Download as pdf or txt
Download as pdf or txt
You are on page 1of 59

Journal Pre-proof

A Review of Recent Developments in Si/C Composite Materials for


Li-ion Batteries

Qitao Shi , Junhua Zhou , Sami Ullah , Xiaoqin Yang ,


Klaudia Tokarska , Barbara Trzebicka , Huy Quang Ta ,
Mark H. Rümmeli

PII: S2405-8297(20)30403-7
DOI: https://doi.org/10.1016/j.ensm.2020.10.026
Reference: ENSM 1392

To appear in: Energy Storage Materials

Received date: 11 August 2020


Revised date: 5 October 2020
Accepted date: 29 October 2020

Please cite this article as: Qitao Shi , Junhua Zhou , Sami Ullah , Xiaoqin Yang , Klaudia Tokarska ,
Barbara Trzebicka , Huy Quang Ta , Mark H. Rümmeli , A Review of Recent Developments
in Si/C Composite Materials for Li-ion Batteries, Energy Storage Materials (2020), doi:
https://doi.org/10.1016/j.ensm.2020.10.026

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier B.V.


A Review of Recent Developments in Si/C Composite Materials for Li-ion Batteries

Qitao Shia, Junhua Zhoua, Sami Ullaha, Xiaoqin Yanga,b, Klaudia Tokarskac, Barbara
Trzebickac, Huy Quang Tad, Mark H. Rümmelia,c,d,e*
a
Soochow Institute for Energy and Materials Innovation, College of Energy,Key Laboratory of
Advanced Carbon Materials and Wearable Energy Technologies of Jiangsu Province,
Soochow University, Suzhou 215006, China
b
School of Energy and Power Engineering, Xi’an Jiaotong University, No. 28, Xianning West
Road, Xi’an, Shanxi 710049, China
c
Centre of Polymer and Carbon Materials, Polish Academy of Sciences,M.
Curie-Sklodowskiej 34, Zabrze 41-819, Poland
d
Institute for Complex Materials, IFW Dresden, 20 Helmholtz Strasse, Dresden 01069,
Germany
e
Department of Chemistry, VŠB-Technical University of Ostrava, Czech Republic
*Corresponding author.

E-mail address: mhr1@suda.edu.cn

Qitao Shi received his bachelor of science degree from the Department of physics and

optoelectronic energy at Soochow University, Suzhou, China, in 2016. Currently, he works as

doctoral researcher at the Soochow Institute for Energy and Materials Innovations (SIEMIS)

and the College of Energy at Soochow University China in Prof. Mark H. Rümmeli‘s group.

His current research focuses on solving the pulverization issues of Si particles as anode

materials by space engineering or structure optimization.


1
Mark H. Rümmeli heads the electron microscopy and LIN labs at the Soochow Institute for

Energy and Materials Innovations (SIEMIS), Soochow University, where he is a full

professor. He is also director of the characterization center at the College of Energy and

SIEMES. Moreover, he is a full professor of the Polish Academy of Sciences (CMPW PAN)

in Zabrze and has full habilitation rights. He obtained his Ph.D. from London Metropolitan

University and then worked as a postdoc at the German Aerospace Center. His research

focuses on the growth mechanisms of 2D nanostructures, their functionalization and

application in energy and biomedicine.

Keywords: Si/C composite materials, advanced preparation methods, in-situ characterization

Abstract: Rechargeable lithium batteries play an increasingly significant role in our daily

lives. Hence, the development of high capacity secondary lithium batteries has become a

research hotspot. In the past decade, silicon has been extensively studied as anode material for

Li-ion batteries because of its extremely high specific capacity. However, the dramatic

volume change and troublesome SEI (solid electrolyte interface) issues during lithiation and

delithiation hinder the commercialisation of Si anode materials. To circumvent these issues,

carbon materials have been widely utilized in composites with Si materials due to their

excellent electrochemical and physical properties. Established preparation methods of Si/C

composite materials facilitate the design of novel Si/C composites. Different forms of carbon
2
can improve the electrochemical performance of silicon materials in different ways.

Advanced characterisation techniques further verify and explain the contribution of carbon

materials to the performance improvement of Si. Si/C composite materials are anticipated to

be the anode material for the next generation of commercial lithium batteries.

1. Introduction

The advent of portable electronic products and alternative fuel vehicles has led to an increased

demand for advanced lithium (Li)-ion batteries. High performance Li-ion batteries provide

electric endurance support in electronic products [1, 2], wherein battery performance is

primarily affected by the battery‘s anode materials. Graphite is a popular low-cost commercial

anode material with high conductivity and stable reversibility. However, the material‘s

relatively low capacity has limited its further development.

Electrochemical alloys of Li metal with other compounds (e.g. Si, Sn, P, and Sb) can be used

as alternatives. Si-based anode materials are a popular candidate for next generation Li-ion

batteries due to an extremely high specific capacity of over 10 times that of commercial

graphite.[3, 4] However, Si-based anodes exhibit large volume changes and form an unstable

solid electrolyte interface (SEI) during electrochemical processes (Figures. 1(a-c)) [5].

Recently, composites of carbon and Si (Si/C) based on various carbon materials and structural

designs have been developed to overcome these issues concerning volume expansion and

continuous SEI formation (Figures. 1(d-f)) [6-8]. Carbon coating and controlled porosity have

been found to prevent the pulverization of Si particles [6]. Carbon coating can also ensure the

stability of the solid electrolyte interphase, thus preventing the consumption of the inner Si by

the continuously formed SEI [7]. Careful design of Si/C material structures can ensure strong

bonding between the electrode materials and current collector [8]. Overall, the improved

performance of Si-based anode materials has been attributed to unique structural designs and

the excellent properties of the carbon materials. Several literature reviews have explored the
3
development of Si/C composite anode materials for Li-ion batteries from different

perspectives. For instance, Dou et al. [9] assessed Si/C composite materials with different

dimensions in great detail, Zhang et al. [10] introduced nanostructured Si/C materials and

related electrolytes and binders, and Shen et al. [11] summarized the progress in Si/C

materials by highlighting different material structures. However, a comprehensive review

introducing various Si/C composite anode materials and their preparation methods along with

advanced characterization techniques has not yet been published. This review focuses on the

use and preparation of carbon materials to enhance the performance of Si materials, giving a

detailed description of one-dimensional (1D) carbon nanofibers (CNFs) and nanotubes

(CNTs), two-dimensional (2D) graphene (G) sheets and transition metal carbides and

carbonitrides (referred to as MXenes), and three-dimensional (3D) graphene shell or

amorphous carbon shell-coated Si particles and Si/graphite composites. The excellent

performance of Si/C composites is further verified and explained using several advanced

characterization techniques. As for zero-dimensional (0D) carbon materials, we only include

carbon black in this category. In addition, research on Si/carbon black composites is rarely

reported, so we will not introduce these materials.

2. Advanced Preparation Methods of Carbon Materials

Different Si materials have been designed and synthesized for Li-ion batteries using various

methods, including Si-nanowire synthesis by vapor-liquid-solid processing [12] and

solvent-mediated phenylsilane decomposition [13], Si-nanosphere growth on SiO2 by

chemical vapor deposition (CVD) [14], Si-nanoparticle synthesis by the reverse micelles

method [15], and Si-film deposition using a radiofrequency/direct current (RF/DC) magnetron

sputtering system [16]. Furthermore, diverse silicon–oxygen compounds could be candidates

for high-capacity anode materials [17-19]. The synthesis of Si materials is well-established

and has been described in several review articles [20, 21]. Thus, the synthesis of Si materials

4
is not discussed in this review. Instead, novel preparation methods of various carbon materials

and Si/C composite materials is explored.

Carbon materials are typically prepared utilizing six methods: electrospinning, Hummer‘s

method, electrochemical exfoliation, CVD, thermal treatment, and hydrothermal

carbonization. The CVD method is generally used to coat graphitic carbon on solid Si

materials, and it can be used to synthesize carbon nanotubes. Thermal treatment is mainly

used to produce amorphous carbon coatings and to convert graphene oxide (GO) coatings

produced by Hummer‘s method to graphene. CNFs stacked between Si nanoparticles can be

synthesized by electrospinning. Less widely utilized methods, such as electrochemical

exfoliation and hydrothermal methods, are also discussed in this section.

2.1. Preparation Methods of 1D Carbon Materials

2.1.1. Electrospinning

Electrospinning is a specialized fibre manufacturing process in which a polymer solution or

melt is sprayed in a strong electric field. The electric field changes the shape of the droplet at

the needle point from spherical to conical, where a fibrous filament extends from the tip of the

cone. These polymer filaments have nanoscale diameters and are thus classified as 1D

nanofibers. A fundamental setup for electrospinning to produce nanofibers is schematically

shown in Figure. 2 [22]. The carbon precursor to produce CNFs requires a careful selection,

as it determines the carbon structure formed upon carbonization. This CNF structure formed

at high temperature can significantly affect the composite material‘s electrical, chemical, and

mechanical properties.

2.2. Preparation Methods of 2D Carbon Materials

2.2.1. Hummer’s Method

5
Hummer‘s method is an effective method for production of graphene sheets and was first

proposed in 1958 by William S. Hummer. JR [23]. Numerous studies have since contributed

to the optimization of this method to produce graphene [24, 25]. Graphene oxide is produced,

and further processing under high temperature is required to produce 2D graphene sheets for

the fabrication of a Si/G composite. Some studies have synthesized graphene directly after the

formation of graphene oxide [26, 27], but most studies have mixed the graphene oxide with Si

particles to strengthen the bond between the two components before further processing. This

mixing has been found to lead to the formation of Si/G composite materials displaying

superior electrochemical properties when applied in Li-ion batteries [25, 28-31].

2.2.2. Electrochemical Exfoliation

Electrochemical exfoliation is an alternative approach to Hummer‘s method for production of

graphene. A potential difference is applied between a graphite anode/cathode in the presence

of an electrolyte-containing medium [32]. Specifically, graphene is typically synthesized by

electrochemical exfoliation of graphite foil in a two-electrode system comprising graphite foil

and platinum wire as the working and counter electrodes, respectively [33]. After exfoliation,

the product is collected by vacuum filtration, washed with deionized water to remove any

remaining residues and sonicated for further removal of non-exfoliated graphite aggregates.

The product is centrifuged to form dispersed graphene sheets. This approach avoids the

conversion of graphene oxide to graphene, which contributes to its considerable advantage

over Hummer‘s method. As graphene oxide is not generated, the product is easily converted

to graphene sheets, and a regular commercial microwave within seconds can provide the

energy for graphene transformation under inert gas protection. Further, electrochemical

exfoliation is relatively simple and is performed rapidly under mild conditions. Lastly, while

graphene produced by Hummer‘s method typically exhibits high sheet resistance due to flake

6
damage caused by the oxidation and exfoliation processes [32]. these effects can be avoided

using electrochemical exfoliating.

2.3. Preparation Methods of 3D Carbon Materials

2.3.1. CVD Method

CVD has gained popularity in the growth of graphene materials for use in electronic and

optoelectronic devices, chemical sensors, nanocomposites, and energy storage [34]. CVD

breaks down gaseous reactants carried into a tube furnace to produce atomic species, which

are subsequently deposited onto a substrate surface. The CVD process is depicted in Figure. 3.

CVD is mainly used to produce graphitic carbon-coated Si and carbon nanotubes, which are

classified as 1D carbon and can be further composited with Si materials. Further, CVD can

deposit Si onto a substrate using SiH4 or SiHCl3. CVD is a relatively convenient and easily

controlled synthesis method because it involves mostly mechanized operation that requires

less manual work.

Most carbon deposition precursors are organic compounds, namely poly(methyl methacrylate)

(PMMA) [35], propylene [36], acetylene [37, 38], methane [39], ethanol [39] and toluene [40].

However, a novel approach has been proposed in which an inorganic CO2 precursor (solid,

liquid, or gaseous) was used for carbon deposition [41, 42]. Under high temperatures, solid

carbon precursors decompose into atomic species, liquid precursors are typically transferred

into the tube using an inert gas, and gaseous precursors are fed directly into the tube furnace.

Regardless of precursor state, the deposition process is the same, and either amorphous carbon

or graphene coatings can be achieved.

2.3.2. Thermal Treatment

Thermal treatment, commonly referred to as annealing, is based on different principles than

the CVD method, but can be performed using the same equipment. Organic carbon precursors
7
are converted to inorganic carbon during thermal treatment by removal of the hydrogen and

oxygen components under high temperature. The process is performed in an inert gas to

prevent the formation of undesirable carbon compounds, while a high temperature is applied

to provide the energy required to remove hydrogen and oxygen from the carbon source. The

organic carbon precursor or organic carbon coated material is placed within a tubular furnace

in advance to ensure that the entire process proceeds without contamination by foreign carbon

sources, as the carbon source can have significant effects on the product. Thermal treatment is

a simple method for coating 3D amorphous carbon on a Si substrate.

A liquid carbon source may also be used in thermal treatment, including sucrose [43],

resorcinol-formaldehyde resin [44, 45], polydopamine [46, 47], sodium dodecyl benzene

sulfonate [48], and styrene [49]. Further, solid carbon sources can be dissolved in an

appropriate solvent to serve as a liquid carbon precursor. Si materials must be coated in the

same solvent before transfer to the furnace tube. Thermal treatment converts the organic

compound into a carbon shell typically comprised of amorphous carbon.

