You are on page 1of 10

Article

Cite This: ACS Appl. Energy Mater. 2019, 2, 1793−1802 www.acsaem.org

Silicon Few-Layer Graphene Nanocomposite as High-Capacity and


High-Rate Anode in Lithium-Ion Batteries
Stefano Palumbo,*,†,‡,§ Laura Silvestri,† Alberto Ansaldo,† Rosaria Brescia,∥
Francesco Bonaccorso,*,†,⊥ and Vittorio Pellegrini†,⊥

Graphene Laboratories and ∥Electron Microscopy Facility, Istituto Italiano di Tecnologia (IIT), Via Morego 30, 16169 Genova,
Italy

Dipartimento di Elettronica e Telecomunicazioni, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy
§
Metrology for Quality of Life Departimento, Istituto Nazionale di Ricerca Metrologica (INRIM), Strada delle Cacce 91, 10135,
Downloaded via KTH ROYAL INST OF TECHNOLOGY on September 20, 2022 at 09:29:45 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Torino, Italy

Bedimensional S.p.A., Via Albisola 121, 16163 Genova, Italy
*
S Supporting Information

ABSTRACT: A silicon-graphene heterostructure provides optimal


electrochemical performance as anode nanomaterial in both half and
full cells with a commercial NMC111 (LiNi1/3Mn1/3Co1/3O2) cathode.
The anode consists of carbon-coated polycrystalline silicon nano-
particles in between a parallel oriented few-layers graphene flakes
(FLG). Electrochemical tests in lithium cells display high capacity
values (∼2300 mAh/g) with a Coulombic efficiency (CE) reaching
99% at current density of 350 mA/g and 1000 mAh/g at current
density values up to 3.5 A/g (CE = 99%). The laminated graphene-based structure yields a protective coating to the silicon
nanoparticles still enabling exposure to lithium ions. The method of production of the laminated silicon-graphene
nanocomposite is scalable and low-cost, offering a practical route to the introduction of high silicon content anodes in lithium-
ion batteries.
KEYWORDS: silicon, graphene, nanostructured composite, large scale production, high capacity anode, lithium-ion batteries

■ INTRODUCTION
The continuous development of lithium-ion battery (LIB)
conversion14 are exploited to overcome these limitations,
leading to higher capacity values, which consequently increase
technology1−4 over the years has led to large improvements in the energy density of the currents LIBs.10,15
terms of specific capacity, energy density, and cycle life Silicon, currently, represents one of the most promising
stability, enabling the introduction of LIBs into the market of alloying anode materials to replace graphite-based ano-
hybrid and electric vehicles.5 For example, LIBs increased their des.13,16−18 Silicon can theoretically react in a lithium cell
efficiency by 5−10% every year for the past 25 years.6 In fact, through an alloying reaction to form Li4.4Si, achieving a specific
while in the 1990s the highest energy density value for LIBs capacity of 4200 mA hg−1.19 However, several issues affect
was ∼90 Wh kg−1,7 currently their value approaches 260 Wh silicon, as well as other alloying materials, hindering its
kg−1.5 implementation in commercial batteries.20 In particular, the
Nevertheless, development of LIBs able to satisfy the main issue of silicon is associated with the significant volume
growing demand in terms of energy densities and fast charging changes (>300%) during the lithiation/delithiation pro-
times poses new challenges, which cannot be overcome by cesses.21 The volume change produces cracks and pulverization
exploiting conventional electrode materials, due to their of the electrode leading to the loss of electrical contact and the
intrinsic theoretical limitations.6,8,9 consequent poor cycling performance.22 The continuous
Conventional LIBs consist of graphite-based anodes and volume expansion/contraction of silicon during cycling also
lithiated transition metal oxide cathodes, such as LiCoO2.2 negatively affects the electrode−electrolyte interface.23 In fact,
These cells typically supply ∼120 Wh kg−1.2 In these the exposure of fresh electrode surface (due to expansion)
electrochemical systems, the redox process is based on determines a continuous electrolyte reduction during the
intercalation/deintercalation of lithium ions between the two subsequent cycles, resulting in an unstable SEI (solid
layered materials,10 a mechanism that displays fundamental electrolyte interface) layer formation.23
limitations preventing further improvements. In fact, consid-
ering the anode, for example, graphite is able to exchange only Received: November 6, 2018
1 Li+ every 6C atoms, achieving a theoretical specific capacity Accepted: February 4, 2019
of 372 mA hg−1.11,12 Alternative processes such as alloying13 or Published: February 4, 2019

