproducts will likely be necessary.

Keywords: cathode electrolyte interphase (CEI), LNMO, operando electrochemical NMR, XPEEM, operando electrochemical EPR

Introduction

The spinel structure with composition LiNi0.5Mn1.5O4 (LNMO) is a promising cathode for Li-ion batteries (LIBs) because of its high energy density of 690 W h/kg1,2 and high rate performance.3,4 In contrast to other commercially available cathode materials currently used in LIBs, LNMO does not contain any Co, making it a potentially cheaper alternative that may also alleviate the serious geopolitical concerns associated with the continued mining of Co in the Democratic Republic of the Congo.5 However, LNMO-containing batteries suffer from poor cycling performance, likely due to the oxidation of organic electrolytes at high voltages (the average operating voltage of LNMO is 4.7 V vs Li/Li+)6,7 and transition metal dissolution from the active material.8 Transition metal dissolution often occurs in tandem with surface reconstruction into an ionically resistive surface layer that results in a loss of active material.911 The dissolved metals can chemically decompose the bulk electrolyte12 and/or deposit on the anode,1320 where Mn is particularly insidious, and can increase the permeability of the solid electrolyte interphase (SEI) to electrolyte.20,21 The byproducts formed during electrolyte oxidation (e.g., HF) may also react with dissolved transition metals to form resistive compounds in the cathode electrolyte interphase (CEI) that deposit on the composite surface and increase the barrier to Li intercalation.11,2227

Despite the clear correlation between transition metal dissolution and LNMO performance degradation, the link between the chemical and structural mechanisms driving this dissolution is not well understood. In the Mn-only spinel, LiMn2O4 (LMO) which contains mixed Mn3+/Mn4+ states, Mn dissolution can occur in the presence of acid (e.g., HF produced during electrolyte oxidation) via the disproportionation of 2Mn3+ centers into Mn2+ and Mn4+.28 Because Mn2+ is more soluble in standard carbonate electrolytes, it tends to dissolve, while Mn4+ forms MnO2-type structures on the active particle surface.11,29 However, in LNMO, 1/4 of the Mn atoms in LMO are replaced with Ni2+, so in principle, all of the Mn is in the 4+ oxidation state and LNMO is not expected to suffer from Mn3+ disproportionation and subsequent dissolution of Mn2+.28,30 In practice, trace Mn3+ is still found in pristine LNMO because of the formation of oxygen vacancies and impurity phases during synthesis,15,3033 which may explain, at least in a small part, why Mn dissolution still occurs in LNMO.7,14,15,26 However, contrary to expectations, multiple reports claim that Mn dissolution proceeds beyond the redox plateau of Mn3+/4+ at 4.1 V vs Li/Li+31 where all of the Mn should be in the 4+ oxidation state.7,29,34,35

In order to explain this anomalous observation, three possible explanations have been proposed in the literature: (1) Mn3+ impurities in the bulk of the active particles are not fully oxidized during charge and allow disproportionation reactions to occur at high voltage, (2) Mn4+ is reduced during oxidation of the electrolyte salt, forming both HF and Mn3+-containing surface phases for the disproportionation reaction, and/or (3) Mn4+ is reduced all the way to Mn2+ during electrolyte solvent oxidation, facilitating direct dissolution. Evidence for mechanism (2) was first reported by Choi et al., who reported that significantly more Mn had dissolved when LiPF6 was present in carbonate solvents during high temperature aging of LMO, indicating that PF6 rather than the electrolyte solvent reduces Mn4+.29 More recently, mechanism (2) was also supported by the observation of a 2 nm thick, Mn3+-containing surface layer on LNMO particles after charge.10 Mechanism (3) was proposed by both Aoshima et al.36 as well as Pieczonka et al.7 who claimed that the correlation between transition metal dissolution and diethyl carbonate (DEC) decomposition into ethanol (detected with gas chromatography–mass spectrometry (GC–MS)) provided evidence of a dissolution reaction by which Mn4+ is directly reduced to Mn2+ by ethanol at the surface of LMO and LNMO cathodes. A recent soft X-ray absorption spectroscopy (sXAS) study by Yang and coworkers34 found higher quantities of Mn2+-containing phases on the LNMO surface after the first charge compared to the first discharge, which led them to believe that surface Mn4+ was directly reduced to Mn2+ upon reaction with the electrolyte at high voltage, supporting mechanism (3). While mechanisms (2) and (3) are similar in that they both agree that dissolution is enabled by the reactivity of Mn4+ with the electrolyte, they differ in whether dissolution occurs by HF-driven Mn3+ disproportionation or via direct dissolution of Mn2+ in the absence of HF.36 Because these reaction mechanisms have overlap between reactants and products (e.g., dissolved Mn2+, HF, and other electrolyte decomposition species), it is difficult to discern experimentally which mechanism is occurring with traditional ex situ techniques. Additionally, while it is generally understood that HF leads to the dissolution of transition metal (e.g., the disproportionation of Mn3+), there is some debate on whether HF is present in the electrolyte because of trace water (from hydrolysis of LiPF6) or if it is generated due to electrolyte oxidation reactions (that also produce water, but lead to increasing quantities of HF at high voltage because hydrolysis is exacerbated).7,3739

To address this gap in understanding, we use a combination of operando 19F NMR and EPR to simultaneously track HF formation and Mn2+ dissolution as a function of charging voltage in an LNMO-containing battery. Here, the operando experiment was key to tracking the kinetics and concentrations both HF and Mn2+ species as they were generated during the first charge at room temperature, in contrast to ex situ methods,7,9,19,29,4045 many of which rely on high temperature aging procedures and postmortem analyses, in which dissolved Mn species have been shown to react during sample preparation.46 With the operando measurements, we found a strong correlation between HF formation and dissolved Mn concentration that increased with voltage, suggesting that Mn dissolution is driven by HF produced by electrolyte oxidation. This reaction begins ≤ 1 h into charging, suggesting that the LNMO particles themselves lower the activation barrier to electrolyte decomposition. At the end of charge, electrolyte oxidation products, such as vinyl compounds, were detected with in situ 1H NMR spectroscopy. We used a combination of X-ray photoemission electron microscopy (XPEEM) and X-ray photoelectron spectroscopy (XPS) to detect changes in the particle surface structure and oxidation state, including the emergence of Mn2+/3+-containing phases that form during charge and feed the transition metal disproportionation/dissolution reaction at high voltage. In addition, XPEEM, XPS, and solution NMR spectroscopy were used to characterize the chemical composition of the resulting CEI and show that this layer is rich in MnF2, which covers the entire composite surface. While it is well known that metal fluorides in the CEI hinder Li ion transport to/from the active particles, we use four-line probe measurements to demonstrate that MnF2 also disrupts the electronic conductivity of the LNMO cathode film, more so than LiF, possibly also contributing to cell degradation.

Materials and Methods

Materials

Li metal ribbon (0.75 mm thick), 1 M LiPF6 in ethylene carbonate:dimethyl carbonate (EC:DMC 1:1 v/v, LP30, battery grade), N-methyl-2-pyrrolidinone (NMP, anhydrous, 99.5%), LiF (≥99.9%), and MnF2 (≥98%) were purchased from Sigma-Aldrich. DMSO-d6 (≥99.9%) was purchased from Cambridge Isotope Laboratories. Prior to use, DMSO-d6 was dried over molecular sieves in an Ar-filled glovebox (O2 < 0.1 ppm, H2O < 0.5 ppm) for 48 h. The surface oxide layer on Li metal was brushed off prior to assembly in Li/LNMO half cells. All other chemicals were used as received. LNMO cathode powder was purchased from MTI Corporation and stored in an Ar-filled glovebox. Carbon super P C45 and polyvinylidene fluoride (PVDF) binder were purchased from MTI Corporation and used as received.

Electrode Fabrication

LNMO cathode films were prepared from a slurry of 8:1:1 LNMO:C45:PVDF. The black powders, LNMO and C45, were first hand-ground in a mortar and pestle for 10 min. This mixture was then added to a solution of the PVDF binder in NMP to create a viscous slurry. The slurry was then cast onto an Al current collector (25 μm thick, MTI Corporation) using a 150 μm doctor blade and dried at 100 °C under vacuum overnight. The dried film was punched into 12.7 mm diameter disks to use in cell assembly. Once dried, cathode films were stored in an Ar-filled glovebox to minimize exposure to air and moisture. Typical mass loadings of active material per cathode were 4–9 mg cm–2. These electrodes were used for all electrochemical testing and extracted for postmortem characterization (see XPEEM and XPS sections below).

Electrochemistry

Electrochemical tests for ex situ XPEEM and XPS were conducted using 2032 coin cells assembled in an Ar-filled glovebox with a 12.7 mm Li metal disc as the anode and LNMO films as the cathode. Each cell used a Whatman glass microfiber (GF/A) separator and was cycled in ∼0.2 mL of battery-grade, LP30 electrolyte. Galvanostatic cycling experiments were performed using C-rates of C/10 or C/20 (based on theoretical capacities of 147 mAh/g for LNMO) between 3.5 and 5 V vs Li/Li+.

Ex Situ and Operando Solution NMR

All solution NMR experiments were performed on a Bruker Avance III 400 spectrometer equipped with a triple resonance broadband observe (TBO) probe head. One-dimensional (1D) 1H (30° single pulse, 1 s recycle delay, 16 scans, internally referenced to residual DMSO solvent at 2.5 ppm) and 19F (30° single pulse, 2 s recycle delay, 32 scans, internally referenced to PF6 electrolyte salt at −74.5 ppm) were recorded at room temperature.

For ex situ solution NMR measurements, Li/LNMO half cells were disassembled in the glovebox after cycling. The glass fiber separator was soaked in 1 mL of DMSO-d6 for 5 min and compressed using clean tweezers to fully extract the electrolyte from the separator. The resulting solution was pipetted into a 5 mm airtight J-Young NMR tube for analysis.

For operando solution NMR, measurements were collected in a previously reported operando electrochemical NMR tube cell37 (shown in Figure S1). In brief, the cell contains a LNMO cathode, a Cu anode, and a LP30 electrolyte (300 μL) assembled in a 3.65 mm o.d. FEP (fluorinated ethylene polypropylene copolymer) NMR tube liner (Wilmad LabGlass, dried at 60 ° C overnight prior to use). The tube cell is submerged into a 5 mm o.d. glass NMR tube containing 0.1 mL of DMSO-d6 (for locking and shimming) and secured in place with Teflon tape, allowing us to collect high-resolution 1H and 19F NMR spectra during electrochemical cycling. Electrodes were placed into the tube approximately 25 mm from the bottom of the FEP liner, completely submerged in the electrolyte. The cathode used in the operando experiments is a 20 × 2.5 mm wide rectangular film cast on an 8 in long strip of Al foil, where only the end section containing the LNMO cathode material is submerged in the electrolyte and the bare Al foil runs to the end of the tube as a current collector and electrical lead. A 20 × 2.5 mm piece of Cu film (6 μm thick) was used as the anode, which was rinsed in acetone and dried prior to use. The Cu anode was connected via heat-shrink chemically resistant electrical insulation tubing to a 22 cm long pure Cu wire, whereby upon heat-treating the tubing, the shrinkage electrically connects the Cu anode to the Cu wire within the tubing. The heat-shrink tubing prevented cell shorting along the length of the NMR tube. The Cu anode was wrapped in a single layer of Celgard 2325 separator (dried at 60 °C for 24 h prior to use). The active material loading was estimated such that the first cycle was consistent with C/10 cycling (∼2–3 mg), which was typically ∼30–40 μA.

