You are on page 1of 12

Journal of Energy Chemistry 64 (2022) 451–462

Contents lists available at ScienceDirect

Journal of Energy Chemistry


journal homepage: www.elsevier.com/locate/jechem

Anion effect on Li/Na/K hybrid electrolytes for Graphite//NCA


(LiNi0.8Co0.15Al0.05O2) Li-ion batteries
Aiman Jrondi a, Georgios Nikiforidis a,b, Mérièm Anouti a,b,⇑
a
Laboratoire PCM2E, Université de Tours, Faculté des sciences et technologie, Parc de Grandmont, 37200 Tours, France
b
LE STUDIUM Institute for Advanced Studies, 45000 Orléans, France

a r t i c l e i n f o a b s t r a c t

Article history: The electrolyte is an essential component of a battery system since it is responsible for the conduction of
Received 23 March 2021 ions between the electrodes. In the quest for cheaper alternatives to common organic electrolytes for
Revised 1 May 2021 lithium-ion batteries (LIB), we formulated hybrid electrolytes comprising a mixture of Na, K, and Li alka-
Accepted 4 May 2021
line salts with ethylene carbonate (EC), ethyl methyl carbonate (EMC), and lithium hexafluorophosphate
Available online 14 May 2021
(LiPF6), giving a total salt concentration of 1.5 M; we determined their physicochemical properties and
investigated their electrochemical behavior on a nickel cobalt aluminum oxide (NCA) cathode and gra-
Keywords:
phite (Gr) anode. The electrolytes demonstrated a melting transition peak (Tm), eutectic behavior, and
Hybrid electrolyte
Alkali salts
ionic conductivities (~13 mS cm1) close to those of a commercial LIB electrolyte (SE, EC/EMC + 1 M
NCA//Li cell LiPF6) and activation energies of ca. 3 kJ mol1. The half-cell coin cells revealed high coulombic efficiency
Li-ion battery (99%), specific capacity (175 mAh g1 at C/10), and capacity retention (92% for NaCF3SO3) for the NCA
cathode and a moderate performance (coulombic efficiency of 98% for 20 cycles) on the graphite anode
after the formation of the SEI layer. The hybrid electrolytes were cycled at 25 °C in a Gr//NCA cell yielding
specific capacities of ca. 225 mAh g1 at a C/5 rate, corroborating that the anion plays a key role and high-
lighting their potential for energy storage applications.
Ó 2021 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published
by ELSEVIER B.V. and Science Press. All rights reserved.

1. Introduction affects the safety and lifespan of the batteries, and the high cost
of the electrodes’ raw materials (e.g., alkali (Li) and transition met-
The difficulty of integrating renewable energy, such as wind and als (Co, Ni)) [9].
solar, into the power grid is that it is unpredictable and intermit- In this vein, advances in material and electrolyte design are
tent. To compensate for its uncertainty and fluctuation, and to indispensable. Batteries consist of three main components: anode
modernize the current grid, battery energy storage systems (BESS) (negative electrode), cathode (positive electrode), and electrolyte.
are pivotal, as they allow the energy output to be stored and dis- The anode material of choice is lamellar graphite, with a high
tributed upon demand. The energy can then be used for a wide reversible capacity (372 mAh g1) and a low volume expansion
range of applications, including transportation, portable, and sta- (<10%) [10]. At high operating voltages, solvent reduction and inor-
tionary backup power [1]. The prevalent battery technology for ganic salt decomposition occur at the anode-electrolyte interface,
such applications is based on lithium-ion chemistry; lithium-ion resulting in the formation of a surface film, commonly referred to
batteries (LIBs) have a long-cycle life (>500 cycles), high energy as the solid electrolyte interface (SEI) [11]. This passivation layer
density (up to 300 Wh kg-1), and moderate cost (US$127 per is critical for battery longevity and performance as it allows
kWh) [2]. Nonetheless, they require further development to sus- lithium ions access to graphite while acting as a barrier for further
tain their advantage against other electrochemical energy storage electrolyte reduction [12].
systems (i.e., supercapacitors [3], Na-ion [4,5], NaS [6], LiS [7], At the cathode, layered transition metal oxides are adopted in
and redox flow batteries [8]), lower their cost and improve their electric vehicle batteries (i.e., Tesla [6]), portable electronics [13],
efficiency and safety. The major hurdles of LIBs are ascribed to and stationary power [14], namely NCA (Li(Ni1xyCoxAly)O2) and
the degradation of the electrolyte and electrode materials, which NMC (Li1+a(NixMnyCoz)1aO2), with working potentials reaching
up to 5 V vs. Li/Li+ [11,15–17]. These materials exhibit good elec-
⇑ Corresponding author. trochemical performance (i.e., reversible specific capacities nearing
E-mail address: meriem.anouti@univ-tours.fr (M. Anouti). 200 mAh g1) on account of their enhanced structure crystallinity

https://doi.org/10.1016/j.jechem.2021.05.004
2095-4956/Ó 2021 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published by ELSEVIER B.V. and Science Press. All rights reserved.
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

