Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Subscriber access provided by CARLETON UNIVERSITY

Energy, Environmental, and Catalysis Applications


Covalently-bonded Si-polymer Nanocomposites Enabled by
Mechanochemical Synthesis as Durable Anode Materials
Wenyue Shi, Hao Bin Wu, Jesse Baucom, Xianyang Li, Shengxiang Ma, Gen Chen, and Yunfeng Lu
ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.0c09938 • Publication Date (Web): 05 Aug 2020
Downloaded from pubs.acs.org on August 5, 2020

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 17 ACS Applied Materials & Interfaces

1
2
3
4
Covalently-bonded Si-polymer Nanocomposites Enabled by Mechanochemical
5
6
Synthesis as Durable Anode Materials
7
8
9 Wenyue Shia, Hao Bin Wuc, Jesse Baucoma, Xianyang Lia, Shengxiang Maa, Gen Chena,b,* and
10
11 Yunfeng Lua,*
12
13
14 a Department
15 of Chemical and Biomolecular Engineering, University of California, Los Angeles,
16 Los Angeles, CA 90095, USA
17
18 b School of Materials Science and Engineering, Central South University, Changsha, Hunan
19
20 410083, China
21 c School of Materials Science and Engineering, Zhejiang University, Hangzhou, Zhejiang 310027,
22
23 China
24
25 * Corresponding authors: Gen Chen (Email: geenchen@gmail.com), Yunfeng Lu
26
27 (luucla@ucla.edu)
28
29
30 KEYWORDS
31
32
33 Silicon/polymer composite; Resilient polymer coating; High-energy ball-milling; Lithium-ion
34
35 battery; Silicon anode material;
36
37
38
39
ABSTRACT
40 Silicon is one of the most promising anode materials for lithium-ion batteries due to its
41
42 high theoretical capacity and low cost. However, significant capacity fading caused by severe
43
44 structural degradation during cycling limits its practical implication. To overcome this barrier, we
45
design a covalently-bonded nanocomposite of silicon and poly(vinyl alcohol) (Si-PVA) by high-
46
47 energy ball-milling of a mixture of micron-sized Si and PVA. The obtained Si nanoparticles are
48
49 wrapped by resilient PVA coatings that covalently bonds to the Si particles. In such nanostructure,
50
51 the soft PVA coatings can accommodate the volume change of the Si particles during repeated
52 lithiation and delithiation. Simultaneously, as formed covalent bonds enhance the mechanical
53
54 strength of the coatings. Due to the significantly improved structural stability, the Si‒PVA
55
56 composite delivers a lifespan of 100 cycles with a high capacity of 1526 mAh g‒1. In addition, a
57
58
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 2 of 17

1
2
3 high initial Coulombic efficiency over 86% and an average value of 99.2% in subsequent cycles
4
5 can be achieved. This reactive ball milling strategy provide a low-cost and scalable route to
6
7 fabricate high performance anode materials.
8
9
10
11
12 1. INTRODUCTION
13
14 The fast-growing demand for electric vehicles, portable devices and grid-scale energy
15
16 storage urges the development of next-generation lithium-ion batteries (LIBs) with higher power
17
18 and energy density1–4. Considerable research efforts hence have been devoted to the engineering
19
of silicon, a promising anode material with moderate operating potential (0.1~0.4 V) and high
20
21 specific capacity (3578 mAh g‒1 for Li3.75Si) in comparison with the current commercial graphite
22
23 (0.05 V and 372 mAh g‒1).5–10 Bulk silicon resources are abundant and cost-effective, yet its severe
24
25 pulverization and repeated formation of solid-electrolyte interphase (SEI) caused by the huge
26 volume expansion (300%) result in irreversible capacity loss and poor cycling stability.11–13
27
28 Therefore, its practical application has been greatly restricted.
29
30
Aiming at improving the structural integrity, a variety of synthetic strategies have been
31
32 developed to fabricate silicon nanostructures.14–19 Significantly, special attention has been put into
33
34 the protection of silicon with functional coating layers, involving different methods such as
35
36 CVD,20–22 heat treatment,14,23–25 sol-gel26,27 and so on.28–31 However, the complexity, inefficiency
37 and high-cost generally impede their large-scale production and adaption. High-energy ball-
38
39 milling (HEBM) has been reported as a low-cost, high-yield and scalable method for the efficient
40
41 production of silicon nanomaterials, which is feasible for industrial manufacture.32 Different from
42
43
mixing ball-milling, energy can be re-dispersed and transported by HEBM with strong friction and
44 collision. More interestingly, the precedents of organic reactions and synthesis of metal-organic
45
46 frameworks demonstrate the feasibility of building chemical bonds through HEBM.33–35 As for
47
48 the preparation of Si-based anode materials, HEBM has mostly been engaged in size breaking-
49
down alone. With the smaller size benefited from HEBM, nanoporous Si particles can be obtained
50
51 with further wet etching process.36,37 Combined with necessary post-processing treatment, robust
52
53 chemical connection with protective layers can be constructed.38–41 However, the concept of
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 3 of 17 ACS Applied Materials & Interfaces

