Revealing the Dynamic Formation Process and Mechanism of Hollow

Dec 30, 2017 - Herein, we chose glucose as a carbon precursor and double surfactants poly(ethylene glycol)-block-poly(propylene glycol)-block-poly(eth...
0 downloads 5 Views 2MB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

Article

Revealing the Dynamic Formation Process and Mechanism of Hollow Carbon Spheres: from Bowl to Sphere’s Shape Xin Liu, Pingping Song, Jiahui Hou, Bo Wang, Feng Xu, and Xueming Zhang ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.7b04634 • Publication Date (Web): 30 Dec 2017 Downloaded from http://pubs.acs.org on January 3, 2018

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

ACS Sustainable Chemistry & Engineering is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

1

Revealing the Dynamic Formation Process and Mechanism of

2

Hollow Carbon Spheres: from Bowl to Sphere’s Shape

3

Xin Liu, Pingping Song, Jiahui Hou, Bo Wang, Feng Xu and Xueming Zhang*

4 5

Beijing Key Laboratory of Lignocellulosic Chemistry, Beijing Forestry University, 35 Qinghua

6

East Road, Haidian District, Beijing, P. R. China, 100083.

7 8

*Corresponding author

9

E-mail: [email protected] (Xueming Zhang). Tel. and Fax: +86-01062336189.

10 11 12 13 14 15 16 17 18 19 20 21 22 1

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

23

ABSTRACT

24

Hollow carbon spheres are attracting great attention due to their great potential utilizations in

25

drug delivery, energy storage and catalysis. However, the formation process and mechanism of

26

the hollow carbon spheres were still unclear. Herein, we chose glucose as carbon precursor,

27

double surfactants PEO–PPO–PEO triblock copolymers and sodium oleate as soft template, the

28

synthesis process of hollow carbon spheres was investigated in the coupling of soft templating

29

method and hydrothermal carbonization system by regulating the reaction time. A dynamic

30

formation process of the hollow carbon spheres was identified based on the results from SEM and

31

TEM images, in which it experienced three evolution stages including hollow carbon bowls,

32

capsules and spheres. In addition, the formation mechanism was also presumed: During the

33

synthesis process, the double surfactant interacted with each other to act as the soft template, and

34

the glucose underwent hydration, polymerization and aromatization stages. When the

35

concentration of aromatic compounds reached the critical supersaturation, the nucleation took

36

place from a point and extended outward gradually along the interface to widen and thicken the

37

carbon shell, resulting in different morphologies’ hollow structured carbon particles were formed

38

successively by controlling the reaction time. Furthermore, the resultant hollow structured carbon

39

particles were stable and uniform, and we made preliminary explorations on their biochemical

40

and electrochemical performance.

41

KEYWORDS: hollow carbon spheres, glucose, soft templating method, hydrothermal

42

carbonization

43 44

INTRODUCTION 2

ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

45

Hollow carbon spheres (HCSs) have attracted increasing attention in recent years. Owing to

46

their unique structure and properties such as encapsulation ability, high surface-to-volume ratio,

47

surface functionality, excellent thermal and chemical stabilities1-3, they have been widely studied

48

for adsorption/separation, catalysis, drug delivery, synthesis template, micro-reactor, energy

49

storage and conversion applications4-10.

50

Much effort has been devoted to the synthesis of HCSs by various approaches. The most

51

commonly and widely used method is the template-assisted methods11-13. Generally, the

52

template-assisted methods can be divided into hard templating method and soft templating

53

method according to the types of core templates used. In the hard templating method, some

54

specially prepared rigid particles are employed as core templates. When the carbon shells around

55

the cores are generated, the templates are generally sacrificed by high temperature, dissolution, or

56

acid/alkaline etching, resulting in the formation of the hollow carbon structure14. This method has

57

the advantages of better control of the morphology and the size of products, while it generally

58

includes the drawbacks of multi-steps, time-consuming and non-environmentally friendly 15-18. In

59

comparison, the soft templating method, based on the self-assemble of the precursor molecules

60

and the colloidal systems as the core-template, is more convenient. Organic additives, block

61

copolymers, surfactants, nanoemulsion droplets or gas bubbles always act as the core-templates,

62

which can be easily consumed or removed in the subsequent process, such as pyrolysis or

63

extraction19,20. However, the precursors in the soft-template method usually tend to polymerize at

64

low temperatures. Additionally, many carbon precursors are highly dependent on fossil-based

65

compounds or carbonaceous polymers (ie. benene, phenol, ethylene, resorcinol, formaldehyde,

66

dihydroxybenzoic acid, cokes, polystyrene, polyacrylonitrile)13,21-23, which are generally 3

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

67

corrosive, toxic or expensive.

68

Considering the environmental concern and economic value, it is preferable to prepare HCSs

69

from renewable resources in a highly efficient and environmental friendly way24,25. Hydrothermal

70

carbonization (HTC), as a facile, low-cost and environmentally friendly technique, is used for

71

preparation of functional carbon materials, and the obtained carbon materials are considered as

72

sustainable alternatives to traditional carbons6,24,25. The hydrothermal carbonization process has a

73

couple of merits, including the use of renewable resources (either from isolated carbohydrates or

74

crude plants), facile instrumentation and techniques, and a high energy and atom economy7,26.

