Sie sind auf Seite 1von 9

Journal of Power Sources 331 (2016) 426e434

Contents lists available at ScienceDirect

Journal of Power Sources


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

Towards Li(Ni0.33Mn0.33Co0.33)O2/graphite batteries with ionic liquidbased electrolytes. I. Electrodes' behavior in lithium half-cells
E. Simonetti a, G. Maresca a, G.B. Appetecchi a, *, G.-T. Kim b, c, N. Loefer b, c,
S. Passerini b, c, **
a
b
c

ENEA, Agency for New Technologies, Energy and Sustainable Economic Development, SSPT-PROMAS-MATPRO, Via Anguillarese 301, Rome, 00123, Italy
Helmholtz Institute Ulm - Karlsruhe Institute of Technology, Helmholtzstrasse 11, 89081, Ulm, Germany
Karlsruhe Institute of Technology (KIT), P.O. Box 3640, 76021, Karlsruhe, Germany

h i g h l i g h t s
 LiTFSI-PYR13TFSI-PYR13FSI ionic liquid electrolyte mixture properly designed and developed.
 Large capacity of graphite and high voltage of NMC combined with the safety of IL electrolytes.
 Electrodes based on water-soluble, natural CMC binder; H2O as the only processing solvent.
 Performance improvement with respect to electrolytes based on a single ionic liquid.

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 15 July 2016
Received in revised form
12 September 2016
Accepted 14 September 2016

Lithium cells based on NMC cathodes or graphite anodes and ionic liquid-based electrolyte mixtures are
investigated. The electrode tapes, using water-soluble natural binders, as well as the ionic liquid materials, are prepared through eco-friendly routes involving H2O as the only processing solvent. The Li/NMC
and Li/graphite half-cells are studied by cyclic voltammetry, impedance spectroscopy and galvanostatic
cycling tests at different temperatures. The results herein reported, demonstrate the performance
improvement in terms of cycling behavior and ageing resistance, granted by the ionic liquid mixtures
with respect to the electrolytes reported in literature based on a single ionic liquid.
2016 Elsevier B.V. All rights reserved.

Keywords:
Ionic liquid electrolyte mixtures
Water-soluble binder
NMC cathodes
Graphite anodes
Safer Li-ion batteries

1. Introduction
Rechargeable lithium batteries are excellent candidates for the
next generation of electrochemical energy storage systems due
to their high gravimetric and volumetric energy densities [1].
However, applications as automotive, storage of energy from
renewable sources, smart grids and also consumer electronics
require higher and higher energy density devices, pushing

* Corresponding author. ENEA, Agency for New Technologies, Energy and Sustainable Economic Development, SSPT-PROMAS-MATPRO, Via Anguillarese 301,
Rome, 00123, Italy.
** Corresponding author. Helmholtz Institute Ulm - Karlsruhe Institute of Technology, Helmholtzstrasse 11, 89081, Ulm, Germany.
E-mail addresses: gianni.appetecchi@enea.it (G.B. Appetecchi), stefano.
passerini@kit.edu (S. Passerini).
http://dx.doi.org/10.1016/j.jpowsour.2016.09.078
0378-7753/ 2016 Elsevier B.V. All rights reserved.

research activities towards large capacity anodic and high voltage


and capacity cathodic materials. While graphite still represents
the best choice as the anode active material due to its
low operative voltage and large capacity, the need for wellperforming cathode materials is more stringent. Lithiated
nickel-manganese-cobalt oxide (NMC) is one of the most investigated, if not the most investigated, high operative voltage
(above 4 V vs Li/Li) cathode active material [2]. Additionally, this
material offers comparable large charge/discharge capacities and
allows to reduce the amount of expensive and toxic cobalt in
lithium battery systems.
Ionic liquids (ILs), molten salts at room temperature constituted
by organic cations and inorganic/organic anions, represent a very
interesting new class of room temperature uids with very unique
properties such as ame retardant, negligible vapor pressure in
conjunction with ambient or sub-ambient melting temperature,