and, in turn, improved cyclability [18–20]. Nevertheless, degrada- via the selection of intrinsically stable components, or kinetically
tion phenomena in nickel-rich cathodes remain an issue when they via the passivation layer formation in the first cycles. We anticipate
are exposed to high oxidation potentials (i.e., >4.9 V). Such events that the addition of Na+ and K+ will better polarize the interface at
affect the cycling stability, inducing gradual phase transformations high potentials, as they are more resistant to oxidation than Li+.
(i.e., layered R3̅m to rock-salt Fm3̅m), transition metal (TM) disso- With the smaller desolvation energies of Na and K, we expect a
lution, and electrolyte reactions with TMs in high oxidation states beneficial kinetic effect on the electrode–electrolyte interface’s
(e.g., electrolyte oxidation with highly active Ni4+) [21]. Electrolyte charge transfer reactions. This assumption is supported by previ-
oxidation reactions at the cathode-electrolyte interface (CEI) [22] ous work. For instance, a study on propylene carbonate revealed
have also been linked to cell impedance increase and performance that despite the larger ionic radius of Na+ (1.02 Å) and K+ (1.38
deterioration [23]. Structural changes due to rock-salt phase trans- Å) than that of Li+’s (0.76 Å), the former delivered smaller Strokes’
formation have been known to release oxygen, exacerbating these radii because of their weaker Lewis acidity, which enables them to
oxidation reactions at the cathode surface [24]. Therefore, surface exhibit faster mobility and smaller ionic diffusion activation
modification to produce self-limiting CEIs by chemical coating of energy in the electrolyte [40]. In addition, it has been proven that
NCA particles (e.g., TiO2, MnO2, ZrO2) has been widely adopted as the diffusion coefficient of K+ is about three times larger than that
an efficient method to prevent and suppress side reactions with of Li+, as investigated by ab initio molecular dynamics simulations
the electrolyte [25,26]. [41]. What’s more, a fluorinated LiF/NaF hybrid interphase derived
Alternatively, the understanding and enhancement of the from an electrolyte-soaking approach exhibited improved ionic
oxidative stability of electrolyte solvents as a function of elec- conductivity, high mechanical strength, and extended dendrite-
trolyte composition (e.g., salt concentration) has been the subject free cycling of metallic Li in Li symmetrical cells [42].
of numerous studies [27,28] as an approach to hamper such reac- In this study, we investigated the comparative effect of Li+, Na+,
tions. Electrolytes are integral components of batteries; their con- and K+ cations and their counter-anions on the physicochemical
ductivities and electrochemical windows dictate the extent of behavior of the electrolytes. The anions selected in the alkali salts
operating potentials, electrode utilization, and power density of were tetrafluoroborate (BF-4), trifluoromethanesulfonate (CF3SO-3),
the batteries, hence their capacities, charging rates, and perfor- and methanesulfonate (CH3SO-3). BF-4 is not uncommon in LIBs; it
mance. The standard composition adopted by most commercial has often been implemented as an additive [43,44], while CF3SO-3
LIBs includes ethyl carbonate (EC) in conjunction with a linear and CH3SO-3 anions have been used in both Li and other battery sys-
alkyl carbonate or ester to lower the viscosity of the solution tems (i.e., zinc and vanadium redox flow batteries [45–47]). We
and, finally, a lithium salt, commonly 1 M lithium hexafluorophos- first determined the ionic conductivity and thermal stability of
phate (LiPF6) [15]. The concentration range of LiPF6 is optimized to the electrolytes. Then, we examined the cycling behavior of
ensure good thermal stability, transport properties, and electro- NCA//Li, Gr//Li half-cells, and the Gr//NCA cell with SE and HE at
chemical performance for alkyl carbonate salt solvent mixtures 25 °C by cyclic voltammetry and galvanostatic charge–discharge
at 1 M < C < 1.5 M [29–31]. Searching for alternatives to common cycling.
organic electrolytes, and taking into account the fact that their
oxidative stability depends on the coupling between anion and sol- 2. Experimental
vents, herein we propose a series of hybrid electrolytes (HE) com-
prising a mixture of alkaline salts (e.g., Li, Na, K) at a total salt 2.1. Chemical compounds and electrolyte formulation
concentration of 1.5 M and compare them against a standard com-
mercial electrolyte (SE) composed of ethylene carbonate (EC): All electrolyte components were battery-grade and used as
ethyl methyl carbonate (EMC) 1:1 wt/wt and 1 M LiPF6. Sodium received. Ethylene carbonate, EC, (99.9%), and ethyl methyl carbon-
ions (Na+, 2.71 vs. SHE for Na/Na+) and potassium ions (K+, ate, EMC, (99.9%) were purchased from Sigma Aldrich. Lithium
2.93 vs. SHE for K/K+) have a lower reduction potential than hexafluorophosphate salt (LiPF6, Sigma Aldrich, battery grade),
lithium ions (Li+, 3.04 vs. SHE for Li/Li+), yet, they are widely used KBF4 (Sigma Aldrich, 99.99%), NaBF4 (Sigma Aldrich, >98%), NaCF3-
in electrochemical energy storage devices due to their abundance SO3 (Sigma Aldrich, 98%), and KCH3SO3 (Sigma Aldrich, >98%) were
and low cost [1,32]. The addition of Na+ and K+ cations in Li- used without prior treatment. KCF3SO3 and NaCH3SO3 salts were
based electrolytes has been demonstrated before and does not formulated by equimolar acid-base neutralization of the corre-
pose an issue at low reduction potentials due to the formation of sponding acids with concentrated aqueous KOH or NaOH solutions.
the SEI layer (or on the lithium metal [33–35]). The SEI (protective The obtained salts were then dried to eliminate residual water and
film on the graphite or Li) constitutes a barrier to the subsequent stored in a glovebox.
degradation of Na+ cations and solvents. In alkylcarbonate based For the standard electrolyte (SE), a binary solvent mixture of EC/
electrolytes, the potential of solvated Na+ and K+ ions shift depend- EMC with a mass ratio of 1/1 was formulated in the glove box at
ing on the nature of the Li-(SEI) surface quality and electrolyte 25 °C (MBraun) and mixed with 1.0 M LiPF6. The quantity of alkali
composition. In-situ Tof-SIMS (i.e., Time-of-Flight Secondary Ion salt (i.e., LiPF6, NaBF4, NaCH3SO3, NaCF3SO3, KBF4, KCH3SO3 and
Mass Spectrometry) from Stark et al. [36] has shown that Na blocks KCF3SO3) added in the EC/EMC + 1 M LiPF6, standard electrolyte
the accelerated dendrite growth and yields a dimpled and (SE), was 0.5 M. For simplification, we will identify the electrolytes
dendrite-free morphology. Xu et al also demonstrated that the formulated by the added salt in the following manner: SE + 0.5 M
addition of Na+ to a carbonate-based electrolyte (i.e., 1 mol L-1 KBF4 is noted as ‘‘KBF004 and SE + 0.5 M LiPF6, as ”LiPF006 . The water
lithium difluoro(oxalato)borate/ethylene carbonate/dimethyl car- content of each solution was less than 20 ppm, determined by
bonate) could reduce electrode polarization and favor Li deposi- using an 831 Karl-Fisher Coulometer (Metrohm). The electrolytes
tion/stripping cycling [37]. were prepared using a Sartorius 1602 MP balance with ± 10-4 g
The effect of anion composition on the lithium salts is high- accuracy inside an argon-filled MBraun glovebox, at 25 °C, with
lighted in its importance in determining the lithium transference less than 5 ppm moisture content.
number and ionic conductivities [38]. The anion radius affects
the transference number and promotes or demotes the migration 2.2. Physicochemical characterization
of Li+ in the carbonate electrolyte [39].
Electrochemical stability under high voltage conditions can be The ionic conductivity of the electrolytes was measured
achieved either thermodynamically, through electrolyte design from 20 °C to 100 °C with a 10-slot (WTSH 10) multichannel
452
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

