sulfur co-doped carbon as high performance anode materials for lithium-ion batteries

sulfur co-doped carbon as high performance anode materials for lithium-ion batteries

Accepted Manuscript Ultrahigh level nitrogen/sulfur co-doped carbon as high performance anode materials for lithium-ion batteries Zhaozheng Qiu, Yemao...

5MB Sizes 0 Downloads 50 Views

Accepted Manuscript Ultrahigh level nitrogen/sulfur co-doped carbon as high performance anode materials for lithium-ion batteries Zhaozheng Qiu, Yemao Lin, Hailin Xin, Pei Han, Dongzhi Li, Bo Yang, Pengchong Li, Shahid Ullah, Haosen Fan, Caizhen Zhu, Jian Xu PII:

S0008-6223(17)30983-1

DOI:

10.1016/j.carbon.2017.09.100

Reference:

CARBON 12430

To appear in:

Carbon

Received Date: 27 July 2017 Revised Date:

24 September 2017

Accepted Date: 28 September 2017

Please cite this article as: Z. Qiu, Y. Lin, H. Xin, P. Han, D. Li, B. Yang, P. Li, S. Ullah, H. Fan, C. Zhu, J. Xu, Ultrahigh level nitrogen/sulfur co-doped carbon as high performance anode materials for lithium-ion batteries, Carbon (2017), doi: 10.1016/j.carbon.2017.09.100. 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.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

Ultrahigh level nitrogen/sulfur co-doped carbon as high

2

performance anode materials for lithium-ion batteries

3

Zhaozheng Qiu,a, # Yemao Lin,a, # Hailin Xin,a Pei Han,a Dongzhi Li,a Bo Yang,a Pengchong Li,a

4

Shahid Ullah,a Haosen Fan,a Caizhen Zhu,a* and Jian Xub

AC C

EP

TE D

M AN U

SC

RI PT

1

* Corresponding author. E-mail address: [email protected] (C. Zhu). #

These authors contributed equally to this work.

ACCEPTED MANUSCRIPT 5

a

6

PR China

7

b

8

Sciences, Beijing, 100190, PR China

College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen, 518060,

RI PT

Beijing National Laboratory for Molecular Sciences Institute of Chemistry, Chinese Academy of

9

Abstract

11

Ultrahigh level nitrogen and sulfur co-doped disordered porous carbon (NSDPC) was

12

facilely synthesized and applied as anode materials for lithium-ion batteries (LIBs).

13

Benefiting from high nitrogen (14.0 wt%) and sulfur (21.1 wt%) doping, electrode

14

fabricated from NS1/3 showed a high reversible capacity of 1188 mA h g−1 at 0.1 A g−1

15

in the first cycle with a high initial columbic efficiency (>75 %). In addition,

16

prolonged life over 500 cycles and excellent rate capability of 463 mA h g−1 at 5 A g−1

17

have been realized. The preeminent electrochemical performance is attributed to three

18

effects: (1) the high level of sulfur and nitrogen; (2) the synergic effect of dual-doping

19

heteroatoms in cooperation with each other; (3) the large quantity of edge defects and

20

abundant micropores and mesopores that can provide extra Li storage regions. These

21

unique features of NSDPC electrodes suggest that they can serve as a practical

22

substitute for graphite as a high performance anode material in LIBs.

23

Keywords: Electrode materials; Co-doping; Porous carbon; Anode; Lithium

AC C

EP

TE D

M AN U

SC

10

24

25

1. Introduction

ACCEPTED MANUSCRIPT With the rapid development of portable devices, renewable energy harvesting, and

27

electric vehicles[1, 2], safe energy storage devices with improved power and energy

28

densities are becoming increasingly important[3]. Among these devices, lithium-ion

29

batteries (LIBs) are promising choices for their high energy density and power density,

30

additionally with long lifespans[4]. However, application of LIBs in large-scale

31

electric energy storage requires further improvements in energy density, rate

32

capability, safety, and electrode durability. To a large extent, these issues rely on

33

several determinants. One of them is the development of novel anode materials.

34

Graphite is the main commercial anode material used in currently available LIBs. But

35

the low theoretical specific capacity of 372 mA h g−1 cannot meet the increasing

36

demands of rapidly developing markets[5]. Therefore, it’s urgently to develop new

37

anode materials with high theoretical capacities and excellent rate and cycling

38

performance for the next-generation LIBs.

39

Various nanostructured products containing Si[6-8], P[9], or Sn[10-12], as well as

40

alloy[13, 14] and some transition metal oxides/sulfide[15-23], have been reported as

41

high capacity anode materials for LIBs. Unfortunately, intrinsic problems such as

42

huge volume expansion, relatively low conductivity, and large voltage hysteresis,

43

exist in these materials during the lithiation process[24]. These deficiencies severely

44

limit their commercialization in LIBs. Therefore, novel carbon-based anode materials,

45

with enhanced electrochemical performance for lithium storage, are still a major focus

46

of research worldwide[25]. Applied them as anode materials, high energy LIBs with

47

elevated specific capacity, excellent cycling stability and rate performance have been

AC C

EP

TE D

M AN U

SC

RI PT

26

ACCEPTED MANUSCRIPT realized. Significant efforts have been made to design novel nanostructures of

49

carbonaceous materials with larger specific capacity for lithium storage [25-27].

50

However, for amorphous carbon materials, although activating agents (such as KOH)

51

has been applied to increase the specific surface area, inferior stability and poor rate

52

performance still occur due to the low degree of graphitization[28]. Therefore, a novel

53

strategy of artificial doping with heteroatoms, such as N[26], B[29] or S[30], have

54

being proposed. With this method, the electronic properties and electrochemical

55

activity of amorphous carbon materials can be effectively improved, leading to

56

excellent Li ion storage capacity. Nevertheless, most reported results related to this

57

field are based on the single element doping but co-doping and their synergistic

58

influence to the electrochemical property is less studied.

59

Although nitrogen doped carbon-based materials have been extensively studied, sulfur

60

and especially nitrogen/sulfur dual-doping are far less exhaustive[31]. When doped

61

with nitrogen the electronic properties of carbon-based materials can be improved[32].

