You are on page 1of 9

Journal of Power Sources 405 (2018) 61–69

Contents lists available at ScienceDirect

Journal of Power Sources


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

Clay minerals derived nanostructured silicon with various morphology: T


Controlled synthesis, structural evolution, and enhanced lithium storage
properties
Qingze Chena,b, Shaohong Liuc, Runliang Zhua,∗, Dingcai Wuc, Haoyang Fua,b, Jianxi Zhua,
Hongping Hea,b
a
CAS Key Laboratory of Mineralogy and Metallogeny, Guangdong Provincial Key Laboratory of Mineral Physics and Materials, Guangzhou Institute of Geochemistry,
Chinese Academy of Sciences, Guangzhou, 510640, China
b
University of Chinese Academy of Sciences, Beijing, 100049, China
c
Materials Science Institute, PCFM Lab and GDHPRC Lab, School of Chemistry, Sun Yat-sen University, Guangzhou, China

H I GH L IG H T S G R A P H I C A L A B S T R A C T

• Three Si nanostructures with various


morphology were successfully synthe-
sized.
• The morphology of nanostructured Si
greatly depended on nature of clay
minerals.
• 2D silicon from montmorillonite ex-
hibited the best electrochemical per-
formance.
• The structure-property relationship of
various Si nanostructures was estab-
lished.

A R T I C LE I N FO A B S T R A C T

Keywords: Nanostructuring is an effective strategy to enhance the structural and cycling stability of silicon anodes in li-
Nanostructured silicon thium-ion batteries. However, a controllable and cost-effective method for synthesizing nanostructured silicon
Clay minerals with various morphology is still a challenge. Herein, we synthesize zero-dimensional, two-dimensional, and
Controlled synthesis three-dimensional silicon nanostructures directly using low-cost and abundant clay minerals as precursors
Structural evolution
without any pretreatment and templates. Our results show that the morphology and microstructure of the re-
Lithium-ion battery anodes
sulting nanostructured silicon strongly depend on the architectural features of clay minerals, i.e., zero-dimen-
sional silicon from palygorskite, two-dimensional silicon from montmorillonite, and three-dimensional silicon
from halloysite. The silicon nanostructures show large specific surface area (over 80 m2 g−1) and hierarchical
pore structure. As anodes in lithium-ion batteries, two-dimensional nanostructured silicon from montmorillonite
exhibits the best electrochemical performance (i.e., 1369 mAh g−1 at 1.0 A g−1 with a capacity retention of 78%
over 200 cycles). This work provides a universal guideline from clay minerals to various silicon nanostructures
via an economical and scalable strategy, and reveals the fundamental structure-property relationship of different
silicon nanostructures synthesized under the same condition, which would contribute to the large-scale pro-
duction of high-performance and low-cost silicon-based anodes in lithium-ion batteries.


Corresponding author.
E-mail address: zhurl@gig.ac.cn (R. Zhu).

https://doi.org/10.1016/j.jpowsour.2018.10.031
Received 25 July 2018; Received in revised form 18 September 2018; Accepted 9 October 2018
Available online 13 October 2018
0378-7753/ © 2018 Elsevier B.V. All rights reserved.
Q. Chen et al. Journal of Power Sources 405 (2018) 61–69