conductivity meter (BioLogic), based on a frequency response ana- investigated anions were calculated using Density Functional The-
lyzer (MCM 10), and connected to a Peltier-based cooling system. ory (DFT) with the B3LYP functional- and 6–311 + G(d) basis set
Simultaneous measurements were performed using sealed cells (Mopac software).
with Pt parallel electrodes protecting the samples from air expo-
sure. The conductometer was calibrated using standard KCl solu-
tions. The thermal stability of electrolyte solutions was
3. Results and discussion
investigated by differential scanning calorimetry (DSC) using a
Perkin-Elmer DSC-4000 at a heating rate of 5 °C min1 from
3.1. Electrolyte formulation and characterization
60° C to 250 °C under nitrogen flow (20 ml min1).
Given that the proposed anions are associated with alkali met-
2.3. Electrochemical characterization als, they can be either solvated or polarized. They react in a differ-
ent way subject to the extent of the association. Here, we mix two
The half-cell tests were performed in standard CR2032-type different salts (AB and XY) with two common linear carbonate sol-
coin cells, where the active mass (NCA or graphite) was used as vents (EC, EMC). For instance, if AB = LiPF6 and XY = KCH3SO3, we
the working electrode, and a piece of lithium metal foil (0.5 mm obtain four types of associations (AB, AY, BX, and XY) and eight
thick, Alfa Aesar) was the counter and reference electrode. Glass possible solvation complexes (A(EC)n, A(EMC)n, X(EC)n, X(EMC)n,.
microfiber filter paper (WhatmanTM GF/C) was employed as the Y(EMC)n, Y(EC)n, B(EMC)n, and Y(EC)n,). The formation of quater-
separator. Full-cell tests were performed in CR2032-type coin cells nary systems (two salts with two solvents) is highly complex and
with graphite and NCA as working and counter electrodes, respec- can lead to high conductivities and improved electrochemical
tively. All the cells were assembled in an Ar-filled glove box behavior.
(MBraun, H2O < 0.1 ppm, O2 < 0.1 ppm) and sealed by a press Table 1 displays the Van der Waals (VdW) volume, partial
(Shenzhen Teensky Technology Co., Shenzhen, China). All elec- charge, electrostatic potential, and HOMO, LUMO energy levels of
trodes and separators were dried under vacuum at 80 °C for 12 h the anions investigated in this study. The VdW volume is defined
and placed in the glovebox without exposure to ambient air. The as the space occupied by a molecule, which is impenetrable to
NCA and graphite electrodes were acquired from TCI chemicals. other molecules with normal thermal energies [48]. Here the low-
The weight ratios of conductive carbon black (CB), Polyvinylidene est VdW volume is reported for BF4. The low (red) and high (blue)
fluoride (PVDF), and active material were 90:5:5 for graphite electrostatic potentials indicate the electron-rich and deficient
(porosity of 43%, a surface mass of 3.45 mg cm2, and 96.25% of regions. The F atom, due to its electronegativity and weak polarity
active material) and 80:10:10 for NCA (viz. active mass of 11 mg/ stemming from the presence of seven electrons in its outermost
cm2). The graphite electrodes were equilibrated with NCA elec- electron orbital [49], withdraws electrons in the vicinity; still, it
trodes (N/P = 1.2). does not contribute to the highest occupied molecular orbital
Cyclic voltammetry (CV) and electrochemical impedance spec- (HOMO) and lowest occupied molecular orbital (LUMO) due to
troscopy (1 MHz-10 Hz, 10 mV amplitude) curves were performed the strong C-F bond that induces localized electronic states ener-
on a VMP3 potentiostat (Biologic Science Instrument, France). getically far below the delocalized HOMO [50]. CF3SO-3 and CH3SO-3
The galvanostatic charge/discharge was measured using a VMP display similar properties (e.g., HOMO, LUMO, and VdW) differing
multichannel potentiostat (Biologic Science Instrument, France). from those of BF-4, the latter exhibiting a lower LUMO energy level,
Finally, the spatial distributions of electrostatic potentials of the indicative of a high electron affinity. Ab initio calculations of

Table 1
Structure, Van der Waals volume (VdW), dipolar moment, and partial negative charge of the anions under study.

Anions CH3SO-3 CF3SO-3 BF-4 PF-6


Structure

VDW volume (Å)3 69.4 85.0 49.9 74.9


Electrostatic potential

HOMO, LUMO orbitals HOMO: 0.58 HOMO: 0.52 HOMO: 0.66 HOMO: 0.66
(Hartree) LUMO: 0.48 LUMO: 0.49 LUMO: 0.02 LUMO: 0.18
Dipolar moment l = 4.509 l = 4.865 apolar apolar
(Debye)
Partial negative charge of anion 0.576 0.433 0.239 0.479

Quantum chemical calculations were performed using the Gaussian 09 W package. Orbital molecular optimization for HOMO LUMO was obtained with B3PW91/6-31G (d)
and calculated with a larger basis set 6–311++G (d, p).

453
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

halogens and BF-4 as their super halogen building block have conductivities in the presence of Na+. The polarity of the species
revealed high electron affinities[51]. plays no role in electrolytic media where the induced dipoles, by
solvation (or polarized interfaces such as electrodes), are more
relevant.
3.1.1. Ionic conductivity of electrolytes Fig. 2 compares the r of the hybrid electrolytes based on the Li/
Ionic conductivity (r) is the property most prone to optimiza- K-salt (Fig. 2a) and Li/Na-salt (Fig. 2b) within a 120 °C temperature
tion during the formulation of electrolytes. Fig. 1 displays a com- range (viz. 20 to 100 °C). The SE exhibits an ionic conductivity of
parison of the conductivities of the hybrid electrolytes with the 22.34 mS cm1 at 100 °C, while ‘‘KCF3SO003 and ”NaCH3SO003 conduc-
standard electrolyte at 60 °C. The hybrid electrolytes SE + 0.5 M tivities are ca. 16.60 and 14.89 mS cm1, respectively. The 33% salt
KBF4 (‘‘KBF004 ), SE + 0.5 M NaCH3SO3 (”NaCH3SO003 ) and SE + 0.5 M increase, either Na or K, leads to cluster formation of the alkali ions
NaBF4 (‘‘NaBF004 ) show comparable r to SE (i.e., ~14 mS cm1) at present with the EC/EMC molecules and decreases the electrolyte
60 °C, with the lowest recorded conductivity for the NaCF3SO3 salt, conductivity at high temperatures. At low temperatures (viz.
reaching 10 mS cm1. Since the conductivity is associated with the 20 °C), the hybrid electrolytes yield similar r values to the SE,
anions’ charge, BF-4, being the least charged anion, exhibits the with the relative deviation of the conductivity is less than 5%
most conducting properties in the presence of the most polarizable between 20 and 5 °C. The trend of increased ionic conductivities
cation (K+) at 60 °C followed by the standard electrolyte (Fig. 1). In with elevated temperatures for cyclic and linear carbonate binary
contrast, the mobility of large anions such as CF3SO-3 is more slug- solutions with Li salt reported in the literature [52] is also evi-
gish, as denoted by the VdW volume (Table 1), and leads to low denced here.
The plots of Fig. 2 adopt a non-Arrhenius behavior and are mod-
eled by the Vogel-Tamman-Fulcher model (VTF, Fig. 3 a-d) [53],
yielding a goodness of fit (R2) of ca. 0.999 for the HEs. Table 2 sum-
marizes r and the ideal glass transition temperature, T0, which is
defined as the temperature where the ion mobility tends to zero.
The activation energy, Ea, represented by parameter B, is related
to the free volume required for Li+ ions transportation through
the electrolyte and corresponds to the slope of the VTF plots
[54]. T0 ranged between 155 and 187 K at 80 °C. The activation
energy measurements exhibit a maximum deviation of
0.7 kJ mol1 from the standard electrolyte. A low Ea value of ca.
2.5 KJ mol1 was found in the case of CH3SO-3 anions for Na as
opposed to 2.9 KJ mol1 for K, while for the case of CF3SO3, the
opposite trend was observed (i.e., Ea KCF3SO3 < Ea NaCF3SO3). This
can be explained by the ionic association between sodium cations
(Na+) and sulfonate ions (SO-3), whose partial charges are different
in the case of CF3SO-3 (-0.576) and CH3SO-3 (-0.433), as indicated in
Table 1. Overall, the results summarized in Table 2 indicate the via-
bility of HEs for LIB application with respect to ionic conductivity.
Ionic conductivity is associated with (i) the solvation of ions,
which is linked to the size and charge density of the cation and
the nature of the solvent, and (ii) the ionic associations between
Fig. 1. Conductivity (r) of the standard electrolyte (1 M LiPF6 in EC/EMC (1:1, wt)) cation and anion, which hinder the diffusion of the cation by the
and its homologue hybrid electrolytes, SE + 0.5 M AB salt at 60 °C. interactions of Li+, Na+ or K+ with the counter-anion. These interac-

Fig. 2. Ionic conductivity (r) vs. temperature for hybrid electrolytes (a) (Li/K) and (b) (Li/Na)-based salts. Both graphs include SE (1 M LiPF6 in EC/EMC: 1:1, wt) as
comparison.

454
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

Fig. 3. Arrhenius and VTF plots showing the ionic conductivity (r) as a function of temperature for hybrid electrolytes (a, b) Na and (c, d) K-based salts.