Prior to the experiment, the tube cell was connected to double-shielded wires which were connected to an electromagnetic shield (NMR Service, GmbH). Once the cell was inserted into the NMR magnet, the shield was placed to lie on top of the upper magnet bore. The double-shielded wires extended out from the other side of the shield to connect to a Biologic SP-150 potentiostat for electrochemical cycling. Prior to acquisition, the sample was locked to DMSO-d6 and shimmed, which was key for detecting the dilute species reported in this work. NMR acquisition parameters are the same as described above for ex situ solution NMR, except 128 scans were used for 19F acquisition, making each acquisition 7 min and 2 s long, defining the time resolution of the operando experiment. All operando measurements were repeated at least twice. The resulting 2D NMR experiments were analyzed in Python by importing the data with the Python package Nmrglue.47 The concentration of HF in the 19F operando measurement was determined by comparing the integrated area of the HF doublet at each time point to that of the LiPF6 doublet in the 19F NMR taken just before the electrochemical measurement started, which corresponds to 1 M LiPF6 (pristine LP30). The complete charge profile for the electrochemistry shown in Figure 1 is shown in Figure S2 to confirm that the operando cell exhibits the expected phase transitions in the LNMO charge curve, where there is a slight plateau beginning at 4.1 V for a small amount of Mn3+/4+ redox and another set of plateaus beginning at approximately 4.7 and 4.75 V for the Ni2+/3+ and Ni3+/4+ redox reactions48,49 (Figure S3).

Figure 1.

Figure 1

(a) Charge profile of the operando LNMO/Cu cell at a rate of C/10 used in the operando solution NMR experiment shown in (b). This charge profile is also representative of the tube cell used for the operando EPR experiment shown in (c), where the tube cells were charged to 5 V vs Li/Li+ at a rate of C/10. Circles are placed along the voltage profile at the values that correspond to select voltages of each color-coded NMR/EPR spectrum shown in panel (b) and (c) as follows: OCV, 4.68, 4.73, 4.76, 4.80, 4.90, and 5 V vs Li/Li+. (b) Representative 19F solution NMR spectra extracted from the operando solution NMR experiment showing HF production during charge. (c) Representative EPR spectra from the operando EPR experiment showing Mn2+ dissolution into the electrolyte during charge. (d) Concentration of HF and Mn2+ detected during charge of the operando cell, which were calculated from area integrals of all spectra recorded for the experiments shown in (b,c). For the EPR, the integrals of the absolute value of the spectra were used. The cycling profile from (a)/Figure S2 is reprinted for visual aid.

EPR Spectroscopy

X-band EPR measurements were performed on an ESR-5000 spectrometer operating at a microwave frequency of 9.47 GHz manufactured by Freiberg Instruments for Bruker, sold by Rotunda Scientific Technologies. All measurements used a B-field sweep from 300 to 370 mT, a microwave power of 5 mW, a 0.5 mT modulation amplitude with 100 kHz frequency, and a 60 s scan time for a single scan. All EPR spectra were processed using background subtraction to remove the slanted baseline.

For operando EPR measurements, EPR spectra were acquired every 15 min, capturing a 60 s snapshot of the chemistry in the cell at a given voltage. The operando EPR measurements used the same electrochemical tube cell as the operando NMR measurements described in the previous sections with the exception that no glass NMR tube containing DMSO-d6 for locking and shimming was required (i.e., the FEP tube cell was inserted into the EPR cavity as is). The cell was directly connected to the Biologic SP-150 potentiostat via alligator clips. The concentration of Mn2+ in the EPR operando measurement was quantified by comparing the integral of the absolute value of each spectrum in the operando experiment to that of the integrated area of the absolute value of a spectrum of the Mn2+ signal from a standard sample of known Mn2+ concentration (300 μL of 200 ppm Mn(II) triflate dissolved in LP30 in an FEP tube liner).

X-Ray Photoemission Electron Microscopy

Prior to electrochemical cycling, LNMO cathode films fabricated as described in the “Electrode Fabrication” section were pressed at 2 tons (40 MPa) in a hydraulic press (Specac) to minimize sample roughness for analysis with XPEEM. After cycling, cathode films were removed from the coin cells in the glovebox and dipped twice into 1 mL of DMC for 3 s each to remove residual salts and minimize surface layer outgassing in the XPEEM vacuum chamber. Cathode films were then dried under vacuum overnight to remove residual solvent. Samples were double-sealed in airtight pouches under Ar and transferred to the XPEEM facility at the Center for Functional Nanomaterials at Brookhaven National Laboratory (BNL) for measurement. Samples were mounted on XPEEM holders in an Ar-filled glovebox and quickly transferred to the XPEEM vacuum chamber, where air exposure during the transfer was less than 20 s.

Pixel-wise X-ray absorption spectra (XAS) were obtained by recording a series of XPEEM images at each energy in each absorption edge range at sequential increments of 0.2 eV. Mn L-edge spectra were referenced to the low energy peak characteristic of Mn2+, which has a reported value of 640 eV.34 Other edges were translated accordingly, except the O K-edge which was referenced to the sharp low energy peak occurring from LNMO particles at 529.5 eV50 and the C K-edge which referenced the lowest energy peak, assigned to conductive carbon, at 285.4 eV.51,52 Local XAS spectra can be extracted from the XPEEM images by plotting the intensity of a given spatial region (set of pixels) at each energy. For example, for Mn L-edge, 112 images were acquired from 635 to 657 eV. All local XAS spectra extracted were normalized individually to their highest intensity value, unless otherwise noted. XPEEM elemental contrast images were obtained by subtracting the pre-edge image from the desired X-ray absorption energy. The C K-edge image at 293.0 eV was extracted to display the spatial distribution of conductive carbon. The pre-edge image at 280.0 eV was then subtracted, yielding the final elemental map of conductive carbon. The same procedure was conducted for other elemental maps, as specified in Figures 2 and 3. Representative C K-edge spectra are shown in Figure S4, where we followed the same normalization procedure as outlined by El-Kazzi and co-workers51 and our previous study.37 In brief, we normalized our measured C-K edge spectra by that of the C K-edge spectrum of an Au microparticle deposited on a pristine LNMO sample that was imaged in XPEEM. The C K-edge spectra in Figure S4 on conductive carbon regions exhibit characteristic absorption peaks at 285.4 and 291.9 eV that correspond to C=C π* and σ* transitions.51

Figure 2.

Figure 2

(a) XPEEM elemental images of a pristine LNMO cathode film (top) and an LNMO cathode film that underwent 50 cycles (cycled at C/10, disassembled at 5 V vs Li/Li+). LNMO particles are mapped in magenta via the image corresponding to 643.0 eV and conductive carbon is mapped in cyan using the image at 293.0 eV. Dark regions indicate surface roughness caused by indirect scattering. Upon probing the F K-edge energies, no fluorine absorption signal was detected within the detection depth of the technique (∼2–10 nm), suggesting that PVDF was not present within the first few nanometers of sample imaged here. (b) Local O K-edge and (c) local Mn L-edge XAS spectra extracted for LNMO particles on the pristine sample or the cycled sample in (a), where the color of the spectrum corresponds to the location of the marker of the same color in (a). All images scale to an area of 30 × 30 μm. In the XAS spectra, dashed lines are marked along distinctive features for visual aid.

Figure 3.

Figure 3

XPEEM elemental images of pristine LNMO (a) before cycling, reproduced from Figure 2a and (b) 50 cycles (cycled at C/10, disassembled at 3.5 V), where LNMO particles are mapped in magenta via the image corresponding to 643 eV and conductive carbon is mapped in cyan using the image at 293.0 eV. (c) Reproduced image from (b) overlaid with an elemental map of Mn2+ (yellow, image corresponding to 640 eV). (d) Local Mn L-edge XAS spectra extracted from the pristine sample, regions as marked in (a). (e) Local Mn L-edge XAS spectra extracted from the 50 cycles sample, regions as marked in (b). (f) Local F K-edge XAS extracted from the 50 cycles sample, regions as marked in (b). All images scale to an area of 30 × 30 μm. In the XAS spectra, dashed lines are marked along distinctive features for visual aid.

X-Ray Photoelectron Spectroscopy

After cycling, the cathodes were removed from coin cells in the glovebox and dipped twice into 1 mL of DMC for 3 s each to remove residual salts and minimize sample outgassing. Cathode films were then dried under vacuum overnight to remove residual solvent. The samples were brought in airtight containers to the XPS instrument, then quickly mounted on the XPS chamber (air exposure <30 s) prior to pumping down the XPS antechamber. Spectra were collected using a PHI 5600 XPS system with a hemispherical analyzer and an Al X-ray source with XPS base chamber pressure <1.0 × 10–8 Torr. XPSPEAK software was used to fit spectra. The adventitious carbon peak in the C 1s spectrum of each sample was referenced to 284.8 eV. All peaks were fit using a Shirley baseline correction with two constraints: (i) the Gaussian–Lorentzian ratio was the same for all peaks in each orbital and (ii) the full width at half maximum (FWHM) was the same for all peaks in each orbital and constrained to < 1.7 eV.

Electronic Conductivity Measurements

We followed the procedure reported by Peterson and Wheeler53 to construct a four-line probe to measure the electronic conductivity of LNMO composite cathode films to determine the impact of different surface coatings. LNMO cathode slurries were combined with various quantities of inorganic CEI species, such as MnF2 and LiF, to see how these components altered electronic conductivity. In these experiments, LNMO powder and C45 were preground for 10 min in a mortar and pestle before combining with PVDF in NMP as described in the Electrode Fabrication section above. At this stage, various quantities of metal fluorides (5.0, 9.7, 23 and 6.4 mg of MnF2 and LiF, respectively) were added to the slurry and mixed in a mortar and pestle for 2 min prior to casting as a 22 × 22 mm square-shaped film using a 150 μm doctor blade onto a 22 × 75 × 1 mm glass microscope slide or PTFE sheet (for MnF2 samples, to avoid reaction with glass). The final film thickness ranged from 100 to 150 μm and typical mass loading ranged from 45 to 65 mg.

The film-containing slide was placed face down onto the four-line probe. A 500 g weight was placed on top of the sample to ensure good electrical contact between the probe and the film. A current of 0.3 mA was run through the outer leads using an Arbin LBT20084 battery cycler. The voltage between the inner two probe lines was measured using a Keysight Technologies 34461A 6 1/2 Digital Multimeter. The electronic conductivity was then measured using eq 1, where L is the length of the sample along the direction of the current, I is the current being passed through the sample, and S is the cross-sectional area of the sample perpendicular to the direction of the current.

graphic file with name mg2c00060_m001.jpg 1

Four measurements were made on a sample to compute the displayed error bars. The sample was rotated clockwise 90° between each of the four measurements.

Scanning Electron Microscopy

Scanning electron microscopy (SEM) images of pristine LNMO composite cathode films used in this study (pressed at 2 tons for XPEEM measurements) were taken using a Zeiss Sigma VP Schottky Thermal Field emission SEM with a Gemini objective lens.

Results and Discussion

Operando Monitoring of HF Formation and Mn2+ Dissolution

We performed operando 19F solution NMR and operando EPR experiments to determine how HF formation is coupled to Mn dissolution during the first charge of LNMO in a working battery (Figure 1). At the open circuit voltage (OCV), there is little to no visible 19F NMR signal in the region corresponding to HF, nor Mn in the EPR, indicating that neither HF nor Mn are present in the electrolyte (or that they are below the detection limit of the measurement) upon battery assembly (Figure 1, bottom black spectra). In 19F NMR, a doublet centered at −191.3 ppm assigned to HF54,55 is observed, where the doublet indicates 1H–19F coupling (J = 474 Hz) (Figure 1b). HF prominently appears upon charging to 4.68 V vs Li/Li+, at approximately 61 min into the 10 h long charge experiment and persists at higher voltages (1st spectrum from bottom, Figure 1b), although fluctuations in HF concentration can be observed at earlier time points (< 1 h, Figure 1d). This coincides with the onset of electrochemical oxidation of the electrolyte and is intriguing because the operando measurement indicates that the kinetics of this process is much faster than that previously reported (on the order of 50 h,56 this point will be discussed later). Meanwhile in the EPR, a sextet from the hyperfine interactions characteristic of Mn2+,18,57,58 is first detected at 4.73 V (2nd spectrum from the bottom, Figure 1c), suggesting that HF formation closely coincides with transition metal dissolution in LNMO. As charging continues, the signal intensity associated with the spectroscopic signatures of both HF and Mn2+ continue to increase (Figure 1d), indicating that their concentrations rise in tandem with one another. The HF and Mn2+ concentrations shown in Figure 1d were determined by relating the area integrals of the NMR and EPR spectra from the experiment to the spectra of species of known concentration (see the corresponding Experimental sections). When first detected, HF and Mn2+ are around 15 ppm and by the end of the charge, the concentration of HF is approximately 120 ppm while the concentration of Mn2+ is about 90 ppm.