75

However, the main problems of HTC carbons are their intrinsic low porosity and conductivity6,27,

76

and it has been reported that the template-assisted methods is a powerful tool to generate the

77

porous materials28,29. Notably, coupling the template-assisted method with the hydrothermal

78

carbonization process has shown powerful capability in controlling the synthesis of various

79

carbonaceous nanostructures with special morphology and property7,29,30.

80

In a previous study by Sun and Li4, monodisperse colloidal carbon microspheres were prepared

81

from aqueous glucose solutions by hydrothermal synthesis, and encapsulated noble-metal

82

nanoparticles were also obtained by this procedure. In addition, in Wang et al. work17, the

83

formation of carbon structure was investigated based on the hydrothermal treatment of glucose on

84

amino-functionalized surfaces of silica-based templates, in which the templates were inversely

85

replicated. After heat treatment and aqueous hydrofluoric acid treatment, the well-developed

86

porous shell hollow carbon spheres were obtained and they were applicable to catalyst supports.

87

Recently, Wang et al.23 used PEO-PPO-PEO triblock copolymers (P123) and sodium oleate (SO)

88

as double surfactants and 2,4-dihydroxybenzoic acid and hexamethylenetetramine as the polymer 4

ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

89

precursors to synthesize uniform hollow polymer spheres (HPSs) based on hydrothermal

90

treatment. Moreover, PtCo bimetallic nanoparticles within the hollow carbon spheres

91

((PtCo@HCS-500) were prepared, which showed outstanding catalytic performance in the

92

hydrogenolysis of 5-hydroxymethylfurfural (HMF) to 2,5-dimethylfuran (DMF).

93

Although some researches focused on the synthesis of hollow carbon spheres and their

94

potential utilizations, few studies aimed at investigating their formation process and mechanism,

95

especially under the coupling of soft templating method and hydrothermal carbonization system.

96

Therefore, further investigation into the evolution states and forming mechanism of hollow

97

carbon spheres is required. In this paper, Glucose, an inexpensive, nontoxic and sustainable

98

carbon-rich biomass material, was chosen as the carbon precursor. The biocompatible and

99

commercially available double surfactants PEO-PPO-PEO triblock copolymers (P123) and

100

sodium oleate (SO) were used as the soft template. Since no hazardous chemicals involved in the

101

synthesis process, and the coupling of soft templating method and hydrothermal carbonization

102

approach was simple and facile, it was believed that our study met the concept of sustainable and

103

green chemistry. A dynamic formation process of the hollow carbon spheres was studied by

104

regulating the reaction time. Depending on the reaction time, this process experienced the

105

dynamic formation of hollow carbon bowls (HCBs), capsules (HCCs), and spheres (HCSs). The

106

formation mechanism of hollow carbon structures has also been assumed, which would be

107

valuable for illustrating the hollow carbon spheres growth in the hydrothermal carbonization and

108

soft templating method synergistic system. Additionally, the resultant hollow structured carbon

109

particles (HCPs) were stable and uniform, and their performances in potential applications for

110

biomedical materials and supercapacitors were simultaneously evaluated. 5

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 31

111 112

EXPERIMENTAL SECTION

113

Chemicals

114

The PEO–PPO–PEO triblock copolymers (P123) was purchased from Aldrich. Sodium oleate

115

(SO) and glucose were purchased from Beijing Ouhe Technology Co., Ltd and Sinopharm

116

Chemical Reagent

117

polytetrafluoroethylene (PTFE) were supplied from Beijing Lanyi Chemical Products Co., Ltd.

118

All the chemicals above were analytical grade and used without further purification.

119

Synthesis of hollow carbon spheres (HCSs)

Co.,

Ltd,

respectively.

Other

chemicals including graphite

and

120

Typically, 0.016 mol glucose was dissolved in 40 mL deionized water. Then, 20 mL of an

121

aqueous solution containing 0.12 mmol SO and 0.0075 mmol P123 was added. After stirring for

122

30 min at room temperature, the solution was transferred into a 75 mL autoclave and

123

hydrothermal treated at 160 °C for different reaction times (3, 6, 12, 18 and 24 h). Then the dark

124

brown products were collected after centrifugation and washed several times with deionized

125

water, and dried at 60 °C overnight. Based on the different morphologies of the obtained particles,

126

we named them as hollow carbon bowls (HCBs), capsules (HCCs), and spheres (HCSs),

127

separately. And all of these were collectively called the hollow structured carbon particles

128

(HCPs).

129

Synthesis of carbonized hollow carbon spheres (CHCSs)

130

The HCSs sample was undergone a temperature-programmed calcination, in which the specific 6

ACS Paragon Plus Environment

Page 7 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

131

program was set as follows. The HCSs were heated to 200 °C and maintained for 30 min, then

132

calcined to 600 °C and held for 1 h, subsequently the HCSs were further heated to 900 °C and

133

kept for 1 h. Whole process was performed under N2 atmosphere with a heating rate of 5 °C/min.

134

The obtained black powder was named as carbonized hollow carbon spheres (CHCSs).