E. Simonetti et al. / Journal of Power Sources 331 (2016) 426e434

remarkable ionic conductivity, wide thermal/chemical/electrochemical stability window, low heat capacity, ability to dissolve
inorganic (e.g., lithium salts), organic and polymeric materials and,
in some cases, hydrophobicity [3e5]. In addition, ILs can be favorably combined to obtain mixtures with improved properties, often
not exhibited by single ionic liquid materials. In the last years, ILs
have been widely investigated as safe electrolyte components to
replace the volatile and hazardous alkyl carbonates commonly used
in commercial lithium batteries. Particularly, ILs based on N-alkylN-methylpyrrolidinium cations, (PYR1A) (the subscripts indicate
the number of carbon atoms in the alkyl side chains), and bis(triuoromethanesulfonyl)imide (TFSI) or bis(uorosulfonyl)imide
(FSI) anions have been favorably proposed [6].
In this work we combine the large capacity of graphite and the
high voltage and capacity of NMC with the safety advantages of
ionic liquid electrolytes [7]. Moreover, both cathode and anode
electrodes were made using uorine-free, water-soluble, natural
binders, such as carboxymethyl cellulose sodium salt (CMC),
instead of the more expensive and less environmentally
friendly polyvinyliden-di-uoride (PVdF). Thus, the electrode
manufacturing involved the use of water as the only processing
solvent [2,8,9], replacing the hazardous, toxic and expensive
N-methyl-pyrrolidone (NMP). The use of CMC allows not only low
cost and eco-friendly manufacturing processes, but also easier
recycling of the battery components. The dissolution of the electrodes' binder in water allows, for example, the easy separation and
full recovery of the metallic current collectors and the electrodes'
powder components [8,9].
In the present manuscript we report the electrochemical performance of NMC cathodes and graphite anodes in ionic liquidbased, LiTFSI-PYR13TFSI-PYR13FSI ternary electrolyte mixtures.
Tests in organic electrolyte solutions were also carried out for
comparison purposes.
2. Experimental
2.1. Synthesis of ionic liquid materials
N-methyl-N-propylpyrrolidinium bis(triuoromethanesulfonyl)
imide
(PYR13TFSI)
and
N-methyl-N-propylpyrrolidinium
bis(uorosulfonyl)imide (PYR13FSI), both ionic liquids, were synthesized through a procedure reported elsewhere [10]. Water was
employed as the only processing solvent. The chemicals, i.e.,
N-methylpyrrolidine (Aldrich, 98 wt%), 1-bromopropane (Aldrich,
99 wt%), activated charcoal (Aldrich, Darco-G60), alumina (Aldrich,
acidic, Brockmann I), LiTFSI (3M, 99.9 wt%) and NaFSI salts (Solvionic 99.9 wt%), were used as received. The synthesized ionic
liquids exhibited moisture, Li and Br contents below 2 ppm
as determined by Karl-Fischer titrations and atomic absorption
measurements.
The 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI ternary mixture (where
0.1, 0.3 and 0.6 represent the mole fractions, respectively), used as
the electrolyte [11,12], was prepared by dissolving proper amounts
of the three components at 50  C). Prior to mixing, LiTFSI was
vacuum dried overnight at 120  C.
2.2. Preparation of NMC cathodes
Composite cathode tapes [2] were fabricated by blending
LiNi0.33Mn0.33Co0.33O2 (NMC, TODA) as active material (88 wt%),
sodium carboxymethyl cellulose (CMC, Dow Wolff Cellulosics,
Walocel CRT 2000 PPA 12 with a degree of substitution equal
to 1.2) as the binder (5 wt%) and carbon black (Super C45,
IMERYS, primary average particle size: 30 nm) as the conducting
agent (7 wt%).

427

CMC binder was rstly dissolved in deionized water by magnetic stirring (300 rpm) overnight at room temperature. Then, the
carbon black (7 wt% of the solid components) and phosphoric acid
(1 wt% of NMC) were added to the aqueous CMC solution and
dispersed for 3 h (1000 rpm) by a Dispermat dissolver. Successively, NMC was added and the so-obtained cathode slurry was
homogenized through high-speed mixing (1000e1500 rpm) by the
Dispermat for 2e3 h. Finally, vacuum (around 100 mPa) was
applied for 10e30 min to remove trapped air from the slurry.
The cathode slurry was cast onto aluminum foil (20 mm thickness). The coated electrode was immediately pre-dried (atmospheric oven) at 80  C for 30 min. The dried cathode electrodes
were cut into disks with a diameter of 1.2 cm. The disk electrodes
were vacuum dried at 170  C for 12 h. The average mass loading of
the NMC electrodes was tuned up to 4.1 mg cm2 (corresponding to
a theoretical capacity of 0.8 mA h cm2).
2.3. Preparation of graphite anodes
Composite anode tapes [2] were fabricated by blending graphite
(SLP 30, IMERYS) (91 wt%), sodium-carboxymethylcellulose (CMC,
5 wt%) and carbon black (Super C45, 4 wt%).
CMC was rstly dissolved in deionized water by magnetic stirring (300 rpm) overnight at room temperature. Then, the solution
was transferred into a stainless steel vessel and the proper carbon
amount was added. Further stirring (by a Dispermat) was carried
out at medium speed (2000 rpm) for about 1.5 h under vacuum.
Afterwards the proper graphite amount was added and the soobtained slurry was homogenized (2000 rpm) for 2 h (20  C) and
for further 10e30 min under vacuum (200 rpm). Successively, the
anode slurry was transferred into a pre-pilot automated coating
line and cast onto copper foil (thickness: 10 mm). The coated electrode was immediately dried in the heating zone (120  C) of the
coating machine in ambient air. The dried anode electrode (active
material mass loading equal to 3.0 mg cm2, corresponding to a
theoretical capacity of 1.1 mA h cm2) was cut into disks of 1.2 cm
diameter nally dried at 170  C under vacuum for 12 h.
2.4. Cell manufacturing
The electrochemical performance of the electrodes was evaluated in Li/NMC (cathode) and Li/graphite (anode) half-cells. The
cells, manufactured in the dry-room (R.H. at 20  C < 0.01%), were
realized by sandwiching a 50 mm lithium disk, a 25 mm polymer
separator (Asahi Kasei, Hipore SV718) and either NMC or graphite
electrode. The ionic liquid electrolyte was loaded into the porous
separator before cell assembling. Successively, the cells were
housed in soft-pouch envelopes, evacuated for 1 h and, then,
vacuum-sealed. For comparison purpose, the electrode tapes were
also investigated in (1M)LiPF6-EC:DMC (1:1 in weight) (Merck,
battery grade) organic electrolytes by manufacturing (in dry-room)
2016 coin-type cells.
2.5. Electrochemical tests
Consecutive cyclic voltammetries were conducted scanning
(0.1 mV s1) the cell voltage from the OCV value towards a) (Li/NMC
cells) more positive voltages and, successively, between 3.0 and
4.3 V (vs Li/Li) and b) (Li/graphite cells) more negative voltages
and, successively, between 0.01 and 2.0 V (vs Li/Li). The measurements were performed at 0, 20 and 40 1  C (the cells were
stored in climatic chambers) using a VMP3 (Bio-Logic SAS) galvanostat/potentiostat.
Impedance spectroscopy measurements were carried out using
the same instrument on Li/NMC and Li/graphite half-cells before

