Triphenylphosphine Oxide as Highly Effective Electrolyte Additive for

6 hours ago - Nickel-rich layered oxide materials (LiNixMnyCo1-x-yO2, x ≥ 0.8, LiNMC) attract great interest for application as positive electrode i...
0 downloads 12 Views 3MB Size
Subscriber access provided by UNIVERSITY OF TOLEDO LIBRARIES

Triphenylphosphine Oxide as Highly Effective Electrolyte Additive for Graphite/NMC811 Lithium Ion Cells Kolja Beltrop, Sven Klein, Roman Nölle, Andrea Wilken, Juhyon J. Lee, Thomas K.-J. Köster, Jakub Reiter, Liang Tao, Chengdu Liang, Martin Winter, Xin Qi, and Tobias Placke Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.8b00413 • Publication Date (Web): 28 Mar 2018 Downloaded from http://pubs.acs.org on March 28, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Triphenylphosphine Oxide as Highly Effective Electrolyte Additive for Graphite/NMC811 Lithium Ion Cells Kolja Beltropa, Sven Kleina, Roman Nöllea, Andrea Wilkena, Juhyon J. Leeb, Thomas K.-J. Kösterb, Jakub Reiterb, Liang Taob, Chengdu Liangc, Martin Wintera,d, Xin Qia,*, Tobias Plackea,* a

University of Münster, MEET Battery Research Center, Institute of Physical Chemistry, Corrensstr. 46, 48149 Münster, Germany b

c

BMW Group, Research Battery Technology, Petuelring 130, 80788 München, Germany

Contemporary Amperex Technology Limited, No.1 Xingang Rd., Jiaocheng District, Ningde, 352100, Fujian Province, China d

Helmholtz Institute Münster, IEK-12, Forschungszentrum Jülich GmbH, Corrensstr. 46, 48149 Münster, Germany

ABSTRACT Nickel-rich layered oxide materials (LiNixMnyCo1-x-yO2, x ≥ 0.8, LiNMC) attract great interest for application as positive electrode in lithium ion batteries (LIBs) due to high specific discharge capacities at moderate upper cut-off voltages below 4.4 V vs. Li/Li+. However, the comparatively poor cycling stability as well as inferior safety characteristics prevent this material class from commercial application so far. Against this background, new electrolyte formulations including additives are a major prerequisite for a sufficient electrochemical performance of Ni-rich NMC materials. In this work, we introduce triphenylphosphine oxide (TPPO) as electrolyte additive for the application in graphite/LiNi0.8Mn0.1Co0.1O2 (NMC811) cells. The addition of only 0.5 wt.% TPPO into a carbonate-based electrolyte (LiPF6 in EC:EMC) significantly increases the 1st cycle Coulombic efficiency (CE) as well as the reversible specific capacity and improves the capacity retention of the LIB full cell cycled between 2.8 - 4.3 V. Electrochemical results indicate that the full cell capacity fade is predominantly caused by active lithium loss at the negative electrode. In this contribution, X-ray photoelectron spectroscopy and inductively coupled plasma-mass spectrometry analysis confirm the participation of the electrolyte additive in the solid electrolyte interphase (SEI) formation on the negative electrode as well as in the cathode electrolyte interphase (CEI) formation on the positive electrode, thus, effectively reducing the active lithium loss during cycling. Furthermore, the performance of 1 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the TPPO additive is compared to literature known electrolyte additives including triphenylphosphine (TPP), vinylene carbonate (VC) and diphenyl carbonate (DPC) demonstrating the outstanding working ability of TPPO in graphite/NMC811 cells.

*Corresponding authors: Dr. Tobias Placke

Dr. Xin Qi

[email protected]

[email protected]

Tel.: +49 251 83-36826

Tel.: +49 251 83-36716

Fax: +49 251 83-36032

Fax: +49 251 83-36032

TOC (Graphical abstract)

2 ACS Paragon Plus Environment

Page 2 of 39

Page 3 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

1. Introduction In order to advance lithium ion batteries (LIBs) with respect to energy density, power density, lifetime and safety, many efforts have been made to substantially expand the application areas of LIBs. Especially the increasing demands on both high gravimetric and volumetric energy density in LIBs for automotive applications, raise the research efforts all over the world.1-6 With respect to active material development, further development of cathode materials is still considered as the main driver to improve energy density, thus, novel cathode materials with enhanced capacity and/or operating voltage are pursued. Unfortunately, high voltage cathode materials are highly restricted by the narrow electrochemical stability window of the state-ofthe-art LiPF6 in carbonate-based electrolytes (≈1.0 – 4.4 V vs. Li/Li+)7-9 and related side reactions, such as metal dissolution10, 11. Instead of focusing only on the high-voltage cathode materials, an alternative way to enhance the energy density of the system is to use cathode materials with high specific capacity at moderate working potentials, i.e. below 4.4 V vs. Li/Li+. By changing the content of each transition metal, a variety of different LiNixMnyCo1-x-yO2 (LiNMC)

materials

LiNi1/3Mn1/3Co1/3O2

have

been

(NMC111),

synthesized

and

LiNi0.5Mn0.3Co0.2O2

characterized, (NMC532),

including

e.g.

LiNi0.6Mn0.2Co0.2O2

(NMC622) and LiNi0.8Mn0.1Co0.1O2 (NMC811).12-15 While an increase in the Ni content leads to an enhanced capacity at the same cut-off potential, it is believed that the thermal stability will decrease in the opposite way and may result in safety issues.12, 16 Kasnatscheew et al. emphasized the dependency of the structural stability of various LiMO2 electrode materials on the respective Li+ extraction ratio, which is related to the specific charge capacity and charge cut-off potential.7 Currently, NMC532 and/or NMC622 can be considered as state-of-the-art cathode materials for commercial high-energy LIBs used for automotive applications.17 In particular, high Ni-content LiNixCoyMn1-x-yO2 materials with x ≥ 0.6, i.e. NMC811, have attracted more attention recently due to their high specific capacity of ≈200 mAh g-1 at an upper cut-off potential of 4.3 V vs. Li/Li+.12 However, the overall cell performance suffers from severe capacity fading caused by ongoing chemical electrolyte decomposition18,

19

and/or intrinsic structural changes of the material, such as large c-axis shrinkage at high cathode potentials, and limit its commercial application so far.20 The increased amount in highly reactive Ni4+ surface sites in the de-lithiated state of NMC811 are suggested to be the predominant cause for the parasitic side reactions between the electrolytes and the NMC surface.21 However, recent studies reveal that the release of reactive oxygen from the NMC surface is causing chemical oxidation of the electrolyte, yielding in increased CO2 and CO 3 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

evolution and that a transition metal induced catalytic effect on the electrochemical electrolyte oxidation does not exist.19 Many efforts have been made to improve the electrochemical performance of the LiNMC systems, including doping as well as coating of the active material with several components to increase the intrinsic structure and to circumvent the above mentioned failure mechanisms.22, 23

Further attempts to stabilize the cathode/electrolyte interface have been made by tailoring

the electrolyte formulations, in order to avoid ongoing electrolyte decomposition and transition metal ion dissolution. The state-of-the-art carbonate-based electrolyte solutions display only a compromise in terms of cycling stability, thermal stability and safety issues because of the manifold advantages and disadvantages of each single component.24-27 A very promising approach to master the above mentioned challenges lies in the addition of small amounts of other components into the electrolyte system to tailor specific targeted properties of the electrolyte without changing the bulk properties.28, 29 These additives are able to create a specific electrode/electrolyte interface thus reducing the chemical/electrochemical side reactions. In addition, the electrolyte properties in terms of electrochemical stability, ionic conductivity and thermal stability can also be enhanced.28, 30 Various literature reports deal with the influence of these electrolyte additives, such as HFscavengers, oxidative stability enhancers, flame retardants or overcharge protection additives, on the overall cell performance, demonstrating the great relevance of this specific research field.31-39 Furthermore, film forming additives are expected to be either oxidized or reduced prior to the electrolyte solvent in order to build a thin and homogeneous passivation film (SEI or CEI) on both the anode and cathode surface to prevent the electrolyte from ongoing (electro-) chemical decomposition.39-44. In general, it has to be considered that the electrolyte can react with both, anode and cathode and that these reactions mutually impact each other in the complete battery cell, depending on capacity and mass balancing.25, 45 Against this backdrop, it has been shown, that the NMC811 is a unique material and that a variety of famous electrolyte additives, like vinylene carbonate (VC), prop-1-ene-1,3-sultone (PES) or tris(trimethylsilyl) phosphite (TMSPi), which improve the cycling stability of graphite/NMC111 or graphite/NMC442 cells do not necessarily enhance the overall cell performance of graphite/NMC811 cells.46 Interestingly, it was proposed that the addition of VC and PES can effectively suppress the formation of a rock salt surface layer on the NMC811 particles, which is often claimed to be the major failure mechanism during longterm cycling.47 Recently, Qiu et al. stated on the application of methyl phenyl carbonate (MPC) and diphenyl carbonate (DPC) as electrolyte additives either as single component or in 4 ACS Paragon Plus Environment

Page 4 of 39

Page 5 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

combination with methylene methyl disulfonate and TMSPi in graphite/NMC811 pouch cells.46 It could be shown that the addition of 1 wt.% DPC leads to a decreased gas evolution, smaller voltage drop during storage, lower impedance increase and better capacity retention during long-term cycling. Phenyl carbonates are supposed to act as a kind of solid electrolyte interphase (SEI) modifier, rather than SEI formers because of the absence of a clear reduction peak prior to EC. Furthermore, the additives seem to have an impact on the positive electrode, as the charge endpoint capacity slippage rate is reduced. The work of Li et al. focused on the failure mechanism of graphite/NMC811 cells with and without the use of different additives including VC, PES, TMSPi and methylene methanedisulfonate (MMDS).21 Even though the addition of 2 wt.% VC to the electrolyte showed much better capacity retention and lower voltage drop compared to the PES system, severe impedance growth was monitored for all cells cycled above 4.2 V indicating an unsatisfactory effect of these additives on the total cell performance. Yim et al. reported on the effect of a bifunctionalized divinyl sulfone (DVS) as electrolyte additive for NMC-712 that improves the stabilization of the electrolyte-electrode interface by forming a durable protective layer on the cathode surface.48 As stated above, several famous electrolyte additives such as VC, PES and TMSPi were proven not to sufficiently work in graphite/NMC811 cells, demonstrating the unique surface chemistry in this system.46 In this contribution and to the best of our knowledge, DPC shows the best results among the literature reported electrolyte additives so far.46 In this contribution, we introduce triphenylphoshine oxide (TPPO) from the phosphine oxide family as an electrolyte additive for the application in graphite/NMC811 lithium ion cells. The chemical structure as well as some major general properties of TPPO are illustrated in Figure 1. Against this background, TPPO combines the most beneficial properties such as low costs, non-toxicity (classified only as harmful)49 and results in a greatly enhanced cell performance. Therefore, it can be concluded that this additive is highly interesting for industry applications and further fundamental studies to advance LIBs.