135

Characterization

136

Scanning electron microscopy (SEM) micrographs were taken with a MERLIN VP compact field

137

emission scanning electron microscope (ZEISS, Germany) using an accelerating voltage of 15 kV.

138

Transmission electron microscopy (TEM) and high resolution transmission electron microscopy

139

(HRTEM) micrographs were recorded on JEM-2100F (JEOL, Japan) operating at 100 kV with an

140

energy dispersive X-ray spectrometer (EDX). The dynamic light scattering (DLS) and zeta

141

potential measurements were measured with a zeta meter (Zeta Master, Malvern Instruments,

142

UK). The crystal structure of samples was recorded on an X-ray diffractometer (Smartlab, Rigaku,

143

Japan) with Cu Kα radiation (λ=0.154 nm). The Raman spectra were performed on a Raman

144

Spectrometer (JY HR-800, HORIBA Ltd., France) at room temperature. Fourier-transform

145

infrared (FT-IR) spectra were measured by a Perkin Elmer Spotlight 400 imaging system (Perkin

146

Elmer, UK). Surface tension measurements were obtained on a contact angle meter SL200KS

147

(Kino, USA). The thermogravimetric analysis (TGA) was carried out using a TG-DTA 7300

148

analyzer (SⅡ, Japan) from 20 to 900 °C in N2 flow of 50 mL/min at a heating rate of 10 °C/min.

149

The N2 adsorption−desorption isotherms and pore size distribution analysis were performed by

150

using a physisorption analyser (SSA-7000, Builder, China).

151

Cytotoxicity evaluation

7

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

152

The cytotoxicity of HCBs was evaluated by a cell counting kit-8 (CCK-8) assay. Mouse

153

fibroblast-cells (L929) were seeded at a density of 4 × 103 cells per well in 180 mL culture

154

medium and incubated for 24 h. Then, the cells were treated with different concentrations of

155

HCBs solution at 37 °C in a humidified incubator with 5% CO2 for 24 h. Then, 20 µL of CCK-8

156

solution was added to each well and incubated for 1 h at 37 °C. The absorbance value was

157

measured with an infinite M200 microplate spectrophotometer (Tecan, Switzerland) at 450 nm

158

wavelength. The cell viability was expressed as percentage of absorbance relative to control, and

159

the control was obtained in the absence of HCBs.

160

Electrochemical characterization

161

The electrode material was a mixture of the active material, graphite and PTFE with the mass

162

ratio of 85:10:5. At room temperature, alcohol was added to the electrode material and stirred

163

continuously to form homogeneous slurry8. Then the slurry was coated on the nickel foam and

164

pressed under 10 MPa for 5min. After being dried to evaporate the solvent, the working electrode

165

was obtained.

166

The electrochemical measurements were performed by using two-electrode test system

167

consisting of two symmetric CHCSs-based electrodes in 6 M KOH electrolyte. All the

168

electrochemical measurements were carried on Autolab 302N electrochemical workstation

169

(Metrohm Ltd, Switzerland) at room temperature. The cyclic voltammetry (CV) was obtained in

170

the voltage range between 0 and 1 V at scan rates from 10 to 200 mV/s. Galvanostatic

171

charge−discharge (GCD) measurements were carried out at current density from 0.1 to 5 A/g in

172

the voltage range between 0 and 1 V. The Nyquist plot was performed at open circuit potential

8

ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

173

over the frequency range from 0.01 to 100 KHz with an amplitude of 5 mV, and it could be fitted

174

to an equivalent circuit including an electrolyte resistance (Rs), charge transfer resistance (Rct),

175

Warburg element (Zw) and pseudocapacitance (Cp)31.

176 177

The specific capacitance of the symmetric supercapacitor was determined from the GCD measurements using the equation: ଶ୍×∆୲

178

C୫ = ∆୚×୫

(1)

179

where I (A) is the discharge current, ∆t (s) is the discharge time, ∆V (V) is the voltage change

180

(excluding the iR drop) within the discharge time, and m (g) is the mass of the active materials on

181

the two electrodes.

182 183

The energy density (E, Wh/kg) and power density (P, W/kg) derived from the GCD curves were calculated by the following equations: (∆୚)మ

184

E = C୫ × ଼×ଷ.଺ ୉

185 186

P = ∆୲

(2) (3)

The controlled experiment with the pure graphite as the active material was tested according to

187

the above method.

188

RESULTS AND DISCUSSION

189

Our research aimed at revealing the formation process and synthesis mechanism of the hollow

190

carbon spheres (HCSs) using glucose as the carbon precursor and coupling both the soft

191

templating method and hydrothermal carbonization process. The double surfactants

192

PEO-PPO-PEO triblock copolymers (P123) and sodium oleate (SO) were chosen as the soft

193

template. The P123 is known to have strong interaction with ionic surfactants, leading to the 9

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

194

formation of mixed micelles in aqueous media32. As shown in the Figure S1, the dynamic light

195

scattering (DLS) results showed the sizes of SO and P123 micelles in H2O were around 150 nm

196

and 25 nm, respectively. When mixing the P123 with SO in the solution, the micelle size

197

dramatically decreased to 3.5 nm. It was also noted that the critical micelle concentration of

198

mixed micelles was less than that from pure SO (Figure S2), which confirmed the formation of

199

P123/SO micelles due to the strong interaction between P123 and SO.

