Accepted Manuscript TiO2-Sn/C composite nanofibers with high-capacity and long-cycle life as anode materials for sodium ion batteries Su Nie, Li Liu, Junfang Liu, Jing Xia, Yue Zhang, Jianjun Xie, Min Li, Xianyou Wang PII:
S0925-8388(18)33273-0
DOI:
10.1016/j.jallcom.2018.09.044
Reference:
JALCOM 47463
To appear in:
Journal of Alloys and Compounds
Received Date: 12 May 2018 Revised Date:
31 August 2018
Accepted Date: 4 September 2018
Please cite this article as: S. Nie, L. Liu, J. Liu, J. Xia, Y. Zhang, J. Xie, M. Li, X. Wang, TiO2-Sn/C composite nanofibers with high-capacity and long-cycle life as anode materials for sodium ion batteries, Journal of Alloys and Compounds (2018), doi: 10.1016/j.jallcom.2018.09.044. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. 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.
ACCEPTED MANUSCRIPT TiO2-Sn/C composite nanofibers with high-capacity and long-cycle life as anode
2
materials for sodium ion batteries
3
Su Niea, Li Liua,b*, Junfang Liua, Jing Xiaa, Yue Zhanga, Jianjun Xiea,
4
Min Lia, Xianyou Wanga
RI PT
1
5
a
6
Joint Engineering Laboratory for Key materials of New Energy Storage Battery,
7
Hunan Province Key Laboratory of Electrochemical Energy Storage and Conversion,
8
School of Chemistry, Xiangtan University, Xiangtan 411105, China
9
b
M AN U
SC
National Base for International Science & Technology Cooperation, National Local
Key Laboratory of Advanced Energy Materials Chemistry (Ministry of Education),
10
Nankai University, Tianjin 300071, China
11
Abstract:
The novel TiO2-Sn/C composite nanofibers have been successfully fabricated by
13
a simple and facile electrospinning process. A small amount of metal tin interacts with
14
TiO2 nanoparticles in a carbon matrix, which makes TiO2-Sn/C nanofibers have the
15
stability of TiO2 and the high capacity of Sn. At the same time, the TiO2-Sn/C
16
nanofibers reveal an improved diffusion coefficient of sodium ions due to a small
17
amount of Sn nanoparticles incorporation. Compared with TiO2/C nanofibers, the
18
TiO2-Sn/C nanofibers electrode shows significantly improved specific capacity,
19
substantial cycling stability, and remarkable rate capability. It delivers high reversible
20
capacity of 255 mA h g-1 at at current densities of 0.05 A g-1 in the range of 0.01-2.5V
21
vs. Na/Na+, and has specific capacities of 214 and 147 mA h g-1 at current densities of
AC C
EP
TE D
12
*Corresponding author at: College of Xiangtan University. China.
E-mail addresses:
[email protected] (L. Liu). 1
ACCEPTED MANUSCRIPT 0.5 and 4 A g-1, respectively. Furthermore, TiO2-Sn/C nanofibers electrode
23
demonstrates a discharge capacity 190.8 mA h g-1 with 95.4 % retention after 1000
24
cycles at 1 A g-1 (the initial and second discharge capacity is 483 and 201 mA h g-1,
25
respectively). Even up to 5 A g-1, the discharge capacity of 134.3 mA h g-1 with
26
83.9 % retention is obtained after 1000 cycles. Outstanding electrochemical
27
performance makes TiO2-Sn/C nanofibers as a hopeful anode material for sodium-ion
28
batteries and to be applied in the field of large-scale energy storage.
29
Keywords:
30
Nanofibers; Titanium dioxide; Metal tin; Electrospinning; Sodium-ion batteries;
31
Introduction
M AN U
SC
RI PT
22
In recent years, lithium-ion batteries (LIBs) have been widely applied to portable
33
equipment, electric vehicles and other fields as an important energy storage device
34
due to their merits of high energy density, long cycle life[1–4]. Nevertheless, the high
35
cost limits the application of LIBs in large-scale energy storage for the limited and
36
uneven distribution of lithium resources. Compared with lithium resources, sodium
37
resources have the advantages of low cost and abundant resources. Therefore,
38
sodium-ion batteries (SIBs) have attracted extensive interest and been regarded as a
39
promising source to satisfy the requirement of large-scale energy storage and smart
40
grid applications in recent years.[5–7] However, due to sodium ions have a large
41
radius compared to lithium ions (ionic radius, 1.02 Å versus 0.76 Å), so it’s hard to
42
discover applicable materials for reversible and rapid ion insertion/extraction.
43
AC C
EP
TE D
32
Among a variety of reported anode materials, titanium dioxide (TiO2) has been
2
ACCEPTED MANUSCRIPT regarded as a hopeful anode material for SIBs due to the merits of low cost,
45
structurally stable and environmental friendliness[8-10]. At present, there are various
46
polymorphs of TiO2, for instance anatase[11,12], rutile[13], TiO2(B)[14], TiO2(H)[15],
47
have been researched as anode materials for NIBs. Nonetheless, these materials show
48
large differences in the electrochemical performance of sodium storage. The
49
electrochemical capabilities of various phases of TiO2 nanospheres have been reported
50
by Wang et al. [16], and concluded that anatase TiO2 had a preferable Na-storage
51
ability compared with mixed anatase/rutile or amorphous TiO2. However, the poor
52
electrical conductivity (~10-12 S cm-1) and low ion diffusion coefficients of TiO2 lead
53
to inferior rate performance, which hinders its practical application.[17] Moreover, it
54
is well known that TiO2 anodes show unattractive specific capacity, which is usually
55
below 200 mA h g-1. These disadvantages greatly restrict the research an attractive
56
and practical application of TiO2 anodes.
TE D
M AN U
SC
RI PT
44
Another type of anodes, based on the alloying materials[18, 8] and conversion
58
materials[19, 20] (e.g. red P, Sb2S3, Fe2O3, MoS2 and so on) have high specific
59
capacities. However, large volume expansion result in the crushing of the electrode
60
material and subsequently generate the irreversible capacity loss and inferior
61
cycle performance during the charge and discharge processes. Particularly, metallic tin
62
(Sn) has received widespread attention as a sodium ion battery anode matter for high
63
theoretical capacity (847 mA h g-1 for Na15Sn4) as well as excellent electronic
64
conductivity[21]. However, the pulverization of the electrode material caused by
65
serious volume changes (~520 %) hinders the practical application of Sn.