Table 2 with different LiPF6 concentrations (0.25–3 M) in EC and EMC have


Comparison of the VTF-parameters of the ionic conductivity between SE and derived indicated that the reactivity of the electrolyte increases with
electrolytes at 25 °C. R2 describes the goodness of fit. higher salt concentrations [55,56]. Fig. 4 shows the DSC profiles
Electrolyte r(mS cm1) T0 (K) Ea (R  B)KJ mol1 R2 of the SE and its hybrid electrolytes where for most systems, a
melting transition peak (Tm) and eutectic behavior are evident.
SE 9.01 158.8 3.2 0.998
LiPF6 7.55 187.3 2.6 0.999 Yet, the onset temperature of the endothermic evaporation peaks
NaBF4 8.14 161.7 3.1 0.997 varies. This behavior is attributed primarily to EMC (Tm EMC = 107
KBF4 8.82 163.7 3.1 0.998 °C), whose interaction with alkali cations and counter anion delays
NaCF3SO3 5.50 155.7 3.5 0.997
KCF3SO3 7.60 167.8 2.7 0.999
NaCH3SO3 9.99 166.6 2.5 0.991
KCH3SO3 7.80 163.4 2.9 0.999

tions can be collaborative or competitive. In essence, the ionic con-


ductivity depends on cation, anion, and solvent interactions, which
in turn, rely on temperature. The ionic conductivity at 25 °C follows
the order ‘‘NaCH3SO003 < SE < ”KBF004 < ‘‘NaBF004 < ”KCH3SO003 < ‘‘KCF3SO003
< ”LiPF006 < ‘‘NaCF3SO003 . At elevated temperatures (80 °C), there is an
inversion, LiPF6 1.5 M becomes the most conductive (Fig. 2), prob-
ably because the temperature lowers the ionic interactions and
boosts the diffusion of Li+ ions. In the latter case, there is a high
molecular association between the carbonates and the salt,
prompting the alkali salts to move in a more coupled manner.

3.1.2. Thermal properties


The formulated electrolytes were measured by differential
scanning calorimetry (DSC) to delineate the role of the added salt Fig. 4. DSC thermograms of a LIB standard electrolyte (1 M LiPF6 in EC/EMC
in the thermal properties of the electrolytes. Typical DSC studies (1:1, wt) together with its homologue hybrid electrolytes SE + 0.5 M AB salt.

455
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

the evaporation process. EC has been reported to show an exother- [61]. From the first CV cycle, we can see that all the nickel is oxi-
mic decomposition peak between 205 and 250 °C [57]. dized at 4.5 V vs. Li/Li+, in line with reports attesting that at poten-
The presence of a eutectic diagram in the EC/EMC mixture tials of 4.5 V, Ni4+ ions would not oxidize any further [20].
stems from the interactions between the two solvents, which pre- After the activation occurring in the first cycle, the CVs embody
vent their crystallization (Tc = -14 °C for EMC and 36 °C for EC). In a sequential rearrangement of the lattice since Li is extracted dur-
turn, the eutectic equilibrium and its coordinates are affected by ing oxidation and reinserted during reduction. The pairs of redox
the nature and the concentration of the salts present. For instance, peaks in the CV curves (cycle 2 and 3) stem from the oxidation
the electrolyte containing CF3SO-3 anion for both Na and K metals of Ni from Ni3+ to Ni4+, involving the phase transitions from hexag-
boils at 150 and 161 °C, respectively; this is the highest tempera- onal to monoclinic (H1 M M) and monoclinic to hexagonal
ture amongst all hybrid electrolytes, followed by KBF4 (the (M M H2), and hexagonal (H2) to hexagonal (H3) during lithium
endothermic peak could be due to the polymorphic transition of extraction [62]. The three prominent redox peaks of the CV in
KBF4 from the orthorhombic to the cubic phase [58]) and NaBF4. Fig. 5a were previously assigned to phase transfer among four dif-
These electrolytes evaporate later than the SE (149 °C), validating ferent phases of LiNiO2-based cathode materials during cycling
their superior thermal stability. On the other hand, the hybrid elec- [63]. No significant change in cycles 2 and 3 was observed, suggest-
trolytes with CF3SO-3, reveal no evaporation peak. Here, the broad ing good stability of the NCA cathode during the lithiation and
endothermic peaks corresponding to eutectic melting lie between delithiation process in the presence of hybrid electrolytes. The
65 and 115 °C. good superposition of the different peaks during the successive
There are two elements to consider with thermal stability at cycles demonstrates good reversibility of the NCA material’s elec-
increasing temperatures. Firstly, solvent evaporation is lowered trochemical process in the HEs, which in turn implies an absence
by interactions due to ion solvation (Raoult’s law [59]). Secondly, of capacity loss and activation within the operating potential
thermal degradation of the LiPF6 salt, known for its deleterious range, viz. 3.0–4.5 V. H1 exhibits the highest oxidation and reduc-
effects in the presence of traces of water on the electrolytes by tion peaks, indicating that the Li is extracted from the Ni4+ rich
thermal decomposition, i.e., PF-6 + H+ ?HF + PF5. When considering phase surface of the electrode, boosting the electronic conductivity
the solvent evaporation and thus vapor pressure, the 1.5 M LiPF6 (i.e., smaller cation mixing in favor of higher electrochemical activ-
mixture is more stable. Nevertheless, taking into account the ity[64]). The potential difference (DE) between the main redox
chemical behaviour of the latter, the electrolytes with Na and K- couple (H1 M M) in the CV indicates the polarization degree [65].
based salts with anions other than PF-6 display superior thermal The DE value found here (110 mV) is in line with other reported
and chemical stability. values of NCA//Li systems [66], suggesting that in concentrated
solutions, there is high diffusion of Li + cations in the NCA.
3.2. Electrochemical performance of Li, Na, and K-based electrolytes on NCA cathodes mediate electrochemical reactions via intercala-
the NCA //Li system tion, where Li+ are extracted from/reinserted into the original
structure with minimum structural changes [67]. The NCA activa-
3.2.1. NCA activation step in the presence of alkali salt additives tion process allows the material to cross the energy gap necessary
The use of alkaline salt mixtures in standard battery electrolytes for its restructuring by reducing the energy barrier required to pro-
decreases the conductivity (i.e., hinders the transport properties of duce the conductive R2 phase (i.e., increased electronic conductiv-
Li+). Yet, we anticipate that they wild better polarize the interface ity due to the formation of Ni4+ ions [60]). This process is
at high potentials as they are more resistant to oxidation. To this demonstrated by the initial increase in potential (within 0.25
end, the electrochemical performance of NCA//Li cells was first mAh g1 of the charge capacity) followed by a relaxation and sta-
investigated by cyclic voltammetry at 25 °C in coin cells, at a scan bilization of the potential (0.25–1.5 mAh g1) once the energy bar-
rate of 0.05 mV s1, equivalent to a C/10 rate. A slow scan rate can rier for the activation is reached. The nature of the electrolyte
delineate the cation mixing effect on the phase transformations affects the activation rate and stabilization potential (i.e., peak
during the charge–discharge process. Fig. 5a shows the first three broadening) of NCA. The magnitude of this initial overvoltage has
cyclic voltammetry cycles in a solution containing 1.5 M LiPF6 been reported to reflect the amount of water-induced surface spe-
(SE + 0.5 M LiPF6). It allows a qualitative analysis of the potentials cies on NCA [68]. Fig. 5b and c show the initial stage of the galvano-
at which chemical reactions accompany electron transfer in the static curve of the NCA activation and confirm that the addition of
NCA electrode. The first cycle corresponds to an irreversible activa- salt, whatever its nature, does not accentuate the energy barrier,
tion process of the NCA lamellar structure [60], viz. inner structural which remains below 4.39 V vs. Li/Li+ before stabilizing
arrangements accompanying the departure of oxygenated species at ~ 4.05 V (Fig. 5b and c). The stabilization plateau is lower for
and the SEI film formation at the anode during the first cycle the Na salts (Fig. 5b) than for K (Fig. 5c). When lithium salt is added

Fig. 5. Cyclic voltammetry of a NCA//Li half-cell at 25 °C and a scan rate of 0.05 mV s1 in a solution containing EC/EMC and 1.5 M LiPF6. Snapshot of the initial galvanostatic
charge profiles of NCA//Li half-cells in the presence of Na-salt (b) and K-salt (c).