The operando NMR and EPR measurements shown in Figure 1 indicate that HF formation is strongly correlated with Mn2+ dissolution from the LNMO cathode. Although the increase in Mn dissolution with voltage is in line with previous reports for LNMO,7,38 it is surprising that it is so strongly correlated with HF in the electrolyte because this finding suggests that Mn dissolution occurs via acid-driven disproportionation reaction at high voltage when Mn should be in the 4+ oxidation state.28 As stated previously, for transition metal dissolution to occur via this route, the Mn at the particle surface must be in the 3+ oxidation state. In addition, the HF concentration continues to increase at a constant rate, even though HF is likely quickly consumed at the particle surface during disproportionation to produce soluble Mn2+ (which is observed simultaneously). Continuous HF production is consistent with oxide surface etching reactions proposed by Benedek and Thackeray (that form dissolved transition metal cations or transition metal fluorides) that result in H2O formation as a byproduct28 and go on to participate in subsequent hydrolysis reactions with LiPF6 to replenish the HF in a cyclic manner.59,60 It is also possible that the local HF concentration closest to the electrodes (which is outside of the measurement area) differs from the average, bulk concentration that we measure in the experiment.

Structural Rearrangements at the Particle Surface Feed the Mn Disproportionation Reaction

To determine if the LNMO active particles exhibit unexpected changes in the Mn oxidation state at a high potential, we performed a series of XPEEM and XPS measurements. XPEEM allows us to extract soft X-ray absorption (sXAS) spectra from different positions on the composite cathode surface before and after cycling with a lateral resolution of approximately 100 nm (Figure 2). XPEEM images acquired for a pristine LNMO film and a film cycled for 50 cycles that was disassembled at 5 V vs Li/Li+ are shown in Figure 2a. Representative 1st and 50th cycle electrochemical charge curves of the cells used in both XPEEM and XPS are included in Figure S5. LNMO particles are shown in magenta and conductive carbon regions are mapped in cyan, while dark regions indicate surface roughness that caused indirect scattering. SEM images of pristine pressed cathode films are shown in Figure S6, which support the spatial arrangement of LNMO and conductive carbon seen in the XPEEM maps. Although Mn2+, Mn3+, and Mn4+ share absorption peaks at several energies in Mn L-edge and O K-edge XAS, one can identify the oxidation states present via qualitative inspection of the relative intensity of peaks within each spectrum and comparison to reference spectra of standard Mn oxides, which are provided in Figures S7 and S8.34,61 Extraction of O K-edge and Mn L-edge sXAS spectra from regions that contain LNMO particles in the pristine sample (Figure 2a, top) yields absorption spectra (Figure 2b,c, black) that are consistent with a mixed Mn3+/4+ oxidation state expected for Mn4+ in uncycled LNMO (e.g., MnO6 coordination environments) and associated Mn3+ impurities from synthesis.34,50,62 The assignments from XPEEM are consistent with findings from XPS, where we see mixed Mn3+/4+ oxidation states upon discharging the cells to 3.5 V vs Li/Li+ (Figure S9).

In contrast, the XPEEM images of the cycled sample that was charged to 5 V (Figure 2a, bottom) yield multiple unique O K-edge and Mn L-edge absorption spectra for individual LNMO particles, suggesting that the Mn oxidation state at the particle surface is heterogeneous after cycling (Figure 2b,c). As expected for the charged state of LNMO, most of the Mn on the surface of the LNMO particles is present in the 4+ oxidation state (Figure 2, gray). The Mn4+ detected here corresponds to fully delithiated LNMO at the particle surface (Ni0.5Mn1.5O4) and/or a surface reconstruction phase that contains Mn4+ (e.g., λ-MnO2 as proposed by Aurbach and co-workers11,63).10,61,62,64 Other particles that show similar O K-edge and Mn L-edge spectra are marked in Figure S10 of the SI. However, we find that some LNMO particles in the charged state exhibit O K-edge and Mn L-edge spectra that are consistent with mixed Mn2+/Mn3+ or Mn3+ phases, such as Mn3O4-like and Mn2O3-like compounds, respectively (Figure 2, light and dark green).61 In particular, the O K-edge spectrum corresponding to the light green region has a broad feature centered at 537 eV that is larger in intensity than the low energy transitions at 529.5 and 531 eV, consistent with that observed for Mn2O3-like phases.61 In the Mn L-edge spectrum for this region (Figure 2c, light green), the ratio of the peak at ∼641.5 eV has slightly increased relative to the peak at ∼643 eV in the pristine sample, which is also consistent with a Mn2O3-like (Mn3+) phase.61 In comparison, O K-edge and Mn L-edge spectra extracted from the dark green region are indicative of a Mn3O4-like (Mn2+/3+) phase (Figure 2, dark green). In the O K-edge spectrum extracted for the dark green region, this Mn3O4-like phase is supported by the increased intensity of the peak at 531 eV with respect to the peak at 529.5 eV (Figure 2b, dark green).61 This observation is complemented by the corresponding Mn L-edge spectrum which shows that the sharp peak at 640 eV has increased in relative intensity to those at ∼641.5 and ∼643 eV, which is associated with the absorption of Mn2+ (Figure 2c, dark green and Figure S8).34 We also note that the presence of reduced Mn species on the LNMO particle surface is supported by changes in the spectral region ranging from approximately 650–655 eV (Figures 2c and S8). With the ∼2–10 nm detection depth of this experiment, we may also be capturing some of the Mn4+ from the underlying particle in these particular regions, but the changes in the relative peak intensities indicate that there is a lower oxidation state on the particle surface. Likewise, XPS measurements from cells disassembled at 5 V after electrochemical cycling also show substantial quantities of both Mn3+ and Mn4+ (Figure S9), supporting that LNMO particles undergo surface structural rearrangement at high voltage. The presence of Mn3+-containing phases on the surface of LNMO particles in the charged state is theoretically unexpected, as all Mn3+ impurities should oxidize to Mn4+ during the plateau in the LNMO charge curve at ∼4.1 V.31 However, this observation is in agreement with more recent scanning transmission electron microscopy (STEM) experiments that confirm the existence of a thin (∼2 nm) Mn3O4 phase on the surface of LNMO upon charging10 as well as sXAS results that detected Mn3+ species on the surface of LNMO after charging.34 Taken together, our results combined with others provide evidence that structural rearrangements at the LNMO particle surface produce a source of Mn3+ that can participate in the acid-driven disproportionation reaction, producing Mn2+ species in the electrolyte.

Spatially Resolved Characterization of the Cathode Electrolyte Interphase on LNMO

In addition to structural changes at the active particle surface, we also used a combination of XPEEM, XPS, and solution NMR to characterize the composition and arrangement of the CEI produced after electrochemical cycling. For analysis of the CEI, XPEEM images were acquired for LNMO films cycled for 50 cycles and disassembled at 3.5 V and compared to the pristine LNMO sample (Figure 3a,b). We note that for CEI characterization, it is critical to analyze the cathode at discharge (i.e., the lithiated state) because of the instability of the CEI at high voltage,65,66 which has been observed for this material in particular.23,38,67 The discharged state is also where we expect the CEI to be the thickest because of chemical cross-talk between the anode and cathode.23,68,69 Observing the CEI when it is thickest will minimize spectral overlap with the LNMO particles for surface-sensitive techniques like XPEEM and XPS, both of which have a maximum penetration depth of ∼10 nm. For completeness, additional samples analyzed on the 50th charge (disassembled at 5 V), as well as the 1st charge (5 V) and discharge (3.5 V), are included in Figure S11.

After 50 electrochemical charge cycles and disassembly at 3.5 V vs Li/Li+, we find that LNMO particles (Figure 3b, green box) are in the Mn3+/Mn4+ mixed oxidation state at 3.5 V. In addition, we find that the LNMO particles also exhibit a sharp absorption at 640 eV, indicating that a layer of Mn2+ has formed on the particles. When we extract local Mn L-edge spectra from areas of the surface containing conductive carbon (Figure 3b, gray and black boxes), we detect absorption spectra consistent with Mn2+ in both regions. Local F K-edge XAS spectra from the conductive carbon indicates that the Mn2+ is likely deposited as MnF2 (shoulder at 689.6 eV,70Figure 3f) because transition metal fluoride species are distributed throughout the composite surface as well, although most concentrated on active particles (Figure S12). The deposition of MnF2 on the cathode surface is consistent with previous reports on LiMn2O422,64,71,72 and LNMO.7,34,50Figure 3c reproduces the image shown in Figure 3b, but with the addition of an elemental map that corresponds to Mn2+ (yellow), where we can see that Mn2+ is detected on almost the entire composite surface. In addition to finding MnF2 in the CEI, F K-edge measurements also detect LiF (peak at 702 eV)73 distributed broadly on the surface after one and 50 cycles (Figure S13). The peak at 692 eV may originate from the PVDF binder, LiF, or residual LiPF6.73,74 We note that while it is likely that the CEI also contains NiF2 in the metal fluoride region, this cannot be distinguished from the Ni in the LNMO at the Ni L-edge to confirm the assignment. O K-edge measurements detect Li2CO374 distributed broadly on the surface after one and 50 cycles (Figure S14). In contrast to the discharged state, XPEEM measurements in the charged state did not detect signals from CEI species such as metal fluorides in the F K-edge or Li2CO3 in the O K-edge (Figure S15). Mn L-edge measurements, with high sensitivity compared to lighter elements, detected Mn2+ on the surface composite on charge, but in very dilute concentrations compared to on discharge (Figures S16 and S17). Less CEI on the composite in the charged state indicates that some CEI compounds may not be stable at 5 V.

XPS measurements of cycled LNMO cathodes disassembled at both charge and discharge confirm our XPEEM observations, which show that the quantities of metal fluorides in the CEI fluctuate with voltage (Figure S18). In the F 1s orbital spectra, a peak at 685.9 eV assigned to metal fluorides (e.g., LiF, MeFx)16,64,72 is detected on discharge but not on charge. This apparent dissipation of metal fluorides upon charging is consistent with previous reports that find that the CEI thickens upon discharge,23,73 possibly due to voltage instability.23,38,67 To explain this phenomenon, Oh and co-workers67 hypothesized that these components may be soluble in HF. Conveniently, the absence of metal fluorides at 5 V allowed for the detection of changes in oxidation state due to crystallographic changes at the particle surface with XPEEM (Figure 2) and CEI compositional changes at discharge (Figure 3).