428

E. Simonetti et al. / Journal of Power Sources 331 (2016) 426e434

and after cyclic voltammetries. The AC tests were run in the


100,000e1 Hz frequency range with a DV amplitude of 10 mV.
The performance of the lithium half-cells was investigated
through galvanostatic cycling using a Maccor 4000 battery cycler.
The tests were run at different current rates (0.1Ce1C) and temperatures (from 0 to 50  C) within the 3.0e4.3 V (vs Li/Li) (Li/NMC
cells) and 0.01e2.0 V (vs Li/Li) (Li/graphite) voltage ranges,
respectively. Li/graphite cells were subjected to a potentiostatic
step at 0.01 V (vs Li/Li) until the residual current decrease below
1/10 of the nominal value, in order to promote lithium intercalation, especially at higher rates.
3. Results and discussion
3.1. Ionic liquid electrolyte
The 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI electrolyte [11,12], was
designed with the aim to optimally combine the properties of
PYR13TFSI (wide thermal and electrochemical stabilities) with
those of PYR13FSI (fast ion transport, low melting temperature,
protective lm-forming ability). In particular, the overall TFSI:FSI
mole ratio was chosen equal to 2:3 [11,12] and the LiTFSI mole

fraction was xed to 0.1 [12,13]. The so-obtained IL ternary electrolyte shows ionic conductivities approaching 103 S cm1
at 20  C and electrochemical stability window up to 5 V [12].
3.2. NMC cathodes
The lithium intercalation process of NMC cathodes in 0.1LiTFSI0.3PYR13TFSI-0.6PYR13FSI electrolyte was investigated through cyclic voltammetries conducted at temperatures ranging from 0 to
40  C. The results, reported in Fig. 1, show how at 20 (panel C) and
40  C (panel E) the trace obtained during the rst anodic (Li
deintercalation) sweep differs somewhat with respect to the
following ones, which display good reproducibility as well as
the proles generated in the cathodic (Li intercalation) scans.
Conversely, at 0  C irreproducible current/voltage proles are
observed during the rst three (consecutive) voltammetric cycles.
This behavior is likely ascribable to a not well-dened, poorly
ionically and/or electronically conductive layer, previously coating
the NMC active material particles [2], which prevents the Li
deintercalation. As reported by several authors [14e17] waterbased processing of common cathode materials leads to the formation of insulating inorganic compounds (e.g., lithium carbonate

Fig. 1. Cyclic voltammetries (panels A, C and E) and Nyquist plots (panels B, D and F) of Li/NMC half-cells in 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI ionic liquid electrolyte at 0 (panels A
and B), 20 (panels C and D) and 40  C (panels E and F). Scan rate (CVs): 0.1 mV s1. The AC measurements were taken before (open circles) and after (solid circles) the cyclic
voltammetries. The inserts in panels B, D and F magnify the high frequency portion of the AC plots.

E. Simonetti et al. / Journal of Power Sources 331 (2016) 426e434

(Li2CO3), lithium hydroxide (LiOH) on the active material surface.


These compounds are evidenced to decompose at potentials above
4.0 V vs Li/Li at room temperature [18,19]. Therefore, above 4 V
and depending on the temperature, e.g., from 4.05 to 4.2 V in
passing from 40 to 0  C, such a passivating layer is seen (at least
partially) to decompose, allowing the Li extraction. The removal of
this passivating lm is witnessed by a 100e150 mV shift towards
lower voltages of the maximum current peak, found to increase
with increasing the temperature (from 0.2 to 0.6 mA cm2 in passing
from 0 to 40  C), in the following anodic scans with respect to the
initial one. Therefore, the Li intercalation process within NMC is
allowed to proceed reversibly. However, it should be mentioned
that simultaneously to the decomposition of supercial Li2CO3 and
LiOH, the SEI (Solid Electrolyte Interphase) forms on the surface of
cathode particles, contributing to the rst anodic sweep. AC measurements, carried out on the same Li/NMC cells before and after
the cyclic voltammetries, are displayed in Fig. 1 (panels B, D and F)
as Nyquist plots [20]. They show sensible decrease of the overall
interfacial resistance (as indicated by the reduction of the
medium-high frequency semicircle [20]), after the cyclic voltammetries at 20  C (from 550 to 350 U) and, especially, 0  C
(from 2500 to 1000 U). This suggests for the improvement of
the electrolyte/NMC interface due to removal of the passive
coating onto the active material surface. Conversely, the opposite
behavior (from 100 to about 150 U) is observed at 40  C, likely
associated to moderate formation of SEI (Solid Electrolyte Interphase) onto NMC and Li electrodes. However, no practical electrolyte degradation is observed as indicated by the neglectable
shift of the high frequency intercept of the impedance plots towards the real axes (inserts in Fig. 1B, D and F), which is indicative
of the bulk electrolyte impedance. The interfacial resistance
is seen to decrease one-order of magnitude in passing from
0 (panel B) to 40  C (panel F), e.g., in agreement with the current
density increase observed in the voltammetry cycles conducted at
the same temperatures.
The voltage vs capacity prole of selected charge-discharge
cycles of Li/NMC cells at different temperatures (see legend in the
panels) is reported in Fig. 2. For comparison purpose, voltage.