456
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

(+LiPF6), two activation peaks are present. We will, therefore, cycling of NCA//Li half cells was carried out at 25 °C and different
essentially retain that the activation process is not disturbed by C-rates for 20 cycles. All cells containing the hybrid electrolytes (K
the addition of salts beyond 1 M and that both Na and K are rather and Na) exhibited similar charge–discharge profiles between 3 and
favorable towards this step. 4.5 V, typical of the delithiation/lithiation profile of NCA cathodes
[70,71], suggesting a similar charge–discharge mechanism.
3.2.2. Comparative galvanostatic characterization of electrolytes For the hybrid electrolytes containing NaCF3SO3 salt (Fig. 6a),
containing alkali salt additives in NCA//Li half-cells the capacity obtained in the first cycle at C/20 reaches 221 mAh
Previous reports have shown that nickel-rich cathodes possess g1 and 191 mAh g1 during charge and discharge, respectively.
high specific capacity and discharge voltage but suffer from severe The dissimilar specific capacities found here indicate that during
cycling degradation, attributed to rapid volume contraction during the first cycle (formation of the cathode electrolyte interphase
the structural transformation from H2 to H3 [69]. The galvanostatic (CEI) layer), more lithium ions are consumed during charge, and

Fig. 6. Voltage specific capacity profiles of the NCA//Li half-cells at cycle 1 and 20 for hybrid electrolytes containing (a, c) Na and (b, d) K salts together with the standard
electrolyte. Current rate is C/20, and the operating temperature is 25 °C.

Table 3
Comparison of the electrochemical performance of the NCA//Li half-cell in the presence of SE and HE at 25 °C.

E + 0.5 M (APF6) QCycle 1 QCycle 20 EDcycle 1(Wh kg1) EDcycle 20 gC (%)


(A = Li, Na, K) (mAh.g1) (mAh.g1) (Wh kg1)
E = 1.0 M LiPF6
Li E 168 154 636 583 99.5
LiPF6 187 177 708 670 98.4
Na BF4 159 168 602 636 99.3
CH3SO3 177 159 670 602 99.4
CF3SO3 191 175 723 662 99.3
K BF4 133 115 503 435 98.9
CH3SO3 183 175 693 662 99.2
CF3SO3 162 145 613 598 98.5

The discharge specific capacity (Q) is measured for a C/10 rate. ED: energy density. gC: average Coulombic efficiency at C/10.

457
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

along with the high potentials, lead to lower discharge capacity. All [73] whilst higher LiPF6 concentrations could provide stronger oxi-
Na hybrid electrolytes show superior capacity in both the 1st and dation resistance as suggested by previous studies [72]).
20th cycle compared to the SE (Fig. 6b), validating that in the pres- Similar findings were recorded for the case of K-based hybrid
ence of these hybrid electrolytes, Li+ can efficiently extract/insert electrolytes except for ‘‘KBF004 which shows poor cycling (Fig. 6c).
from/in the NCA cathode. For the best performing HE (NaCF3SO3), The best performing, HE contains ”KBF004 ‘‘KCH3SO003 and yields a
the capacity retention was ca. 92% at the end of the 20th cycle, con- capacity of 230 mAh g1 and 180 mAh g1 during charge and dis-
sistent with the one recorded for the SE (91.6%), and lower when charge, respectively, with a capacity retention of 93% (Fig. 6d).
compared to 1.5 M LiPF6. The effect of a higher LiPF6 concentration The NCA//Li half cells were tested at four different C-rates (e.g.,
(1 vs. 1.5 M) has been previously reported [72]; here, we propose C/20, C/10, C/5, and 1C, Fig. 7a and b). As expected, higher C-rates
to use cheaper alkali salts (e.g., NaBF4 and KBF4) than LiPF6[15] (1C) moderated the cell’s capacity to 100 mAh g1 (e.g., ‘‘NaCF3SO3-
to achieve comparable LIB performances. ”, ‘‘NaCH3SO3” and ‘‘KCH3SO3”) and even to 50 mAh g1 for both
A summary of the results for the rest of the electrolytes is given electrolytes KCH3SO3BF-4. The performance of the HEs containing
in Table 3. The positive correlation of the addition of Na salt in the NaCF3SO3, NaCH3SO3, and KCH3SO3 was superior to the SE for all
SE with capacity from Table 3, together with the unchanged dis- C-rates. The galvanostatic results of the NCA//Li cells indicate that
charge profiles of NCA in most HEs (apart from KBF4), indicate that the addition of the aforementioned alkali salts improves the capac-
the lamellar structure of NCA is adequately protected by this alkali ity of the cell by an average of 10%. Likewise, the coulombic effi-
salt at high potentials (i.e., 4.5 V vs. Li/Li+) as opposed to the SE (i.e., ciency (gC) remained at 99% (Table 3), while the specific energy
the direct contact between Co, Ni and 1 M LiPF6 gives rise to disso- density based on the average operating voltage of the cell
lution of these transition metals and in turn to capacity fade of NCA remained between 600 and 670 Wh Kg1.

Fig. 7. Cycling performance of the NCA//Li half-cells in different hybrid electrolytes containing (a) Na and (b) K salts on top of the standard electrolyte (SE), EC/EMC + 1 M
LiPF6. The operating temperature is 25 °C.

Scheme 1. Comparative HOMO and LUMO energy levels of CF3SO3,CH3SO-3 BF-4 and PF-6 anions. The values are extracted from Table 1.