In addition to the inorganic CEI, we can also assign changes in the organic CEI with C 1s and O 1s spectra from XPS (Figure S18). At both charge and discharge, we see peaks that correspond to adventitious/conductive carbon (284.8 eV) and the PVDF binder (286.5, 290.4, and 292.2 eV16) in the C 1s spectra that can also be identified in the pristine cathode film. The peak at 288.2 eV in all of the samples corresponds to C=O environments that may arise from residual Li2CO3 (as seen in the pristine sample), Li alkyl carbonates, other carbonate decomposition compounds, and/or residual EC/DMC solvent.75,76 The O 1s spectra for the charged sample show further evidence of solvent decomposition in the cycled samples, with a peak at 531.6 eV for C=O (e.g., from Li2CO3),23 532.5 eV for O–H or O–Li species (e.g., ROCO2Li, LiOH, or MnOH),76,77 533.8 eV for C–O-containing compounds,23 and 535.5 eV that may arise from salt decomposition products78 (e.g., LixPFyOz79 or O-CF2-containing ethers80). For the discharged sample, the C 1s spectrum shows the same peaks and relative compositions as the charged spectrum, except there is no 292.2 eV peak for the binder, indicating a thicker CEI on discharge. The O 1s spectrum on discharge shows that ROCO2Li (C–O) species, hydroxyl species, and polyethers (C–O) are also present as they were on charge.

The electrolyte oxidation reactions that occur at high voltage were also examined via additional analysis of the operando NMR measurements as well as further in and ex situ NMR experiments because these eliminate the need for sample washing prior to analysis when compared to X-ray techniques. One hour after the operando 19F solution NMR experiment shown in Figure 1 concluded (i.e., once the cell finished charging to 5 V), in situ 1H NMR was acquired to detect the organic decomposition products produced during charge (Figure S19). In this measurement, we found soluble species indicative of EC decomposition, such as vinyl compounds and Li formate. Unfortunately, NMR signals from 3.5 to 4.5 ppm overlap with EC/DMC peaks from the bulk electrolyte,37 so these measurements must be complemented with ex situ methods. Ex situ solution NMR of the spent electrolyte extracted from an LNMO/Li coin cell after 50 cycles exhibited 1H resonances at 3.71 and 4.29 ppm corresponding to DMDOHD,37 an oligomer formed from solvent oxidation (Figure S20).

Further inspection of the full spectra acquired during the operando 19F solution NMR experiment presented in Figure 1 reveals several fluorine species characteristic of LiPF6 degradation (see Figure 4). The pristine electrolyte shows a doublet centered at −85.2 ppm (d, JP–F = 938 Hz) that is assigned to PF2O2, a product from PF6 breakdown. This doublet grows in intensity with voltage, indicating that PF6 decomposes at high voltage. A doublet at −90.1 ppm (d, JP–F = 1065 Hz) emerges at 4.68 V and then another at −88.5 ppm (d, JP–F = 1004 Hz) appears at 4.76 V that correspond to POF381 and OPF2(OCH3), respectively.78 Both doublets increase in intensity with voltage after initial detection, suggesting that the formation of these compounds coincides directly with the decomposition of LiPF6 at high voltage. POF3 and HF are simultaneously detected at 4.68 V, which raises questions regarding the kinetics of the proposed equations (eqs 2 and 3) below for PF6 hydrolysis reactions that produce these products:29,39,82,83

graphic file with name mg2c00060_m002.jpg 2
graphic file with name mg2c00060_m003.jpg 3

Figure 4.

Figure 4

Selected spectra reproduced from the operando 19F NMR experiment in Figure 1b showing the chemical shift range of −82 to −93 ppm. The baseline appears slanted because of the overlap with the wide FWHM of the much larger LiPF6 peak nearby at −74.5 ppm.

The operando measurements performed here offer the time resolution necessary to clearly show that HF and POF3 production occur within 1 h of charging (Figures 1 and 4), rather than multiple hours as previously reported.56,59,8486 Using both experimental methods and molecular dynamics (MD) simulations, Liu et al.85 found that under high voltage conditions, the oxidation barrier for PF6-H2O coordination complexes can be lowered to ∼1 eV for a one electron oxidation reaction shown in eq 4, yet this still corresponds to a reaction time > 1 h.

graphic file with name mg2c00060_m004.jpg 4

The faster kinetics observed in our experiments can be rationalized in two possible ways: (i) eqs 24 are accurate, but in a real system are catalyzed by reactive metal centers on the LNMO particle surface during battery operation or (ii) in addition to eqs 24, an entirely different (and faster) reaction occurs at the cathode that feeds the production of HF. In scenario (i), Mn and/or Ni sites on LNMO may lower the activation energy for LiPF6 hydrolysis or oxidation, leading to faster kinetics during charging. Several reports have hypothesized that the partially or fully delithiated structure, Ni0.5Mn1.5O4 (isostructural with λ-MnO2), is catalytic,6,7,29,34,36,43,87 but this property is still not well understood. In addition to water-driven hydrolysis, the LNMO particles themselves may oxidize or hydrolyze PF6 in the absence of H2O that leads to the production of HF (scenario (ii)). Choi et al. suggested that direct PF6 oxidation at the particle surface causes concomitant surface transition metal reduction and proton abstraction from a nearby solvent molecule or water via eq 5,29 but the kinetics of this process at the LNMO surface are also not well understood.

graphic file with name mg2c00060_m005.jpg 5

This general mechanism was supported by Solchenbach et al.,86 who recently consolidated several density functional theory (DFT) predictions8890 with their experimental findings to propose that oxidized EC fragments act as proton donors (Brønsted acids) that allow for the formation of HF independent to H2O, following a direct PF6 oxidation process similar to eq 5. Furthermore, previous DFT studies predict that Mn-containing cathodes may be terminated with acidic Mn-OH functional groups upon solvent oxidation,6,91,92 supporting that PF6 hydrolysis (eq 6) or direct oxidation (eq 7) proceeds via an alternative proton source other than water that can lower the activation energy of the reaction because of the more acidic nature of the Mn-OH group. The −OH environments seen in O 1s XPS (Figure S18) could come from Mn-OH surface hydroxyls, supporting this mechanism. Equations 67 are supported by Figure 4, which are either associated with the increase in HPO2F2 or PF5 with voltage (which may be detected as POF3 or HPO2F2 via eq 3 or eqs 8 and 9, see below).

graphic file with name mg2c00060_m006.jpg 6
graphic file with name mg2c00060_m007.jpg 7

Both mechanism (i) and (ii) are possible and may be occurring in tandem, as several electrolyte oxidation pathways generate water39,93 in addition to the HF etching reactions of the LNMO surface28 and explain the cyclic production of the reaction products (e.g., increasing HF concentration seen in Figure 1b, while Karl Fischer titration of the pristine electrolyte yielded < 10 ppm H2O and HF is below the NMR detection limit in the pristine electrolyte). Similarly, HPO2F2 is associated with the formation of HF, as shown in eqs 8 and 9.59,93

graphic file with name mg2c00060_m008.jpg 8
graphic file with name mg2c00060_m009.jpg 9

In any case, we strongly believe that the active material surface plays a role in accelerating the kinetics of electrolyte decomposition and that this finding warrants further investigation.

Influence of Metal Fluoride Deposition on Cathode Film Properties

Four-line probe measurements53 were conducted to measure the sheet conductivity of the porous cathode films to investigate the impact of metal fluorides on the electronic conductivity of the LNMO composite to understand the functionality of CEI (Figure 5). We found that the pristine LNMO film had a sheet conductivity of 523 ± 26.1 mS/cm. Upon adding dilute amounts of MnF2 (5 mg, 4.7% of the total sample composition), the sheet conductivity dropped more than twofold to 200 ± 73 mS/cm. The conductivity dropped to 90.5 ± 20.0 mS/cm and then to 49.6 ± 1.40 mS/cm when 9.7 and 23 mg of MnF2 were added, respectively. LiF was also measured as a reference compound, where we matched the volume of LiF in an LNMO film to the volume of MnF2 in a separate LNMO film for direct comparison (i.e., 6.4 mg of LiF corresponds to the same volume as 9.7 mg of MnF2 based on the densities of LiF (2.64 g/cm3) and MnF2 (3.98 g/cm3)). From this measurement, we find that MnF2 is more electronically insulating than LiF for a given volume (290 ± 22.9 mS/cm for 6.4 mg of LiF and 90.5 ± 20.0 mS/cm for 9.7 mg of MnF2). These measurements show that MnF2 in the CEI is more electronically disruptive to the bulk properties of the electrode than other components, such as LiF, providing insight that is only gained from a quantitative understanding of the compositional heterogeneity of the CEI.

Figure 5.

Figure 5

Four-line probe measurements of different LNMO cathode films (8:1:1 LNMO:C:PVDF), each with dilute amounts (as specified on the x-axis) of LiF or MnF2 added to the slurry prior to casting and drying.

Mechanistic Understanding of Transition Metal Dissolution, CEI Formation, and Impact on Film Properties in LNMO

The combination of time resolved and spatially resolved characterizations used in this study lend insight into the complex degradation reactions at play in LNMO batteries. With time resolved NMR (i.e., the operando solution NMR), we discovered that electrolyte oxidation (e.g., transformation of PF6 into POF3) occurs at the same potential as HF formation (4.68 V). These coupled processes occur on timescales that are much faster than previously thought (min vs h), underscoring the critical role of the active cathode itself in accelerating electrolyte decomposition reactions. XPEEM analyses showed that this electrolyte oxidation resulted in the formation of Mn3+-rich surface layers during charge, which are susceptible to acid-driven disproportionation reactions at high voltage. These findings agree with our operando EPR measurements that detect Mn2+ dissolution as the voltage is raised above 4.7 V during battery charging. The fact that LiPF6 decomposes into HF and POF3 at the same time that Mn dissolution occurs provides new evidence that LiPF6 oxidation plays a major role in transition metal dissolution in LNMO batteries. These observations align with studies showing that dissolution in LNMO is greatly mitigated in aging experiments using LiPF6-free vs LiPF6-containing electrolytes.29,36 Together, these findings indicate that Mn dissolution should, in theory, be mitigated by using Li salts that have higher anodic stability and/or are less prone to generating HF, such as LiFSI.8,94

The understanding that Mn dissolution at high voltages in LNMO is driven by electrolyte-facilitated reduction of surface transition metal centers can help us understand the similarities and differences in transition metal dissolution between spinels and layered oxide cathode materials (e.g., NMC) that also have highly reactive surfaces to facilitate electrolyte decomposition and undergo structural rearrangement during electrochemical cycling.95,96 During NMC charging, electrolyte oxidation at the particle surface results in a concomitant release of oxygen that results in the production of HF37,97,98 as well as a reduction of surface transition metals that forms an oxygen-depleted rock salt layer composed of Ni2+, Mn2+/3+, and Co2+.99101 Thus, in both cathode structures, the surface reconstruction process provides both a source of reduced transition metals and HF needed for metal dissolution, but the extent of dissolution will be highly dependent on the exact composition of the surface layer, which is typically determined by the original stoichiometry of the active material.99,100,102,103

However, in Ni-rich NMCs like NMC811, where only 10% of the transition metal layer contains Mn, it is not entirely clear how Mn dissolution, or transition metal dissolution in general, impacts performance decline. Yet, deposition of transition metals on graphite anodes in NMC cells is correlated with capacity fade in in LIBs.99,104,105 Preferential Mn dissolution over Ni and Co has been observed in NMC cells and often deviates from what might be expected based on the stoichiometry of the material.9799,105,106 This preferential Mn dissolution has been attributed to the higher solubility of Mn2+ in nonaqueous electrolytes compared to Ni and Co10,98,99 as well as the tendency of surface Ni in Ni-rich NMCs to form a stable NiO rocksalt phase.105,107 One study on the comparative aging between spinels and layered cathodes showed that LMO/LNMO exhibit at least 3× more transition metal dissolution compared to NMC-type cathodes during aging,40,44 but how transition metal dissolution proceeds with different cathodes during electrochemical cycling is still not well understood. In addition, if surface hydroxyl groups are responsible for accelerating the production of HF that ultimately etches surface Mn species as proposed in eqs 6 and 7, the crystallographic structure and its termination will be critical for determining transition metal dissolution rates from the cathode. For this reason, more studies focused on the direct comparison of cycling-induced dissolution behavior between spinel-type cathodes (LMO, LNMO) and layered transition metal oxides (e.g., NMC) as well as a more robust understanding of their surface chemistries are needed to better understand the relative magnitude of dissolution in these systems.