© 2019 American Chemical Society 1793 DOI: 10.1021/acsaem.8b01927


ACS Appl. Energy Mater. 2019, 2, 1793−1802
ACS Applied Energy Materials Article

To overcome the aforementioned issues several strategies of silicon nanoparticles onto the graphene surface,78,79,85
have been adopted mainly focusing on the design of advanced demonstrated to be more effective in optimizing the perform-
electrode structures,24 including, in particular, the exploitation ance in a lithium cell, but with the downside of being hardly
of silicon nanoparticles25 and their encapsulation into a carbon compatible with large scale industrial production.86−88
matrix.26 Here, we present a laminated silicon-FLG composite to be
Several experimental works25,27,28 have proven the existence used as anodic material in a lithium cell that combines a facile
of a relationship between silicon particle size and the and low-cost method for preparation with high and stable
electrochemical performance, demonstrating that the use of electrochemical performance.
non-agglomerated nanosized silicon can suppress particle Pristine materials include commercial silicon nanoparticles
cracking upon lithiation. However, besides controlling the and FLG flakes produced by a wet jet mill (WJM) process,
size of silicon particles, their morphology (i.e., nano- which enables the scaling-up of the production of high quality
spheres,29−31 nanowires32 or nanotubes33) also is known to few-layers graphene.89 Electrode preparation exploits a three-
affect the electrochemical properties. step process. First, by using ultrasonication, FLG flakes and
In addition to the nanostructuring approach,29−33 the silicon nanoparticles are dispersed in a common solvent. Then,
realization of a composite of silicon nanoparticles with a the obtained slurry is directly deposited on the copper current
conductive carbon material represents another established way collector. Finally, the electrode is subjected to a thermal
to mitigate the volumetric changes, thus improving the treatment in a reducing atmosphere.90
performance of the electrode in terms of both cycle life and We show that the as-prepared electrodes are formed by FLG
rate, ensuring the electrical contact of the electrode.34−36 This flakes with a preferential orientation parallel to the electrode
approach proved to be effective only when a carbon surface, that entrap polycrystalline carbon-coated silicon
nanostructure is able to enclose the silicon nanoparticles.34−36 nanoparticles with overall silicon content as high as 57%.
In fact, only minor improvements are observed by simply The resulting laminated silicon-FLG electrode shows high
mixing silicon nanoparticles with an amorphous carbon or by specific capacity values (>2000 mAhg−1), being able to sustain
carbon coating methods.17,37 In contrast, more complex current densities up to 7 A/g, with a Coulombic efficiency
strategies, such as the encapsulation of silicon into carbon (CE) of 99%. Furthermore, we also demonstrate the use of
nanostructures with empty spaces,38 as well as their infiltration silicon-FLG electrode in a full lithium-ion cell using NMC111
into a nanoporous matrix, allow silicon nanoparticles to freely (LiNi1/3Mn1/3Co1/3O2) as the positive electrode. Electro-
expand and contract, with a positive effect on cyclability.18 chemical tests at C/2 show that the cell can exchange 1800
Among the explored conductive carbon structures,34−36 mAhg−1 (referred to silicon mass) reaching a CE of 99%.
graphene39−41 has recently emerged as a promising silicon These results surpass previous attempts to develop commer-
partnering material owing to its excellent mechanical proper- cially suitable solutions of high silicon content composites for
ties42 and electrical conductivity.43 Several approaches have Li-ion cells.
been proposed over the last years to successfully integrate
graphene and silicon nanoparticles to obtain anodes with long
cycle life.44−56
■ MATERIAL AND METHODS
Materials. Silicon nanoparticles (<100 nm diameter) were
So far, the most common graphene-based material used in purchased from Alfa-Aesar. Few-layer graphene flakes were prepared
combination with silicon is reduced graphene oxide by WJM, following the procedure developed in our laboratories89 (see
(RGO).57−61 In fact, the presence of an oxygen-containing next section for details). NMC111 coated on Al foil (16 mg/cm2) was
functional group makes the RGO flakes hydrophilic, thus provided by Varta Micro Innovation gmbH.
enabling the possibility of using a large variety of solvents, e.g., Wet-Jet Mill Exfoliation of Graphite. 200 g of graphite flakes
water.41,62 This means that through simple synthetic strategies, (+100 mesh, Sigma-Aldrich) was mixed with 20 L of N-methyl-2-
pyrrolidone (NMP) (>97%, Sigma-Aldrich) by a mechanical stirrer
graphene can be functionalized or modified with electroactive (Eurostar digital Ika-Werke).89 A hydraulic mechanism and a piston
species through covalent or noncovalent bond.63−65 Another supplied a pressure between 180 and 250 MPa to push the as-
strategy relies on the use of single- or few-layer graphene (SLG prepared mixture into a set of 5 different perforated and
and FLG, respectively), which, on one hand, accommodates interconnected disks, called processor. Two jet streams were
the electroactive material and, on the other hand, provides a generated at the second disk, which have two holes with diameters
much more effective pathway for electron conductivity.66,67 of 1 mm. Then, the jet streams collided between the second and the
Further examples to improve the performance of silicon-based third disk, which consists of a nozzle (i.e., a half-cylinder channel) of a
electrodes make use of nitrogen-doped graphene.68 In fact, it diameter of 0.3 mm. The shear force generated by the solvent when
the sample passes through such nozzle promotes the graphite
was experimentally demonstrated69−72 that the doping of
exfoliation.89 The sample was then separated in two jet streams,
graphene flakes with nitrogen atoms modifies the graphene recombining in the last disk before leaving the processor. Immediately
electronic structure73 and improves the lithium diffusion after the processor, the sample was cooled down by means of a chiller.
kinetics and transfer,74 resulting in a beneficial effect on its The processed sample was then collected. The wet-jet milling process
electrochemical performance.75−77 was repeated passing the sample through the 0.15 mm nozzle. Finally,
Different methods have been explored to produce silicon- a third exfoliation step was carried out by adjusting the diameter of
graphene composites.26,66,67,78−83 Although the simple physical the nozzle to 0.10 mm. The dispersion in NMP was then subjected to
mixing of silicon nanoparticles and graphene flakes could rotovapor and freeze-drying to completely remove the solvent.
represent an easy way to have a low cost and scalable method Electrode Preparation. Silicon-FLG electrodes were prepared by
dispersing equal masses of silicon nanoparticles, FLG flakes and
for the industrial production of electrodes,80,81 it has not poly(acrylic acid) (PAA) in ethanol by ultrasonication for 90 min.
always been validated by the cell electrochemical perform- The mass loading of the active material was 30 mg/mL.
ance.82 Alternative methodologies enabling the in situ growth 50 μL of the obtained dispersion (corresponding to a mass of 1.5
of graphene in order to obtain a controlled morphology of the mg of active material) was subsequently deposited, by drop casting,
silicon-graphene composite26,66,67,83,84 as well as the synthesis onto a copper disk (15 mm diameter) and then dried in air for 5 min.

1794 DOI: 10.1021/acsaem.8b01927


ACS Appl. Energy Mater. 2019, 2, 1793−1802
ACS Applied Energy Materials Article

Figure 1. (a−c) Cross section and top view SEM micrographs of the silicon-FLG electrode before and (d−f) after annealing.

Finally, the electrodes were annealed in a tubular furnace under a two components of the doublet and a doublet separation of 0.6 eV
controlled atmosphere (vacuum + 5 sccm of H2) for 30 min at 750 °C have been assumed.92
with a heating rate of 15 °C/min up to 700 °C slowing down to 5 °C/ To evaluate the silicon content in the composite, a thermogravi-
min to reach 750 °C. To preserve the samples from an undesired metric analysis (TGA) was performed using a TGA Q500-TA
bending of the copper disk, a steel foil was superimposed during the Instrument. As analytic gas, air with a flux of 50 mL/min was implied.
process. The method used during the measurements provides an initial
Physical-Chemical Characterization. The as-prepared silicon- stabilization of the sample at 30 °C for 5 min, and then a heat rump of
FLG electrode was characterized via a set of techniques. Scanning 5 °C/min up to a final temperature of 800 °C maintained for 1 h.
electron microscopy (SEM) images of the electrodes were acquired by For electrochemical characterization, coin cells 2032 were
a JEOL JSM-6490LA instrument. High-resolution transmission assembled facing the as-prepared silicon-graphene electrode with a
electron microscopy (HRTEM) and energy-filtered TEM (EFTEM) lithium foil. A solution consisted of LP30 (1 M LiPF6 in DMC:EC)
analyses were carried out on an image-corrected JEOL JEM-2200FS and 10% v/v FEC (fluoro-ethylene carbonate) embedded in a
TEM (Schottky emitter source), operated at 200 kV, equipped with Whatman borosilicate separator was used as electrolyte. The
an in-column image filter (Ω-type) and a Bruker XFlash 5060 SDD assembled lithium cells were tested in a Biologic VMP3 potentiostat
energy-dispersive X-ray spectrometer (EDS). EFTEM maps reported through galvanostatic cycling and potentiodynamic cycling with
here were obtained using the three-window method at the C K and O galvanostatic acceleration (PCGA) measurements. Galvanostatic
K ionization edges. To prepare the grid for TEM analysis, the cycling was carried out using the potential range of 0.1−1.0 V at
electrodes were first sonicated in ethanol and then 100 μL of each different current density from 0.35 A/g to 7 A/g. Electrochemical
dispersion was drop cast onto a holey amorphous carbon-coated impedance spectra were acquired upon cycling by applying a potential
copper grid. signal of 5 mV amplitude in the frequency range 100 kHz to 100
Raman measurements were carried out for silicon-FLG electrodes mHz. The PCGA measurements were performed setting a potential
before and after the annealing process. The Raman spectra were step of 5 mV and a cutoff of C/20 between 0.1 and 1.0 V. Before any
acquired by using a Renishaw inVia micro-Raman spectrometer with electrochemical test, a formation cycle was performed between 0.01
the 514.5 nm line of an Ar+ laser. The incident laser was focused by a and 2.0 V and applying a current density of 0.35A/g, corresponding to
50× optical microscope objective with a numerical aperture of 0.95 a C rate of ∼C/10.
and then the scattered light was detected in a backscattering geometry Finally, a complete lithium ion cell was assembled using silicon-
dispersed by a 2400 l/mm holografic grating. All the spectra were FLG electrode as anode, a commercial NMC111, with a nominal
collected by acquiring 3 accumulations of 30 s each. Acquired spectra capacity of 2 mAh/cm2 electrode as cathode and LP30 + 10% v/v of
FEC as electrolyte. The electrochemical behavior of the as-assembled
were analyzed by OriginLab OriginPro 2016 using Voight curves in
cell was tested by galvanostatic cycling in the potential range of 2.8−
order to deconvolute the main Lorentzian components.
4.0 V, using a C-rate of C/2.
X-ray diffraction (XRD) patterns were recorded on a PANalytical
In order to provide a stable electrode−electrolyte interphase, a
Empyrean X-ray diffractometer equipped with a 1.8 kW CuKα
formation procedure was carried out before cycling the Li-ion cell and
ceramic X-ray tube, PIXcel3D 2 × 2 area detector and operating at 45
consisting of (i) a first charge to 4.0 V at C/10, followed by a
kV and 40 mA. The diffraction patterns were collected in air at room potentiostatic step with a current cutoff corresponding to C/20; (ii)
temperature using Parallel-Beam (PB) geometry and symmetric first discharge at 2.8 V at C/10 followed by a rest time of 3 h; then
reflection mode. X-ray diffraction data analysis was carried out using (iii) two charge−discharge cycles between 4.0 and 2.8 V at C/2, with
HighScore 4.7 software91 from PANalytical. a potentiostatic step at 4.0 V with a current cutoff corresponding to
X-ray photoelectron spectroscopy (XPS) measurements were C/20.
performed on a Kratos Axis UltraDLD spectrometer, using a For all the electrochemical tests, specific capacity is referred to the
monochromatic Al Kα source operated at 15 kV and 20 mA. High- mass of silicon.