62

As a complement to nitrogen, sulfur doping has attracted increasing attention in recent

63

carbon materials research. While sulfur has a larger atomic radius, it can enlarged the

64

interlayer spacing(d002) of the carbon matrix[31]. The doped sulfur substantially

65

increases the charge capacity with the enlarged graphite crystallite size (Lc) and

66

creates more micropores, improving the electrochemical properties of the carbon[30].

67

Meanwhile, its easily polarizable lone pairs can change the charge state of

68

neighboring carbon atoms. Hence, it can be used to tune chemical reactivity and

69

catalytic activity of the carbon materials. Therefore, as a doping heteroatom, sulfur

AC C

EP

TE D

M AN U

SC

RI PT

48

ACCEPTED MANUSCRIPT can be incorporated into carbon-based anode materials, improving their reversible

71

capacity in LIBs[33]. Importantly, when both highly interactive elements of nitrogen

72

and sulfur are simultaneously doped into the carbon matrix, synergistic effects can be

73

aroused[34]. In addition, it has been reported that the binding state of sulfur and

74

nitrogen in the carbon matrix can be tuned when different sulfur and nitrogen sources

75

were used[31]. Most studies property on sulfur/nitrogen dual-doped carbon materials

76

is about their enhanced electrocatalyst ability[31, 35]. Few reports focus on

77

dual-doped carbons for LIBs. Therefore, we were inspired to investigate the favorable

78

influences of dual-doped heteroatoms in LIBs systems.

M AN U

SC

RI PT

70

Herein, we applied a “dual-doping” strategy to synthesize an amorphous porous

80

carbon with nitrogen and sulfur at an ultrahigh doping level. In this way, the

81

synergistic effects of heteroatom co-doping and a disordered porous structure has

82

been achieved. The nitrogen and sulfur co-doped disordered porous carbon (NSDPC)

83

was employed as a preeminent anode material for high-performance LIBs. By a facile

84

synthesis method, an amorphous porous structure was formed. In the sample of NS1/3,

85

micropores and mesopores were homogenously embedded with a nitrogen content of

86

14.0 wt% and a sulfur content of 21.1 wt%. Benefitting from the convenient transport

87

pathway for Li-ions and electron, as well as the abundant pores for extra Li storage in

88

the carbon framework, the NS1/3 electrode exhibits excellent rate performance with

89

superior cycling stability. The reversible specific capacity achieved as high as 1188

90

mA h g−1 at a current density of 0.1 A g−1, and even after increasing the rate to 5 A g−1

91

high capacity of 463 mA h g−1 can still be obtained, the capacity still remains at 653

AC C

EP

TE D

79

ACCEPTED MANUSCRIPT mA h g−1 after 500 cycles at 1A g−1.

93

2. Experimental Section

94

2.1. Synthesis of cystine aggregates

95

10 mM L-Cysteine was dissolved in 100 mL ultrapure water (18.0 MΩ cm−1) under

96

ultrasonication for about 10 min. The pH value of the L-Cysteine aqueous solution

97

was rapidly adjusted to 8.0 using Na2CO3 aqueous solution. 5 mL H2O2 solution (30

98

wt%) was added drop by drop slowly to the L-Cysteine solution with a syringe (10

99

mL). The whole reaction process was kept at room temperature with stirring. After 10

100

min, the resulting solution was incubated at room temperature without interruption for

101

24 h. The resulting precipitation was centrifuged and washed several times with water

102

and ethanol and finally the white powder of cystine aggregates were obtained after

103

dried under oven at 60 °C for 24 h.

104

2.2. Preparation of NSDPC

105

In order to adjust the elemental composition of the obtained NSHPC, varied

106

concentrations of melamine and sulfur were added to a constant amount of cystine

107

aggregates harvested in front. The nitrogen/sulfur co-doped carbon was synthesized

108

by the following synthetic route. In a typical experiment: 8.0 g of cystine aggregates

109

and x g (x = 3.0, 2.4, 2.0, 1.6, or 1.0) of melamine and y g (y = 1.0, 1.6, 2.0, 2.4, or 3

110

corresponding to x) of sulfur were mixed thoroughly. Followed by annealing at

111

500 °C for 2 hours in the atmosphere of Ar with a heating ramp of 2 °C min−1 and the

112

obtained sample were labeled according to the following scheme: NSX/Y where X and

113

Y denotes the integer proportion of x and y. NS3/1 hence corresponds to an experiment

AC C

EP

TE D

M AN U

SC

RI PT

92

ACCEPTED MANUSCRIPT in which 3.0 g of melamine and 1.0 g sulfur were added to 8.0 g of cystine aggregates.

115

2.3. Material characterization

116

The powder X-ray diffraction (XRD) patterns of all samples were recorded with an

117

X-ray diffractometer (D8 Advance of Bruker, Germany) with filtered Cu Kα radiation

118

over the 2θ range of 10-60°. Field emission scanning electron microscopy (FE-SEM)

119

images were collected on a JSM-7800F scanning electron microscope. The amounts

120

of doped nitrogen and sulfur in the synthesized materials were determined by a

121

CHNOS Elemental Analyzer (vario EL cube, Germany). Transmission electron

122

microscopy (TEM) images were taken on a JEM-2100 transmission electron

123

microscope using an accelerating voltage of 200 kV, and high-resolution transmission

124

electron microscope (HRTEM) (JEOL-2011) was operated at an acceleration voltage

125

of 200 kV. The specific surface area was evaluated at 77 K (Quantachrome

126

NOVA1200e) using the Brunauer-Emmett-Teller (BET) method, while the pore

127

volume and pore size were calculated according to the Barrett-Joyner-Halenda (BJH)

128

formula applied to the adsorption branch. Thermogravimetric analysis (TGA) was

129

carried out using a STA409PC from 0 °C to 800 °C at a heating rate of 10 °C min−1

130

under Ar. X-ray Photoelectron Spectrum (XPS) was performed on an ESCALAB 250

131

X-ray Photoelectron Spectrometer with Al Ka radiation. Raman spectra were obtained

132

using a Digilab FTS3500 from Bio-Rad with a laser wavelength of 632.8 nm.