Because LNMO and NMC811 both undergo transition metal dissolution driven by HF formation (albeit to different extents), metal fluoride species resulting from these reactions like MnF2, NiF2, and LiF have been observed in the CEI of both materials.7,8,22,23,34,50,67,108110 However, we note that LNMO produces measurably higher quantities of MnF2 compared to NMC811, even when cycled to high voltage (we have used 4.6 V vs Li/Li+ in previous work on NMC811 to study the CEI).37 In recent work, we found that the CEI layer on NMC811 (50 cycles, C/10, 3–4.6 V) is rich in LiF but contained no detectable quantities of MnF2 compared to LiF and PVDF.37 These observations are in stark contrast to the XPEEM results presented here for LNMO (50 cycles, C/10, 3.5–5 V, disassembled at 3.5 V) which show a CEI that contains high quantities of MnF2 in addition to LiF, possibly due to differences in HF formation and Mn dissolution during electrochemical cycling at voltages ≥ 4.7 V.38,39 The reaction energy associated with MnF2 formation is favored over the dissolved transition metal and corresponding fluoride anion in highly acidic environments, which may explain the differences in CEI composition between the layered oxides and the spinels.28

The result of large quantities of MnF2 deposition on the cathode film surface is a decrease in both the electronic conductivity (assessed with four-line probe measurements) as well as the ionic conductivity (which is widely reported in the literature64,71,72,111,112). However, the scale of the MnF2 particles (micron-sized) used in this experiment differs from those expected to form on the CEI in a working system (nanometer-sized), but given the high band gap of this material (between 9 and 10 eV113), we expect that the native CEI will still be electronically insulating. While these conductivity measurements may initially suggest that Mn-rich composites may require higher quantities of conductive carbon, adding much more will decrease the energy density of the cell and possibly lead to other adverse interactions at the cathode/electrolyte interface. As MnF2 has been previously observed before in the CEI for a wide variety of cathode chemistries that perform well in existing Li-ion cells, what these results instead indicate is that the carbon content in composite electrodes must overcompensate for these losses during electrochemical cycling. If we can prevent the formation of MnF2 (and other insulators) in the first place, we may have the opportunity to reduce the carbon content in the cell and improve the overall energy density.

Conclusions

We used operando EPR and NMR to correlate transition metal dissolution with HF generation in the electrolyte during the first charge of an LNMO battery. The operando experiments showed that HF is generated because of electrochemical oxidation of the electrolyte salt and that the amount of HF and dissolved Mn are strongly coupled during the first charge, within the first hour. The rapid HF generation reaction during charging indicates that the active particles play a role catalyzing electrolyte decomposition. The detection of HF and Mn dissolution occurred simultaneously at 4.7 V when the cathode is expected to have oxidized fully to Mn4+, indicating that Mn3+ must be produced during charge to enable the Mn disproportionation reaction to occur. This was confirmed via XPEEM and XPS analyses that detected Mn3+-containing surface phases, supporting that surface Mn4+ is likely reduced during electrolyte oxidation, allowing for transition metal dissolution to occur at high voltages. The simultaneous detection of POF3 and HF at the same voltage with operando solution NMR indicates that PF6 oxidation largely drives transition metal dissolution and indicates that this dissolution can be mitigated by avoiding the formation of HF during cycling. Chemical and spatial characterization of the CEI indicates that these reactions lead to large amounts of MnF2 deposits not only on LNMO particles but also on the carbon additives that decrease the electronic conductivity of the LNMO cathode film and likely alter Li transport to the active particles. However, because MnF2 desorbs from the surface at high voltage, high voltage pulses or holds may be able to disrupt these phases on the cathode surface. Together, these results highlight the importance of developing advanced electrolytes with improved anodic stability to prevent Mn4+ reduction and enable low- or no-Co cathode active materials for Li batteries.

Acknowledgments

This work was supported by National Science Foundation CAREER Award (CBET-2045262). This research used resources (XPEEM/LEEM end station of the ESM beamline) of the Center for Functional Nanomaterials and the National Synchrotron Light Source II, which are the U.S. Department of Energy (DOE) Office of Science facilities at Brookhaven National Laboratory, under Contract No. DE-SC0012704. We also acknowledge resources made available through BNL/LDRD#19-013. J.C.H. is supported by the National Science Foundation Graduate Research Fellowship Program (NSF GRFP # 2021278071). R.M. is supported by the U.S. Department of Defense through the National Defense Science and Engineering Graduate (NDSEG) Fellowship Program.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsmaterialsau.2c00060.

  • Operando NMR/tube cell design diagram; electrochemical cycling data; solution NMR spectra; SEM images; XPS spectra; XPEEM results and associated XAS spectra; and NMR spectral data sets presented in this study can be accessed freely on the Open Science Framework (OSF) (PDF)

Author Contributions

CRediT: Julia C. Hestenes conceptualization (equal), data curation (lead), formal analysis (lead), investigation (lead), methodology (lead), visualization (equal), writing-original draft (lead), writing-review & editing (equal); Jerzy T. Sadowski formal analysis (supporting), investigation (equal), methodology (lead), resources (lead), writing-review & editing (supporting); Richard May formal analysis (supporting), writing-review & editing (supporting); Lauren E Marbella conceptualization (equal), data curation (supporting), formal analysis (supporting), funding acquisition (lead), investigation (supporting), methodology (supporting), project administration (lead), resources (lead), software (lead), supervision (lead), validation (equal), visualization (equal), writing-original draft (supporting), writing-review & editing (equal).

The authors declare no competing financial interest.

Supplementary Material

mg2c00060_si_001.pdf (7.6MB, pdf)