resolution narrow scans were acquired on Si 2p core levels at constant
pass energy of 10 eV in steps of 0.1 eV. The photoelectrons were RESULTS AND DISCUSSION
detected at a takeoff angle of Φ = 0° with respect to the surface
normal. The data were processed with Casa XPS software v 2.3.17. Morphological and Structural Characterization of the
Silicon spectra fitting has been performed by considering Voigt Silicon-FLG Electrodes. The structural and morphological
profiles; for elemental silicon, an intensity ratio of 2:1 between the properties of the FLG produced by WJM are reported in ref
1795 DOI: 10.1021/acsaem.8b01927
ACS Appl. Energy Mater. 2019, 2, 1793−1802
ACS Applied Energy Materials Article

Figure 2. (a) Superposition of the normalized Raman spectra for the Si-FLG anode pre-annealing (red) and post-annealing (blue). (b−c) HRTEM
images and corresponding fast Fourier transforms (FFTs) for regions in the (b) pristine and (c) annealed Si-FLG electrode on top of an
amorphous carbon film. The FFTs indicate a [001]-oriented graphite (ICSD 76767) pattern, due to the FLG flakes, and randomly oriented cubic
silicon (ICSD 51688) nanometer-sized domains.

89. Here we report the morphological characterization of the An explanation of these features can be linked to the carbon-
as-prepared silicon-FLG electrodes carried out by SEM. Figure content present in the binder that turned into a carbon coating
1 reports the comparison of the electrodes before (a−c) and after the annealing process. This hypothesis was then further
after (d−f) the annealing process. explored by means of HRTEM and EFTEM imaging. A deeper
Figure 1a,b shows the cross section of the electrode once analysis on the Raman spectra is reported in Figure S3.
deposited onto the copper foil. We can clearly distinguish the The fast Fourier transform (FFT) of the HRTEM image,
two components, with the isotropic silicon nanoparticles in shown in Figures 2b,c, clearly show the [001] orientation of
between the FLG flakes. Observing the top view of the the FLG flakes, while the silicon spots are arranged in rings,
electrode in Figure 1c, we can see also on the surface of the corresponding to polycrystalline silicon.
electrode the presence of silicon nanoparticles and the FLG Further details of the electrodes have been highlighted by
flakes. EFTEM elemental mapping and XPS measurements.
After the annealing process, the electrode morphology is Figure 3 shows the EFTEM analysis for the silicon-graphene
electrode pre-annealing (Figure 3a−c) and post-annealing
significantly changed. The electrode cross section in Figure 1d
shows that the graphene flakes are oriented parallel one to each
other, the dimension of the silicon aggregates appear reduced
and they are homogeneously distributed between the FLG
flakes. As evident in Figure 1e, FLG flakes entrap the silicon
nanoparticles creating a well-organized and ordered structure.
Differences between electrodes before and after annealing are
also evident in the surface of the electrode (Figure 1f), where
FLG flakes cover a higher surface compared to the electrode
without annealing.
The aim of the annealing process is twofold: (i) achieve a
higher conductivity of the composite by the formation of a
carbon coating over the Si-NPs and in between, i.e., a better
electrical contact between the Si-NPs and the FLG flakes, and
(ii) increase the anode specific capacitance by removing the
binder. To validate this hypothesis, Raman spectra in the
1100−1900 cm−1 range were acquired. Figure 3. (a,d) Zero-loss filtered Bright Field TEM (BF-TEM) image
The spectra were then normalized to the intensity of the and (b,c, e,f) corresponding EFTEM elemental maps for carbon (red)
carbon G peak and the peaks were deconvoluted in their and oxygen (blue) on regions of (a−c) pristine and (d−f) annealed
Lorentzian components. In Figure 2a the typical Raman Si-FLG electrode, partly suspended on a hole in the holey carbon
spectra of pristine and annealed samples are reported. support film.
On the pristine sample spectrum, the typical components of
graphene spectrum in the range acquired (D, G, and D′93) can
be clearly identified. The peaks are well-defined and separated (Figure 3d−f). By comparing the C maps in Figure 3b and e it
and the I(G)/I(D) ratio is 0.064. On the contrary, after the is possible to note an inhomogeneous 3-nm-thick carbon
annealing process, the intensity of the D peak was strongly coverage all over the silicon nanoparticles, in the case of the
increased, and the D and G peaks are partially overlapping. annealed sample. On the contrary, in the pristine sample there
By the deconvolution of the Raman spectrum, five different is a low amount of carbon. Furthermore, a clear surface
components can be identified. Of them, three can be ascribed oxidation is observed in the annealed sample (Figure 3f), less
to the graphene flakes, i.e., D, G, and D′ peaks, while the other prominent in the pristine sample (Figure 3c). From the EDS
two reveal the typical features of the amorphous carbon. quantification, the annealed sample is clearly much richer in
1796 DOI: 10.1021/acsaem.8b01927
ACS Appl. Energy Mater. 2019, 2, 1793−1802
ACS Applied Energy Materials Article

carbon and slightly richer in oxygen compared to the pristine


sample (Figure S2).
The electrode surface has been further analyzed by XPS
measurements. Figure 4 reports the spectra obtained for the

Figure 5. PCGA response of the annealed silicon-graphene electrode


in lithium cell. Voltage range used for the formation cycle is 0.01−2 V,
while for the other cycles it is in the 0.1−1 V range.