133

2.4. Electrochemical measurements

134

The electrochemical performance of the NSDPC was examined by using CR 2032

135

coin-type cells which assembled in an argon-filled glovebox, using lithium foils as the

AC C

EP

TE D

M AN U

SC

RI PT

114

ACCEPTED MANUSCRIPT counter electrodes, Celgard 2500 membrane as the separator, 1 M LiPF6 in a 1:1 (v/v)

137

mixture of ethylene carbonate and dimethyl carbonate as the electrolyte. To prepare a

138

working electrode, active materials (NSDPCs, 70 wt%), conductive material

139

(acetylene black, 20 wt%), and binder (Sodium Alginate (SA), 10 wt%) were milled

140

in ultrapure water to form slurries and then coated onto the surface of a copper foil

141

current collector. The electrode capacity was measured by a galvanostatic discharge–

142

charge method in the voltage range between 0.01 and 3.0 V on a battery test system

143

(Land CT2001A, China). Cyclic voltammetry (CV) from 0.01 to 3.0 V (vs. Li/Li+) at

144

0.1 mV s−1 and electrochemical impedance spectroscopy (EIS) in the frequency range

145

100 MHz to 0.01 Hz and with an amplitude of 5 mV were performed using an

146

electrochemical workstation (CHI660A). All electrochemical measurements were

147

carried out at room temperature.

148

3. Results and Discussion

149

3.1. Synthesis and characterizations of the NSDPC

150

It has been demonstrated that H2O2 can oxidize the –SH group of cysteine to form

151

disulfide bonds and lead to the formation of cystine[36]. The development of cystine

152

aggregates is a result of intermolecular interactions, mainly hydrogen bonding and

153

electrostatic interactions[37]. The morphology of the obtained cystine aggregates was

154

thin slice shape hexagonal crystal (Fig. 1S). Fig. 1 shows the illustration of the

155

synthesis of NSDPC. NSDPC was synthesized facilely by annealing the mixture of

156

as-prepared cystine aggregates, melamine, and sulfur in a flowing atmosphere of Ar at

157

500 °C for 2 h. The cystine aggregates with N content of 11.6 wt% and S content of

AC C

EP

TE D

M AN U

SC

RI PT

136

ACCEPTED MANUSCRIPT 26.4% were selected as the substrate material. Melamine with a high N content (66.7

159

wt%) and sulfur were used as excess N and S sources, respectively.

160

Thermogravimetric analysis (TGA) and differential thermal analysis (DTA) of the

161

mixture of as-prepared cystine aggregates, melamine, and sulfur (mass ratio of 8: 1: 3)

162

were performed in an Ar atmosphere. It indicated that the carbon sulfurization

163

temperature was approximately 300 °C (Fig. 2a). When the temperature was increased

164

to 500 °C, the carbon sulfurization and carbonization of the as-prepared cystine

165

aggregates and melamine were accomplished, and the excess sulfur was evaporated

166

off.

169

SC

M AN U TE D

168

Fig. 1. The schematic of the synthesis of NSDPC composites

EP

167

RI PT

158

The results obtained from XRD measurements are summarized in Fig. 2b and Table 1.

171

Samples of NS3/1, NS3/2, NS3/3, NS2/3, and NS1/3 exhibited broad peaks at around 23.5°

172

and weak peaks at 43°, which were readily indexed to the representative (002) and

173

(101) planes of amorphous carbon, respectively. According to Bragg’s equation, the

174

interlayer spacing (d002) of the five samples was calculated to be approximately 0.38

175

nm. The larger d002 values of the NSDPC materials in comparison with that of

176

graphite (0.335 nm) implied that the intercalation of sulfur result in an enlarged

AC C

170

ACCEPTED MANUSCRIPT 177

interlayer distance of porous carbon, facilitating the diffusion and insertion/extraction

178

of Li+ in carbon matrix[38]. Table 1. Summary of the d-spacing of the NSDPC samples

180 181 182

NS3/1

NS3/2

NS3/3

NS2/3

NS1/3

23.29

23.66

23.35

23.23

23.66

d002(nm) 0.381

0.376

0.380

0.382

2θ (°)

183

RI PT

179

0.376

To further examine the structural characteristic, the five samples were measured using

185

N2 adsorption–desorption isotherms. As displayed in Fig. 2c and Table S1 (in

186

Supplementary Material), the specific Brunauer-Emmett-Teller (BET) surface area of

187

the NS3/1 is 34.6 m2 g−1. Whereas the specific surface areas of the NS3/2, NS3/3 and

188

NS2/3 are increased significantly to 42.8, 129.2 and 249.6 m2 g−1, respectively. As for

189

the sample of NS1/3, it reaches the maximum of 341.0 m2 g−1 at the ratio of

190

melamine(X)/sulfur(Y) reach 1: 3. Moreover, NS1/3 possesses the highest total pore

191

volume, being 0.474 cm3 g−1 (Table S1). This demonstrates that the effectiveness of a

192

suitable ratio of melamine and sulfur in developing more pores in NSDPC skeleton.

193

The NSDPC samples showed mesopores are dominant in the samples, which covered

194

a range between 2∼10 nm (Fig. 2d).

AC C

EP

TE D

M AN U

SC

184

SC

RI PT

ACCEPTED MANUSCRIPT

M AN U

195

Fig. 2. (a) TG and DTA curves of the mixture of cystine aggregates, melamine, and

197

sulfur with a mass ratio of 8: 1: 3 in the temperature range of 20–800 °C with a

198

heating rate of 10 °C min−1 in an Ar atmosphere; (b) XRD patterns of NSDPC

199

samples over the 2θ range of 10–60°; (c) N2 adsorption–desorption isotherm and (d)

200

BJH pore-size distribution of NSDPC samples.

201

TE D

196

The morphologies of the as-prepared NSDPC samples were characterized by

203

field-emission scanning electron microscopy (FESEM) as shown in Fig. 3a–f. The

204

SEM images of the resulting carbon materials exhibited flake-like morphology. Many

205

interconnected pores existed in the carbon sheets, especially for the NS1/3 sample.

206

High magnification observation of the sheets revealed that the sheets had numerous

207

nanoscale pores within them. This 3D structure has many interpenetrative channels

208

with large internal spaces, which are mainly formed by the volatilization of excess

209

sulfur, H2O, CO2, H2S, NH3, and CO during the carbonization process[38]. The novel

210

structure in this carbon framework ensures facile access of electrolytes, efficient

AC C

EP

202

ACCEPTED MANUSCRIPT transfer of lithium ions and supply additional lithium storage space, fascinating rapid

212

charging/discharging of the electrode. The HRTEM image (Fig. 3g) reveals the rough

213

surface of the amorphous carbon with nanopores. TEM images further confirm the

214

interconnected porous structure. TEM elemental mapping of the NSDPCs shown in

215

Fig. 3g indicates that the distribution of C, N, and S is very homogenous.