458
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

In general, the stability of the electrolyte is associated with the The Li+ insertion is regarded as a ‘‘charging process” where dur-
highest occupied molecular orbital (HOMO) for the oxidation pro- ing the first cycle (at C/20, Fig. 8a and 8b), the SEI is formed, fol-
cess and the lowest unoccupied molecular orbital (LUMO) energy lowed by lithium de-insertion (discharge). The electrolyte
levels for the reduction. From Scheme 1, we note that PF-6 and containing excess Li (1.5 M LiPF6) exhibits the highest charge
BF-4 exhibit the lowest HOMO energy (0.66 Hartree 17.96 eV), capacity, ca. 350 mAh g1, and the best cycling behavior (upon
indicative of high oxidation stability that arises from the high elec- 20 cycles, Fig. 9), followed by ‘‘KCH3SO003 and SE with 310 and 300
tronegativity of the fluorine atom. These HOMO levels however, mAh g1, respectively. After the formation of the SEI layer, the
don’t explain the performances observed on the NCA cathode Gr//Li half-cells show high coulombic efficiency, over 98%, for the
(Fig. 7). Additionally, taking into account the potential reduction remaining 20 cycles. It should be noted that high C-rates restrict
of anions on the Li metal, the CF3SO3 anion exhibits the lowest (C/1) the capacity to values under 100 mAh g1 except for the case
LUMO, making it more susceptible to reduction than the other of SE and 1.5 M LiPF6. Here, the Li+ intercalation into the 3D stacked
three anions (e.g., BF4, CH3SO3, and LiPF6). These anion potential structure and the coexistence of two-phase lithiated graphite [74]
windows do not justify the observed performance. Therefore, there is enhanced and stabilized with an excess amount of lithium salt
is no correlation between the intrinsic oxidation–reduction stabil- [75,76]. Comparable performance to that of SE is apparent for the
ity of salts and the electrolyte’s performance; this stability is linked ”NaCH3SO003 , where a capacity of 309 mAh g1 at a C/10 rate was
to the solvent after the solvation of ions. observed after 20 cycles (i.e., 97% capacity retention). The results
of the rate capability are summarized in Table 4, along with the
gC values.
3.2.3. Comparative galvanostatic characterization of electrolytes
containing alkali salt additives in Gr//Li half-cells
During the operation of LIBs, a solid electrolyte interphase (SEI) 3.2.4. Performance of electrolytes containing alkali salt additives in
layer forms on the graphite surface by the reductive decomposition Gr//NCA full-cells
of carbonate solvents. Hence, the electrochemical properties of gra- When characterizing a material in a half-cell configuration, one
phite (Gr) in the hybrid electrolytes were investigated by con- cannot neglect the lithium metal’s surface changes during cycling
structing Gr//Li half cells using lithium metal as the counter and propagating salt deposits, mainly when the electrolyte con-
electrode. tains cations other than Li+. To mitigate this issue, we compiled

Fig. 8. Voltage specific capacity profiles of the Gr//Li half-cells at cycle 1 and 20 in the presence of hybrid electrolytes containing (a) Na and (b) K salts and compared against
SE. Current rate is C/20, and the operating temperature is 25 °C.

Fig. 9. Cycling performance of the Gr//Li half-cells at room temperature in hybrid electrolytes containing (a) Na and (b) K salts on top of the standard electrolyte (SE).

459
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

Gr//NCA full cells and carried out galvanostatic charge–discharge rates. The discharge capacity remained over 150 mAh g1 when the
measurements at 25 °C for the HE. Voltage-capacity profiles for cell was discharged in 1 hour for NaBF4 and KBF4 HE.
the HE containing NaBF4 and KBF4, respectively, are illustrated in The capacities shown in Fig. 10 are greater than those obtained
Fig. 10. The cycling protocol commenced with the SEI layer forma- in the half-cell configuration (Table 4), indicating that Na+ and K+
tion at a slow rate (C/20) followed by ramping through different C- are not beneficial to the Li metal anode when in contact with the
HE (possibly creating a poor electrode/electrolyte interface at these
concentrations) leading to poor overall cell performance. To this
Table 4 end, reducing the amount of extra salt could promote a better
Comparison of the electrochemical performance of the Gr//Li half-cell in the presence
interface. A recently formulated electrolyte with 1 M LiPF6 and
of SE and HE at 25 °C.
0.02 M NaPF6 yielded low overpotentials on Li symmetric cells
E + 0.5 M (APF6) QCycle 1 QCycle 20 gC (%) [77], stemming from an isolating and self-healing electrostatic
E = 1.0 M LiPF6 (mAh.g1) (mAh.g1)
shield of Na+.
(A = Li, Na, K)
The NCA//Gr full-cells were cycled at different C-rates up to 1C
Li SE 332 329 98.50
(Fig. 11a and 11b) for Na-based and K-based HE, respectively. HEs
LiPF6 395 376 98.89
Na BF4 316 310 98.31 containing KBF4, NaBF4, KCH3SO3, and 1.5 M LiPF6 exhibited supe-
CH3SO3 316 309 98.41 rior performance to that of the standard electrolyte throughout all
CF3SO3 332 329 98.3 C-rates and good capacity retention (viz. 85–92%, Table 5). Other
K BF4 290 298 99.3 than the initial SEI formation cycle, the cells’ coulombic efficiencies
CH3SO3 362 360 98.2
were high, remaining between 98 to 99.5% for most of the HEs, and
CF3SO3 119 nw* nw*
are in line with SE values. The HE containing KCH3SO3 reaches an
The discharge specific capacity (Qdischarge) is measured for a C/10 rate. * Not energy density of 777 Wh kg1 after 40 cycles, the highest of all
working. ED: Energy density. gC: Average Coulombic efficiency at C/10.
investigated samples and more than double when compared to

Fig. 10. Voltage specific capacity profiles of the NCA//Gr cell in hybrid electrolytes containing (a) 1 M LiPF6 (EC/EMC) + 0.5 M NaBF4 and (b) 1 M LiPF6 (EC/EMC) + 0.5 M KBF4
at 25 °C. The current rate C/20 represents the initial formation cycle.

Fig. 11. Cycling performance of the Gr//NCA cells under different hybrid electrolytes containing (a) Na and (b) K salts and compared against SE. Operating temperature is
25 °C.

460
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

Table 5
Comparison of the electrochemical performance of the Gr//NCA cell in the presence of SE and HE at 25 °C.

E + 0.5 M (APF6) QCycle 4 (C/5) QCycle 40 EDcycle 4 EDcycle 40 gC (%)


(A = Li, Na, K) mAh.g1 (C/5) Wh kg1 Wh kg1
E = 1.0 M LiPF6 mAh.g1 (DE = 3.7 V) (DE = 3.7 V)
Li E 125 104 462 384 98.50
LiPF6 160 157 592 580 99.20
Na BF4 182 179 673 662 99.11
CH3SO3 46 36 170.2 133 97.67
K BF4 187 173 691.9 640 98.3
CH3SO3 190 210 703 777 98.9
CF3SO3 152 95 562 351 99.8

The discharge specific capacity (Qdischarge) is measured for a C/5 rate. ED: energy density. gC: average Coulombic efficiency at C/5.

the SE. KBF4 and KCF3SO3 yielded slightly lower efficiencies, Declaration of Competing Interest
between 92 and 95%. The results obtained with the Gr//NCA full
cell suggest that these hybrid electrodes with mixed alkali salts The authors declare that they have no known competing finan-
are suitable for operation in this system at room temperature. On cial interests or personal relationships that could have appeared
the contrary, the HE with KCF3SO3, and NaCH3SO3 exhibit low to influence the work reported in this paper.
specific capacities. These results do not reflect the classification
observed in the half Gr//Li cells and corroborate the deleterious
role of lithium metal, which does not make it possible to account Acknowledgments
for the actual performance of materials when they are in the pres-
ence of Na+ and K+ (particularly at such concentrations) while The authors would like to thank ‘‘La Région Centre Val de Loire”
specifying that the counterion plays a non-negligible role in this for financial support through the Obama project under the Lavoi-
effect. sier regional program.

