Sie sind auf Seite 1von 9

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

pubs.acs.org/cm

Efficient Aluminum Chloride−Natural Graphite Battery


Kostiantyn V. Kravchyk,†,‡,§ Shutao Wang,†,‡,§ Laura Piveteau,†,‡ and Maksym V. Kovalenko*,†,‡

Laboratory of Inorganic Chemistry, Department of Chemistry and Applied Biosciences, ETH Zürich, Vladimir-Prelog-Weg 1,
CH-8093 Zürich, Switzerland

Laboratory for Thin Films and Photovoltaics, Empa-Swiss Federal Laboratories for Materials Science and Technology,
Ü berlandstrasse 129, CH-8600 Dübendorf, Switzerland
*
S Supporting Information

ABSTRACT: The quest for low-cost and large-scale stationary


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

storage of electricity has led to a surge of reports on novel


batteries comprising exclusively highly abundant chemical
elements. Aluminum-based systems, inter alia, are appealing
because of the safety and affordability of aluminum anodes. In
Downloaded via 83.47.141.56 on June 24, 2018 at 10:40:43 (UTC).

this work, we examined the recently proposed aluminum−ionic


liquid−graphite architecture. Using 27Al nuclear magnetic
resonance, we confirmed that AlCl4− acts as an intercalating
species. Although previous studies have focused on graphitic
cathodes, we analyzed the practicality of achievable energy
densities and found that the AlCl3-based ionic liquid is a capacity-limiting anode material. By focusing on both the graphitic
cathode and the AlCl3-based anode, we improved the overall energy density. First, high cathodic capacities of ≤150 mAh g−1 and
energy efficiencies of 90% at high electrode loadings of at least 10 mg cm−2 were obtained with natural, highly crystalline graphite
flakes, which were subjected to minimal mechanical processing. Second, the AlCl3 content in the ionic liquid was increased to its
maximal value, which essentially doubled the energy density of the battery, resulting in a cell-level energy density of ≤62 Wh
kg−1. The resulting batteries were also characterized by high power densities of at least 489 W kg−1.

■ INTRODUCTION
Secondary (i.e., rechargeable) batteries are deeply integrated
Coulombic efficiency of >99.5%, without formation of
dendrites.35−38 Batteries employing chloroaluminate ionic
into every aspect of human life, including portable devices liquids have received a great deal of attention since a
(smartphones, tablets, and laptops) and transportation publication by the group of Dai in 2015,15 wherein a metallic
(electrical cars, buses, bikes, etc.). Additionally, batteries are Al anode, two synthetic forms of graphite as cathodes (CVD-
increasingly seen as global players in the efficient management grown graphitic foam and pyrolytic graphite), and an
of electricity from sustainable and renewable sources as well as AlCl3:EMIMCl (1-ethyl-3-methylimidazolium chloride) ionic
for the stabilization and decentralization of the electrical grid. liquid were combined into a battery, which is schematically
The simplest management concept is a household utility-level depicted in Figure 1a. This battery design assumes that the
system combining a stationary battery with solar or wind intercalating species are AlCl4− ions, as described in early
electricity.1−3 Even for medium-level storage, the kilowatt-hour electrochemistry reports from the late 1970s.16,39,40 This
cost per charging cycle, in addition to the gravimetric and assumption is in agreement with our 27Al nuclear magnetic
volumetric energy densities, becomes a major consideration for resonance (NMR) spectroscopy results presented below. The
the commercial success of stationary storage systems.4,5 The graphitic cathodes in ref 15 delivered cathodic capacities of 67
volumetric energy densities are the highest for Li-ion batteries. mAh g−1 at an average discharge voltage of 2.0 V. The authors
However, for eventual deployment on the terawatt-hour scale, of ref 15 initially stated that natural forms of graphite exhibited
which is the scale of current worldwide hydroelectric storage, performances much poorer than those of both tested synthetic
low manufacturing costs and high natural abundances of all the graphitic forms. The high porosity of the synthetic graphite was
chemical elements constituting the battery become critical considered an important factor for the fast and highly reversible
factors. intercalation of AlCl4− ions. Subsequent studies by the same
With these goals in mind, battery technologies based on group and by others further increased the capacities and power
highly abundant metals, such as Na,6−10 Mg,11−14 and Al,15−30 densities using porous synthetic graphitic materials, such as
as charge carriers and/or electrodes have become a major graphitic foams41 and graphene nanoribbons (record-high
research focus. In particular, batteries that employ metallic Al as capacity values of 148 mAh g−1).42 However, the use of costly
an anode can harness numerous advantages, such as its high
natural abundance, high charge-storage capacity, and Received: March 14, 2017
safety.31−34 In addition, Al can be reversibly deposited and Revised: May 1, 2017
stripped in chloroaluminate ionic liquids with a high Published: May 1, 2017

© 2017 American Chemical Society 4484 DOI: 10.1021/acs.chemmater.7b01060


Chem. Mater. 2017, 29, 4484−4492
Chemistry of Materials Article

Figure 1. Aluminum chloride−graphite battery. (a) Schematics of the charging process. (b) Comparison of the calculated (curves) and experimental
(data points) cell-level energy densities. The curves are computed from eq I, which describes the dependency of the energy density vs Cc at various
values of r assuming an average discharge voltage of 2 V. The experimental data points represent this work or estimates from reported voltages and
Cc.15,42−45 (c) Graphite ore. (d) Large natural graphite flakes produced commercially from graphite ore.