200 201

Figure 1. (a−e) TEM and (f−j) SEM images of the HCPs prepared with different reaction times.

202

To understand the formation and evolution process of the HCSs, we characterized the physical

203

morphology and structure of the HCPs obtained at different reaction times from 3 to 24 h by

204

TEM and SEM images (Figure 1). The images in Figures 1a and 1f revealed that the bowl-like

205

HCPs began to take shape even though the thin shell did not have enough mechanical strength to

206

retain the plump structure (3 h). For the samples synthesized for 6h (Figures 1b and 1g), the

207

morphology of bowl-like HCPs were more regular, but deflated structure and fragments still

208

existed. When the reaction time prolonged to 12 h (Figures 1c and 1h), HCBs with diameter

209

around at 400 nm and shell thickness of about 60 nm formed eventually, which became more

210

uniform and monodisperse. After 18 h reaction, the HCCs with an opening formed as shown in

10

ACS Paragon Plus Environment

Page 10 of 31

Page 11 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

211

Figures 1d and 1i, the average diameter was between 500 and 550 nm and shell thickness was in

212

the range of 80-100 nm. Finally, the uniform HCSs with diameter of 550-600 nm and shell

213

thickness of about 180 nm were formed after hydrothermal treatment for 24 h as shown in

214

Figures 1e and 1j. Obviously, we could conclude that the diameter of hollow void first increased

215

and then decreased and shell thickness increased with the increase of the reaction time, which

216

indicated that the swelling of the nanoemulsion template and the polymerization in the core kept

217

happening during the hydrothermal carbonization process.

218 219

Figure 2. TEM and SEM images of the corresponding morphologies synthesized with different molar ratios of P123

220

and SO: (a) 0:16 (without P123), (b) 0.5:16, (c) 2:16, (d) 4:16, (e) 1:0 (without SO).

221 222

Figure 3. TEM and SEM images of the corresponding morphologies synthesized with different amounts of glucose (a)

223

and (c): 0.008 mol, (b) and (d): 0.032 mol.

224

In order to investigate the effect of different molar ratios of P123 and SO and amounts of

11

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

225

glucose on the morphologies of HCSs, some contrast experiments were also carried out as shown

226

in Figures 2 and 3. It was clearly noted that many small spherical carbon particles with around

227

180 nm diameter and some large HCCs were formed with the absence of P123 in the reaction

228

system (Figures 2a and 2f). However, when only the P123 was added, the carbon particles with a

229

wide size distribution from 80 nm to 2 µm could be obtained (Figures 2e and 2j). As the ratios of

230

P123 to SO were 0.5:16 (Figures 2b and 2g) and 1:16 (Figures 1c and 1h), the relatively regular

231

HCBs formed, which indicated the P123 interacted with SO and served as morphology controller.

232

The successive increasing of P123 (2:16) resulted in the formation of larger HCBs with uneven

233

shell thickness as shown in Figures 2c and 2h. However, too much P123 (Figures 2d and 2i) led

234

to the formation of much bigger bowl-like HCPs and the size distribution was very broad, which

235

might be attributed to the instability of the emulsions in hydrothermal process23.Therefore, it

236

could be concluded that the optimal molar ratio between P123 and SO was 1:16 in our study.

237

What’s more, we also adjusted the amount of carbon precursor (glucose) as shown in Figure 3.

238

When the amount of glucose was reduced to 0.008 mol, the deflated HCBs were formed with 300

239

nm in average diameter and 80 nm in shell thickness (Figures 3a and 3c). With increasing of

240

glucose to 0.032 mol, the HCSs formed with uneven size and shell thickness as depicted in

241

Figures 3b and 3d.

12

ACS Paragon Plus Environment

Page 12 of 31

Page 13 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

242 243

Figure 4. FT-IR spectra of the glucose, HCBs and HCSs.

244

The FT-IR spectra corresponding to HCBs, HCSs and glucose raw material are shown in

245

Figure 4. It was clear that nearly no differences between HCBs and HCSs samples were observed.

246

The bands at approximately 2900 cm-1 corresponded to stretching vibrations of aliphatic C-H33.

247

And the band at 1000–1460 cm-1 were assigned to C-O stretching vibrations in hydroxyl, ester or

248

ether and O-H bending vibrations34. Notably, the intensities of the bands corresponding to the

249

hydroxyl groups (3000–3700 and 1000–1450 cm-1)35 in the samples became weaker compared

250

with the raw glucose, and new vibration band of C=O groups appeared at 1707 cm-1, which was

251

attributed to the dehydration reactions. Meanwhile, the new vibration peaks at 1615, 1510, and

252

1450 cm-1 unambiguously indicated the existence of benzene ring structure in the samples36,37,

253

and the bands at 875–750 cm-1 were assigned to aromatic C-H out-of-plane bending vibrations38.

254

Based on the above analysis, it was demonstrated that the aromatization of glucose occurred

255

during hydrothermal treatment. Additionally, the zeta-potential of HCSs was negative value

256

(-18.4 mV), which also confirmed the hydrophilic groups attached to the carbon framework39.