2. Experimental

2.1 Electrode and electrolyte preparation The electrode materials and formulations applied in this work are summarized in Table 1. The anode contains 95.4 wt. % graphite (FSNC-1; Shanshan Technology; D50 = 15.0 ± 2.0 µm; 5 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

BET surface area = 1.3 ± 0.3 m2 g-1), 1.5 wt. % carbon black (Super C65; Imerys), 2.5 wt. % SBR (SB 5521; LIPATON; Polymer latex GmbH) and 0.6 wt. % Na-CMC (Walocel CRT 2000 PPA12; Dow Wolff Cellulosics) and was prepared in the laboratory. Deionized water was used as dispersant and the electrode paste was homogenized with a dispermat machine (10.000 rpm; 60 min; Dissolver Dispermat LC30, VMA-Getzmann GmbH) and cast onto the dendritic copper current collector (Carl Schlenk AG; purity: 99.9%). The areal capacity of the graphite anode was ≈2.6 mAh cm-2 and the density was 1.4 g cm-3, as determined by the capacity of the cathodes in order to achieve a suitable anode/cathode capacity balancing. For the full cell investigations, the capacity ratio between the cathode and anode was set as 1:1.30, to avoid lithium metal plating at the graphite anode. The NMC811-based cathodes were prepared in large scale with the battery line in the MEET institute. Carbon black (Super C65; Imerys) and a PVdF solution in N-methyl-2-pyrrolidone (Solef 5130, Solvay; 6 wt. %) were first added into an air-tight container and homogenized by a shear mixer at 2500 rpm for half an hour. Afterwards, the NMC811 material (provided by CATL; BET surface area= 0.64 m2 g-1; D90= 18.3 µm) was added and the paste was mixed for another 1.5 hours. The solid loading of the electrode paste was 50% and the viscosity after dispersion was 1500 Pa s at a shear rate of 10 s-1. The container was set under reduced pressure to avoid the generation of gas bubbles. A water-cooling system was applied to remove the heat generated during the mixing process. The paste was cast on the aluminum foil (Evonik Industries; purity: 99.9%) and the dimension of each electrode sheet was 20 · 20 cm2. The electrodes were cut into 12 mm ø disks (1.13 cm2), and the areal capacity of all the electrodes was controlled as ≈2.1 – 2.2 mAh cm-2 with a standard deviation of less than 0.03 mAh cm-2. The influence of the electrolyte additives on the cycling stability and Coulombic efficiency of the graphite/NMC811 system was evaluated within the benchmark electrolytes 1 M LiPF6 in EC/EMC 3:7 by weight (Solvionic; purity: battery grade), which was defined as baseline electrolyte 1 (BL 1) and 1 M LiPF6 in EC/EMC 1:1 by wt. (Solvionic purity: battery grade), defined as baseline electrolyte 2 (BL 2), respectively. In each case 0.5 wt. % of the corresponding electrolyte additive, i.e. triphenylphosphine oxide (TPPO; Sigma Aldrich; purity: 98.5%), triphenylphosphine (TPP; Sigma Aldrich; purity: 99%), vinylene carbonate (VC; Sigma Aldrich; purity: 99.5%) or diphenyl carbonate (DPC; Sigma Aldrich; purity: 99%) was used. The electrolyte preparation and storage as well as the cell manufacturing were carried out in an argon-filled glove box (H2O and O2 contents < 0.1 ppm).

6 ACS Paragon Plus Environment

Page 6 of 39

Page 7 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

2.2 Cell preparation and electrochemical characterization In order to monitor the potential changes of the anode and cathode vs. the reference electrode

(lithium

metal;

Albemarle

Corporation;

purity:

battery

grade)

in

the

charge/discharge process, a three-electrode set-up (Swagelok type T-cell) was applied for the electrochemical characterization. For the investigation of the graphite/NMC811 cells, a polypropylene separator (PP, Freudenberg 2190, 6 layers) soaked with 240 µL of the corresponding electrolyte was used. To ensure a sufficient wetting of the composite electrodes (surface area: 1.13 cm2) by the electrolytes, the assembled cells were equilibrated for 24 hours at 20 °C. It has to be kept in mind that a much larger amount of electrolyte is used in these laboratory three-electrode cells compared to commercial cells, therefore, the percentage of the studied electrolyte additive most likely needs to be adjusted when going to commercial cells formats. The electrochemical charge/discharge cycling stability of the different electrolyte formulations was examined via constant current charge/discharge cycling (CCC), performed on a Maccor 4000 battery analysis system within climatic chambers set to 20 °C. The cells were charged/discharged between 2.8 V and 4.3 V, with a specific current of 19.6 mA g-1 (0.2 mA cm-2, 0.1C) for two formation cycles in order to provide a homogeneous SEI formation. A current density of 2.0 mA cm-2 is defined as 1C. Afterwards, the cells were cycled with 98 mA g-1 (1 mA cm-2, 0.5C). A constant voltage charge step at 4.3 V after each constant current charge process was applied until the current dropped below 0.01C. The calculation of the capacity retention after representative cycles is based on the discharge capacity of the third cycle after the formation process. The oxidative stability limits of various electrolyte compositions were obtained by means of linear sweep voltammetry (LSV) using a VSP instrument (Bio-Logic Science Instruments) with a scan rate of 1 mV s‐1. All measurements were carried out using metallic lithium (Albemarle Corporation; purity: battery grade) as counter and reference electrodes. For the oxidative scan, a platinum working electrode (WE; geometry: round-shaped; surface area: 7.85 · 10-3 cm2) was used. The anodic limits were arbitrarily defined as the potential at which the current density reached 0.025 mA cm-2. Cyclic voltammetry (CV) of the graphite anodes was performed on a VSP instrument (Bio-Logic Science Instruments).The scan rate was set to 0.05 mV s-1 within a potential range of 0.005 V and 1.8 V vs. Li/Li+. Graphite was used as WE and metallic lithium as counter and reference electrodes.

7 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2.3 Characterization of the specific surface composition via XPS analysis X-ray photoelectron spectroscopy (XPS) measurements were performed using a monochromatic Al Kα source (hν = 1486.6 eV) at a 10 mA filament current and a 12 kV filament voltage source energy. In order to compensate for the charging of the sample, a charge neutralizer was used. The measurement was carried out applying a pass energy of 40 eV at a 0° angle of emission. The measurement time was 600 s, and the lateral resolution was 3 µm. The pressure within the analysis chamber was 10−7 Pa. The fitting was performed using CasaXPS. The data fitting was carried out as described in detail elsewhere.50, 51 For the positive electrodes, the energy of the C 1s peak (conductive carbon at a binding energy (BE) of 284.5 eV) was used as internal standard for the calibration of the BE of the measured spectra. For the negative electrodes, the energy scale of the measured spectra was adjusted using the energy of the C 1s peak (graphite at BE = 284.5 eV) as internal reference. The harvested positive and negative electrodes were taken from discharged cells after 102 cycles in BL 1 and BL 1 + 0.5 wt. % TPPO. Prior to the measurement, the harvested electrodes were rinsed with dimethylcarbonate (3 × 1 mL) to remove electrolyte residue.

2.4 Determination of active lithium-loss via ICP-MS analysis The harvested graphite electrodes were taken from discharged cells after two formation cycles in BL 1 and BL 1 + 0.5 wt. % TPPO. Prior to the measurement, the harvested electrodes were rinsed with dimethyl carbonate (3 × 1 mL) to remove electrolyte residue. The gathered values for the lithium content display an average of three measurements. The graphite electrodes were weighed into PTFE vessels and digested in a Multiwave PRO microwave reaction system SOLV (Anton Paar, Graz, Austria) with 3 mL nitric acid and 3 mL hydrochloric acid. After cooling down, the samples were filled up to 25 mL with deionized water and diluted 1:10 for analysis. Calibration solutions were prepared in a range from 0.01 µg/L to 500 µg/L containing 1 Vol.-% of nitric acid. An Agilent 7900 inductively coupled plasma-mass spectrometer (ICP-MS) with Pt-cones from Agilent Technologies was used for total determination of lithium. The system was controlled by the software Mass Hunter 4.3 Workstation Software for ICP 7900 (Agilent Technologies). For sample introduction, a SPS 4 auto sampler (Agilent Technologies) was preconnected to the peristaltic pump of the system combined with a MicroMist nebulizer. The torch position, ion lenses, gas output, resolution axis, and background were optimized daily with the tuning solution (1 mg/L Ce, Co, Li, Tl, Y) to perform a short-term stability analysis of the instrument, to maximize ion signals and to 8 ACS Paragon Plus Environment

Page 8 of 39

Page 9 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

minimize interference effects due to high oxide levels (CeO+/ Ce+ < 1.2%) and doubly charged ions (Ce2+/Ce+ < 1%). Linearity response in the pulsed and analogue modes (P/A factor determination) was verified weekly using PA tuning solutions. Detailed information about instrumental settings for ICP-MS analysis can be found in Table S1 (supporting information).

3. Results and discussion

3.1 Influence of TPPO as electrolyte additive on the electrolyte stability The results from the linear sweep voltammetry measurements for BL 1 and BL 2 electrolytes without (a, c) and with (b, d) 0.5 wt.% TPPO are depicted in Figure 2, respectively. The results indicate that the addition of TPPO to the baseline electrolytes decreases the oxidative stability of both electrolyte systems (stability decrease of ≈0.16-0.23 V). In the TPPO molecule, the phosphorus atom exhibits the highest possible oxidation state (+V). Thus, the observed increased current flow is most likely induced by the electrochemical oxidation of the phenyl moieties and/or P-C bonds. Due to the reduced electrolyte stability, an oxidative decomposition and contribution of TPPO in the CEI formation process appears reasonable and will be the subject of the following discussion. According to literature, the current density increase, starting at a potential of ≈4.2 V vs. Li/Li+, is attributed to the oxidative decomposition of the carbonates.[4] Furthermore, a possible migration of electrochemical reduction products of TPPO from the negative electrode through the electrolyte and subsequent oxidation on the cathode surface, often called as cross-talk between anode and cathode52, 53, is conceivable, however not terminable within this measurement. Nevertheless, it also has to be kept in mind that the validity of the stability data can be questioned, as it is not sure whether the obtained results can be transferred from the LSV measurement set-up to a practical LIB full cell set-up, as discussed in a previous publication.54 The CV diagrams for the graphite negative electrodes in BL 1 and BL 2 electrolytes with and without 0.5 wt.% TPPO are illustrated in Figure 3. The CV diagrams for the BL 2 electrolyte were implemented to evaluate the influence of an increased ethylene carbonate (EC) content (50 wt.% in BL 2 compared to 30 wt.% in BL 1) on the reductive electrolyte stability and to further study the working principle of the TPPO additive. From the measurements, the onset potential for electrolyte reduction was arbitrarily defined as the potential at which the specific current reached 0.003 mA mg-1 (as seen in the inset images).