associated with the redox process of FLG, are observable at 0.1


and 0.15 V, followed by two large peaks at 0.3 and 0.45 V
(labeled C and D, respectively) corresponding to the multistep
delithiation process of amorphous Li−Si alloys.96 In the
Figure 4. XPS spectrum comparison of the silicon particles and of the second cycle, limiting the potential cut-offs to 0.1−1 V, we also
silicon-FLG electrode post-annealing. limited the contribution of FLG flakes into redox process,
while the two-anodic process centered at 0.25 and 0.1 V,
according to a-Si + xLi+ → a-LixSi; a-LixSi + yLi+ → a-
annealed silicon-FLG electrode in comparison to the non- Li(y+x)Si,96 and two cathodic peaks at 0.3 and 0.45 V, related to
annealed one and the pristine silicon sample. Both spectra are silicon delithiation process, are still observable.
characterized by the presence of a doublet of narrow peaks, Figure 6a shows the cycling performance of the annealed
with the most intense one centered at (99.4 ± 0.2) eV, typical silicon-FLG electrode, obtained in galvanostatic mode with a
of elemental silicon,92 accompanied by a broad peak centered current density of 0.35 A/g. When discharged at 0.01 V, the
at (103.4 ± 0.2) eV, corresponding to SiO2.92 The relative silicon-FLG electrode achieves a specific capacity >3500 mAh/
intensity of the oxide peak increased with the annealing g, almost reaching the theoretical one. The recharge to 2 V
process, passing from 19.6% to 28.3% of the whole silicon shows a CE of almost 89%. In the other cycles, limiting the
content, as obtained by the fitting procedure (Figure S1). potential range between 0.1 and 1 V, the tested electrode
Besides, there is no significant difference between the pristine discharges a capacity of ∼2300 mAh/g with an initial
silicon and the silicon-FLG electrode not subjected to the Coulombic efficiency (ICE) of 95% that reach 99% after 12
annealing process. cycles. The specific capacity achieved at the second cycle
The increase of oxygen content might be due to annealing remains rather stable during cycling (with a fading per cycle of
conditions and by the presence of oxygen atoms in the PAA. 0.13%), delivering 1800 mAh/g after 90 cycles, which
Structural and morphological analyses have demonstrated correspond to ∼78% of the initial value. Figure 6b shows the
that with the combined approach of ultrasonication and potential profiles of some representative cycles. We can
annealing, we obtained a silicon-FLG electrode with a well- observe that the charge−discharge curves exhibit a similar
organized structure. The annealing step leads to the formation shape in every cycle, confirming the reversibility of the process.
of a carbon coating onto the silicon surface, which could be the Furthermore, we can distinguish the typical sloping curve
main cause of the partial oxidation of the silicon (less than associated with the alloying/de-alloying of silicon.93−96 In fact,
10%). Thermogravimetric analysis revealed silicon content of in agreement with the PCGA analysis, discharge curves are
57% (Figure S4). In summary, the as-prepared electrodes characterized by two slopes at 0.20 and 0.1 V, while recharge
consist of a laminated structure in which small silicon processes show an initial plateau at 0.15 V, and two slopes
aggregates are entrapped between FLG flakes. around 0.3 and 0.4 V. Finally, the formation of SEI has been
Electrochemical Characterization. Potentiodynamic cy- evaluated acquiring impedance spectra at different cycles
cling with galvanostatic acceleration (PCGA) measurements (Figure S5). The sequence of Nyquist plots in Figure S5a
were performed to provide insight into the redox processes describes a rather regular evolution of a nonblocking
related to the as-prepared silicon-FLG electrodes. Figure 5 interphase to a blocking one, compatible with a reactive
reports the differential potential profile obtained during the interphase which is progressively passivated from electrolyte
formation cycles (blue line) and the subsequent cycles (red decomposition products. Upon cycling, the resistance of the
line). SEI rises from 13 Ω to 38 Ω after 100 cycles (Figure S5b).
In agreement with previous work,94 the formation cycle For comparison, in Figure S6 is the behavior of the same
shows a completely different feature when compared to the electrode when it is not subjected to the annealing step. This
other cycles. During the discharge, it is possible to observe the electrode shows a lower specific capacity, not being able to
occurrence of multiple processes occurring below 0.13 V, sustain prolonged cycling. In fact, in a few cycles, it is possible
associated with the contemporary intercalation of lithium into to observe a loss of specific capacity of ∼40%. Moreover, after
FLG layers and the complete lithiation of Si. As already less than 50 cycles the material starts to degrade, with
reported,94,95 the lithiation of the crystalline silicon phase irreversible damage after only 100 cycles.
initially form amorphous LixSi alloys, according to Si + xLi+ + In summary, the electrode after annealing demonstrated
xe− → LixSi (0 ≤ x ≤ 4.4) followed by the formation of Li15Si4 enhanced electrochemical properties compared to the
below 0.06 V. In the recharge process, two small peaks, reference electrode. On one hand, the annealing process
1797 DOI: 10.1021/acsaem.8b01927
ACS Appl. Energy Mater. 2019, 2, 1793−1802
ACS Applied Energy Materials Article

discharge−recharge cycles, it is possible to observe an initial


drop of the specific capacity (from 1500 mAh/g to less than
1000 mAh/g), corresponding to a loss of ∼30% of the initial
value. Then, in the subsequent cycles the specific capacity
reaches a stable value of ∼1000 mAh/g. After 250 charge/
discharge cycles the specific capacity reached a final value of
750 mAh/g, with capacity retention of 75%. Discharge/charge
profiles in Figure S7b show the signature of the electro-
chemical lithiation/delithiation of silicon, as evident from the
shapes of the curves, showing behavior similar to the electrode
cycled at lower currents.
Ex-Situ Morphological Analysis. In order to study the
changes occurring in the electrode material upon cycling, an
ex-situ morphological and structural analysis has been
performed by the use of SEM and HRTEM. For the analysis,
the electrodes were first cycled under galvanostatic condition
in lithium cell using a current density of 0.35A/g. Afterward,
the tests were stopped at different cycle numbers (i.e., 1, 10,
50, and 100) and the electrode recovered and washed with
anhydrous DMC to eliminate the residual lithium salt on the
electrode surface. The micrographs obtained for the electrode
after 100 cycles are reported in Figure 7.

Figure 6. Electrochemical performance of the annealed silicon-FLG


electrode in lithium cell: (a) cycling performance; (b) potential
profiles obtained at 0.35 A/g; and (c) rate capability measured at
different current density values. For all measurements, before cycling,
a formation cycle (indicated as cycle 0) has been performed between
0.01 and 2 V at 0.35 A/g.

determines the creation of a well-organized structure. The


graphene flakes entrapping silicon particles can buffer the
volumetric expansion/contraction of silicon during cycling, Figure 7. (a) HRTEM image and (inset) corresponding FFT, (b)
beneficial for the electrochemical performance. On the other zero-loss filtered BF-TEM image and corresponding elemental maps
hand, annealing also causes the pyrolysis of the non-conductive of (c) C and (d) O for Si nanoparticles from the post-mortem
PAA, creating a carbonaceous network that improve the electrode (after 100 cycles), suspended on holes in the carbon
conductivity and physically sustain the silicon-graphene support film.
structure.
To evaluate the rate capability, the annealed silicon-FLG After the first forming cycle, the composite does not show
electrode was further cycled increasing the current density any appreciable changes in the structure nor in the
from 0.35 A/g (corresponding to C/10) to 7 A/g (∼2 C). As morphology. The latter is similar to that of the post-annealing
reported in Figure 6c, the annealed silicon-FLG electrode electrode (Figure 1a). The graphene flakes and the silicon
exhibits excellent rate performance. However, we noticed that nanoparticles are clearly visible (Figure S8a,e). According to
the exchanged capacity decreases from 2300 mAh/g at 0.35A/ EFTEM mapping (Figure S9a), silicon nanoparticles are still
g to 750 mAh/g at 7 A/g (2 C). In addition, when the current covered by a carbon coating (Figure S9b) of ∼3 nm thick
density returns to the initial value of 0.35 A/g, the specific together with an oxide-reach shell (Figure S9c). HRTEM
capacity raises again to ∼2000 mAh/g, which is the same value images analysis (Figure S10a,e) shows that silicon nano-
obtained during first cycles. particles are still polycrystalline, as already observed just after
Since the annealed silicon-FLG electrode demonstrated the annealing process.
cyclability at high current densities, it was decided to perform a Since the tenth cycle, silicon undergoes a major structural
prolonged discharge/recharge test of the anode at 3.5 A/g. The change, turning from polycrystalline to amorphous, as pointed
obtained results are shown in Figure S7a. During the first 50 out by the HRTEM analysis (Figure S10b,f). This can explain
1798 DOI: 10.1021/acsaem.8b01927
ACS Appl. Energy Mater. 2019, 2, 1793−1802
ACS Applied Energy Materials Article