429

proles obtained with (1M)LiPF6-EC:DMC (1:1 in weight) electrolyte are also presented. As also observed in the organic electrolyte, the 1st charge is found to be different with respect to the
following ones; in particular, the cell voltage is seen to suddenly
increase up to about 4 V and, successively, to quickly decrease down
to 3.8 V showing the typical plateau corresponding to Li deintercalation process from NMC material [2]. This behavior,
observed at all the investigated temperatures, agrees with the results obtained from the cyclic voltammetry tests reported in Fig. 1.
The passive layer, coating the NMC particles, prevents lithium ion
extraction out of the active material, leading to increased electrolyte/electrode interfacial resistance and, therefore, sharp voltage
increase. This is especially observed at the lower temperatures,
once more in agreement with the results showed in Fig. 1. Above
4 V, the insulating passive layer decomposes, allowing easier Li
deintercalation, as indicated by the voltage decay followed by a
plateau. Such a behavior accounts for the irreversible capacity
detected in the rst charge/discharge cycle [2,21]. The initial, sudden voltage rise upon charge is more marked with the ionic liquid
electrolyte, likely due to its lower ionic conductivity and higher
viscosity. In fact, it is seen to increase with temperature decrease.
The discharge half-cycles show the (expected) plateau due to Li
intercalation into NMC. It is worth to note that both charge (with
the exception of the rst half-cycle) and discharge features are
practically overlapping, indicating high reversibility of the lithium
intercalation process and high coulombic efciency from 0 to 50  C.
Such a behavior supports for the feasibility and reliability of Limetal cells constituted of NMC electrodes based on a natural
binder, and the 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI ionic liquid
electrolyte. The results clearly evidence how NMC cathodes are able
of delivering more than 90% of the capacity discharged in organic
solutions with a somewhat similar operative voltage. A progressive
capacity decrease is observed in passing from 50 to 0  C, however,
still interesting capacity values (about 75 mA h g1, corresponding
to one half of the nominal capacity) and good capacity retention are
achieved at the lowest temperature.
Fig. 3 compares the cycling performance detected, upon 100%
DOD tests, at 23 (upper panel) and 50  C (lower panel) and different

Fig. 2. Voltage vs capacity proles of Li/NMC half-cells in 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI ionic liquid electrolyte at 0.1C and different temperatures. Panel A: 23  C; panel B:
50  C; panel C: 10  C; panel D: 0  C. The voltage vs capacity proles, obtained in (1M)LiPF6-EC/DMC (1:1 in weight) conventional organic electrolyte at 0.1C and 23  C, are reported
for comparison purpose in panel A.

430

E. Simonetti et al. / Journal of Power Sources 331 (2016) 426e434

Fig. 3. Discharge capacity evolution of Li/NMC half-cells in 0.1LiTFSI-0.3PYR13TFSI0.6PYR13FSI ionic liquid (full squares) and (1M)LiPF6-EC/DMC (1:1 in weight) conventional organic (open squares) electrolytes at different current rates. T 23 (upper
panel) and 50  C (lower panel).

current rates with 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI (solid


squares) and (1M)LiPF6 in EC/DMC (1:1 in weight) (open squares)
electrolytes. At room temperature, the NMC electrodes show a
nominal reversible capacity at 0.1C approaching 135 mA h g1 with
very good capacity retention. A relevant capacity (110 mA h g1) is
delivered even at medium rate (0.2C) whereas a more marked
decay (not detected in organic electrolytes) is observed at higher
current rates. This is likely ascribed to both the higher viscosity of
the ionic liquid electrolyte (with respect to the organic one) and,
especially, to formation of the eutectic compound (Li:PYR13 mol
ratio 1:2) [13]. The latter, promoted by higher currents leading to
increase of the Li concentration at one electrolyte/electrode
interface (and contemporaneous Li impoverishment at the other
one), is solid at room temperature, thus sharply and, disadvantageously, decreasing the (room temperature) ionic conductivity.
However, a similar behavior is not observed at 50  C because of the
melting of the eutectic compound. Therefore, at medium temperatures the NMC performance in the IL-based electrolyte approaches
that observed in organic electrolyte, e.g., more than 110 mA h g1
were still discharged at 1C, corresponding to about 73% of the initial
(nominal) capacity recorded at 0.1C. Thus, Li/NMC cells with ILbased electrolytes exhibit good capacity retention from room to
medium temperature even at high rates. Further tests performed at
low-medium rates (0.1C) after fast (1C) charge-discharge cycling,
indicate that these cells are able to readily recover the initial
capacity.
The performance of NMC electrodes in the ionic liquid electrolyte was investigated under prolonged cycling tests (0.1 C and 100%
DOD) at 23 and 40  C. The electrodes were also tested at 40  C using
(1M)LiPF6-EC/DMC (1:1 in weight) organic electrolyte for