1. Introduction natural clay minerals with different structure, nanostructured Si with


various morphology, and the corresponding electrochemical properties.
Silicon is regarded as an attractive candidate for next-generation Si nanostructures with different morphology were synthesized directly
anode of lithium-ion batteries (LIBs) owing to its exceptional theore- using three typical clay minerals (i.e., chain-layered palygorskite (Pal),
tical capacity (∼4200 mAh g−1 with a Li22Si4 phase) and low discharge layered montmorillonite (Mt), and tubular halloysite (Hal)) without
potential (∼0.2 V vs. Li+/Li) compared with conventional graphite pretreatment. The method involved in a modified magnesiothermic
anode materials [1–3]. However, the Si anode undergoes serious ca- reduction, in which a molten salt (i.e., NaCl) was employed as a heat
pacity fading because of its huge volume change during the lithiation/ scavenger and a dispersing agent simultaneously. The molten salt might
delithiation process, which then causes the loss of electrical con- provide a controllable high-temperature liquid environment for the
ductivity and the increase of variable solid electrolyte interphase (SEI) synthesis of nanoparticles by adsorbing the massive heat generated
area [4–8]. Nanostructuring has been proven to be an effective strategy from the exothermic reaction [33,34], consequently preventing the
for enhancing the structural and cycling stability of Si anodes [9,10]. formation of the unwanted byproducts. Our results indicated that the
Various Si nanostructures, including zero-dimensional (0D) nano- microstructure and morphology of as-prepared nanostructured Si
particles [11,12], one-dimensional (1D) nanotubes and nanowires highly depended on the microstructures of clay minerals. When used as
[13,14], two-dimensional (2D) nanosheets [15,16], and three-dimen- anodes in LIBs, the 2D nanostructured Si from Mt exhibited the best
sional (3D) porous nanostructures [17–19], have been developed as the electrochemical performance in terms of high specific capacity, good
anodes of LIBs and exhibited good electrochemical performance. rate capability, and stable cycling performance, which could be at-
Despite these remarkable advances, the fabrication of nanos- tributed to the limited variable SEI area and efficient transport property
tructured Si with specific morphology always depends greatly on ex- of 2D Si nanostructure.
pensive silicon precursors, complicated procedures, and delicate
equipment, which severely hinder the practical application of Si an- 2. Experimental section
odes. For example, traditional methods, such as electroless or electro-
chemical etching of silicon wafers [20,21] and chemical vapor de- 2.1. Materials
position of silane [22,23], are controllable in designing Si
nanostructures but generally associated with low yield and high cost. Pal (purity > 93%, Anhui, China) and calcium-rich Mt (purity >
Due to the facile operation and low processing temperature, magne- 95%, Inner Mongolia, China) in this study were used as received
siothermic reduction of silica is a promising and scalable method to without further purification. Hal (Shanxi, China) was collected via
synthesize nanostructured Si [24]. However, to control the morphology simple sedimentation (purity > 98%), and then dried overnight at
of Si nanostructures, complex designs involving in additional templates 80 °C for the following experiments. Chemical compositions determined
are usually introduced [25,26], which make the procedures energy- by XRF results of the three clay minerals (wt.%) were listed (Table S1).
intensive and time-consuming. As such, a facile, low-cost, and con- Magnesium (Mg, analytical grade) was purchased from Shanghai
trollable strategy for fabricating the nanostructured Si with different Aladdin Chemical Reagent Co., Ltd and sodium chloride (NaCl, analy-
morphology is of high significance. tical grade) from Shanghai RichJoint Chemical Reagents Co., Ltd. Both
With high silicon content, uniform nano-sized architectures, low hydrofluoric acid (HF, 40 wt.%) and hydrochloric acid (HCl, 37 wt.%)
price, and natural abundance [27–29], clay mineral recently has been were obtained from Guangzhou Chemical Reagent Factory.
regarded as one of the most attractive precursors for the synthesis of
nanostructured Si via magnesiothermic reduction [30–32]. As the metal 2.2. Preparation of nanostructured silicon
ions (especially Al) in clay minerals may easily react with other oxides
to form high temperature byproducts (e.g., spinel) during the heavily The nanostructured Si was synthesized by a combination of mag-
exothermic reaction, previous studies always first transformed clay nesiothermic reduction reaction and salt melt synthesis. According to
minerals to amorphous SiO2 (to remove metal ions) by strong acid the results of preliminary experiments (Fig. S1), the mass ratio of 1: 0.8:
etching [31,32]. Obviously, this method will severely destroy the ori- 5 for clay: Mg: NaCl was selected. In a typical procedure, 1 g of clay
ginal nanostructures of clay minerals and can hardly tailor the mor- mineral (e.g., Mt), 0.8 g of Mg powders, and 5 g of NaCl were uniformly
phology of nanostructured Si. Meanwhile, Adpakpang et al. also suc- mixed and then sealed into a stainless steel reactor in a glovebox filled
cessfully synthesized porous Si nanoplate without the generation of with Ar. This reactor was transferred to a corundum tube furnace and
high temperature phases by directly employing Al-free exfoliated clay heated with a heating ramp rate of 5° min−1 and maintained at 650 °C
(Laponite) nanosheet as precursor [30]. for 5 h in a high-purity Ar flow. After cooling down naturally, the ob-
However, considering the diverse structures and complex chemical tained products were rinsed with distilled water to remove NaCl (which
compositions of clay minerals, a universal guidance from clay minerals could be easily recycled and reused by evaporating the supernatant).
to various Si nanostructures is of high necessity but remains a chal- Thereafter, the above products were further washed with 1 M HCl for
lenge. Theoretically, choosing the proper heat absorbent could lower 3 h to eliminate MgO and other impurities (e.g., Mg2Si), followed by
the system temperature and thus prevent the formation of the un- leaching with 1% HF. The resulting nanostructured Si was thoroughly
favorable high temperature phases. As such, clay minerals with a washed with distilled water and then vacuum-dried at 60 °C. For the
variety of special nanostructures wherein element Si is evenly dis- sake of convenience, the final Si product using Mt as precursor was
tributed might be in situ converted to various Si nanostructures without denoted as Si(Mt) and the others were denoted in the same way.
any pretreatment and templates. On the other hand, although various Si
nanostructures individually exhibited excellent electrochemical per- 2.3. Characterization methods
formance, there have been very limited experimental studies on dis-
tinguishing the electrochemical performance of Si nanostructures with Powder X-ray diffraction (XRD) patterns were collected on a Bruker
various morphology synthesized under the same conditions (e.g., with D8 ADVANCE X-ray diffractometer (Bruker AXS, Germany) at 40 kV
the same preparation method and from precursors with similar che- and 40 mA with Ni-filtered CuKα radiation (λ = 0.154 nm). Raman
mical compositions). By combining the abundant nanostructures of the spectra were recorded on a Renishaw 2000 confocal micro-Raman
synthesized Si (from clay minerals) with their electrochemical perfor- Spectrometer (Renishaw, UK) with an Ar-ion laser (λ = 514.5 nm) and
mance, one may reveal the essential structure-property relationship of an air-cooled CCD detector. X-ray photoelectron spectroscopy (XPS)
various Si nanostructures. measurements were performed on a Thermo Fisher K-Alpha XPS in-
This work, for the first time, established the relationships among strument (Thermo Fisher Scientific, UK) with a monochromatic Al-Kα