1
2
3 reactive high-energy ball-milling (RBM) has rarely been applied for the direct production of
4
5 silicon nanocomposites with chemically bonded coating layers.
6
7
Herein, we report polymer-coated silicon nanocomposites with in situ generation of
8
9 covalent bonds between Si nanoparticles and poly(vinyl alcohol) (PVA) by a solvent-free, one-pot
10
11 and no-post-process RBM method. PVA acts as the resilient component in the composites because
12
13 of its high tensile strength and flexibility. The flexible covalently-bonded coating buffers the
14 volume expansion of the silicon during lithiation process, which effectively prevents the particles
15
16 from pulverization and maintains the integrity of electrode. The size of Si nanoparticles can be
17
18 adjusted by regulating the content of PVA, among which silicon nanoparticles coated by 5% PVA
19
20
deliver the optimal electrochemical properties. A reversible specific capacity of 1526 mAh g‒1 is
21 obtained at 200 mA g‒1 after 100 cycles with loading of 1 mg cm‒2. A high initial Coulombic
22
23 efficiency (ICE) over 86% and an average of 99.2% have been achieved.
24
25
26
27
28 2. EXPERIMENTAL SECTION
29
30
2.1 Synthesis for PVA Coated Silicon Composite
31
32
33 Silicon (99%, ~325 mesh) and poly(vinyl alcohol) (MW 89,000-98,000, 99% hydrolyzed)
34
35 were purchased from Sigma-Aldrich without further purification. Si and different amount of PVA
36 (0, 2 wt%, 5 wt%, 10 wt%) were mixed in the jars with the ball-to-powder ratio of ~12, sealed in
37
38 argon atmosphere, and milled by the high-energy ball-miller (Retsch, High Energy Ball Mill Emax)
39
40 at 1000 rpm. After 10 h, all powder was collected without further treatment.
41
42 2.2 Structural Characterization
43
44
45
Crystalline structures of the Si-PVA powder were determined by the Rigaku powder X-ray
46 diffractometer (XRD) using Kα radiation (λ = 1.54 Å). Surface morphology and particle sizes
47
48 were investigated by scanning electron microscopy (Nova 230 Nano SEM), transmission electron
49
50 microscopy (FEI T12 Quick CryoEM and CryoET TEM) and High-resolution transmission
51
electron microscopy (FEI Titan S/TEM). Raman spectra was conducted on a Renishaw 2000
52
53 System with a He/Ne laser at a wavelength of 633 nm. Infrared spectra experiments were carried
54
55 out by a transmission mode on a Jasco 420 Fourier transform infrared (FTIR) spectrophotometer.
56
57
58
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 4 of 17

1
2
3 Thermogravimetric analysis (TGA) was performed in argon atmosphere by a ramping rate of 10
4
5 oC min‒1. Gas sorption measurements were conducted using a Micromeritics ASAP 2020 system
6
7 at 77 K. Prior to gas adsorption/desorption measurement, all power was degassed at 120 °C for
8
9
12 h. For X-ray photoelectron spectroscopy (XPS) studies, AXIS Ultra DLD was used for analysis.
10 All the spectra were fitted to Gaussian–Lorentzian functions and a linear-type background using
11
12 CasaXPS software. The binding energy values were all calibrated using C 1s peak at 284.5 eV.
13
14 2.3 Electrochemical Characterization
15
16
17 The electrochemical performance was evaluated using 2032-type coin cells with the Li foil
18
19
as the counter/reference electrode. The working electrodes were comprised of active materials
20 (Si-PVA composite), acetylene black, and poly(acrylic acid) (PAA) binder with the weight ratio
21
22 7:2:1 and were fabricated using a typical slurry method. The mass loading of the active material
23
24 was ~1 mg cm−2 for each electrode. The electrolyte was 1 M LiPF6 in a mixture of ethylene
25
carbonate (EC)/diethyl carbonate (DEC) with a volume ratio of 1:1 (LP40, BASF), and 5 vol%
26
27 fluoroethylene carbonate (FEC, Sigma-Aldrich) was used as an additive. The assembled cells
28
29 were galvanostatically cycled between 0.05 and 2 V vs Li+/Li at different current densities using
30
31 a LAND battery testing system (China). The specific capacities were calculated based on the total
32 mass of the Si-PVA powder. Cyclic voltammetry (CV) experiments for the first cycles were
33
34 performed using the electrochemical station Solartron 1860/1287 with a scan rate of 0.1 mV s−1
35
36 from open-circuit voltage (OCV, ∼3 V vs Li/Li+) to 0 and back to 2 V. Electrochemical impedance
37
38
spectroscopy (EIS) data were recorded ranging from 100 kHz to 10 mHz with a voltage amplitude
39 of 10 mV.
40
41
42
43
44 3. RESULTS AND DISCUSSION
45
46
47 The synthetic strategy is schematically illustrated in Figure 1. The bulk silicon powder
48
and PVA without any solvent are sealed in argon atmosphere in a ball-milling jar. Bearing the
49
50 impact from the steel balls, the micron-sized silicon is ground into nanoparticles. Simultaneously,
51
52 the mechanical energy also triggers the chemical reaction between freshly exposed surface of
53
54 silicon nanoparticles and reactive hydroxyl groups of PVA. Due to the low dissociation energy of
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 5 of 17 ACS Applied Materials & Interfaces

1
2
3 O-H bonds42, the hydroxyl groups on PVA cleave and form Si-O-C bonds, generating
4
5 nanocomposites of PVA coated silicon nanoparticles with robust chemical bonding.
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26 Figure 1. Schematic diagram of the reactive ball-milling (RBM) process starting with bulk silicon and the
27
28 product of poly(vinyl alcohol) (PVA) coated silicon nanoparticles with the covalent bonds between PVA
29
and the silicon on the surface.
30
31
32
33 The samples with different amount of PVA relative to silicon are denoted as RBM Si-
34 2%PVA, RBM Si-5% PVA and RBM Si-5%PVA, respectively. Pure ball-milled Si without
35
36 adding PVA is denoted as HEBM Si. The Raman spectrum of crystalline silicon generally contains
37
38 one sharp peak at 520 cm‒1, which will redshift in a size-dependent manner.43 Compared to raw
39
40
silicon materials, the characteristic peak of ball-milled silicon-PVA nanocomposite in Raman
41 spectrum shifts from 520 to 504 cm‒1 (Figure 2a) with different content of PVA, indicating the
42
43 formation of Si nanocrystals. The corresponding peak of HEBM Si, however, appears at a slightly
44
45 higher wavenumber, suggesting a larger particle size than that of the PVA coated silicon.
46
47 X-ray diffraction (XRD) patterns presents accordant results (Figure 2b). Increasing the
48
49 content of PVA in the composite significantly broadens the widths of diffraction peaks of (111),
50
51 (220) and (311) of Si-PVA composites, which directly confirms the reduced crystalline size in the
52 presence of PVA. Due to the turbulent crushing during RBM, the crystalline silicon experiences
53
54 a deformation into amorphous structure on the edge of the particles as suggested by the reduced
55
56 peak intensity in the XRD patterns, which might benefit the kinetics of lithium ions transport.8 N2
57
58
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 6 of 17