CVD processes could pose an unsurmountable obstacle for the reduced to Al atoms for electrodeposition and simultaneously
further practical use of such batteries, fully neutralizing the cost generates AlCl4− for intercalation. Therefore, severe require-
benefits associated with the simplicity of the battery and the ments are posed on the amount of used ionic liquid, as the
abundance of the constituent elements. In parallel, similar ionic liquid, in fact, acts as a capacity-limiting liquid anode and
charge-storage capacities were obtained from electrodes made not just as an electrolyte. Thus, we tested AlCl3-rich
of carbon paper,25,27,43 which presumably was produced from formulations of this battery up to the highest achievable
natural graphite,27 and from a few-layer graphene aerogel.44 AlCl3:EMIMCl ratio of 2 and thereby improved the cell-level
The latter was also derived from natural graphite through its energy density to 62 Wh kg−1. We discuss the reasons why a
oxidation into graphite oxide followed by reduction. further substantial increase in energy density is likely
Encouraged by both the initial report and some of these impossible. However, we highlight the high energy efficiencies
recent publications, herein, we sought to thoroughly test the (≤91%) and excellent power densities of the Al chloride−
most basic and inexpensive form of natural graphite, graphite graphite batteries with values of at least 489 W kg−1 at high
flakes, as the cathode material in such Al batteries. We found areal loadings of graphite (≥10 mg cm−2).
that natural graphite flakes, with minimal processing by
sonication, delivered a charge-storage capacity of 148 mAh
g−1, at an average voltage of 2 V. We observed that open
■ RESULTS AND DISCUSSION
Operation Mechanism and Energy Density of the Al
graphite edges and flaky morphology with preserved pristine Chloride−Natural Graphite Flake Battery. Contrary to a
crystalline structures (low density of crystalline defects) were Li-ion battery utilizing graphite as an anode, an Al chloride
crucial for achieving such high charge-storage capacity. Notably, battery exploits the reversible oxidation of the graphite
Dai et al. also reported on natural graphite flakes as cathodes network, i.e., its cathodic functionality. The non-rocking-chair,
during the preparation of this work, using a somewhat different i.e., mixed-ion, operation of the battery is clearly seen from
electrode formulation and achieving a high charge-storage Figure 1a. The following half-reactions take place upon
capacity of 110 mAh g−1 at a current density of 99 mA g−1.45 charging:
We also present direct spectroscopic evidence of the on anode: 4Al 2Cl 7− + 3e− ↔ 7AlCl4 − + Al (I)
intercalation of AlCl4− using solution- and solid-state 27Al
NMR spectroscopy. Furthermore, we analyze and discuss the
on cathode: xC + AlCl4 − ↔ Cx (AlCl4 −) + e− (II)
challenges related to the eventual practical use of this battery
technology. Moreover, we emphasize that an Al chloride− There is no unidirectional flow of Al ions or any other Al
graphite battery is not a rocking-chair battery, contrary to Li- species from one electrode to the other. For this reason, it is
ion batteries, and should not be called an “Al-ion battery” misleading to call this battery an “Al-ion battery”, although this
because the species that intercalate and/or deintercalate into term has become commonplace since the publication by Dai et
the graphitic cathode (AlCl4−) are different from those involved al.15 The actual operation mechanism is, however, fundamen-
in the simultaneous anodic process (i.e., exclusive deposition tally different from that of metal-ion batteries: Al species are
and stripping of Al atoms and/or ions from the Al foil). This is depleted from the liquid phase during the charge process
not only a matter of correct definitions. A portion of AlCl3 acts (Figure 1a). AlCl3:EMIMCl (i.e., EMIM+ + AlCl4−), a mixture
as an anode material during the charging process, whereby it is of two solids that become a liquid at room temperature (an
4485 DOI: 10.1021/acs.chemmater.7b01060
Chem. Mater. 2017, 29, 4484−4492
Chemistry of Materials Article

Figure 2. Electrochemical characterization of the pristine flakes (black) and processed smaller flakes (blue) compared to potato-shaped particles
produced commercially (red) and in house (light green). (a) Representative SEM images of the pristine graphite flakes and of two products of their
mechanical processing by sonication (smaller flakes) and by knife-milling (partially spheroidized, i.e., potato-shaped). (b and d) Galvanostatic charge
and discharge voltage curves and (c and e) cycle stability measurements for the pristine, processed smaller flakes and both potato-shaped specimens
with respect to Li+- or AlCl4−-ion storage. The same color in the legend represents the same material throughout all plots. Electrochemical
measurements in Li-ion half-cells [with 1 M LPF6 in ethylene carbonate/dimethyl carbonate (1:1 vol %) as an electrolyte and metallic lithium as a
counter electrode] and in AlCl3−graphite batteries (r = 1.3). Panel c also plots the Coulombic efficiency, whereas panel e presents the energy
efficiency. Typical Coulombic efficiencies of the AlCl3−graphite batteries are presented in Figure S3.

ionic liquid) due to the acid−base reaction forming AlCl4−, is battery an Al chloride−graphite battery, highlighting the fact
not only an electrolyte but also a source of electroactive species. that AlCl3 can be viewed as an actual anode material dissolved
AlCl3 is added in excess to allow Al2Cl7− ions to form. The in a stoichiometric mixture of AlCl3 and EMIMCl. Therefore,
electrodeposition of Al ions from chloride-based ionic liquids anodic reaction I can be reduced to
has been thoroughly studied in the past, leading to the 3 − 3 1
consensus that Al2Cl7− ions allow the electroplating of AlCl3 + e ↔ AlCl4 − + Al
4 4 4 (III)
Al.35,36,46−53 Al does not deposit from neutral or basic melts
(excess of EMIMCl). Clearly, the amount of Al2Cl7− ions, and The quantity of the neutral AlCl3:EMIMCl melt is linked to
hence the whole mass of the liquid phase, must match the the excessive AlCl3 via r. Considering the whole mass of the
capacity of the graphitic cathode. For each electrodeposited Al liquid anode, the theoretical charge-storage capacity of this
atom, three AlCl4− anions simultaneously intercalate into battery on a cell level can be determined starting from the
graphite. The charging process ends when there are only AlCl4− standard relationship Ctotal = (CACC)/(CA + CC),55,56 wherein
ions left, i.e., the neutral melt forms (AlCl3:EMIMCl = 1), or the weights of both the cathode and liquid anolyte are factored
when the maximal capacity of the graphitic cathode is reached. in
The highest molar ratio (r) of AlCl3 to EMIMCl that still forms Fx(r − 1)CC
a usable ionic liquid is ∼2:1; above this ratio, AlCl3 no longer C total =
Fx(r − 1) + CC(rMAlCl3 + MEMIMCl)
dissolves in the solution. The operation of such a battery is not
uniquely limited to EMIM-based formulations: AlCl3−urea28,54 20.1(r − 1)CC × 103
or fully inorganic NaAlCl4 melts27 have been recently presented =
20.1(r − 1) × 103 + CC(rMAlCl3 + MEMIMCl)
as cost-efficient alternatives. We also note that starting with an
Al anode is not a necessity. Any current collector supporting (1)
−1
the initial electroplating of Al or coated with a minimally thin where F = 26.8 × 10 mAh mol (Faraday constant), x = 3/4
3

Al film as a seed layer is sufficient. Therefore, we call this (number of electrons used to reduce 1 mol of the anodic
4486 DOI: 10.1021/acs.chemmater.7b01060
Chem. Mater. 2017, 29, 4484−4492
Chemistry of Materials Article