comparison purpose. The results, reported in Fig. 4, show how


about 98% of the initial capacity (panel A) is recovered upon 100
consecutive charge/discharge cycles conducted in ionic liquid
electrolyte at 23  C (triangles), corresponding to a moderate fade,
i.e., lower than 0.03% per cycle, with a coulombic efciency quickly
overcoming 99% (panel B). At 40  C, very good capacity retention
(more than 96% of initial capacity delivered after one hundred cycles) and reversibility (efciency approaching 98%) are still
observed in 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI (black squares).
Conversely, pronounced capacity fading (i.e., more than 16% of the
initial capacity is lost after 100 cycles) and lower coulombic efciency (progressively decreasing, upon a few cycles, down to a
stable value around 90%) are detected with the organic electrolyte
(red squares) at 40  C. Such a decay in performance might be ascribable to electrolyte decomposition (e.g., in particular, LiPF6,
which leads to HF formation [22]), resulting in progressively irreversible degradation (promoted by the high operative voltage) of
the NMC material [23], and/or progressive worsening of the electrolyte/Li metal interface. Therefore, at T  40  C, the LiTFSIPYR13TFSI-PYR13FSI ionic liquid electrolyte system performs better,
in terms of cycling behavior and ageing resistance, than the conventional organic ones, resulting appealing for lithium batteries at
medium-high temperature working conditions.
Finally, Fig. 5 summarizes the capacity behavior, obtained at
different current rates and temperatures, for a NMC electrode in the
0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI electrolyte mixture (results are
also shown in Table 1). At low (0.1C) current rates, good capacities,
i.e., from 70 mA h g1 and above, are delivered even at 0  C, which is
one of the highest, if not the highest, capacity values observed for a
cell using an ionic liquid electrolyte well below room temperature
[6]. Therefore, high voltage NMC cathodes can be successfully
combined with LiTFSI-PYR13TFSI-PYR13FSI ionic liquid ternary

Fig. 4. Prolonged cycling performance (panel A) and coulombic efciency evolution


(panel B) of Li/NMC half-cells in: 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI ionic liquid
electrolyte at 23 (triangles) and 40  C (squares); and (1M)LiPF6-EC/DMC (1:1 in
weight) conventional organic electrolyte (circles) at 23  C.

E. Simonetti et al. / Journal of Power Sources 331 (2016) 426e434

Fig. 5. Discharge capacity vs current density dependence of Li/NMC half-cells in


0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI ionic liquid electrolyte at different temperatures.
The current rates are also reported.

electrolytes for realizing safer, high energy density lithium batteries


to be addressed to low-medium power devices operating in wide
temperature ranges. The rate capability at and below room temperature, however, is higher with respect to other ionic liquid
electrolyte systems [6] and expected to be enhanced by incorporating ionic liquids with more proper architecture and/or suitable
non-volatile additives (e.g., glyme) [24]. At medium-high rates
(0.5C) remarkable capacity performance is observed above room
temperature, indicating the feasibility to sustain higher current
densities. Notably, the superior properties in terms of safety, ageing
resistance and cycling behavior, make the LiTFSI-PYR13TFSIPYR13FSI ionic liquid electrolyte system more appealing than
organic solutions for lithium batteries operating at temperatures
40  C.
3.3. Graphite anodes
The lithium intercalation process in graphite anodes from
0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI ionic liquid electrolyte was
investigated by cyclic voltammetries at different temperatures, as
reported in Fig. 6. In the rst cathodic sweep, two broad features,
centered at about 1.0 and 1.5 V vs Li/Li, are observed in the whole
temperature range investigated (from 0 to 40  C). These correspond
to the growth of a protective, lithium ion conducting layer (SEI)
which avoids insertion of (PYR13) cations into the layered graphite
structure [25]. The disappearance of the above mentioned peaks in
the following cathodic sweeps (panels A, C and E) indicates that
these processes (which are irreversible since no corresponding
peaks are recorded in the anodic sweeps) take place only in the rst
cathodic sweep (black trace). Previous work [6,26,27] has indicated
that FSI-containing ionic liquids play a key role with carbonaceous
anodes, because the decomposition products at low potentials
[28,29], are able to form a stable SEI lm [6,26,27]. Below 0.2 V the
Table 1
Summary of the capacity values of NMC cathode tapes in 0.1LiTFSI-0.3PYR13TFSI0.6PYR13FSI ionic liquid electrolyte at different current rates and temperatures.
T/ C

0
10
23
50

Delivered capacity/mA h g1


0.02C

0.035C

0.05C

0.1C

0.2C

0.5C

1C

124.8
153.0
n. a.
n. a.

109.7
143.5
n. a.
n. a.

98.6
132.6
n. a.
n. a.