M AN U

SC

RI PT

211

217

AC C

EP

TE D

216

218

Fig. 3. SEM images of NSDPC samples: (a) NS3/1, (b) NS3/2, (c) NS3/3, (d) NS2/3, and

219

(e, f) NS1/3; (g) HRTEM images and TEM-EDS elemental mapping of NS1/3.

220 221

Elemental analysis (EA) was employed to quantify the N and S doping levels in

222

carbon bulk. The results demonstrated that the contents of N and S in the NSDPC

223

materials could be tailored by varying the ratio of melamine and sulfur (Table 2).

ACCEPTED MANUSCRIPT Pristine NSDPC materials derived from cystine aggregates showed a substantial

225

presence of nitrogen and sulfur (10.2 wt% and 7.6 wt%, respectively). N and S

226

concentration increased drastically with additional melamine and sulfur. For example,

227

the N and S levels reached to 16.3 wt% (N) and 13.1 wt% (S) with the ratio of

228

melamine and sulfur to be 3: 1. When the ratio was changed to 1: 3, the content of N

229

decreased to 14.0 wt%, while the content of S increased to 21.1 wt%.

RI PT

224

230

SC

Table 2. Elemental composition of the NSDPC samples Sample C (%) S (%) N (%) O (%) H (%) S/N NS 61.1 7.6 10.2 18.8 2.3 0.75 NS3/1 57.8 13.1 16.3 10.5 2.2 0.80 NS3/2 59.0 13.2 15.6 9.9 2.2 0.85 59.7 14.4 13.4 10.3 2.3 1.07 NS3/3 NS2/3 56.8 18.0 13.5 9.7 2.1 1.33 54.1 21.1 14.0 9.0 1.9 1.51 NS1/3

M AN U

231

TE D

232

Raman spectroscopy was carried out to examine the degree of graphitization and

234

imperfections in the obtained carbon samples (Fig. 4a–f). Two characteristic peaks at

235

approximately 1403 and 1560 cm−1 correspond to the D and G peaks, respectively.

236

The D peak indicate to structure defects and disordered structure in the graphite layer

237

due to the co-doping of nitrogen and sulfur. While the G peak is related to the sp2 C–

238

C bonds of perfect graphite layer[27]. The intensity ratio of the G to D band (ID/IG) is

239

generally employed to evaluate the extent of structural disorder for the carbon

240

materials[39]. The ID/IG ratios for NS3/1, NS3/2, NS3/3, NS2/3, and NS1/3 were 1.322,

241

1.393, 1.440, 1.477, and 1.558, respectively. Obviously, an increasing trend in the

242

ID/IG ratio was observed after more S was doped. The NS1/3 sample showed the

AC C

EP

233

ACCEPTED MANUSCRIPT highest ID/IG ratio (1.558), indicating that higher S-doping lead to more defect and

244

disorder in the carbon framework. The results agree with the previous report on

245

sulfur-doped porous carbon[40]. The disordered structure and numerous defects are

246

supposed to be benefit to the insertion/extraction and additional storage of Li ions[41].

247

As shown in Fig. 4f, G band has a trend of shifted down as the ratio of N and S

248

precursor adjust from 3: 1 to 1: 3, which can be attributed to the recovery of the

249

conjugated structure[42] and the electron donation of heteroatoms[43].

250

TE D

M AN U

SC

RI PT

243

Fig. 4. Raman spectra at the D (disordered) band and G (graphitic) band of (a) NS3/1,

252

(b) NS3/2, (c) NS3/3, (d) NS2/3, and (e) NS1/3; (f) spectra of the NSDPC samples.

AC C

253

EP

251

254

X-ray photoelectron spectroscopy (XPS) was used to investigate the chemical status

255

of the elements and the surface chemistry of the NSDPCs, as shown in Fig. 5. The

256

XPS survey spectrum shows peaks for C1s (284.6 eV), N1s (400.5 eV), O1s (531.5

257

eV), and S2p (164.5 eV) (Fig. 5a), confirming the existence of elements detected by

258

EA and the elemental mapping results (Fig. 3g). The N and S concentrations (at%) on

259

the surface of each NSDPC sample are shown in Table S2(in Supplementary Material).

ACCEPTED MANUSCRIPT The sulfur content of 8.1 at% (equivalent to 18.4 wt% in Table S3) in NS1/3 calculate

261

from the XPS data was generally in concordance with the EA analysis result (21.1

262

wt%) (Table 2), indicating uniform distribution of heteroatoms in the bulk materials.

263

The high-resolution C1s spectrum was fitted by two sub-peaks, including C–C (ca.

264

284.0 eV), and C–S/C–N (ca. 285.5 eV) (Fig. 5b) [26, 27, 44]. The deconvoluted XPS

265

spectrum of N1s contains two characteristic sub-peaks at 397.8, and 399.6 eV (Fig.

266

5c), corresponding to pyridinic-type, and pyrrolic-type, respectively (Table S4) [45].

267

The pyridinic and pyrrolic nitrogen species can improve the conductivity in the

268

carbon matrix, thanks to their p-electron pair that donate to the π-conjugated system in

269

the graphene layers[46, 47]. The high resolution S2p spectrum is shown in Fig. 5d,

270

which exhibits three sub-peaks at binding energies of 163.0, 164.0, and 167.8 eV

271

(Table S4). The former two can be attributed to the S2p3/2 and S2p1/2 of the –C–Sx–C–

272

(x =1–2) covalent bond of the thiophene-S, while the weak peak at 167.8 eV

273

corresponds to C–SOx–C (x = 2–4) groups[38].

AC C

EP

TE D

M AN U

SC

RI PT

260

274 275

Fig. 5. (a) Full-scan XPS of NSDPCs and (b–d) high resolution C1s, N1s, and S2p

ACCEPTED MANUSCRIPT 276

XPS scans of NS1/3. XPS N1s, and S2p spectra for the other samples are provided in

277

the ESI. (High resolution N1s, and S2p XPS scans of the other four samples were

278

showed in Fig. S2 in Supplementary Material.)