9 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

In general, the addition of TPPO to the baseline electrolytes leads to a shift of the onset potential for electrolyte reduction to higher values, indicating the reductive decomposition of TPPO prior to the electrolyte solvent molecules and effectively reducing the electrolyte decomposition in the ongoing scan. In this contribution, the current density from the reductive decomposition of BL 2 (Figure 3c), in the first cycle (onset potential: ≈ 0.61 V vs. Li/Li+) is significantly higher than for BL 1 (onset potential: ≈ 0.79 V vs. Li/Li+; Figure 3a), indicating an enhanced decomposition of the BL 2 electrolyte, most likely caused by the increased EC content and further evidenced below. Against this background, the addition of TPPO to the baseline electrolytes has a strong impact on the lithiation/de-lithiation behavior of graphite-based anodes as denoted by the reduced current peak area in the first charge process (Figure 3b and d). For example, the current density increase during lithiation in the first reductive scan in the TPPO-containing BL 2 electrolyte starts at ≈ 0.79 V vs. Li/Li+ compared to 0.61 V vs. Li/Li+ in the pure BL 2 electrolyte. The reductive peak at ≈ 0.8 V vs. Li/Li+ is most likely caused by the electrochemical reduction of TPPO molecules on the electrode surface, whereas the broad reductive peak at ≈ 0.6 V vs. Li/Li+ can be related to the electrochemical reduction of EC molecules, both forming a solid electrolyte interphase (SEI) layer on the graphite active material29, 55, and especially on the “non-basal plane” surfaces of graphite.56-59 A particularly conspicuous aspect in this contribution is the prevention of the severe electrolyte decomposition between 0.6 and 0.2 V vs. Li/Li+ by the addition of TPPO, evidencing the effective contribution of the additive in the SEI formation process on graphite. For all electrolyte systems, the discussed reduction peaks vanished in the 2nd cycle, indicating the formation of an effective SEI on graphite.

3.2 Influence of TPPO addition on the Coulombic efficiency and capacity retention in graphite/ NMC811 cells In order to study the influence of the electrolyte additive TPPO on electrochemical performance of the graphite/ NMC811 cells, constant current cycling experiments were conducted. The cathode/anode capacity ratio is a critical parameter to achieve an optimized cell performance45 and was set as 1:1.30, with respect to the properties of the active materials and composite electrodes. According to the CV measurements (Figure 3), the BL 2 electrolyte was studied to further evaluate the influence of a higher EC content on the overall cell performance. The corresponding results are depicted in Figure 4.

10 ACS Paragon Plus Environment

Page 10 of 39

Page 11 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

In general, graphite/NMC811 cells with BL 1 electrolyte show a higher discharge capacity as well as higher Coulombic efficiency (CE) in the first and ongoing cycles compared to BL 2 electrolyte (cf. Table 2). A possible reason for this behavior can be assigned to the differences in the SEI formation at the graphite anode. As stated above, the BL 2 electrolyte contains a higher amount of EC and most likely forms a thicker SEI at the anode, which consumes more Li+ from the NMC811 cathode. As a result, less active lithium would remain in the system to participate in the electrochemical reaction in the following charge/discharge processes. Thus, the 1st cycle CE and the specific discharge capacity are directly affected, i.e. BL 1 displays an enhanced 1st cycle CE (≈70%) compared to BL 2 (≈62%, Table 2). By adding only 0.5 wt.% TPPO to the electrolytes, the overall discharge capacity, CE and capacity retention can be greatly increased (≈26% discharge capacity gain in BL 1 and ≈44% discharge capacity gain in BL 2 electrolyte after 100 cycles) indicating a participation of the additive in the formation of the corresponding surface layers on the anode and/or cathode. Most likely, due to an increased active lithium loss during the first cycles in BL 2 compared to BL 1 electrolyte, the overall discharge capacity is reduced and, thus, the deviation in discharge capacity between BL 1 and BL 2 electrolytes with and without TPPO as additive is more severe (Figure 4). In general, active lithium loss which results in capacity fading is considered as one major failure mechanism for LIB full cells and has to be minimized, in particular in the formation cycles but also during ongoing cycling.60-65 As a consequence, the BL 1 electrolyte was applied as benchmark electrolyte in this work, due to the superior cycling performance and, thus, the following discussion focuses only on the influence of TPPO as electrolyte additive in the BL 1 electrolyte (=1 M LiPF6, EC/EMC 3:7 by wt.). The charge/discharge potential profiles of graphite/NMC811 cells, taken from the 1st cycle in BL 1 and BL 1 + 0.5wt.% TPPO as electrolyte are depicted in Figure 5a and Figure 5b, respectively. Figure 5c and d show the enlarged potential profiles of the graphite anodes. In the first charge process, both cathodes reach nearly the same cut-off potential against the reference electrode (≈4.38 V vs. Li/Li+), while both anodes reach ≈0.08 V vs. Li/Li+. Judging from the anode potential profiles during charge, both electrolytes start decomposition from ≈0.8 V vs. Li/Li+, resulting in the formation of the SEI layer on graphite. The anode potential incline between 0.8 V – 0.2 V vs. Li/Li+ in the BL 1 electrolyte is much lower than for the one with TPPO, indicating a higher degree of decomposition caused by the reduction of the BL 1 electrolyte (Figure 5c, d). This phenomenon directly affects the cathode and anode potential profiles during discharge. As mentioned above, the decomposition of electrolyte components, 11 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

mainly from the reduction of EC is most likely accompanied with an increased active lithium loss during SEI formation. As a consequence, in the BL 1 electrolyte, insufficient active lithium is available to fully re-lithiate the NMC811, indicated by the rise in the anode potential together with the missing drop in the cathode potential at the end of discharge (Figure 5a). A closer look on the anode potential profiles in the low specific capacity area (0100 mAh g-1) during charge is illustrated in Figure 5d. In the potential range between 0.8 V – 0.55 V vs. Li/Li+, the reductive current increase is more pronounced in the TPPO containing electrolyte as also observed in CV measurements, indicating the reductive decomposition of TPPO for SEI formation (cf. Figure 3c). However, in this electrolyte, the anode potential drops much faster to 0.2 V vs. Li/Li+ and the overall first lithiation plateau (≈0.2 V – 0.1 V

vs. Li/Li+) is enlarged compared to the BL 1 electrolyte, resulting in an increased Li+ intercalation into graphite and, thus, leading to a higher specific capacity during discharge. According to the discharge potential profile, a specific discharge capacity of 197 mAh g-1 can be obtained for the electrolyte with 0.5 wt.% TPPO, whereas the BL 1 electrolyte shows only a minor discharge capacity of 167 mAh g-1. Thus, it can be assumed that more active lithium is present in the TPPO system after SEI formation which, in turn, leads to a higher discharge capacity in the ongoing cycles. This assumption is further supported by comparing the 1st cycle CE of the electrolyte systems. The addition of TPPO leads to an increase of the 1st cycle CE from 73% to 86%. In this context, it is of great importance to distinguish between a specific capacity loss caused by active lithium loss i.e. during SEI formation and/or kinetic limitations during the re-lithiation of the NMC active material.8 Consequently, if the observed specific capacity loss in BL 1 electrolyte is due to kinetic limitations, a constant voltage step (CV-step) at the discharge cut-off voltage in the 3rd cycle would result in an increased specific discharge capacity. In Figure S1 (supporting information), it becomes clearly visible that no discharge capacity gain was achieved during the CV-step, evidencing that the observed specific capacity loss is caused by active lithium loss rather than by kinetic limitations. In conclusion, the electrolyte additive TPPO strongly suppresses faradaic parasitic side reactions, presumably by forming a more effective, less lithium consuming and electronically insulating, but lithium ion conducting SEI layer. By comparing the constant voltage behavior during discharge in the work of Kasnatscheew et al. 8and in this work, it can be furthermore concluded, that the electrolyte composition does have a strong influence on the impact of kinetic limitations on the observed specific capacity loss. In order to gain a better understanding of the working mechanism of TPPO in graphite/NMC811 cells, the change of the upper cut-off potential of the cathode material 12 ACS Paragon Plus Environment

Page 12 of 39

Page 13 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

during cycling was monitored over time in the two different electrolytes. The results for the BL 1 electrolyte and the TPPO-based electrolyte are depicted in Figure 6a and b, respectively. For the baseline electrolyte, an ongoing increase in the cathode potential is observed, indicating most likely a continuous loss of active lithium.53, 61 The increased cathode potential (≈20 mV) is directly accompanied with an increased de-lithiation degree of the NMC structure during charge, which then may suffer from a surface reconstruction to a rock salt structure and/or oxidative electrolyte decomposition.20,

21

By the addition of TPPO to the

electrolyte, this increase in cathode potential can be effectively suppressed. After a small initial increase (≈5 mV), the cathode potential remains stable over the whole cycling (300 cycles) and helps to improve the stability of the system. The long-term cycling behavior of NMC 811/graphite cells with and without TPPOcontaining electrolytes is depicted in Figure 7. By addition of TPPO, an overall discharge capacity gain of ≈30% based on the 350th cycle with a CE exceeding 99.9% can be achieved. Furthermore, the capacity retention after 295 cycles in the TPPO containing electrolyte was still above 80% state of health (SOH, based on the discharge capacity after the 2nd cycle) whereas the BL 1 electrolyte reached the end of life (EOL) criteria after 167 cycles. This results in a remarkable capacity retention increase of ≈77% for the TPPO-based cell. According to the above discussed electrochemical investigations, it can be concluded that TPPO effectively acts as a SEI modifier in order to decrease faradaic parasitic side reactions with – most likely - less active lithium loss in the first and ongoing cycles. Furthermore, a positive effect of TPPO on the overall cathode performance cannot be excluded, especially regarding the strong capacity retention increase. Further electrochemical and spectroscopic investigations were conducted and are presented in the following to elucidate the working mechanism of TPPO as electrolyte additive in graphite/ NMC811 cells.