References

  1. Goodenough J. B.; Kim Y. Challenges for Rechargeable Li Batteries. Chem. Mater. 2010, 22, 587–603. 10.1021/cm901452z. [DOI] [Google Scholar]
  2. Liang G.; Peterson V. K.; See K. W.; Guo Z.; Pang W. K. Developing High-Voltage Spinel LiNi0.5Mn1.5O4 cathodes for High-Energy-Density Lithium-Ion Batteries: Current Achievements and Future Prospects. J. Mater. Chem. A 2020, 8, 15373–15398. 10.1039/d0ta02812f. [DOI] [Google Scholar]
  3. Jung H. G.; Jang M. W.; Hassoun J.; Sun Y. K.; Scrosati B. A High-Rate Long-Life Li4Ti5O12/Li[Ni0.45Co0.1Mn1.45]O4 Lithium-Ion Battery. Nat. Commun. 2011, 2, 516. 10.1038/ncomms1527. [DOI] [PubMed] [Google Scholar]
  4. Kim J. H.; Myung S. T.; Yoon C. S.; Kang S. G.; Sun Y. K. Comparative Study of LiNi0.5Mn1.5O4-δ and LiNi0.5Mn1.5O4 Cathodes Having Two Crystallographic Structures: Fd3m and P4332. Chem. Mater. 2004, 16, 906–914. 10.1021/cm035050s. [DOI] [Google Scholar]
  5. Gourley S. W. D.; Or T.; Chen Z. Breaking Free from Cobalt Reliance in Lithium-Ion Batteries. iScience 2020, 23, 101505 10.1016/j.isci.2020.101505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Leung K.; Rosy; Noked M. Anodic Decomposition of Surface Films on High Voltage Spinel Surfaces - Density Function Theory and Experimental Study. J. Chem. Phys. 2019, 151, 234713. 10.1063/1.5131447. [DOI] [PubMed] [Google Scholar]
  7. Pieczonka N. P. W.; Liu Z.; Lu P.; Olson K. L.; Moote J.; Powell B. R.; Kim J. H. Understanding Transition-Metal Dissolution Behavior in LiNi0.5Mn1.5O4 High-Voltage Spinel for Lithium Ion Batteries. J. Phys. Chem. C 2013, 117, 15947–15957. 10.1021/jp405158m. [DOI] [Google Scholar]
  8. Alvarado J.; Schroeder M. A.; Zhang M.; Borodin O.; Gobrogge E.; Olguin M.; Ding M. S.; Gobet M.; Greenbaum S.; Meng Y. S.; Xu K. A Carbonate-Free, Sulfone-Based Electrolyte for High-Voltage Li-Ion Batteries. Mater. Today 2018, 21, 341–353. 10.1016/j.mattod.2018.02.005. [DOI] [Google Scholar]
  9. Liu T.; Dai A.; Lu J.; Yuan Y.; Xiao Y.; Yu L.; Li M.; Gim J.; Ma L.; Liu J.; Zhan C.; Li L.; Zheng J.; Ren Y.; Wu T.; Shahbazian-Yassar R.; Wen J.; Pan F.; Amine K. Correlation between Manganese Dissolution and Dynamic Phase Stability in Spinel-Based Lithium-Ion Battery. Nat. Commun. 2019, 10, 4721. 10.1038/s41467-019-12626-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Lin M.; Ben L.; Sun Y.; Wang H.; Yang Z.; Gu L.; Yu X.; Yang X. Q.; Zhao H.; Yu R.; Armand M.; Huang X. Insight into the Atomic Structure of High-Voltage Spinel LiNi0.5Mn1.5O4 Cathode Material in the First Cycle. Chem. Mater. 2015, 27, 292–303. 10.1021/cm503972a. [DOI] [Google Scholar]
  11. Aurbach D.; Markovsky B.; Talyossef Y.; Salitra G.; Kim H. J.; Choi S. Studies of Cycling Behavior, Ageing, and Interfacial Reactions of LiNi0.5Mn1.5O4 and Carbon Electrodes for Lithium-Ion 5-V Cells. J. Power Sources 2006, 162, 780–789. 10.1016/j.jpowsour.2005.07.009. [DOI] [Google Scholar]
  12. Wang C.; Xing L.; Vatamanu J.; Chen Z.; Lan G.; Li W.; Xu K. Overlooked Electrolyte Destabilization by Manganese (II) in Lithium-Ion Batteries. Nat. Commun. 2019, 10, 3423. 10.1038/s41467-019-11439-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Mahootcheianasl N.; Kim J. H.; Pieczonka N. P. W.; Liu Z.; Kim Y. Multilayer Electrolyte Cell: A New Tool for Identifying Electrochemical Performances of High Voltage Cathode Materials. Electrochem. Commun. 2013, 32, 1–4. 10.1016/j.elecom.2013.03.031. [DOI] [Google Scholar]
  14. Fang X.; Ding N.; Feng X. Y.; Lu Y.; Chen C. H. Study of LiNi0.5Mn1.5O4 Synthesized via a Chloride-Ammonia Co-Precipitation Method: Electrochemical Performance, Diffusion Coefficient and Capacity Loss Mechanism. Electrochim. Acta 2009, 54, 7471–7475. 10.1016/j.electacta.2009.07.084. [DOI] [Google Scholar]
  15. Kim J. H.; Pieczonka N. P. W.; Li Z.; Wu Y.; Harris S.; Powell B. R. Understanding the Capacity Fading Mechanism in LiNi0.5Mn1.5O4/Graphite Li-Ion Batteries. Electrochim. Acta 2013, 90, 556–562. 10.1016/j.electacta.2012.12.069. [DOI] [Google Scholar]
  16. Aktekin B.; Lacey M. J.; Nordh T.; Younesi R.; Tengstedt C.; Zipprich W.; Brandell D.; Edström K. Understanding the Capacity Loss in LiNi0.5Mn1.5O4-Li4Ti5O12 Lithium-Ion Cells at Ambient and Elevated Temperatures. J. Phys. Chem. C 2018, 122, 11234–11248. 10.1021/acs.jpcc.8b02204. [DOI] [Google Scholar]
  17. Lu D.; Xu M.; Zhou L.; Garsuch A.; Lucht B. L. Failure Mechanism of Graphite/LiNi0.5Mn1.5O4 Cells at High Voltage and Elevated Temperature. J. Electrochem. Soc. 2013, 160, A3138–A3143. 10.1149/2.022305jes. [DOI] [Google Scholar]
  18. Shkrob I. A.; Kropf A. J.; Marin T. W.; Li Y.; Poluektov O. G.; Niklas J.; Abraham D. P. Manganese in Graphite Anode and Capacity Fade in Li Ion Batteries. J. Phys. Chem. C 2014, 118, 24335–24348. 10.1021/jp507833u. [DOI] [Google Scholar]
  19. Lee Y. K.; Park J.; Lu W. A Comprehensive Study of Manganese Deposition and Side Reactions in Li-Ion Battery Electrodes. J. Electrochem. Soc. 2017, 164, A2812–A2822. 10.1149/2.1851712jes. [DOI] [Google Scholar]
  20. Harris O. C.; Lin Y.; Qi Y.; Leung K.; Tang M. H. How Transition Metals Enable Electron Transfer through the SEI: Part I. Experiments and Butler-Volmer Modeling. J. Electrochem. Soc. 2020, 167, 013502 10.1149/2.0022001jes. [DOI] [Google Scholar]
  21. Harris O. C.; Leung K.; Tang M. H. How Transition Metals Enable Electron Transfer through the SEI: Part II. Redox-Cycling Mechanism Model and Experiment. J. Electrochem. Soc. 2020, 167, 013503 10.1149/2.0032001jes. [DOI] [Google Scholar]
  22. Scipioni R.; Isheim D.; Barnett S. A. Revealing the Complex Layered-Mosaic Structure of the Cathode Electrolyte Interphase in Li-Ion Batteries. Appl. Mater. Today 2020, 20, 100748 10.1016/j.apmt.2020.100748. [DOI] [Google Scholar]
  23. Li Q.; Wang Y.; Wang X.; Sun X.; Zhang J. N.; Yu X.; Li H. Investigations on the Fundamental Process of Cathode Electrolyte Interphase Formation and Evolution of High-Voltage Cathodes. ACS Appl. Mater. Interfaces 2020, 12, 2319–2326. 10.1021/acsami.9b16727. [DOI] [PubMed] [Google Scholar]
  24. Yang L.; Ravdel B.; Lucht B. L. Electrolyte Reactions with the Surface of High Voltage LiNi0.5Mn1.5O4 Cathodes for Lithium-Ion Batteries. Electrochem. Solid-State Lett. 2010, 13, A95. 10.1149/1.3428515. [DOI] [Google Scholar]
  25. Demeaux J.; Caillon-Caravanier M.; Galiano H.; Lemordant D.; Claude-Montigny B. LiNi 0.4Mn1.6O4/Electrolyte and Carbon Black/Electrolyte High Voltage Interfaces: To Evidence the Chemical and Electronic Contributions of the Solvent on the Cathode-Electrolyte Interface Formation. J. Electrochem. Soc. 2012, 159, A1880–A1890. 10.1149/2.052211jes. [DOI] [Google Scholar]
  26. Norberg N. S.; Lux S. F.; Kostecki R. Interfacial Side-Reactions at a LiNi0.5Mn1.5O4 Electrode in Organic Carbonate-Based Electrolytes. Electrochem. Commun. 2013, 34, 29–32. 10.1016/j.elecom.2013.04.007. [DOI] [Google Scholar]
  27. Erickson E. M.; Li W.; Dolocan A.; Manthiram A. Insights into the Cathode-Electrolyte Interphases of High-Energy-Density Cathodes in Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2020, 12, 16451–16461. 10.1021/acsami.0c00900. [DOI] [PubMed] [Google Scholar]
  28. Benedek R.; Thackeray M. M. Reaction Energy for LiMn2O4 Spinel Dissolution in Acid. Electrochem. Solid-State Lett. 2006, 9, A265. 10.1149/1.2188071. [DOI] [Google Scholar]
  29. Choi N. S.; Yeon J. T.; Lee Y. W.; Han J. G.; Lee K. T.; Kim S. S. Degradation of Spinel Lithium Manganese Oxides by Low Oxidation Durability of LiPF6-Based Electrolyte at 60 °C. Solid State Ionics 2012, 219, 41–48. 10.1016/j.ssi.2012.05.012. [DOI] [Google Scholar]
  30. Xiao J.; Chen X.; Sushko P. V; Sushko M. L.; Kovarik L.; Feng J.; Deng Z.; Zheng J.; Graff G. L.; Nie Z.; Choi D.; Liu J.; Zhang J. G.; Whittingham M. S. High-Performance LiNi0.5Mn1.5O4 Spinel Controlled by Mn3+ Concentration and Site Disorder. Adv. Mater. 2012, 24, 2109–2116. 10.1002/adma.201104767. [DOI] [PubMed] [Google Scholar]
  31. Cabana J.; Casas-Cabanas M.; Omenya F. O.; Chernova N. A.; Zeng D.; Whittingham M. S.; Grey C. P. Composition-Structure Relationships in the Li-Ion Battery Electrode Material LiNi0.5Mn1.5O4. Chem. Mater. 2012, 24, 2952–2964. 10.1021/cm301148d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Shaju K. M.; Bruce P. G.; Bruce P. G. Nano- LiNi0.5Mn1.5O4 Spinel: A High Power Electrode for Li-Ion Batteries. Dalton Trans. 2008, 40, 5471–5475. 10.1039/b806662k. [DOI] [PubMed] [Google Scholar]
  33. Mao J.; Dai K.; Zhai Y. Electrochemical Studies of Spinel LiNi0.5Mn1.5O4 Cathodes with Different Particle Morphologies. Electrochim. Acta 2012, 63, 381–390. 10.1016/j.electacta.2011.12.129. [DOI] [Google Scholar]
  34. Qiao R.; Wang Y.; Olalde-Velasco P.; Li H.; Hu Y. S.; Yang W. Direct Evidence of Gradient Mn(II) Evolution at Charged States in LiNi0.5Mn1.5O4 Electrodes with Capacity Fading. J. Power Sources 2015, 273, 1120–1126. 10.1016/j.jpowsour.2014.10.013. [DOI] [Google Scholar]
  35. Wang L.-F.; Ou C.-C.; Striebel K. A.; Chen J.-S. Study of Mn Dissolution from LiNi0.5Mn1.5O4 Spinel Electrodes Using Rotating Ring-Disk Collection Experiments. J. Electrochem. Soc. 2003, 150, A905. 10.1149/1.1577543. [DOI] [Google Scholar]
  36. Aoshima T.; Okahara K.; Kiyohara C.; Shizuka K. Mechanisms of Manganese Spinels Dissolution and Capacity Fade at High Temperature. J. Power Sources 2001, 97-98, 377–380. 10.1016/S0378-7753(01)00551-1. [DOI] [Google Scholar]
  37. Hestenes J. C.; May R.; Sadowski J. T.; Munich N.; Marbella L. E. Resolving Chemical and Spatial Heterogeneities at Complex Electrochemical Interfaces in Li-Ion Batteries. Chem. Mater. 2022, 34, 232–243. 10.1021/acs.chemmater.1c03185. [DOI] [Google Scholar]
  38. Yoon T.; Soon J.; Lee T. J.; Ryu J. H.; Oh S. M. Dissolution of Cathode–Electrolyte Interphase Deposited on LiNi0.5Mn1.5O4 for Lithium-Ion Batteries. J. Power Sources 2021, 503, 230051 10.1016/j.jpowsour.2021.230051. [DOI] [Google Scholar]
  39. Metzger M.; Sicklinger J.; Haering D.; Kavakli C.; Stinner C.; Marino C.; Gasteiger H. A. Carbon Coating Stability on High-Voltage Cathode Materials in H2O-Free and H2O-Containing Electrolyte. J. Electrochem. Soc. 2015, 162, A1227–A1235. 10.1149/2.0461507jes. [DOI] [Google Scholar]