RI PT

279

3.2. Electrochemical performances

281

The cyclic voltammetry (CV) profile of the obtained NS1/3 electrode in the initial

282

three cycles at a scan rate of 0.1 mV s−1 is shown in Fig. 6a. In the first cycle, a

283

prominent reduction peak appear at 0.65 V, and almost disappeared in the following

284

cycles, indicating decomposition of the electrolyte and the formation of a solid

285

electrolyte interface (SEI) film on the electrode surface[33]. The peak near 0 V

286

corresponds to the insertion of Li-ion into carbonaceous materials, consistent with that

287

of a reported graphene anode[48]. The oxidation peak located at 2.34 V can be

288

attribute to the transformation of LixS into polysulfides[49]. In the subsequent cycles,

289

the CV curves almost overlapped, implying excellent cycling stability of the electrode.

290

Correlative plateau regions can be observed in the charge-discharge profiles (Fig. 6b).

291

The charge/discharge curves of the NS1/3 electrode was tested using a constant current

292

density of 0.1 A g−1. The first discharge and charge profiles of it showed high specific

293

capacities of 1572 and 1188 mA h g−1, respectively, with an initial coulombic

294

efficiency of 75.6%. It was much higher than that of the 3D NS-GSs derived from

295

poly(vinylpyrrolidone) and (NH4)2S2O8 (CE = 43.8%, 4.6 wt% N, and 1.5 wt% S) [50].

296

In the first cycle, the formation of the SEI layer and the irreversible storage of Li-ions

297

into special positions in the porous carbon material make the irreversible capacity

AC C

EP

TE D

M AN U

SC

280

ACCEPTED MANUSCRIPT inevitable[50]. Exactly, the voltage plateau between 0.6 and 0.8 V in the first

299

discharge curve can be attributed to electrolyte decomposition and SEI film formation

300

on the electrode surface, agreeing with the CV curve. The cycling performances of the

301

obtained carbon materials were evaluated at 0.1 A g−1 over a range of 0.01–3.0 V

302

versus Li/Li+. Stable reversible capacities were obtained after the first ten cycles. After

303

50 cycles, the NS1/3 electrode still maintained a discharge capacity of 864 mA h g−1

304

(Fig. 6c), indicating the excellent cycling stability of the material.

305

The cycling performances of the electrodes were also investigated at a larger current

306

density of 1A g−1 (Fig. 6d). For the NS1/3 electrode, the reversible capacity is stable at

307

653 mA h g−1 after 500 cycles, suggesting extraordinarily stable electrochemical

308

performance. The coulombic efficiencies of these anodes remained greater than 98%

309

after 10 initial cycles at 0.1 and 1 A g−1, indicating their highly reversible nature for

310

efficient Li-ion insertion/extraction. To further understand the electrochemical

311

performance of the as-prepared NSDPCs, we examined their rate performances. The

312

rate capabilities and cycle performances of the NSDPC electrodes at various current

313

densities (from 0.1 to 5 A g−1) are shown in Fig. 6e. At current densities of 0.1, 0.2,

314

0.5, 1, 2, and 5 A g−1, stable charge capacities of NS1/3 were 991, 864, 743, 656, 582,

315

and 463 mA h g−1, respectively. When the current density was restored to 0.2 A g−1

316

and 0.1 A g−1, the capacities rapidly recovered to 873 and 917 mA h g−1, respectively.

317

This result demonstrates that the as-prepared N/S co-doped carbon has excellent

318

potential as a high-rate anode material for LIBs. As shown in Table 1, N doping level

319

for these five samples were comparable, indicating that the enhanced capacity is

AC C

EP

TE D

M AN U

SC

RI PT

298

ACCEPTED MANUSCRIPT mainly contributed by S doping. Although many studies report that N doping is an

321

efficient way to improving the Li storage capability[51, 52], our results show that S

322

doping is more efficient for capacity enhancement. The electrochemical performance

323

of dual-doped carbon materials with varying sulfur content was further evaluated. As

324

shown in Fig. 6c and d, increasing sulfur content from 7.6 to 21.1 wt% result in the

325

continuous increase of reversible capacity. NS1/3 (with N and S content of 14.0% and

326

21.1 wt%, respectively) exhibited the best lithium storage performance. In order to

327

possess an overall understanding of the recent development of N/S co-doped carbon

328

anodes for LIBs, the cycling performance and rate capabilities of these anodes have

329

been summarized and listed in Tables S6.

M AN U

SC

RI PT

320

330

In addition, electrochemical impedance spectroscopy of LIBs was performed in the

332

frequency range from 100 kHz to 0.01 Hz (displayed as Nyquist plots). The superior

333

cycle performance of NS1/3 can be attributed to its excellent conductivity (confirmed

334

by the results shown in Fig. 6f). The Nyquist plots display a semicircular loop in the

335

high-frequency region and a sloped line in the low-frequency region. The diameter of

336

the semicircular loop corresponds to the resistance of charge transfer (Rct) at the

337

interface in the LIBs. The sloped line in the low-frequency region represents the

338

Warburg impedance, which is related to diffusion of Li-ion in the electrode materials.

339

The Nyquist plots were analyzed and fitted by an equivalent circuit model (Fig. 6f).

340

The electrode made with NS1/3 exhibited a much lower SEI film resistance (RSEI, 32.0

341

Ω) and charge transfer resistance (Rct, 7.9 Ω) in comparison with the other electrodes

AC C

EP

TE D

331

ACCEPTED MANUSCRIPT based on equivalent circuit simulation (Table S5). This result suggest that the NS1/3

343

electrode has a thinner SEI film, favoring rapid Li+ insertion/extraction and facile

344

charge transfer at the electrode/electrolyte interface. Moreover, the synergistic effect

345

of N and S co-doping favors NS1/3 with its enhanced electronic conductivity,

346

providing an alternative route for electron transfer and guaranteeing continuous and

347

rapid electron transport.

RI PT

342

349

Fig.

Electrochemical

performance

351

voltammograms of NS1/3 at a scan rate of 0.1 mV s−1; (b) charge-discharge curves of

352

NS1/3 at 0.1 A g−1; cycling performance at a current density of 0.1 A g−1 (c) and 1 A g−1

353

(d); (e) capacity over cycling at different current densities; (f) Nyquist plots of

354

NSDPC samples after ten initial cycles.

of

NSDPC

electrodes.