material, i.e., AlCl3), r is the AlCl3:EMIMCl molar ratio, CC and have focused exclusively on natural graphite flakes and the use
CA are the specific capacities of the cathode and anode, of minimal processing or the simplest processing of these flakes
respectively, in milliampere-hours per gram, MAlCl3 is the molar by mechanical means. We started with large commercial natural
mass of AlCl3 in grams per mole, and MEMIMCl is the molar graphite flakes (pristine; 0.2−1 mm in lateral size), which were
mass of EMIMCl or of any other Cl− source (i.e., HCl in the commercially produced from graphite ore by mechanical means
simplest case) in grams per mole. Ctotal is an upper bound for a (crushing, wet milling, and floatation). The production is often
capacity, as in the commercial battery it is further reduced by combined with acidic chemical treatment for additional
25−50% because of the weight of current collectors, separators, purification. In addition to using these pristine flakes as
and packaging (determined by the battery design). received, we also fragmented them into smaller particles,
Using eq I, we analyzed the effects of increasing r and CC on resulting in either a retained flat morphology with unfolded
the overall energy density of this battery (E = CtotalV). In Figure edges [processing by sonication; ∼40 μm in lateral size (Figure
1b, we assume an average voltage of the battery to be constant 2a, left)] or an imparted partial spheroidization [∼50 μm;
at 2 V for all theoretical curves. This value of the average processed by knife-milling (Figure 2a, right)]. The particle
discharge voltage is the highest value we obtained exper- shape after spheroidization resembled the “potato” graphite
imentally at r = 1.3. At r > 1.3, the measured average voltage particles commonly used in Li-ion batteries.
dropped progressively by up to 0.2 V [from ∼2 to ∼1.8 V A custom-made cell design with Swagelok fittings was
(Figure S1)]. The primary conclusion is that parameter r is the employed for the electrochemical tests (Figure S2). The cell
most useful variable, in addition to the cathodic capacity of consisted of aluminum (anodic side) and tungsten or glassy
graphite, for improving the energy density of the battery carbon (cathode side) current collectors combined inside a
because of the capacity-limiting effect of the ionic liquid. This fully plastic housing. The cathodes consisted of pure graphite
capacity-limiting effect can be illustrated by calculating the (10 mg cm−2), without the use of any binder or conductive
theoretical CA, considering the whole mass (volume) of the additives, to ensure that the measured electrochemical
ionic liquid as a liquid anode (i.e., as an anolyte). The resulting characteristics could be fully attributed to graphite itself. The
values are 48 mAh g−1 (63 Ah L−1) and 19 mAh g−1 (24 Ah electrodes were moderately pressed against each other with a
L−1) for r values of 2 and 1.3, respectively. Hence, to achieve an glass-fiber separator between them. No preferential orientation
optimal energy density, the highest capacity of the cathode of the graphite flakes with respect to the current collectors was
(CC) must be harnessed and a concomitant increase in r must observed by optical microscopy.
also be included. As a reference point, we analyzed all available All graphite electrodes were tested not only as cathodes in
literature on AlCl3−graphite batteries and identified the highest the AlCl3−graphite cells but also in standard Li-ion cells. All
experimental energy density of ∼45 Wh kg−1 for an such comparative tests were conducted at moderate current
AlCl3:EMIMCl anolyte at r = 1.8, a graphitic capacity of 57 densities of 100 mA g−1. Our findings are indeed orthogonal
mAh g−1 (obtained at r = 1.8), and a corresponding discharge between the experiments of Li-ion storage (Figure 2b,c) and
voltage of 1.88 V (see the Supporting Information of ref 15). intercalation of AlCl4− ions (Figure 2d,e; all Al cells tested at r
All other reported data in the literature were obtained at r = 1.3. = 1.3). In particular, the Li-ion tests readily confirm that the
At this lower value of r, the energy density does not exceed 30 commercial potato-shaped particles and our spheroidized
Wh kg−1 for experimental graphitic capacities of 70−110 mAh particles (also termed potato-shaped in Figure 2) deliver
g−1. An increase in r leads to a concomitant substantial decrease capacities of at least 360 mAh g−1, near the theoretical charge-
in CC and the average discharge voltage, as also mentioned in storage capacity (372 mAh g−1). The 2−3-fold poorer Li-ion
ref 15, and is also apparent from our measurements (Figure storage in the flaky graphite particles is in good agreement with
S1). As detailed below, we have experimentally examined this results of numerous kinetic studies of the effects of their
trade-off for natural graphite flakes, leading to the highest- orientation and microstructure,57−61 which have all highlighted
reported energy density of 62 Wh kg−1 (91 Wh L−1) calculated the accessibility of the edges and particle sizes as key factors for
from a capacity of 124 mAh g−1 (273 Ah L−1) at r = 2 and an Li-ion storage. In this context, a recent study by Billaud et al.62
averaged discharge voltage of 1.77 V (Figure S1). For showed that dense vertical staking of graphite flakes
comparison, the highest graphitic capacity of ∼150 mAh g−1 perpendicular to the current collectors allowed for fast
(330 Ah L−1) was obtained at r = 1.3; nevertheless, this intercalation into laterally large flakes. In the absence of such
corresponds to a much lower energy density of 33 Wh kg−1 (45 favorable orientations, the potato-shaped particles remain an
Wh L−1). optimal morphology for industrial applications. However, a
With the most inexpensive form of graphite (natural flakes) strikingly different scenario is found for the insertion of AlCl4−
as a cathode and all other components being composed of ions. With large pristine flakes, a high capacity of 95 mAh g−1
highly abundant elements (Al, Cl, C, N, and H), the AlCl3− (Figure 2d,e) and a high energy efficiency [90% (Figure 2e)] at
graphite battery might become a viable technology for large- an average discharge voltage of ∼2 V were observed. A further
scale stationary storage purposes. In this domain, this battery capacity increase to 132 mAh g−1 was obtained with smaller
may find use as a nontoxic alternative to the primary battery flakes, retaining the superb flatness of the voltage profiles. On
technologies of similar energy densities: lead−acid (30−50 Wh the other hand, commercial and in-house potato-shaped
kg−1; 50−80 Wh L−1) and vanadium redox-flow batteries graphite particles exhibited lower capacities for AlCl4− ions,
(VRB; 10−30 Wh kg−1; 50−80 Wh L−1).4 AlCl3−graphite which is in striking contrast to the results for Li-ion storage.
batteries exhibit high energy efficiencies (∼90%), which are, Importantly, while the graphite anodes are prone to irreversible
again, on par with those of lead−acid batteries (90%) and VRB capacity loss, as seen from the reduced Coulombic efficiencies
(85%) or even Li-ion batteries (90−95%). of 70−80% (Figure 2d) in the initial cycles as a result of the
Testing Natural Graphite Flakes as Inexpensive formation of a solid−electrolyte interface (SEI), the operation
Cathodes. To comply with stringent cost requirements, we of AlCl3−graphite batteries proceeds from the beginning with
4487 DOI: 10.1021/acs.chemmater.7b01060
Chem. Mater. 2017, 29, 4484−4492
Chemistry of Materials Article