257

The presence of hydrophilic groups would greatly improve the stability of HCPs for enhancing 13

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

258

their dispersions in aqueous media, and be beneficial for the introduction of additional functional

259

groups or materials during hydrothermal reaction or post functionalization, resulting in further

260

improvement in the physical and chemical properties of HCPs7, 40,41.

261 262

Scheme 1. Schematic illustration of the formation process of HCSs.

263

A possible mechanism and formation process of HCSs were proposed as illustrated in Scheme

264

1. Firstly, the SO and P123 as the double surfactants formed mixed micelles due to the

265

hydrophobic interaction between the alkyl chains of SO and PO blocks in P123 in aqueous

266

media23, 42. As the temperature increased and time prolonged, the oleate would convert to oleic

267

acid and form oleic acid emulsion core43. Meanwhile, the glucose interacted with

268

PEO–PPO–PEO triblock copolymer template through the hydrogen bond between hydroxyl

269

groups and EO repeating units20, 44, and the glucose underwent dehydration reactions and led to

270

the formation of soluble products, such as furfural-like compounds (i.e. 5-hydroxymethyl furfural,

271

furfural) and some organic acids (i.e. formic, acetic and lactic acids)4, 38, 45. These acids lowered

14

ACS Paragon Plus Environment

Page 14 of 31

Page 15 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

272

the pH of the system and acted as the catalysts for the subsequent further dehydration and

273

degradation reaction26,29,36. In addition, the PPO segments became more hydrophobic led to the

274

swelling of the nanoemulsion and the PEO segments contributed to their stability during this

275

stage18, 44. The subsequent reaction stage has been denoted as the “polymerization” stage, which

276

may be induced by the intermolecular dehydration, fragmentation, aldol condensation and

277

resulted in the formation of soluble polymers. Subsequently, the aromatization of polymers

278

occurred. When the concentration of aromatic compounds reached its critical supersaturation, a

279

burst nucleation took place and grew towards the interface of the emulsion droplets and water,

280

leading to the formation of carbon shell. Based on the SEM and TEM images from different

281

reaction time, the nucleation started from a point and extended outward gradually along the

282

interface to widen and thicken the carbon shell until the final size and morphology were achieved.

283

And a dynamic formation process of hollow carbon spheres was identified, which experienced

284

three evolution stages including hollow carbon bowls, capsules and spheres.

285 286

Figure 5. Cell viability of L929 cells in the presence of different concentration of HCBs.

287

To explore the potential application of the HCPs in biochemical application, the cell counting

288

kit-8 (CCK-8) assay was used to evaluate the cytotoxicity of the HCPs to living cells. As shown 15

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

289

in Figure 5, take the HCBs as an example, the L929 cell viabilities against samples were

290

estimated to be greater than 80% after 24 h incubation with the concentration ranged from 10 to

291

1000 µg/ mL, which indicated the excellent biocompatibility, low toxicity and safety in vitro and

292

in vivo applications of the as-prepared products46-47. Additionally, the HCPs had quite uniform

293

shell thickness and easily controlled void size, also possessed appropriate surface functional

294

groups, which may be helpful in realizing identical behavior in biochemistry and clinical

295

diagnostics prospects36, 48.

296 297

Figure 6. (a) TEM images of CHCSs and (b-d) Element mapping overlap, C and O elements in CHCSs.

298

The carbonized hollow carbon spheres (CHCSs) were obtained by a temperature-programmed

299

calcination of HCSs as depicted in the experimental. As shown in Figure 6a, the original

300

morphology of HCSs were still remained after high-temperature treatment. The elemental

301

mapping images (Figures 6b-6d) showed that all the expected elements, including carbon and

302

oxygen elements, could be detected and matched well with CHCSs’ structure. In order to

303

investigate whether the soft template core in the HCSs was removed after hydrothermal process

304

or not, the pure oleic acid, P123 and HCSs were analysed by thermogravimetric (TG) analysis

305

(Figure S3). It was shown that the onset of thermal degradation of pure oleic acid and P123 was

306

around 200 °C, which was higher than the hydrothermal process (160 °C). Moreover, the

307

maximum rate of weight loss was observed at the second pyrolysis stage, in which over 96%

16

ACS Paragon Plus Environment

Page 16 of 31

Page 17 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

308

weight was pyrolyzed between 200 and 400 oC. Therefore, it was tentatively concluded that the

309

soft template core was still retained after hydrothermal process, while it was completely removed

310

in

311

adsorption–desorption isotherms and pore size distribution of CHCSs are shown in Figure S4.

312

According to the International Union of Pure and Applied Chemistry (IUPAC) classification, the

313

N2 adsorption–desorption isotherms of the CHCSs (Figure. S4a) were identified as type I,

314

supporting the microporous nature of the interconnected microporous carbon framework49. The

315

CHCSs displayed a specific BET surface area of 871.02 m2/g and a pore volume of 0.59 cm3/g,

316

respectively. As shown in Figure S4b, the pore size distribution concentrated in micropores region,

317

the existence of micropores was beneficial for increasing the effective contact area between

318

electrolyte ion and electrode49,50. On the contrary, if the pores are too small to allow easy access

319

of electrolyte ions, they will not contribute to double-layer capacitance and will weaken the

320

ability to store charge51. Additionally, the hollow core in CHCSs can further shorten the ion

321

diffusion distance and accelerate the ion response rate18.

the

temperature-programmed

carbonization

process23.