(a)

Cyclic

AC C

EP

350

355

6.

TE D

M AN U

SC

348

356

The Li storage mechanism of NSDPC materials is as follows. The relatively large

357

specific surface area and hierarchical porous structure, with coexisting micropores

358

and mesopores, providing a large quantity of sites for Li-ion storage. Moreover, with

ACCEPTED MANUSCRIPT ultrahigh levels of heteroatom doping, abundant defect and enlarged interlayer

360

spacing in the carbon matrix facilitate absorption and insertion/extraction of Li-ion. In

361

this case, Li-ion can diffuse through defects perpendicular to the interlayer plane, with

362

additional sites in the interlayer space for accommodation of Li-ion[50]. The 3D

363

porous networks serves as a reservoir for the storage of Li-ion and reduces the

364

diffusion distance for them, while the mechanical stability of the nanosheets ensures

365

superior

366

performance of the NSDPC materials is the result of their novel porous structure,

367

appropriate co-doping of heteroatoms, and the proper degree of graphitization. In

368

addition to increasing conductivity, the doped N (in the forms of pyridinic-N,

369

pyrrolic-N) act as electrophilic atoms due to their higher electronegativity than C

370

atoms, causing the nearby C atoms to be polarized and have more electrochemical

371

activity. Furthermore, in addition to increasing the interlayer spacing and defects in

372

the carbon matrix, sulfur atoms (in the forms of thiophenic-S species) covalently

373

bonded to the pyrolytic carbon can serve as accommodation sites for Li-ion, leading

374

to high capacity even at high rates. As a consequence, dual doping have a synergistic

375

effect for the Li-ion storage compare to the doping by single atom[34]. Therefore,

376

NSDPC materials are able to accept more charge than undoped carbon materials.

377

Recent theoretical studies also prove that doped carbon shows better storage of Li-ion

378

due to the increased number of defects in the graphene plane, consistent with our

379

results[26, 52].

380

Conclusions

Therefore,

the

outstanding

SC

performance.

electrochemical

AC C

EP

TE D

M AN U

cycling

RI PT

359

ACCEPTED MANUSCRIPT In summary, carbon-based materials are attracting extensive attention in the secondary

382

battery industry due to their low-cost and facile preparation. We employed a one-step

383

N/S co-doping synthesis process to convert cystine aggregates into heteroatom-doped

384

interconnected porous carbon nanosheets. The NS1/3 specimen exhibited a large

385

reversible lithium storage capacity of 864 mA h g−1 after 50 cycles at 0.1 A g−1,

386

excellent cycle performance (653 mA h g−1 for the 500th cycle at 1 A g−1), and a

387

superior rate capability (463 mA h g−1 at 5 A g−1). This obtained carbon materials

388

show great promise as an anode material for high-performance LIBs. The super

389

electrochemical performance of NSDPC materials originates from their unique

390

interconnected porous structure, appropriate pore distribution, and high inherent N

391

and S contents, which shorten the diffusion distance of lithium ions and provide a

392

large number of lithium storage sites.

393

Acknowledgements

TE D

M AN U

SC

RI PT

381

This work was supported by National Natural Science Foundation of China

395

(51673117), National High-tech R&D Program (863) (2015AA03A204), the Science

396

and Technology Innovation Commission of Shenzhen (JCYJ20140418091413553,

397

JCYJ20150625102750478,

398

JCYJ20160520163535684, JCYJ20160422144936457), Foundation for Distinguished

399

Young Talents in Higher Education of Guangdong, China(2013LYM_0080), Special

400

Program for Applied Research on Super Computation of the NSFC-Guangdong Joint

401

Fund (the second phase).

AC C

EP

394

JCYJ20150529164656097, JSGG20160226201833790,

ACCEPTED MANUSCRIPT References

403

[1] L. Lu, X. Han, J. Li, J. Hua, M. Ouyang, A review on the key issues for lithium-ion battery

404

management in electric vehicles, J. Power Sources 226 (2013) 272-288.

405

[2] A. Farmann, W. Waag, A. Marongiu, D.U. Sauer, Critical review of on-board capacity

406

estimation techniques for lithium-ion batteries in electric and hybrid electric vehicles, J. Power

407

Sources 281 (2015) 114-130.

408

[3] Z. Yang, J. Zhang, M.C. Kintner-Meyer, X. Lu, D. Choi, J.P. Lemmon, et al., Electrochemical

409

energy storage for green grid, Chem. Rev. 111(5) (2011) 3577-3613.

410

[4] K. Zhang, Z. Hu, Z. Tao, J. Chen, Inorganic & organic materials for rechargeable Li batteries

411

with multi-electron reaction, Sci. China Mater. 57(1) (2014) 42-58.

412

[5] H. Li, Z.X. Wang, L.Q. Chen, X.J. Huang, Research on Advanced Materials for Li-ion

413

Batteries, Adv. Mater. 21(45) (2009) 4593-4607.

414

[6] C.L. Wang, Y.G. Fang, Y. Xu, L.Y. Liang, M. Zhou, H.P. Zhao, et al., Manipulation of

415

Disodium Rhodizonate: Factors for Fast-Charge and Fast-Discharge Sodium-Ion Batteries with

416

Long-Term Cyclability, Adv. Funct. Mater. 26(11) (2016) 1777-1786.

417

[7] S. Licht, A. Douglas, J.W. Ren, R. Carter, M. Lefler, C.L. Pint, Carbon Nanotubes Produced

418

from Ambient Carbon Dioxide for Environmentally Sustainable Lithium-Ion and Sodium-Ion

419

Battery Anodes, Acs Central Sci. 2(3) (2016) 162-168.

420

[8] Y.H. Huang, Q. Bao, J.G. Duh, C.T. Chang, Top-down dispersion meets bottom-up synthesis:

421

merging ultranano silicon and graphene nanosheets for superior hybrid anodes for lithium-ion

422

batteries, J. Mater. Chem. A 4(25) (2016) 9986-9997.

423

[9] L. Chen, G. Zhou, Z. Liu, X. Ma, J. Chen, Z. Zhang, et al., Scalable Clean Exfoliation of

424

High-Quality Few-Layer Black Phosphorus for a Flexible Lithium Ion Battery, Adv. Mater. 28(3)

425

(2015) 510-517.