3.3 Analysis of the interphases and active lithium loss: XPS and ICP-MS investigations of pristine and harvested electrodes The chemical and electrochemical anodic and cathodic decomposition of electrolyte components on the specific positive and negative electrode surface results in the formation of unique interphases at both electrode/electrolyte interfaces, strongly affecting the overall cell performance.31 Against this background, the evaluation of a possible participation of TPPO in the corresponding cathode electrolyte interphase (CEI), as well as in the SEI formation is of great interest. Therefore, XPS analysis was conducted to prove the assumption made above of 13 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the SEI modifying ability and to verify a possible influence of TPPO on the cathode performance. The surface analysis may provide a deeper understanding of the working principle of TPPO as electrolyte additive in graphite/NMC811 cells. In the following, XPS spectra and elemental concentrations of pristine and harvested positive and negative electrodes from discharged cells after 102 cycles in BL 1 and BL 1 + 0.5 wt.% TPPO are presented in their entirety.

3.3.1 XPS analysis of the NMC811 positive electrodes The F 1s, Mn 2p, O 1s, C 1s, P 2p and Li 1s spectra of the pristine and harvested positive electrodes from discharged cells with and without 0.5 wt.% TPPO are displayed in Figure 8, respectively. Additionally, the corresponding elemental concentrations (at.%) are summarized in Table 3 and the mean atomic concentration percentages (at.%) of the specific surface layers are listed in Table S2. The results from the pristine cathode surface are discussed for the purpose of quantitative interpretation of the electrochemically treated samples. The surface of the pristine NMC811 cathode is mainly composed of species related to the conductive carbon and to the PVdF binder. From the F 1s spectrum, the peak at ≈687.5 eV can be attributed to the C-F bonds in PVdF.50 In addition, also a small peak is observed at ≈684.7 eV which is presumably caused by the presence of LiF on the cathode/electrolyte interface as remaining impurity of the electrode paste preparation process and due to parasitic dehydrofluorination of PVdF.66, 67 A small intensity signal at ≈642.3 eV is present in the Mn 2p spectrum and is related to the NMC811 active material. The O 1s spectrum displays two peaks at ≈531.4 and ≈529.1 eV and a small shoulder at ≈533.4 eV, which is presumably caused by chemically formed R2CO3, low coordination oxygen atoms68 and possibly by Li2CO369 as remaining surface impurity from the synthesis process. The peak at ≈532.4 eV can be related to oxidized conductive carbon and the signal at ≈529.1 eV can be assigned to O2- anions in NMC811.70 The C 1s spectrum contains multiple contributions, in particular from the CF2 (CF2CH2; at ≈290.7 eV) and CH2 (CH2CF2; at ≈286.5 eV) groups of PVdF, respectively, and an additional large peak at ≈284.5 eV that results from C-H and C-C groups of the conductive carbon.71, 72 As expected, no peaks are present in the P 2p and Li 1s spectra. After constant current cycling in BL 1 (middle) and BL 1 + 0.5 wt.% TPPO-based electrolyte (bottom), the specific surface composition on NMC811 changes significantly. In this contribution, the decrease of the Mn 2p peak (≈643 eV) in combination with the decrease of the O 1s peak at ≈529 eV (O2- anions in the NMC811 active material) and the decrease of the C 1s peak at ≈284 eV (conductive carbon) for the cycled electrodes compared to the pristine 14 ACS Paragon Plus Environment

Page 14 of 39

Page 15 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

electrodes, indicates the formation of a compact surface film on the cathode, thus, diminishing the peak intensities of the active and inactive cathode materials.71 Due to the same mean atomic concentration percentages (≈0.20 at.%) of Mn 3/2, it can be assumed that the formed surface layers possess the same thickness for the different electrolyte formulations and allows, therefore, a precise quantification and interpretation of the specific elemental concentrations. From the F 1s spectrum as well as from the mean atomic concentration percentages (cf. Table S2), a distinct increase in the LiF content, which has been identified as a common decomposition product of LiPF6 73, on the surface of the cycled electrodes is clearly visible. However, no difference in the accumulated F 1s elemental concentration of both electrolytes is observed. Judging from the O 1s spectra of the cycled electrodes, the increased peak intensity at ≈533.4 eV arises from the electrochemical decomposition of the electrolyte solvent and salt.74 Even though, the mean atomic concentration percentages for Li2CO3 remain the same for the pristine and the cycled electrodes, a clear difference is found for the alkyl carbonates. With an alkyl carbonate mean value of ≈2.28 at.% for the BL 1 compared to

≈2.97 at.% for the TPPO-containing electrolyte, the latter electrolyte formulation shows the highest value for the alkyl carbonates and indicates the CEI modifying ability of TPPO. These findings are further supported by the increased carbonate and phosphate content seen in the C 1s and P 2p spectra as well as in the corresponding mean atomic concentration percentages. The peak at ≈134 eV in the P 2p spectrum is assigned to a P atom that is bound to a less electronegative element than F, such as O, representing the TPPO molecule or its decomposition products. In this context, a peak shift to lower binding energy (BE) is present in the spectrum for the TPPO-based system. Furthermore, a significant increase in the LiPOxFy (≈134.3 eV) content from 0.14 at.% in BL 1, to 0.23 at.% in the TPPO containing electrolyte was observed and clearly indicates the participation of the TPPO additive in the CEI layer composition. However, no clear oxidation of TPPO was observed in the LSV measurements (on platinum), as discussed above. Further, the F 1s spectra of LiPOxFy overlap with the peak coming from the PVdF binder and prevents a further identification of these decomposition species. As stated in literature, an increased amount in LiPOxFy species was proven to have a beneficial effect on the electrochemical stability of the CEI surface layer.75 The Li 1s spectra and the elemental concentration of lithium species show no difference for the harvested electrodes. Finally, the XPS results of pristine and harvested positive electrodes may be summarized as follows: i) after 102 cycles in electrolytes with and without TPPO, a compact surface film on the cathode, formed by electrolyte solvent and salt decomposition products was confirmed; ii) the resulting surface films possess the same thickness for the 15 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

different electrolyte formulations and allow, therefore, a precise quantification and interpretation of the specific elemental concentrations of the surface layers; iii) the surface film resulting from the electrochemical decomposition of the TPPO containing electrolyte shows an increased amount of alkyl carbonates as well as LiPOxFy species, indicating the participation of TPPO in the CEI formation process. Furthermore, the resulting surface layers are still thin enough (< 5 nm) to allow the detection of several electrode components such as conductive carbon, PVdF and NMC811.

3.3.2 XPS analysis of the graphite negative electrodes The specific F 1s, O 1s, C 1s, P 2p and Li 1s spectra of the pristine and harvested negative electrodes from discharged cells after 102 cycles in BL 1 and BL 1 + 0.5 wt.% TPPO are displayed in Figure 9, respectively. In addition, the corresponding elemental concentrations (at.%) are summarized in Table 4 and the mean atomic concentration percentages (at.%) of the specific surface layers are listed in Table S3. In the F 1s spectra of the cycled electrodes in both electrolytes, two peaks are present at

≈684.7 eV and ≈686.2 eV and can be assigned to the formation of LiF/POyFz76 species for the former peak and LiPFx species for the latter one, respectively. Interestingly, no difference in the elemental concentrations of P 2p species (cf. Table 4) is observed, indicating that the significant increase in the F 1s peak at ≈684.7 eV for the TPPO containing electrolyte is predominately related to an increase in LiF rather than of POxFy- species in the SEI layer. The O 1s spectra of the pristine anode shows a broad peak at ≈533.0 eV corresponding to small amounts of oxygen, most likely from the graphite prismatic surfaces and contributions from the oxygen containing CMC binder.51, 56 For the cycled electrodes, the peak intensity at

≈533.0 eV increases primarily for the BL 1 electrolyte and additionally, a significant increase in the peak at ≈531.3 eV, attributed to Li2CO3 formation, is seen in the O 1s spectra for both electrolyte formulations. However, the asymmetric peak shape of the O 1s spectra as well as overlapping of peak intensities coming from possible lithium alkyl carbonates (RLiCO3), R2CO3, CMC and Li2CO3 species creates difficulties in a precise identification of major parts of the SEI.77 Unfortunately, the same issue is present in the C 1s spectra. The strong asymmetric peak shape allows only a rough assessment of trends and prevents an exact differentiation between the different species in the C 1s spectra.51 After cycling, the graphite peak intensity at ≈284.5 eV decreases in both electrolyte systems and shows the lowest value for the electrodes cycled in the BL 1 electrolyte. As the intensity of the C 1s peak at

≈284.5 eV is also affected by possible decomposition products, a reliable statement on the 16 ACS Paragon Plus Environment

Page 16 of 39

Page 17 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

SEI thickness based on the graphite signal is not possible without additional sputter depth profiling measurements. The most prominent change in the C 1s spectra is observed in the BL 1 electrolyte. A new peak at ≈282.2 eV appears and is related to the formation of a lithium/graphite species (LiCx). The identification of a metal / carbon species on an electrochemically treated graphite negative electrode, owning a negative BE shift with respect to the graphite signal, was also reported by Niehoff et al.51 This may explain the increased irreversible active lithium loss in the first and ongoing cycles in the BL 1 electrolyte and further explains the reduced discharge capacity compared to the TPPO-containing electrolyte. Combining the results from the O 1s and C 1s spectra, it can be concluded that the nature of the formed SEI in the baseline electrolyte without TPPO contains more organic compounds due to increased solvent reduction products and supported by the higher amounts of R2CO3, carbonates and PEO species, as found in the measured surface layer composition. The unique nature of the SEI formed in the TPPO-based electrolyte is more inorganic, as a result of increased salt reduction/decomposition products such as LiF. This assumption is further supported by analyzing the Li 1s spectra. The highest peak intensity at ≈55.2 eV together with the highest elemental concentration of Li 1s was observed for the TPPO-containing electrolyte. Finally, the following conclusion of the discussed XPS results for pristine and cycled negative electrodes in different electrolyte formulations appears reasonable: i) the SEI formed in the baseline electrolyte (BL 1) consists of more organic reduction/decomposition species, predominantly from solvent molecules; ii) metal/carbon species were found in the surface layer formed in the TPPO-free electrolyte, likely explaining the increased lithium-loss observed during cycling; iii) the SEI resulting from the BL 1+TPPO-based electrolyte is mainly

composed

of

inorganic

species

such

as

LiF,

resulting

from

the

reduction/decomposition of conductive salt species.