73.7
114.3
133.5
151.5

15.9
84.1
114.6
143.6

9.5
12.0
57.3
130.0

5.6
6.2
20.0
116.8

431

Li intercalation process takes place. In the following anodic scan,


three features at about 0.1 (only a shoulder), 0.2 and 0.3 V (which
are shifted at 0  C towards higher voltages due to increased IL
electrolyte and interfacial resistances), conrm the reversibility of
the lithium insertion in the LiTFSI-PYR13TFSI-PYR13FSI ionic liquid
electrolyte (e.g., each peak corresponds to the formation of a specic stage in the lithiated graphite anode).
After the voltammetry tests depicted in Fig. 6 (panels A, C
and D), impedance measurements (panels B, D and F) evidence a
remarkable decrease of the overall interfacial resistance with no
relevant change of the electrolyte ionic resistance (e.g., identied
by the high frequency intercept of the AC plots with the real axes
[20]). For instance, around one-order of magnitude decrease are
observed at 0  C (from 3000 to 500 U), 20  C (from 400 to 40 U) and
40  C (from 100 to 10 U). This support for the optimization of the
electrolyte/electrode interfaces upon cycling. For instance, the SEI
passive layer is likely ascribable to improved wettability of the
graphite anode towards the ionic liquid electrolyte, as also suggested by the progressive current rise of the intercalation/deintercalation features recorded.during the rst voltammetric cycles
(panels A, C and E of Fig. 6). In addition, an improvement of the
Li/electrolyte interface may be taken into account as very low
(interfacial) resistances were observed in FSI-based ionic liquid
electrolytes during lithium plating/stripping tests [30]. Notably, a
considerable decrease of the cell impedance, e.g., from 500 to 15 U,
occurs on going from 0 to 40  C, as also indicated by the current
density increases observed in the cyclic voltammetry curves.
The voltage vs capacity proles of selected discharge-charge
cycles, performed on graphite electrodes in different ionic liquid
electrolyte systems (panels from B through D) at 0.1C and 23  C, is
plotted in Fig. 7. The recorded electrode behavior in (1M)LiPF6-EC/
DMC (1:1 in weight) (panel A) is also reported for comparison
purpose. As expected, upon the initial discharge (lithiation), a
feature is detected with all investigated electrolytes, accounting for
the SEI formation onto the graphite electrodes which results in
irreversible capacity loss (electrochemical decomposition of the
electrolyte).
However, in the following lithiation, this irreversible feature
disappears, leaving the typical Li intercalation plateaus (<0.2 V vs
Li/Li) into graphite anodes [6,8,9]. This behavior indicates the good
lm-forming ability of the investigated ionic liquid electrolyte
systems, in good agreement with the voltammetry results of Fig. 6.
Similarly, the charge (delithiation) half-cycles show the corresponding (expected) plateaus due to Li deintercalation. As
observed before for NMC cathodes, graphite anodes show very
good reproducibility upon consecutive cycles (with the exception of
the rst half-cycle), including very good capacity retention and
coulombic efciencies, conrming the high reversibility of the
lithium intercalation process. This behavior, shown in all investigated ionic liquid electrolytes, once more supports for the feasibility and reliability of the natural binder-based anodes as well as
for their eco-manufacturing and compatibility with IL materials. As
observed in panel B, the reversible capacity is found not to exceed
220 mA h g1 in the 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI electrolyte
mixture, showing, however, good cycling ability. Aiming to increase
the anode performance, the PYR13TFSI was replaced with ethyl
methyl imidazolium (EMI) TFSI, thus resulting in capacity increase
up to 280 mA h g1 (panel C) likely ascribable to the better lmforming ability of the imidazolium cation [6] with respect to the
pyrrolidinium one, which derives from the lower cathodic stability
of (EMI) versus (PYR13) [31]. A further improvement is detected
upon addition of modest (5% in weight) ethylene carbonate (EC)
amount to the 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI electrolyte
(panel D), which is seen to push the reversible capacity above
350 mA h g1, i.e., approaching that observed in organic electrolytes

432

E. Simonetti et al. / Journal of Power Sources 331 (2016) 426e434

Fig. 6. Cyclic voltammetries (panels A, C and E) and Nyquist plots (panels B, D and F) of Li/graphite half-cells in 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI ionic liquid electrolyte at
0 (panels A and B), 20 (panels C and D) and 40  C (panels E and F). Scan rate (CVs): 0.1 mV s1. The AC measurements were taken before (open circles) and after (solid circles) the
cyclic voltammetries. The inserts in panels B, D and F magnify the high frequency portion of the AC plots.

(panel A). EC is well known to play a key role during the SEI forming
step onto carbonaceous electrodes [32], allowing growth of more
stable and/or compact passive layers and, therefore, resulting in
benecial effects on the cycling performance. In addition, small EC
contents were seen not depleting the electrolyte safety [33], i.e., the
organic additive is mainly consumed during the rst half-cycle.
The cycling performance of graphite anodes, at 0.1C and 23  C, in
different ionic liquid electrolytes is depicted in Fig. 8. Good capacity
retention is observed for the 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI
electrolyte (black triangles) followed by progressive decay upon 20
cycles, but still keeping remarkable coulombic efciency values. As
also reported in Fig. 7, the replacement of the (PYR13) cation and
(TFSI) anion with (EMI) and (FSI), respectively, leads to capacity
(Table 2) and cycling performance (especially in LiTFSI-EMIFSI)
improvements, once more ascribable to the better lm-forming
ability of the imidazolium and bis(uorosulfonyl)imide ions,
which avoid the unwanted co-intercalation of pyrrolidinium cations. A contemporaneous lowering in initial coulombic efciency is
noticed, e.g., from 88 to 80% for the rst half-cycle, which most
likely results from the more relevant SEI growth. However, the
charge/discharge efciency accounts to 99% at the 10 and 50 cycle

in 0.1LiTFSI-0.3EMITFSI-0.6PYR13FSI (black squares) and 0.1LiTFSI0.9EMIFSI (blues squares) electrolytes, respectively. Therefore,
despite the appealing achieved improvements, pure ionic liquidbased electrolytes are still not able to allow an optimal SEI forming in graphite anodes, thus detrimentally affecting their cycling
performance. A relevant improvement in terms of both the
reversible capacity and the cycling performance is obtained upon
incorporation of just 5 wt% of EC (red squares) into the 0.1LiTFSI0.3PYR13TFSI-0.6PYR13FSI electrolyte. About 350 mA h g1 (Table 2)
were recorded, e.g., approaching the capacity values exhibited in
organic solutions, with very good capacity retention, i.e., about
99% of the initial capacity was delivered after more than 50
consecutive charge/discharge cycles, suggesting the establishment
of a stable, well-protective SEI during the initial discharge of the
graphite anode. In addition, the behavior in 0.1LiTFSI-0.3PYR13TFSI0.6PYR13FSI 5 wt% EC exhibits a stable trend even during the rst
cycles, likely ascribable to a better wettability of this electrolyte
towards graphite anodes with respect to the other investigated IL
electrolyte systems. This represents one of the best results, in terms
of cycling behavior, regarding carbonaceous anodes in ionic liquid
electrolytes systems [6], once more demonstrating the feasibility of