AC C
EP
57
3
ACCEPTED MANUSCRIPT 66
Nanocrystallization of tin particles is an effective strategy to prevent the pulverization
67
of Sn, which could reduce the mechanical stress produced during the alloying and
68
de-alloying reaction to sodium ions[22]. Electrospinning technology is widely used in the preparation of one-dimensional
70
nanostructured lithium/sodium ion battery electrode material due to its simple and
71
versatile characteristics.[23, 24] At present, there have been a lot of research reports
72
on the preparation of high-performance electrode materials (such as Sn/C, Fe3O4/C,
73
TiO2/C) by electrospinning technology.[25-27] The advantages of electrospinning are
74
mainly reflected in the following two aspects. On one hand, the obtained
75
one-dimensional nanofibers have large specific surface area and a high surface area to
76
volume ratio, providing many active sites. On the other hand, the porous structure
77
effectively alleviates the large volume change during the reaction of the battery, which
78
is beneficial to the sufficient wetting of the electrolyte and the increase of the
79
potential, thereby improving the energy storage efficiency.[28, 29]
TE D
M AN U
SC
RI PT
69
Herein, we first report the preparation of TiO2-Sn/C composite nanofibers
81
(denoted as TiO2-Sn/C NFs) through simple electrospinning technique. The
82
as-prepared TiO2-Sn/C NFs combine the advantages of TiO2 and Sn, that a small
83
amount of Sn nanoparticles increases the specific capacity of the electrode material
84
obviously while retaining the excellent cycling stability of TiO2. The one-dimensional
85
structure of nanofibers can provide a large specific surface area for sufficient
86
contacting with the electrolyte and short ions/electrons transmission paths[30, 31].
87
Furthermore, anatase TiO2 and Sn nano-particles are uniformly embedded in the
AC C
EP
80
4
ACCEPTED MANUSCRIPT carbon matrix, which not only enhances stability but also improves the overall
89
conductivity. As expected, TiO2-Sn/C NFs electrode exhibits high reversible capacity
90
255 mA h g-1 at 0.05 A g-1, superior rate property (214 and 147 mA h g-1 at current
91
densities of 0.5 and 4 A g-1, respectively), and outstanding long cycling performance
92
(134.3 mA h g-1 with 83.9 % retention at 5 A g-1 over 1000 cycles) as anode materials,
93
offering a promising application in constructing low cost and high capacity SIBs.
94
2. Experimental Section
95
2.1. Preparation of Materials
M AN U
SC
RI PT
88
The TiO2-Sn/C NFs were manufactured by electrospinning technology followed
97
by post-calcination treatment. The precursor solution for electrospinning was made as
98
follows: First, 8 ml N, N-dimethylformamide (DMF, Kermel, 99.5 %) and 1.5 ml
99
acetic acid glacial (CH3COOH, Kermel, 99.5 %) were mixed, then adding with 0.1 g
100
tin dichloride dihydrate (SnCl2·2H2O, Kermel, 99.5 %) and 1.5 ml tetra-n-butyl
101
titanate (C16H36O4Ti, Kermel, 99 %). After that, 0.75 g polyvinylpyrrolidone (PVP,
102
Mw=1,300,000, Alfa-Aesar) was added into the above mixed solution under magnetic
103
stirring for 8 h to obtain a precursor solution. The precursor solution was loaded into a
104
10 mL syringe connected with a blunt-tip stainless-steel needle and spun on an
105
electrospinning unit with the applied voltage of 15 kV. The needle-to-collector
106
distance was 15 cm and the flow rate was 0.38 ml h-1. The as-collected nanofibers
107
were first dried at 70 °C for 8 h in a vacuum oven and then pre-calcined at 200 °C for
108
2 h with a heating rate of 4 °C min-1. Finally, The TiO2-Sn/C composite nanofibers
109
can be obtained by carbonization at 600 °C for 4 h in Ar/H2 atmosphere with a heating
AC C
EP
TE D
96
5
ACCEPTED MANUSCRIPT rate of 4°C min-1. For comparison, the pristine TiO2/C NFs and other two kinds of
111
TiO2-Sn/C composite nanofibers were prepared with similar methods by adjusting the
112
additive amounts of SnCl2·2H2O to 0.0, 0.05, and 0.2 g, respectively.
113
2.2. Structure and Physical Characterization
RI PT
110
The structure and crystallinity of the sample were determined by X-ray powder
115
diffraction (XRD) in a Rigaku D/Max-2500 powder diffractometer with Cu-Kα
116
radiation (λ=1.5418 Å). The surface morphologies and detailed microstructures of the
117
synthesized products were observed by using a scanning electron microscope (SEM,
118
JEOL, SM-71480) and a transmission electron microscopy (TEM, JEOL,
119
JEM-100CX). X-ray photoelectron spectroscopy (XPS, ThermoFisher, Escalab 250XI)
120
was
121
Thermogravimetry
122
adsorption-desorption isotherms were measured with TriStar II3020. Inductively
123
Coupled Plasma Optical Emission Spectrometer (ICP-OES, Optima 5300DV) is used
124
to determine the content of elements in composite fibers.
125
2.3. Electrochemical Measurements
examine was
M AN U
to
the
valence
state
executed
through
TGA
of
as-prepared
(Q600,
SDT).
materials. Nitrogen
EP
TE D
performed
AC C
126
SC
114
The working electrode consisted of active materials (TiO2-Sn/C NFs or TiO2/C
127
NFs), carbon black, polyvinylidene fluoride (PVDF) binder in a mass ratio of
128
70:20:10, dissolved in an appropriate amount of N-Methyl-2-pyrrolidinone (NMP)
129
and the copper foil was employed as the current collector. For SIBs, the half-cells
130
were assembled with pure sodium foil as the counter electrode and a glass fiber
131
(Whatman, GF/C) as separator in an argon-filled glove box. The electrolyte solution
6
ACCEPTED MANUSCRIPT was composed of 1 M NaClO4 dissolved in propylene carbonate/ethylene carbonate
133
(PC/EC, 1:1 in volume). The loading mass of the active material is about 1.0 ± 0.1 mg
134
cm-2. The coin cells were cycled by galvanostatic discharge–charge measurements
135
with a battery testing system (Neware, China) at room temperature in the voltage
136
interval of 0.01 and 2.5 V (vs. Na+/Na). Both cyclic voltammetry (CV) tests and
137
electrochemical impedance spectra (EIS) experiments were performed on a CHI604E
138
electrochemistry workstation within a potential window of 0.01–2.5 V. The signal of
139
alternating current (AC) perturbation was ±5 mV and the frequency range was 100
140
mHz-105 Hz. For LIBs, the testing batteries were assembled with Celgard 2300 film
141
as the separator, lithium metal as the counter electrode, and 1 M LiPF6 dissolved in
142
dimethyl carbonate/ethylene carbonate (DMC/EC, 1:1 in volume) as the electrolyte.
143
Similarly, the coin cells were tested by galvanostatic discharge–charge measurements
144
with a battery testing system (Neware, China) at room temperature in the voltage
145
interval of 0.01 and 3.0 V (vs. Li+/Li).
146
3. Results and discussion
147
3. 1. Structure and morphology
SC
M AN U
TE D
EP
AC C
148
RI PT
132
In order to analyze the phase composition and structures of the TiO2/C NFs and
149
TiO2-Sn/C NFs, X-ray diffraction (XRD) measurements were implemented. All the
150
diffraction peaks of TiO2/C NFs are well indexed to anatase TiO2 (JCPDS No.