322 323

Figure 7. (a) Raman spectra and (b) XRD pattern of CHCSs and HCSs.

17

ACS Paragon Plus Environment

What’s

more,

the

N2

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

324

The Raman spectra of CHCSs and HCSs (Figure 7a) showed two distinct peaks at 1587 cm-1

325

(G-band) and 1341 cm-1 (D-band), which were attributed to in-plane vibrations of crystalline

326

graphite and disordered amorphous carbon, respectively52. Both the D band and G band of HCSs

327

displayed a broad shape and relative low intensity, indicating the high content of amorphous

328

carbon. While the D band and G band of CHCSs showed sharper and high intensity peaks, which

329

indicated the improved graphitization degree after calcination53. The IG/ID intensity ratio of

330

CHCSs was calculated to be 1.04, indicating that the CHCSs was partially graphitized54. In

331

addition, the weak broad 2D peak observed at ~2800 cm-1 was scribed to the second-order zone

332

boundary phonons55. This kind of carbons with partial graphitization is highly desirable for the

333

application as electrode materials56. Figure 7b showed the XRD patterns of CHCSs and HCSs.

334

Before the carbonization, the XRD pattern of HCSs showed a broad peak at 2θ = 21.30°, which

335

was attributed to highly disordered and low crystallined carbon atoms. After carbonization, two

336

obvious broad diffraction peaks (around at 24° and 44°) were observed in the XRD pattern of

337

CHCSs, which were assigned to the stacking carbon layer structure (002) and ordered graphitic

338

carbon structure (100), suggesting the carbonization degree increased after carbonization process

339

and coincided with the Raman result57,58. Additionally, an obvious increase in the intensity at the

340

low-angle region of the XRD pattern also demonstrates a high porosity in the sample as

341

mentioned above49,59.

18

ACS Paragon Plus Environment

Page 18 of 31

Page 19 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

342 343

Figure 8. (a) CV curves at various sweep rates and (b) GCD curves at different current densities for CHCSs.

344

The electrochemical performance of the CHCSs was measured with symmetric two-electrode

345

system in 6 M KOH solution. Obviously, comparing with the pure graphite electrode (Figure S5),

346

the performance of CHCSs electrode has been improved significantly. The cyclic voltammetry

347

(CV) curves in Figure 8a showed nearly symmetrical rectangular shapes in the range from 10 to

348

200 mV/s and displayed a more rectangular shape at a low scan rate. The existence of Equivalent

349

Series Resistance (Figure S6) and pore structure may cause the slight deviation of the CV

350

curves53. The galvanostatic charge/discharge (GCD) profiles (Figure 8b) exhibited nearly linearly

351

symmetric triangular shape and the columbic efficiency can still reach up to 94% when the

352

current density was 1 A/g, indicating electrode’s charge–discharge process was highly reversible.

353

The slight curvature shown on the lines might be ascribed to a certain degree of pseudocapacitive

354

behavior of the electrode due to the existence of the oxygen-containing groups as revealed in the

355

EDX and FT-IR results60. The specific capacitance of CHCSs was calculated as 116 F/g at 0.1 A/g,

356

and the CHCSs displayed an energy density of 4.03 Wh/kg and a power density of 25.1 W/kg

357

under the same condition. These results indicated an efficient double layer and fast ion transport

358

within the working electrodes was formed and the CHCSs possessed the potential in 19

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

359

supercapacitor application.

360 361

CONCLUSION

362

In summary, the formation process and mechanism of hollow carbon spheres were investigated

363

based on the coupling the soft templating method with hydrothermal carbonization process.

364

During this process, the double surfactant interacted with each other, resulting in forming mixed

365

micelles and acting as the soft template. Glucose underwent dehydration reactions and led to the

366

formation of soluble products, such as furfural-like compounds and some organic acids. These

367

acids lowered the pH of the system and acted as the catalysts for the subsequent further

368

dehydration and degradation reaction. Then the soluble derivatives went through polymerization

369

and aromatization stages. When the concentration of aromatic compounds reached its critical

370

supersaturation, a burst nucleation took place from a point and extended outward gradually along

371

the interface to widen and thicken the carbon shell until the final size and morphology were

372

achieved. By regulating the reaction time, a dynamic formation process of hollow carbon spheres

373

was identified, in which it experienced three evolution stages including hollow carbon bowls,

374

capsules and spheres. Furthermore, the resultant hollow structured carbon particles were stable

375

and uniform, and showed good performance in potential applications for biochemistry materials

376

and supercapacitors.

377 378

379

ACKNOWLEDGMENT

We are grateful for the financial support of this research from the Fundamental Research Funds 20

ACS Paragon Plus Environment

Page 20 of 31

Page 21 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

380

for the Central Universities (No. 2016JX03), National Key Research and Development Program

381

of China (2017YFD0601004) and Natural Science Foundation of China (No. 31470606).

382 383

384 385

Note

The authors declare no conflict of interest including any financial, personal or other relationships with other people or organizations.