References
4. Conclusion
[1] G. Nikiforidis, M.C.M. van de Sanden, M.N. Tsampas, RSC Adv. 9 (2019) 5649–
In this study, we have synthesized and tested a series of hybrid 5673.
electrolytes comprising a mixture of alkali salts (Li, Na, K) at a total [2] T. Kim, W. Song, D.-Y. Son, L.K. Ono, Y. Qi, Mater. Chem. A. 7 (2019) 2942–2964.
[3] P.K. Panda, A. Grigoriev, Y.K. Mishra, R. Ahuja, Nanoscale Adv. 2 (2020) 70–108.
salt concentration of 1.5 M for LIB applications. Their ionic conduc- [4] T. Perveen, M. Siddiq, N. Shahzad, R. Ihsan, A. Ahmad, M.I. Shahzad, Renew.
tivity at 60 °C is between 10 and 15 mS cm1 and follows the order: Sustain. Energy Rev. 119 (2020) 109549.
NaCH3SO3 <SE<KBF4 <NaBF4 <KCH3SO3 <KCF3SO3 <LiPF6 <NaCF3SO3. [5] X. Chen, X. Shen, T.-Z. Hou, R. Zhang, H.-J. Peng, Q. Zhang, Chem. 6 (2020)
2242–2256.
The majority of the HEs demonstrated a melting transition peak [6] G. Nikiforidis, G.J. Jongerden, E.F. Jongerden, M.C.M. van de Sanden, M.N.
(Tm) and eutectic behavior, with the onset of the endothermic Tsampas, J. Electrochem. Soc. 166 (2019) A135–A142.
evaporation peak varying due to their interaction with EMC. [7] S. Phadke, E. Coadou, M. Anouti, J. Phys. Chem. Lett. 8 (2017) 5907–5914.
[8] P. Leung, X. Li, C. Ponce de León, L. Berlouis, C.T.J. Low, F.C. Walsh, RSC Adv. 2
Molecular simulations revealed that the CF3SO-3 and CH3SO-3 anions (2012) 10125–10156.
display similar properties (e.g., HOMO, LUMO, and VdW) as [9] X. Zeng, M. Li, D. Abd El-Hady, W. Alshitari, A.S. Al-Bogami, J. Lu, K. Amine, Adv.
opposed to BF-4. The activation energies of the Na and K salt homo- Energy Mater. 9 (2019) 1900161.
[10] B. Flamme, D.J. Yeadon, S. Phadke, M. Anouti, J. Energy Chem. 52 (2021) 332–
logues depend strongly on the nature of the anions, reaching val- 342.
ues of 3.2 kJ mol1 with a maximum deviation of ± 0.5 kJ mol1. [11] M. Raghibi, B. Xiong, S. Phadke, M. Anouti, Electrochim. Acta. 362 (2020)
Upon assembling coin cells with the proposed HEs, half-cell studies 137214.
[12] A. Würsig, H. Buqa, M. Holzapfel, F. Krumeich, P. Novák, Electrochem. Solid
confirmed the NCA activation, which displayed adequate cyclabil-
State Lett. 8 (2004) A34.
ity (20 cycles), high coulombic efficiency (99%), and high capacity [13] Y. Liang, C.-Z. Zhao, H. Yuan, Y. Chen, W. Zhang, J.-Q. Huang, D. Yu, Y. Liu, M.-M.
retention (e.g., 92% for NaCF3SO3) compared to a standard com- Titirici, Y.-L. Chueh, H. Yu, Q. Zhang, InfoMat. 1 (2019) 6–32.
mercial electrolyte, signifying that the lamellar structure of NCA [14] S. Kandhasamy, G. Nikiforidis, G.J. Jongerden, F. Jongerden, M.C.M. van de
Sanden, M.N. Tsampas, ChemElectroChem. 8 (2021) 1156–1166.
is adequately protected by the alkali salts at high potentials [15] G. Nikiforidis, M. Raghibi, A. Sayegh, M. Anouti, J. Phys. Chem. Lett. 12 (2021)
(4.5 V vs. Li/Li+). In addition, after the formation of the SEI layer, 1911–1917.
the Gr//Li half-cells exhibited high coulombic efficiency (>98%) [16] S. Phadke, M. Anouti, Electrochim. Acta. 223 (2017) 31–38.
[17] R. Robert, C. Villevieille, P. Novák, J. Mater. Chem. A. 2 (2014) 8589–8598.
for 20 cycles but showed moderate resilience to high current rates, [18] G.E. Blomgren, J. Electrochem. Soc. 164 (2016) A5019.
suggesting that the coexistence of two-phase lithiated graphite is [19] N. Ding, X. Wang, Y. Hou, S. Wang, X. Li, D.W.H. Fam, Y. Zong, Z. Liu, Solid State
solely stabilized with an excess amount of lithium salt. Since only Ion. 323 (2018) 72–77.
[20] B. Larhrib, G. Nikiforidis, M. Anouti, Electrochim. Acta. (2021) 137841.
the Li+ ions are intercalated in the insertion materials (Gr and [21] J.R. Croy, D.C. O’Hanlon, S. Sharifi-Asl, M. Murphy, A. Mane, C.-W. Lee, S.E.
NCA), the counterions have no role in the insertion mechanism. Trask, R. Shahbazian-Yassar, M. Balasubramanian, Chem. Mater. 31 (2019)
Aside from the HE with KCF3SO3, and NaCH3SO3, the performance 3891–3899.
[22] W. Wang, Q. Yang, K. Qian, B. Li, J. Energy Chem. 47 (2020) 72–78.
of the Gr//NCA full cell with the HEs is typical for this battery sys- [23] K. Edström, T. Gustafsson, J.O. Thomas, Electrochim. Acta. 50 (2004) 397–403.
tem (i.e., 40 cycles, four different C-rates, capacity retention of 90%, [24] J. Vetter, P. Novák, M.R. Wagner, C. Veit, K.C. Möller, J.O. Besenhard, M. Winter,
coulombic efficiency of 98%), validating the suitability of these M. Wohlfahrt-Mehrens, C. Vogler, A. Hammouche, J. Power Sources. 147
(2005) 269–281.
electrolytes for battery applications. The conclusion of this work
[25] L. Zhang, J. Fu, C. Zhang, Nanoscale Res. Lett. 12 (2017) 1–7.
is therefore that the cations (Li, Na, or K) added to Li+ (standard [26] B. Huang, X. Li, Z. Wang, H. Guo, L. Shen, J. Wang, J. Power Sources. 252 (2014)
electrolyte) do not disturb or improve the performance during 200–207.
intercalation of Li+ cations in NCA and graphite, but improve the [27] J. Kasnatscheew, B. Streipert, S. Röser, R. Wagner, I. Cekic Laskovic, M. Winter,
Phys. Chem. Chem. Phys. 19 (2017) 16078–16086.
thermal stability, conductivity, and potentially the cost of elec- [28] S. Perez Beltran, X. Cao, J.-G. Zhang, P.B. Balbuena, Chem. Mater. 32 (2020)
trolytes because Na and K salts are more accessible and cheaper. 5973–5984.