151
21-1272) in Figure 1a. The weak intensity of the TiO2 peaks in XRD patterns may be
152
ascribed to TiO2 nanoparticles embedded in the carbon matrix[32]. As shown in
153
Figure 1b, the main diffraction peaks of TiO2-Sn/C NFs can be well indexed to the
7
ACCEPTED MANUSCRIPT standard XRD patterns of anatase TiO2 (JCPDS No. 21-1272) and minor peaks are
155
correspond to tetragonal Sn (JCPDS No. 04-0673), which prove that TiO2-Sn/C NFs
156
are made up of a lot of anatase TiO2 and a little amount of metal tin. In addition, the
157
XRD pattern of TiO2-Sn/C NFs with different amounts of SnCl2·2H2O is illustrated in
158
Figure S1a. It can be observed that the diffraction peak of tetragonal Sn is becoming
159
sharper with the amount of SnCl2·2H2O increases, signifying a rise of Sn content in
160
TiO2-Sn/C NFs. The XRD patterns of precursors of TiO2-Sn/C NFs and PVP NFs are
161
shown in Figure S1b. Neither the precursor nor the pre-calcined sample can find the
162
diffraction peak of SnCl2.2H2O and its peaks is consistent with the peaks of PVP
163
precursor. The reason for this phenomenon may be that the content of SnCl.2H2O is
164
small or TiO2 is an amorphous state, so that it can be hidden and cannot be found.
M AN U
SC
RI PT
154
Raman spectra in Figure 1c reveal two prominent peaks at 1350 and 1600 cm-1,
166
corresponding to a disorder-induced feature (D-band) and the E2g mode of graphite
167
(G-band), respectively. The low intensity ratio of D versus G band (ID/IG<1.0) can be
168
used to indicate a certain degree of graphitization in a carbon matrix[33]. The values
169
of ID/IG for TiO2/C NFs and TiO2-Sn/C NFs are 0.84 and 0.82 respectively, which
170
represent high-quality electrical conductivity[34, 35]. Based on the weight losses in
171
the TG curve under air atmosphere(Figure 1d), the total TiO2 and TiO2-Sn content in
172
the as-prepared samples were determined to be 69.6 wt % and 71.2 wt %. The weak
173
content difference may be related to the addition of SnCl2.2H2O in the precursor
174
solution. Where, the reaction occurring between predominant weight loss step
175
between 400˚C to 500˚C is oxidized of carbon. The thermogravimetric curves of the
AC C
EP
TE D
165
8
ACCEPTED MANUSCRIPT 176
precursor and the annealed sample under N2 atmosphere are shown in Figure S1c. It
177
can be seen that the temperature at which PVP is carbonized is about 450 ℃. The scanning electron microscopy (SEM) images for TiO2/C NFs (Figure 2a, b)
179
and TiO2-Sn/C NFs (Figure 2c, d) are exhibited in Figure 2. It can be seen that both
180
TiO2/C NFs and TiO2-Sn/C NFs show long and continuous uniform diameter of about
181
110 nm. Figure 2b, d shows high magnification SEM images of TiO2/C NFs and
182
TiO2-Sn/C NFs, respectively. Smooth surface and interconnected networks are
183
observed in the obtained fibers. There are no aggregated particles can be discovered
184
on the surface of the fibers. It implies that all TiO2 and Sn particles are embedded in
185
the carbon matrix, which is corresponding with the XRD results. Further, the N2
186
adsorption–desorption
187
Brunauer-Emmett-Teller (BET) surface area and pore volume of TiO2-Sn/C NFs and
188
TiO2/C NFs. As shown in Figure S2. The Brunauer–Emmett–Teller (BET) surface
189
area and the pore volume of TiO2-Sn/C NFs are 136.9 m2 g-1 and 0.107 cm3 g-1
190
relative to the TiO2/C NFs, which showed BET surface area and pore volume of 117.9
191
m2 g-1 and 0.094 cm3 g-1. The increased specific surface area and pore volume may be
192
because of gas generation during the reduction of tin dioxide to generate micropores,
193
shown in the following formula[36]: S nO 2 + C → Sn + CO 2 (1)
194
The larger specific surface area ensures enough contact between the electrode material
195
and the electrolyte, thereby shortening the transport path to accelerate the rapid
196
transfer of Na+ ions.
197
M AN U
SC
RI PT
178
are
performed
to
determine
the
AC C
EP
TE D
measurements
Transmission electron microscopy (TEM) was further implemented to explore
9
ACCEPTED MANUSCRIPT the morphology and microstructure of obtained fibers. As shown in Figure 3a, TiO2
199
nanoparticles are dispersed in the carbon matrix to form TiO2/C composite nanofibers.
200
In addition, it can be clearly observed in Figures 3b-c that the particles are dispersed
201
uniformly in the carbon matrix, which not only greatly increases the conductivity of
202
TiO2 but also prevents the aggregation of Sn particles. Simultaneously, the TEM
203
mapping images of TiO2-Sn/C NFs in Figure 3d also show that the homogeneous
204
distribution of C, O, Sn and Ti elements in the nanofibers. The distinct lattice fringes
205
with a d-spacing of 0.35 nm and 0.15 nm can be observed in HRTEM image (Figure
206
3e) and interplanar crystal spacing statistical tables (Figure 3f), corresponding to the
207
(1 0 1) and (2 1 3) faces of anatase TiO2, respectively. However, it is regrettable that
208
the lattice fringes of Sn could not be found, which may be related to the fact that the
209
content of Sn is too little, or the rather smaller size of Sn nanoparticles and their
210
lattice fringes were covered by the lattice fringes of TiO2[37]. As shown in the EDS
211
spectrum (Figure S3), the content of Sn is only about 6.87 wt.%. In order to further
212
understand the composite ratio of Sn and TiO2, ICP-OES tests were conducted on
213
different samples. As shown in Table 1, the content of Sn in the composite fiber
214
increases with the increase of SnCl2.2H2O addition.
SC
M AN U
TE D
EP
AC C
215
RI PT
198
To further gain the elemental composition and valence, X-ray photoelectron
216
spectroscopy (XPS) was applied to survey the as-fabricated TiO2/C NFs and
217
TiO2-Sn/C NFs. The various elemental composition of two samples is marked in the
218
wide spectrum (Figure 4a) and the high-resolution spectra of C, Ti, Sn are further
219
displayed in Figure 4b-d, respectively. As shown in Figure 4b, there is no change of C
10
ACCEPTED MANUSCRIPT 1s peaks in TiO2/C NFs and TiO2-Sn/C NFs. The curves at 284.8 eV, 286.2 eV and
221
289.1 eV can be assigned to the C–C, C–O, and O–C=O bonds, respectively[38, 39].
222
XPS results of Ti 2p of TiO2/C NFs and TiO2-Sn/C NFs are emerged in Figure 4c.