62
Q. Chen et al. Journal of Power Sources 405 (2018) 61–69

X-ray source (1468.6 eV). Elemental components were determined by a attributed to the decrease in crystal sizes of the obtained Si. According
Rigaku ZSX 100e X-ray fluorescence spectrometer (XRF, Rigaku, Japan) to the Scherrer equation, the calculated crystal sizes of Si(Hal), Si(Mt),
and a GVI IsoProbe multicollector inductively-coupled plasma mass and Si(Pal) were ∼90, 49, and 24 nm, respectively.
spectrometry (MC-ICPMS, GV Instruments, UK). N2 adsorption-deso- The Raman spectra of the resulting nanostructured Si showed the
rption isotherms were carried out at liquid nitrogen temperature Raman-scattering bands at ∼507 cm−1 (Fig. 1b) arising from the
(−196 °C) on a Micromeritics ASAP 2020 system (Micromeritics, USA). scattering of the first-order optical phonon of Si [37]. Compared with
The samples were degassed at 200 °C under vacuum for 12 h before the intense Raman band (521 cm−1) of bulk crystalline Si (Fig. S4),
measurement. The specific surface area (SBET) of the sample was cal- those of nanostructured Si shifted to a lower frequency and became
culated via the multi-point Brunauer–Emmett–Teller (BET) equation, broad and asymmetric, which could be owing to the decrease of dia-
and the pore size distribution (PSD) was evaluated by the Barrett- meters (i.e., the size-confinement effect as reported in previous studies)
Joyner-Halenda (BJH) method. Scanning electron microscopy (SEM) [37–39]. Similarly, the successively weakened Raman-scattering bands
and energy dispersive X-ray spectroscopy (EDS) were measured via an of Si(Hal), Si(Mt), and Si(Pal) implied the decrease in sizes of the na-
SU-8010 cold field emission scanning electron microscope (Hitachi, nostructured Si, as observed in the XRD results.
Japan). Transmission electron microscopy (TEM) and selected area The high resolution Si 2p XPS spectra of the nanostructured Si can
electronic diffraction (SAED) were recorded using an FEI Talos F200S be deconvolved into two primary peaks at ∼99.9 and ∼99.3 eV, which
field-emission transmission electron microscope (FEI Co., USA) with could be assigned to Si 2p1/2 and Si 2p3/2 of elemental Si [40], re-
SAED attachments at an acceleration voltage of 200 kV. spectively (Fig. 1c). Several minor peaks at ∼100.5, ∼101.5, and
∼103.0 eV, corresponding to SiOx derivatives with different oxidation
2.4. Electrochemical measurements states (Si+, Si2+, and Si3+, respectively) [30,41], may be attributed to
the post-oxidation of Si nanocrystals when exposed to the air. No trace
The electrochemical performance of the resulting nanostructured Si of SiO2 peak arising at ∼103.5 eV suggested the complete reduction of
was evaluated in a coin-type cell (CR2032) assembled in an Ar-filled the clay minerals.
glovebox (< 0.1 ppm of H2O and O2). The working electrode was The specific surface area and porosity of the nanostructured Si were
prepared by casting the well-mixed slurry on a clean cooper foil, and analyzed by N2 adsorption-desorption isotherms (Fig. 1d). All of the
then vacuum-dried at 60 °C for 12 h. The slurry consisted of 60% active nanostructured Si exhibited the same IV-type isotherms and H3-type
materials, 20% acetylene black (Super P), and 20% sodium alginate hysteresis loops which were indicative of slit-like mesopores of Si na-
(SA, analytical grade) binder by mass. The typical mass loading density noparticles based on the IUPAC classification [42]. In contrast to Si(Mt)
of the nanostructured Si was about 0.6–0.8 mg cm−2. Then, the elec- and Si(Hal), Si(Pal) had better N2 adsorption capacity in the high re-
trode sheet was cut into disks for coin-type cell fabrication. Lithium lative pressure range, indicating the larger external surface area re-
foils were used as the counter and reference electrodes, and poly- sulting from the interspaces between agglomerated Si nanoparticles.
ethylene films (Celgard 2400) as separators. The electrolyte was com- The SBET of Si(Hal), Si(Mt), and Si(Pal) were calculated to be ∼80, 83,
posed of 1 M LiPF6 solution in 1:1 v/v ethylene carbonate/dimethyl and 116 m2 g−1, respectively. Furthermore, the PSD curves manifested
carbonate (EC/DMC) and 10 wt.% fluorinated ethylene carbonate (FEC) that all of the Si samples displayed similar broad distribution of pore
additives. Galvanostatic charge-discharge measurements were con- size in the range of 2–100 nm centered at ∼25 nm, whereas the pore
ducted using a multichannel Battery Test System (LANHE, CT2001A) at volume increased in the order Si(Hal) < Si(Mt) < Si(Pal) (insert of
room temperature with a fixed voltage range of 0.001–1.5 V (vs. Li/ Fig. 1d).
Li+). The cyclic voltammetry (CV) plots were performed on a CHI660D The morphology and microstructure of the obtained Si were further
electrochemical workstation at a scan rate of 0.1 mV s−1 in the voltage investigated by SEM and TEM images. Si(Mt) exhibited a lamellar and
range of 0.001–1.5 V (vs. Li/Li+). The electrochemical impedance porous structure (Fig. 2d), somewhat maintaining the original 2D
spectroscopy was conducted in the same equipment by using an alter- structure of Mt, while Si(Hal) and Si(Pal) showed a 3D interconnected
nating current perturbation of 5 mV in the frequency range from 0.01 to framework and a loose texture composed of 0D nanoparticles (Fig. 2a,
100 kHz. The specific capacity in this study was evaluated based on the g), respectively, totally different from the original morphology of their
mass of active materials only. The cells after cycling were disassembled precursors. The corresponding EDS data and XPS survey spectra re-
in an Ar-filled glovebox, and the working electrodes were thoroughly vealed a high purity of the resulting nanostructured Si (the existence of
washed with acetonitrile to remove residual electrolytes, and then dried Si and O and the absence of metal elements) (Fig. S5), which was
for further measurements. confirmed by the elemental analysis results based on MC-ICPMS (Table
S2). The TEM images further indicated that the Si nanoparticles were
3. Results and discussion intrinsically interconnected by bridging or overlapping with each other,
forming abundant mesopores and macropores (Fig. 2b, e, and h), con-
3.1. Morphological and structural analysis of nanostructured silicon sistent with the results of pore structure analysis. The SAED patterns
and the interplanar spacing values of ∼0.31 and 0.19 nm (well
As shown by the crystal structure diagrams (Figs. S2a, d, g), the clay matching with the (111) and (220) planes of polycrystalline Si) verified
minerals in this work were phyllosilicates in which the individual layers the generation of high crystalline Si (Fig. 2c, f, i) [43,44]. Besides, the
mainly consisted of a silicon-oxygen tetrahedral (T) sheet and an alu- HRTEM images demonstrated that the crystal sizes of the nanos-
minum- or magnesium-oxygen/-hydroxide octahedral (O) sheet in ei- tructured Si decreased in the order Si(Hal) > Si(Mt) > Si(Pal), which
ther a 1:1 (Hal) or 2:1 (Mt and Pal) proportion [27]. Accordingly, en- was in good agreement with XRD and Raman results. These results
tirely different morphologies were observed in the SEM and TEM above clearly showed that the silicon precursors had a great influence
images (Fig. S2), i.e., nano-sized tubes for Hal, 2D layers for Mt, and on the morphology and microstructure of the corresponding reduction
fibrous structures for Pal, respectively. These three clay minerals with products.
high purity (Fig. S3) were employed as the starting materials without
pretreatment. 3.2. Formation mechanism of nanostructured silicon with different
The structure and composition of the resulting Si products were morphology
summarized (Fig. 1). The sharp characteristic reflections of crystalline
Si appeared in the XRD patterns of all the resulting products (Fig. 1a) The ex situ XRD patterns of intermediates obtained from clay mi-
[35,36]. Noticeably, the intensity of X-ray reflections gradually de- nerals showed that MgO, Mg2Si, and NaCl appeared simultaneously in
creased in the order Si(Hal) > Si(Mt) > Si(Pal), which could be the intermediate at 500 °C (Fig. 3a and Fig. S6). The characteristic