1
2
3 sorption isotherms indicate a rising BET surface area with the addition of PVA due to the reduced
4
5 particle size (Figure S1). Starting from micron-size silicon with a BET surface area of only 5.1
6
7 m2 g‒1, Si-PVA nanocomposites deliver much enlarged BET surface areas of 11.1, 17.4 and 19.2
8
9
m2 g‒1 for Si-2%PVA, Si-5%PVA and Si-10%PVA, respectively. The different volume of these
10 samples (Figure 2c) with the same weight indicate a reduced tap density and thus a decreasing
11
12 particle size. Based on the tap density of Pristine Si, the tap densities of ball-milled powders are
13
14 calculated accordingly and presented in Table S1.
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37 Figure 2. Characterizations of bulk silicon, HEBM Si, and RBM Si-PVA nanocomposites. (a) Raman
38 spectrum, (b) XRD pattern, (c) photographs comparing the tap density, TEM images of (d) HEBM Si and
39
40 (e) RBM Si-5% PVA and (f) HR-TEM image of RBM Si-5% PVA.
41
42
43 SEM characterizations were carried out to examine the morphology of samples after ball
44
45 milling. Owing to the exposure of the high-energy surface of the particles during HEBM, the
46
47
secondary aggregation of the nanoparticles is usually severe and inevitable. Without additional
48 PVA, the finely crushed primary HEBM Si particles naturally aggregate into secondary particles
49
50 up to 10 μm, which dispossesses the advantages of nanomaterials and renders a drastic structural
51
52 transition during lithiation and delithiation. By forming covalent bonds between the PVA and Si
53
particles, the active surface is chemically quenched, which effectively mitigates the formation of
54
55 aggregates. The RBM Si-5%PVA primary particles are typically below 200 nm (Figure S2a),
56
57
58
59
60 ACS Paragon Plus Environment
Page 7 of 17 ACS Applied Materials & Interfaces

1
2
3 notably smaller than the HEBM Si (Figure S2b). Most of the secondary particles of RBM Si-
4
5 5%PVA fall in the range of 0.5-1 μm. TEM images display more distinct differences. The RBM
6
7 Si-5% PVA particles are around 200 nm and well-dispersed (Figure 2e). In contrast, the HEBM
8
9
Si appears as an aggregation with size over 2 μm (Figure 2d). The sizes of RBM Si-2%PVA and
10 RBM Si-10%PVA are consistent with the BET results as well (Figure S3). The diffraction spots
11
12 corresponding to (111), (220) and (311) lattice planes of polycrystalline silicon can be found in
13
14 the SAED pattern (Figure S4). The (111) lattice of silicon with an interplanar spacing 0.32 nm
15
can be determined in the HRTEM image (Figure 2f). More importantly, a layer of amorphous
16
17 domain is observed around the Si nanocrystals, which may be assigned to the polymer coating
18
19 and/or amorphous Si. The Si nanoparticles are effectively separated by the in situ formed polymer
20
21 framework and the agglomeration is therefore remarkably suppressed during HEBM. Such unique
22 structure might accommodate the volume change of Si, and also prevent the Si particles from
23
24 aggregation upon repeated lithiation/delithiation.
25
26
The formation of covalent bonds between Si and PVA during the ball-milling process is
27
28 further supported by the X-ray photoelectron spectroscopy (XPS) profiles. Typical peaks for Si in
29
30 Si 2p spectrum located at 99.6, 100.3, 100.9 and 103.5 eV represent Si 2p3/2, Si 2p1/2, Si-O and
31
32 SiO2, respectively (Figure 3a).44 All these peaks remain in the spectrum of RBM Si-5%PVA. In
33 addition, a new peak at 102.5 eV corresponding to Si-O-C bonding appears,45 probably due to the
34
35 reaction between hydroxyl groups (-OH) from PVA and freshly exposed surface of silicon particles
36
37 during RBM (Figure 3b). After the RBM process, the C-H/C-C peak (284.8 eV) from PVA
38
39
polymer chain remains in C 1s spectrum, while the original C-OH peak at 286.3 eV is replaced by
40 a new peak at 286.9 eV for Si-O-C (Figure 3c, d). Consistently, in O 1s spectrum, the Si-O-C peak
41
42 (532.2 eV) in RBM Si-5%PVA completely replaces the C-OH peak (533.4 eV) in PVA (Figure
43
44 3e, f, and S5).46 These observations imply that hydroxyl groups in PVA have fully reacted with
45
silicon during RBM and a layer of Si-O-C coating layer has formed on the surface of silicon. The
46
47 FTIR result provides direct evidences of the covalent bonds on the coated silicon particles surface
48
49 (Figure 3g). The peak of Si-O-Si stretching around 1150-1200 cm‒1 for HEBM Si 47 also appears
50
51 in the spectrum of RBM Si-5%PVA, while the C-O peak in PVA for RBM Si-5%PVA shows up
52 at around 1100 cm‒1. More importantly, the characteristic peak of Si-O-C locates at around 1020
53
54 cm‒1. Additionally, the chemical linking between PVA and Si particles after RBM is confirmed
55
56 by TGA. Both RBM Si-5%PVA and HEBM Si exhibit extraordinary thermal stability up to 1000
57
58
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 8 of 17