386 387

ASSOCIATED CONTENT

388

Supporting Information

389

The size distribution and surface tension measurements of the P123, SO and P123/SO

390

micelle/emulsion in H2O. The TG and DTG curves of P123, oleic acid and HCSs. N2

391

adsorption-desorption isotherms and pore size distribution of CHCSs. Some supplementary

392

electrochemical results for pure graphite and CHCSs electrodes. The Supporting Information is

393

available free of charge on the ACS Publications website.

394 395

AUTHOR INFORMATION

396

Corresponding Author

397

*Tel. /Fax: +86-01062336189. E-mail: [email protected].

398

399 21

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 31

400

REFERENCES

401

(1) Zhou, W.; Xiao, X.; Cai, M.; Yang, L. Polydopamine-coated, nitrogen-doped, hollow

402

carbon-sulfur double-layered core-shell structure for improving lithium-sulfur batteries. Nano

403

Lett. 2014, 14 (9), 5250-5256, DOI: 10.1021/nl502238b.

404

(2) Liang, J.; Yu, X. Y.; Zhou, H.; Wu, H. B.; Ding, S.; Lou, X. W. Bowl-like SnO2@carbon

405

hollow particles as an advanced anode material for lithium-ion batteries. Angew. Chem., Int. Ed.

406

2014, 53 (47), 12803-12807, DOI: 10.1002/anie.201407917.

407

(3) Liang, J.; Hu, H.; Park, H.; Xiao, C.; Ding, S.; Paik, U.; Lou, X. W. Construction of hybrid

408

bowl-like structures by anchoring NiO nanosheets on flat carbon hollow particles with enhanced

409

lithium

410

10.1039/c5ee01125f.

411

(4) Sun, X.; Li, Y. Colloidal carbon spheres and their core/shell structures with noble-metal

412

nanoparticles. Angew. Chem., Int. Ed. 2004, 43 (5), 597-601, DOI: 10.1002/anie.200352386.

413

(5) Gröger, H.; Kind, C.; Leidinger, P.; Roming, M.; Feldmann, C. Nanoscale hollow spheres:

414

microemulsion-based synthesis, structural characterization and container-type functionality.

415

Materials 2010, 3 (8), 4355-4386, DOI: 10.3390/ma3084355.

416

(6) Wang, J.; Nie, P.; Ding, B.; Dong, S.; Hao, X.; Dou, H.; Zhang, X. Biomass derived carbon

417

for energy storage devices. J. Mater. Chem. A 2017, 5 (6), 2411-2428, DOI: 10.1039/c6ta08742f.

418

(7) Hu, B.; Wang, K.; Wu, L.; Yu, S. H.; Antonietti, M.; Titirici, M. M. Engineering carbon

419

materials from the hydrothermal carbonization process of biomass. Adv. Mater. 2010, 22 (7),

420

813-828, DOI: 10.1002/adma.200902812.

storage

properties.

Energy

Environ.

Sci.

2015,

22

ACS Paragon Plus Environment

8

(6),

1707-1711,

DOI:

Page 23 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

421

(8) Fang, B.; Kim, J. H.; Kim, M.; Kim, M.; Yu, J. S. Hierarchical nanostructured hollow

422

spherical carbon with mesoporous shell as a unique cathode catalyst support in proton exchange

423

membrane fuel cell. Phys. Chem. Chem. Phys. 2009, 11 (9), 1380-1387, DOI: 10.1039/b816629c.

424

(9) Arif, A. F.; Kobayashi, Y.; Balgis, R.; Ogi, T.; Iwasaki, H.; Okuyama, K. Rapid

425

microwave-assisted synthesis of nitrogen-functionalized hollow carbon spheres with high

426

monodispersity. Carbon 2016, 107, 11-19, DOI: 10.1016/j.carbon.2016.05.048.

427

(10) Arif, A. F.; Kobayashi, Y.; Schneider, E. M.; Hess, S. C.; Balgis, R.; Izawa, T.; Iwasaki, H.;

428

Taniguchi, S.; Ogi, T.; Okuyama, K.; Stark, W. J. Selective Low-Energy Carbon Dioxide

429

Adsorption Using Monodisperse Nitrogen-Rich Hollow Carbon Submicron Spheres. Langmuir

430

2017, DOI: 10.1021/acs.langmuir.7b01353.

431

(11) Wang, D.; Yu, Y.; He, H.; Wang, J.; Zhou, W.; Abruna, H. D. Template-free synthesis of

432

hollow-structured Co3O4 nanoparticles as high-performance anodes for lithium-ion batteries. ACS

433

nano 2015, 9 (2), 1775-1781, DOI: 10.1021/nn400731g.

434

(12) Liu, J.; Wickramaratne, N. P.; Qiao, S. Z.; Jaroniec, M. Molecular-based design and

435

emerging applications of nanoporous carbon spheres. Nat. Mater. 2015, 14 (8), 763-774, DOI:

436

10.1038/nmat4317.

437

(13) Li, S.; Pasc, A.; Fierro, V.; Celzard, A. Hollow carbon spheres, synthesis and applications–a

438

review. J. Mater. Chem. A 2016, 4 (33), 12686-12713, DOI: 10.1039/c6ta03802f.