426

[10] J. Kong, Z. Liu, Z. Yang, H.R. Tan, S. Xiong, S.Y. Wong, et al., Carbon/SnO2/carbon

427

core/shell/shell hybrid nanofibers: tailored nanostructure for the anode of lithium ion batteries

428

with high reversibility and rate capacity, Nanoscale 4(2) (2012) 525-30.

429

[11] M.S.A.S. Shah, S. Muhammad, J.H. Park, W.S. Yoon, P.J. Yoo, Incorporation of PEDOT:PSS

430

into SnO2/reduced graphene oxide nanocomposite anodes for lithium-ion batteries to achieve

AC C

EP

TE D

M AN U

SC

RI PT

402

ACCEPTED MANUSCRIPT ultra-high capacity and cyclic stability, Rsc Adv. 5(18) (2015) 13964-13971.

432

[12] M. Wang, H. Yang, X. Zhou, W. Shi, Z. Zhou, P. Cheng, Rational design of SnO2@C

433

nanocomposites for lithium ion batteries by utilizing adsorption properties of MOFs, Chem.

434

Commun. 52(4) (2015) 717-20.

435

[13] C.M. Park, J.H. Kim, H. Kim, H.J. Sohn, Li-alloy based anode materials for Li secondary

436

batteries, Chem. Soc. Rev. 39(8) (2010) 3115-3141.

437

[14] D.D. Tao, Z.X. Fang, M. Qiu, Y. Li, X. Huang, K.N. Ding, et al., First-principles study of

438

Na2+xTi7O15 as anode materials for sodium-ion batteries, J. Alloy. Compd. 689 (2016) 805-811.

439

[15] P. Poizot, S. Laruelle, S. Grugeon, L. Dupont, J.M. Tarascon, Nano-sized transition-metal

440

oxides as negative-electrode materials for lithium-ion batteries, Nature 407(6803) (2000) 496-499.

441

[16] J. Chen, L.N. Xu, W.Y. Li, X. Gou, α-Fe2O3 Nanotubes in Gas Sensor and Lithium-Ion

442

Battery Applications, Adv. Mater. 17(5) (2005) 582-586.

443

[17] S.J. Yang, S. Nam, T. Kim, J.H. Im, H. Jung, J.H. Kang, et al., Preparation and exceptional

444

lithium anodic performance of porous carbon-coated ZnO quantum dots derived from a

445

metal-organic framework, J. Am. Chem. Soc. 135(20) (2013) 7394-7397.

446

[18] F.C. Zheng, D.Q. Zhu, X.H. Shi, Q. Chen, Metal–organic framework-derived porous

447

Mn1.8Fe1.2O4 nanocubes with an interconnected channel structure as high-performance anodes for

448

lithium ion batteries, J. Mater. Chem. A 3 (2015) 2815–2824.

449

[19] B. Liu, X.B. Zhang, H. Shioyama, T. Mukai, T. Sakai, Q. Xu, Converting cobalt oxide

450

subunits in cobalt metal-organic framework into agglomerated Co3O4 nanoparticles as an electrode

451

material for lithium ion battery, J. Power Sources 195 (2010) 857–861.

452

[20] P.L. Taberna, S. Mitra, P. Poizot, P. Simon, J.M. Tarascon, High rate capabilities Fe3O4-based

453

Cu nano-architectured electrodes for lithium-ion battery applications, Nat. Mater. 5(7) (2006)

454

567-573.

455

[21] M. Depardieu, R. Demir-Cakan, C. Sanchez, M. Birot, H. Deleuze, M. Morcrette, et al., On

456

the effect of gold nanoparticles loading within carbonaceous macro-mesocellular foams toward

457

lithium-sulfur battery performances, Solid State Sci. 55 (2016) 112-120.

458

[22] M. Marinescu, T. Zhang, G.J. Offer, A zero dimensional model of lithium-sulfur batteries

459

during charge and discharge, Phys. Chem. Chem. Phys. 18(1) (2016) 584-593.

460

[23] K. Zhang, X.P. Han, Z. Hu, X.L. Zhang, Z.L. Tao, J. Chen, Nanostructured Mn-based oxides

AC C

EP

TE D

M AN U

SC

RI PT

431

ACCEPTED MANUSCRIPT for electrochemical energy storage and conversion, Chem. Soc. Rev. 44 (2015) 699-728.

462

[24] D. McNulty, D.N. Buckley, C. O'Dwyer, Synthesis and electrochemical properties of

463

vanadium oxide materials and structures as Li-ion battery positive electrodes, J Power Sources

464

267 (2014) 831-873.

465

[25] Z. Li, Z. Xu, X. Tan, H. Wang, C.M.B. Holt, T. Stephenson, et al., Mesoporous nitrogen-rich

466

carbons derived from protein for ultra-high capacity battery anodes and supercapacitors, Energy

467

Environ. Sci. 6(3) (2013) 871-878.

468

[26] Y. Yang, F.C. Zheng, G.L. Xia, Z.Y. Lun, Q. Chen, Experimental and theoretical

469

investigations of nitro-group doped porous carbon as a high performance lithium-ion battery

470

anode, J. Mater. Chem. A 3(36) (2015) 18657-18666.

471

[27] J. Zhang, Z.X. Yang, J.Y.C. Qiu, H.W. Lee, Design and synthesis of nitrogen and sulfur

472

co-doped porous carbon via two-dimensional interlayer confinement for a high-performance

473

anode material for lithium-ion batteries, J. Mater. Chem. A 4(16) (2016) 5802-5809.

474

[28] J. Wang, S. Kaskel, KOH activation of carbon-based materials for energy storage, J.

475

Mater.Chem. 22(45) (2012) 23710-23725.

476

[29] C. Fang, Y.H. Huang, W.X. Zhang, J.T. Han, Z. Deng, Y.L. Cao, et al., Routes to High Energy

477

Cathodes of Sodium-Ion Batteries, Adv. Energy Mater. 6(5) (2016).

478

[30] Y.P. Wu, S.B. Fang, Y.Y. Jiang, R. Holze, Effects of doped sulfur on electrochemical

479

performance of carbon anode, J. Power Sources 108(1-2) (2002) 245-249.

480

[31] S.A. Wohlgemuth, F. Vilela, M.M. Titirici, M. Antonietti, A one-pot hydrothermal synthesis

481

of tunable dual heteroatom-doped carbon microspheres, Green Chem. 14(3) (2012) 741-749.