1
2
3
C, compared to pure PVA with major weight loss occurring at around 250 C (Figure 3h).
4
5 Different from extensively reported physical coating/wrapping on the surface of Si nanoparticles,
6
7 RBM is capable to achieve in situ generation of covalent bonds, inducing a robust shell to protect
8
9 silicon particles. Furthermore, the complete removal of reactive groups in the raw materials might
10 also improve the electrochemical stability.
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31 Figure 3. Characterizations of the chemical structure of the HEBM Si and RBM Si-5%PVA. Si 2p XPS
32 spectra for (a) HEBM Si and (b) RBM Si-5% PVA. C 1s XPS spectra for (c) PVA and (d) RBM Si-5%
33
34 PVA. O 1s XPS spectra for (e) PVA and (f) RBM Si-5% PVA. (g) FTIR and (h) TGA for PVA, HEBM Si
35 and RBM Si-5% PVA.
36
37
38
39 The electrochemical lithiation/delithiation performances of the as-prepared silicon
40
particles were investigated in half-cell configurations. Figure 4a displays the CV feature of the
41
42 RBM Si-5%PVA for the first four cycles. In the first cycle, the distinctive cathodic peak around
43
44 0.05 V reveals the initial formation of LixSi (0x3.75) alloy process. The two anodic peaks at
45
46 0.36 V and 0.53 V can be assigned to the phase transformation from high-Li-content to low-Li-
47
content LixSi alloys and eventually to amorphous Si. The subsequent scans exhibit similar CV
48
49 profiles, with another cathodic peak emerging at 0.18 V. The current increases steadily as the
50
51 scans proceed, implying an activation process and formation of stable SEI on the electrode surface.
52
53 Besides the well-established lithiation/delithiation peaks, no side reaction was detected, which is
54 attributed to the elimination of reactive groups during RBM process.
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 9 of 17 ACS Applied Materials & Interfaces

1
2
3 Galvanostatic charging/discharging at 0.2 A g‒1 for RBM Si-5%PVA discloses stable
4
5 voltage profiles with well-defined charging/discharging plateaus. With the increasing content of
6
7 PVA in the active materials, the specific capacity slightly decreases. Meanwhile, stability of the
8
9
electrodes is notably improved, leading to higher capacity retention upon cycling. The initial
10 capacity for bare HEBM Si is 3255 mAh g‒1 with an ICE of only 83.0%, along with fast capacity
11
12 fading (Figure 4b). The stability is slightly improved with 2% of PVA. Nonetheless, the RBM
13
14 Si-5%PVA electrode exhibits initial specific capacity of around 3000 mAh g‒1 and an ICE of 86.1%
15
with negligible capacity loss in the first three cycles (Figure 4c). The RBM Si-10%PVA shows a
16
17 similar stability compared with RBM Si-5%PVA while the initial specific capacity drops to 2434
18
19 mAh g‒1 due to the higher weight ratio of inactive PVA (Figure S6). Through the PVA coating,
20
21 the stability and the ICE of the electrodes are greatly improved with minimal sacrifice on the
22 specific capacity.
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 10 of 17

1
2
3 Figure 4. Electrochemical characterizations of the RBM Si-PVA anodes. (a) CV for RBM Si-5%PVA
4
5 sample. Galvanostatic charging/discharging profiles for (b) HEBM Si and (c) RBM Si-5%PVA. (d) Rate
6
7 performance of HEBM Si and RBM Si-5%PVA electrodes with the loading of 1 mg cm‒2. (e) Long-term
8 cycling performances of ball-milled silicon materials at 0.2 A g‒1 with the mass loading of 1 mg cm‒2. (f)
9
10 EIS profiles of RBM Si-5%PVA electrode before and after cycling. Inset: magnified EIS profiles of the
11 high frequency region. (g) Fast charging/discharging (3.0 A g‒1) performance of RBM Si-5%PVA electrode
12
13 with the mass loading of 1 mg cm‒2.
14
15
16 Figure 4d compares the rate performance of RBM Si-5%PVA and HEBM Si electrodes.
17
18 Ranging from 0.15 to 3.0 A g‒1, the capacity of electrode drops stepwise to 801 mAh g‒1 and
19
20 recovers to 2234 mAh g‒1 once the current returns to 0.15 A g‒1. Figure 4e shows the galvanostatic
21
cycling stability of the ball-milled silicon materials with a loading of 1 mg cm-2 and a current
22
23 density of 0.2 A g‒1. In contrast to the drastic capacity fading of HEBM Si, Si-PVA
24
25 nanocomposites synthesized through RBM present enhanced stability and higher capacity
26
27 retention. It is noteworthy that the RBM Si-5%PVA offers the best improvement. Besides a high
28 ICE of 86.1% as mentioned earlier, the Coulombic efficiency (CE) rapidly reaches over 97% from
29
30 the second cycle onwards (Table S2). After being activated in the first 10 cycles, the electrode
31
32 steadily becomes stabilized and remains stable for later cycles, with 1526 mAh g‒1 reversible
33
34
specific capacity retained after 100 cycles and an average CE of 99.2%. A substantial capacity
35 loss to 1395 mAh g‒1 occurs in the first 5 cycles for RBM Si-10%PVA electrode, ending up with
36
37 a reversible capacity of 798 mAh g‒1 after 100 cycles. We speculate that the severe polymer
38
39 solvation and swelling from the higher content of PVA leads to a more significant structural change,
40
thus causing the quick capacity fade in the first several cycles. In addition, HEBM Si is simply
41
42 stirred and mixed with 5% of PVA to demonstrate the vital role of RBM. Owing to the side
43
44 reactions with the electrolytes induced by the reactive groups on PVA, the ICE is hardly over 70 %.
45
46 Without the robust chemical interaction to preserve the integrity, the electrode quickly fails. EIS
47 tests were conducted simultaneously at different cycle numbers. The diameters of the semicircles
48
49 in the plots were evaluated with the measurement and indicate the charge transfer resistance (Rct)
50
51 of the cell.48,49 For fresh electrodes, the resistance first gradually decreases from 316 Ω (HEBM
52
53
Si) to 240 Ω (RBM Si-5%PVA), due to the improved ion transportation kinetics benefited from
54 the reduced particle sizes. However, when the PVA content has reached 10%, the charge transfer
55
56 resistance increases back to 340 Ω, indicating the role of non-conductive additive PVA has
57
58
59
60 ACS Paragon Plus Environment
Page 11 of 17 ACS Applied Materials & Interfaces