223
There are Ti 2p1/2 and Ti 2p3/2 peak at ca.458.9 eV and ca.464.7 eV respectively,
224
suggesting that the predominant state of Ti element in TiO2/C NFs is present in the
225
form of Ti4+[40]. For TiO2-Sn/C NFs, the corresponding peaks shift slightly to the
226
lower binding energy of 458.7 and 464.5 eV, which is related to the formation of
227
oxygen vacancies that lead to changes in the chemical environment[41].The Sn 3d
228
peak of TiO2-Sn/C NFs can be deconvoluted into Sn 3d3/2 (494.5 eV) and Sn 3d5/2
229
(486.1 eV) in Figure 4d. Among them, the binding energy of Sn 3d5/2 (486.1 eV) is
230
higher than 484.9 eV (Sn0) but lower that 486.3 (Sn2+), which means that a part of Sn
231
particles are covered with a tin suboxide layer[42, 43]. Similarly, Figure S4 shows the
232
O1s trend of TiO2-Sn/C NFs and TiO2/C NFs. It can be seen that only the appearance
233
of Ti-O and Ti-OH appears in the O 1s diagram, which means that Sn is mainly in Sn0.
234
In addition, both the main peaks of Ti-O and Ti-OH have the same tendency of slight
235
deviation towards low binding energy as Ti 2p, which may be attributed to changes in
236
the chemical environment.
237
3.2. Electrochemical Performance
SC
M AN U
TE D
EP
AC C
238
RI PT
220
Galvanostatic discharge–charge measurements were applied to examine the
239
electrochemical performance of TiO2-Sn/C NFs and TiO2/C NFs. As indicated in
240
Figure 5a, the specific discharge capacity of TiO2/C NFs electrode is only 341.3 mA h
241
g-1 in the first discharge process and 171.4 mA h g-1 in the second cycle at 1 A g-1.
11
ACCEPTED MANUSCRIPT After 1000 cycles, the capacity of TiO2/C NFs is 100.2 mA h g-1 and the
243
corresponding capacity retention rate is 58.5 % (based on the capacity of the second
244
cycle). Under the same condition, the TiO2-Sn/C NFs electrode displays a higher
245
capacity of 483.1 mA h g-1 at the first discharge process and can be maintain
246
reversible capacity of 190.8 mA h g-1 after 1000 cycles at a current density of 1 A g-1,
247
corresponding to 95.4 % capacity retention (according to the capacity of the second
248
cycle (200.1 mA h g-1)). Correspondingly, the overall profiles for TiO2-Sn/C NFs in
249
the voltage range of 0.01–2.5 V (vs. Na/Na+) are exhibited in Figure 5b. There was no
250
significant change in the charging and discharging platform with the increase of the
251
number of cycle, corresponding to the glorious cycling stability. The large capacity
252
loss of the premiere cycle is principally connected with the interfacial reaction of the
253
electrode with the electrolyte leading to the formation of the solid-electrolyte
254
interfacial (SEI) film[44].
TE D
M AN U
SC
RI PT
242
The rate performance of two samples is expressed in Figure 5c and the
256
TiO2-Sn/C NFs electrode reveals more excellent rate performance than TiO2/C NFs.
257
TiO2-Sn/C NFs provide a discharge capacity of 255 mA h g-1 at 0.05 A g-1. The
258
discharge capacities keep at 244, 227, 214, 195, 165, 156, 147 and 139 mA h g-1 when
259
the current density increases to 0.1, 0.2, 0.5, 1, 2, 3, 4, and 5 A g-1, respectively. The
260
specific capacity can still be maintained at 245 mA h g-1 in the following cycles when
261
the current densities recovered to 0.05 A g-1. Ulteriorly, the rate capability of
262
TiO2-Sn/C NFs and TiO2/C NFs is directly compared through capacity remaining ratio
263
(shown in Figure S5). It is clearly seen that the TiO2-Sn/C NFs electrode releases
AC C
EP
255
12
ACCEPTED MANUSCRIPT much higher specific capacity than the TiO2/C NFs electrode at large current densities;
265
e.g., 139 and 41.6 mA h g-1 at 5 A g-1 are acquired for TiO2-Sn/C NFs and TiO2/C NFs,
266
corresponding to capacity retention rates of 49.6 % and 20.6 % (compared to the
267
capacity of the second cycle at 0.05 A g-1), respectively. The high rate (or fast) charge
268
and discharge performance of the TiO2-Sn/C NFs can be highly enhanced by the
269
improved kinetics provided for electronically conducting metal tin nanoparticles
270
reside in TiO2-Sn/C NFs. The discharge/charge curves emerge into Figure 5d, the
271
specific discharge capacity gradually decrease with the increasing current density.
272
However, the TiO2-Sn/C NFs electrode shows a less polarization at high current
273
densities than TiO2/C NFs (Figure S6a), which is consistent with excellent rate
274
performance. This excellent rate performance may originate from the interaction of
275
TiO2 nanoparticles and a small amount of Sn nanoparticles in the carbon matrix.
TE D
M AN U
SC
RI PT
264
To explore the detailed information of the outstanding electrochemical
277
performance of TiO2-Sn/C NFs, the detailed charge and discharge diagrams from the
278
second cycle to the fifth cycle at a small current of 0.05 A g-1 and cyclic voltammetry
279
was carried out. Compared to TiO2/C NFs (Figure S6b), there is one unspectacular
280
plateau in the discharge curves of TiO2-Sn/C NFs at 0.35 V appears on the Figure 5e,
281
corresponding to the sodiation course of metal tin[21]. A small anodic peak near 0.35
282
V is observed in the CV plot of TiO2-Sn/C NFs for the second cycle at 0.1 mV s-1
283
relative to TiO2/C NFs (shown in Figure 5f). This peak is well consistent with the
284
platform of charge and discharge curves. From the above studies, it can be easily
285
inferred that the outstanding electrochemical performance of TiO2-Sn/C NFs is
AC C
EP
276
13
ACCEPTED MANUSCRIPT derived from the electrochemical reaction of Sn. Concretely, the CV curves of the first
287
four cycles of TiO2-Sn/C NFs and TiO2/C NFs are presented in Figure S7. There is a
288
broad and weak reduction peak of TiO2-Sn/C NFs appears at ca. 0.97 V,
289
corresponding to the typical irreversible reaction of Sn in the first sodication scan[45].
290
In addition, the CV curve of TiO2-Sn/C NFs (Figure S7a) has a good coincidence after
291
the initial cycle, which is in consonance with glorious cycling stability. To better
292
understand the different sodium-alloying, CV test was performed at a small scanning
293
speed of 0.05 mV s-1. As shown in Figure S8, no more peaks were found to
294
correspond to sodium-Sn alloying except 0.35 V, which is the main peak of the
295
sodium alloying reaction of Sn.[46, 47] This may be related to the low content of Sn
296
in the composite fiber.
M AN U
SC
RI PT
286
For the sake of evidence that TiO2-Sn/C NFs have excellent electrochemical
298
performance, the long cycle testing was implemented at high current density of 5 A
299
g-1 (Figure 5g). It should be emphasized that there is 83.9 % retention (based on the
300
capacity of the second cycle (150.9 mA h g-1)) with specific capacity remaining at
301
134.3 mA h g-1 even for 1000 cycles. Besides, to confirm the effect of Sn addition on
302
the TiO2-Sn/C NFs composite, the electrochemical properties of samples with
303
different amounts of SnCl2·2H2O (0.0 g, 0.05 g, 0.1 g and 0.2 g) are shown in Figure
304
S9. It can be seen the addition of tin increases could improve remarkably its capacity.