  40. Choi W.; Manthiram A. Comparison of Metal Ion Dissolutions from Lithium Ion Battery Cathodes. J. Electrochem. Soc. 2006, 153, A1760. 10.1149/1.2219710. [DOI] [Google Scholar]
  41. Zhan C.; Lu J.; Jeremy Kropf A.; Wu T.; Jansen A. N.; Sun Y. K.; Qiu X.; Amine K. Mn(II) Deposition on Anodes and Its Effects on Capacity Fade in Spinel Lithium Manganate-Carbon Systems. Nat. Commun. 2013, 4, 2437. 10.1038/ncomms3437. [DOI] [PubMed] [Google Scholar]
  42. Tsunekawa H.; Tanimoto A. S.; Marubayashi R.; Fujita M.; Kifune K.; Sano M. Capacity Fading of Graphite Electrodes Due to the Deposition of Manganese Ions on Them in Li-Ion Batteries. J. Electrochem. Soc. 2002, 149, A1326. 10.1149/1.1502686. [DOI] [Google Scholar]
  43. Kuppan S.; Duncan H.; Chen G. Controlling Side Reactions and Self-Discharge in High-Voltage Spinel Cathodes: The Critical Role of Surface Crystallographic Facets. Phys. Chem. Chem. Phys. 2015, 17, 26471–26481. 10.1039/c5cp04899k. [DOI] [PubMed] [Google Scholar]
  44. Su C.-C.; He M.; Amine R.; Chen Z.; Yu Z.; Rojas T.; Cheng L.; Ngo A. T.; Amine K. Unveiling Decaying Mechanism through Quantitative Structure-Activity Relationship in Electrolytes for Lithium-Ion Batteries. Nano Energy 2021, 83, 105843 10.1016/j.nanoen.2021.105843. [DOI] [Google Scholar]
  45. Quilty C. D.; Wheeler G. P.; Wang L.; McCarthy A. H.; Yan S.; Tallman K. R.; Dunkin M. R.; Tong X.; Ehrlich S.; Ma L.; Takeuchi K. J.; Takeuchi E. S.; Bock D. C.; Marschilok A. C. Impact of Charge Voltage on Factors Influencing Capacity Fade in Layered NMC622: Multimodal X-Ray and Electrochemical Characterization. ACS Appl. Mater. Interfaces 2021, 13, 50920–50935. 10.1021/acsami.1c14272. [DOI] [PubMed] [Google Scholar]
  46. Wandt J.; Freiberg A.; Thomas R.; Gorlin Y.; Siebel A.; Jung R.; Gasteiger H. A.; Tromp M. Transition Metal Dissolution and Deposition in Li-Ion Batteries Investigated by Operando X-Ray Absorption Spectroscopy. J. Mater. Chem. A 2016, 4, 18300–18305. 10.1039/C6TA08865A. [DOI] [Google Scholar]
  47. Helmus J. J.; Jaroniec C. P. Nmrglue: An Open Source Python Package for the Analysis of Multidimensional NMR Data. J. Biomol. NMR 2013, 55, 355–367. 10.1007/S10858-013-9718-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Lee E. S.; Nam K. W.; Hu E.; Manthiram A. Influence of Cation Ordering and Lattice Distortion on the Charge-Discharge Behavior of LiMn1.5Ni0.5O4 Spinel between 5.0 and 2.0 V. Chem. Mater. 2012, 24, 3610–3620. 10.1021/cm3020836. [DOI] [Google Scholar]
  49. Song J.; Shin D. W.; Lu Y.; Amos C. D.; Manthiram A.; Goodenough J. B. Role of Oxygen Vacancies on the Performance of Li[Ni0.5-xMn1.5+x]O4 (x = 0, 0.05, and 0.08) Spinel Cathodes for Lithium-Ion Batteries. Chem. Mater. 2012, 24, 3101–3109. 10.1021/cm301825h. [DOI] [Google Scholar]
  50. Qiao R.; Wray L. A.; Kim J. H.; Pieczonka N. P. W.; Harris S. J.; Yang W. Direct Experimental Probe of the Ni(II)/Ni(III)/Ni(IV) Redox Evolution in LiNi0.5Mn1.5O4 Electrodes. J. Phys. Chem. C 2015, 119, 27228–27233. 10.1021/acs.jpcc.5b07479. [DOI] [Google Scholar]
  51. Mirolo M.; Leanza D.; Höltschi L.; Jordy C.; Pelé V.; Novák P.; El Kazzi M.; Vaz C. A. F. Post Mortem and Operando XPEEM: A Surface-Sensitive Tool for Studying Single Particles in Li-Ion Battery Composite Electrodes. Anal. Chem. 2020, 92, 3023–3031. 10.1021/acs.analchem.9b04124. [DOI] [PubMed] [Google Scholar]
  52. Zhuo Z.; Lu P.; Delacourt C.; Qiao R.; Xu K.; Pan F.; Harris S. J.; Yang W. Breathing and Oscillating Growth of Solid-Electrolyte-Interphase upon Electrochemical Cycling. Chem. Commun. 2018, 54, 814–817. 10.1039/c7cc07082a. [DOI] [PubMed] [Google Scholar]
  53. Peterson S. W.; Wheeler D. R. Direct Measurements of Effective Electronic Transport in Porous Li-Ion Electrodes. J. Electrochem. Soc. 2014, 161, A2175–A2181. 10.1149/2.0661414jes. [DOI] [Google Scholar]
  54. Wiemers-Meyer S.; Winter M.; Nowak S. Mechanistic Insights into Lithium Ion Battery Electrolyte Degradation-a Quantitative NMR Study. Phys. Chem. Chem. Phys. 2016, 18, 26595–26601. 10.1039/c6cp05276b. [DOI] [PubMed] [Google Scholar]
  55. Rinkel B.; Hall D. S.; Temprano I.; Grey C. P. Electrolyte Oxidation Pathways in Lithium-Ion Batteries. J. Am. Chem. Soc. 2020, 142, 15058–15074. 10.1021/jacs.0c06363. [DOI] [PubMed] [Google Scholar]
  56. Plakhotnyk A. V.; Ernst L.; Schmutzler R. Hydrolysis in the System LiPF6 - Propylene Carbonate - Dimethyl Carbonate - H2O. J. Fluorine Chem. 2005, 126, 27–31. 10.1016/j.jfluchem.2004.09.027. [DOI] [Google Scholar]
  57. Banerjee A.; Shilina Y.; Ziv B.; Ziegelbauer J. M.; Luski S.; Aurbach D.; Halalay I. C. On the Oxidation State of Manganese Ions in Li-Ion Battery Electrolyte Solutions. J. Am. Chem. Soc. 2017, 139, 1738–1741. 10.1021/jacs.6b10781. [DOI] [PubMed] [Google Scholar]
  58. Nguyen H.; Clément R. J. Rechargeable Batteries from the Perspective of the Electron Spin. ACS Energy Lett. 2020, 5, 3848–3859. 10.1021/acsenergylett.0c02074. [DOI] [Google Scholar]
  59. Stich M.; Göttlinger M.; Kurniawan M.; Schmidt U.; Bund A. Hydrolysis of LiPF6 in Carbonate-Based Electrolytes for Lithium-Ion Batteries and in Aqueous Media. J. Phys. Chem. C 2018, 122, 8836–8842. 10.1021/acs.jpcc.8b02080. [DOI] [Google Scholar]
  60. Rodrigues M. T. F.; Liao C.; Kalaga K.; Shkrob I. A.; Abraham D. P. Dehydration Rather Than HF Capture Explains Performance Improvements of Li-Ion Cells by Ceramic Nanoparticles. ACS Appl. Energy Mater. 2019, 2, 5380–5385. 10.1021/acsaem.9b00976. [DOI] [Google Scholar]
  61. Gilbert B.; Frazer B. H.; Belz A.; Conrad P. G.; Nealson K. H.; Haskel D.; Lang J. C.; Srajer G.; De Stasio G. Multiple Scattering Calculations of Bonding and X-Ray Absorption Spectroscopy of Manganese Oxides. J. Phys. Chem. A 2003, 107, 2839–2847. 10.1021/jp021493s. [DOI] [Google Scholar]
  62. Zhou J.; Hong D.; Wang J.; Hu Y.; Xie X.; Fang H. Electronic Structure Variation of the Surface and Bulk of a LiNi0.5Mn1.5O4 Cathode as a Function of State of Charge: X-Ray Absorption Spectroscopic Study. Phys. Chem. Chem. Phys. 2014, 16, 13838–13842. 10.1039/c4cp01436g. [DOI] [PubMed] [Google Scholar]
  63. Talyosef Y.; Markovsky B.; Salitra G.; Aurbach D.; Kim H. J.; Choi S. The Study of LiNi0.5Mn1.5O4 5-V Cathodes for Li-Ion Batteries. J. Power Sources 2005, 146, 664–669. 10.1016/j.jpowsour.2005.03.064. [DOI] [Google Scholar]
  64. Kim D.; Park S.; Chae O. B.; Ryu J. H.; Kim Y.-U.; Yin R.-Z.; Oh S. M. Re-Deposition of Manganese Species on Spinel LiMn2O4 Electrode after Mn Dissolution. J. Electrochem. Soc. 2012, 159, A193–A197. 10.1149/2.003203jes. [DOI] [Google Scholar]
  65. Cherkashinin G.; Nikolowski K.; Ehrenberg H.; Jacke S.; Dimesso L.; Jaegermann W. The Stability of the SEI Layer, Surface Composition and the Oxidation State of Transition Metals at the Electrolyte-Cathode Interface Impacted by the Electrochemical Cycling: X-Ray Photoelectron Spectroscopy Investigation. Phys. Chem. Chem. Phys. 2012, 14, 12321–12331. 10.1039/c2cp41134b. [DOI] [PubMed] [Google Scholar]
  66. Yoon T.; Lee T.; Soon J.; Jeong H.; Jurng S.; Ryu J. H.; Oh S. M. The Investigation of Electrolyte Oxidation and Film Deposition Characteristics at High Potentials in a Carbonate-Based Electrolyte Using Pt Electrode. J. Electrochem. Soc. 2018, 165, A1095–A1098. 10.1149/2.0931805jes. [DOI] [Google Scholar]
  67. Yoon T.; Kim D.; Park K. H.; Park H.; Jurng S.; Jang J.; Ryu J. H.; Kim J. J.; Oh S. M. Compositional Change of Surface Film Deposited on LiNi0.5Mn1.5O4 Positive Electrode. J. Electrochem. Soc. 2014, 161, A519–A523. 10.1149/2.019404jes. [DOI] [Google Scholar]
  68. Fang S.; Jackson D.; Dreibelbis M. L.; Kuech T. F.; Hamers R. J. Anode-Originated SEI Migration Contributes to Formation of Cathode-Electrolyte Interphase Layer. J. Power Sources 2018, 373, 184–192. 10.1016/j.jpowsour.2017.09.050. [DOI] [Google Scholar]
  69. Hestenes J. C.; Ells A. W.; Navarro Goldaraz M.; Sergeyev I. V.; Itin B.; Marbella L. E. Reversible Deposition and Stripping of the Cathode Electrolyte Interphase on Li2RuO3. Front. Chem. 2020, 8, 681. 10.3389/fchem.2020.00681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Qiao R.; Chin T.; Harris S. J.; Yan S.; Yang W. Spectroscopic Fingerprints of Valence and Spin States in Manganese Oxides and Fluorides. Curr. Appl. Phys. 2013, 13, 544–548. 10.1016/j.cap.2012.09.017. [DOI] [Google Scholar]
  71. De Taeye L. L.; Vereecken P. M. Detrimental MnPO4F and MnF2 formation on LiMn2O4 in the 3 V Region. J. Mater. Chem. A 2021, 9, 23256–23268. 10.1039/d1ta06227a. [DOI] [Google Scholar]
  72. Quinlan F. T.; Sano K.; Willey T.; Vidu R.; Tasaki K.; Stroeve P. Surface Characterization of the Spinel LixMn2O4 Cathode before and after Storage at Elevated Temperatures. Chem. Mater. 2001, 13, 4207–4212. 10.1021/cm010335v. [DOI] [Google Scholar]
  73. Chung K. Y.; Yoon W. S.; Kim K. B.; Cho B. W.; Yang X. Q. Formation of an SEI on a LiMn2O4 Cathode during Room Temperature Charge-Discharge Cycling Studied by Soft X-Ray Absorption Spectroscopy at the Fluorine K-Edge. J. Appl. Electrochem. 2011, 41, 1295–1299. 10.1007/s10800-011-0344-6. [DOI] [Google Scholar]
  74. Qiao R.; Lucas I. T.; Karim A.; Syzdek J.; Liu X.; Chen W.; Persson K.; Kostecki R.; Yang W. Distinct Solid-Electrolyte-Interphases on Sn (100) and (001) Electrodes Studied by Soft X-Ray Spectroscopy. Adv. Mater. Interfaces 2014, 1, 1300115 10.1002/admi.201300115. [DOI] [Google Scholar]
  75. Zhang J. N.; Li Q.; Wang Y.; Zheng J.; Yu X.; Li H. Dynamic Evolution of Cathode Electrolyte Interphase (CEI) on High Voltage LiCoO2 Cathode and Its Interaction with Li Anode. Energy Storage Mater. 2018, 14, 1–7. 10.1016/j.ensm.2018.02.016. [DOI] [Google Scholar]
  76. Verma P.; Maire P.; Novák P. A Review of the Features and Analyses of the Solid Electrolyte Interphase in Li-Ion Batteries. Electrochim. Acta 2010, 55, 6332–6341. 10.1016/j.electacta.2010.05.072. [DOI] [Google Scholar]
  77. Luo H. M.; Sun Y. X.; Zhang F. B.; Zhang D. Y.; Zhang Y. F. Y. J. Q. Preparation and Electrochemical Applications of Spherical Maltose-Based Templated Carbon/MnOx Composite Materials. Electrochim. Acta 2014, 148, 228–236. 10.1016/j.electacta.2014.10.044. [DOI] [Google Scholar]