2.3.3. Hydrothermal Method

Hydrothermal reaction processing is a general term for chemical reactions conducted in water,

aqueous solutions, or steam under a certain temperature and pressure. Hydrothermal

processing is generally performed in an autoclave, which can withstand temperatures of up to

1100 ℃ and pressures of up to 1 GPa owing to its reliable sealing system and explosion-proof

design. A typical hydrothermal approach was used by Jeong et al [50]. Si was ultra-sonicated

in distilled water with sucrose, oxalic acid, cetyltrimethylammonium bromide, and polyvinyl

pyrrolidone, transferred to a Teflon-lined stainless steel autoclave, and held at 200 °C for 10 h.

The solution was cooled, filtered, washed with distilled water, and carbonized under argon

atmosphere to produce Si-embedded carbon microspheres. These spheres included void

tunnels and interconnected amorphous carbon-coated Si particles. This structure maintained


8
the electrochemical activity of the pulverized Si particles by holding the cracked pieces of the

particles in the void space within the carbon skeleton. Moreover, the carbon skeletons

enhanced the conductivity of the electrode.

3. Si/C Composite Materials

3.1. Si/1D Carbon Composite Materials

Carbon nanotubes (CNTs) and CNFs are 1D carbon materials that can be used to form

Si/carbon nanotube and nanofiber composite materials. CNTs are widely used in photonics,

optoelectronics, catalysis, and battery applications. Specifically, CNTs composited with Si

materials show great promise for use in Li-ion batteries due to several advantages, including

high electrical conductivity and impressive mechanical and thermal stabilities [51]. These

properties are crucial for satisfactory electrode performance during electrochemical

processing.

Wang et al. reported, for the first time, the synthesis of heterostructures comprising vertical

multiwall CNTs directly deposited by amorphous/crystalline Si nanodroplets, which were

produced using a simple two-step liquid injection CVD process (Figure. 4(a)) [52]. The space

between the multiwall CNTs facilitated homogenous deposition of the Si nanoparticles on the

CNT surfaces, and this prevention of Si nanoparticle agglomeration alleviated the strain

caused by the volume change of Si. Further, the hybrid Si/CNT nanostructure provided rapid

electron and ion transportation. The hybrid anode exhibited a high reversible capacity of

~2000 mAh/g and high Columbic efficiency over the first 25 discharge/charge cycles (Figure.

4(b)). Thereafter, the cycling tests revealed a rapid loss in performance due to fragmentation

of the attached Si particles, weakening of the Si-CNTs interface, and continuous SEI growth.

Thus, this hybrid structure prevented the agglomeration of Si particles and supported electron

and ion transportation but did not solve the pulverization and SEI issues. These issues remain

a huge challenge in the development of Si/1D carbon composite materials.


9
The Si pulverization and SEI issues may be solved by coating Si particles with a carbon shell

before compositing with CNTs. Su et al. designed 3D porous anode materials comprising Si

and SiO2 decorated in-situ with CNTs (Figure. 5(a)) [53]. SiO2 was heated with Mg to convert

SiO2 to Si clusters, leaving behind a small amount of SiO2. Fe was deposited on the surface of

the product (so called pSS particles) and served as a catalyst. Subsequently, the pSS powder

was loaded into a tube furnace under Ar/H2 atmosphere carrying toluene at 750 ℃. The

toluene interacted with the surface of the Fe particles and led to the in situ growth of CNTs.

This carbon source was in turn deposited on the surface of the Si particles. The pSS/CNT

composite anode performed better than a regular pSS anode (Figure. 5(b)). Electrochemical

impedance spectroscopy (EIS) indicated that the carbon layer and interconnected CNTs

contributed to the overall improvement in the conductivity of the pSS-based material (Figure.

5(c)). The pSS/CNT composite exhibited more stable long-term cycling performance than the

Si/CNT composite [52] due to the additional carbon coating on the Si particles. This carbon

coating acted as an electrolyte blocking layer and prevented continuous SEI formation, thus

facilitating high Coulombic efficiency [54]. Further, the layer alleviated the strain caused by

the volume change, which prevented the typical rapid cracking and pulverization of Si

particles. Yi et al. proposed a similar method for growing CNTs on porous Si [55], where Si

micro- and nanospheres were cross-linked by highly elastic carbon nanotubes. While CNTs

are often directly composited with Si materials, they can also be used as a soft membrane

shield for traditional polymer separators [56], thus providing additional space to buffer the

dramatic volume change of a Si anode.

Recently, Zhang et al. designed a robust hierarchical 3D Si/CNT composite with a void and

carbon shell as Li-ion battery anodes [57]. CNTs and commercial Si nanoparticles were

interwoven into a 3D structure, in which the Si nanoparticles were electrically interconnected

well by the CNTs. The extra carbon shell outside the Si/CNT composite particle effectively

prevented the electrolyte from contacting the Si nanoparticles and provided channels for
10
rapidly passing Li ions. In addition, the void inside the carbon shell adequately

accommodated the Si volume increase during the lithiation process to avoid damage. The 3D

Si/CNTs-composite exhibited excellent cycling performance, delivering 943 mAh/g after

1000 cycles at the rate of C/5, which was attributed to its special structure.

To take advantage of the inner space of the CNTs, Cui et al. used anodic aluminium oxide

(AAO) as a template to grow CNT arrays (AAO/CNTs) using a CVD method [58]. In

addition, they used liquid paraffin as an inexpensive carbon resource. Si nanoparticles were

then injected into the CNT arrays by the pressure difference, with the as-grown AAO/CNT

membrane serving as a barrier layer to separate the device into two parts, thus isolating the

silicon suspension from the air. Next, another carbon layer was deposited to completely wrap

the Si nanoparticles. Finally, a CNTs–Si–CNTs sandwich structure was obtained after etching

away the AAO template, and the resulting composite anode manifested a high specific

capacity of 1050 mAh/g after 500 cycles owing to its structural stability and the excellent

electrical properties of the double layer of CNTs sandwiching the Si nanoparticles.

CNFs were first prepared by Thomas Edison in 1879, where cotton and bamboo were

carbonized. The technique has been further developed and applied in both fundamental

scientific research and practical settings [59, 60]. The structure of a CNF is dependent on the

morphology of its precursor and the processing conditions. CNFs generally exhibit high

thermal and chemical stabilities, good thermal and electrical conductivities, and excellent

stress resistance, thereby leading to broad application prospects in composite materials, Li-ion

battery anode materials, nanoelectronic devices, and hydrogen storage materials. CNFs may

be synthesized using CVD [61], electrospinning, or solid-phase synthesis [62], where the vast

majority of CNFs used for Si/C composite anode materials are synthesized by electrospinning.

Thus, only electrospinning will be discussed herein.

Xu et al. reported 3D Si/C fibre paper electrodes for Li-ion batteries fabricated using a

combined electrospray/electrospinning technique [63]. The CNF electrospinning precursor


11
comprised polyacrylonitrile (PAN) dissolved in dimethylformamide (DMF) solution.

Electrospinning was used to produce PAN-coated Si nanoclusters interconnected by the PAN

nanofiber paper (Figure. 6(a)), which was carbonized to form carbon-coated Si clusters

between CNFs. The use of PAN as a carbon precursor led to a superior CNF structure [59].

The Si/C fibre paper acted as a free-standing electrode with superior electrochemical

performance and was a viable alternative current collector to replace Cu metal, offering light

weight, excellent electric conductivity, and good mechanical properties [64]. The cycling

performance of the Si/C fibre papers with various Si mass loadings were evaluated (Figure.

6(d)). The electrode exhibited stable cycling behaviour and high Coulombic efficiency, even

at the highest mass loading of 1.7 mg/cm2, thereby demonstrating the robustness of the 3D

Si/C paper architecture. An impressive capacity was achieved under different discharge and

charge rates, even at a high current density of 8 A/g (Figure. 6(b)). The electrochemistry was

evaluated on the basis of cyclic voltammetry (CV) curves (Figure. 6(c)) and was found to be

consistent with that reported previously [65]. A similar free-standing electrode composed of

multichannel carbon fibres confining Si nanoparticles (Si@MC-CNFs) was also recently

reported by Chen et al [66]. In this study, Si nanoparticles were directed inside the CNFs, and

the multichannel structure greatly benefitted the ion/electron transport.

While CNFs can act as a tight net to hold and electrically connect Si particles, there are many

other ways to composite Si materials with CNFs. These approaches include Si nanoparticles

anchored onto CNFs [67, 68], branched CNFs on Si nanowires [69], branched CNFs on Si

particles [70], and Si core-carbon shell fibres [71]. Scalable production of Si core-carbon shell

fibres by electrospinning with a dual nozzle was proposed by Hwang et al [71]. Si

nanoparticles mixed with PMMA were injected into the core channel of the dual nozzle, while

PAN dissolved in DMF was injected into the shell channel of the nozzle. Electrospinning and

carbonization were conducted to form the SiNPs@C core−shell 1D fibres (Figure. 7(a)).

SiNPs@C exhibited excellent cycling performance, where only 1% of the initial capacity was
12
lost after 300 cycles at a high current density of 3 C (2.75 A/g) (Figure. 7(b)). Further, the

Coulombic efficiency was often greater than 100%, indicating that a stable SEI was formed

on the SiNPs@C surface. The abnormally high efficiency was linked to continuous changes

in the arrangement of Si nanoparticles during cycling, which caused a fluctuation in the ionic

and electric conductivities. Therefore, some residual Li stored in previous cycles can be

extracted additionally. The EIS measurements further confirmed the presence of a stable SEI

(Figure. 7(c)). Testing under different current densities revealed that the composite anode

exhibited stable capacity, even at 12 C (11 A/g) (Figure. 7(d)). The remarkable performance

of the material was attributed to the excellent mechanical and electrochemical properties of

the carbon fibres and the porous structure of the Si nanoparticles.

Kong‘s group designed necklace-like Si@CNFs as robust anode materials by using SiO2

nanospheres as the Si source for electrospinning [72]. The SiO2 nanospheres were synthesized

by a Stöber process [73]. After electrospinning a Si@polymer network, the resulting

composite was treated by a magnesiothermic reduction reaction to synthesize the

necklace-like Si@CNFs. During this process, SiO2 nanospheres were reduced to smaller Si

nanoparticles, and the polymer wrapping the original SiO2 nanospheres became a CNF shell

encapsulating the Si nanoparticles. The diversity of carbon and silicon sources makes

Si/CNFs synthesized by electrospinning worth further exploration.

Most reports on Si/1D carbon composite materials involved the exposure of Si particles, with

and without a thin carbon layer coating, to the electrolyte [52, 67, 69, 74-77]. Consequently,

these electrodes suffer from varying degrees of capacity fading due to continuous SEI

formation, where the Si content of the composite material is consumed. Thus, an appropriate

carbon coating on the surface of Si particles is required to protect the material from the

electrolyte and enhance the composite electrode performance.

3.2. Si/2D Carbon Composite Materials


13
The most common form of 2D carbon is graphene, which is the thinnest and strongest known

material [78]. Graphene exhibits excellent electrical, mechanical, and thermal properties due

to its unique single atom thickness [79]. Graphene is widely used for energy storage,

especially in Li-ion batteries, Na-ion batteries, electrochemical capacitors, metal-air batteries,

and Li-S batteries [80]. The use of chemically doped graphene has attracted much research

interest, where a band gap is created by doping with elements such as boron and nitrogen to

produce more useful properties [81]. Consequently, doped graphene has been incorporated in

composite anode materials.

Graphene sheets in Si/2D graphene composite materials for Li-ion batteries are primarily

synthesized by Hummer‘s method owing to cost and processing considerations. A

modification of Hummer‘s method was proposed by Xin for production of a unique Si/G

nanocomposite with a novel 3D porous architecture, comparable to a regular Si@G nanosheet

[28]. Graphene oxide sheets were formed, and a modified sol-gel method was used to produce

SiO2@GO with tetraethyl orthosilicate (TEOS) as the SiO2 precursor. SiO2@GO was treated

with a magnesium thermal reduction method to convert the graphene oxide sheets to graphene

(G) sheets, thus forming the unique Si@G (Si/G) nanocomposite. To ensure a tight Si@G

composite structure, the graphene oxide solution was mixed, stirred, ultrasonicated, spray

dried, and annealed to form the 3D porous graphene coated Si@G (Figure. 8(a)). The Si/G

microspheres consistently delivered a capacity of up to 1000 mAh/g over 30 cycles, which

was much higher than that in the case of the Si@G nanosheets (Figure. 8(b)). This superior

electrochemical performance was attributed to the material‘s 3D porous architecture.

Additionally, the Si/G nanocomposite exhibited increasing capacity in the initial 15 cycles

due to reconstruction of a crystal structure near the Si surface, leading to an activation

phenomenon [64]. This was also confirmed by the CV measurements (Figure. 8(c)). The

Nyquist plots (Figure. 8(d)) illustrated that the charge transfer resistance of the Si/G

nanocomposite was much smaller than that in the Si@G nanosheets, which was attributed to
14
the 3D architecture and additional graphene. The Si particles in the simpler Si@G nanosheets

were exposed to the electrolyte, and the Si particles and graphene sheets were not strongly

bound, leading to the loss of Si particles. These issues led to rapid capacity fading, but were

resolved by the 3D architecture and extra graphene sheets in the Si/G nanocomposite.