305
Among those samples, the composite added with 0.1 g SnCl2·2H2O shows the best
306
electrochemical performance (delivers high reversible capacity of 232 mA h g-1 after
307
100 cycles at 0.1 A g-1 ).
AC C
EP
TE D
297
14
ACCEPTED MANUSCRIPT In order to further highlight the meliority of TiO2-Sn/C NFs, some outstanding
309
and representative reports about anatase TiO2 based anodes compared with TiO2-Sn/C
310
NFs for sodium-ion storage performance are presented in Figure 6. It can be seen
311
intuitively that TiO2-Sn/C NFs electrode has higher discharge capacity than many
312
other researches at different current densities. Although nanoporous anatase TiO2
313
exhibits superior specific capacity at a large current density, its cycle performance is
314
dissatisfactory. It has a discharge specific capacity of only 163 mA h g-1 after 500
315
cycles at a current density of 1C (335 mA g-1)[48].
M AN U
SC
RI PT
308
Based on the outstanding sodium-storage performance, it is significant to try to
317
extend the lithium-storage properties of the TiO2-Sn/C NFs sample. As shown in
318
Figure S10a, the TiO2-Sn/C NFs electrode also displays more outstanding rate
319
performance than TiO2/C NFs. The TiO2-Sn/C NFs electrode provides a discharge
320
capacity of 472.1 mA h g-1 at a current density of 0.05 A g-1. When the current density
321
increases to 0.1, 0.2, 0.5, 1, 2, 3, 4, and 5 A g-1, the discharge capacity moderately
322
decreases to 407.7, 357.8, 294.4, 247.1, 204.1, 182.4, 164.4, and 148.6 mA h g-1,
323
respectively. When the current density decreases to 50 mA g-1, the reversible capacity
324
can recover to 450.4 mA h g-1. Besides, the TiO2-Sn/C NFs electrode shows a higher
325
capacity of 762.2 mA h g-1 at the first discharge process and can stabilize the
326
discharge capacity at 284.7 mA h g-1 after 10 cycles at a current density of 1 A g-1.
327
Even after 500 cycles, the reversible capacity of 247.6 mA h g-1 can be maintained
328
(Figure S10b). The charge/discharge voltage profiles of the TiO2-Sn/C NFs and
329
TiO2/C NFs at different current densities in the voltage range of 0.01-3.0 V (vs. Li/Li+)
AC C
EP
TE D
316
15
ACCEPTED MANUSCRIPT are presented in Figure S10c, d, respectively. Compared with TiO2/C NFs, the
331
TiO2-Sn/C NFs electrode exhibits a less polarization as the current density increases.
332
From discussed above, tin metal has a significant effect on the TiO2-Sn/C electrode
333
for lithium ion battery as well as sodium ion battery.
RI PT
330
EIS measurement was introduced to explore dynamical properties of the
335
TiO2-Sn/C NFs electrode. As shown in Figure 7a, the EIS patterns are mainly
336
composed of three parts: small intercept at the high frequency region (Rs), semicircle
337
in the high frequency region (Rct) and sloping line in the low frequency region
338
(Zw)[55]. The Rs, Rct and Zw represent the resistance of the electrolyte contacting with
339
particles, the charge-transfer resistance and Na+ ion diffusion in the anode active
340
material, respectively. According to the fitting experimental data (see Table 2),
341
TiO2-Sn/C NFs show a lower Rct value of 242.8 Ω compared to TiO2/C NFs with a Rct
342
value of 297.1 Ω, indicative of the introduction of a small amount of Sn can achieve a
343
kinetics increase in the electrochemical reaction. The diffusion coefficient of sodion
344
ions (DNa+) could be figured out according to the following formulas[56, 57]:
M AN U
TE D
EP
D Na + =
R 2T 2 (2) Z ' = Rs + Rct + σ w ω -1/2 (3) 2 A2 n 4 F 4 c 2σ 2w
AC C
345
SC
334
346
where R is the gas constant, T is the Kelvin temperature, A is the surface area of the
347
electrode, n is the number of electrons per molecule during the electronic transfer
348
reaction, F is the Faraday constant, c is concentration of sodium ions, and σw is the
349
Warburg factor which could be gained by calculating the slope of the line Z'~ω-1/2 as
350
shown in Figure. 7b. By using Eq. (2) and (3), the sodium ion diffusion coefficients
351
(DNa+) of TiO2-Sn/C NFs after the 1st cycle increases to 6.9×10-13 cm2 s-1 compared to 16
ACCEPTED MANUSCRIPT the TiO2/C NFs value of 2.4×10-13 cm2 s-1. The increased diffusion coefficient of
353
sodium ions (DNa+) may be attributed to a small amount of Sn nanoparticles
354
incorporation, and bring about better electrochemical performance for TiO2-Sn/C NFs
355
electrode.
356
4. Conclusions
RI PT
352
In summary, electrospun TiO2-Sn/C nanofibers (TiO2-Sn/C NFs) show high
358
specific capacity, excellent cycle performance, and outstanding rate capability due to
359
the synergistic effect of anatase titanium dioxide and metal tin. Besides, compared to
360
pristine TiO2/C NFs, TiO2-Sn/C NFs show a larger specific surface area, which
361
promotes the contact with the electrolyte and shortens the diffusion route for both
362
electrons and ions. EIS results also reveal an increased diffusion coefficient of Na+
363
ions (DNa+) in TiO2-Sn/C NFs electrode owing to the introduction of metallic tin. As a
364
consequence, TiO2-Sn/C NFs electrode demonstrates a higher discharge capacity of
365
255 mA h g-1 at 0.05 A g-1 and can be kept a reversible capacity of 190.8 mA h g-1
366
with 95.4 % retention after 1000 cycles at 1 A g-1. Even the current density increases
367
to 5 A g-1, it still has the discharge capacity of 134.3 mA h g-1 with 83.9 % retention
368
after 1000 cycles, indicating prominent cycle stability and excellent rate capability. In
369
addition, the results show that metal tin also has a significant impact on the TiO2-Sn/C
370
electrode for lithium ion battery. The present work demonstrates that a small amount
371
of metal tin incorporation is an efficient way to improve the electrochemical
372
performance of TiO2/C nanofibers, and this strategy could promote the application of
373
TiO2 anodes for sodium ion batteries and lithium ion batteries.
AC C
EP
TE D
M AN U
SC
357
17
ACCEPTED MANUSCRIPT 374
Acknowledgements This work was supported financially by the National Science Natural Foundation
376
of China (Grant No. 51672234), the Research Foundation for Hunan Youth
377
Outstanding People from Hunan Provincial Science and Technology Department
378
(2015RS4030), Hunan 2011 Collaborative Innovation Center of Chemical
379
Engineering & Technology with Environmental Benignity and Effective Resource
380
Utilization, Program for Innovative Research Cultivation Team in University of
381
Ministry of Education of China (1337304), and the 111 Project (B12015).