higher Coulombic efficiencies of >92% (Figure S3), suggesting charge, followed by various discharge rates). As seen in the
SEI-free operation. standard CC measurements (Figure 3c; same rates for charge
CCCV Charging Protocol for Combining High Capaci- and discharge), higher discharge capacities were obtained at
ties and High Power Densities. Having established that the charging/discharging current densities of ≤200 mA g−1, which
mechanically processed graphite flakes offer higher charge- decreased at higher rates. A slower CCCV (50 mAh g−1 for a
storage capacities, we examined this form of graphite in greater CC step) protocol indicates the unique rate capability of these
detail (Figure 3). First, we found that the CCCV (constant batteries: the same high discharge capacities were obtained up
to the highest tested discharge rates of 2.5 A g−1, i.e., 3.6 min
for the complete discharge. The corresponding power density
was as high as 489 W kg−1, which compares favorably with
those of the other stationary battery technologies such as lead
acid (75−300 W kg−1)4 and VRB (60−100 W L−1).63
We then examined the effect of the AlCl3 molar fraction by
adjusting parameter r in the range of 1.1−2.0 (highest possible)
while applying the CCCV charging protocol. A clear trend was
observed in which a higher r resulted in a greater loss of voltage
(by ≤0.2 V). Also, the graphitic capacities dropped from ∼150
mAh g−1 (r = 1.3) to ∼124 mAh g−1 (r = 2). These losses are
lower than the theoretically expected energy densities (stars as
data points in Figure 1b). Despite these effects, the highest
energy density of 62 Wh kg−1 was obtained at r = 2,
highlighting the importance of the concerted focus on both the
cathode and the anode.
Benefits of the Flaky Morphology for Achieving
Higher Capacities from Natural Graphite. Powder X-ray
diffraction (XRD) patterns of the charged graphite flakes [a
capacity of ≤132 mAh g−1 (Figure 4a)] indicate that the (002)
graphite peak fully vanishes, and the appearance of three
additional peaks indicates that the intercalation of the ions
occurs in an alternating manner. Such a “staging” intercalation
mechanism is known for various intercalation compounds with
graphite (see the illustration of stages 0−6 in Figure S4).64,65 A
lower stage number indicates a higher intercalant concen-
tration. Recent theoretical calculations24 have shown that
intercalation of AlCl4− into graphite can be observed by the
formation of stages 4, 3, and 2. Simulated XRD peak positions
qualitatively fit the experimental pattern, suggesting a mixed
stage 3−4 intercalation (Figure 4a). The insertion of ions leads
to the expansion of the graphite volume. Larger AlCl4− anions
(295 pm) lead to a drastic expansion by up to 80%, as seen
from recent experimental and theoretical works,24,42,45,54,66−68
which is in sharp contrast to the minor (∼10%) effect from
smaller Li+ ions (90 pm).69 This may explain the highly
beneficial role of the flake morphology: large volumetric
Figure 3. Electrochemical performance of the processed graphite changes are easily accommodated by the homogeneous increase
flakes in the AlCl3−graphite battery (r = 1.3). (a and b) Galvanostatic in thickness. On the other hand, folding, bending, or other
charge−discharge voltage curves and cycling performance measured geometric constrains present in other shapes, such as in potato-
using CC and CCCV protocols. (c) Rate capability measurements shaped graphite, might hamper the high loading of intercalating
obtained with two protocols: standard (same current densities for anions.
charge and discharge during the cycle) and CCCV (50 mAh g−1 It is natural to expect that variations in the synthesis/
CCCV charge and different discharge currents).
processing conditions not only modify the overall shape of the
graphite particles but also introduce various degrees of
current−constant voltage) charging protocol, which involves structural disorder into the idealized lattice of crystalline
constant charging up to 1.9 V followed by one or more graphite, which might also contribute to the electrochemical
constant voltage steps (in the range of 1.9−2.1 V; terminated at performance. For example, powder XRD patterns of pristine,
a current drop of 90%), boosts the subsequent discharge flake, and spherical graphite particles revealed an increase in the
capacity to 150 mAh g−1. It should be noted that constant interplanar graphene−graphene distance from 3.354 to 3.369 Å
voltage steps at higher voltages (>2.1 V) were detrimental (see Table S1), as can be clearly observed from the shift of the
because of side reactions that reduced the Coulombic (002) peak to smaller 2Θ values (Figure 4b).
efficiency. Constant voltage steps at lower voltages (<1.9 V) Similarly, the Raman spectra of the spherical particles were
were of limited effect. Additionally, the rate capability characterized by an increased magnitude of the signal at 1250−
measurements were performed using conventional constant 1400 cm−1 (D-band), which is usually attributed to structural
current (CC) and CCCV protocols (50 mAh g−1 CCCV disorder. Higher defectiveness, as seen from the larger d spacing
4488 DOI: 10.1021/acs.chemmater.7b01060
Chem. Mater. 2017, 29, 4484−4492
Chemistry of Materials Article

capable of removing the ionic liquid. Retention of the charged


state was confirmed by the observation of the same XRD
pattern after the washing step (Figure S5). Importantly, other
solvents that can solubilize the ionic liquid, such as acetonitrile,
methanol, toluene, pentane, tetrahydrofuran, hexane, and
trichloroethylene, expectedly reduced the graphite framework
[i.e., chemically discharged (see further details in Figure S6)].
The 13C NMR spectra showed the absence of CCl4 and
EMIMCl signals after the washing and drying procedures
(Figure S7). The 27Al solid-state NMR spectra of the charged
and washed graphite (Figure 5) exhibited a single peak at 103

Figure 5. 27Al solid-state NMR spectra of charged (blue) and


noncharged (black) small flakes of graphite particles. The noncharged
particles were prepared by impregnation with the AlCl3:EMIMCl
electrolyte, followed by washing with CCl4.

ppm, confirming AlCl4− as the single or dominant form of Al


species that intercalated. This assignment was further
Figure 4. Characterization of the crystalline perfection of various corroborated by a comparison with literature values for Al
graphite materials. (a) X-ray diffraction patterns of sonicated graphite chloride-based ionic liquids71,72 and by the 27Al NMR
flakes before and after charging at a rate of 50 mA g−1 to a capacity of measurements of our reference solutions of acidic (r = 2)
132 mAh g−1. (b and c) XRD and Raman spectroscopy measurements and neutral (r = 1) AlCl3:EMIMCl melts (see Figures S8 and
of the pristine graphite flakes before and after their processing by S9 for further discussions). In another reference measurement
sonication into smaller flakes. For comparison, data for the potato- using 27Al solid-state NMR, no signal was detected on graphite
shaped particles, prepared in house by knife-milling of the pristine
flakes or obtained from the commercial source, are also shown.
flakes that were simply immersed in the electrolyte overnight
and washed with CCl4 (Figure 5).

and more intense Raman D-band (Figure 4b,c), might be the


key factor that explains why the commercial potato-shaped
■ CONCLUSIONS
We have examined the electrochemical behavior of AlCl3−
particles exhibit the lowest charge-storage capacities for AlCl4− graphite batteries, employing natural graphite flakes in
ions (in agreement with the data from ref 45). Overall, despite powdered form as cathodes. We found that the electrochemical
the fact that intercalation of both Li+ and AlCl4− ions into oxidation of the graphite network is accompanied by the
graphite occurs through the edges,70 bending and folding of the intercalation of AlCl4− ions, as evidenced by 27Al NMR
layers as well as atomic and crystalline disorder appear to be spectroscopy. The major factors governing the efficient uptake
much stronger negative factors for the insertion of larger AlCl4− of AlCl4− ions are the flake morphologies of the graphite
ions. particles and the high atomistic structural quality of these
Spectroscopic Evidence of the Incorporation of AlCl4− particles. The highest capacities of ≤150 mAh g−1 were
Ions into Graphite. The conclusion that AlCl4− species obtained with sonicated natural graphite flakes. Considering
intercalate into graphite was corroborated thus far not only by that this battery is not a “rocking-chair” system, we have
XRD measurements but also by X-ray photoelectron (XPS), focused on the question of the practically achievable energy
Auger electron (AES), and X-ray absorption (XAS) spectros- densities. The liquid phase of this battery was estimated to be a
copy measurements.15,27,28,41,43,45 Herein, we present another capacity-limiting liquid anode, which means that the energy
direct spectroscopic probe of the fate of Al species using density is primarily determined by the concentration of the
solution- and solid-state 27Al NMR measurements. For these electroactive species in the ionic liquid rather than by the
measurements, small graphite flakes were charged at 50 mA g−1 cathodic capacity of the graphite. AlCl3−graphite batteries were
up to 2.45 V and then washed from the ionic liquid using characterized by high energy efficiencies of ≤91% and by high
tetrachloromethane (CCl4) as an oxidatively stable solvent power densities (rate capabilities) of 489 W kg−1 at the full
4489 DOI: 10.1021/acs.chemmater.7b01060
Chem. Mater. 2017, 29, 4484−4492
Chemistry of Materials Article