1
2
3 overruled the slight size decrease compared to RBM Si-5%PVA (Figure S7). This result provides
4
5 another explanation for the superior cell performance with RBM Si-5%PVA. After the first cycle
6
7 of galvanostatic charging/discharging for the RBM Si-5%PVA electrode, the impedance reduces
8
9
dramatically to 56 Ω, suggesting the formation of ionic conductive SEI layer. The impedance of
10 the cell continues to decrease until it reaches 21 Ω after the 50th cycle, and remains stable at 21 Ω
11
12 to the 100th cycle (Figure 4f). At the beginning of cell cycling, the formation of SEI layers and
13
14 improved interfacial contact that can be ascribed to the activation and wetting process significantly
15
enhanced the charge transfer capability. Eventually, the Rct maintains constant, revealing a
16
17 stabilized SEI layer and mitigated Si pulverization. To strike a balance between the capacity and
18
19 stability, RBM Si-5%PVA is preferentially selected for the additional cycling tests. Further
20
21 increasing the current density to 3.0 A g‒1, the RBM Si-5%PVA electrode delivers a reversible
22 specific capacity 750 mAh g‒1 after 150 cycles (Figure 4g). Through the covalent linking between
23
24 silicon and PVA protective layer by RBM, the electrochemical performances in terms of CE and
25
26 cycling stability of the electrode are greatly improved.
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 12 of 17

1
2
3 Figure 5. (a) Schematic diagram of the lithiation/delithiaion process of the ball-milled silicon materials.
4
5 (b-g) Morphology of cycled silicon electrodes. SEM images of (b, c, d) HEBM Si electrodes and (e, f, g)
6
RBM Si-5%PVA after (b, e) one, (c, f) five and (d, g) ten cycles, respectively.
7
8
9
10
11
Figure 5a illustrates the working mechanism of the resilient PVA coating that promotes
12 the cycling stability of the silicon electrodes. On the surface of silicon, the flexible polymer PVA
13
14 with a linear chain structure can stretch and contract along with the volume expansion and
15
16 shrinkage of silicon particles during lithiation and delithiation. Unlike solely physical contact, the
17 stable chemical bonding between the PVA chains and particles surface prevents phase separation
18
19 during the volume variation of silicon. The PVA coating acts as a resilient buffer layer to
20
21 accommodate the volume expansion and thus prevents the particles from pulverization. As a result,
22
23
the integrity and thus cycling stability of the electrode would be notably improved. SEM images
24 of cycled electrodes provide a visible demonstration of the improved mechanical stability of Si-
25
26 PVA nanocomposites. After only one cycle of lithiation/delithiation, the surface of HEBM Si
27
28 electrode becomes rough with deep cracks with a typical width of 10.5 μm (Figure 5b). The
29
electrode film continuously breaks down upon cycling (Figure 5c), leading to obvious
30
31 disintegration after 10 cycles with almost 20 μm width cracks (Figure 5d). On the other hand, only
32
33 minor crevices typically 2.0 μm in width on the surface of electrode is observed for RBM Si-
34
35 5%PVA (Figure 5e), and the electrode remains generally intact from the 5th to the 10th cycle (Figure
36 5f, g). The RBM Si-10%PVA electrode shows a similar result (Figure S8). With only 5% PVA
37
38 coating, we successfully mitigate the pulverization and thus enhanced the cycling stability of the
39
40 silicon anodes.
41
42
43
44
45 4. CONCLUSIONS
46
47 In conclusion, we successfully developed an efficient RBM strategy for large-scale
48
49 preparation of Si-PVA nanocomposites as anode materials for LIBs by using in situ chemical
50
51 reaction between silicon particles and the hydroxyl groups on PVA chains. The Si-PVA
52
nanocomposite with 5% PVA exhibits a primary size of Si within 200 nm coated by PVA through
53
54 Si-O-C covalent bond on the surface. By varying the weight ratio of PVA, the sizes of the particles
55
56 can be regulated, among which 5% PVA coated silicon particles demonstrate optimal
57
58
59
60 ACS Paragon Plus Environment
Page 13 of 17 ACS Applied Materials & Interfaces