The relatively simple Hummer‘s method can be creatively adapted to design uniquely

structured Si/G composites. A material comprising alternating strata of Si nanoparticles and

graphene was designed by Sun et al [29]. Graphene oxide was synthesized by oxidizing

graphite using Kovtyukhova‘s method [24], a well-known modification of Hummer‘s method.

The dried graphene oxide and Si nanoparticles were scaled, uniformly dispersed in deionized

water, and vacuum filtered. This filtering–transferring process was repeated several times to

prepare a multi-layered Si–GO structure. The composite was immersed in HBr and held at

120 ℃ for 2 h to reduce the graphene oxide and obtain the Si–re-G alternating strata

structured composite, where Si nanoparticles were sandwiched between graphene sheets

(Figure. 9(a)). The Si–re-G composite was cut into individual electrodes using a razorblade.

The CV curves of the composite (Figure. 9(b)) were in good agreement with those of

previously reported Si–C composites [82]. Excellent capacity retention of about 1500 mAh/g

was observed after 100 cycles (Figure. 9(c)), with no obvious capacity decline. This

remarkable performance was further demonstrated under testing at different current densities

(Figure. 9(d)). The performance of the Si–re-G alternating strata composite was far superior to

previously reported electrochemical behaviour of mixed Si–re-G anodes, which tended to

exhibit a continuous capacity decay [83, 84]. The alternating strata structure facilitated

electron movement between the Si particles and served as a flexible cushion to alleviate the

strain caused by volume change. Moreover, the strata structure maintained the integrity of the

Si nanoparticles by preventing agglomeration. A similar structure composed of

layer-by-layer-engineered silicon-based sandwich nanomats can be found in Zhou et al.‘s

recent work [85], wherein graphite microsheets acted as mats to sandwich Si nanoparticles.
15
In Si/G composites produced from graphene synthesized by Hummer‘s method, graphene

production is one of the simplest production steps. Further processing of graphene oxide and

Si particles to form a functional Si/G composite material is often more complex. To ensure

secure attachment between the Si particles and reduced graphene oxide, Si particles are often

functionalized before mixing with the graphene oxide, typically through organic

functionalization of the non-oxidized Si surfaces [86]. Si particles can be mixed with various

organic compounds, such as PDDA [82, 87], PVP [88], APTMS [89], and APS [90].

Graphene oxide typically contains hydroxyl (-OH) and carboxyl (-COOH) groups, and the

organic species on the surface of functionalized Si particles bind to the -OH and -COOH

groups by chemical bonding. Functionalized Si particles and reduced graphene oxide have

been found to form a strong bond to prevent the loss of Si particles, which improves the

performance of the composite electrode. The self-assembly Si and GO solution is generally

vaporized by spray drying or freeze drying [91-94], while graphene oxide is often reduced to

graphene by thermal reduction. Microwave irradiation can also be used for graphene oxide

reduction and carbonization [33].

The two applications based on Hummer‘s method that have been presented utilized

free-standing electrodes, where the composites were used directly as an anode material in

Li-ion batteries without a current collector. This ability was attributed to the tight bonding

achieved between the graphene oxide and Si particles, as well as the tough structure of the

reduced graphene oxide and good overall structural design of the composite. Si/G composite

materials have been reported in which Si particles were deposited onto graphene sheets and

bonded by covalent immobilization [33, 95]. However, the electrode comprising Si deposited

onto graphene sheets did not perform as well as the other free-standing electrodes. The raw

structure has a higher exposure of the Si particles to the electrolyte, which led to capacity

fading. Although 2D graphene provides a better overall electrode conductivity and better

distribution of the Si particles, protecting the Si particles from being consumed by the
16
electrolyte and preventing Si vacancies in the graphene sheets remains a challenge. Thus,

control over the weight ratio of Si to graphene, selection of an appropriately sized Si particle,

and careful design of the material‘s structure should be considered.

In contrast with the 2D graphene listed above, Lin et al. grew graphene nanowalls (GNWs) on

Ni foams using a radio-frequency (RF) plasma-enhanced horizontal tube furnace deposition

system [96]. Then, the Si material was deposited onto the as-grown GNWs to synthesize a

GNWs@Si composite. The tube furnace, GNWs, and GNWs@Si are shown schematically in

Figure. 10(a–c). The Raman spectra (Figure. 10(d)) indicated that the GNWs exhibited a

relatively low degree of graphitization and possessed abundant intergranular defects, which

could introduce a large number of charge carriers into the conductive network formed by the

2D GNWs. Owing to its stable structure, flexibility, and excellent conductivity without

additives or other complex fabrication processes, the GNWs@Si composite anode featured a

good specific capacity, as shown in Figure. 10(e).

In addition to graphene, MXenes are another popular 2D carbon material [97] that consist of

transition metal carbides, nitrides, or carbonitrides with a thickness of several atomic layers.

MXenes are produced by the selective etching of the A element from MAX phases, which are

metallically conductive, layered solids connected by strong metallic, ionic, and covalent

bonds, such as Ti2AlC, Ti3AlC2, and Ta4AlC3. These new 2D materials were named MXenes

to emphasize the loss of the A element from the MAX parent phase and to highlight their 2D

nature, which is similar to graphene. Their 2D morphologies, combined with their metallic

electrical conductivities, have considerably promoted their application in battery materials. In

batteries, MXenes can be used as current collectors [98], multifunctional binders [99, 100],

free-standing agents [101, 102] , and simple composite materials [103, 104].

In the typical MXene/Si@SiOx@C layer-by-layer superstructure designed by Guo‘s group

[103], layered Ti3C2Tx MXenes were formed by etching away Al layers from the Ti3AlC2

MAX phases using hydrofluoric acid (HF). TEOS was used as the silicon source to deposit
17
SiO2 on MXene, followed by magnesiothermic reduction. After pyrolyzing the PMMA

polymer on the surface of the Si nanoparticles using urea as a nitrogen source,

MXene/Si@SiOx@C was obtained. Figure. 11(a) depicts a schematic diagram of the synthesis

of the MXene/Si@SiOx@C layer-by-layer superstructure, which was further characterized by

scanning electron microscopy (SEM) and high-resolution transmission electron microscopy

(HRTEM) to verify the structure (Figure. 11(b-g)). Figure. 11(h) shows that

MXene/Si@SiOx@C exhibited stable and high cycling performance, which varied with the

different amounts of silicon. When the composite contained 72.8 wt% silicon, the

MXene/Si@SiOx@C-2 anode had a Coulombic efficiency of approximately 100% and

delivered a stable capacity of 1547 mAh/g after 200 cycles. This anode also exhibited a high

rate capability, as shown in Figure. 11(i). Its extraordinary performance was attributed to the

favourable electronic conductivity, rapid Li ion mobility, flexible layer spacing, and rich

covalent bonds provided by the MXene layers.

Because MXene has excellent properties similar to those of graphene, it would be worth

exploring; furthermore, the knowledge gained from the development of Si/graphene

composites could be used to design and synthesize related Si/MXene composite materials.

Such composite materials have already been well explored as free-standing,

current-collector-free, and binder-free anodes, and the results well met the expectations.

However, as MXene is generally synthesized by top-down methods, MXene has not yet been

coated (deposited) onto a silicon surface.

3.3. Si/3D Carbon Composite Materials

The main types of 3D carbon materials are graphene and amorphous carbon-coated Si

particles, porous carbon, and graphite. However, Si particles wrapped in graphene sheets are

not regarded as Si/3D carbon composite materials.

Carbon-coated Si particles are generally synthesized by CVD or thermal treatment, where a

Si/C composite with a core-shell (yolk-shell) structure is obtained. This carbon shell can
18
effectively prevent the degradation of the Si core by the electrolyte. Furthermore, this

yolk-shell structure reserves space to allow for Si volume expansion, thus ensuring that the Si

core‘s electrochemical activity and electrode‘s integrity are maintained.

A typical Si core-graphene shell structure was developed by Shi et al. by direct vertical

graphene (vG) growth on commercial SiO microparticles (d-SiO@vG) using CVD with

methane as the carbon source [39]. During graphene synthesis, the SiO microparticles were

simultaneously transformed to silicon and silica through a disproportionation reaction. An

amorphous SiO2 surface layer was formed during the initial heating of the reaction, thus

providing numerous catalytic nucleation sites for the synthesis of graphene. This layer also

prevented the formation of SiC. The unique vertical graphene coating led to a significant

improvement in electrical conductivity of the SiO on the single-particle level and electrode

level (Figure. 12(a)). Further, the structure provided transport channels for lithium ions. In

addition, the structure of the vertical graphene encapsulated SiO particles was maintained

after repeated lithiation and delithiation (Figure. 12(b)). The cathodic scan exhibited a peak at

0.65 V in the first cycle, which subsequently disappeared in the following cycles. This

observation was attributed to the irreversible formation of a stable SEI layer (Figure. 12(c)).

The stability of this SEI film was confirmed by the Nyquist plot based on EIS (Figure. 12(d)).

At a high areal mass loading of 1.5 mg/cm2, the d-SiO@vG electrode exhibited high specific

capacity with minimal fading over 100 cycles (Figure. 12(e)). Moreover, the d-SiO@vG

electrode exhibited superior capacity (~450 mAh/g) when compared to that of commercial

graphite at a high current density of 1600 mA/g (Figure. 12(f)). The stable rate performance

of the electrode further demonstrated the great advantages of vertically grown graphene.

A granadilla-like Si/C composite was proposed by Zhang et al., where clusters of yolk-shell

units were synthesized by a combination of a modified templating method and CVD [37].

Each yolk-shell unit comprised a Si particle and carbon shell separated by a void space. The

yolk-shell units were produced by coating Si particles with CaCO3 (CaCO3@Si) by


19
co-precipitation to provide a sacrificial layer and control void space formation. The

CaCO3@Si composite was placed into a tube furnace for CVD treatment with an acetylene

carbon source. The sacrificial CaCO3 layer disappeared to produce a Si core-carbon shell

cluster with interconnected void spaces. Typically, an SiO2 sacrificial layer is initially

produced by the hydrolysis of tetraethyl orthosilicate (TEOS) [44, 46] to form the void space.

However, SiO2 is expensive, complex to work with, and not environmentally friendly, while

the use of CaCO3 is less expensive, simpler, and less toxic. The granadilla-like Si composite

was assembled by the individual encapsulation of the Si nanoparticles in the CaCO 3

framework, carbon deposition, and removal of the carbonate templates using HCl (Figure.

13(a)). EIS plots of the granadilla-like Si composites with different Si contents were evaluated

(Figure. 13(b)). The 30@Si-granadilla composite had a high carbon content and exhibited the

lowest resistance to Li-ion transfer and highest conductivity. However, the high carbon

content led to a lower overall specific capacity of the electrode (Figure. 13(c)). The

50@Si-granadilla composite exhibited the highest cycling capacity, which was attributed to

its high Si content and optimal structure. All the granadilla-like Si composite samples

exhibited good cycling stability over 200 cycles, which was linked to a good rate performance

(Figure. 13(d)). The unique structure of the composite allowed for Si volume expansion and

avoided the loss of Si particles from the current collector after cracking. Further, the Si core

did not form a thick SEI. The well-connected carbon network maintained good electrode

integrity and provided great conductivity.

The novelty of some Si/3D carbon composites lies in the unique structure of the Si or carbon

components. Another granadilla-like porous Si/C structure proposed by Guan [105] was

synthesized by mixing SiO2 coated Si (SiO2/Si) with PVP to form a homogeneous suspension.

Carbonization and SiO2 removal by chemical etching were performed to produce a

granadilla-like porous Si/C composite. The porous structure of the Si microparticles can be

achieved with an appropriate precursor, such as SiO2 [106], zeolite SSZ-13 [107], or Mg2Si
20
[108]. Carbon coating is typically conducted to further enhance the electrical properties of the

composite and avoid continuous SEI formation.

In addition to carbon surface coatings, porous carbon could provide fast electron and ion

transport channels along with abundant void spaces. Generally, porous carbon frameworks are

derived from various organic carbon materials, such as pitch filtration and decomposition

[109], calcination of a mixture of glucose and calcium carbonate [110], gelation of multiple

organic species [111], and the etching and calcination of natural rape pollen [112].