charge. Increasing the AlCl3:EMIMCl ratio to 2:1 and using Assembly and Testing of Li-Ion Batteries. The same plastic
graphite flakes as an anode, we reached an energy density of 62 cells, separators, and graphite loadings (≥10 mg cm−2) were utilized.
Wh kg−1, which is on par with major stationary battery LiPF6 (1 M) in ethylene carbonate and dimethyl carbonate (1:1 vol %)
was used as the Li electrolyte, and metallic Li foils were used as the
technologies. Our analysis shows that these energy densities are
counter and reference electrodes.
difficult to further improve, and we anticipate a practical limit of Structural Characterization of Graphite Particles. Scanning
70−80 Wh kg−1. On one side, the AlCl3:EMIMCl ratio at r = 2 electron microscopy (SEM) images were obtained with a model M400
already contains 64.5 wt % AlCl3. Higher contents of AlCl3 will SEM microscope. The XRD patterns were recorded using a Bruker
be difficult to realize. As an upper bound to this content limit, AXS D8 Advance X-ray diffractometer with Ni-filtered (2 mm
we took the proton instead of the EMIM+ cation in Figure 1b, thickness) Cu Kα radiation (λ = 1.5406 Å) operating at 40 kV and 40
which yielded 88 wt % AlCl3. On the other side, graphitic mA. Each sample was run at least twice using silicon as the internal
capacities of >150 mAh g−1 would have a marginal effect on the standard. For ex situ X-ray diffraction (XRD) studies, fully charged
overall energy density. We suggest that future work should graphite flakes were removed from the cell inside the glovebox and
washed with dried CCl4. To isolate them from the atmosphere, the
focus on the other issues associated with this technology, one charged graphite samples were sealed between Scotch tape and
being the incompatibility of most metallic current collectors exposed to the air shortly prior to the XRD measurements. The sizes
with the corrosive AlCl3-based ionic liquids. Even gold slowly of the potato- and flake-shaped graphite particles were determined by a
dissolves in AlCl3:EMIMCl ionic liquid when electrochemically small-angle light scattering (SALS) instrument, Mastersizer 2000
polarized up to 2.5 V versus Al. Thus far, only tungsten and (Malvern). The sample was diluted with anhydrous ethanol without
molybdenum have been identified as electrochemically stable being filtered. Raman spectroscopy measurements were performed on
current collectors in such batteries. These metals do not fulfill a high-resolution confocal Raman microscope (Ntegra Spectra, NT-
the criterion of high natural abundance, when compared to the MDT) equipped with a 632.8 nm HeNe laser at an incident power of
1.7 mW for excitation at room temperature. Nitrogen adsorption
abundance of the other components of the AlCl3−graphite isotherms were measured at 77 K between 10−3 and 100 kPa using a
battery or to Li ions. There is a pressing need to identify fully computerized Micromeritics ASAP 2020 instrument. Surface
current collectors made of alternative alloys or compounds areas were calculated by the Brunauer−Emmett−Teller (BET)
comprising elements of higher natural abundance and available method between P/P0 values of 0.005 and 1. Prior to the
at low manufacturing costs. measurements, samples were degassed under vacuum (10−1 mbar) at


473 K for 12 h using a Micromeritics VacPrep 061 Sample Degas
System.
EXPERIMENTAL SECTION NMR Spectroscopy Measurements. All solid-state NMR spectra
Chemicals and Battery Components. Pristine graphite flakes were recorded on a Bruker 16.4 T NMR spectrometer equipped with
(99.9%, ∼10 mesh, Alfa Aesar), spherical graphite powder (AE 104L, an Avance III console and a double-resonance 2.5 mm solid-state
TIMCAL), 1-ethyl-3-methylimidazolium chloride (EMIMCl, 99%, probe head. Solution-state NMR spectra were recorded on a Bruker
Iolitec), AlCl3 (99%, granules, Acros), Al foil (MTI Corp.), a W plate 11.7 T spectrometer equipped with a PABBO gradient field three-
(Bocheng Molybdenum Co., Ltd), a glassy carbon plate (Good- channel liquid-state probe head for 5 mm NMR tubes. Experiments
Fellow), and a glass microfiber separator (GF/D, catalog no. 1823- were performed at room temperature without spinning. Chemical
257, Whatman) were used as received. shifts were referenced to Al(NO3)3 in D2O (27Al) and Si(CH3)4 (13C).
Preparation of Electrolytes. Ionic liquid electrolytes based on The numbers of transients were 1024 for the 27Al NMR spectra and
EMIMCl were prepared by slowly mixing EMIMCl solid powder and 10240 for the 13C NMR spectra. A classical Hahn-echo pulse sequence
AlCl3 granules in an argon-filled glovebox. During mixing, a highly was used for the 27Al spectra, and a simple 30°-one-pulse excitation
isothermal reaction occurs, eventually forming a light-yellow liquid. sequence was used for the 13C NMR spectra. All spectra were recorded
Subsequently, the ionic liquid electrolyte was treated with Al foil at without decoupling. The recycle delay was set to 2 s, the echo delay to
150 °C for 6 h until a nearly colorless liquid was seen. 0.2 ms, and the pulse length for a 90° flip angle to 3.4 μs for the 27Al
Sonication of Pristine Graphite Flakes. Pristine graphite flakes NMR spectra. A recycle delay of 2 s was used with a pulse length for a
(0.2 g) were placed into a 4 mL glass vial, and the vial was filled with 30° flip angle of 1.1 μs for the 13C NMR spectra. The number of
3.5 mL of ethanol and sonicated for 30 min using a model HD2200 transients for the solution-state NMR spectra varied from 1 to 16 scans
Sonopuls ultrasonic homogenizer operated at 30% power. The for 27Al NMR and from 64 to 128 for the 13C spectra. The 27Al and
resulting smaller flakes were washed three times with ethanol and 13
C 90° solution-state excitation pulse lengths were 9.5 and 9.8 μs,
dried under vacuum at 80 °C for 12 h. respectively. During the acquisition of the solution-state NMR spectra,
Knife-Milling of Pristine Graphite Flakes (spheroidization). decoupling was performed using a waltz16 sequence while a recycle
Pristine graphite flakes (10 g) were placed in a kitchen blender (Betty delay of 2 s was applied.
Bossi from Fust MixFIT) and knife-milled for 30 min. During the
milling, a nitrogen flow was added to the container to cause more
efficient stirring of the flakes. The knife-milled graphite particles were
dried under vacuum at 80 °C for 12 h.