the formation of a small valley in the cycling performance of atmosphere. The latter process allows the pyrolysis of
Figure 6a. This amorphous structure is preserved also for poliacrylic acid, which increases the electrical conductivity of
higher cycles (Figure 7a). the silicon nanoparticles and the bond with the graphene
From SEM imaging (Figure S8 in SI), a gradual evolution of flakes.
the Si nanoparticles into a flat, continuous layer is observed. The as-prepared nanostructured electrodes have shown
After 50 cycles, SEM images (Figure S8c,g) show how the excellent electrochemical performances, achieving high specific
cluster of silicon has almost covered the entire electrode capacity in lithium cell (>2000 mAh/g), able to sustain high
interface. It is still possible to point the most superficial currents (over 750 mAh/g at 7 A/g) maintaining high
individual graphene flakes. Moreover, the morphology of the reversibility (Coulombic efficiency −CE− ≈ 99%). These
silicon nanoparticles becomes more and more porous with the results are in line with those already reported in the literature
cycling and can no longer be neglected after 50 cycles (Figure (Table S1).
S9d,e). In the same way, the oxidation of the silicon Finally, we demonstrated the use of silicon-FLG electrode in
nanoparticles seems to increase with the cycling, with the a complete lithium-ion battery. Our full cell delivered a first
first appreciable evidence only after 50 cycles (Figure S9i). cycle specific capacity of 1800 mAh/g and a good stability (CE
For higher cycling, the aggregation and deaggregation of 99%).
silicon nanoparticles leads to a new structure, as shown in the We believe that such laminated silicon-FLG graphene
top view by Figure S8d. The silicon nanoparticles now are nanocomposite represents a potential hybrid morphology for
aggregated in individual cluster, and it is again possible to high silicon loading Li-ion cells displaying high-rate capabilities
observe graphene flakes on the surface. and excellent stability. The developed nanostructured silicon-
By contrast, the graphene flakes are not significantly affected FLG anode paved the way for further optimization of the
by the cycling (Figure S10d,h), an expected result which electrode by, for example, exploring different morphologies of
confirms that the silicon nanoparticles are the active material. silicon nanoparticles in combination with FLG.
Development of Lithium-Ion Battery. After having
demonstrated the electrochemical activity of the silicon-
graphene electrode in a lithium cell, this electrode was tested
■ ASSOCIATED CONTENT
* Supporting Information
S
in a complete lithium-ion cell. The Supporting Information is available free of charge on the
To this end, a commercial NMC111 with a nominal capacity ACS Publications website at DOI: 10.1021/acsaem.8b01927.
of 2 mAh/cm2 was chosen as the positive electrode, while as Additional XPS spectra and galvanostatic cycling
electrolyte the same mixture of LP30 and 10% v/v of FEC used measurements as well as EDS maps, TGA curves and
in half-cell experiments, was adopted. Anode and cathode tables (PDF)


masses were balanced considering the anode reversible
capacity (2300 mAh/g) and the reversible capacity of the
AUTHOR INFORMATION
cathode (137 mAh/g as reported in Figure S8).
After the formation procedure (Figure S9), the cell achieved Corresponding Authors
a specific capacity of ∼1800 mAh/g (a value referred to the *E-mail: stefano.palumbo@iit.it.
silicon mass) with a CE > 99% (see Figure 8). After 20 cycles, *E-mail: francesco.bonaccorso@iit.it.
ORCID
Alberto Ansaldo: 0000-0002-3493-4157
Francesco Bonaccorso: 0000-0001-7238-9420
Funding
We acknowledge the European Union’s Horizon 2020 research
and innovation program under grant agreement no. 785219,
GrapheneCore2.
Notes
The authors declare no competing financial interest.
Figure 8. (a) Specific capacity vs cycle number plot and (b) cell
voltage profiles in a silicon-FLG/NMC111 cell. ■ ACKNOWLEDGMENTS
We thank Sergio Marras and Mirko Prato from the Materials
Characterization Facility at the Fondazione Istituto Italiano di
the cell’s specific capacity is still higher than 1600 mAh/g.
Tecnologia for help with X-ray Diffraction measurements and
Unfortunately, the specific capacity decreases over prolonged
XPS analysis. We thank Simone Lauciello for SEM imaging,
cycling, and after 200 cycles only 33% of the initial capacity is
Luca Gagliani and Manuel Crugliano for their contribution to
retained (Figure S9).


the production of few-layer graphene flakes, Giammarino
CONCLUSION Pugliese for thermal analysis, and Christoph Stangl of Varta
Micro Innovation gmbH for useful discussion.


We designed and developed a nanostructured silicon-few-layer
graphene (FLG) electrode to be exploited as anode for lithium REFERENCES
ion battery. The nanostructured silicon-FLG electrode (1) Armand, M.; Tarascon, J.-M. Building Better Batteries. Nature
comprises horizontally oriented FLG flakes entrapping the 2008, 451, 652−657.
silicon nanoparticles. The realization of such nanostructured (2) Scrosati, B.; Garche, J. Lithium Batteries: Status, Prospects and
anode materials is rather simple, being directly fabricated onto Future. J. Power Sources 2010, 195, 2419−2430.
the copper current collector, exploiting the use of ultra- (3) Yoshino, A., Sanechika, K., Nakajima, T. Secondary Battery;
sonication followed by an annealing process in reducing Japanese Patent 1,989,293, 1985.