Recently, a gigaporous carbon microsphere implanted with Si particles and encapsulated by a

carbon nanoshell was proposed by Li‘s group [113], and their preparation method is shown in

Figure. 14(a). Poly(styrene-co-vinylbenzyl chloride) (PSV), Si powder, and oleic acid were

mixed to form Si-embedded porous microsphere via emulsification. The porous microspheres

were then encapsulated by poly(urea–formaldehyde) (PUF) via in-situ polymerization. After

adding an extra carbon source and controlling the pH to complete the carbon encapsulation,

the as-formed product was carbonized to obtain the final carbon-encapsulated porous Si/C

microsphere. Figure. 14(b) shows the morphology of the porous microspheres before

implantation with Si, exhibiting abundant micropores. Figure. 14(c–d) illustrates the

microspheres implanted with micron-sized Si powder with increasing content. They explored

both micron-sized Si particles (u-Si) and nano-sized Si particles (n-Si). The

carbon-encapsulated porous Si/C microsphere anode far outperformed the bare Si anode, as

shown in Figure. 14(e-f). The superior retention of n-Si compared with that of u-Si should

correlate with the smaller volume variation of the smaller n-Si. In addition, the extra carbon

coating on the Si/C microsphere further maintained the integrity of the electrode. Owing to

the excellent porous carbon framework, the carbon-encapsulated n-Si/C microsphere

exhibited a high initial capacity of 3320 mAh/g and a good retention of 90% after 100 cycles.

In summary, porous carbon materials provide high electron and ion conductivity and void

21
space, all of which are vital to improving Si anodes, and the extra carbon coating on Si

particles or Si/C composites further decreases the influence of the SEI.

Compared with the above 3D carbon materials, Si/graphite composites are more attractive

because they have been commercialized. Future applications of graphite anodes are limited by

their low specific capacity, and the severe volume change and SEI issues associated with Si

anodes also hinder their commercialization. Thus, the Si/graphite composite has been deemed

to be the most appropriate approach for realizing a high energy density in current commercial

LIB (lithium-ion battery) systems with a Si anode. To commercialize high-energy-density

Si/graphite composite anodes for next-generation LIB, in addition to solving the problems

associated with Si anode, other aspects to consider are ensuring tight contact between the

silicon and graphite and finding the best weight ratio of silicon and graphite. Very recently,

Cho et al. comprehensively reviewed the research progress in integrated graphite and silicon

anodes and reviewed their prospects for their commercialization in high-energy Li-ion

batteries [114]. Therefore, we will not further review Si/graphite composites herein.

Compared to the common conductive agent carbon black, whose contact with Si is

point-to-point, 1D, 2D and 3D carbon materials possess better electrical conductivity over

electrical contact with Si materials because of increased contact sites. 1D carbon materials can

form conductive networks that improve electron and Li-ion transport pathways. Some 1D

carbon materials release stress caused by volume change and decrease weak electrical contact

between Si nanoparticles by wrapping Si nanoparticles. 2D graphene sheets show better

electron conductivity and higher thermal and mechanical stability than other carbon materials.

Thus, Si electrodes composited with 2D graphene sheets have increased electrochemical

stability, and in some specially designed electrodes, graphene sheets can work as the current

collector. This facilitates more active material within the limited cell space. Compared to 1D

carbon materials, 2D graphene sheets provide more contact sites because of extremely high

specific surface. Although the contact between Si materials and 1D/2D carbon materials is
22
relatively intimate, poor structural design of Si/1D carbon or Si/2D carbon composites

continues to allow exposure of Si particles to electrolyte. The resulting Si particle loss and

continuous SEI formation consumes the outer layer of Si particles and leads to capacity fading.

3D graphene and carbon-coated on Si surface are designed to protect Si particles from wetting

by electrolyte. During electrochemical processes, electrolyte forms a stable SEI film with the

graphene/carbon shell instead of consuming Si content. Furthermore, the tough

graphene/carbon shell alleviates the volume expansion of Si so that Si particles will not be

pulverized. The unique yolk-shell structure of Si/3D graphene/carbon composites could even

reserve proper space for Si particle expansion and contraction instead of departure from the

electrode and subsequent loss of electrochemical activity. Apparently, 3D graphene or carbon

coating is quite effective in protecting Si anodes and achieving high battery performance.

However, the cost to prepare 3D graphene/carbon-coated Si is usually higher, and the

preparation is harder to control for optimization. To easily compare the performance of

different Si/C composites anodes, we tabulated the capacities of the various Si/C composite

anodes mentioned above, along with their structures and synthetic methods, in Table 1.

Table 1. Electrochemical performance and synthesis methods of Si/C composite anodes

materials in LIBs.

Initial Mass
Cycling Stability
Anodes Capacity Initial CE Loading Structure Synthetic Method Ref.
mAh/g
mAh/g mg/cm2

~1000 after Silicon/C two-step liquid

Si/CNTs 2552 80.3% 100cycles at 0.2 — nanotube injection CVD [52]

A/g heterostructures process

943 after 1000 hierarchical 3D spray drying and


Si/CNTs-C — 56% 0.9 [57]
cycles at 0.84 A/g Si/CNTs thermal treatment

1050 after 500 tubular sandwich


Si/CNTs 2097 63% 2.0 CVD [58]
cycles at 1 A/g structure

1178 after 100 3D Si/C fiber electrospray and


Si/CNFs 1996 80% 1.3 [63]
cycles at 0.5 A/g paper electrospinning

Si/CNFs ~1400 87.5% ~1250 after 100 0.6 core−shell fibers electrospinning [71]

23
cycles at 0.24 A/g

Stöber method,

710 after 200 necklace-like electrospinning,


Si/CNFs ~3000 ~87% 1.0 [72]
cycles at 0.5 Ag Si@C nanofibers magnesiothermic

reduction

Kovtyukhova‘s
1500 after 100 alternating
Si/graphene over 5000 47% 2.7 method, filtering– [29]
cycles at 1.35 Ag stratum
transferring

GNWs@Si
1116 after 200
Si/graphene 3798 48.3% — nanowall PECVD [96]
cycles
network

1040 after 150 3D conductive HF etching,


Si@C/Mxene 2276 73% 1.2 [99]
cycles at 0.42 A/g framework vacuum filtration

~1500 after 210 free-standing HF etching,


Si/Mxene ~3500 71% 1.1−1.3 [101]
cycles at 1 A/g paper vacuum filtration

ball milling, HF
1547 after 200
Si/Mxene ~1800 — ~1 layer-by-layer etching, Stöber [103]
cycles at 0.84 A/g
method

866 after 100 vertical graphene


Si@graphene ~2100 ~95% 1.5 CVD [39]
cycles at 0.32 A/g coating

1080 after 200 granadilla-like coprecipitation


Si@a-C ~1643 80% — [37]
cycles at 0.25 A/g C@void@Si process, CVD

1160 after 1000 pomegranate like Stöber method,


Si@a-C 2350 82% 0.2 [44]
cycles at 2.1 A/g yolk-shell CVD

hollow porous surface


1043 after 200
Si/p-C 1960 79.4% 0.85 carbon functionalization, [112]
cycles at 0.4 A/g
encapsulated Si thermal treatment

emulsification,

2985 after 100 in-situ


Si/p-C 3320 81% <1 porous carbon [113]
cycles at 0.42 A/g polymerization,

thermal treatment

reflux technique,
~500 after 50 Si coated
catalytic
Si/graphite 525 93.8% cycles at 1.75 — edge-activated [115]
hydrogenation,
mA/cm graphite
CVD

high-energy

Si/graphite 711 90.5% — — Si/C microsphere ball-milling [116]

method

4. Advanced Characterization Techniques


24
Electrode materials may be evaluated based on macroscopic performance testing and

microscopic composition and structure characterization, where the macroscopic properties

and performance of a material are highly dependent on its microstructure. Long-term cycling

performance, rate performance, cyclic voltammetry measurements, and EIS are common

macroscopic performance tests used in the study of Li-ion batteries. These parameters provide

a good indication of the material‘s electrochemical performance during charging and

discharging. Further characterization is required to better understand the material‘s elemental

composition and structure, including thermal gravimetric analysis (TGA), Brunauer–Emmett–

Teller specific surface area measurements (BET), Fourier-transform infrared spectroscopy

(FTIR), Raman spectroscopy, X-ray diffraction spectroscopy (XRD), X-ray photoelectron

spectroscopy (XPS), scanning electron microscopy (SEM), transmission electron microscopy

(TEM), cryo-electron microscopy (cryo-EM).

As a representative ex-situ characterization technique, cryo-EM is an ultralow-temperature

sample preparation and transportation technique used for electron microscopy, which can be

used to directly observe liquid, semi-liquid, and electron-beam-sensitive samples. Huang et al.

visualized the structure and chemistry of the SEI on silicon anodes using atomic-resolution

cryogenic (scanning) transmission electron microscopy (cryo-(S)TEM) and electron energy

loss spectroscopy (EELS) [117]. In a standard EC (ethylene carbonate) /DEC (diethyl

carbonate) electrolyte, the lithiated Si (Li15Si4) was interfaced with a bilayer SEI, consisting

of an inner amorphous LixSiOy layer from the lithiation of the SiOx and an outer layer

consisting of amorphous LEDC (Li ethylene dicarbonate) and crystalline Li2O arising from

the decomposition of EC. Upon delithiation, Li2O first reacted with Si to form LixSiOy, which

could be further delithiated, and organic carbonates such as LEDC were oxidized. In addition

to observing the SEI, cryo-EM can also be used to investigate other sensitive materials such

as lithium dendrites [118]. Thus, cryo-EM can be expected to promote the understanding of

electrochemical processes.
25
However, ex-situ characterization does not provide insight into the electrochemical process.

Owing to the widespread use, convenience, and ease of use of ex-situ characterization

techniques, these are not detailed in this review. Recent advances in in-situ characterization

have allowed for the real-time observation of structural changes and compound evolution

during electrochemical processing. Compared to ex-situ characterization techniques, in-situ

analysis can provide more in-depth information.

4.1. In-situ TEM

TEM provides atomic-level resolution and has become increasingly popular in material

characterization. In-situ TEM has made great advances, thereby facilitating the real-time

observation of a wide variety of details during the electrochemical process. This is a distinct

advantage over traditional ex-situ TEM.

A SiC-free graphene-coated Si composite material was designed by Son et al., and the

lithiation and volume expansion of the graphene-coated Si particles was monitored in real

time using in-situ TEM analysis (Figure. 15(a)) [119]. The schematic of the in situ TEM

device is shown in Figure. 15(b). The graphene-coated Si particles were placed onto a

gold-fixed electrode and lithiated using an electrode with a Li/LiO2 tip, which caused the

particles to swell. The particles fully encapsulated by graphene with no obvious defects (red

and orange circles in Figure. 15(a)) exhibited a diameter increase of ~30%. Particles with a

defective graphene coating exhibited pulverization of the inner Si particle and rupture through

the defect (green and blue circles in Figure. 15(a)). The interlayer distance of graphene layers

increased by the end of lithiation due to Li-ion intercalation [120]. Liu et al. also applied

in-situ TEM to study the structural change of pomegranate-inspired yolk-shell clusters of Si/C

composite [44]. The in-situ TEM data shows that initial well-defined void spaces and chosen

nanoparticles can ensure Si particles expand inside the carbon shell with little change to both

carbon shell and Si yolk.


26
In-situ TEM was used by Liu et al. to reveal that individual Si nanoparticles below a critical

size of ~150 nm neither cracked nor fractured after the first lithiation process, while larger

particles exhibited surface cracking and particle fracture [121]. This novel finding provided

important insight for the selection of Si nanoparticles in the design of Si/C composite anode

materials.

4.2. In-situ XRD

In-situ XRD measurements rely on the design of the in-situ cell, as a poorly designed cell can

give rise to signal artefacts that interfere with the signals of interest. Several different types of

in-situ cells have been reported for in-situ XRD, such as modified coin cells and modified

pouch cells. An in situ/operando electrochemical coin cell designed for synchrotron-based

X-ray characterization is shown in Figure. 16(a). Further, cells used for synchrotron-based

in-situ XRD should be designed as simple as possible to allow easy assembly and disassembly

in the synchrotron facility [122]. One should also be able to incorporate the cell into the

beamline setup and accommodate the X-ray optics. The cell should be highly reproducible

and the cell components and materials must be carefully selected and constructed to avoid

interference from inactive cell components. In addition to external noise, beam damage

should also be considered. Intermittent probing of the sample can reduce the total amount of

X-ray radiation during acquisition. While performing this technique can be complex, in-situ

XRD is a powerful tool to study crystal, structural, and phase transformations within electrode

materials during the electrochemical process.

Synchrotron-based in-situ XRD was used by Yang to study the structural changes in

graphite-mixed carbon-coated Si used as lithium intercalation materials [123].

Synchrotron-based X-ray sources provide higher intensities and greater photon energies,

which facilitate higher penetration power and better signal-to-noise ratios for favourable

in-situ capabilities [124]. However, battery materials are easily damaged by the high-energy
27
beam and analysis should be conducted with care [122]. The in-situ XRD patterns of a

graphite-mixed carbon-coated Si electrode during the first discharge exhibited significant

peak broadening and intensity reduction of the (111), (220), and (311) reflexes (Figure. 16(b))

[123]. This indicated that the lithium intercalation of the Si powder involved the conversion of

the Si crystals to either the amorphous state or Li–Si alloy nanocrystals. Moreover, the

changes in graphite‘s (002) reflection indicated that lithium first intercalated into graphite,

and then into Si. In-situ XRD characterization of a bare Si electrode for comparison revealed

that the Li–Si alloy formed in the carbon-coated composite was distributed more uniformly

with a lower concentration of lithium. This led to higher utilization of the active Si materials

and better cycling performance of the carbon-coated Si particles.