382
References
383
[1] J. Liang, H. Hu, H. Park, C. Xiao, S. Ding, U. Paik, X.W. (David) Lou,
384
Construction of hybrid bowl-like structures by anchoring NiO nanosheets on flat
385
carbon hollow particles with enhanced lithium storage properties, Energy
386
Environ. Sci. 8 (2015) 1707–1711.
SC
[2] M. V. Reddy, G. V. Subba Rao, B.V.R. Chowdari, Metal oxides and oxysalts as anode materials for Li ion batteries, Chem. Rev. 113 (2013) 5364–5457.
391 392
[3] M. Ebner, F. Marone, M. Stampanoni, V. Wood, Visualization and quantification
AC C
390
EP
388 389
M AN U
TE D
387
RI PT
375
of electrochemical and mechanical degradation in Li-ion batteries, Sci. (80-. ). 342 (2013) 716–720.
[4]
C. Zhao, Y. Cai, K. Yin, H. Li, D. Shen, Carbon-Bonded, Oxygen-deficient TiO2,
393
nanotubes with hybridized phases for superior Na-ion storage. Chem. Eng. J.
394
(2018).
395
[5] N. Yabuuchi, K. Kubota, M. Dahbi, S. Komaba, Research development on
18
ACCEPTED MANUSCRIPT 396
sodium-ion batteries, Chem. Rev. 114 (2014) 11636–11682. [6] T. Chen, Y. Liu, L. Pan, T. Lu, Y. Yao, Z. Sun, D.H.C. Chua, Q. Chen,
398
Electrospun carbon nanofibers as anode materials for sodium ion batteries with
399
excellent cycle performance, J. Mater. Chem. A. 2 (2014) 4117.
401 402
[7] J.-M. Tarascon, Key challenges in future Li-battery research, Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 368 (2010) 3227–3241.
[8] J. Xie, L. Liu, J. Xia, Y. Zhang, Template-free synthesis of Sb2S3 hollow microspheres
as
anode
materials
for
404
batteries, Nano-Micro Lett. 10.1(2018):12.
Lithium-ion
and
Sodium-ion
M AN U
403
SC
400
RI PT
397
[9] F.F. Cao, Y.G. Guo, S.F. Zheng, X.L. Wu, L.Y. Jiang, R.R. Bi, L.J. Wan, J. Maier,
406
Symbiotic coaxial nanocables: Facile synthesis and an efficient and elegant
407
morphological solution to the lithium storage problem, Chem. Mater. 22 (2010)
408
1908–1914.
TE D
405
[10] Z. Yan, L. Liu, J. Tan, Q. Zhou, Z. Huang, One-pot synthesis of bicrystalline
410
titanium dioxide spheres with a core–shell structure as anode materials for
411
lithium and sodium ion batteries, J. Power Sources 269.269(2014):37-45.
AC C
EP
409
412
[11] L. Wu, D. Buchholz, D. Bresser, L. Gomes Chagas, S. Passerini, Anatase TiO2
413
nanoparticles for high power sodium-ion anodes, J. Power Sources. 251 (2014)
414
379–385.
415
[12] X. Yang, C. Wang, Y. Yang, Y. Zhang, X. Jia, J. Chen, X. Ji, Anatase TiO2
416
nanocubes for fast and durable sodium ion battery anodes, J. Mater. Chem. A. 3
417
(2015) 8800–8807.
19
ACCEPTED MANUSCRIPT 418
[13] H. Usui, S. Yoshioka, K. Wasada, M. Shimizu, H. Sakaguchi, Nb-doped rutile
419
TiO2: A potential anode material for Na-ion battery, ACS Appl. Mater. Interfaces.
420
7 (2015) 6567–6573. [14] M. Søndergaard, K.J. Dalgaard, E.D. Bøjesen, K. Wonsyld, S. Dahl, B.B. Iversen,
422
In situ monitoring of TiO2(B)/anatase nanoparticle formation and application in
423
Li-ion and Na-ion batteries, J. Mater. Chem. A. 3 (2015) 18667–18674.
RI PT
421
[15] J.C. Pérez-Flores, C. Baehtz, A. Kuhn, F. García-Alvarado, Hollandite-type TiO2 :
425
a new negative electrode material for sodium-ion batteries, J. Mater. Chem. A. 2
426
(2014) 1825–1833.
M AN U
SC
424
[16] D. Su, S. Dou, G. Wang, Anatase TiO2 : better anode material than amorphous
428
and rutile phases of TiO2 for Na-ion batteries, Chem. Mater. (2015)
429
150818142825005.
430
TE D
427
[17] YB. He, M. Liu, ZL. Xu, B. Zhang, B. Li, Li℃ion reaction to improve the rate performance
432
Techno. 1.11(2013):668-674.
434 435
nanoporous
anatase
TiO2
anodes. Energy
[18] Y. Kim, Y. Park, A. Choi, N.S. Choi, J. Kim, J. Lee, J.H. Ryu, S.M. Oh, K.T. Lee,
AC C
433
of
EP
431
An amorphous red phosphorus/carbon composite as a promising anode material for sodium ion batteries, Adv. Mater. 25 (2013) 3045–3049.
436
[19] X. Liu, T. Chen, H. Chu, L. Niu, Z. Sun, L. Pan, C.Q. Sun, Fe2O3-reduced
437
graphene oxide composites synthesized via microwave-assisted method for
438
sodium ion batteries, Electrochim. Acta. 166 (2015) 13–16.
439
[20] L. David, R. Bhandavat, G. Singh, MoS2/graphene composite paper for
20
ACCEPTED MANUSCRIPT 440
sodium-ion battery electrodes, ACS Nano. 8 (2014) 1759–1770. [21] Y. Liu, N. Zhang, L. Jiao, Z. Tao, J. Chen, Ultrasmall Sn nanoparticles embedded
442
in carbon as high-performance anode for sodium-ion batteries, Adv. Funct. Mater.
443
25 (2015) 214–220.
RI PT
441
[22] Y. Zhu, X. Han, Y. Xu, Y. Liu, S. Zheng, K. Xu, L. Hu, C. Wang, Electrospun
445
Sb/C fibers for a stable and fast sodium-ion battery anode, ACS Nano. 7 (2013)
446
6378–6386.
SC
444
[23] Y. Bai, Y. Liu, Y. Li, L. Ling, F. Wu, Mille-feuille shaped hard carbons derived
448
from polyvinylpyrrolidone via environmentally friendly electrostatic spinning for
449
sodium ion battery anodes, Rsc Adv. 7.9(2017):5519-5527.
M AN U
447
[24] Y. Xu, Y. Zhou, F. Han, 3D Si/C fiber paper electrodes fabricated using a
451
combined electrospray/electrospinning technique for Li℃ion batteries. Adv.
452
Energy Mater. 5.1(2015).
TE D
450
[25] X. Xia, X. Wang, H. Zhou, The effects of electrospinning parameters on coaxial
454
Sn/C nanofibers: Morphology and lithium storage performance, Electro.