461
A. Jrondi, G. Nikiforidis and Mérièm Anouti Journal of Energy Chemistry 64 (2022) 451–462

[29] C.L. Berhaut, D. Lemordant, P. Porion, L. Timperman, G. Schmidt, M. Anouti, RSC [55] G.G. Botte, R.E. White, Z. Zhang, J. Power Sources. 97 (2001) 570–575.
Adv. 9 (2019) 4599–4608. [56] G.G. Eshetu, S. Grugeon, S. Laruelle, S. Boyanov, A. Lecocq, J.-P. Bertrand, G.
[30] S. Amara, J. Toulc’Hoat, L. Timperman, A. Biller, H. Galiano, C. Marcel, M. Marlair, Phys. Chem. Chem. Phys. 15 (2013) 9145–9155.
Ledigabel, M. Anouti, ChemPhysChem. 20 (2019) 581–594. [57] Y.-S. Duh, C.-Y. Lee, Y.-L. Chen, C.-S. Kao, Thermochim. Acta. 642 (2016) 88–94.
[31] C.L. Berhaut, P. Porion, L. Timperman, G. Schmidt, D. Lemordant, M. Anouti, [58] V.I. Saldin, V.V. Sukhovey, N.N. Savchenko, A.B. Slobodyuk, L.N. Ignatieva, Russ.
Electrochim. Acta. 180 (2015) 778–787. J. Inorg. Chem. 61 (2016) 630–637.
[32] W. Zhang, Y. Liu, Z. Guo, Science Adv. 5 (2019) eaav7412. [59] P. Costa, M.A. Teixeira, G. Mestre, L. Carneiro, J.M. Loureiro, A.E. Rodrigues, Ind.
[33] V. Nilsson, A. Kotronia, M. Lacey, K. Edström, P. Johansson, A.C.S. Appl, Energy Eng. Chem. Res. 56 (2017) 8767–8777.
Mater. 3 (2020) 200–207. [60] X. Yang, J. Chen, Q. Zheng, W. Tu, L. Xing, Y. Liao, M. Xu, Q. Huang, G. Cao, W. Li,
[34] A.J. Naylor, M. Carboni, M. Valvo, R. Younesi, A.C.S. Appl, Mater. Interfaces. 11 J. Mater. Chem. A. 6 (2018) 16149–16163.
(2019) 45636–45645. [61] H. Xie, K. Du, G. Hu, J. Duan, Z. Peng, Z. Zhang, Y. Cao, J. Mater. Chem. A. 3
[35] G.G. Eshetu, T. Diemant, M. Hekmatfar, S. Grugeon, R.J. Behm, S. Laruelle, M. (2015) 20236–20243.
Armand, S. Passerini, Nano Energy. 55 (2019) 327–340. [62] J.D. Steiner, L. Mu, J. Walsh, M.M. Rahman, B. Zydlewski, F.M. Michel,
[36] J.K. Stark, Y. Ding, P.A. Kohl, J. Electrochem. Soc. 160 (2013) D337–D342. H.L. Xin, D. Nordlund, F. Lin, A.C.S. Appl, Mater. Interfaces. 10 (2018) 23842–
[37] Q. Xu, Y. Yang, H. Shao, Electrochim. Acta. 271 (2018) 617–623. 23850.
[38] V. Aravindan, J. Gnanaraj, S. Madhavi, H.-K. Liu, Chem. Eur. J. 17 (2011) 14326– [63] S. Deng, Y. Li, Q. Dai, J. Fu, Y. Chen, J. Zheng, T. Lei, J. Guo, J. Gao, W. Li, Sustain
14346. Energ Fuels. 3 (2019) 3234–3243.
[39] Y. Tominaga, K. Yamazaki, V. Nanthana, J. Electrochem. Soc. 162 (2015) [64] B.-S. Liu, Z.-B. Wang, F.-D. Yu, Y. Xue, G.-J. Wang, Y. Zhang, Y.-X. Zhou, RSC Adv.
A3133–A3136. 6 (2016) 108558–108565.
[40] K. Kubota, M. Dahbi, T. Hosaka, S. Kumakura, S. Komaba, Chem. Rec. 18 (2018) [65] Z. Huang, Z. Wang, Q. Jing, H. Guo, X. Li, Z. Yang, Electrochim. Acta. 192 (2016)
459–479. 120–126.
[41] T.A. Pham, K.E. Kweon, A. Samanta, V. Lordi, J.E. Pask, J. Phys. Chem. C. 121 [66] G. Cao, J. Zhu, Y. Li, Y. Zhou, Z. Jin, B. Xu, C. Hai, J. Zeng, RSC Adv. 10 (2020)
(2017) (1920) 21913–21922. 9917–9923.
[42] Y. Wang, F. Liu, G. Fan, X. Qiu, J. Liu, Z. Yan, K. Zhang, F. Cheng, J. Chen, J. Am. [67] S. Hwang, E.A. Stach, J. Phys. D: Appl. Phys. 53 (2020) 113002.
Chem. Soc. 143 (2021) 2829–2837. [68] M. Hofmann, F. Nagler, M. Kapuschinski, U. Guntow, G.A. Giffin,
[43] T. Rüther, A.I. Bhatt, A.S. Best, K.R. Harris, A.F. Hollenkamp, Batter. Supercaps. 3 ChemSusChem. 13 (2020) 5962–5971.
(2020) 793–827. [69] H.-J. Noh, S. Youn, C.S. Yoon, Y.-K. Sun, J. Power Sources. 233 (2013)
[44] X. Zuo, C. Fan, J. Liu, X. Xiao, J. Wu, J. Nan, J. Electrochem. Soc. 160 (2013) 121–130.
A1199–A1204. [70] J. Li, W. Li, S. Wang, K. Jarvis, J. Yang, A. Manthiram, Chem. Mater. 30 (2018)
[45] G. Nikiforidis, A. Belhcen, M. Anouti, J. Energy Chem. 57 (2021) 238–246. 3101–3109.
[46] G. Nikiforidis, W.A. Daoud, Electrochim. Acta. 141 (2014) 255–262. [71] H. Lyu, Y. Li, C.J. Jafta, C.A. Bridges, H.M. Meyer, A. Borisevich, M.P.
[47] G. Nikiforidis, W.A. Daoud, J. Electrochem. Soc. 162 (2015) A809–A819. Paranthaman, S. Dai, X.-G. Sun, J. Power Sources. 412 (2019) 527–535.
[48] R. Simha, G. Carri, J. Polym. Sci., Part B: Polym. Phys. 32 (1994) 2645–2651. [72] Q. Liu, H. Xu, F. Wu, D. Mu, L. Shi, L. Wang, J. Bi, B. Wu, A.C.S. Appl, Energy
[49] N. Xu, J. Shi, G. Liu, X. Yang, J. Zheng, Z. Zhang, Y. Yang, J. Power Sources Adv. 7 Mater. 2 (2019) 8878–8884.
(2021) 100043. [73] H. Zhou, Z. Yang, D. Xiao, K. Xiao, J. Li, J Mater Sci-Mater El.29 (2018) 6648–
[50] S. Park, J. Jeong, G. Hyun, M. Kim, H. Lee, Y. Yi, Sci. Rep. 6 (2016) 35262. 6659.
[51] C. Paduani, P. Jena, J. Phys. Chem. A. 116 (2012) 1469–1474. [74] A.J. Blake, H. Huang, J. Chem. Educ. 92 (2015) 355–359.
[52] J. Landesfeind, H.A. Gasteiger, J. Electrochem. Soc. 166 (2019) A3079. [75] Z. Zeng, V. Murugesan, K.S. Han, X. Jiang, Y. Cao, L. Xiao, X. Ai, H. Yang, J.-G.
[53] K.M. Diederichsen, E.J. McShane, B.D. McCloskey, ACS Energy Lett. 2 (2017) Zhang, M.L. Sushko, J. Liu, Nat. Energy. 3 (2018) 674–681.
2563–2575. [76] Y. Yamada, K. Furukawa, K. Sodeyama, K. Kikuchi, M. Yaegashi, Y. Tateyama, A.
[54] K.M. Diederichsen, H.G. Buss, B.D. McCloskey, Macromolecules. 50 (2017) Yamada, J. Am. Chem. Soc. 136 (2014) 5039–5046.
3831–3840. [77] J. Liu, Y. Wang, F. Liu, F. Cheng, J. Chen, J. Energy Chem. 42 (2020) 1–4.

462

You might also like