4.3. In-situ Raman

Raman spectroscopy is based on the Raman scattering effect discovered by Indian scientist

C.V. Raman. The scattering spectra are measured at frequencies differing from the incident

light to study molecular vibration and rotation, thus characterizing molecular structure.

Inelastic scattering caused by the interaction of molecular vibrations, optical phonons in solids,

and laser excitation is referred to as Raman scattering. Many structural, mechanical, and

chemical changes within electrodes during electrochemical cycling can be studied by in-situ

Raman spectroscopy [125, 126]. Raman spectroscopy does not require long-range structural

ordering, and amorphous compounds or electrode materials with poor crystallinity can be

studied, which is often not possible with XRD [124].

Earlier in-situ Raman studies were performed on a composite electrode comprising

microcrystalline Si dispersed homogeneously on a carbon matrix (Figure. 17) [127]. The

Raman spectra provided direct evidence of Si particle lithiation during the first discharge,

where the Si signal completely disappeared as the voltage dropped below 0.1 V as Li 3.5Si was

formed [128]. The gradual loss of Raman Signal intensity below 0.4 V was attributed to
28
multiple factors, including the skin depth effect and initial insertion of lithium on the Si

surface. The inset in Figure 10c shows the Raman spectra of the Si/C composite during the

first charge cycle, where peaks attributed to crystalline structures related to the formation of

Li–Si alloy phase were not observed. These observations revealed that the electrode

comprised mixed phases (amorphous Si and Li–Si alloy) after the first cycle.

4.4. In-situ AFM

Atomic force microscopy (AFM) uses a microcantilever to observe and magnify the force

between the tip probe on the cantilever and the atoms of the sample to achieve atomic-scale

resolution. Owing to the special working principle, AFM can observe both conductor and

non-conductor, which makes up for the limitations of scanning electron microscopy.

Recently, the development of in situ AFM has promoted studies of Si-based materials,

especially the analysis of volume evolution. Breitung et al [129]. applied in situ AFM to study

both the macroscopic (volume change) and microscopic (cracks and pinholes) mechanical

degradation that occurs in individual Si particles (Figure. 18(a)). AFM revealed that during

the first lithiation cycle, significant structural changes caused capacity loss (Figure. 18(b));

later, the capacity did not further fade, and coulombic efficiency was well retained, which

accords with the observation that nanoparticles can withstand stress induced cracking (Figure.

18(c)). Additionally, in-situ AFM was applied to analyse both lateral and vertical expansion of

Si [130], detect the thin SEI film during electrochemical cycling [131], and investigate

nanomechanical properties of Si-based anodes [132].

4.5. In-situ NMR

Nuclear magnetic resonance (NMR) is a physical process in which a nucleus with non-zero

magnetic moment is affected by the application of an external magnetic field. Application of a

magnetic field populates a low-energy spin state of the atomic nucleus; when a second
29
magnetic field is applied, nuclei in a low-energy spin state absorbed energy and are converted

to a high-energy spin state. Net energy absorption is called resonance. Nuclear resonance is

dependent on the molecular environment of the molecule; therefore, NMR spectroscopy

provides structural information. Absorption is electronically detected and plotted as resonance

frequency vs. Energy absorbed. Schematics of the a typical electrochemical cell and the set-up

for the in situ NMR measurements is shown in Figure. 19 [133]. In Si-based anode materials,

NMR can reveal the lithiation/delithiation behaviors of Si which involve crystalline to

amorphous phase transitions. Ogata et al. utilized in situ 7Li NMR spectroscopy to investigate

the second and subsequent lithiation behaviors of amorphous Si in SiNWs [134]. The

electrochemical and NMR sprectra of the process revealed the over-lithiation of c-Li3.75Si that

occurs below 50 mV and the formation of small clusters within the Li3.75Si phase on charge

which was only formed during the second electrochemical cycle. The defects formed within

this phase can act as nuclei for the growth of larger Si clusters, which can then grow and form

crystalline Si.

Besides the mentioned in-situ characterization techniques, in situ X-ray and neutron

radiography was combined to detect the Li–Si distribution and the alloying-induced volume

changes. It was visualized that many Si particles did not undergo lithiation during the initial

lithiation inside Si-based LIBs [135]. In-situ neutron reflectometry was also used to evaluate

SEI formed on silicon electrodes, indicating that the contracted by up to 22% during lithiation

and swelled during delithiation, slowly approaching an equilibrium thickness [136]. To

further understand the chemical composition of the SEI on silicon electrodes, it was

investigated using in-situ attenuated total reflection–Fourier-transform infrared (ATR-FTIR)

spectroscopy with an adjustable penetration depth [137]. This investigation revealed that

effects of the SEI on the capacity degradation include both chemical degradation (i.e. a

chemical reaction between electrolyte and electrode surface) during initial cycling and

chemical-mechanical degradation (i.e. from the repeated SEI formation and particle fracture)
30
after prolonged cycling. Meanwhile, to apply in-situ XPS to observe the lithiation and

delithiation reactions of Si anode, Endo et al. deposited a Si film on a solid electrolyte, which

could compensate for the shortcomings of XPS [138]. The observed phase transformations

were consistent with those in previous reports [139, 140]. In summary, the various in-situ

characterization techniques are powerful tools for studying the structural and chemical

evolution of materials in real time.

Advanced characterization techniques are powerful tools to understand the macroscopic

properties of Si/C composite materials when applied to batteries or other fields. The

complementary combination of ex-situ and in-situ characterization techniques can provide

detailed information before and during electrochemical processes. A sufficient understanding

of a material‘s properties and its chemical and structural changes can guide the practical

applications of battery materials for superior performance.

5. Conclusion and Perspective

A wide range of strategies can be applied to synthesize Si/C composite materials. CVD and

electrospinning methods are typically used to produce 1D carbon nanofiber and carbon

nanotubes; Hummer‘s method is predominantly used to produce 2D graphene sheets from

graphite, and CVD or thermal treatment are used to coat 3D carbon on the surface of Si.

Neither Si nor carbon electrodes alone meet the requirements of next generation commercial

batteries, but the Si content can be optimized to enhance the overall capacity of most Si/C

composite electrode materials, while carbon content stabilizes the performance of Si. These

properties can be combined to produce efficient commercial battery materials with stable and

high-performance capabilities.

In-situ characterization techniques present unparalleled advantages for the real-time

observation of structural, mechanical, and chemical changes during electrochemical

processing, thus providing more insight than ex-situ characterization. However, in-situ
31
characterization is associated with higher costs than ex-situ characterization techniques.

Moreover, there is an increased risk of equipment and sample damage during operation.

Many types of carbon precursors can be used to produce Si/C composite materials. Some of

these chemicals are harmful to the environment and human body, including methane and

toluene. Thus, the use of a non-toxic carbon source is encouraged. The quality of the

synthesized carbon material is largely dependent on the carbon source, and is thus another

important consideration. Among the wide variety of carbon sources for carbon material

synthesis, non-toxic, low-cost, and high quality carbon precursors must be further developed.

Contact between the Si particles and carbon materials should be optimized as the capacity of

the Si/C composite electrode is mainly attributed to its Si component. Stronger physical

contact facilitates electrochemical contact, which must be preserved during repeated

electrochemical cycling to ensure that the conductivity of the electrode is maintained.

The formation of void spaces, namely, pores in the electrode and voids within the structure, is

unavoidable in these electrode materials. The controlled design of void spaces can provide

room to accommodate the stress attributed to Si volume expansion and shrinkage. Thus, the

loss of electrode materials from the electrode is minimized. Furthermore, proper void space

ensures optimal contribution of active materials and, in turn, stable high performance. SEI can

accumulate inside pores upon cycling if the pore size is too large, subsequently inhibiting

ionic percolation. Thus, the pore size and number should be carefully controlled.

The total volume capacity of a Si/C electrode may be improved by increasing the Si/C weight

ratio, but one must ensure that other changes do not affect the Si performance. The

performance of the electrode must be maximized and stabilized for commercial application,

and the Si/C ratio should be optimized based on different carbon materials and structure

designs.

32
Declaration of interests
☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.
☐The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:

References

[1] J.-M. Tarascon, M. Armand, Issues and challenges facing rechargeable lithium batteries,
Nature 414 (2001) 359-367.
[2] B. Dunn, H. Kamath, J.-M. Tarascon, Electrical energy storage for the grid: a battery of
choices, Science 334 (2011) 928-935.
[3] H. Wu, Y. Cui, Designing nanostructured Si anodes for high energy lithium ion batteries,
Nano Today 7 (2012) 414-429.
[4] M. N. Obrovac, V. L. Chevrier, Alloy negative electrodes for Li-ion batteries, Chem.
Rev. 114 (2014) 11444-11502.
[5] Y. Jin, B. Zhu, Z. Lu, N. Liu, J. Zhu, Challenges and recent progress in the development
of Si anodes for lithium-ion battery, Adv. Energy Mater. 7 (2017) 1700715.
[6] J. Wu, X. Qin, C. Miao, Y. He, G. Liang, D. Zhou, M. Liu, C. Han, B. Li, F. Kang, A
honeycomb-cobweb inspired hierarchical coreeshell structure design for electrospun
silicon/carbon fibers as lithium-ion battery anodes, Carbon 98 (2016) 582-591.
[7] G. Hou, B. Cheng, Y. Cao, M. Yao, B. Li, C. Zhang, Q. Weng, X. Wang, Y. Bando, D.
Golberg, F. Yuan, Scalable production of 3D plum-pudding-like Si/C spheres: towards
practical application in Li-ion batteries, Nano Energy 24 (2016) 111-120.
[8] Q. Yun, X. Qin, W. Lv, Y. He, B. Li, F. Kang, Q. Yang, ‗‗Concrete‘‘ inspired
construction of a silicon/carbon hybrid electrode for high performance lithium ion
battery, Carbon 93 (2015) 59-67.
[9] F. Dou, L. Shi, G. Chen,D. Zhang, Silicon/carbon composite anode materials for
lithium‑ion batteries, Electrochem. Energy Rev. 2 (2019) 149-198.
[10] M. Zhang, T. Zhang, Y. Ma, Y. Chen, Latest development of nanostructured Si/C
materials for lithium anode studies and applications, Energy Storage Mater. 4 (2016)
1-14.
[11] X. Shen, Z. Tian, R. Fan, L. Shao, D. Zhang, G. Cao, L. Kou, Y. Bai, Research progress
on silicon/carbon composite anode materials for lithium-ion battery, J. Energy Chem. 27
(2018) 1067–1090.

33
[12] C. K. Chan, H. Peng, G. Liu, K. McIlwrath, X. Zhang, R. A. Huggins, Y. Cui,
High-performance lithium battery anodes using silicon nanowires, Nat. Nanotechnol. 3
(2008) 31-35.
[13] C. K. Chan, R. N. Patel, M. J. O‘Connell, B. A. Korgel, Y. Cui, Solution-grown Silicon
nanowires for lithium-ion battery anodes, ACS Nano 4 (2010) 1443-1450.
[14] Y. Yao, M. T. McDowell, I. Ryu, H. Wu, N. Liu, L. Hu, W. D. Nix, Y. Cui,
Interconnected silicon hollow nanospheres for lithium-ion battery anodes with long cycle
life, Nano Lett. 11 (2011) 2949-2954.
[15] H. Kim, M. Seo, M.-H. Park, J. Cho, A critical size of silicon nano-anodes for lithium
rechargeable batteries, Angew. Chem. Int. Ed. 122 (2010) 2192-2195.
[16] M.T. Demirkan, L. Trahey, T. Karabacak, Cycling performance of density modulated
multilayer silicon thin film anodes in Li-ion batteries, J. Power Sources 273 (2015)
52-61.
[17] F. Dou, L. Shi, P. Song, G. Chen, J. An, H. Liu, D. Zhang, Design of orderly carbon
coatings for SiO anodes promoted by TiO2 toward high performance lithium-ion battery,
Chem. Eng. J. 338 (2018) 488–495.
[18] F. Dou, Y. Weng, G. Chen, L. Shi, H. Liu, D. Zhang, Volume expansion restriction
effects of thick TiO2/C hybrid coatings on micro-sized SiOx anode materials, Chem. Eng.
J. 387 (2020) 124106.
[19] J. Tu, Y. Yuan, P. Zhan, H. Jiao, X. Wang, H. Zhu, S. Jiao, Straightforward approach
toward SiO2 nanospheres and their superior lithium storage performance, J. Phys. Chem.
C 118 (2014) 7357−7362.
[20] L. Zhang, X. Liu, Q. Zhao, S. Dou, H. Liu, Y. Huang, X. Hu, Si-containing precursors
for Si-based anode materials of li-ion batteries: a review, Energy Storage Mater. 4 (2016)
92-102.
[21] J. Cho, Porous Si anode materials for lithium rechargeable batteries, J. Mater. Chem. 20
(2010) 4009-4014.
[22] M. Inagaki, Y. Yang, F. Kang, Carbon nanofibers prepared via electrospinning, Adv.
Mater. 24 (2012) 2547-2566.
[23] JR. W. S. Hummers, R. E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc.
80 (1958) 1339.
[24] N. I. Kovtyukhova, P. J. Ollivier, B. R. Martin, T. E. Mallouk, S. A. Chizhik, E. V.
Buzaneva, A. D. Gorchinskiy, Layer-by-layer assembly of ultrathin composite films