63
Q. Chen et al. Journal of Power Sources 405 (2018) 61–69

Fig. 1. Characterization of nanostructured silicon samples from different clay minerals (i.e., Si(Pal), Si(Mt), and Si(Hal)). (a) XRD patterns, (b) Raman spectra, (c)
High resolution Si 2p XPS spectra, and (d) N2 adsorption-desorption isotherms (inset: the corresponding BJH pore size distributions).

reflection arising from Si became evident since 550 °C. With the tem- during the reduction process, which was attributed to the endothermic
perature further increasing, the peak intensity for MgO and Si gradually effect via the fusion of NaCl [47]. Noticeably, the reflections of Al were
increased, while that for Mg2Si decreased. These phenomena suggested exclusively discernible in the XRD pattern of intermediate from Hal
that Mg2Si acted as an intermediate phase during the reduction process, (Fig. 3 and Fig. S6), arising from the reaction of Al2O3 (s) + 3 Mg (s, l)
which were further supported by the ex situ XRD patterns of the in- → 2Al (s, l) + 3MgO (s). However, element Al could be detected in all
termediates at 650 °C for different times (Fig. S7). Generally, Si was of the EDS mapping of intermediates, which was ascribed to the
formed via the direct magnesiothermic reduction reaction of SiO2 amorphous Al2O3 converted from the O sheets during the exothermic
(s) + 2 Mg (s, l) → Si (s) + 2MgO (s). The above results implied an- reaction [48].
other way for the formation of Si, i.e., via the reactions of 4 Mg (s, Based on above analyses, the reaction mechanism of nanostructured
l) + SiO2 (s) → Mg2Si (s) + 2MgO (s) and Mg2Si (s) + SiO2 (s) → 2Si Si with different structure could be deduced as follows: The magne-
(s) + 2MgO (s), which were thermodynamically possible [45,46]. The siothermic reduction reaction between the external silicon-oxygen T
similar phenomenon was also found in the synthesis of silicene flowers sheets and Mg was initiated at ∼550 °C. The released massive heat from
[24]. Moreover, the SEM images and EDS analyses of intermediates the exothermic reaction could be efficiently absorbed by the heat sca-
showed that the final Si products were fully covered by the byproducts venger NaCl, consequently preventing the reaction system from being
(e.g., MgO) and NaCl (Fig. 3b and Fig. S6), which was further supported overheated. Unlike those SiO2 precursors without any pores, the clay
by the interlaced and/or overlapped lattice fringes of NaCl(111), Si minerals possessed a large amount of pores (e.g., the nanoscale lumens
(111), and MgO(200) in the HRTEM images (Fig. 3d). After subsequent of Hal, the interlayer spaces of Mt, and the structural channels of Pal),
selective etching, a high-purity nanostructured Si was obtained. When and the molten salts and liquid Mg could easily enter into the interlayer
the clay minerals were directly used as precursors via the magne- spaces and/or structural channels of clay minerals to ensure a complete
siothermic reaction without introducing NaCl, the highly exothermic reaction (Fig. S9). For 1: 1 type Hal, the intercalated Mg could si-
process would easily lead to the formation of high temperature phases multaneously react with both T and O sheets to produce Si and Al
(e.g., mullite and spinel) (Fig. S8), owing to the complex chemical (Fig. 3a), respectively, and the newly formed Al further reacted with
compositions of the clay minerals [27]. As such, the absence of any neighbor T sheet via the reaction of 4Al (s, l) + 3SiO2 (s) → 2Al2O3
impurity phases in the nanostructured Si synthesized in the case of NaCl (s) + 3Si (s) [49]. As a result, a 3D interconnected silicon framework
as heat absorbent convincingly verified the relatively low temperature with large pores was obtained after eliminating the molten salts and

64
Q. Chen et al. Journal of Power Sources 405 (2018) 61–69

Fig. 2. Morphological characterization of nanostructured silicon samples from different clay minerals. SEM (left), TEM (middle), and HRTEM images (right) of Si
(Hal) (a–c), Si(Mt) (d–f), and Si(Pal) (g–i). The insets in the right panel represent the corresponding SAED patterns and magnified TEM images showing crystalline
structure.