1
2
3 electrochemical performances. At 0.2 A g‒1, the RBM Si-5%PVA delivers a reversible capacity
4
5 over 1526 mAh g‒1 after 100 cycles, with an ICE over 86.1% and an average CE of 99.2%. Raising
6
7 the current to 3.0 A g‒1, a capacity around 800 mAh g‒1 after 150 cycles can be maintained. The
8
9
superior electrochemical performance can be attributed to the following aspects: (1) nanosized Si
10 particles from ball milling greatly shorten the diffusion pathway of Li+, leading to enhanced
11
12 kinetics; (2) polymer coating effectively reduces the size of primary nanoparticles; (3) the resilient
13
14 PVA coating layer with robust covalent bond is adaptive to the volume change of silicon particles.
15
Therefore, the severe pulverization of Si-based anode is effectively alleviated. In addition, the
16
17 synthetic strategy features one-pot and solvent-free reaction without post-procedure, as well as
18
19 low cost of the raw materials, thus being suitable for industrial manufacturing. Thus, the RBM
20
21 method developed in this work offers a feasible approach to fabricate Si-based anode materials for
22 next-generation LIBs.
23
24
25
26
27 ASSOCIATED CONTENT
28
29 Supporting Information
30
31
32
The Supporting Information is available free of charge on the ACS
33
34 Publications website at http://pubs.acs.org.
35
36 Supporting information includes materials and electrochemical
37
38 characterizations.
39
40
41
42
43 AUTHOR INFORMATION
44
45
46 Corresponding Author
47
48 * E-mail: geenchen@gmail.com, luucla@ucla.edu
49
50
51
52
53 Acknowledgement
54
55
56 We are grateful for the support of the UCLA-Dynavolt Research Center.
57
58
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 14 of 17

1
2
3 References
4
5
6 (1) Nitta, N.; Wu, F.; Lee, J. T.; Yushin, G. Li-Ion Battery Materials: Present and Future.
7 Materials Today 2015, 18, 252–264.
8
9
(2) Liu, Z.; Yu, Q.; Zhao, Y.; He, R.; Xu, M.; Feng, S.; Li, S.; Zhou, L.; Mai, L. Silicon
10 Oxides: A Promising Family of Anode Materials for Lithium-Ion Batteries. Chem. Soc.
11 Rev. 2019, 48, 285–309.
12 (3) Ma, J.; Sung, J.; Hong, J.; Chae, S.; Kim, N.; Choi, S.-H.; Nam, G.; Son, Y.; Kim, S. Y.;
13 Ko, M.; Cho, J. Towards Maximized Volumetric Capacity via Pore-Coordinated Design
14 for Large-Volume-Change Lithium-Ion Battery Anodes. Nat. Commun. 2019, 10, 475.
15
(4) Li, M.; Lu, J.; Chen, Z.; Amine, K. 30 Years of Lithium-Ion Batteries. Adv. Mater. 2018,
16
17 30, 1800561
18 (5) Mukanova, A.; Jetybayeva, A.; Myung, S.-T.; Kim, S.-S.; Bakenov, Z. A Mini-Review on
19 the Development of Si-Based Thin Film Anodes for Li-Ion Batteries. Materials Today
20 Energy 2018, 9, 49–66.
21 (6) Lu, J.; Chen, Z.; Pan, F.; Cui, Y.; Amine, K. High-Performance Anode Materials for
22 Rechargeable Lithium-Ion Batteries. Electrochem. Energ. Rev. 2018, 1, 35–53.
23
(7) Feng, K.; Li, M.; Liu, W.; Kashkooli, A. G.; Xiao, X.; Cai, M.; Chen, Z. Silicon-Based
24
25 Anodes for Lithium-Ion Batteries: From Fundamentals to Practical Applications. Small
26 2018, 14, 1702737
27 (8) Xu, Z.-L.; Liu, X.; Luo, Y.; Zhou, L.; Kim, J.-K. Nanosilicon Anodes for High
28 Performance Rechargeable Batteries. Prog. Mater. Sci. 2017, 90, 1–44.
29 (9) Jin, Y.; Zhu, B.; Lu, Z.; Liu, N.; Zhu, J. Challenges and Recent Progress in the
30 Development of Si Anodes for Lithium-Ion Battery. Adv. Energy Mater. 2017, 7,
31
32
1700715.
33 (10) Wang, B.; Ryu, J.; Choi, S.; Zhang, X.; Pribat, D.; Li, X.; Zhi, L.; Park, S.; Ruoff, R. S.
34 Ultrafast-Charging Silicon-Based Coral-Like Network Anodes for Lithium-Ion Batteries
35 with High Energy and Power Densities. ACS Nano 2019, 13, 2307–2315.
36 (11) Choi, M.; Kim, J.-C.; Kim, D.-W. Waste Windshield-Derived Silicon/Carbon
37 Nanocomposites as High-Performance Lithium-Ion Battery Anodes. Sci. Rep. 2018, 8,
38
960.
39
40 (12) Su, X.; Wu, Q.; Li, J.; Xiao, X.; Lott, A.; Lu, W.; Sheldon, B. W.; Wu, J. Silicon-Based
41 Nanomaterials for Lithium-Ion Batteries: A Review. Adv. Energy Mater. 2014, 4,
42 1300882.
43 (13) Choi, S.; Kwon, T.-W.; Coskun, A.; Choi, J. W. Highly Elastic Binders Integrating
44 Polyrotaxanes for Silicon Microparticle Anodes in Lithium Ion Batteries. Science 2017,
45 357, 279–283.
46
(14) Yoo, J.-K.; Kim, J.; Jung, Y. S.; Kang, K. Scalable Fabrication of Silicon Nanotubes and
47
48 Their Application to Energy Storage. Adv. Mater. 2012, 24, 5452–5456.
49 (15) Chen, H.; Hou, X.; Chen, F.; Wang, S.; Wu, B.; Ru, Q.; Qin, H.; Xia, Y. Milled Flake
50 Graphite/Plasma Nano-Silicon@carbon Composite with Void Sandwich Structure for
51 High Performance as Lithium Ion Battery Anode at High Temperature. Carbon 2018,
52 130, 433–440.
53 (16) Yoon, T.; Bok, T.; Kim, C.; Na, Y.; Park, S.; Kim, K. S. Mesoporous Silicon Hollow
54
55
Nanocubes Derived from Metal-Organic Framework Template for Advanced Lithium-Ion
56 Battery Anode. ACS Nano 2017, 11, 4808–4815.
57
58
59
60 ACS Paragon Plus Environment
Page 15 of 17 ACS Applied Materials & Interfaces