E. Simonetti et al. / Journal of Power Sources 331 (2016) 426e434

433

Fig. 7. Voltage vs capacity prole of Li/graphite cells in (1M)LiPF6-EC/DMC (1:1 in weight) (panel A), 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI (panel B), 0.1LiTFSI-0.3EMITFSI-0.6PYR13FSI
(panel C) and 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI 5 wt% EC (panel D) at 0.1C and 23  C. In panel D the proles recorded at 40  C (red traces) are reported. (For interpretation of the
references to colour in this gure legend, the reader is referred to the web version of this article.)

realizing safer, but still highly performing lithium batteries. Additional work is in progress with the aim of further optimizing the
nature and amount of the additive.
The results of preliminary cycling tests, carried out on graphite/
NMC full cells (cathode limited), are reported as charge/discharge
voltage proles in Fig. SI1 (Supplementary Information). The
0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI 5 wt% EC quaternary mixture
was used as the electrolyte. A nominal capacity exceeding
100 mA h g1 is observed with a moderate initial efciency, likely
due to not optimal anode-cathode mass balance, which is seen to
approach 99% in the following cycles. This conrms suitable formation of SEI onto the graphite anode in the initial charge step,
evidenced by the plateau (missing in the following cycles) around
2.3 V. About 60% of capacity delivered at 0.05C is seen to be discharged at 1C, indicating good rate capability. It is to note that both
the cell design and the electrode weight balance are, however,

susceptible of further optimization. Extended results on full cells


will be reported in coming soon paper.
4. Conclusions
Composite NMC cathodes and graphite anodes, prepared
through an eco-friendly procedure using a water-soluble natural
binder, were studied in 0.1LiTFSI-0.3PYR13TFSI-0.6PYR13FSI ionic
liquid electrolytes at temperatures ranging from 0 to 50  C.
Cyclic voltammetries evidenced good reversibility of the lithium
intercalation process both in NMC and graphite electrodes. The
passivating, poorly ionic conductive layer onto NMC material was
seen to be removed upon the rst cycle above 4 V whereas a protective SEI is stably formed onto the graphite particles, as also
conrmed by impedance measurements.
NMC cathodes exhibited very good cycling behavior up to 50  C
with capacity fading lower than 0.03% per cycle and coulombic
efciency overcoming 99%. At current rate 0.1C capacity values
from 70 to 125 mA h g1 were recorded even at 0  C, not reported
till now in ionic liquid electrolytes below room temperature. At
medium-high rates (0.5C) above 120 mA h g1 were observed at
50  C, indicating the possibility to sustain high current densities. It
is worth to highlight the better performance, in terms of safety,
ageing resistance and cycling behavior, of the LiTFSI-PYR13TFSIPYR13FSI ionic liquid electrolyte system in NMC high voltage
cathodes with respect to organic solutions at temperatures 40  C.
Table 2
Summary of the reversible capacity delivered by graphite anode tapes at 23  C in
different electrolyte systems. Current rate: 0.1C.

Fig. 8. Cycling performance of Li/graphite half-cells in different ionic liquid electrolytes at 0.1C and 23  C. Open data markers: discharge step. Solid data markers: charge
step.

Electrolyte system

Capacity/mA h g1

LiTFSI-PYR13TFSI-PYR13FSI
LiTFSI-EMITFSI-PYR13FSI
LiTFSI-EMIFSI
LiTFSI-PYR13TFSI-PYR13FSI 5 wt% EC
LiPF6-EC-DMC

223.6
276.5
260.5
352.1
371.2

434

E. Simonetti et al. / Journal of Power Sources 331 (2016) 426e434

Graphite anodes were found to exhibit good reversible capacity


(220 mA h g1) with coulombic efciency leveling 100%. Relevant
improvements both in capacity and cycling performance were
detected upon progressive replacement of (PYR13) and (TFSI) with
(EMI) and (FSI) respectively, i.e., up to 280 mA h g1 in 0.1LiTFSI0.3EMITFSI-0.6PYR13FSI and 0.1LiTFSI-0.9EMIFSI, and, especially,
upon incorporation of small ethylene carbonate (EC) amounts
(5 wt%), i.e., above 352 mA h g1 in 0.1LiTFSI-0.3PYR13TFSI0.6PYR13FSI 5 wt%. These results, due to the good lm-forming
ability of (EMI), (FSI) and EC, are among the best ones observed
in carbonaceous anodes with ionic liquid electrolyte systems.
Acknowledgements
The authors wish to acknowledge the EU for funding within the
GREENLION Project (FP7 - Contract no. 285268).
Appendix A. Supplementary data
Supplementary data related to this article can be found at http://
dx.doi.org/10.1016/j.jpowsour.2016.09.078.
References
[1] B. Scrosati, J. Garche, J. Power Sources 195 (2010) 2419.
[2] N. Loefer, J.V. Zamory, N. Laszczynski, I. Doberdo, G.-T. Kim, S. Passerini,
J. Power Sources 248 (2014) 915.
[3] J.R.D. Rogers, K.R. Seddon, Ionic Liquids: Industrial Application to Green
Chemistry (ACS Symposium Series 818), American Chemical Society, Washington (, 2002.
[4] C. Chiappe, D. Pieraccini, J. Phys. Org. Chem. 18 (2005) 275.
[5] H. Ohno (Ed.), Electrochemical Aspects of Ionic Liquids, John Wiley & Sons Inc.,
Hoboken, New Jersey (, 2005.
[6] G.B. Appetecchi, M. Montanino, S. Passerini, Ionic liquid-based electrolytes for
high-energy lithium batteries in ionic liquids: science and applications, in:
A.E. Visser, N.J. Bridges, R.D. Rogers (Eds.), ACS Symposium Series, vol.1117, Oxford
University Press, Inc., American Chemical Society, Washington, DC, USA, 2013.
[7] F. Mueller, N. Loefer, G.-T. Kim, T. Diemant, R.J. Behm, S. Passerini,
ChemSusChem 9 (11) (2016) 1290.
[8] G.T. Kim, S.S. Jeong, M. Joost, E. Rocca, M. Winter, S. Passerini, A. Balducci,
J. Power Sources 195 (2010) 6130.
[9] S.F. Lux, F. Schappacher, A. Balducci, S. Passerini, M. Winter, J. Electrochem.
Soc. 157 (3) (2010) A320.