1799 DOI: 10.1021/acsaem.8b01927


ACS Appl. Energy Mater. 2019, 2, 1793−1802
ACS Applied Energy Materials Article

(4) Tarascon, J. M.; Armand, M. Issues and Challenges Facing Evaluated by in-Operando Dilatometry and Acoustic Emission. J.
Rechargeable Lithium Batteries. Nature 2001, 414, 359−367. Power Sources 2016, 330, 253−260.
(5) Foundations for the Future. Nat. Energy 2016, 1, 16147. (29) Yang, J. Small Particle Size Multiphase Li-Alloy Anodes for
(6) Evarts, E. C. Lithium Batteries: To the Limits of Lithium. Nature Lithium-Ionbatteries. Solid State Ionics 1996, 90, 281−287.
2015, 526, S93−S95. (30) Kwon, Y.; Kim, H.; Doo, S. G.; Cho, J. Sn0.9Si0.1/Carbon Core-
(7) Brandt, K. Historical development of secondary lithium batteries. Shell Nanoparticles for High-Density Lithium Storage Materials.
Solid State Ionics 1994, 69, 173−183. Chem. Mater. 2007, 19, 982−986.
(8) Janek, J.; Zeier, W. G. A Solid Future for Battery Development. (31) Kwon, Y.; Cho, J. High Capacity Carbon-Coated Si70Sn30
Nat. Ener. 2016, 500, 16141. Nanoalloys for Lithium Battery Anode Material. Chem. Commun.
(9) Goodenough, J. B.; Kim, Y. Challenges for Rechargeable Li 2008, 0, 1109−1111.
Batteries. Chem. Mater. 2010, 22, 587−603. (32) Peng, K.; Jie, J.; Zhang, W.; Lee, S. T. Silicon Nanowires for
(10) Etacheri, V.; Marom, R.; Elazari, R.; Salitra, G.; Aurbach, D. Rechargeable Lithium-Ion Battery Anodes. Appl. Phys. Lett. 2008, 93,
Challenges in the Development of Advanced Li-Ion Batteries: A 033105.
Review. Energy Environ. Sci. 2011, 4, 3243−3262. (33) Wu, H.; Chan, G.; Choi, J. W.; Ryu, I.; Yao, Y.; Mcdowell, M.
(11) Bruce, P. G.; Scrosati, B.; Tarascon, J. M. Nanomaterials for T.; Lee, S. W.; Jackson, A.; Yang, Y.; Hu, L.; Cui, Y. Stable Cycling of
Rechargeable Lithium Batteries. Angew. Chem., Int. Ed. 2008, 47, Double-Walled Silicon Nanotube Battery Anodes through Solid-
2930−2946. Electrolyte Interphase Control. Nat. Nanotechnol. 2012, 7, 310−315.
(12) Aricò, A. S.; Bruce, P.; Scrosati, B.; Tarascon, J. M.; Van (34) Zhang, Y.; Zhang, X. G.; Zhang, H. L.; Zhao, Z. G.; Li, F.; Liu,
Schalkwijk, W. Nanostructured Materials for Advanced Energy C.; Cheng, H. M. Composite Anode Material of Silicon/Graphite/
Conversion and Storage Devices. Nat. Mater. 2005, 4, 366−377. Carbon Nanotubes for Li-Ion Batteries. Electrochim. Acta 2006, 51,
(13) Obrovac, M. N.; Chevrier, V. L. Alloy Negative Electrodes for 4994−5000.
Li-Ion Batteries. Chem. Rev. 2014, 114, 11444−11502. (35) Dimov, N.; Kugino, S.; Yoshio, M. Carbon-Coated Silicon as
(14) Poizot, P.; Laruelle, S.; Grugeon, S.; Dupont, L.; Tarascon, J. Anode Material for Lithium Ion Batteries: Advantages and
M. Nano-Sized Transition-Metal Oxides as Negative-Electrode Limitations. Electrochim. Acta 2003, 48, 1579−1587.
Materials for Lithium-Ion Batteries. Nature 2000, 407, 496−499. (36) Lee, H. Y.; Lee, S. M. Carbon-Coated Nano-Si Dispersed
(15) Choi, J. W.; Aurbach, D. Promise and Reality of Post-Lithium- Oxides/Graphite Composites as Anode Material for Lithium Ion
Ion Batteries with High Energy Densities. Nat. Rev. Mater. 2016, 1, Batteries. Electrochem. Commun. 2004, 6, 465−469.
16013. (37) Lee, J. K.; Oh, C.; Kim, N.; Hwang, J.-Y.; Sun, Y.-K. Rational
(16) Li, W.; Sun, X.; Yu, Y. Si-, Ge-, Sn-Based Anode Materials for Design of Silicon-Based Composites for High-Energy Storage Devices.
Lithium-Ion Batteries: From Structure Design to Electrochemical J. Mater. Chem. A 2016, 4, 5366−5384.
Performance. Small Methods 2017, 1, 1600037. (38) Hertzberg, B.; Alexeev, A.; Yushin, G. Deformations in Si-Li
(17) Casimir, A.; Zhang, H.; Ogoke, O.; Amine, J. C.; Lu, J.; Wu, G. Anodes upon Electrochemical Alloying in Nano-Confined Space. J.
Silicon-Based Anodes for Lithium-Ion Batteries: Effectiveness of Am. Chem. Soc. 2010, 132, 8548−8549.
Materials Synthesis and Electrode Preparation. Nano Energy 2016, 27, (39) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.;
359−376. Zhang, Y.; Dubonos, S. V.; Grigorieva, I. V.; Firsov, A. A. Electric
(18) Liang, B.; Liu, Y.; Xu, Y. Silicon-Based Materials as High Field in Atomically Thin Carbon Films. Science 2004, 306, 666−669.
Capacity Anodes for next Generation Lithium Ion Batteries. J. Power (40) Quesnel, E.; Roux, F.; Emieux, F.; Faucherand, P.; Kymakis, E.;
Sources 2014, 267, 469−490. Volonakis, G.; Giustino, F.; Martín-García, B.; Moreels, I.; Gürsel, S.
(19) Winter, M.; Besenhard, J. O. Electrochemical Lithiation of Tin A.; Yurtcan, A. B.; Di Noto, V.; Talyzin, A.; Baburin, I.; Tranca, D.;
and Tin-Based Intermetallics and Composites. Electrochim. Acta 1999, Seifert, G.; Crema, L.; Speranza, G.; Tozzini, V.; Bondavalli, P.;
45, 31−50. Pognon, G.; Botas, C.; Carriazo, D.; Singh, G.; Rojo, T.; Kim, G.; Yu,
(20) Dey, A. N. Electrochemical Alloying of Lithium in Organic W.; Grey, C. P.; Pellegrini, V. Graphene-Based Technologies for
Electrolytes. J. Electrochem. Soc. 1971, 118, 1547. Energy Applications, Challenges and Perspectives. 2D Mater. 2015, 2,
(21) Winter, M.; Besenhard, J. O.; Spahr, M. E.; Novák, P. Insertion 030204.
Electrode Materials for Rechargeable Lithium Batteries. Adv. Mater. (41) Bonaccorso, F.; Colombo, L.; Yu, G.; Stoller, M.; Tozzini, V.;
1998, 10, 725−763. Ferrari, A. C.; Ruoff, R. S.; Pellegrini, V. Graphene, Related Two-
(22) McDowell, M. T.; Lee, S. W.; Nix, W. D.; Cui, Y. 25th Dimensional Crystals, and Hybrid Systems for Energy Conversion
Anniversary Article: Understanding the Lithiation of Silicon and and Storage. Science 2015, 347, 1246501.
Other Alloying Anodes for Lithium-Ion Batteries. Adv. Mater. 2013, (42) Papageorgiou, D. G.; Kinloch, I. A.; Young, R. J. Mechanical
25, 4966−4985. Properties of Graphene and Graphene-Based Nanocomposites. Prog.
(23) Boukamp, B. A. All-Solid Lithium Electrodes with Mixed- Mater. Sci. 2017, 90, 75−127.
Conductor Matrix. J. Electrochem. Soc. 1981, 128, 725. (43) Hussain, S.; Iqbal, M. W.; Park, J.; Ahmad, M.; Singh, J.; Eom,
(24) Chan, C. K.; Peng, H.; Liu, G.; McIlwrath, K.; Zhang, X. F.; J.; Jung, J. Physical and Electrical Properties of Graphene Grown
Huggins, R. A.; Cui, Y. High-Performance Lithium Battery Anodes under Different Hydrogen Flow in Low Pressure Chemical Vapor
Using Silicon Nanowires. Nat. Nanotechnol. 2008, 3, 31−35. Deposition. Nanoscale Res. Lett. 2014, 9, 546.
(25) Liu, X. H.; Zhong, L.; Huang, S.; Mao, S. X.; Zhu, T.; Huang, J. (44) Zhang, M.; Zhang, T.; Ma, Y.; Chen, Y. Latest Development of
Y. Size-Dependent Fracture of Silicon Nanoparticles during Nanostructured Si/C Materials for Lithium Anode Studies and
Lithiation. ACS Nano 2012, 6, 1522−1531. Applications. Energy Storage Mater. 2016, 4, 1−14.
(26) Ding, X.; Wang, H.; Liu, X.; Gao, Z.; Huang, Y.; Lv, D.; He, P.; (45) Wehling, T. O.; Novoselov, K. S.; Morozov, S. V.; Vdovin, E.
Huang, Y. Advanced Anodes Composed of Graphene Encapsulated E.; Katsnelson, M. I.; Geim, A. K.; Lichtenstein, A. I. Molecular
Nano-Silicon in a Carbon Nanotube Network. RSC Adv. 2017, 7, Doping of Graphene. Nano Lett. 2008, 8, 173−177.
15694−15701. (46) Boukhvalov, D. W.; Katsnelson, M. I. Chemical Functionaliza-
(27) Schott, T.; Robert, R.; Pacheco Benito, S.; Ulmann, P. A.; Lanz, tion of Graphene with Defects. Nano Lett. 2008, 8, 4374−4379.
P.; Zürcher, S.; Spahr, M. E.; Novák, P.; Trabesinger, S. Cycling (47) Banhart, F.; Kotakoski, J.; Krasheninnikov, A. V. Structural
Behavior of Silicon-Containing Graphite Electrodes, Part B: Effect of Defects in Graphene. ACS Nano 2011, 5, 26−41.
the Silicon Source. J. Phys. Chem. C 2017, 121, 25718−25728. (48) Araujo, P. T.; Terrones, M.; Dresselhaus, M. S. Defects and
(28) Tranchot, A.; Idrissi, H.; Thivel, P. X.; Roué, L. Influence of the Impurities in Graphene-like Materials. Mater. Today 2012, 15, 98−
Si Particle Size on the Mechanical Stability of Si-Based Electrodes 109.