1
2
3 (17) Liu, N.; Wu, H.; McDowell, M. T.; Yao, Y.; Wang, C.; Cui, Y. A Yolk-Shell Design for
4
5
Stabilized and Scalable Li-Ion Battery Alloy Anodes. Nano Lett. 2012, 12, 3315–3321.
6 (18) Zhang, L.; Wang, C.; Dou, Y.; Cheng, N.; Cui, D.; Du, Y.; Liu, P.; Al-Mamun, M.;
7 Zhang, S.; Zhao, H. A Yolk-Shell Structured Silicon Anode with Superior Conductivity
8 and High Tap Density for Full Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2019, 58,
9 8824–8828.
10 (19) Li, P.; Chen, G.; Lin, Y.; Chen, F.; Chen, L.; Zhang, N.; Cao, Y.; Ma, R.; Liu, X. 3D
11
Network Binder via In Situ Cross‐Linking on Silicon Anodes with Improved Stability for
12
13
Lithium‐Ion Batteries. Macromol. Chem. Phys. 2020, 221, 1900414.
14 (20) Yao, Y.; McDowell, M. T.; Ryu, I.; Wu, H.; Liu, N.; Hu, L.; Nix, W. D.; Cui, Y.
15 Interconnected Silicon Hollow Nanospheres for Lithium-Ion Battery Anodes with Long
16 Cycle Life. Nano Lett. 2011, 11, 2949–2954.
17 (21) Jia, H.; Gao, P.; Yang, J.; Wang, J.; Nuli, Y.; Yang, Z. Novel Three-Dimensional
18 Mesoporous Silicon for High Power Lithium-Ion Battery Anode Material. Adv. Energy
19
Mater. 2011, 1, 1036–1039.
20
21 (22) Fu, K.; Xue, L.; Yildiz, O.; Li, S.; Lee, H.; Li, Y.; Xu, G.; Zhou, L.; Bradford, P. D.;
22 Zhang, X. Effect of CVD Carbon Coatings on Si@CNF Composite as Anode for
23 Lithium-Ion Batteries. Nano Energy 2013, 2, 976–986.
24 (23) Wu, H.; Chan, G.; Choi, J. W.; Ryu, I.; Yao, Y.; McDowell, M. T.; Lee, S. W.; Jackson,
25 A.; Yang, Y.; Hu, L.; Cui, Y. Stable Cycling of Double-Walled Silicon Nanotube Battery
26 Anodes through Solid-Electrolyte Interphase Control. Nat. Nanotechnol. 2012, 7, 310–
27
28
315.
29 (24) Liu, J.; Kopold, P.; van Aken, P. A.; Maier, J.; Yu, Y. Energy Storage Materials from
30 Nature through Nanotechnology: A Sustainable Route from Reed Plants to a Silicon
31 Anode for Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2015, 54, 9632–9636.
32 (25) Wu, J.; Qin, X.; Miao, C.; He, Y.-B.; Liang, G.; Zhou, D.; Liu, M.; Han, C.; Li, B.; Kang,
33 F. A Honeycomb-Cobweb Inspired Hierarchical Core–shell Structure Design for
34
Electrospun Silicon/Carbon Fibers as Lithium-Ion Battery Anodes. Carbon 2016, 98,
35
36
582–591.
37 (26) Xin, X.; Zhou, X.; Wang, F.; Yao, X.; Xu, X.; Zhu, Y.; Liu, Z. A 3D Porous Architecture
38 of Si/Graphene Nanocomposite as High-Performance Anode Materials for Li-Ion
39 Batteries. J. Mater. Chem. 2012, 22, 7724-7730.
40 (27) Chae, C.; Noh, H.-J.; Lee, J. K.; Scrosati, B.; Sun, Y.-K. A High-Energy Li-Ion Battery
41 Using a Silicon-Based Anode and a Nano-Structured Layered Composite Cathode. Adv.
42
Funct. Mater. 2014, 24, 3036–3042.
43
44 (28) He, Y.; Yu, X.; Wang, Y.; Li, H.; Huang, X. Alumina-Coated Patterned Amorphous
45 Silicon as the Anode for a Lithium-Ion Battery with High Coulombic Efficiency. Adv.
46 Mater. 2011, 23, 4938–4941.
47 (29) Lee, J.-I.; Ko, Y.; Shin, M.; Song, H.-K.; Choi, N.-S.; Kim, M. G.; Park, S. High-
48 Performance Silicon-Based Multicomponent Battery Anodes Produced via Synergistic
49 Coupling of Multifunctional Coating Layers. Energy Environ. Sci. 2015, 8, 2075–2084.
50
51
(30) Liang, Z.; Lin, D.; Zhao, J.; Lu, Z.; Liu, Y.; Liu, C.; Lu, Y.; Wang, H.; Yan, K.; Tao, X.;
52 Cui, Y. Composite Lithium Metal Anode by Melt Infusion of Lithium into a 3D
53 Conducting Scaffold with Lithiophilic Coating. Proc. Natl. Acad. Sci. 2016, 113, 2862–
54 2867.
55 (31) An, W.; Gao, B.; Mei, S.; Xiang, B.; Fu, J.; Wang, L.; Zhang, Q.; Chu, P. K.; Huo, K.
56
57
58
59
60 ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 16 of 17