[10] M. Montanino, F. Alessandrini, S. Passerini, G.B. Appetecchi, Electrochim. Acta


96 (2013) 124.
fer, N. Laszczynski, S. Passerini,
[11] O. Miguel, I. Cendoya, G.-T. Kim, N. Lo
P.M. Schweizer, F. Castiglione, A. Mele, G.B. Appetecchi, M. Moreno,
M. Brandon, T. Kennedy, E. Mullane, K.M. Ryan, M. Olive, I. de Meatza,
GREENLION Project: advanced manufacturing processes for low cost Greener
Li-Ion batteries, in: Electric Vehicle Batteries: Moving from Research towards
Innovation, Book ID 329226, Springer International Publishing, Switzerland,
2014.
[12] M. Moreno, E. Simonetti, G.B. Appetecchi, M. Carewska, M. Montanino, G.-T.
Kim, N. Loefer, S. Passerini, J. Electrochem. Soc. (Special Issue IMLB2106) A1
(2016) 164.
[13] W.A. Henderson, S. Passerini, Chem. Mater 16 (2004) 2881.
[14] J. Kim, Y. Hong, K.S. Ryu, M.G. Kim, J. Cho, Electrochem. Solid-State Lett. 9 (1)
(2006) A19.
[15] G.V. Zhuang, G. Chen, J. Shim, X. Song, P.N. Ross, T.J. Richardson, J. Power
Sources 134 (2004) 293.
[16] X. Zhang, W.J. Jiang, X.P. Zhu, A. Mauger, Qilu, C.M. Julien, J. Power Sources 196
(2011) 5102.
[17] X. Xiong, Z. Wang, P. Yue, H. Guo, F. Wu, J. Wang, X. Li, J. Power Sources 222
(2013) 318.
[18] R. Wang, X. Yu, J. Bai, H. Li, X. Huang, L. Chen, X. Yang, J. Power Sources 218
(2012) 113.
[19] C. Ling, R. Zhang, K. Takechi, F. Mizuno, J. Phys. Chem. C 118 (2014) 26591.
[20] J.R. MacDonald, Impedance Spectroscopy, John Wiley & Sons Editor, New
York, USA, 1987.
[21] F. Lin, I.M. Markus, D. Nordlund, T.-C. Weng, M.D. Asta, H.L. Xin, M.M. Doeff,
Nat. Comm. 5 (2014) 3529.
[22] A.M. Anderson, K. Edstrom, J. Electrochem. Soc. 148 (2001) A1100.
[23] S.B. Chikkannanavar, D.M. Bernardia, L. Liu, J. Power Sources 248
(2014) 91.
[24] D. Shanmukaraj, S. Grugeon, S. Laruelle, M. Armand, ChemSusChem 8 (16)
(2015) 2691.
[25] T. Sugimoto, M. Kikuta, E. Ishiko, M. Kono, M. Ishikawa, J. Power Sources 183
(2008) 436.
[26] G.B. Appetecchi, M. Montanino, A. Balducci, S.F. Lux, M. Winter, S. Passerini,
J. Power Sources 192 (2009) 599.
[27] T. Sugimoto, Y. Atsumi, M. Kikuta, E. Ishiko, M. Kono, M. Ishikawa, J. Power
Sources 189 (2009) 802.
[28] E. Paillard, Q. Zhou, W.A. Henderson, G.B. Appetecchi, M. Montanino,
S. Passerini, J. Electrochem. Soc. 156 (2009) A891.
[29] G.B. Appetecchi, M. Montanino, M. Carewska, F. Alessandrini, S. Passerini,
ECST 25 (36) (2010) 49.
[30] G.-T. Kim, G.B. Appetecchi, M. Montanino, F. Alessandrini, S. Passerini, ECST 25
(36) (2010) 127.
[31] G.B. Appetecchi, M. Montanino, M. Carewska, M. Moreno, F. Alessandrini,
S. Passerini, Electrochim. Acta 56 (2011) 1300.
[32] J.M. Tarascon, M. Armand, Nature 414 (2001) 359.
[33] M. Montanino, M. Moreno, M. Carewska, G. Maresca, E. Simonetti, R. Lo Presti,
F. Alessandrini, G.B. Appetecchi, J. Power Sources 269 (2014) 608.

Das könnte Ihnen auch gefallen