3.3.3 Data from the ICP-MS analysis In accordance to the electrochemical investigations which identified the observed specific capacity loss after formation as a result of active lithium loss, rather than kinetic limitations during the re-lithiation of the NMC active material (cf. chapter 3.2 and Figure S1), also the XPS results from the negative electrodes indicate an increased active lithium loss during the SEI formation process. This process is assumed to be the main reason for the deviation in discharge capacity in the baseline electrolyte compared to the TPPO-based electrolyte. To further verify this suspicion, inductively coupled plasma-mass spectrometry investigations 17 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(ICP-MS) from harvested graphite negative electrodes after formation (2 cycles; completely discharged state) were conducted to quantify the active lithium loss and to correlate these results with the deviation of the discharge capacity in the specific electrolytes (Figure 10). The active lithium loss (defined as the remaining mean specific lithium content; average of three cells) of the anode is ≈ 62% higher in the BL 1 electrolyte (0.53%) compared to the TPPO-containing electrolyte (0.20%). This result clearly evidences an increased active lithium loss during SEI formation in the BL 1 electrolyte, negatively influencing the 1st cycle CE (≈70% in BL 1compared to ≈86% in BL 1+TPPO electrolyte) and discharge capacity after the 2nd cycle (≈157 mAh g-1 in BL 1 compared to ≈187 mAh g-1 in BL 1+TPPO electrolyte;

cf. Table S4). According to Faradays law, it is possible to calculate the charge (Q) in dependence of the amount of the remaining lithium content in the specific electrodes after formation:

ܳ = ݊∗‫ܨ∗ݖ‬

(1)

where Q is the total electric charge in coulombs (C), n is the amount of the substance (mol), z is the valence number of the substance and F is the Faraday constant (F=96485 C mol-1). As a result of the calculation according to equation (1) (cf. Table S4), the remaining lithium content for the BL 1 electrolyte without and with TPPO accounts to 47.4 mAh g-1 and 17.9 mAh g-1, respectively. These findings results in a difference of 29.5 mAh g-1 for the electrolyte systems which is very close to the difference in the specific discharge capacity observed in the CCC experiments (30.0 mAh g-1). Hence, the deviation in discharge capacity in the electrolyte systems is completely related to active lithium loss during SEI formation. Within these results, the outstanding SEI modifying ability of TPPO can be evidenced. Summarizing the above made discussion, the addition of TPPO to the baseline electrolyte effectively reduces the active lithium loss by the formation of effective passivation layer on the electrodes surfaces and greatly enhances the overall cell performance.

3.6 Correlation and comparison of TPPO to other electrolyte additives As stated above, several famous electrolyte additives such as VC, PES and TMSPi were proven not to sufficiently work in graphite/NMC811 cells, demonstrating the unique surface chemistry in this system (cf. chapter 1).46 In this contribution and to the best of our knowledge, DPC shows the best results among the literature reported electrolyte additives so far.46 To ensure a fair comparison, 0.5 wt.% of prominent electrolyte additives including triphenylphosphine (TPP), VC and DPC were added to the BL 1 electrolyte, respectively. According to the molecular weight (MW) of each additive, it is obvious that TPPO, as the 18 ACS Paragon Plus Environment

Page 18 of 39

Page 19 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

molecule with the highest MW, displays the lowest molality in the specific electrolyte formulation (Table S5). In comparison, the molality of VC is approximately three times higher. The specific cycling stability is illustrated in Figure 11 and the corresponding values for the discharge capacity and CE of selected cycles as well as the capacity retention after 100 cycles are summarized in Table 5. The TPPO-containing electrolyte shows the highest discharge capacity, 1st cycle CE and capacity retention after 100 cycles among the investigated electrolyte additives. However, a further optimization in view of the ideal amount of each compound in the specific electrolyte formulation and cycling condition (e.g. temperature) was not evaluated within this work. To the best of our knowledge, the presented results for the TPPO containing electrolyte are among the best performances of a graphite/NMC811 cell system reported so far. The addition of at least 0.5 wt.% of TPPO to the control electrolyte greatly enhances the overall cell performance. In addition, the additive is of low costs and is commercially available as it is a by-product of many useful reactions in organic synthesis, including the Wittig, Staudinger and Mitsunobu reactions. Furthermore, this compound is beneficial in terms of safety concerns, as it is listed to be slightly hazardous and not highly toxic or toxic as many other established electrolyte additives known from the literature, like sultones and phosphates.49, 75, 78

4. Conclusion In this study, the electrochemical performance of graphite/LiNi0.8Mn0.1Co0.1O2 (NMC811) cells cycled between 2.8 and 4.3 V in carbonate-based electrolytes, particularly in 1M LiPF6 in EC:EMC (3:7), with respect to the impact of triphenylphosphine oxide (TPPO; 0.5 wt.%) as electrolyte additive was comprehensively examined. The addition of TPPO to the electrolyte greatly enhances the overall cell performance of graphite/NMC811 cells and particularly results in a higher discharge capacity, an increased 1st cycle Coulombic efficiency (CE) as well as a remarkable improvement of the capacity retention (80% after 295 cycles). Furthermore, the TPPO-based system was compared to different

promising

electrolyte

additives

including

vinylene

carbonate

(VC),

triphenylphosphine (TPP) and diphenyl carbonate (DPC) in this cell set-up. The TPPOcontaining electrolyte showed the highest discharge capacity, highest 1st cycle CE and capacity retention after 100 cycles among these investigated electrolyte additives. To the best of our knowledge, the presented results for the TPPO containing electrolyte are among the best performances of a graphite/NMC811 cell system reported so far. However, in future 19 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

studies, an optimization with respect to the ideal additive amount of each compound has to be evaluated. Furthermore, we could confirm that TPPO and/or its decomposition products can participate in the formation of both interphases, i.e. the solid electrolyte interphase (SEI) and cathode electrolyte interphase (CEI), leading to the assumption that the additive results in a decreased active lithium loss. In electrochemical analysis, TPPO displayed a decreased oxidative stability as well as reduced charge consumption during SEI formation at the graphite negative electrode. From XPS analysis, we found that the surface film on NMC811, resulting from the electrochemical decomposition of the TPPO-containing electrolyte, showed an increased amount in alkyl carbonates as well as LiPOxFy species demonstrating the participation of TPPO in the CEI formation. The SEI layer formed in the TPPO-containing electrolyte was mainly composed of inorganic species, such as LiF, resulting from the reduction of conductive salt species. ICP-MS analysis also gave evidence for an increased active lithium loss during SEI formation in the pure electrolyte compared to the TPPO-containing electrolyte, confirming the lower discharge capacity and lower 1st cycle CE of the pure carbonate-based electrolyte system. The electrolyte additive TPPO combines the most beneficial properties such as low costs and non-toxicity as well as greatly enhanced cell performance for graphite/NMC811 cells. Therefore, it could be concluded that this additive is highly interesting for commercial applications. However, further fundamental studies are needed to fully understand the working mechanism of this electrolyte additive and will be the subject in future works.

Associated content The Supporting Information is available free of charge on the ACS Publications website at DOI: XXXXXX Experimental details, supporting figures and tables for electrochemical data, XPS and ICPMS analysis

Acknowledgements The authors wish to thank the BMW Group and Contemporary Amperex Technology Limited (CATL) for funding this work and for the material supply. We want to thank Dr. Stephan Röser for fruitful discussions and Dr. Philip Niehoff for his help with the XPS measurements. Our gratitude goes also to Gültekin Göl for the preparation of the NMC-based cathodes.

20 ACS Paragon Plus Environment

Page 20 of 39

Page 21 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure and Table Captions Figure 1: General chemical and physicochemical properties of the electrolyte additive triphenylphosphine oxide (TPPO). Figure 2: LSV diagrams for BL 1 (=1 M LiPF6, EC/EMC 3:7 by wt.; a), BL 1 + 0.5 wt.% TPPO (b), BL 2 (=1 M LiPF6, EC/EMC 1:1 by wt.; c) and BL 2 + 0.5 wt.% TPPO (d) electrolyte. Scan rate: 1 mV s-1; anodic limit: 0.025 mA cm-2. Figure 3: Cyclic voltammograms of graphite/lithium metal cells with BL 1 (=1 M LiPF6, EC/EMC 3:7 by wt.; a), BL 1 + 0.5 wt.% TPPO (b), BL 2 (=1 M LiPF6, EC/EMC 1:1 by wt.; c) and BL 2 + 0.5 wt.% TPPO (d). Cut-off potential: 0.005 – 1.4 V vs. Li/Li+; Scan rate: 0.05 mV s-1. Figure 4: Influence of the addition of 0.5 wt.% TPPO on the charge/discharge cycling behavior of graphite/NMC811 cells in BL 1 (=1 M LiPF6, EC/EMC 3:7 by wt.; a, b) and BL 2 (=1 M LiPF6, EC/EMC 1:1 by wt.; c, d) electrolyte. Depicted are the discharge capacities (a ,b) as well as the CEs (c, d). Cut-off voltage: 2.8 V-4.3 V. Figure 5: (a, b): Anode and cathode potential profiles vs. Li/Li+ of graphite/NMC811 cells during charge (black) and discharge (red). Cut-off voltage: 2.8 V-4.3 V; Electrolyte: BL 1 (=1 M LiPF6, EC/EMC 3:7 by wt.; a); BL 1 + 0.5 wt. % TPPO (b). (c, d): The corresponding anode potential profiles vs. Li/Li+ during charge and discharge in BL 1 (black) and BL 1 + 0.5 wt.% TPPO electrolyte (red). Figure 6: Change of the upper cut-off potential of the NMC811 cathode during cycling in the BL 1 electrolyte (=1 M LiPF6, EC/EMC 3:7 by wt.; a) and in the TPPO-containing electrolyte (b) taken from CCC experiments. Figure 7: Influence of the addition of 0.5 wt. % TPPO on the long-term cycling behavior of graphite/NMC811 cells in BL 1 (=1 M LiPF6, EC/EMC 3:7 by wt.) and BL 1 + 0.5 wt.% TPPO electrolyte during charge and discharge. Depicted are the discharge capacities (a) as well as the CEs (b). The discharge capacities until SOH 80% are depicted in green. The calculation of the capacity retention after representative cycles is based on the discharge capacity of the 3rd cycle after the formation process. Cut-off voltage: 2.8 V-4.3 V. Figure 8: XPS spectra (F 1s, Mn 2p, O 1s, C 1s, P 2p and Li 1s) of pristine (top) and harvested NMC811 electrodes taken from discharged cells after 102 cycles in BL 1 (=1 M LiPF6, EC/EMC 3:7 by wt.; middle) and BL 1 + 0.5 wt. % TPPO electrolyte (bottom). Figure 9: XPS spectra (F 1s, O 1s, C 1s, P 2p and Li 1s) of pristine (top) and harvested graphite negative electrodes taken from discharged cells after 102 cycles in BL 1 (=1 M LiPF6, EC/EMC 3:7 by wt.; middle) and BL 1 + 0.5 wt. % TPPO electrolyte (bottom). 21 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 10: Remaining specific Li content of harvested and fully discharged graphite anodes after the 2nd cycle in BL 1 electrolyte (=1 M LiPF6, EC/EMC 3:7 by wt.) with and without 0.5 wt. % TPPO. Figure 11: Comparison of the influence of the addition of 0.5 wt.% TPPO, TPP, VC and DPC into the BL 1 electrolyte (=1 M LiPF6, EC/EMC 3:7 by wt.) on the cycling behavior of graphite/NMC811 cells during charge and discharge. Depicted are the discharge capacities (a) as well as the CEs (b). Cut-off voltage: 2.8 V - 4.3 V.