1800 DOI: 10.1021/acsaem.8b01927


ACS Appl. Energy Mater. 2019, 2, 1793−1802
ACS Applied Energy Materials Article

(49) Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J. W.; Potts, J. R.; (71) Wu, Y. P.; Jiang, C. Y.; Wan, C. R.; Fang, S. B.; Jiang, Y. Y.
Ruoff, R. S. Graphene and Graphene Oxide: Synthesis, Properties, Nitrogen-Containing Polymeric Carbon as Anode Material for
and Applications. Adv. Mater. 2010, 22, 3906−3924. Lithium Ion Secondary Battery. J. Appl. Polym. Sci. 2000, 77, 1735−
(50) Allen, M. J.; Tung, V. C.; Kaner, R. B. Honeycomb Carbon: A 1741.
Review of Graphene. Chem. Rev. 2010, 110, 132−145. (72) Reddy, A. L. M.; Srivastava, A.; Gowda, S. R.; Gullapalli, H.;
(51) Raccichini, R.; Varzi, A.; Passerini, S.; Scrosati, B. The Role of Dubey, M.; Ajayan, P. M. Synthesis of Nitrogen-Doped Graphene
Graphene for Electrochemical Energy Storage. Nat. Mater. 2015, 14, Films for Lithium Battery Application. ACS Nano 2010, 4, 6337−
271−279. 6342.
(52) Cai, X.; Lai, L.; Shen, Z.; Lin, J. Graphene and Graphene-Based (73) Wei, D.; Liu, Y.; Wang, Y.; Zhang, H.; Huang, L.; Yu, G.
Composites as Li-Ion Battery Electrode Materials and Their Synthesis of N-Doped Graphene by Chemical Vapor Deposition and
Application in Full Cells. J. Mater. Chem. A 2017, 5, 15423−15446. Its Electrical Properties. Nano Lett. 2009, 9, 1752−1758.
(53) Levchenko, I.; Ostrikov, K.; Zheng, J.; Li, X.; Keidar, M.; Teo, (74) Wang, H.; Zhang, C.; Liu, Z.; Wang, L.; Han, P.; Xu, H.;
K. B. K. Scalable Graphene Production: Perspectives and Challenges Zhang, K.; Dong, S.; Yao, J.; Cui, G. Nitrogen-Doped Graphene
of Plasma Applications. Nanoscale 2016, 8, 10511−10527. Nanosheets with Excellent Lithium Storage Properties. J. Mater.
(54) Diamond, W. Do. Mass production of high quality graphene: An Chem. 2011, 21, 5430−5434.
analysis of worldwide patents, 2012; https://www.nanowerk.com/ (75) Lin, J.; He, J.; Chen, Y.; Li, Q.; Yu, B.; Xu, C.; Zhang, W.
spotlight/spotid=25744.php. Pomegranate-Like Silicon/Nitrogen-Doped Graphene Microspheres
(55) Bonaccorso, F.; Lombardo, A.; Hasan, T.; Sun, Z.; Colombo, as Superior-Capacity Anode for Lithium-Ion Batteries. Electrochim.
L.; Ferrari, A. C. Production and processing of graphene and 2d Acta 2016, 215, 667−673.
crystals. Mater. Today 2012, 15, 564−589. (76) Tang, X.; Wen, G.; Zhang, Y.; Wang, D.; Song, Y. Novel Silicon
(56) Chen, K.; Song, S.; Liu, F.; Xue, D. Structural Design of Nanoparticles with Nitrogen-Doped Carbon Shell Dispersed in
Graphene for Use in Electrochemical Energy Storage Devices. Chem. Nitrogen-Doped Graphene and CNTs Hybrid Electrode for Lithium
Soc. Rev. 2015, 44, 6230−6257. Ion Battery. Appl. Surf. Sci. 2017, 425, 742−749.
(57) Ji, L.; Meduri, P.; Agubra, V.; Xiao, X.; Alcoutlabi, M. (77) Tang, X.; Wen, G.; Song, Y. Stable Silicon/3D Porous N-
Graphene-Based Nanocomposites for Energy Storage. Adv. Energy Doped Graphene Composite for Lithium-Ion Battery Anodes with
Mater. 2016, 6, 1502159. Self-Assembly. Appl. Surf. Sci. 2018, 436, 398−404.
(58) Xu, C.; Xu, B.; Gu, Y.; Xiong, Z.; Sun, J.; Zhao, X. S. Graphene- (78) Li, N.; Jin, S.; Liao, Q.; Cui, H.; Wang, C. X. Encapsulated
Based Electrodes for Electrochemical Energy Storage. Energy Environ. within Graphene Shell Silicon Nanoparticles Anchored on Vertically
Sci. 2013, 6 (5), 1388. Aligned Graphene Trees as Lithium Ion Battery Anodes. Nano Energy
(59) Lv, W.; Li, Z.; Deng, Y.; Yang, Q. H.; Kang, F. Graphene-Based 2014, 5, 105−115.
Materials for Electrochemical Energy Storage Devices: Opportunities (79) Li, F.; Yue, H.; Yang, Z.; Li, X.; Qin, Y.; He, D. Flexible Free-
Standing Graphene Foam Supported Silicon Films as High Capacity
and Challenges. Energy Storage Mater. 2016, 2, 107−138.
Anodes for Lithium Ion Batteries. Mater. Lett. 2014, 128, 132−135.
(60) Yang, Y.; Han, C.; Jiang, B.; Iocozzia, J.; He, C.; Shi, D.; Jiang,
(80) Loveridge, M. J.; Lain, M. J.; Huang, Q.; Wan, C.; Roberts, A.
T.; Lin, Z. Graphene-Based Materials with Tailored Nanostructures
J.; Pappas, G. S.; Bhagat, R. Enhancing Cycling Durability of Li-Ion
for Energy Conversion and Storage. Mater. Sci. Eng., R 2016, 102, 1−
Batteries with Hierarchical Structured Silicon-Graphene Hybrid
72.
Anodes. Phys. Chem. Chem. Phys. 2016, 18, 30677−30685.
(61) Wu, S.; Xu, R.; Lu, M.; Ge, R.; Iocozzia, J.; Han, C.; Jiang, B.;
(81) Huang, Q.; Loveridge, M. J.; Genieser, R.; Lain, M. J.; Bhagat,
Lin, Z. Graphene-Containing Nanomaterials for Lithium-Ion R. Electrochemical Evaluation and Phase-Related Impedance Studies
Batteries. Adv. Energy Mater. 2015, 5, 1−40. on Silicon-Few Layer Graphene (FLG) Composite Electrode
(62) Bonaccorso, F.; Bartolotta, A.; Coleman, J. N.; Backes, C. 2D- Systems. Sci. Rep. 2018, 8, 1386.
Crystal-Based Functional Inks. Adv. Mater. 2016, 28, 6136−6166. (82) Chou, S. L.; Wang, J. Z.; Choucair, M.; Liu, H. K.; Stride, J. A.;
(63) Chen, D.; Feng, H.; Li, J. Graphene Oxide: Preparation, Dou, S. X. Enhanced Reversible Lithium Storage in a Nanosize
Functionalization, and Electrochemical Applications. Chem. Rev. Silicon/Graphene Composite. Electrochem. Commun. 2010, 12, 303−
2012, 112, 6027−6053. 306.
(64) Gao, W. The Chemistry of Graphene Oxide. In Graphene (83) Son, I. H.; Park, J. H.; Park, S.; Park, K.; Han, S.; Shin, J.; Doo,
Oxide: Reduction Recipes, Spectroscopy, and Applications; Springer S. G.; Hwang, Y.; Chang, H.; Choi, J. W. Graphene Balls for Lithium
International Publishing: Cham, 2015; pp 61−95. Rechargeable Batteries with Fast Charging and High Volumetric
(65) Dreyer, D. R.; Todd, A. D.; Bielawski, C. W. Harnessing. The Energy Densities. Nat. Commun. 2017, 8, 1−10.
Chemistry of Graphene Oxide. Chem. Soc. Rev. 2014, 43, 5288−5301. (84) Son, I. H.; Park, J. H.; Kwon, S.; Park, S.; Rümmeli, M. H.;
(66) Suresh, S.; Wu, Z. P.; Bartolucci, S. F.; Basu, S.; Mukherjee, R.; Bachmatiuk, A.; Song, H. J.; Ku, J.; Choi, J. W.; Choi, J. M.; Doo, S.
Gupta, T.; Hundekar, P.; Shi, Y.; Lu, T. M.; Koratkar, N. Protecting G.; Chang, H. Silicon Carbide-Free Graphene Growth on Silicon for
Silicon Film Anodes in Lithium-Ion Batteries Using an Atomically Lithium-Ion Battery with High Volumetric Energy Density. Nat.
Thin Graphene Drape. ACS Nano 2017, 11, 5051−5061. Commun. 2015, 6, 7393.
(67) Ding, X.; Liu, X. X.; Huang, Y.; Zhang, X.; Zhao, Q.; Xiang, X.; (85) Chen, S.; Bao, P.; Huang, X.; Sun, B.; Wang, G. Hierarchical
Li, G.; He, P.; Wen, Z.; Li, J.; Huang, Y. Enhanced Electrochemical 3D Mesoporous Silicon@graphene Nanoarchitectures for Lithium
Performance Promoted by Monolayer Graphene and Void Space in Ion Batteries with Superior Performance. Nano Res. 2014, 7, 85−94.
Silicon Composite Anode Materials. Nano Energy 2016, 27, 647−657. (86) Ferrari, A. C.; Bonaccorso, F.; Falko, V.; Novoselov, K. S.;
(68) Ji, D.; Yang, Z.; Xiong, L.; Luo, H.; Xiong, G.; Zhu, Y.; Wan, Y. Roche, S.; Bøggild, P.; Borini, S.; Koppens, F. H. L.; Palermo, V.;
Effect of Si Content on Structure and Electrochemical Performance of Pugno, N.; Garrido, J. A.; Sordan, R.; Bianco, A.; Ballerini, L.; Prato,
Ternary Nanohybrids Integrating Si Nanoparticles, N-Doped Carbon M.; Lidorikis, E.; Kivioja, J.; Marinelli, C.; Ryhänen, T.; Morpurgo, A.;
Shell, and Nitrogen-Doped Graphene. RSC Adv. 2017, 7, 4209−4215. Coleman, J. N.; Nicolosi, V.; Colombo, L.; Fert, A.; Garcia-
(69) Cho, Y. J.; Kim, H. S.; Im, H.; Myung, Y.; Jung, G. B.; Lee, C. Hernandez, M.; Bachtold, A.; Schneider, G. F.; Guinea, F.; Dekker,
W.; Park, J.; Park, M.-H.; Cho, J.; Kang, H. S. Nitrogen-Doped C.; Barbone, M.; Sun, Z.; Galiotis, C.; Grigorenko, A. N.;
Graphitic Layers Deposited on Silicon Nanowires for Efficient Konstantatos, G.; Kis, A.; Katsnelson, M.; Vandersypen, L.; Loiseau,
Lithium-Ion Battery Anodes. J. Phys. Chem. C 2011, 115, 9451−9457. A.; Morandi, V.; Neumaier, D.; Treossi, E.; Pellegrini, V.; Polini, M.;
(70) Wu, Y.; Fang, S.; Jiang, Y. Effects of Nitrogen on the Carbon Tredicucci, A.; Williams, G. M.; Hong, B. H.; Ahn, J.-H.; Kim, J. M.;
Anode of a Lithium Secondary Battery. Solid State Ionics 1999, 120, Zirath, H.; van Wees, B. J.; van der Zant, H.; Occhipinti, L.; Di
117−123. Matteo, A.; Kinloch, I. A.; Seyller, T.; Quesnel, E.; Feng, X.; Teo, K.;