byproducts (Fig. 4a), as observed in the SEM and TEM images (Fig. 2). occurred in the cathodic scan of first cycle but disappeared in sub-
However, the interlamellar Mg could only contact with the adjacent T sequent cycles, which could be ascribed to the decomposition of elec-
sheets for 2: 1 type clay minerals. For Mt (Fig. 4b), the O sheets served trolytes and the generation of the SEI membrane [50,51]. A sharp
as isolation barriers to prevent the random bonding of the unsaturated cathodic peak at less than 0.1 V in the first and following scans corre-
bonds on newly generated Si during the reduction reaction, conse- sponded to the transformation of crystalline Si to an amorphous phase.
quently maintaining a similar 2D structure as the original Mt. Com- Meanwhile, a new reductive peak at ∼0.2 V emerged gradually, owing
pared with layered Mt, chain-layered Pal lacks the 2D continuous O to the lithiation process of amorphous Si formed from the initial cycle
sheets due to the periodically reversal of T sheets [27]. Therefore, [52]. In the first anodic scan, two distinct peaks centered at 0.37 and
without available isolation barriers, Si(Pal) tended to form a loose 0.51 V were observed, which were attributed to the extraction of li-
texture composed of 0D nanoparticles (Fig. 4c). Besides the direct thium ions from LixSi alloy to generate amorphous Si [53,54]. The
magnesiothermic reduction reaction, Si could be also synthesized via an enhanced intensities of both oxidative and reductive peaks over the first
intermediate phase Mg2Si in all of the reaction systems above (Fig. 4). few cycles indicated an activation process of nanostructured Si, which
The mechanism of magnesiothermic reduction from these three clay was consistent with the previously reported Si-based anode materials
minerals to nanostructured Si could also be extended to designing Si [55–57]. In addition, the CV patterns of Si(Hal) and Si(Pal) were quite
nanostructure with specific morphology and structure from other clay similar to those of Si(Mt) (Figs. S10a and c), suggesting the similar
minerals. fundamental lithiation and delithiation processes.
The galvanostatic discharge/charge curves of the Si(Mt) anode for
the first and second cycles at a current density of 0.2 A g−1 in the
3.3. Electrochemical performance of nanostructured silicon voltage range of 0.001–1.5 V were depicted (Fig. 5b). All of the dis-
charge/charge voltage plateaus matched well with the reductive/oxi-
The electrochemical performance of the resulting nanostructured Si dative peaks in the CV curves. The first discharge curve exhibited a
as anode material of LIBs was evaluated using CR2032 coin-type half sloping voltage plateau between 0.1 and 1.3 V and a long flat one at
cells. Typical CV curves for the first five cycles of the Si(Mt) anode at a ∼0.1 V, generally associated with the formation of the stable SEI film
scan rate of 0.1 mV s−1 were recorded to investigate the redox reactions and the lithiation process of crystalline Si [58,59], respectively. The
related to the charge and discharge processes (Fig. 5a). Two irreversible later discharge plateau moved to a higher voltage (∼0.2 V) in the
peaks (i.e., a weak one at ∼1.26 V and a broad one at ∼0.80 V)

65
Q. Chen et al. Journal of Power Sources 405 (2018) 61–69

Fig. 3. Characterization of intermediates obtained from Hal (without acid washing). (a) Ex situ XRD patterns of as-reduced Hal at different temperatures. (b) SEM
image and EDS mapping data, (c) TEM image, and (d) the corresponding HRTEM image of the intermediate synthesized at 650 °C.

Fig. 4. Schematic illustration showing the different reaction processes between the metallic reducing agent Mg and clay minerals (i.e., Hal (a), Mt (b), and Pal (c)).

second cycle, which was due to the phase transition from crystalline Si mAh g−1, corresponding to an initial Coulombic efficiency (CE) of
to amorphous structure in the first cycle [60]. On the other hand, the 72.1%. The capacity loss was generally assigned to the irreversible
charge voltage plateaus of the first two cycles were ∼4.5 V, suggesting formation of SEI film and the side reactions between active materials
the delithiation process of LixSi alloy. Moreover, the Si(Mt) anode dis- and electrolyte, which usually happened for most anode materials
played the initial specific discharge/charge capacities of 2853/2057 [61–63]. Compared to Si(Mt), both Si(Hal) and Si(Pal) possessed the

66
Q. Chen et al. Journal of Power Sources 405 (2018) 61–69

Fig. 5. (a) CV curves of the Si(Mt) electrode at a scan rate of 0.1 mV s−1 in the voltage range of 0.001–1.5 V (vs. Li/Li+); (b) galvanostatic discharge/charge curves of
the Si(Mt) electrode at a current density of 0.2 A g−1 between 0.001 and 1.5 V (vs. Li/Li+); (c) rate capability of the Si(Mt) electrode at various current densities from
0.1 to 10 A g−1; (d) comparative cycling performance and the corresponding Coulombic efficiencies of the Si(Mt), Si(Pal), and Si(Hal) electrodes at current rates of
0.2 A g−1 for the initial three cycles and 1.0 A g−1 for the remaining cycles.

analogous discharge/charge feature, with the initial CEs of 74.1% and high CE over 99% after 40 cycles, while Si(Pal) and Si(Hal) required
66.3% (Figs. S10b and d), respectively. The lowest initial CE of Si(Pal) near 70 and 90 cycles to reach such a level of CE, respectively (Fig. 5d).
may be due to its highest specific surface area, which could facilitate The improved CE of Si(Mt) confirmed that 2D Si nanostructures could
the generation of SEI membrane. restrain the irreversible depletion of lithium ions. Recently, the variable
The Si(Mt) anode delivered the reversible specific discharge capa- SEI surface areas per unit mass of Si nanostructures with different
cities of 2453, 2218, 1949, 1657, 1385, 1001, and 562 mAh g−1 at morphology during cycling were calculated theoretically to increase in
current densities of 0.1, 0.2, 0.5, 1.0, 2.0, 5.0, and 10 A g−1 (Fig. 5c), the order of 2D < 0D < 3D by the impressive work by Zhang et al.
respectively. Remarkably, the specific capacity retained 562 mAh g−1 [24]. In view of the large variable SEI surface areas during cycling, the
even at 10 A g−1, which was still higher than the calculated capacity of utilization of 0D Si(Pal) and 3D Si(Hal) inevitably needed additional
commercial graphite anode (372 mAh g−1). When the current density interfacial stabilization. By contrast, the 2D structure of Si(Mt) could
was changed from 10 to 0.1 A g−1, the capacity could be almost fully effectively accommodate the volume change of Si with introducing the
returned to the original value, implying an outstanding reversibility of most limited variable SEI area, which consequently enhanced the sta-
Si(Mt) anode. In addition, Si(Hal) and Si(Pal) also showed parallel rate bilization of the interface and Si nanostructures. On the other hand, in
capacities to Si(Mt) (Fig. S11). The good rate performance of all the spite of the smaller size and larger SBET and pore volume, 0D Si(Pal) still
nanostructured Si from clay minerals could be ascribed to well-crys- showed poorer cycling performance than 2D Si(Mt). This results de-
tallized Si nanostructures with large SBET and pore volume, which could monstrated that regulating the morphology had a stronger influence
alleviate the large strain resulting from the repeated lithiation and than downscaling the particle size to improve the cycling stability of Si
delithiation processes. The hierarchical pore structures (the in- nanostructures when the size was reduced to a certain extent (e.g., <
corporation of macro-/mesopores) also contributed to improving the 100 nm).
electrode structure stability and electrochemical kinetics [56]. As shown by the TEM images of the nanostructured Si after cycling
The cycling performance of the nanostructured Si was further ex- (Fig. S12), all of the Si nanostructures exhibited the large volume
amined (Fig. 5d). All of the electrodes were first activated at 0.2 A g−1 change and became amorphous. However, Si(Mt) did not break into
for the initial three cycles and then cycled at 1.0 A g−1 for the re- smaller particles and remained intact after cycling compared to Si(Hal)
maining cycles. With the similar initial specific capacity after the ki- and Si(Pal), suggesting the better structural and interfacial stability of
netic enhancement, the Si(Mt), Si(Pal), and Si(Hal) anodes exhibited 2D Si nanostructure during cycling. This character was further con-
reversible capacities of 1369, 1163, and 928 mAh g−1 after 200 cycles, firmed by the Nyquist plots based on electrochemical impedance
corresponding to capacity retentions of 78%, 72%, and 56%, respec- spectroscopy (Fig. S13). The charge transfer resistance at the surface of
tively. The descending cycling performance of the above three Si an- anode and the interfacial resistance between electrolyte and electrode
odes could be attributed to the distinct interfacial characteristic of could be estimated by the diameter of the high-frequency semicircle
various Si nanostructures. Moreover, the Si(Mt) electrode showed a [64–67]. The high-frequency semicircle diameter of the Si(Mt)