Table 1: Electrode composition and constitution for the graphite/NMC811 cell investigations. Table 2: Discharge capacity and Coulombic efficiency data of investigated electrolytes taken from CCC of graphite/NMC811 cells of selected cycles (cf. Figure 4) and the calculation of the capacity retention after 100 cycles, based on the third cycle. Error range for the discharge capacity values is ± 4 mAh g-1 and for the Coulombic efficiency ± 0.9% for the first and ± 0.05% for ongoing cycles. Table 3: Mean atomic concentration percentages (at. %) (F 1s, Mn 2p, O 1s, C 1s, P 2p and Li 1s) of pristine (top) and harvested NMC811 electrodes taken from discharged cells after 102 cycles in BL 1 (=1 M LiPF6, EC/EMC 3:7 by wt.; middle) and BL 1 + 0.5 wt. % TPPO electrolyte (bottom). Table 4: Elemental concentrations (F 1s, Mn 2p, O 1s, C 1s, P 2p and Li 1s) of pristine (top) and harvested negative electrodes taken from discharged cells after 102 cycles in BL 1 (=1 M LiPF6, EC/EMC 3:7 by wt.; middle) and BL 1 + 0.5 wt.% TPPO electrolyte (bottom). Table 5: Discharge capacity and Coulombic efficiency data of investigated electrolytes taken from CCC of graphite/NMC811 cells of selected cycles (cf. Figure 10) and the calculation of the capacity retention after 100 cycles, based on the 3rd cycle. Error range for the discharge capacity values is ± 4 mAh g-1 and for the Coulombic efficiency ± 0.9% for the first and ± 0.05% for ongoing cycles. .

22 ACS Paragon Plus Environment

Page 22 of 39

Page 23 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 1

Figure 2

23 ACS Paragon Plus Environment

Chemistry of Materials

Figure 3

Figure 4 Cycle number 10

20

30

40

50

60

Cycle number 70

80

90

100

0

Discharge

200 175 150 125 100 75

BL 1 + 0.5 wt. % TPPO BL 1

a)

capacity / mAh g

-1

225

BL 1 + 0.5 wt. % TPPO BL 1

b)

99.0 0

20

30

40

50

60

70

80

90

100

50

60

70

80

90

100

50

60

70

80

90

100

125 100 75

BL 2 + 0.5 wt. % TPPO BL 2

a)

99.6 99.4 BL 2 + 0.5 wt. % TPPO BL 2

99.2

b)

99.0 10

40

150

99.8

99.2

30

175

99.8

99.4

20

200

50 100.0

99.6

10

225

50 100.0

Coulombic efficiency / %

Discharge

capacity / mAh g

-1

0

Coulombic efficiency / %

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 39

0

10

20

Cycle number

30

40

Cycle number

24 ACS Paragon Plus Environment

Page 25 of 39

Figure 5

Figure 6 4.410

4.410

b)

b)

Cathode Potential

Cathode Potential

4.405

Potential vs. Li/Li+ / V

4.405

Potential vs. Li/Li+ / V

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

BL 1

4.400 4.395 4.390 4.385 4.380 4.375

4.400 BL 1 + 0.5 wt.% TPPO 4.395 4.390 4.385 4.380 4.375

4.370

4.370 0

200

400

600

800

1000

1200

0

200

Time / h

400

600

Time / h

25 ACS Paragon Plus Environment

800

1000

1200

Chemistry of Materials

Figure 7 Cycle number capacity / mAh g

Discharge

-1

0

25

50

75

100

125

150

175

200

225

250

275

300

325

350

300

325

350

200

80% SOH 175 150 125

/

a)

100

/

BL 1 + 0.5 wt. % TPPO BL 1

80% SOH

100.0

Coulombic efficiency / %

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 39

99.8 99.6 99.4

BL 1 + 0.5 wt. % TPPO

99.2

b)

99.0 0

BL 1 25

50

75

100

125

150

175

200

Cycle number

26 ACS Paragon Plus Environment

225

250

275

Page 27 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 8

27 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 9

28 ACS Paragon Plus Environment

Page 28 of 39

Page 29 of 39

Figure 10

Figure 11 0

capacity / mAh g

Discharge

-1

Cycle number 10

20

30

40

50

60

70

80

90

100

225 200 175 150 125 100 75

a)

50 100.0

Coulombic efficiency / %

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

BL 1 + 0.5 wt. % TPPO BL 1 + 0.5 wt. % TPP BL 1 + 0.5 wt. % VC BL 1 + 0.5 wt. % DPC BL 1

99.8 99.6 BL 1 + 0.5 wt. % TPPO BL 1 + 0.5 wt. % TPP BL 1 + 0.5 wt. % VC BL 1 + 0.5 wt. % DPC BL 1

99.4 99.2

b)

99.0 0

10

20

30

40

50

60

Cycle number

29 ACS Paragon Plus Environment

70

80

90

100

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 39

Table 1 Properties

Negative electrode 95.4 wt.% graphite; 2.5 wt.% SBR; 1.5 wt.% Super C65; 0.6 wt.% Na-CMC Water ≈2.6 mAh cm-2 ≈7.7 mg cm-2 Dendritic copper (20 µm) 30-35%

Composition

Processing solvent Average areal capacity Electrode loading Current collector Porosity

Positive electrode 93 wt.% NMC811; 4 wt.% Super C65; 3 wt.% PVdF

N-methyl-2-pyrrolidone ≈2.1-2.2 mAh cm-2 ≈ 10.6 mg cm-2 Aluminum (15 µm) 30-35%

Table 2 Electrolyte

Discharge capacity / mAh g-1 st 1 5th 100th cycle cycle cycle 162.9 155.5 135.1

1st cycle 70.5

5th cycle 99.4

100th cycle 99.9

86

BL 1 + TPPO

197.5

187.0

171.6

86.2

99.8

99.9

92

BL 2

143.1

135.8

118.3

62.4

99.2

99.86

86

BL 2 + TPPO

196.8

186.4

172.1

85.8

99.8

99.9

92

BL 1

Coulombic efficiency / %

Capacity retention [100th / 3th] / %

Table 3 Element region / %at F 1s

Mn 2p

O 1s

C 1s

P 2p

Li 1s

Pristine electrode

21.78

0.33

3.94

73.94

0

0

BL 1

23.16

0.22

5.22

67.33

0.24

3.83

BL 1 + TPPO

22.74

0.22

6.46

66.47

0.31

3.81

30 ACS Paragon Plus Environment

Page 31 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Table 4 Element region / %at F 1s

O 1s

C 1s

P 2p

Li 1s

Pristine electrode

0

3.46

96.54

0

0

BL 1

6.81

21.43

55.91

1.37

14.48

BL 1 + TPPO

11.00

18.69

48.28

1.33

20.71

Table 5 Electrolyte

BL 1 BL 1 + 0.5 wt. % TPPO BL 1 + 0.5 wt. % TPP BL 1 + 0.5 wt. % VC BL 1 + 0.5 wt. % DPC

Discharge capacity / mAh g-1 st 1 5th 100th cycle cycle cycle 162.9 155.5 135.1

Coulombic efficiency / % st th 1 5 100th cycle cycle cycle 70.5 99.4 99.9

197.5

187.0

171.6

86.2

99.8

99.9

92

200.2

188.8

164.7

84.3

99.5

99.8

87

188.7

181.3

158.2

82.3

99.4

99.9

86

159.7

151.6

123.1

69.4

99.3

99.8

80

31 ACS Paragon Plus Environment

Capacity retention / [100th / 3th] / %

86

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References 1.

Goodenough, J. B.; Kim, Y., Challenges for Rechargeable Li Batteries. Chem. Mater.

2010, 22, 587-603. 2.

Armand, M.; Tarascon, J. M., Building better batteries. Nature 2008, 451, 652-657.

3.

Blomgren, G. E., The Development and Future of Lithium Ion Batteries. J.

Electrochem. Soc. 2017, 164, A5019-A5025. 4.

Placke, T.; Kloepsch, R.; Dühnen, S.; Winter, M., Lithium ion, lithium metal, and

alternative rechargeable battery technologies: the odyssey for high energy density. J. Solid

State Electrochem. 2017, 21, 1939-1964. 5.

Andre, D.; Kim, S.-J.; Lamp, P.; Lux, S. F.; Maglia, F.; Paschos, O.; Stiaszny, B.,

Future generations of cathode materials: an automotive industry perspective. J. Mater. Chem.

A 2015, 3, 6709-6732. 6.

Wagner, R.; Preschitschek, N.; Passerini, S.; Leker, J.; Winter, M., Current research

trends and prospects among the various materials and designs used in lithium-based batteries.

J. Appl. Electrochem. 2013, 43, 481-496. 7.

Kasnatscheew, J.; Evertz, M.; Kloepsch, R.; Streipert, B.; Wagner, R.; Cekic

Laskovic, I.; Winter, M., Learning from Electrochemical Data: Simple Evaluation and Classification of LiMO2-type-based Positive Electrodes for Li-Ion Batteries. Energy Technol. 2017, 5, 1670-1679. 8.

Kasnatscheew, J.; Evertz, M.; Streipert, B.; Wagner, R.; Klopsch, R.; Vortmann, B.;

Hahn, H.; Nowak, S.; Amereller, M.; Gentschev, A. C.; Lamp, P.; Winter, M., The truth about the 1st cycle Coulombic efficiency of LiNi1/3Co1/3Mn1/3O2 (NCM) cathodes. Phys. Chem.

Chem. Phys. 2016, 18, 3956-3965. 9.

Kasnatscheew, J.; Evertz, M.; Streipert, B.; Wagner, R.; Nowak, S.; Cekic Laskovic,

I.; Winter, M., Changing Established Belief on Capacity Fade Mechanisms: Thorough Investigation of LiNi1/3Co1/3Mn1/3O2 (NCM111) under High Voltage Conditions. J. Phys.

Chem. C 2017, 121, 1521-1529. 10.