455
Acta 121.3(2014):345-351.
AC C
EP
453
456
[26] L. Wang, Y. Yu, PC. Chen, Electrospinning synthesis of C/Fe3O4, composite
457
nanofibers and their application for high performance lithium-ion batteries, J.
458 459
Power Sources 183.2(2008):717-723.
[27] Y. Xiong, J. Qiao, Y. Chao, Electrospun TiO2/C Nanofibers As a high-capacity
460
and
cycle-stable
anode
461
Interfaces. 8.26(2016):16684.
for
sodium-ion
21
batteries. Acs
Appl.
Mater.
ACCEPTED MANUSCRIPT 462
[28] Z. Zhang, X. Li, C. Wang, Polyacrylonitrile and carbon nanofibers with
463
controllable
nanoporous
structures
464
Eng. 294.10(2009):673–678.
by
electrospinning, Macroml.
Mater.
[29] T. Chen, Y. Liu, L. Pan, Electrospun carbon nanofibers as anode materials for
466
sodium ion batteries with excellent cycle performance, J. Mater. Chem.
467
A 2.12(2014):4117-4121.
RI PT
465
[30] L. Wang, K. Zhang, Z. Hu, W. Duan, F. Cheng, J. Chen, Porous CuO nanowires
469
as the anode of rechargeable Na-ion batteries, Nano Res. 7 (2014) 199–208.
470
[31] L. Ji, Z. Lin, M. Alcoutlabi, X. Zhang, Recent developments in nanostructured
471
anode materials for rechargeable lithium-ion batteries, Energy Environ. Sci. 4
472
(2011) 2682.
M AN U
SC
468
[32] Y.E. Zhu, L. Yang, J. Sheng, Y. Chen, H. Gu, J. Wei, Z. Zhou, Fast sodium
474
storage in TiO2@CNT@C nanorods for high-performance Na-ion capacitors,
475
Adv. Energy Mater. 7 (2017) 1–9.
TE D
473
[33] Y. Mao, H. Duan, B. Xu, L. Zhang, Y. Hu, C. Zhao, Z. Wang, L. Chen, Y. Yang,
477
Lithium storage in nitrogen-rich mesoporous carbon materials, Energy Environ.
AC C
478
EP
476
Sci. 5 (2012) 7950–7955.
479
[34] Y. Zhong, M. Yang, X. Zhou, Y. Luo, J. Wei, Z. Zhou, Orderly packed anodes for
480
high-power lithium-ion batteries with super-long cycle life: Rational design of
481
MnCo3/large-area graphene composites, Adv. Mater. 27 (2015) 806–812.
482
[35] L. Sun, C. Tian, Y. Fu, Y. Yang, J. Yin, L. Wang, H. Fu, Nitrogen-doped porous
483
graphitic carbon as an excellent electrode material for advanced supercapacitors,
22
ACCEPTED MANUSCRIPT 484
Chem. - A Eur. J. 20 (2014) 564–574. [36] H. Wang, P. Gao, S. Lu, H. Liu, G. Yang, J. Pinto, X. Jiang, The effect of tin
486
content to the morphology of Sn/carbon nanofiber and the electrochemical
487
performance as anode material for lithium batteries, Electrochim. Acta. 58 (2011)
488
44–51.
RI PT
485
[37] N. Chen, D. Deng, Y. Li, X. Xing, X. Liu, X. Xiao, Y. Wang, The xylene sensing
490
performance of WO3 decorated anatase TiO2 nanoparticles as a sensing material
491
for a gas sensor at a low operating temperature, RSC Adv. 6 (2016)
492
49692–49701.
M AN U
SC
489
[38] Y.L. Huang, S.M. Yuen, C.C.M. Ma, C.Y. Chuang, K.C. Yu, C.C. Teng, H.W.
494
Tien, Y.C. Chiu, S.Y. Wu, S.H. Liao, F.B. Weng, Morphological, electrical,
495
electromagnetic interference (EMI) shielding, and tribological properties of
496
functionalized multi-walled carbon nanotube/poly methyl methacrylate (PMMA)
497
composites, Compos. Sci. Technol. 69 (2009) 1991–1996.
TE D
493
[39] L. Shao, S. Quan, Y. Liu, Z. Guo, Z. Wang, A novel “gel-sol” strategy to
499
synthesize TiO2 nanorod combining reduced graphene oxide composites, Mater.
AC C
500
EP
498
Lett. 107 (2013) 307–310.
501
[40] S. Anwer, Y. Huang, J. liu, J. Liu, M. Xu, Z. Wang, R. Chen, J. Zhang, F. Wu,
502
Nature-inspired Na2Ti3O7 nanosheets-formed three-dimensional microflowers
503
architecture as a high-performance anode material for rechargeable sodium-ion
504
batteries, ACS Appl. Mater. Interfaces. 9 (2017) 11669–11677.
505
[41] H. Wang, Y. Liu, M. Li, H. Huang, H.M. Xu, R.J. Hong, H. Shen,
23
ACCEPTED MANUSCRIPT 506
Multifunctional TiO2 nanowires-modified nanoparticles bilayer film for 3D
507
dye-sensitized solar cells, Optoelectron. Adv. Mater. Rapid Commun. 4 (2010)
508
1166–1169. [42] R. Dai, W. Sun, Y. Wang, Ultrasmall tin nanodots embedded in nitrogen-doped
510
mesoporous carbon: metal-organic-framework derivation and electrochemical
511
application as highly stable anode for lithium ion batteries, Electrochim. Acta.
512
217 (2016) 123–131.
SC
RI PT
509
[43] H. Ying, S. Zhang, Z. Meng, Z. Sun, W. Q. Han, Ultrasmall Sn nanodots
514
embedded inside N-doped carbon microcages as high-performance lithium and
515
sodium ion battery anodes, J. Mater. Chem. A. 5 (2017) 8334–8342.
M AN U
513
[44] J. Xia, L. Liu, J. Xie, H. Yan, Y. Yuan, M. Chen, C. Huang, Y. Zhang, S. Nie, X.
517
Wang, Layer-by-layered SnS2/graphene hybrid nanosheets via ball-milling as
518
promising anode materials for lithium ion batteries, Electrochim. Acta. 269
519
(2018) 452–461.
522 523
EP
521
[45] M. Sha, H. Zhang, Y. Nie, K. Nie, X. Lv, N. Sun, X. Xie, Y. Ma, X. Sun, Sn nanoparticles@nitrogen-doped
carbon
nanofiber
composites
as
AC C
520
TE D
516
high-performance anodes for sodium-ion batteries, J. Mater. Chem. A. 5 (2017) 6277–6283.
524
[46] Y. Xu, Y. Zhou, Y. Liu, CS. Wang, Electrochemical Performance of Porous
525
Carbon/Tin Composite Anodes for Sodium-Ion and Lithium-Ion Batteries, Adv.
526
Energy Mater. 3.1(2013):128-133.