34
from micron-sized graphite oxide sheets and polycations, Chem. Mater. 11 (1999)
771-778.
[25] D. P. Wong, R. Suriyaprabha, R. Yuvakumar, V. Rajendran, Y.-T. Chen, B.-J. Hwang,
L.-C. Chen, K.-H. Chen, Binder-free rice husk-based silicon–graphene composite as
energy efficient Li-ion battery anodes, J. Mater. Chem. A 2 (2014) 13437-13441.
[26] M. Ko, S. Chae, S. Jeong, P. Oh, J. Cho, Elastic a‑Silicon nanoparticle backboned
graphene hybrid as a self-compacting anode for high-rate lithium ion batteries, ACS
Nano 8 (2014) 8591-8599.
[27] J. Ren, Q. Wu, G. Hong, W. Zhang, H. Wu, K. Amine, J. Yang, S.-T. Lee, Silicon–
graphene composite anodes for high-energy lithium batteries, Energy Technol. 1 (2013)
77-84.
[28] X. Xin, X. Zhou, F. Wang, X. Yao, X. Xu, Y. Zhu, Z. Liu, A 3D porous architecture of
Si/graphene nanocomposite as high-performance anode materials for Li-ion batteries, J.
Mater. Chem. 22 (2012) 7724-7730.
[29] F. Sun, K. Huang, X. Qi, T. Gao, Y. Liu, X. Zou, X. Wei, J. Zhong, A rationally
designed composite of alternating strata of Si nanoparticles and graphene: a
high-performance lithium-ion battery anode, Nanoscale 5 (2013) 8586-8592.
[30] X. Gao, J. Li, Y. Xie, D. Guan, C. Yuan, A multilayered silicon-reduced graphene oxide
electrode for high performance lithium-ion batteries, ACS Appl. Mater. Interfaces 7
(2015) 7855-7862.
[31] H. Jiang, X. Zhou, G. Liu, Y. Zhou, H. Ye, Y. Liu, K. Han, Free-standing Si/graphene
paper using Si nanoparticles synthesized by acid-etching Al-Si alloy powder for
high-stability Li-ion battery anodes, Electrochim. Acta 188 (2016) 777-784.
[32] A. M. Abdelkader, A. J. Cooper, R. A. W. Dryfe, I. A. Kinloch, How to get between the
sheets: A review of recent works on the electrochemical exfoliation of graphene
materials from bulk graphite, Nanoscale 7 (2015) 6944-6956.
[33] J. M. Kim, D. Ko, J. Oh, J. Lee, T. Hwang, Y. Jeon, W. H. Antinka, Y. Piao,
Electrochemically exfoliated graphene as a novel microwave susceptor: the ultrafast
microwave-assisted synthesis of carbon-coated silicon−graphene film as a lithium-ion
battery anode, Nanoscale 9 (2017) 15582-15590.
[34] V. Singh, D. Joung, L. Zhai, S. Das, S. I. Khondaker, S. Seal, Graphene based materials:
past, present and future, Prog. Mater Sci. 56 (2011) 1178-1271.
[35] Z. Sun, Z. Yan, Jun Yao, E. Beitler, Y. Zhu, J. M. Tour, Growth of graphene from solid
carbon sources, Nature 468 (2010) 549-552.

35
[36] A. Magasinski, P. Dixon, B. Hertzberg, A. Kvit, J. Ayala, G. Yushin, High-performance
lithium-ion anodes using a hierarchical bottom-up approach, Nat. Mater. 9 (2010)
353-358.
[37] L. Zhang, R. Rajagopalan, H. Guo, X. Hu, S. Dou, H. Liu, A green and facile way to
prepare granadilla-like silicon-based anode materials for Li-ion batteries, Adv. Funct.
Mater. 26 (2016) 440-446.
[38] M. Ko, S. Chae, J. Ma, N. Kim, H.-W. Lee, Y. Cui, J. Cho, Scalable synthesis of
silicon-nanolayer-embedded graphite for high-energy lithium-ion batteries, Nat. Energy
1 (2016) 1-8.
[39] L. Shi, C. Pang, S. Chen, M. Wang, K. Wang, Z. Tan, P. Gao, J. Ren, Y. Huang, H. Peng,
Z. Liu, Vertical graphene growth on SiO microparticles for stable lithium ion battery
anodes, Nano Lett. 17 (2017) 3681-3687.
[40] C. Gao, H. Zhao, P. Lv, T. Zhang, Q. Xia, J. Wang, Engineered Si sandwich electrode: Si
nanoparticles/graphite sheet hybrid on Ni foam for next-generation high-performance
lithium ion batteries, ACS Appl. Mater. Interfaces 7 (2015) 1693-1698.
[41] Y. Zhang, N. Du, Y. Chen, Y. Lin, J. Jiang, Y. He, Y. Lei, D. Yang, Carbon dioxide as a
green carbon source for the synthesis of carbon cages encapsulating porous silicon as
high performance lithium-ion battery anodes, Nanoscale 10 (2018) 5626-5633.
[42] J. Zhou, S. Jiang, Y. Li, Z. Pan, Y. Qian, Y. Zhao, N. Lin, Y. Qian, Chemical fixation of
CO2 on activated Si: producing graphitic carbon-stabilized Si particles for Li-storage,
Energy Storage Mater. 31 (2020) 36–43.
[43] D. Sui, Y. Xie, W. Zhao, H. Zhang, Y. Zhou, X. Qin, Y. Ma, Y. Yang, Y. Chen, A
high-performance ternary Si composite anode material with crystal graphite core and
amorphous carbon shell, J. Power Sources 384 (2018) 328-333.

[44] N. Liu, Z. Lu, J. Zhao, M. T. McDowell, H.-W. Lee, W. Zhao, Y. Cui, A


pomegranate-inspired nanoscale design for large-volume-change lithium battery anodes,
Nat. Nanotechnol. 9 (2014) 187-192.

[45] C. Shen, X. Fang, M. Ge, A. Zhang, Y. Liu, Y. Ma, M. Mecklenburg, X. Nie, C. Zhou,
Hierarchical carbon-coated ball-milled silicon: synthesis and applications in
free-standing electrodes and high-voltage full lithium-ion batteries, ACS Nano 12 (2018)
6280-6291.
[46] N. Liu, H. Wu, M. T. McDowell, Y. Yao, C. Wang, Y. Cui, A yolk-shell design for
stabilized and scalable Li-ion battery alloy anodes, Nano Lett. 12 (2012) 3315-3321.

36
[47] X. Zhuang, P. Song, G. Chen, L. Shi, Y. Wu, X. Tao, H. Liu, D. Zhang, Coralloid-like
nanostructured c‑nSi/SiOx@Cy anodes for high performance lithium ion battery, ACS
Appl. Mater. Interfaces 9 (2017) 28464−28472.

[48] J. Han, G. Chen, T. Yan, H. Liu, L. Shi, Z. An, J. Zhang,D. Zhang, Creating
graphene-like carbon layers on SiO anodes via a layer-by-layer strategy for lithium-ion
battery, Chem. Eng. J. 347 (2018) 273–279.
[49] H. Su, A. A. Barragan, L. Geng, D. Long, L. Ling, K. N. Bozhilov, L. Mangolini, J. Guo,
Colloidal synthesis of silicon@carbon composite materials for lithium-ion batteries,
Angew. Chem. Int. Ed. 129 (2017) 10920-10925.
[50] M.-G. Jeong, H. L. Du, M. Islam, J. K. Lee, Y.-K. Sun, H.-G. Jung, Self-rearrangement
of silicon nanoparticles embedded in micro-carbon sphere framework for high-energy
and long-life lithium ion batteries, Nano Lett. 17 (2017) 5600-5606.
[51] Y. Yan, J. Miao, Z. Yang, F. Xiao, H. Yang, B. Liu, Y. Yang, Carbon nanotube catalysts:
recent advances in synthesis, characterization and applications, Chem. Soc. Rev. 44
(2015) 3295-3346.
[52] W. Wang, P. N. Kumta, Nanostructured hybrid silicon/carbon nanotube heterostructures:
reversible high-capacity lithium-ion anodes, ACS Nano 4 (2010) 2233-2241.
[53] J. Su, J. Zhao, L. Li, C. Zhang, C. Chen, T. Huang, A. Yu, Three-dimensional porous Si
and SiO2 with in situ decorated carbon nanotubes as anode materials for Li-ion batteries,
ACS Appl. Mater. Interfaces 9 (2017) 17807-17813.
[54] L. Zong, Y. Jin, C. Liu, B. Zhu, X. Hu, Z. Lu, J. Zhu, Precise perforation and scalable
production of Si particles from low grade sources for high-performance lithium ion
battery anodes, Nano Lett. 16 (2016) 7210-7215.
[55] Z. Yi, N. Lin, Y. Zhao, W. Wang, Y. Qian, Y. Zhu, Y. Qian, A flexible
micro/nanostructured Si microsphere cross-linked by highlyelastic carbon nanotubes
toward enhanced lithium ion battery anodes, Energy Storage Mater. 17 (2019) 93-100.
[56] D. Wei, X. Gao, S. Zeng, H. Li, H. Li, W. Li, X. Tao, L. Xu, P. Chen, Improving the
performance of micro-silicon anodes in lithium-ion batteries with a functional carbon
nanotube interlayer, ChemElectroChem 5 (2018) 3143-3149.
[57] H. Zhang, X. Zhang, H. Jin, P. Zong, Y. Bai, K. Lian, H. Xu, F. Ma, A robust
hierarchical 3D Si/CNTs composite with void and carbon shell as Li-ion battery anodes,
Chem. Eng. J. 360 (2019) 974–981.

37
[58] S. Cui, S. Chen,L. Deng, Si nanoparticles encapsulated in CNTs arrays with tubular
sandwich structure for high performance Li ion battery, Ceram. Int. 46 (2020) 3242–
3249.
[59] X. Huang, Fabrication and properties of carbon fibers, Materials 2 (2009) 2369-2403.
[60] C. Xiang, N. Behabtu, Y. Liu, H. G. Chae, C. C. Young, B. Genorio, D. E. Tsentalovich,
C. Zhang, D. V. Kosynkin, J. R. Lomeda, C.-C. Hwang, S. Kumar, M. Pasquali, J. M.
Tour, Graphene nanoribbons as an advanced precursor for making carbon fiber, ACS
Nano 7 (2013) 1628-1637.
[61] M. Li, N. Li, W. Shao, C. Zhou, Synthesis of carbon nanofibers by CVD as a catalyst
support material using atomically ordered Ni3C nanoparticles, Nanotechnology 27 (2016)
505706.
[62] T. Hyeon, S. Han, Y.-E. Sung, K.-W. Park, Y.-W. Kim, High-performance direct
methanol fuel cell electrodes using solid-phase-synthesized carbon nanocoils, Angew.
Chem. Int. Ed. 42 (2003) 4352-4356.
[63] Y. Xu, Y. Zhu, F. Han, C. Luo, C. Wang, 3D Si/C fiber paper electrodes fabricated using
a combined electrospray/electrospinning technique for Li-ion batteries, Adv. Energy
Mater. 5 (2014) 1400753.
[64] A. Sun, J. Guo, C. Wang, A porous silicon–carbon anode with high overall capacity on
carbon fiber current collector, Electrochem. Commun. 12 (2010) 981-984.
[65] Y. Wen, Y. Zhu, A. Langrock, A. Manivannan, S. H. Ehrman, C. Wang,
Graphene-bonded and -encapsulated Si nanoparticles for lithium ion battery anodes,
Small 9 (2013) 2810-2816.
[66] X. Chen, P. Hu, J. Xiang, R. Zhang, Y. Huang, Confining silicon nanoparticles within
freestanding multichannel carbon fibers for high-performance Li-ion batteries, ACS
Appl. Energy Mater. 2 (2019) 5214−5218.
[67] L. Ji, X. Zhang, Evaluation of Si/carbon composite nanofiber-based insertion anodes for
new-generation rechargeable lithium-ion batteries, Energy Environ. Sci. 3 (2010)
124-129.
[68] S. Liu, W. Xu, C. Ding, J. Yu, D. Fang, Y. Ding, H. Hou, Boosting electrochemical
performance of electrospun silicon-based anode materials for lithium-ion battery by
surface coating a second layer of carbon, Appl. Surf. Sci. 494 (2019) 94–100.
[69] T. Song, D. H. Lee, M. S. Kwon, J. M. Choi, H. Han, S. G. Doo, H. Chang, W. I. Park,
W. Sigmund, H. Kim, U. Paik, Silicon nanowires with a carbon nanofiber branch as
lithium-ion anode material, J. Mater. Chem. 21 (2011) 12619-12621.