Evertz, M.; Horsthemke, F.; Kasnatscheew, J.; Borner, M.; Winter, M.; Nowak, S.,

Unraveling transition metal dissolution of Li1.04Ni1/3Co1/3Mn1/3O2 (NCM 111) in lithium ion full cells by using the total reflection X-ray fluorescence technique. J. Power Sources 2016, 329, 364-371. 11.

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 32 ACS Paragon Plus Environment

Page 32 of 39

Page 33 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

dissolution and capacity fading of NCM cathode material. Electrochim. Acta 2014, 134, 393398. 12.

Noh, H.-J.; Youn, S.; Yoon, C. S.; Sun, Y.-K., Comparison of the structural and

electrochemical properties of layered Li[NixCoyMnz]O2 (x = 1/3, 0.5, 0.6, 0.7, 0.8 and 0.85) cathode material for lithium-ion batteries. J. Power Sources 2013, 233, 121-130. 13.

Schipper, F.; Nayak, P.; Erickson, E.; Amalraj, S.; Srur-Lavi, O.; Penki, T.; Talianker,

M.; Grinblat, J.; Sclar, H.; Breuer, O.; Julien, C.; Munichandraiah, N.; Kovacheva, D.; Dixit, M.; Major, D.; Markovsky, B.; Aurbach, D., Study of Cathode Materials for Lithium-Ion Batteries: Recent Progress and New Challenges. Inorganics 2017, 5, 32-61. 14.

Myung, S.-T.; Maglia, F.; Park, K.-J.; Yoon, C. S.; Lamp, P.; Kim, S.-J.; Sun, Y.-K.,

Nickel-Rich Layered Cathode Materials for Automotive Lithium-Ion Batteries: Achievements and Perspectives. ACS Energy Lett. 2016, 196-223. 15.

Manthiram, A.; Song, B.; Li, W., A perspective on nickel-rich layered oxide cathodes

for lithium-ion batteries. Energy Storage Materials 2017, 6, 125-139. 16.

Ryu, H.-H.; Park, K.-J.; Yoon, C. S.; Sun, Y.-K., Capacity Fading of Ni-Rich

Li[NixCoyMn1–x–y]O2 (0.6 ≤ x ≤ 0.95) Cathodes for High-Energy-Density Lithium-Ion Batteries: Bulk or Surface Degradation? Chem. Mater. 2018, 30, 1155-1163. 17.

Schmuch, R.; Wagner, R.; Hörpel, G.; Placke, T.; Winter, M., Nat. Energy 2018, DOI:

10.1038/s41560-018-0107-2. 18.

Downie, L. E.; Hyatt, S. R.; Dahn, J. R., The Impact of Electrolyte Composition on

Parasitic Reactions in Lithium Ion Cells Charged to 4.7 V Determined Using Isothermal Microcalorimetry. J. Electrochem. Soc. 2016, 163, A35-A42. 19.

Jung, R.; Metzger, M.; Maglia, F.; Stinner, C.; Gasteiger, H. A., Chemical versus

Electrochemical Electrolyte Oxidation on NMC111, NMC622, NMC811, LNMO, and Conductive Carbon. J. Phys. Chem. Lett. 2017, 8, 4820-4825. 20.

Shi, H.; Wang, X.; Hou, P.; Zhou, E.; Guo, J.; Zhang, J.; Wang, D.; Guo, F.; Song, D.;

Shi, X.; Zhang, L., Core–shell structured Li[(Ni0.8Co0.1Mn0.1)0.7(Ni0.45Co0.1Mn0.45)0.3]O2 cathode material for high-energy lithium ion batteries. J. Alloys Compd. 2014, 587, 710-716. 21.

Li, J.; Downie, L. E.; Ma, L.; Qiu, W.; Dahn, J. R., Study of the Failure Mechanisms

of LiNi0.8Mn0.1Co0.1O2 Cathode Material for Lithium Ion Batteries. J. Electrochem. Soc. 2015, 162, A1401-A1408. 22.

Zuo, D.; Tian, G.; Li, X.; Chen, D.; Shu, K., Recent progress in surface coating of

cathode materials for lithium ion secondary batteries. J. Alloys Compd. 2017, 706, 24-40.

33 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

23.

Wang, D.; Li, X.; Wang, Z.; Guo, H.; Xu, Y.; Fan, Y.; Ru, J., Role of zirconium

dopant on the structure and high voltage electrochemical performances of LiNi0.5Co0.2Mn0.3O2 cathode materials for lithium ion batteries. Electrochim. Acta 2016, 188, 48-56. 24.

Xu, K., Nonaqueous liquid electrolytes for lithium-based rechargeable batteries.

Chem. Rev. 2004, 104, 4303-4417. 25.

Xu, K., Electrolytes and interphases in Li-ion batteries and beyond. Chem. Rev. 2014,

114, 11503-11618. 26.

Schmitz, R. W.; Murmann, P.; Schmitz, R.; Müller, R.; Krämer, L.; Kasnatscheew, J.;

Isken, P.; Niehoff, P.; Nowak, S.; Röschenthaler, G.-V.; Ignatiev, N.; Sartori, P.; Passerini, S.; Kunze, M.; Lex-Balducci, A.; Schreiner, C.; Cekic-Laskovic, I.; Winter, M., Investigations on novel electrolytes, solvents and SEI additives for use in lithium-ion batteries: Systematic electrochemical characterization and detailed analysis by spectroscopic methods. Prog. Solid

State Chem. 2014, 42, 65-84. 27.

Cekic-Laskovic, I.; von Aspern, N.; Imholt, L.; Kaymaksiz, S.; Oldiges, K.; Rad, B.

R.; Winter, M., Synergistic Effect of Blended Components in Nonaqueous Electrolytes for Lithium Ion Batteries. Top. Curr. Chem. 2017, 375, 37-101. 28.

Zhang, S. S., A review on electrolyte additives for lithium-ion batteries. J. Power

Sources 2006, 162, 1379-1394. 29.

Winter, M., The Solid Electrolyte Interphase - The Most Important and the Least

Understood Solid Electrolyte in Rechargeable Li Batteries. Z. Phys. Chem. 2009, 223, 13951406. 30.

Xu, K.; von Cresce, A., Interfacing electrolytes with electrodes in Li ion batteries. J.

Mater. Chem. 2011, 21, 9849-9864. 31.

Haregewoin, A. M.; Wotango, A. S.; Hwang, B.-J., Electrolyte additives for lithium

ion battery electrodes: progress and perspectives. Energy Environ. Sci. 2016, 9, 1955-1988. 32.

Streipert, B.; Janßen, P.; Cao, X.; Kasnatscheew, J.; Wagner, R.; Cekic-Laskovic, I.;

Winter, M.; Placke, T., Evaluation of Allylboronic Acid Pinacol Ester as Effective Shutdown Overcharge Additive for Lithium Ion Cells. J. Electrochem. Soc. 2017, 164, A168-A172. 33.

Vogl, U.; Schmitz, A.; Stock, C.; Pablo Badillo, J.; Jakob Gores, H.; Winter, M.,

Investigation of N-Ethyl-2-Pyrrolidone (NEP) as Electrolyte Additive in Regard to Overcharge Protecting Characteristics. J. Electrochem. Soc. 2014, 161, A1407-A1414. 34.

Tornheim, A.; He, M.; Su, C.-C.; Zhang, Z., The Role of Additives in Improving

Performance in High Voltage Lithium-Ion Batteries with Potentiostatic Holds. J.

Electrochem. Soc. 2017, 164, A6366-A6372. 34 ACS Paragon Plus Environment

Page 34 of 39

Page 35 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

35.

Tu,

W.;

Ye,

C.;

Yang,

X.;

Xing,

L.;

Liao,

Y.;

Liu,

X.;

Li,

W.,

Trimethylsilylcyclopentadiene as a novel electrolyte additive for high temperature application of lithium nickel manganese oxide cathode. J. Power Sources 2017, 364, 23-32. 36.

von Aspern, N.; Röser, S.; Rezaei Rad, B.; Murmann, P.; Streipert, B.; Mönnighoff,

X.; Tillmann, S. D.; Shevchuk, M.; Stubbmann-Kazakova, O.; Röschenthaler, G.-V.; Nowak, S.; Winter, M.; Cekic-Laskovic, I., Phosphorus additives for improving high voltage stability and safety of lithium ion batteries. J. Fluorine Chem. 2017, 198, 24-33. 37.

Wagner, R.; Kraft, V.; Streipert, B.; Kasnatscheew, J.; Gallus, D. R.; Amereller, M.;

Korth, M.; Cekic-Laskovic, I.; Winter, M., Magnesium-based additives for the cathode slurry to enable high voltage application of lithium-ion batteries. Electrochim. Acta 2017, 228, 9-17. 38.

Qi, X.; Tao, L.; Hahn, H.; Schultz, C.; Gallus, D. R.; Cao, X.; Nowak, S.; Röser, S.;

Li, J.; Cekic-Laskovic, I.; Rad, B. R.; Winter, M., Lifetime limit of tris(trimethylsilyl) phosphite as electrolyte additive for high voltage lithium ion batteries. RSC Adv. 2016, 6, 38342-38349. 39.

Jang, S. H.; Yim, T., Effect of Silyl Ether-functinoalized Dimethoxydimethylsilane on

Electrochemical Performance of a Ni-rich NCM Cathode. ChemPhysChem 2017, 18, 34023406. 40.

Milien, M. S.; Tottempudi, U.; Son, M.; Ue, M.; Lucht, B. L., Development of

Lithium Dimethyl Phosphate as an Electrolyte Additive for Lithium Ion Batteries. J.

Electrochem. Soc. 2016, 163, A1369-A1372. 41.

Rong, H.; Xu, M.; Zhu, Y.; Xie, B.; Lin, H.; Liao, Y.; Xing, L.; Li, W., A novel

imidazole-based electrolyte additive for improved electrochemical performance of high voltage nickel-rich cathode coupled with graphite anode lithium ion battery. J. Power Sources 2016, 332, 312-321. 42.

Zhou, Z.; Ma, Y.; Wang, L.; Zuo, P.; Cheng, X.; Du, C.; Yin, G.; Gao, Y., Triphenyl

phosphite as an electrolyte additive to improve the cyclic stability of lithium-rich layered oxide cathode for lithium-ion batteries. Electrochim. Acta 2016, 216, 44-50. 43.

Xu, C.; Renault, S.; Ebadi, M.; Wang, Z.; Björklund, E.; Guyomard, D.; Brandell, D.;

Edström, K.; Gustafsson, T., LiTDI: A Highly Efficient Additive for Electrolyte Stabilization in Lithium-Ion Batteries. Chem. Mater. 2017, 29, 2254-2263. 44.