67
Q. Chen et al. Journal of Power Sources 405 (2018) 61–69

Fig. 6. Volume expansion tests of the electrodes. Cross-sectional SEM images of the Si(Mt) (a, b), Si(Pal) (c, d), and Si(Hal) (e, f) electrodes before electrochemical
test (a, c, e) and after 100 cycles at 1.0 A g−1 (b, d, f).

electrode showed the least variation from 5 to 100 cycles among the practical viewpoints compared to other recently reported Si anodes
three electrodes, indicating the minimum change of charge-transfer synthesized by other methods (Table S3). In future work, we will in-
characteristics of the Si(Mt) electrode. In addition, the cross-sectional troduce a second phase (e.g., carbon, metal, and conductive polymer)
SEM images of the Si electrodes were measured to evaluate the volume on the surface of nanostructured Si. Such coating is expected to serve as
change in macroscopic view before and after 100 cycles at 1.0 A g−1 a soft media to buffer the stress derived from volume expansion and
(Fig. 6). The Si(Mt) electrode exhibited a relatively minor volume ex- improve the transport property, thus further enhancing the cycling
pansion (∼54%) in comparison with Si(Hal) (> 103%) and Si(Pal) stability of nanostructured Si.
(> 64%), which further manifested the enhanced interfacial stabiliza-
tion of 2D nanostructured Si.
Based on the above analysis, the high specific capacity, good rate 4. Conclusions
capability, and stable cycling performance of 2D Si(Mt) could be pri-
marily attributed to two aspects: First, compared with 0D Si(Pal) and In summary, three Si nanostructures with various morphology were
3D Si(Hal), more limited variable SEI area of 2D nanostructure could successfully synthesized directly using clay minerals as precursors via a
enhance the stabilization of the interface and nanostructures, thus molten salt-modified magnesiothermic reduction reaction. The mor-
minimizing volume variation and maintaining structural integrity phology and microstructure of the resulting nanostructured Si sig-
during cycling. Second, the intrinsically interconnected feature of nificantly depended on the nature of clay minerals. Si(Mt) exhibited a
planar nanostructured Si contributed to robust and fast electron/ion 2D porous structure, maintaining the original morphology of Mt, while
transport paths, which allowed Si(Mt) to stand relatively high current Si(Hal) and Si(Pal) showed a 3D interconnected framework and a loose
density. Without the introduction of complicated preparation and sur- texture composed of 0D nanoparticles, respectively, totally different
face engineering strategies, the Si nanostructures in our work exhibited from their precursors. Without the introduction of sophisticated designs
good electrochemical performance for lithium storage in terms of and surface engineering strategies, all of the Si nanostructures exhibited
good electrochemical performance as the anodes of LIBs due to the