*
ASSOCIATED CONTENT
S Supporting Information

Assembly and Testing of AlCl3−Graphite Batteries. All The Supporting Information is available free of charge on the
graphite samples were dried at 200 °C under vacuum overnight. No ACS Publications website at DOI: 10.1021/acs.chemma-
binders or solvents were used for the preparation of the electrodes. ter.7b01060.
Homemade plastic cells were assembled in an argon-filled glovebox Additional electrochemical, XRD, Raman, BET, and
(<1 ppm O2, <1 ppm H2O) using three layers of a glass-fiber separator NMR data and photos of the cell used for electro-
soaked with AlCl3:EMIMCl ionic liquid (Figure S2). Aluminum foil
chemical measurements (PDF)


served as both the reference and counter electrodes. A tungsten plate
was used as the cathodic current collector. Graphitic material (10 mg
over ∼1 cm2) was homogeneously distributed on the surface of the AUTHOR INFORMATION
glass-fiber separator before it was covered with the tungsten current Corresponding Author
collector. These cells were cycled between 0.01 and 2.45 V on an *E-mail: mvkovalenko@ethz.ch.
MPG2 multichannel workstation (Bio-Logic). A CCCV protocol was
used at 1.92 and 2.07 V until the current dropped to 10% of the initial ORCID
value. Shutao Wang: 0000-0002-1689-2272
4490 DOI: 10.1021/acs.chemmater.7b01060
Chem. Mater. 2017, 29, 4484−4492
Chemistry of Materials Article

Maksym V. Kovalenko: 0000-0002-6396-8938 (15) Lin, M.-C.; Gong, M.; Lu, B.; Wu, Y.; Wang, D.-Y.; Guan, M.;
Angell, M.; Chen, C.; Yang, J.; Hwang, B.-J.; et al. An ultrafast
Author Contributions rechargeable aluminium-ion battery. Nature 2015, 520, 324−328.
§
K.V.K. and S.W. contributed equally to this work. (16) Fouletier, M.; Armand, M. Electrochemical method for
Notes characterization of graphite-aluminium chloride intercalation com-
pounds. Carbon 1979, 17, 427−429.
The authors declare no competing financial interest. (17) Gifford, P. R.; Palmisano, J. B. An aluminum/chlorine

■ ACKNOWLEDGMENTS
We thank Dr. Frank Krumeich for assistance with SEM
rechargeable cell employing a room temperature molten salt
electrolyte. J. Electrochem. Soc. 1988, 135, 650−654.
(18) Levitin, G.; Yarnitzky, C.; Licht, S. Fluorinated graphites as
energetic cathodes for nonaqueous Al batteries. Electrochem. Solid-State
measurements, Dr. Michael Wörle for assistance with XRD Lett. 2002, 5, A160−A163.
measurements, Mr. Alberto Cingolani and Mrs. Lu Jin for (19) Jayaprakash, N.; Das, S. K.; Archer, L. A. The rechargeable
assistance with SALS, and Mr. Feng Shao for assistance with aluminum-ion battery. Chem. Commun. 2011, 47, 12610−12612.
Raman measurements. This work was financially supported by (20) Rani, J. V.; Kanakaiah, V.; Dadmal, T.; Rao, M. S.; Bhavanarushi,
the Swiss Federal Commission for Technology and Innovation S. Fluorinated natural graphite cathode for rechargeable ionic liquid
(CTI) through the CTI Swiss Competence Centers for Energy based aluminum−ion battery. J. Electrochem. Soc. 2013, 160, A1781−
A1784.
Research (SCCER, “Heat and Electricity Storage”) and by the
(21) Wang, W.; Jiang, B.; Xiong, W.; Sun, H.; Lin, Z.; Hu, L.; Tu, J.;
Competence Center for Energy and Mobility (CCEM, Project Hou, J.; Zhu, H.; Jiao, S. A new cathode material for super-valent
SLIB). battery based on aluminium ion intercalation and deintercalation. Sci.