Röser, S.; Lerchen, A.; Ibing, L.; Cao, X.; Kasnatscheew, J.; Glorius, F.; Winter, M.;

Wagner, R., Highly Effective Solid Electrolyte Interphase-Forming Electrolyte Additive Enabling High Voltage Lithium-Ion Batteries. Chem. Mater. 2017, 29, 7733-7739.

35 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

45.

Page 36 of 39

Kasnatscheew, J.; Placke, T.; Streipert, B.; Rothermel, S.; Wagner, R.; Meister, P.;

Laskovic, I. C.; Winter, M., A Tutorial into Practical Capacity and Mass Balancing of Lithium Ion Batteries. J. Electrochem. Soc. 2017, 164, A2479-A2486. 46.

Qiu, W.; Xia, J.; Chen, L.; Dahn, J. R., A study of methyl phenyl carbonate and

diphenyl carbonate as electrolyte additives for high voltage LiNi0.8Mn0.1Co0.1O2/graphite pouch cells. J. Power Sources 2016, 318, 228-234. 47.

Li, J.; Liu, H.; Xia, J.; Cameron, A. R.; Nie, M.; Botton, G. A.; Dahn, J. R., The

Impact of Electrolyte Additives and Upper Cut-off Voltage on the Formation of a Rocksalt Surface Layer in LiNi0.8Mn0.1Co0.1O2 Electrodes. J. Electrochem. Soc. 2017, 164, A655A665. 48.

Yim, T.; Kang, K. S.; Mun, J.; Lim, S. H.; Woo, S.-G.; Kim, K. J.; Park, M.-S.; Cho,

W.; Song, J. H.; Han, Y.-K.; Yu, J.-S.; Kim, Y.-J., Understanding the effects of a multifunctionalized additive on the cathode–electrolyte interfacial stability of Ni-rich materials. J.

Power Sources 2016, 302, 431-438. 49.

https://www.sigmaaldrich.com/catalog/product/aldrich/t84603?lang=de®ion=DE,

Accessed 9th November, 2017. 50.

Niehoff, P.; Winter, M., Composition and Growth Behavior of the Surface and

Electrolyte

Decomposition

Layer

of/on

a

Commercial

Lithium

Ion

Battery

LixNi1/3Mn1/3Co1/3O2 Cathode Determined by Sputter Depth Profile X-ray Photoelectron Spectroscopy. Langmuir 2013, 29, 15813-15821. 51.

Niehoff, P.; Passerini, S.; Winter, M., Interface investigations of a commercial lithium

ion battery graphite anode material by sputter depth profile X-ray photoelectron spectroscopy.

Langmuir 2013, 29, 5806-5816. 52.

Michalak, B.; Berkes, B. B.; Sommer, H.; Brezesinski, T.; Janek, J., Electrochemical

Cross-Talk Leading to Gas Evolution and Capacity Fade in LiNi0.5Mn1.5O4/Graphite FullCells. J. Phys. Chem. C 2017, 121, 211-216. 53.

Krueger, S.; Kloepsch, R.; Li, J.; Nowak, S.; Passerini, S.; Winter, M., How Do

Reactions at the Anode/Electrolyte Interface Determine the Cathode Performance in LithiumIon Batteries? J. Electrochem. Soc. 2013, 160, A542-A548. 54.

Kasnatscheew, J.; Streipert, B.; Roser, S.; Wagner, R.; Cekic Laskovic, I.; Winter, M.,

Determining oxidative stability of battery electrolytes: validity of common electrochemical stability window (ESW) data and alternative strategies. Phys. Chem. Chem. Phys. 2017, 19, 16078-16086.

36 ACS Paragon Plus Environment

Page 37 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

55.

Winter, M.; Appel, W. K.; Evers, B.; Hodal, T.; Moller, K. C.; Schneider, I.; Wachtler,

M.; Wagner, M. R.; Wrodnigg, G. H.; Besenhard, J. O., Studies on the anode/electrolyte interface in lithium ion batteries. Chem. Monthly 2001, 132, 473-486. 56.

Placke, T.; Siozios, V.; Rothermel, S.; Meister, P.; Colle, C.; Winter, M., Assessment

of Surface Heterogeneity: a Route to Correlate and Quantify the 1st Cycle Irreversible Capacity Caused by SEI Formation to the Various Surfaces of Graphite Anodes for Lithium Ion Cells. Z. Phys. Chem. 2015, 229, 1451-1470. 57.

Placke, T.; Siozios, V.; Schmitz, R.; Lux, S. F.; Bieker, P.; Colle, C.; Meyer, H. W.;

Passerini, S.; Winter, M., Influence of graphite surface modifications on the ratio of basal plane to "non-basal plane" surface area and on the anode performance in lithium ion batteries.

J. Power Sources 2012, 200, 83-91. 58.

Olivier, J. P.; Winter, M., Determination of the absolute and relative extents of basal

plane surface area and "non-basal plane surface" area of graphites and their impact on anode performance in lithium ion batteries. J. Power Sources 2001, 97-8, 151-155. 59.

Kohs, W.; Santner, H. J.; Hofer, F.; Schrottner, H.; Doninger, J.; Barsukov, I.; Buqa,

H.; Albering, J. H.; Moller, K. C.; Besenhard, J. O.; Winter, M., A study on electrolyte interactions with graphite anodes exhibiting structures with various amounts of rhombohedral phase. J. Power Sources 2003, 119, 528-537. 60.

Schwieters, T.; Evertz, M.; Mense, M.; Winter, M.; Nowak, S., Lithium loss in the

solid electrolyte interphase: Lithium quantification of aged lithium ion battery graphite electrodes by means of laser ablation inductively coupled plasma mass spectrometry and inductively coupled plasma optical emission spectroscopy. J. Power Sources 2017, 356, 4755. 61.

Holtstiege, F.; Wilken, A.; Winter, M.; Placke, T., Running out of lithium? A route to

differentiate between capacity losses and active lithium losses in lithium-ion batteries. Phys.

Chem. Chem. Phys. 2017, 19, 25905-25918. 62.

Smith, A. J.; Burns, J. C.; Zhao, X.; Xiong, D.; Dahn, J. R., A High Precision

Coulometry Study of the SEI Growth in Li/Graphite Cells. J. Electrochem. Soc. 2011, 158, A447-A452. 63.

Smith, A. J.; Burns, J. C.; Dahn, J. R., A High Precision Study of the Coulombic

Efficiency of Li-Ion Batteries. Electrochem. Solid-State Lett. 2010, 13, A177-A179. 64.

Smith, A. J.; Burns, J. C.; Xiong, D.; Dahn, J. R., Interpreting High Precision

Coulometry Results on Li-ion Cells. J. Electrochem. Soc. 2011, 158, A1136-A1142.

37 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

65.

Vortmann-Westhoven, B.; Winter, M.; Nowak, S., Where is the lithium? Quantitative

determination of the lithium distribution in lithium ion battery cells: Investigations on the influence of the temperature, the C-rate and the cell type. J. Power Sources 2017, 346, 63-70. 66.

Andersson, A. M.; Abraham, D. P.; Haasch, R.; MacLaren, S.; Liu, J.; Amine, K.,

Surface Characterization of Electrodes from High Power Lithium-Ion Batteries. J.

Electrochem. Soc. 2002, 149, A1358-A1369. 67.

Edström, K.; Gustafsson, T.; Thomas, J. O., The cathode-electrolyte interface in the

Li-ion battery. Electrochim. Acta 2004, 50, 397-403. 68.

Li, J.; Xing, L.; Zhang, R.; Chen, M.; Wang, Z.; Xu, M.; Li, W.,

Tris(trimethylsilyl)borate as an electrolyte additive for improving interfacial stability of high voltage layered lithium-rich oxide cathode/carbonate-based electrolyte. J. Power Sources 2015, 285, 360-366. 69.

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-A97. 70.

Nurpeissova, A.; Park, D. I.; Kim, S. S.; Sun, Y. K., Epicyanohydrin as an Interface

Stabilizer Agent for Cathodes of Li-Ion Batteries. J. Electrochem. Soc. 2015, 163, A171A177. 71.

Wang, Z.; Xing, L.; Li, J.; Xu, M.; Li, W., Triethylborate as an electrolyte additive for

high voltage layered lithium nickel cobalt manganese oxide cathode of lithium ion battery. J.

Power Sources 2016, 307, 587-592. 72.

Kusachi, Y.; Zhang, Z.; Dong, J.; Amine, K., Electrode Surface Film Formation in

Tris(ethylene glycol)-Substituted Trimethylsilane–Lithium Bis(oxalate)borate Electrolyte. J.

Phys. Chem. C 2011, 115, 24013-24020. 73.

Bouayad, H.; Wang, Z.; Dupré, N.; Dedryvère, R.; Foix, D.; Franger, S.; Martin, J. F.;

Boutafa, L.; Patoux, S.; Gonbeau, D.; Guyomard, D., Improvement of Electrode/Electrolyte Interfaces in High-Voltage Spinel Lithium-Ion Batteries by Using Glutaric Anhydride as Electrolyte Additive. J. Phys. Chem. C 2014, 118, 4634-4648. 74.

Xiao, A.; Yang, L.; Lucht, B. L.; Kang, S.-H.; Abraham, D. P., Examining the Solid

Electrolyte Interphase on Binder-Free Graphite Electrodes. J. Electrochem. Soc. 2009, 156, A318-A327. 75.

Wagner, R.; Korth, M.; Streipert, B.; Kasnatscheew, J.; Gallus, D. R.; Brox, S.;

Amereller, M.; Cekic-Laskovic, I.; Winter, M., Impact of Selected LiPF6 Hydrolysis Products

38 ACS Paragon Plus Environment

Page 38 of 39

Page 39 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

on the High Voltage Stability of Lithium-Ion Battery Cells. ACS Appl. Mater. Interfaces 2016, 8, 30871-30878. 76.

Gilbert, J. A.; Bareño, J.; Spila, T.; Trask, S. E.; Miller, D. J.; Polzin, B. J.; Jansen, A.

N.; Abraham, D. P., Cycling Behavior of NCM523/Graphite Lithium-Ion Cells in the 3–4.4 V Range: Diagnostic Studies of Full Cells and Harvested Electrodes. J. Electrochem. Soc. 2017, 164, A6054-A6065. 77.

Dedryvère, R.; Martinez, H.; Leroy, S.; Lemordant, D.; Bonhomme, F.; Biensan, P.;

Gonbeau, D., Surface film formation on electrodes in a LiCoO2/graphite cell: A step by step XPS study. J. Power Sources 2007, 174, 462-468. 78.

https://www.sigmaaldrich.com/catalog/product/aldrich/291250?lang=de®ion=DE,

Accessed 9th November, 2017.

39 ACS Paragon Plus Environment