  78. Parimalam B. S.; MacIntosh A. D.; Kadam R.; Lucht B. L. Decomposition Reactions of Anode Solid Electrolyte Interphase (SEI) Components with LiPF6. J. Phys. Chem. C 2017, 121, 22733–22738. 10.1021/acs.jpcc.7b08433. [DOI] [Google Scholar]
  79. Yu Y.; Karayaylali P.; Katayama Y.; Giordano L.; Gauthier M.; Maglia F.; Jung R.; Lund I.; Shao-Horn Y. Coupled LiPF6 Decomposition and Carbonate Dehydrogenation Enhanced by Highly Covalent Metal Oxides in High-Energy Li-Ion Batteries. J. Phys. Chem. C 2018, 122, 27368–27382. 10.1021/acs.jpcc.8b07848. [DOI] [Google Scholar]
  80. Friedman A. K.; Shi W.; Losovyj Y.; Siedle A. R.; Baker L. A. Mapping Microscale Chemical Heterogeneity in Nafion Membranes with X-Ray Photoelectron Spectroscopy. J. Electrochem. Soc. 2018, 165, H733–H741. 10.1149/2.0771811jes. [DOI] [Google Scholar]
  81. Wilken S.; Treskow M.; Scheers J.; Johansson P.; Jacobsson P. Initial Stages of Thermal Decomposition of LiPF6-Based Lithium Ion Battery Electrolytes by Detailed Raman and NMR Spectroscopy. RSC Adv. 2013, 3, 16359–16364. 10.1039/c3ra42611d. [DOI] [Google Scholar]
  82. Freiberg A. T. S.; Sicklinger J.; Solchenbach S.; Gasteiger H. A. Li2CO3 Decomposition in Li-Ion Batteries Induced by the Electrochemical Oxidation of the Electrolyte and of Electrolyte Impurities. Electrochim. Acta 2020, 346, 136271 10.1016/j.electacta.2020.136271. [DOI] [Google Scholar]
  83. Campion C. L.; Li W.; Lucht B. L. Thermal Decomposition of LiPF6-Based Electrolytes for Lithium-Ion Batteries. J. Electrochem. Soc. 2005, 152, A2327. 10.1149/1.2083267. [DOI] [Google Scholar]
  84. Tasaki K.; Kanda K.; Nakamura S.; Ue M. Decomposition of LiPF6 and Stability of PF5 in Li-Ion Battery Electrolytes. J. Electrochem. Soc. 2003, 150, A1628. 10.1149/1.1622406. [DOI] [Google Scholar]
  85. Liu M.; Vatamanu J.; Chen X.; Xing L.; Xu K.; Li W. Hydrolysis of LiPF6-Containing Electrolyte at High Voltage. ACS Energy Lett. 2021, 6, 2096–2102. 10.1021/acsenergylett.1c00707. [DOI] [Google Scholar]
  86. Solchenbach S.; Metzger M.; Egawa M.; Beyer H.; Gasteiger H. A. Quantification of PF5 and POF3 from Side Reactions of LiPF6 in Li-Ion Batteries. J. Electrochem. Soc. 2018, 165, A3022. 10.1149/2.0481813JES. [DOI] [Google Scholar]
  87. Dumaz P.; Rossignol C.; Mantoux A.; Sergent N.; Bouchet R. Kinetics Analysis of the Electro-Catalyzed Degradation of High Potential LiNi0.5Mn1.5O4 Active Materials. J. Power Sources 2020, 469, 228337 10.1016/j.jpowsour.2020.228337. [DOI] [Google Scholar]
  88. Li T.; Xing L.; Li W.; Wang Y.; Xu M.; Gu F.; Hu S. How Does Lithium Salt Anion Affect Oxidation Decomposition Reaction of Ethylene Carbonate: A Density Functional Theory Study. J. Power Sources 2013, 244, 668–674. 10.1016/j.jpowsour.2012.12.062. [DOI] [Google Scholar]
  89. Borodin O.; Behl W.; Jow T. R. Oxidative Stability and Initial Decomposition Reactions of Carbonate, Sulfone, and Alkyl Phosphate-Based Electrolytes. J. Phys. Chem. C 2013, 117, 8661–8682. 10.1021/jp400527c. [DOI] [Google Scholar]
  90. Xing L.; Borodin O.; Smith G. D.; Li W. Density Functional Theory Study of the Role of Anions on the Oxidative Decomposition Reaction of Propylene Carbonate. J. Phys. Chem. A 2011, 115, 13896–13905. 10.1021/jp206153n. [DOI] [PubMed] [Google Scholar]
  91. Østergaard T. M.; Giordano L.; Castelli I. E.; Maglia F.; Antonopoulos B. K.; Shao-Horn Y.; Rossmeisl J. Oxidation of Ethylene Carbonate on Li Metal Oxide Surfaces. J. Phys. Chem. C 2018, 122, 10442–10449. 10.1021/acs.jpcc.8b01713. [DOI] [Google Scholar]
  92. Okuno Y.; Ushirogata K.; Sodeyama K.; Shukri G.; Tateyama Y. Structures, Electronic States, and Reactions at Interfaces between LiNi0.5Mn1.5O4 Cathode and Ethylene Carbonate Electrolyte: A First-Principles Study. J. Phys. Chem. C 2019, 123, 2267–2277. 10.1021/acs.jpcc.8b10625. [DOI] [Google Scholar]
  93. Rinkel B. L. D.; Vivek J. P.; Garcia-Araez N.; Grey C. P. Two Electrolyte Decomposition Pathways at Nickel-Rich Cathode Surfaces in Lithium-Ion Batteries. Energy Environ. Sci. 2022, 15, 3416–3438. 10.1039/D1EE04053G. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Milien M. S.; Beyer H.; Beichel W.; Klose P.; Gasteiger H. A.; Lucht B. L.; Krossing I. Lithium Bis(2,2,2-Trifluoroethyl)Phosphate Li[O2P(OCH2CF3)2]: A High Voltage Additive for LNMO/Graphite Cells. J. Electrochem. Soc. 2018, 165, A2569–A2576. 10.1149/2.0541811jes. [DOI] [Google Scholar]
  95. Giordano L.; Karayaylali P.; Yu Y.; Katayama Y.; Maglia F.; Lux S.; Shao-Horn Y. Chemical Reactivity Descriptor for the Oxide-Electrolyte Interface in Li-Ion Batteries. J. Phys. Chem. Lett. 2017, 8, 3881–3887. 10.1021/acs.jpclett.7b01655. [DOI] [PubMed] [Google Scholar]
  96. Xu C.; Märker K.; Lee J.; Mahadevegowda A.; Reeves P. J.; Day S. J.; Groh M. F.; Emge S. P.; Ducati C.; Layla Mehdi B.; Tang C. C.; Grey C. P. Bulk Fatigue Induced by Surface Reconstruction in Layered Ni-Rich Cathodes for Li-Ion Batteries. Nat. Mater. 2021, 20, 84–92. 10.1038/s41563-020-0767-8. [DOI] [PubMed] [Google Scholar]
  97. Dose W. M.; Temprano I.; Allen J. P.; Björklund E.; O’Keefe C. A.; Li W.; Mehdi B. L.; Weatherup R. S.; De Volder M. F. L.; Grey C. P. Electrolyte Reactivity at the Charged Ni-Rich Cathode Interface and Degradation in Li-Ion Batteries. ACS Appl. Mater. Interfaces 2022, 14, 13206–13222. 10.1021/acsami.1c22812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Gallus D. R.; Schmitz R.; Wagner R.; Hoffmann B.; Nowak S.; Cekic-Laskovic I.; Schmitz R. W.; Winter M. The Influence of Different Conducting Salts on the Metal Dissolution and Capacity Fading of NCM Cathode Material. Electrochim. Acta 2014, 134, 393–398. 10.1016/j.electacta.2014.04.091. [DOI] [Google Scholar]
  99. Björklund E.; Xu C.; Dose W. M.; Sole C. G.; Thakur P. K.; Lee T. L.; De Volder M. F. L.; Grey C. P.; Weatherup R. S. Cycle-Induced Interfacial Degradation and Transition-Metal Cross-Over in LiNi0.8Mn0.1Co0.1O2-Graphite Cells. Chem. Mater. 2022, 34, 2034–2048. 10.1021/acs.chemmater.1c02722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Liu H.; Bugnet M.; Tessaro M. Z.; Harris K. J.; Dunham M. J. R.; Jiang M.; Goward G. R.; Botton G. A. Spatially Resolved Surface Valence Gradient and Structural Transformation of Lithium Transition Metal Oxides in Lithium-Ion Batteries. Phys. Chem. Chem. Phys. 2016, 18, 29064–29075. 10.1039/c6cp05262b. [DOI] [PubMed] [Google Scholar]
  101. Ruff Z.; Xu C.; Grey C. P. Transition Metal Dissolution and Degradation in NMC811-Graphite Electrochemical Cells. J. Electrochem. Soc. 2021, 168, 060518 10.1149/1945-7111/ac0359. [DOI] [Google Scholar]
  102. Hwang S.; Kim S. Y.; Chung K. Y.; Stach E. A.; Kim S. M.; Chang W. Determination of the Mechanism and Extent of Surface Degradation in Ni-Based Cathode Materials after Repeated Electrochemical Cycling. APL Mater. 2016, 4, 096105 10.1063/1.4963723. [DOI] [Google Scholar]
  103. Hwang S.; Jo E.; Chung K. Y.; Hwang K. S.; Kim S. M.; Chang W. Structural Evolution of LixNiyMnzCo1-y-zO2 Cathode Materials during High-Rate Charge and Discharge. J. Phys. Chem. Lett. 2017, 8, 5758–5763. 10.1021/acs.jpclett.7b02579. [DOI] [PubMed] [Google Scholar]
  104. Jung R.; Linsenmann F.; Thomas R.; Wandt J.; Solchenbach S.; Maglia F.; Stinner C.; Tromp M.; Gasteiger H. A. Nickel, Manganese, and Cobalt Dissolution from Ni-Rich NMC and Their Effects on NMC622-Graphite Cells. J. Electrochem. Soc. 2019, 166, A378–A389. 10.1149/2.1151902jes. [DOI] [Google Scholar]
  105. Li W.; Liu X.; Celio H.; Smith P.; Dolocan A.; Chi M.; Manthiram A. Mn versus Al in Layered Oxide Cathodes in Lithium-Ion Batteries: A Comprehensive Evaluation on Long-Term Cyclability. Adv. Energy Mater. 2018, 8, 1703154 10.1002/aenm.201703154. [DOI] [Google Scholar]
  106. Buchberger I.; Seidlmayer S.; Pokharel A.; Piana M.; Hattendorff J.; Kudejova P.; Gilles R.; Gasteiger H. A. Aging Analysis of Graphite/LiNi1/3Mn1/3Co1/3O2 Cells Using XRD, PGAA, and AC Impedance. J. Electrochem. Soc. 2015, 162, A2737–A2746. 10.1149/2.0721514jes. [DOI] [Google Scholar]
  107. Zhan C.; Wu T.; Lu J.; Amine K. Dissolution, Migration, and Deposition of Transition Metal Ions in Li-Ion Batteries Exemplified by Mn-Based Cathodes-A Critical Review. Energy Environ. Sci. 2018, 11, 243–257. 10.1039/c7ee03122j. [DOI] [Google Scholar]
  108. Sun H. H.; Manthiram A. Impact of Microcrack Generation and Surface Degradation on a Nickel-Rich Layered Li[Ni0.9Co0.05Mn0.05]O2 Cathode for Lithium-Ion Batteries. Chem. Mater. 2017, 29, 8486–8493. 10.1021/acs.chemmater.7b03268. [DOI] [Google Scholar]
  109. Li W.; Dolocan A.; Oh P.; Celio H.; Park S.; Cho J.; Manthiram A. Dynamic Behaviour of Interphases and Its Implication on High-Energy-Density Cathode Materials in Lithium-Ion Batteries. Nat. Commun. 2017, 8, 14589. 10.1038/ncomms14589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Kim J. H.; Pieczonka N. P. W.; Lu P.; Liu Z.; Qiao R.; Yang W.; Tessema M. M.; Sun Y. K.; Powell B. R. In Situ Formation of a Cathode-Electrolyte Interface with Enhanced Stability by Titanium Substitution for High Voltage Spinel Lithium-Ion Batteries. Adv. Mater. Interfaces 2015, 2, 1500109 10.1002/admi.201500109. [DOI] [Google Scholar]
  111. Kim J. M.; Kim C.; Yoo S.; Kim J. H.; Kim J. H.; Lim J. M.; Park S.; Lee S. Y. Agarose-Biofunctionalized, Dual-Electrospun Heteronanofiber Mats: Toward Metal-Ion Chelating Battery Separator Membranes. J. Mater. Chem. A 2015, 3, 10687–10692. 10.1039/c5ta02445e. [DOI] [Google Scholar]
  112. Zhu W.; Chong S.; Sun J.; Guo S.; Tai Z.; Wang Y.; Liu Y. The Enhanced Electrochemical Performance of Li1.2Ni0.2Mn0.6O2 through Coating MnF2 Nano Protective Layer. Energy Technol. 2019, 7, 1900443 10.1002/ente.201900443. [DOI] [Google Scholar]
  113. Strehlow W. H.; Cook E. L. Compilation of Energy Band Gaps in Elemental and Binary Compound Semiconductors and Insulators. J. Phys. Chem. Ref. Data 1973, 2, 163–200. 10.1063/1.3253115. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

mg2c00060_si_001.pdf (7.6MB, pdf)

Articles from ACS Materials Au are provided here courtesy of American Chemical Society

RESOURCES