38
[70] S.-M. Jang, J. Miyawaki, M. Tsuji, I. Mochida, S.-H. Yoon, The preparation of a novel
Si–CNF composite as an effective anodic material for lithium–ion batteries, Carbon 47
(2009) 3383-3391.
[71] T. H. Hwang, Y. M. Lee, B.-S. Kong, J.-S. Seo, J. W. Choi, Electrospun core−shell
fibers for robust silicon nanoparticle-based lithium ion battery anodes, Nano Lett. 12
(2012) 802-807.
[72] X. Kong, Y. Zheng, Y. Wang, S. Liang, G. Cao,A. Pan, Necklace-like Si@C nanofibers
as robust anode materials for high performance lithium ion batteries, Sci. Bull. 64 (2019)
261–269.
[73] W. Stöber, A. Fink,E. Bohn, Controlled growth of monodisperse silica spheres in the
micron size range, J. Colloid Interfaces Sci. 26 (1968) 62-69.
[74] Y. Wang, X. Wen, J. Chen, S. Wang, Foamed mesoporous carbon/silicon composite
nanofiber anode for lithium ion batteries, J. Power Sources 281 (2015) 285-292.
[75] W. Wang, I. Ruiz, K. Ahmed, H. H. Bay, A. S. George, J. Wang, J. Butler, M. Ozkan, C.
S. Ozkan, Silicon decorated cone shaped carbon nanotube clusters for lithium ion battery
anodes, Small 10 (2014) 3389-3396.
[76] L. Cui, L. Hu, J. W. Choi, Y. Cui, Light-weight free-standing carbon nanotube-silicon
films for anodes of lithium ion batteries, ACS Nano 4 (2010) 3671-3678.
[77] D. Wei, X. Tao, X. Gao, S. Zeng, L. Xu, P. Chen, Improving the performance of
micro-silicon anodes in lithium-ion batteries with a functional carbon nanotube interlayer,
ChemElectroChem 5 (2018) 3143-3149.
[78] A. K. Geim, Graphene: status and prospects, Science 324 (2009) 1530-1534.
[79] G. Liu, W. Jin, N. Xu, Graphene-based membranes, Chem. Soc. Rev. 44 (2015)
5016-5030.
[80] R. Raccichini, A. Varzi, S. Passerini, B. Scrosati, The role of graphene for
electrochemical energy storage, Nat. Mater. 14 (2015) 271-279.
[81] C.N.R. Rao, K. Gopalakrishnan, A. Govindaraj, Synthesis, properties and applications of
graphene doped with boron, nitrogen and other elements, Nano Today 9 (2014) 324-343.
[82] X. Zhou, Y.Yin, L. Wan, Y. Guo, Self-assembled nanocomposite of silicon nanoparticles
encapsulated in graphene through electrostatic attraction for lithium-ion batteries, Adv.
Energy Mater. 2 (2012) 1086-1090.
[83] H. Xiang, K. Zhang, G. Ji, J. Y. Lee, C. Zou, X. Chen, J. Wu, Graphene/nanosized
silicon composites for lithium battery anodes with improved cycling stability, Carbon 49
(2011) 1787-1796.

39
[84] J. K. Lee, K. B. Smith, C. M. Haynerb, H. H. Kung, Silicon nanoparticles–graphene
paper composites for Li ion battery anodes, Chem. Commun. 46 (2010) 2025-2027.
[85] X. Zhou, Y. Liu, C. Du, Y. Ren, R. Xiao, P. Zuo, G. Yin, Y. Ma, X. Cheng,Y. Gao,
Layer-by-layer engineered silicon-based sandwich nanomat as flexible anode for
lithium-ion batteries, ACS Appl. Mater. Interfaces 11 (2019) 39970−39978.
[86] S. Ciampi, J. B. Harper, J. J. Gooding, Wet chemical routes to the assembly of organic
monolayers on silicon surfaces via the formation of Si–C bonds: surface preparation,
passivation and functionalization, Chem. Soc. Rev. 39 (2010) 2158-2183.
[87] X. Tang, G. Wen, Y. Song, Stable silicon/3D porous N-doped graphene composite for
lithium-ion battery anodes with self-assembly, Appl Surf. Sci. 436 (2018) 398-404.
[88] C. Fang, Y. Deng, Y. Xie, J. Su, G. Chen, Improving the electrochemical performance of
Si nanoparticle anode material by synergistic strategies of polydopamine and graphene
oxide coatings, J. Phys. Chem. C 119 (2015) 1720-1728.
[89] N. Kim, C. Oh, J. Kim, J.-S. Kim, E. D. Jeong, J.-S. Bae, T. E. Hong, J. K. Lee,
High-performance Li-ion battery anodes based on silicon-graphene self-assemblies, J.
Electrochem. Soc. 164 (2017) A6075-A6083.
[90] M. Zhou, F. Pu, Z. Wang, T. Cai, H. Chen, H. Zhang, S. Guan, Facile synthesis of novel
Si nanoparticles–graphene composites as high-performance anode materials for Li-ion
batteries, Phys.Chem.Chem.Phys. 15 (2013) 11394-11401.
[91] X. Zhou, Y. Yin, L. Wan, Y. Guo, Facile synthesis of silicon nanoparticles inserted into
graphene sheets as improved anode materials for lithium-ion batteries, Chem. Commun.
48 (2012) 2198-2200.
[92] K. Feng, W. Ahn, G. Lui, H. W. Park, A. G. Kashkooli, G. Jiang, X. Wang, X. Xiao, Z.
Chen, Implementing an in-situ carbon network in Si/reduced graphene oxide for high
performance lithium-ion battery anodes, Nano Energy 19 (2016) 187-197.
[93] Y. Yan, Z. Xu, C. Liu, H. Dou, J. Wei, X. Zhao, J. Ma, Q. Dong, H. Xu, Y. He, Z. Ma, X.
Yang, Rational design of the robust janus shell on silicon anodes for high-performance
lithium-ion batteries, ACS Appl. Mater. Interfaces 11 (2019) 17375-17383.
[94] Q. Xu, J. Sun, J. Li, Y. Yin, Y. Guo, Scalable synthesis of spherical Si/C granules with
3D conducting networks as ultrahigh loading anodes in lithium-ion batteries, Energy
Storage Mater. 12 (2018) 54-60.
[95] G. Zhao, L. Zhang, Y. Meng, N. Zhang, K. Sun, Decoration of graphene with silicon
nanoparticles by covalent immobilization for use as anodes in high stability lithium ion
batteries, J. Power Sources 240 (2013) 212-218.

40
[96] G. Lin, H. Wang, L. Zhang, Q. Cheng, Z. Gong, K. Ostrikov, Graphene nanowalls
conformally coated with amorphous/ nanocrystalline Si as high-performance binder-free
nanocomposite anode for lithium-ion batteries, J. Power Sources 437 (2019) 226909.
[97] M. Naguib, V. N. Mochalin, M. W. Barsoum, Y. Gogotsi, 25th Anniversary Article:
MXenes: A New Family of Two-Dimensional Materials, Adv. Mater. 26 (2014) 992–
1005.
[98] C. Wang, N. Kurra, M. Alhabeb, J. Chang, H. N. Alshareef, Y. Gogotsi, Titanium
carbide (MXene) as a current collector for lithium-ion batteries, ACS Omega 3 (2018)
12489−12494.

[99] P. Zhang, Q. Zhu, Z. Guan, Q. Zhao, N. Sun, B. Xu, A Fflexible Si@C electrode with
excellent stability employing an MXene as a multifunctional binder for lithium-ion
batteries, ChemSusChem 13 (2020) 1621 –1628.
[100] C. Zhang, S. Park, A. Seral‐Ascaso, S. Barwich, N. McEvoy, C. S. Boland, J. N.
Coleman, Y. Gogotsi, V. Nicolosi, High capacity silicon anodes enabled by MXene
viscous aqueous ink, Nat. Commun. 10 (2019) 849.
[101] Y. Tian, Y. An, J. Feng, Flexible and freestanding silicon/MXene composite papers for
high-performance lithium-ion batteries, ACS Appl. Mater. Interfaces 11 (2019)
10004−10011.

[102] D. Cao, M. Ren, J. Xiong, L. Pan, Y. Wang, X. Ji, T. Qiu, J. Yang, C. Zhang,
Self-assembly of hierarchical Ti3C2Tx-CNT/SiNPs resilient films for high performance
lithium ion battery electrodes, Electrochim. Acta 348 (2020) 136211.
[103] Y. Zhang, Z. Mu, J. Lai, Y. Chao, Y. Yang, P. Zhou, Y. Li, W. Yang, Z. Xia, S. Guo,
MXene/Si@SiOx@C Layer-by-layer superstructure with autoadjustable function for
superior stable lithium storage, ACS Nano 13 (2019) 2167−2175.

[104] X. Zhu, J. Shen, X. Chen, Y. Li, W. Peng, G. Zhang, F. Zhang, X. Fan, Enhanced
cycling performance of Si-MXene nanohybrids as anode for high performance lithium
ion batteries, Chem. Eng. J. 378 (2019) 122212.
[105] P. Guan, J. Li, T. Lu, T. Guan, Z. Ma, Z. Peng, X. Zhu, L. Zhang, Facile and scalable
approach to fabricate granadilla-like porous-structured silicon-based anode for lithium
ion batteries, ACS Appl. Mater. Interfaces 10 (2018) 34283-34290.
[106] B. Ma, B. Lu, J. Luo, X. Deng, Z. Wu, X. Wang, The hollow mesoporous silicon
nanobox dually encapsulated by SnO2/C as anode material of lithium ion battery,
Electrochim. Acta 288 (2018) 61-70.

41
[107] B. Wang, W. Li, T. Wu, J. Guo, Z. Wen, Self-template construction of mesoporous
silicon submicrocube anode for advanced lithium ion batteries, Energy Storage Mater. 15
(2018) 139-147.
[108] W. An, B. Gao, S. Mei, B. Xiang, J. Fu, L. Wang, Q. Zhang, P. K. Chu, K. Huo,
Scalable synthesis of ant-nest-like bulk porous silicon for high-performance lithium-ion
battery anodes, Nat. Commun. 10 (2019) 1447.
[109] G. D. Park, J. H. Choi, D. S. Jung, J.-S. Park, Y. C. Kang, Three-dimensional porous
pitch-derived carbon coated Si nanoparticles-CNT composite microsphere with superior
electrochemical performance for lithium ion batteries, J. Alloy. Comp. 821 (2020)
153224.
[110] J. Liang, Z. Zhang, W. Yang, Y. Liu, X. Zhang, M. Javid, Y. Jung, X. Dong,
Three-dimensional porous carbon skeleton supporting Si nanosheets as anode for
high-performance lithium ion batteries, Ionics 26 (2020) 2233–2245.
[111] G. Hasegawa, K. Kanamori, K. Nakanishi, K. Hayashi, Thermogravimetric evolved gas
analysis and microscopic elemental mapping of the solid electrolyte interphase on
silicon incorporated in free-standing porous carbon electrodes, Langmuir 35 (2019)
12680−12688.
[112] H. Chen, S. He, X. Hou, S. Wang, F. Chen, H. Qin, Y. Xia, G. Zhou, Nano-Si/C
microsphere with hollow double spherical interlayer and submicron porous structure to
enhance performance for lithium-ion battery anode, Electrochim. Acta 312 (2019)
242-250.
[113] X. Wu, C. Yu, C. Li, Carbon-encapsulated gigaporous microsphere as potential Si
anode-active material for lithium-ion batteries, Carbon 160 (2020) 255-264.
[114] S. Chae, S.-H. Choi, N. Kim, J. Sung, J. Cho, Integration of graphite and silicon anodes
for the commercialization of high-energy lithium-ion batteries, Angew. Chem. Int. Ed.
59 (2020) 110 –135.
[115] N. Kim, S. Chae, J. Ma, M. Ko, J. Cho, Fast-charging high-energy lithium-ion batteries
via implantation of amorphous silicon nanolayer in edge-plane activated graphite anodes,
Nat. Commun. 8 (2017) 812.
[116] J. Li, G. Li, J. Zhang, Y. Yin, F. Yue, Q. Xu, Y. Guo, Rational Design of Robust Si/C
microspheres for high-tap-density anode materials, ACS Appl. Mater. Interfaces 11
(2019) 4057−4064.