1801 DOI: 10.1021/acsaem.8b01927


ACS Appl. Energy Mater. 2019, 2, 1793−1802
ACS Applied Energy Materials Article

Rupesinghe, N.; Hakonen, P.; Neil, S. R. T.; Tannock, Q.; Löfwander,


T.; Kinaret, L. Science and technology roadmap for graphene, related
two-dimensional crystals, and hybrid systems. Nanoscale 2015, 7,
4598−4810.
(87) Zurutuza, A.; Marinelli, C. Challenges and Opportunities in
Graphene Commercialization. Nat. Nanotechnol. 2014, 9, 730−734.
(88) Ciriminna, R.; Zhang, N.; Yang, M. Q.; Meneguzzo, F.; Xu, Y.
J.; Pagliaro, M. Commercialization of Graphene-Based Technologies:
A Critical Insight. Chem. Commun. 2015, 51, 7090−7095.
(89) Del Rio Castillo, A. E.; Pellegrini, V.; Ansaldo, A.; Ricciardella,
F.; Sun, H.; Marasco, L.; Buha, J.; Dang, Z.; Gagliani, L.; Lago, E.;
Curreli, N.; Gentiluomo, S.; Palazon, F.; Prato, M.; Oropesa-Nuñez,
R.; Toth, P. S.; Mantero, E.; Crugliano, M.; Gamucci, A.; Tomadin,
A.; Polini, M.; Bonaccorso, F. High-Yield Production of 2D Crystals
by Wet-Jet Milling. Mater. Horiz. 2018, 5, 890−904.
(90) Greco, E.; Nava, G.; Fathi, R.; Fumagalli, F.; Del Rio-Castillo,
A. E.; Ansaldo, A.; Monaco, S.; Bonaccorso, F.; Pellegrini, V.; Di
Fonzo, F. Few-Layer Graphene Improves Silicon Performance in Li-
Ion Battery Anodes. J. Mater. Chem. A 2017, 5, 19306−19315.
(91) Degen, T., Sadki, M., Bron, E., König, U., Nénert, G. The High
Score Suite. In Powder Diffraction; Cambridge University Press, 2014;
Vol 29, pp S13−S18.
(92) NIST X-ray Photoelectron Spectroscopy Database v 4.1; National
Institute of Standards and Technology, Gaithersburg, 2012; http://
srdata.nist.gov/xps/.
(93) Ferrari, A. C.; Basko, D. M. Raman spectroscopy as a versatile
tool for studying the properties of graphene. Nat. Nanotechnol. 2013,
8, 235.
(94) Li, J.; Dahn, J. R. An in situ X-ray diffraction study of the
reaction of Li with crystalline Si. J. Electrochem. Soc. 2007, 154, A156−
A161.
(95) Hatchard, T. D.; Dahn, J. R. In situ XRD and electrochemical
study of the reaction of lithium with amorphous silicon. J. Electrochem.
Soc. 2004, 151 (6), A838−A842.
(96) Obrovac, M. N.; Krause, L. J. Reversible Cycling of Crystalline
Silicon Powder. J. Electrochem. Soc. 2007, 154, A103.

1802 DOI: 10.1021/acsaem.8b01927


ACS Appl. Energy Mater. 2019, 2, 1793−1802

You might also like