68
Q. Chen et al. Journal of Power Sources 405 (2018) 61–69

large SBET and hierarchical pore structure. Compared with Si(Hal) and [22] A. Magasinski, P. Dixon, B. Hertzberg, A. Kvit, J. Ayala, G. Yushin, Nat. Mater. 9
Si(Pal), Si(Mt) exhibited a more stable cycling performance (1369 mAh (2010) 353–358.
[23] X.K. Huang, J. Yang, S. Mao, J.B. Chang, P.B. Hallac, C.R. Fell, B. Metz, J. Jiang,
g−1 at 1.0 A g−1 with an capacity retention of 78% over 200 cycles) P.T. Hurley, J. Chen, Adv. Mater. 26 (2014) 4326–4332.
attributed to the limited variable SEI area of 2D nanostructure and [24] X.H. Zhang, X.Y. Qiu, D.B. Kong, L. Zhou, Z.H. Li, X.L. Li, L.J. Zhi, ACS Nano 11
more efficient transport property. This work, for the first time, estab- (2017) 7476–7484.
[25] Z.Y. Lu, J.X. Zhu, D.H. Sim, W.W. Zhou, W.H. Ship, H.H. Hng, Q.Y. Yan, Chem.
lished the relationships among natural clay minerals with different Mater. 23 (2011) 5293–5295.
structure, nanostructured Si with various morphology, and the corre- [26] X.L. Liu, Y.F. Gao, R.H. Jin, H.J. Luo, P. Peng, Y. Liu, Nanomater. Energy 4 (2014)
sponding electrochemical properties. We disclosed the formation me- 31–38.
[27] F. Bergaya, G. Lagaly, Chapter 1 general introduction: clays, clay minerals, and clay
chanism from clay minerals to nanostructured Si via magnesiothermic science, in: F. Bergaya, B.K.G. Theng, G. Lagaly (Eds.), Developments in Clay
reduction, which could be extended to designing Si nanostructure with Science, Elsevier, 2006, pp. 1–18.
specific morphology using other clay minerals. Moreover, we revealed [28] R.L. Zhu, Q.Z. Chen, Q. Zhou, Y.F. Xi, J.X. Zhu, H.P. He, Appl. Clay Sci. 123 (2016)
239–258.
the fundamental structure-property relationship of different Si nanos-
[29] E. Ruiz-Hitzky, P. Aranda, M. Darder, G. Rytwo, J. Mater. Chem. 20 (2010)
tructures synthesized under the same conditions, which would con- 9306–9321.
tribute to the practical applications of nanostructured Si anodes in LIBs. [30] K. Adpakpang, S.B. Patil, S.M. Oh, J.-H. Kang, M. Lacroix, S.-J. Hwang, Electrochim.
Acta 204 (2016) 60–68.
[31] L. Sun, T.T. Su, L. Xu, H.-B. Du, Phys. Chem. Chem. Phys. 18 (2016) 1521–1525.
Acknowledgments [32] X.Y. Zhou, L.L. Wu, J. Yang, J.J. Tang, L.H. Xi, B. Wang, J. Power Sources 324
(2016) 33–40.
This work was financially supported by the National Natural Science [33] Z. Hu, X. Xiao, H. Jin, T. Li, M. Chen, Z. Liang, Z. Guo, J. Li, J. Wan, L. Huang,
Y. Zhang, G. Feng, J. Zhou, Nat. Commun. 8 (2017) 15630.
Foundation of China (41572031), grants from the National Program for [34] X.F. Liu, C. Giordano, M. Antonietti, J. Mater. Chem. 22 (2012) 5454–5459.
Support of Top-notch Young Professionals, Guangdong Provincial [35] M.-S. Wang, L.-Z. Fan, M. Huang, J. Li, X. Qu, J. Power Sources 219 (2012) 29–35.
Youth Top-notch Talent Support Program (2014TQ01Z249), [36] J. Feng, Z. Zhang, L. Ci, W. Zhai, Q. Ai, S. Xiong, J. Power Sources 287 (2015)
177–183.
Guangdong Special Support Program, Team Project of Natural Science [37] R.P. Wang, G.W. Zhou, Y.L. Liu, S.H. Pan, H.Z. Zhang, D.P. Yu, Z. Zhang, Phys. Rev.
Foundation of Guangdong Province, China (S2013030014241), Science B 61 (2000) 16827–16832.
and Technology Planning Project of Guangdong Province, China [38] W. Luo, X. Wang, C. Meyers, N. Wannenmacher, W. Sirisaksoontorn, M.M. Lerner,
X. Ji, Sci. Rep. 3 (2013) 2222.
(2017B030314175), and the Foundation of Key Laboratory of Clay [39] N.H. Hai, I. Grigoriants, A. Gedanken, J. Phys. Chem. C 113 (2009) 10521–10526.
Mineral Applied Research of Gansu Province, Lanzhou Institute of [40] X.X. Wang, J. Xu, Q.L. Wang, A.Q. Xu, Y.S. Zhai, J.R. Luo, Y. Jiang, N.Y. He,
Chemical Physics, Chinese Academy of Sciences (CMAR-2017-01). The Z.F. Wang, Small 13 (2017) 1603369.
[41] H.G. Song, D. Liu, J. Yang, L. Wang, H.X. Xu, Y.J. Xiong, Chemnanomat 3 (2017)
authors also thank S. S. Wei from GIGCAS in Guangzhou for TEM
22–26.
analysis. This is contribution No. IS-2596 from GIGCAS. [42] M. Thommes, K. Kaneko, A.V. Neimark, J.P. Olivier, F. Rodriguez-Reinoso,
J. Rouquerol, K.S.W. Sing, Pure Appl. Chem. 87 (2015) 1051–1069.
Appendix A. Supplementary data [43] A. Xing, J. Zhang, Z.H. Bao, Y.F. Mei, A.S. Gordin, K.H. Sandhage, Chem. Commun.
49 (2013) 6743–6745.
[44] J. Chun, S. An, J. Lee, J. Mater. Chem. 3 (2015) 21899–21906.
Supplementary data to this article can be found online at https:// [45] W.-S. Kim, Y. Hwa, J.-H. Shin, M. Yang, H.-J. Sohn, S.-H. Hong, Nanoscale 6 (2014)