■ REFERENCES
(1) Alstone, P.; Gershenson, D.; Kammen, D. M. Decentralized
Rep. 2013, 3, 3383.
(22) Hudak, N. S. Chloroaluminate-doped conducting polymers as
positive electrodes in rechargeable aluminum batteries. J. Phys. Chem.
C 2014, 118, 5203−5215.
energy systems for clean electricity access. Nat. Clim. Change 2015, 5, (23) Geng, L.; Lv, G.; Xing, X.; Guo, J. Reversible electrochemical
305−314. intercalation of aluminum in Mo6S8. Chem. Mater. 2015, 27, 4926−
(2) Fairley, P. Energy storage: Power revolution. Nature 2015, 526, 4929.
S102−S104. (24) Jung, S. C.; Kang, Y.-J.; Yoo, D.-J.; Choi, J. W.; Han, Y.-K.
(3) Fares, R. L.; Webber, M. E. The impacts of storing solar energy in Flexible few-layered graphene for the ultrafast rechargeable aluminum-
the home to reduce reliance on the utility. Nat. Energy 2017, 2, 17001. ion battery. J. Phys. Chem. C 2016, 120, 13384−13389.
(4) Chen, H.; Cong, T. N.; Yang, W.; Tan, C.; Li, Y.; Ding, Y. (25) Jiao, S.; Lei, H.; Tu, J.; Zhu, J.; Wang, J.; Mao, X. An
Progress in electrical energy storage system: A critical review. Prog. industrialized prototype of the rechargeable Al/AlCl3−[EMIm]Cl/
Nat. Sci. 2009, 19, 291−312. graphite battery and recycling of the graphitic cathode into graphene.
(5) Chu, S.; Majumdar, A. Opportunities and challenges for a Carbon 2016, 109, 276−281.
sustainable energy future. Nature 2012, 488, 294−303. (26) Stadie, N. P.; Wang, S.; Kravchyk, K. V.; Kovalenko, M. V.
(6) Palomares, V.; Serras, P.; Villaluenga, I.; Hueso, K. B.; Carretero- Zeolite-templated carbon as an ordered microporous electrode for
Gonzalez, J.; Rojo, T. Na-ion batteries, recent advances and present aluminum batteries. ACS Nano 2017, 11, 1911−1919.
challenges to become low cost energy storage systems. Energy Environ. (27) Song, Y.; Jiao, S.; Tu, J.; Wang, J.; Liu, Y.; Jiao, H.; Mao, X.;
Sci. 2012, 5, 5884−5901. Guo, Z.; Fray, D. J. A long-life rechargeable Al ion battery based on
(7) Vogt, L. O.; Marino, C.; Villevieille, C. Electrode engineering of molten salts. J. Mater. Chem. A 2017, 5, 1282−1291.
conversion-based negative electrodes for Na-ion batteries. Chimia (28) Jiao, H.; Wang, C.; Tu, J.; Tian, D.; Jiao, S. A rechargeable Al-
2015, 69, 729−733. ion battery: Al/molten AlCl3-urea/graphite. Chem. Commun. 2017, 53,
(8) Billaud, J.; Clément, R. J.; Armstrong, A. R.; Canales-Vázquez, J.; 2331−2334.
Rozier, P.; Grey, C. P.; Bruce, P. G. β-NaMnO2: a high-performance (29) Choi, S.; Go, H.; Lee, G.; Tak, Y. Electrochemical properties of
cathode for sodium-ion batteries. J. Am. Chem. Soc. 2014, 136, 17243− an aluminum anode in an ionic liquid electrolyte for rechargeable
17248. aluminum-ion batteries. Phys. Chem. Chem. Phys. 2017, 19, 8653−
(9) Singh, G.; Tapia-Ruiz, N.; Lopez del Amo, J. M.; Maitra, U.; 8656.
Somerville, J. W.; Armstrong, A. R.; Martinez de Ilarduya, J.; Rojo, T.; (30) Cohn, G.; Ma, L.; Archer, L. A. A novel non-aqueous aluminum
Bruce, P. G. High voltage Mg-doped Na0.67Ni0.3−xMgxMn0.7O2 (x = sulfur battery. J. Power Sources 2015, 283, 416−422.
0.05, 0.1) Na-ion cathodes with enhanced stability and rate capability. (31) Elia, G. A.; Marquardt, K.; Hoeppner, K.; Fantini, S.; Lin, R.;
Chem. Mater. 2016, 28, 5087−5094. Knipping, E.; Peters, W.; Drillet, J.-F.; Passerini, S.; Hahn, R. An
(10) Walter, M.; Bodnarchuk, M. I.; Kravchyk, K. V.; Kovalenko, M. overview and future perspectives of aluminum batteries. Adv. Mater.
V. Evaluation of metal phosphide nanocrystals as anode materials for 2016, 28, 7564−7579.
Na-ion batteries. Chimia 2015, 69, 724−728. (32) Zafar, Z. A.; Imtiaz, S.; Razaq, R.; Ji, S.; Huang, T.; Zhang, Z.;
(11) Muldoon, J.; Bucur, C. B.; Gregory, T. Quest for nonaqueous Huang, Y.; Anderson, J. A. Cathode materials for rechargeable
multivalent secondary batteries: magnesium and beyond. Chem. Rev. aluminum batteries: current status and progress. J. Mater. Chem. A
2014, 114, 11683−11720. 2017, 5, 5646−5660.
(12) Pan, B.; Huang, J.; Feng, Z.; Zeng, L.; He, M.; Zhang, L.; (33) Fu, L.; Li, N.; Liu, Y.; Wang, W.; Zhu, Y.; Wu, Y. Advances of
Vaughey, J. T.; Bedzyk, M. J.; Fenter, P.; Zhang, Z.; et al. aluminum based energy storage systems. Chin. J. Chem. 2017, 35, 13−
Polyanthraquinone-based organic cathode for high-performance 20.
rechargeable magnesium-ion batteries. Adv. Energy Mater. 2016, 6, (34) Wang, Y.; Chen, R.; Chen, T.; Lv, H.; Zhu, G.; Ma, L.; Wang,
1600140. C.; Jin, Z.; Liu, J. Emerging non-lithium ion batteries. Energy Storage
(13) Aurbach, D.; Lu, Z.; Schechter, A.; Gofer, Y.; Gizbar, H.; Mater. 2016, 4, 103−129.
Turgeman, R.; Cohen, Y.; Moshkovich, M.; Levi, E. Prototype systems (35) Jiang, T.; Chollier Brym, M. J.; Dubé, G.; Lasia, A.; Brisard, G.
for rechargeable magnesium batteries. Nature 2000, 407, 724−727. M. Electrodeposition of aluminium from ionic liquids: Part I
(14) Walter, M.; Kravchyk, K. V.; Ibáñez, M.; Kovalenko, M. V. electrodeposition and surface morphology of aluminium from
Efficient and inexpensive sodium−magnesium hybrid battery. Chem. aluminium chloride (AlCl3)−1-ethyl-3-methylimidazolium chloride
Mater. 2015, 27, 7452−7458. ([EMIm]Cl) ionic liquids. Surf. Coat. Technol. 2006, 201, 1−9.