1
2
3 Scalable Synthesis of Ant-Nest-like Bulk Porous Silicon for High-Performance Lithium-
4
5
Ion Battery Anodes. Nat. Commun. 2019, 10, 1447.
6 (32) Gauthier, M.; Mazouzi, D.; Reyter, D.; Lestriez, B.; Moreau, P.; Guyomard, D.; Roué, L.
7 A Low-Cost and High Performance Ball-Milled Si-Based Negative Electrode for High-
8 Energy Li-Ion Batteries. Energy Environ. Sci. 2013, 6, 2145-2155.
9 (33) Stolle, A.; Szuppa, T.; Leonhardt, S. E. S.; Ondruschka, B. Ball Milling in Organic
10 Synthesis: Solutions and Challenges. Chem. Soc. Rev. 2011, 40, 2317–2329.
11
(34) Yuan, W.; Friscić, T.; Apperley, D.; James, S. L. High Reactivity of Metal-Organic
12
13
Frameworks under Grinding Conditions: Parallels with Organic Molecular Materials.
14 Angew. Chem., Int. Ed. 2010, 49, 3916–3919.
15 (35) Bennett, T. D.; Saines, P. J.; Keen, D. A.; Tan, J.-C.; Cheetham, A. K. Ball-Milling-
16 Induced Amorphization of Zeolitic Imidazolate Frameworks (ZIFs) for the Irreversible
17 Trapping of Iodine. Chem. Eur. J 2013, 19, 7049–7055.
18 (36) Sohn, M.; Park, H.-I.; Kim, H. Foamed Silicon Particles as a High Capacity Anode
19
Material for Lithium-Ion Batteries. Chem. Commun. 2017, 53, 11897–11900.
20
21 (37) Nzabahimana, J.; Chang, P.; Hu, X. Porous Carbon-Coated Ball-Milled Silicon as High-
22 Performance Anodes for Lithium-Ion Batteries. J. Mater. Sci. 2019, 54, 4798–4810.
23 (38) Kasukabe, T.; Nishihara, H.; Iwamura, S.; Kyotani, T. Remarkable Performance
24 Improvement of Inexpensive Ball-Milled Si Nanoparticles by Carbon-Coating for Li-Ion
25 Batteries. J. Power Sources 2016, 319, 99–103.
26 (39) Zhu, B.; Jin, Y.; Tan, Y.; Zong, L.; Hu, Y.; Chen, L.; Chen, Y.; Zhang, Q.; Zhu, J.
27
28
Scalable Production of Si Nanoparticles Directly from Low Grade Sources for Lithium-
29 Ion Battery Anode. Nano Lett. 2015, 15, 5750–5754.
30 (40) Li, X.; Yang, D.; Hou, X.; Shi, J.; Peng, Y.; Yang, H. Scalable Preparation of
31 Mesoporous Silicon@C/Graphite Hybrid as Stable Anodes for Lithium-Ion Batteries. J.
32 Alloys Compd. 2017, 728, 1–9.
33 (41) Shen, C.; Fang, X.; Ge, M.; Zhang, A.; Liu, Y.; Ma, Y.; Mecklenburg, M.; Nie, X.; Zhou,
34
C. Hierarchical Carbon-Coated Ball-Milled Silicon: Synthesis and Applications in Free-
35
36
Standing Electrodes and High-Voltage Full Lithium-Ion Batteries. ACS Nano 2018, 12,
37 6280–6291.
38 (42) Dean, J. A. Lange’s Handbook of Chemistry. Materials and Manufacturing Processes
39 1990, 5, 687–688.
40 (43) Kim, K. H.; Lee, D. J.; Cho, K. M.; Kim, S. J.; Park, J.-K.; Jung, H.-T. Complete
41 Magnesiothermic Reduction Reaction of Vertically Aligned Mesoporous Silica Channels
42
to Form Pure Silicon Nanoparticles. Sci. Rep. 2015, 5, 9014.
43
44 (44) Zhang, L.; Kuramoto, N.; Azuma, Y.; Kurokawa, A.; Fujii, K. Thickness Measurement of
45 Oxide and Carbonaceous Layers on a28 Si Sphere by XPS. IEEE Trans Instrum Meas
46 2017, 66, 1297–1303.
47 (45) Sivaranjini, B.; Mangaiyarkarasi, R.; Ganesh, V.; Umadevi, S. Vertical Alignment of
48 Liquid Crystals over a Functionalized Flexible Substrate. Sci. Rep. 2018, 8, 8891.
49 (46) Khung, Y. L.; Ngalim, S. H.; Scaccabarozzi, A.; Narducci, D. Formation of Stable Si-O-C
50
51
Submonolayers on Hydrogen-Terminated Silicon(111) under Low-Temperature
52 Conditions. Beilstein J Nanotechnol 2015, 6, 19–26.
53 (47) Michalak, D. J.; Amy, S. R.; Estève, A.; Chabal, Y. J. Investigation of the Chemical
54 Purity of Silicon Surfaces Reacted with Liquid Methanol. J. Phys. Chem. C 2008, 112,
55 11907–11919.
56
57
58
59
60 ACS Paragon Plus Environment
Page 17 of 17 ACS Applied Materials & Interfaces

1
2
3 (48) Yohannes, Y. B.; Lin, S. D.; Wu, N.-L. In Situ DRIFTS Analysis of Solid Electrolyte
4
5
Interphase of Si-Based Anode with and without Fluoroethylene Carbonate Additive. J.
6 Electrochem. Soc. 2017, 164, A3641–A3648.
7 (49) Zhang, H.; Qin, X.; Wu, J.; He, Y.-B.; Du, H.; Li, B.; Kang, F. Electrospun Core–shell
8 Silicon/Carbon Fibers with an Internal Honeycomb-like Conductive Carbon Framework
9 as an Anode for Lithium Ion Batteries. J. Mater. Chem. A 2015, 3, 7112–7120.
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
For Table of Contents Only
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment

You might also like