doi.org/10.1016/j.jpowsour.2018.10.031. 4297–4302.
[46] M. Barati, S. Sarder, A. McLean, R. Roy, J. Non-Cryst. Solids 357 (2011) 18–23.
[47] B. Zalba, J.M. Marin, L.F. Cabeza, H. Mehling, Appl. Therm. Eng. 23 (2003)
References 251–283.
[48] Q.Z. Chen, R.L. Zhu, L.Y. Ma, Q. Zhou, J.X. Zhu, H.P. He, Appl. Clay Sci. 135 (2017)
129–135.
[1] W. Luo, X.Q. Chen, Y. Xia, M. Chen, L.J. Wang, Q.Q. Wang, W. Li, J.P. Yang, Adv.
[49] N. Yoshikawa, A. Kikuchi, S. Taniguchi, J. Am. Ceram. Soc. 85 (2002) 1827–1834.
Energy Mater. 7 (2017) 1701083.
[50] N. Lin, J.B. Zhou, L.B. Wang, Y.C. Zhu, Y.T. Qian, ACS Appl. Mater. Interfaces 7
[2] M.N. Obrovac, V.L. Chevrier, Chem. Rev. 114 (2014) 11444–11502.
(2015) 409–414.
[3] X.X. Zuo, J. Zhu, P. Müller-Buschbaum, Y.-J. Cheng, Nanomater. Energy 31 (2017)
[51] J. Li, J.R. Dahn, J. Electrochem. Soc. 154 (2007) A156–A161.
113–143.
[52] F. Lin, D. Nordlund, T.-C. Weng, Y. Zhu, C. Ban, R.M. Richards, H.L. Xin, Nat.
[4] L. Lin, X. Xu, C. Chu, M.K. Majeed, J. Yang, Angew. Chem. Int. Ed. 55 (2016)
Commun. 5 (2014) 3358.
14063–14066.
[53] M. Gauthier, D. Mazouzi, D. Reyter, B. Lestriez, P. Moreau, D. Guyomard, L. Roue,
[5] R. Yi, M.L. Gordin, D.H. Wang, Nanoscale 8 (2016) 1834–1848.
Energy Environ. Sci. 6 (2013) 2145–2155.
[6] J.Y. Luo, J. Gao, A.X. Wang, J.X. Huang, ACS Nano 9 (2015) 9432–9436.
[54] Y. Xing, T. Shen, T. Guo, X. Wang, X. Xia, C. Gu, J. Tu, J. Power Sources 384 (2018)
[7] Q.B. Yun, X.Y. Qin, W. Lv, Y.-B. He, B.H. Li, F.Y. Kang, Q.-H. Yang, Carbon 93
207–213.
(2015) 59–67.
[55] J. Liu, P. Kopold, P.A. van Aken, J. Maier, Y. Yu, Angew. Chem. Int. Ed. 54 (2015)
[8] Y. Zhang, N. Du, Y. Chen, Y. Lin, J. Jiang, Y. He, Y. Lei, D. Yang, Nanoscale 10
9632–9636.
(2018) 5626–5633.
[56] X.X. Zuo, Y.G. Xia, Q. Ji, X. Gao, S.S. Yin, M.M. Wang, X.Y. Wang, B. Qiu, A.X. Wei,
[9] L.H. Liu, J. Lyu, T.H. Li, T.K. Zhao, Nanoscale 8 (2016) 701–722.
Z.C. Sun, Z.P. Liu, J. Zhu, Y.-J. Cheng, ACS Nano 11 (2017) 889–899.
[10] Y. Yu, L. Gu, C.B. Zhu, S. Tsukimoto, P.A. van Aken, J. Maier, Adv. Mater. 22 (2010)
[57] Z. Favors, W. Wang, H.H. Bay, Z. Mutlu, K. Ahmed, C. Liu, M. Ozkan, C.S. Ozkan,
2247–2250.
Sci. Rep. 4 (2014) 5623.
[11] L.L. Wu, J. Yang, X.Y. Zhou, M.F. Zhang, Y.P. Ren, Y. Nie, J. Mater. Chem. 4 (2016)
[58] Y.H. Xu, Y.J. Zhu, F.D. Han, C. Luo, C.S. Wang, Adv. Energy Mater. 5 (2015)
11381–11387.
1400753.
[12] S.-H. Kim, D.H. Lee, C. Park, D.-W. Kim, J. Power Sources 395 (2018) 328–335.
[59] Y. Han, N. Lin, T. Xu, T. Li, J. Tian, Y. Zhu, Y. Qian, Nanoscale 10 (2018)
[13] Y.L. Zhou, X.L. Jiang, L. Chen, J. Yue, H.Y. Xu, J. Yang, Y.T. Qian, Electrochim. Acta
3153–3158.
127 (2014) 252–258.
[60] J. Wang, X.C. Meng, X.L. Fan, W.B. Zhang, H.Y. Zhang, C.S. Wang, ACS Nano 9
[14] X.L. Wang, G. Li, M.H. Seo, G. Lui, F.M. Hassan, K. Feng, X.C. Xiao, Z.W. Chen, ACS
(2015) 6576–6586.
Appl. Mater. Interfaces 9 (2017) 9551–9558.
[61] Y. An, H. Fei, Z. Zhang, L. Ci, S. Xiong, J. Feng, Chem. Commun. 53 (2017)
[15] J. Ryu, D. Hong, S. Choi, S. Park, ACS Nano 10 (2016) 2843–2851.
8360–8363.
[16] P.P. Wang, Y.X. Zhang, X.Y. Fan, J.X. Zhong, K. Huang, J. Power Sources 379
[62] D.B. Kong, H.Y. He, Q. Song, B. Wang, W. Lv, Q.H. Yang, L.J. Zhi, Energy Environ.
(2018) 20–25.
Sci. 7 (2014) 3320–3325.
[17] J.W. Liang, D.H. Wei, N. Lin, Y.C. Zhu, X.N. Li, J.J. Zhang, L. Fan, Y.T. Qian, Chem.
[63] W. Li, M. Li, J.-a. Shi, X. Zhong, L. Gu, Y. Yu, Nanoscale 10 (2018) 12430–12435.
Commun. 50 (2014) 6856–6859.
[64] Y. An, H. Fei, G. Zeng, X. Xu, L. Ci, B. Xi, S. Xiong, J. Feng, Y. Qian, Nanomater.
[18] L. Sun, F. Wang, T.T. Su, H.B. Du, ACS Appl. Mater. Interfaces 9 (2017)
Energy 47 (2018) 503–511.
40386–40393.
[65] M. Chen, B. Li, X. Liu, L. Zhou, L. Yao, J. Zai, X. Qian, X. Yu, J. Mater. Chem. 6
[19] Y. An, H. Fei, G. Zeng, L. Ci, S. Xiong, J. Feng, Y. Qian, ACS Nano 12 (2018)
(2018) 3022–3027.
4993–5002.
[66] Y. An, Z. Zhang, H. Fei, X. Xu, S. Xiong, J. Feng, L. Ci, J. Power Sources 363 (2017)
[20] Z.L. Zhang, Y.H. Wang, W.F. Ren, Q.Q. Tan, Y.F. Chen, H. Li, Z.Y. Zhong, F.B. Su,
193–198.
Angew. Chem. Int. Ed. 53 (2014) 5165–5169.
[67] Y. An, H. Fei, G. Zeng, L. Ci, B. Xi, S. Xiong, J. Feng, J. Power Sources 378 (2018)
[21] M. Thakur, R.B. Pernites, N. Nitta, M. Isaacson, S.L. Sinsabaugh, M.S. Wong,
66–72.
S.L. Biswal, Chem. Mater. 24 (2012) 2998–3003.

69

You might also like