4491 DOI: 10.1021/acs.chemmater.7b01060


Chem. Mater. 2017, 29, 4484−4492
Chemistry of Materials Article

(36) Jiang, T.; Chollier Brym, M. J.; Dubé, G.; Lasia, A.; Brisard, G. (56) Chockla, A. M.; Klavetter, K. C.; Mullins, C. B.; Korgel, B. A.
M. Electrodeposition of aluminium from ionic liquids: Part II - studies Solution-grown germanium nanowire anodes for lithium-ion batteries.
on the electrodeposition of aluminum from aluminum chloride ACS Appl. Mater. Interfaces 2012, 4, 4658−4664.
(AICl3) - trimethylphenylammonium chloride (TMPAC) ionic liquids. (57) Heß, M.; Novák, P. Shrinking annuli mechanism and stage-
Surf. Coat. Technol. 2006, 201, 10−18. dependent rate capability of thin-layer graphite electrodes for lithium-
(37) Auborn, J. J.; Barberio, Y. L. An ambient temperature secondary ion batteries. Electrochim. Acta 2013, 106, 149−158.
aluminum electrode: its cycling rates and its cycling efficiencies. J. (58) Tran, T. D.; Feikert, J. H.; Pekala, R. W.; Kinoshita, K. Rate
Electrochem. Soc. 1985, 132, 598−601. effect on lithium-ion graphite electrode performance. J. Appl.
(38) Lai, P. K.; Skyllas-Kazacos, M. Electrodeposition of aluminium Electrochem. 1996, 26, 1161−1167.
in aluminium chloride/1-methyl-3-ethylimidazolium chloride. J. (59) Ebner, M.; Chung, D.-W.; García, R. E.; Wood, V. Tortuosity
Electroanal. Chem. Interfacial Electrochem. 1988, 248, 431−440. anisotropy in lithium-ion battery electrodes. Adv. Energy Mater. 2014,
(39) Baiker, A.; Habegger, E.; Schlögl, R. Intercalation of graphite by 4, 1301278.
aluminum chloride. Influence of graphite properties on intercalation (60) Yoshio, M.; Wang, H.; Fukuda, K.; Umeno, T.; Abe, T.; Ogumi,
rate. Berich. Bunsen. Gesell. 1985, 89, 530−538. Z. Improvement of natural graphite as a lithium-ion battery anode
(40) Mohandas, K. S.; Sanil, N.; Noel, M.; Rodriguez, P. material, from raw flake to carbon-coated sphere. J. Mater. Chem. 2004,
Electrochemical intercalation of aluminium chloride in graphite in 14, 1754−1758.
the molten sodium chloroaluminate medium. Carbon 2003, 41, 927− (61) Winter, M.; Besenhard, J. O.; Spahr, M. E.; Novák, P. Insertion
932. electrode materials for rechargeable lithium batteries. Adv. Mater.
(41) Wu, Y.; Gong, M.; Lin, M.-C.; Yuan, C.; Angell, M.; Huang, L.; 1998, 10, 725−763.
Wang, D.-Y.; Zhang, X.; Yang, J.; Hwang, B.-J.; et al. 3D graphitic (62) Billaud, J.; Bouville, F.; Magrini, T.; Villevieille, C.; Studart, A. R.
foams derived from chloroaluminate anion intercalation for ultrafast Magnetically aligned graphite electrodes for high-rate performance Li-
aluminum-ion battery. Adv. Mater. 2016, 28, 9218−9222. ion batteries. Nat. Energy 2016, 1, 16097.
(42) Yu, X.; Wang, B.; Gong, D.; Xu, Z.; Lu, B. Graphene (63) Mohamed, M. R.; Sharkh, S. M.; Walsh, F. C. In Redox flow
nanoribbons on highly porous 3D graphene for high-capacity and batteries for hybrid electric vehicles: Progress and challenges. 2009
ultrastable Al-ion Batteries. Adv. Mater. 2017, 29, 1604118. IEEE Vehicle Power and Propulsion Conference, September 7−10;
(43) Sun, H.; Wang, W.; Yu, Z.; Yuan, Y.; Wang, S.; Jiao, S. A new IEEE: New York, 2009; pp 551−557.
aluminium-ion battery with high voltage, high safety and low cost. (64) Dresselhaus, M. S.; Dresselhaus, G. Intercalation compounds of
Chem. Commun. 2015, 51, 11892−11895. graphite. Adv. Phys. 1981, 30, 139−326.
(44) Chen, H.; Guo, F.; Liu, Y.; Huang, T.; Zheng, B.; Ananth, N.; (65) Noel, M.; Santhanam, R. Electrochemistry of graphite
Xu, Z.; Gao, W.; Gao, C. A defect-free principle for advanced graphene intercalation compounds. J. Power Sources 1998, 72, 53−65.
cathode of aluminum-ion battery. Adv. Mater. 2017, 29, 1605958− (66) Bhauriyal, P.; Mahata, A.; Pathak, B. The staging mechanism of
1605965. AlCl4 intercalation in graphite electrode for aluminium-ion battery.
(45) Wang, D.-Y.; Wei, C.-Y.; Lin, M.-C.; Pan, C.-J.; Chou, H.-L.; Phys. Chem. Chem. Phys. 2017, 19, 7980−7989.
Chen, H.-A.; Gong, M.; Wu, Y.; Yuan, C.; Angell, M.; et al. Advanced (67) Wu, M. S.; Xu, B.; Chen, L. Q.; Ouyang, C. Y. Geometry and
rechargeable aluminium ion battery with a high-quality natural fast diffusion of AlCl4 cluster intercalated in graphite. Electrochim. Acta
graphite cathode. Nat. Commun. 2017, 8, 14283. 2016, 195, 158−165.
(46) Lai, P. K.; Skyllas-Kazacos, M. Aluminium deposition and (68) Gao, Y.; Zhu, C.; Chen, Z.; Lu, G. Understanding ultrafast
dissolution in aluminium chloriden-butylpyridinium chloride melts. uechargeable aluminum-ion battery from first-principles. J. Phys. Chem.
Electrochim. Acta 1987, 32, 1443−1449. C 2017, 121, 7131−7138.
(47) Chao-Cheng, Y. Electrodeposition of aluminum in molten (69) Abe, T.; Ogumi, Z. Nano-aspects of carbon negative electrodes
AlCl3-n-butylpyridinium chloride electrolyte. Mater. Chem. Phys. 1994, for Li ion batteries. In Nanoscale Technology for Advanced Lithium
37, 355−361. Batteries; Osaka, T., Ogumi, Z., Eds.; Springer: New York, 2014; pp
(48) Zhao, Y.; VanderNoot, T. J. Electrodeposition of aluminium 31−40.
from nonaqueous organic electrolytic systems and room temperature (70) An, S. J.; Li, J.; Daniel, C.; Mohanty, D.; Nagpure, S.; Wood, D.
molten salts. Electrochim. Acta 1997, 42, 3−13. L. The state of understanding of the lithium-ion-battery graphite solid
(49) Zein El Abedin, S.; Moustafa, E. M.; Hempelmann, R.; Natter, electrolyte interphase (SEI) and its relationship to formation cycling.
H.; Endres, F. Electrodeposition of nano- and microcrystalline Carbon 2016, 105, 52−76.
aluminium in three different air and water stable ionic liquids. (71) Taulelle, F.; Popov, A. I. Aluminum-27 NMR study of some
ChemPhysChem 2006, 7, 1535−1543. AlCl3−MCl molten systems. J. Solution Chem. 1986, 15, 463−471.
(50) Abood, H. M. A.; Abbott, A. P.; Ballantyne, A. D.; Ryder, K. S. (72) Nakayama, Y.; Senda, Y.; Kawasaki, H.; Koshitani, N.; Hosoi, S.;
Do all ionic liquids need organic cations? Characterisation of [AlCl2· Kudo, Y.; Morioka, H.; Nagamine, M. Sulfone-based electrolytes for
aluminium rechargeable batteries. Phys. Chem. Chem. Phys. 2015, 17,
nAmide]+AlCl4− and comparison with imidazolium based systems.
5758−5766.
Chem. Commun. 2011, 47, 3523−3525.
(51) Abbott, A. P.; Harris, R. C.; Hsieh, Y.-T.; Ryder, K. S.; Sun, I. W.
Aluminium electrodeposition under ambient conditions. Phys. Chem.
Chem. Phys. 2014, 16, 14675−14681.
(52) Bakkar, A.; Neubert, V. A new method for practical
electrodeposition of aluminium from ionic liquids. Electrochem.
Commun. 2015, 51, 113−116.
(53) Fang, Y.; Yoshii, K.; Jiang, X.; Sun, X.-G.; Tsuda, T.; Mehio, N.;
Dai, S. An AlCl3 based ionic liquid with a neutral substituted pyridine
ligand for electrochemical deposition of aluminum. Electrochim. Acta
2015, 160, 82−88.
(54) Angell, M.; Pan, C.-J.; Rong, Y.; Yuan, C.; Lin, M.-C.; Hwang,
B.-J.; Dai, H. High coulombic efficiency aluminum-ion battery using an
AlCl3-urea ionic liquid analog electrolyte. Proc. Natl. Acad. Sci. U. S. A.
2017, 114, 834−839.
(55) Reddy, T. Linden’s handbook of batteries, 4th ed.; McGraw-Hill
Education: New York, 2010.

4492 DOI: 10.1021/acs.chemmater.7b01060


Chem. Mater. 2017, 29, 4484−4492

Das könnte Ihnen auch gefallen