482

[32] R. Gokhale, V. Aravindan, P. Yadav, S. Jain, D. Phase, S. Madhavi, et al., Oligomer-salt

483

derived 3D, heavily nitrogen doped, porous carbon for Li-ion hybrid electrochemical capacitors

484

application, CARBON 80 (2014) 462-471.

485

[33] Y.S. Yun, V.D. Le, H. Kim, S.J. Chang, S.J. Baek, S. Park, et al., Effects of sulfur doping on

486

graphene-based nanosheets for use as anode materials in lithium-ion batteries, J. Power Sources

487

262 (2014) 79-85.

488

[34] F. Wu, J. Li, Y. Tian, Y. Su, J. Wang, W. Yang, et al., 3D coral-like nitrogen-sulfur co-doped

489

carbon-sulfur composite for high performance lithium-sulfur batteries, Sci. Rep. 5 (2015)

490

13340-13349.

AC C

EP

TE D

M AN U

SC

RI PT

461

ACCEPTED MANUSCRIPT [35] X.Q. Xie, T. Makaryan, M.Q. Zhao, K.L. Van Aken, Y. Gogotsi, G.X. Wang, MoS2

492

Nanosheets Vertically Aligned on Carbon Paper: A Freestanding Electrode for Highly Reversible

493

Sodium-Ion Batteries, Adv. Energy Mater. 6(5) (2016) 1502161.

494

[36] Z. Zhang, J. Feng, L. Ci, Y. Tian, S. Xiong, Metal-organic framework derived CuO hollow

495

spheres as high performance anodes for sodium ion battery, Mater. Technol. 31(12) (2016)

496

740-740.

497

[37] H. Han, C. Wang, Z. Ma, Z. Su, A facile method to produce highly monodispersed

498

nanospheres of cystine aggregates, Nanotechnology 17(20) (2006) 5163-5166.

499

[38] W. Li, M. Zhou, H.M. Li, K.L. Wang, S.J. Cheng, K. Jiang, A high performance sulfur-doped

500

disordered carbon anode for sodium ion batteries, Energy Environ. Sci. 8 (2015) 2916-2921.

501

[39] J. Ye, J. Zang, Z. Tian, M. Zheng, Q. Dong, Sulfur and nitrogen co-doped hollow carbon

502

spheres for sodium-ion batteries with superior cyclic and rate performance, J. Mater. Chem. A 4

503

(2016) 13223-13227.

504

[40] G.Q. Ning, X.L. Ma, X. Zhu, Y.M. Cao, Y.Z. Sun, C.L. Qi, et al., Enhancing the Li Storage

505

Capacity and Initial Coulombic Efficiency for Porous Carbons by Sulfur Doping, ACS Appl.

506

Mater. Interfaces 6(18) (2014) 15950-15958.

507

[41] C.K. Ho, C.Y.V. Li, K.Y. Chan, Scalable Template-Free Synthesis of Na2Ti3O7/Na2Ti6O13

508

Nanorods with Composition Tunable for Synergistic Performance in Sodium-Ion Batteries, Ind.

509

Eng. Chem. Res.55(38) (2016) 10065-10072.

510

[42] Y.M. Tao, K. Rui, Z.Y. Wen, Q.S. Wang, J. Jin, T. Zhang, et al., FeS2 microsphere as cathode

511

material for rechargeable lithium batteries, Solid State Ion. 290 (2016) 47-52.

512

[43] S.P. Wang, J.X. Yu, Electrochemical Mechanism for FeS2/C Composite in Lithium Ion

513

Batteries with Enhanced Reversible Capacity, Energies 9(4) (2016) 225.

514

[44] R.Y. Li, Y.Y. Jiang, X.Y. Zhou, Z.J. Li, Z.G. Gu, G.L. Wang, et al., Significantly enhanced

515

electrochemical performance of lithium titanate anode for lithium ion battery by the hybrid of

516

nitrogen and sulfur co-doped graphene quantum dots, Electrochim. Acta 178 (2015) 303-311.

517

[45] T.Q. Lin, I.W. Chen, F.X. Liu, C.Y. Yang, H. Bi, F.F. Xu, et al., Nitrogen-doped mesoporous

518

carbon of extraordinary capacitance for electrochemical energy storage, Science 350(6267) (2015)

519

1508-1513.

520

[46] F. Zheng, Y. Yang, Q. Chen, High lithium anodic performance of highly nitrogen-doped

AC C

EP

TE D

M AN U

SC

RI PT

491

ACCEPTED MANUSCRIPT porous carbon prepared from a metal-organic framework, Nat. Commun. 5 (2014) 5261-5270.

522

[47] Y. Meng, T.T. Yu, S. Zhang, C. Deng, Top-down synthesis of muscle-inspired alluaudite

523

Na2+2xFe2-x(SO4)3/SWNT spindle as a high-rate and high-potential cathode for sodium-ion

524

batteries, J. Mater. Chem. A 4(5) (2016) 1624-1631.

525

[48] T.S. Yoder, M. Tussing, J.E. Cloud, Y.A. Yang, Resilient carbon encapsulation of iron pyrite

526

(FeS2) cathodes in lithium ion batteries, J. Power Sources 274 (2015) 685-692.

527

[49] S.S. Zhang, D.T. Tran, Pyrite FeS2 as an efficient adsorbent of lithium polysulphide for

528

improved lithium-sulphur batteries, J. Mater. Chem. A 4(12) (2016) 4371-4374.

529

[50] D. Sun, J. Yang, X. Yan, Hierarchically porous and nitrogen, sulfur-codoped graphene-like

530

microspheres as a high capacity anode for lithium ion batteries, Chem. Commun. 51(11) (2015)

531

2134-2137.

532

[51] Z.L. Wang, D. Xu, H.G. Wang, Z. Wu, X.B. Zhang, In Situ Fabrication of Porous Graphene

533

Electrodes for High-Performance Energy Storage, Acs Nano 7(3) (2013) 2422-2430.

534

[52] Y. Mao, H. Duan, B. Xu, L. Zhang, Y. Hu, C. Zhao, et al., Lithium storage in nitrogen-rich

535

mesoporous carbon materials, Energy Environ. Sci. 5(7) (2012) 7950-7955.

M AN U

SC

RI PT

521

AC C

EP

TE D

536