42
[117] W. Huang, J. Wang, M. R. Braun, Z. Zhang, Y. Li, D. T. Boyle, P. C. McIntyre, Y. Cui,
Dynamic structure and chemistry of the silicon solid-electrolyte interphase visualized by
cryogenic electron microscopy, Matter 1 (2019) 1–14.
[118] M. J. Zachman, Z. Tu, S. Choudhury, L. A. Archer, L. F. Kourkoutis, Cryo-STEM
mapping of solid–liquid interfaces and dendrites in lithium-metal batteries, Nature 560
(2018) 345–349.
[119] H. Son, J. H. Park, S. Kwon, S. Park, M. H. Rümmeli, A. Bachmatiuk, H. J. Song, J. Ku,
J. W. Choi, J.-m. Choi, S.-G. Doo, H. Chang, Silicon carbide-free graphene growth on
silicon for lithium-ion battery with high volumetric energy density, Nat. Commun. 6
(2015) 7393.
[120] R. Yazami, Y. Reynier, Thermodynamics and crystal structure anomalies in
lithium-intercalated graphite, J. Power Sources 153 (2006) 312-318.
[121] X. Liu, L. Zhong, S. Huang, S. X. Mao, T. Zhu, J. Huang, Size-dependent fracture of
silicon nanoparticles during lithiation, ACS Nano 6 (2012) 1522-1531.
[122] S.-M. Bak, Z. Shadike, R. Lin, X. Yu, X. Yang, In situ/operando synchrotron-based
X-ray techniques for lithium-ion battery research, NPG Asia Mater. 10 (2018) 563-580.
[123] X. Yang, J. McBreen, W.-S. Yoon, M. Yoshio, H. Wang, K. Fukuda, T. Umeno,
Structural studies of the new carbon-coated silicon anode materials using
synchrotron-based in situ XRD, Electrochem. Commun. 4 (2002) 893-897.
[124] D. Liu, Z. Shadike, R. Lin, K. Qian, H. Li, K. Li, S. Wang, Q. Yu, M. Liu, S.
Ganapathy, X. Qin, Q. Yang, M. Wagemaker, F. Kang, X. Yang, B. Li, Review of recent
development of in situ/operando characterization techniques for lithium battery research,
Adv. Mater. 31 (2019) 1806620.
[125] R. Baddour-Hadjean, J.-P. Pereira-Ramos, Raman microspectrometry applied to the
study of electrode materials for lithium batteries, Chem. Rev. 110 (2010) 1278-1319.
[126] P. P. R. M. L. Harks, F. M. Mulder, P. H. L. Notten, In situ methods for Li-ion battery
research: a review of recent developments, J. Power Sources 288 (2015) 92-105.
[127] J. Nanda, M. K. Datta, J. T. Remillard, A. O‘Neill, P. N. Kumta, In situ Raman
microscopy during discharge of a high capacity silicon–carbon composite Li-ion battery
negative electrode, Electrochem. Commun. 11 (2009) 235-237.
[128] Y. Xu, G. Yin, P. Zuo, Geometric and electronic studies of Li15Si4 for silicon anode,
Electrochim. Acta 54 (2008) 341-345.

43
[129] B. Breitung, P. Baumann, H. Sommer, J. Janek, T. Brezesinski, In situ and operando
atomic force microscopy of high-capacity nano-silicon based electrodes for lithium-ion
batteries, Nanoscale 8 (2016) 14048-14056.
[130] C. R. Becker, K. E. Strawhecker, Q. P. McAllister, C. A. Lundgren, In situ atomic force
microscopy of lithiation and delithiation of silicon nanostructures for lithium ion
batteries, ACS Nano 7 (2013) 9173-9182.
[131] A. Tokranov, B. W. Sheldon, C. Li, S. Minne, X. Xiao, In situ atomic force microscopy
Study of initial solid electrolyte interphase formation on silicon electrodes for Li-ion
batteries, ACS Appl. Mater. Interfaces 6 (2014) 6672-6686.
[132] M. Haro, V. Singh, S. Steinhauer, E. Toulkeridou, P. Grammatikopoulos, M. Sowwan,
Nanoscale heterogeneity of multilayered Si anodes with embedded nanoparticle scaffolds
for Li‐ion batteries, Adv. Sci. 4 (2017) 1700180.
[133] F. Blanc, M. Leskes, C. P. Grey, In situ solid-state NMR spectroscopy of
electrochemical cells: batteries, supercapacitors, and fuel cells, Acc. Chem. Res. 46
(2013) 1952-1963.
[134] K. Ogata, E. Salager, C.J. Kerr, A.E. Fraser, C. Ducati, A. J. Morris, S. Hofmann, C. P.
Grey, Revealing lithium–silicide phase transformations in nano-structured silicon-based
lithium ion batteries via in situ NMR spectroscopy, Nat. Commun. 5 (2014) 3217.
[135] F. Sun, H. Markötter, I. Manke, A. Hilger, S. S. Alrwashdeh, N. Kardjilov, J. Banhart,
Complementary X-ray and neutron radiography study of the initial lithiation process in
lithium-ion batteries containing silicon electrodes, Appl. Surf. Sci. 399 (2017) 359-366.
[136] T. M. Fears, M. Doucet, J. F. Browning, J. K. S. Baldwin, J. G. Winiarz, H. Kaiser, H.
Taub, R. L. Saccig, G. M. Veith, Evaluating the solid electrolyte interphase formed on
silicon electrodes: a comparison of ex situ X-ray photoelectron spectroscopy and in situ
neutron reflectometry, Phys. Chem. Chem. Phys. 18 (2016) 13927-13940.
[137] F. Shi, P. N. Ross, G. A. Somorjai, K. Komvopoulos, The chemistry of electrolyte
reduction on silicon electrodes revealed by in situ ATR-FTIR spectroscopy, J. Phys.
Chem. C 121 (2017) 14476−14483.
[138] R. Endo, T. Ohnishi, K. Takada, T. Masuda, In situ observation of lithiation and
delithiation reactions of a silicon thin film electrode for all-solid-state lithium-ion
batteries by X‑ray photoelectron spectroscopy, J. Phys. Chem. Lett. 11 (2020)
6649−6654.

44
[139] M. T. McDowell, S. W. Lee, J. T. Harris, B. A. Korgel, C. Wang, W. D. Nix, Y. Cui, In
situ TEM of two-phase lithiation of amorphous silicon nanospheres, Nano Lett. 13 (2013)
758−764.
[140] B. Philippe, R. Dedryv re, J. Allouche, F. Lindgren, M. Gorgoi, H. Rensmo, D.
Gonbeau, K. Edstr m, Nanosilicon electrodes for lithium-ion batteries: interfacial
mechanisms studied by hard and soft X-ray photoelectron spectroscopy, Chem. Mater.
24 (2012) 1107−1115.

Figure 1. (a-c) Fundamental problems associated with Si-based anodes in Li-ion batteries.

Reproduced with permission [5]. Copyright 2017, Wiley-VCH. (d) Hierarchical core-shell

structures used to prevent Si particle pulverization. Reproduced with permission [6].

Copyright 2016, Elsevier. (e) Carbon coating to ensure the stability of SEI layers. Reproduced

with permission [7]. Copyright 2016, Elsevier. (f) Concrete-like structural design to enhance

the strength of the electrode. Reproduced with permission [8]. Copyright 2015, Elsevier.

45
Figure 2. Scheme of fundamental setup for electrospinning. Reproduced with permission [22].
Copyright 2012, Wiley-VCH.

Figure 3. Schematic illustration of a CVD reactor.

46
Figure 4. (a) Schematic of Si/CNT hybrid nanostructure fabrication. (b) Cycling capacity and

Coulombic efficiency curves of a coin cell over 25 cycles. Reproduced with permission [52].

Copyright 2010, American Chemical Society.

Figure 5. (a) Schematic of pSS/CNTs synthesis. (b) Cycling performance of the pSS/CNT

anode. (c) EIS measurements of the pSS/CNT anode. Reproduced with permission [53].

Copyright 2017, American Chemical Society.

47
Figure 6. (a) Schematic of the synthesis and architecture of a flexible 3D Si/C fibre paper

electrode before and after carbonization. (b) Rate capability of the flexible 3D Si/C fibre

paper electrode. (c) Cyclic voltammograms of the flexible 3D Si/C fibre paper electrode in the

1st, 2nd, and 5th cycles. (d) Cycling stability and Coulombic efficiency of flexible 3D Si/C

fibre paper electrodes with different overall loading masses. Reproduced with permission [63].

Copyright 2015, Wiley-VCH.

48
Figure 7. (a) Electrospinning with a dual nozzle and subsequent carbonization. (b) Cycling

performance of SiNPs@C at 3 C rate. (c) EIS measurements of SiNPs@C at the 25th, 50th,

100th, 200th, and 300th cycles. (d) Rate capability test of SiNPs@C. Reproduced with

permission [71]. Copyright 2012, American Chemical Society.

49
Figure 8. (a) Preparation and microstructural characteristics of the Si/G nanocomposite. (b)

Capacity of the Si/G nanocomposite and Si@G nanosheets cycled at 100 mA/g. (c) Cyclic

voltammograms of the Si/G nanocomposite from the first to the sixth cycle. (d) Nyquist plots

of the Si/G nanocomposite and Si@G nanosheets, where solid circles represent results after

the first cycles and the hollow circles represent results after rate performance. Reproduced

with permission [28]. Copyright 2012, The Royal Society of Chemistry.

50
Figure 9. (a) Schematic of a proposed composite comprising alternating strata of Si

nanoparticles and graphene (Si–re-G). (b) Cyclic voltammograms of the first three cycles of

the alternating strata Si–re-G composite. (c) Capacity–cycle profile of the alternating strata

Si–re-G composite anode with nine layers. (d) Specific capacity of five-layered (S1) and

nine-layered (S2) alternating strata Si–re-G composite anodes between 1.2 and 0.01 V at

current densities. Reproduced with permission [29]. Copyright 2013, The Royal Society of

Chemistry.

51
Figure 10. Schematics of (a) the RF plasma-enhanced horizontal tube furnace deposition

system, (b) GNWs prepared on the Ni foam, and (c) GNWs@Si composite prepared on the Ni

foam. (d) Raman spectrum of GNWs synthesized on the Ni foam, (e) cycling performance of

LIBs based on GNWs, pure silicon, and the GNWs@Si composite. Reproduced with

permission [96]. Copyright 2019, Elsevier.

52
Figure 11. (a) Schematic of the preparation procedure for MXene/Si@SiOx@C nanohybrids.

SEM images of (b) MXene, (c) MXene/Si, and (d)MXene/Si@SiOx@C. (e) TEM and (f, g)

HRTEM images of MXene/Si@SiOx@C. (h) Cycling performance of MXene/Si@SiOx@C-

1, MXene/Si@SiOx@C-2, MXene/Si@SiOx@C-3, and bare Si electrodes at 0.2C for 200

cycles. (i) Charge/discharge profiles of MXene/Si@SiOx@C-2 electrode at different C-rates.

Reproduced with permission [103]. Copyright 2019, American Chemical Society.

53
Figure 12. (a) Schematic of SiO microparticles encapsulated with vertical graphene. (b)

Stable electrical connection between SiO particles by interconnection of the vertical graphene

nanosheets. (c) Representative cyclic voltammograms of a d-SiO@vG electrode of the 1st,

2nd, and 5th cycles. (d) Nyquist plots of d-SiO@vG electrodes before and after cycling. (e)

Specific capacity and cycling efficiency of d-SiO and d-SiO@vG electrodes at a current

density of 320 mA/g. (f) Rate performance of the d-SiO@vG electrode, where 1 C =1600

mA/g. Reproduced with permission [39]. Copyright 2017, American Chemical Society.

54
Figure 13. (a) Schematic of the granadilla-like Si composite design. (b) Electrochemical

impedance plots of the 30@Si-granadilla, 40@Si-granadilla, and 50@Si-granadilla anodes.

(c) Cycling properties and (d) rate performance of the 30@Si-granadilla, 40@Si-granadilla,

and 50@Si-granadilla anodes. Reproduced with permission [37]. Copyright 2016,

Wiley-VCH.

Figure 14. (a) Schematic of the fabrication of a carbon-encapsulated porous Si-based

microsphere. SEM images of porous microspheres implanted with various u-Si contents of (b)

0 wt%, (c)25 wt%, (d) 50 wt%. Stability of capacities obtained during the charge process at a

constant current of 420 mA/g and the corresponding CE for (a) @u-Si and u-Si cells and (b)

@n-Si and n-Si cells. Reproduced with permission [113]. Copyright 2020, Elsevier.

55
Figure 15. (a) Gr–Si NPs attached to the surface of an Au wire and a Li/LiO2 electrode before

and after lithiation. Reproduced with permission [119]. Copyright 2015, Springer Nature. (b)

Schematic of the in situ TEM device.

Figure 16. (a) In situ/operando electrochemical coin cell designs for synchrotron-based X-ray

characterization. Reproduced with permission [122]. Copyright 2018, Springer Nature. (b)

In-situ XRD patterns of a graphite-mixed carbon-coated Si electrode during the first discharge.

Reproduced with permission [123]. Copyright 2002, Elsevier.

56
Figure 17. In-situ micro-Raman signal of Si as a function of the first discharge and charge

(inset) voltages, where the solid blue lines are the fit to the experimental spectra. Reproduced

with permission [127]. Copyright 2002, Elsevier.

Figure 18. (a) Schematic showing the setup used for in situ and operando AFM of nano-Si

electrodes containing polymer binder and carbon black additive. Upon lithiation (charging),

the Si particles expand in volume and probably fracture to some extent. In addition, SEI

57
formation on the free electrode surface occurs. (b) SEM-like images of a nano-Si electrode at

different potentials. The dashed circle indicates a surface region with significant structural

changes during the first lithiation cycle. (c) AFM height images of a nano-Si electrode

obtained in situ a) before cycling and in (b, first cycle; d, 2nd cycle) a ―fully‖ lithiated and (c,

first cycle) delithiated state. Each image shows the same area of the top surface. Reproduced

with permission [129]. Copyright 2016, The Royal Society of Chemistry.

Figure 19. Schematics of the electrochemical cell and the set-up for the in situ 7Li NMR

measurements. Reproduced with permission [133]. Copyright 2013, American Chemical

Society.

58

You might also like