527
[47] L. Fan, J. Zhang, J. Cui, Electrochemical performance of rod-like Sb-C
24
ACCEPTED MANUSCRIPT 528
composite as anodes for Li-ion and Na-ion batteries, J. Mater. Chem.
529
A 3.7(2015):3276-3280. [48] L. Ling, Y. Bai, Y. Li, Q. Ni, Z. Wang, F. Wu, C. Wu, Quick activation of
531
nanoporous anatase TiO2 as high-rate and surable anode materials for sodium-ion
532
batteries, ACS Appl. Mater. Interfaces. 9 (2017) 39432–39440.
RI PT
530
[49] J. Wang, G. Liu, K. Fan, D. Zhao, B. Liu, J. Jiang, D. Qian, C. Yang, J. Li,
534
N-doped carbon coated anatase TiO2 nanoparticles as superior Na-ion battery
535
anodes, J. Colloid Interface Sci. 517 (2018) 134–143.
M AN U
SC
533
536
[50] X. Shi, Z. Zhang, K. Du, Y. Lai, J. Fang, J. Li, Anatase TiO2@C composites with
537
porous structure as an advanced anode material for Na ion batteries, J. Power
538
Sources. 330 (2016) 1–6.
[51] Y. Xu, E. Memarzadeh Lotfabad, H. Wang, B. Farbod, Z. Xu, A. Kohandehghan,
540
D. Mitlin, Nanocrystalline anatase TiO2: a new anode material for rechargeable
541
sodium ion batteries, Chem. Commun. 49 (2013) 8973.
TE D
539
[52] J. Chen, G. Zou, H. Hou, Y. Zhang, Z. Huang, X. Ji, Pinecone-like hierarchical
543
anatase TiO2 bonded with carbon enabling ultrahigh cycling rates for sodium
AC C
544
EP
542
storage, J. Mater. Chem. A. 4 (2016) 12591–12601.
545
[53] K.T. Kim, G. Ali, K.Y. Chung, C.S. Yoon, H. Yashiro, Y.K. Sun, J. Lu, K. Amine,
546
S.T. Myung, Anatase titania nanorods as an intercalation anode material for
547
rechargeable sodium batteries, Nano Lett. 14 (2014) 416–422.
548
[54] Y. Zhang, Y. Yang, H. Hou, X. Yang, J. Chen, M. Jing, X. Jia, X. Ji, Enhanced
549
sodium storage behavior of carbon coated anatase TiO2 hollow spheres, J. Mater.
25
ACCEPTED MANUSCRIPT 550
Chem. A. 3 (2015) 18944–18952. [55] J. Xie, Y. Pei, L. Liu, S. Guo, J. Xia, M. Li, Y. Ouyang, X. Zhang, X. Wang,
552
Hydrothermal synthesis of antimony oxychlorides submicron rods as anode
553
materials for lithium-ion batteries and sodium-ion batteries, Electrochim. Acta.
554
254 (2017) 246–254.
RI PT
551
[56] L. Yi, L. Liu, G. Guo, X. Chen, Y. Zhang, S. Yu, X. Wang, Expanded
556
graphite@SnO2@polyaniline composite with enhanced performance as anode
557
materials for lithium ion batteries, Electrochim. Acta. 240 (2017) 63–71.
M AN U
558
SC
555
[57] L.-L. Zhang, S. Duan, X.-L. Yang, G. Peng, G. Liang, Y.-H. Huang, Y. Jiang, S.-B. Ni, M. Li, Reduced graphene oxide modified Li2FeSiO4/C composite with
560
enhanced electrochemical performance as cathode material for lithium ion
561
batteries, ACS Appl. Mater. Interfaces. 5 (2013) 12304–9.
AC C
EP
TE D
559
26
ACCEPTED MANUSCRIPT Table 1. Analysis of the content of Sn and Ti in different samples by ICP-OES. Sn
Ti
TiO2-Sn/C NFs (0.05 g)
3.32 wt%
33.79 wt%
TiO2-Sn/C NFs (0.1 g)
6.08 wt%
32.97 wt%
TiO2-Sn/C NFs (0.2 g)
10.11 wt%
RI PT
Sample
34.98 wt%
TiO2-Sn/C NFs
6.7
TiO2/C NFs
7.4
Rct (Ω)
M AN U
Rs (Ω)
AC C
EP
TE D
Samples
SC
Table 2. EIS parameters of TiO2-Sn/C NFs and TiO2/C NFs samples.
1
DNa+ (cm2 s-1)
242.8
6.9×10-13
297.1
2.4×10-13
EP
TE D
M AN U
SC
Graphical abstract
RI PT
ACCEPTED MANUSCRIPT
AC C
Figure 1. (a) XRD pattern of TiO2/C NFs; (b) XRD pattern of TiO2-Sn/C NFs; (c) Raman spectra of TiO2/C NFs and TiO2-Sn/C NFs; (d) TG curves of TiO2/C NFs and TiO2-Sn/C NFs under air atmosphere.
1
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
Figure 2. (a)(b) SEM images of TiO2/C NFs; (c)(d) SEM images of TiO2-Sn/C NFs.
2
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
Figure 3. (a) TEM image of TiO2/C NFs; (b)(c) TEM image of TiO2-Sn/C NFs; (d)
EP
TEM elemental mapping images of C, O, Sn and Ti elements of TiO2-Sn/C NFs; (e)
AC C
HRTEM image of TiO2-Sn/C NFs with lattice spacing; (f) Interplanar crystal spacing statistical tables.
3
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
Figure 4. (a) XPS wide spectrum of TiO2/C NFs and TiO2-Sn/C NFs; High-resolution (b) C 1s and (c) Ti 2p XPS spectra of TiO2/C NFs and
AC C
EP
TE D
TiO2-Sn/C NFs; (d) High-resolution Sn 3d spectra of TiO2-Sn/C NFs.
4
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
Figure 5. (a) Cycling performance of TiO2/C NFs and TiO2-Sn/C NFs electrode under a current density of 1 A g-1; (b) Continuous discharge and charge curves of TiO2-Sn/C NFs electrode under a current density of 1A g-1; (c) Rate
capability of TiO2/C NFs and TiO2-Sn/C NFs; (d) Charge-discharge curves of TiO2-Sn/C NFs at 0.05-5 A g-1 in the range of 0.01-2.5 V; (e) Charge-discharge curves of TiO2-Sn/C NFs from the second cycle to the fifth
cycle at a current density of 0.05 A g-1; (f) The CV plots of TiO2-Sn/C NFs and
5
ACCEPTED MANUSCRIPT TiO2/C NFs for the second cycle at 0.1 mV s-1; (g) The long cycle performance
M AN U
SC
RI PT
of TiO2-Sn/C NFs under the high current density of 5 A g-1.
Figure 6. Comparison of the electrochemical performance between TiO2-Sn/C NFs
AC C
EP
TE D
with other reported TiO2 in studies as anode materials for NIBs.
Figure 7. (a) Nyquist plots measured for the TiO2-Sn/C NFs and TiO2/C NFs electrodes after first cycle at 0.05 A g-1; (b) The relationship plot between Z' and ω-1/2 of the low-frequency region.
6