439

(14) Valle-Vigón, P.; Sevilla, M.; Fuertes, A. B. Synthesis of uniform mesoporous carbon

440

capsules by carbonization of organosilica nanospheres. Chem. Mater. 2010, 22 (8), 2526-2533,

441

DOI: 10.1021/cm100190a.

442

(15) Li, Y.; Shi, J. Hollow-structured mesoporous materials: chemical synthesis, functionalization 23

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 31

443

and applications. Adv. Mater. 2014, 26 (20), 3176-3205, DOI: 10.1002/adma.201305319.

444

(16) Cheng, W.; Tang, K.; Qi, Y.; Sheng, J.; Liu, Z. One-step synthesis of superparamagnetic

445

monodisperse porous Fe3O4 hollow and core-shell spheres. J. Mater. Chem. 2010, 20 (9),

446

1799-1805, DOI: 10.1039/b919164j.

447

(17) Ikeda, S.; Tachi, K.; Ng, Y. H.; Ikoma, Y.; Sakata, T.; Mori, H.; Harada, T.; Matsumura, M.

448

Selective Adsorption of Glucose-Derived Carbon Precursor on Amino-Functionalized Porous

449

Silica for Fabrication of Hollow Carbon Spheres with Porous Walls. Chem. Mater. 2007, 19 (17),

450

4335-4340, DOI: 10.1021/cm0702969.

451

(18) Li, Y.; Tan, H.; Salunkhe, R. R.; Tang, J.; Shrestha, L. K.; Bastakoti, B. P.; Rong, H.; Takei,

452

T.; Henzie, J.; Yamauchi, Y.; Ariga, K. Hollow carbon nanospheres using an asymmetric triblock

453

copolymer structure directing agent. Chem. Commun. 2016, 53 (1), 236-239, DOI:

454

10.1039/c6cc07360c.

455

(19) Ramasamy, E.; Chun, J.; Lee, J. Soft-template synthesized ordered mesoporous carbon

456

counter electrodes for dye-sensitized solar cells. Carbon 2010, 48 (15), 4563-4565, DOI:

457

10.1016/j.carbon.2010.07.030.

458

(20) Meng, Y.; Gu, D.; Zhang, F.; Shi, Y.; Yang, H.; Li, Z.; Yu, C.; Tu, B.; Zhao, D. Ordered

459

Mesoporous Polymers and Homologous Carbon Frameworks: Amphiphilic Surfactant Templating

460

and

461

10.1002/ange.200501561.

462

(21) Cao, C.; Wei, L.; Su, M.; Wang, G.; Shen, J. “Spontaneous bubble-template” assisted

463

metal–polymeric framework derived N/Co dual-doped hierarchically porous carbon/Fe3O4

464

nanohybrids: superior electrocatalyst for ORR in biofuel cells. J. Mater. Chem. A 2016, 4 (23),

Direct

Transformation.

Angew.

Chem.

2005,

117

24

ACS Paragon Plus Environment

(43),

7215-7221,

DOI:

Page 25 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

465

9303-9310, DOI: 10.1039/c6ta03125k.

466

(22) zhang, H.; Li, X. Interface-mediated fabrication of bowl-like and deflated ballon-like hollow

467

carbon

468

10.1016/j.jcis.2015.04.027.

469

(23) Wang, G. H.; Hilgert, J.; Richter, F. H.; Wang, F.; Bongard, H. J.; Spliethoff, B.;

470

Weidenthaler, C.; Schuth, F. Platinum-cobalt bimetallic nanoparticles in hollow carbon

471

nanospheres for hydrogenolysis of 5-hydroxymethylfurfural. Nat. Mater. 2014, 13 (3), 293-300,

472

DOI: 10.1038/nmat3872.

473

(24)

474

Natural-Precursor-Derived

475

10.1021/acs.chemrev.5b00566.

476

(25) Abbasi, T.; Abbasi, S. A. Biomass energy and the environmental impacts associated with its

477

production and utilization. Renewable Sustainable Energy Rev. 2010, 14 (3), 919-937, DOI:

478

10.1016/j.rser.2009.11.006.

479

(26) Sevilla, M.; Fuertes, A. B. Chemical and structural properties of carbonaceous products

480

obtained by hydrothermal carbonization of saccharides. Chemistry 2009, 15 (16), 4195-4203,

481

DOI: 10.1002/chem.200802097.

482

(27) White, R. J.; Budarin, V.; Luque, R.; Clark, J. H.; Macquarrie, D. J. Tuneable porous

483

carbonaceous materials from renewable resources. Chem. Soc. Rev. 2009, 38 (12), 3401-3418,

484

DOI: 10.1039/b822668g.

485

(28) Yan, H. Soft-templating synthesis of mesoporous graphitic carbon nitride with enhanced

486

photocatalytic H2 evolution under visible light. Chem. Commun. 2012, 48 (28), 3430-3432, DOI:

nanospheres.

Bazaka,

K.;

J.

Colloid

Jacob,

M.

V.;

Nanocarbons.

Interface

Ostrikov,

Sci.

K.

Chem. Rev.

2015,

K.

452,

Sustainable

2016, 116

25

ACS Paragon Plus Environment

(1),

141-147,

Life

DOI:

Cycles

of

163-214, DOI:

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 31

487

10.1039/c2cc00001f.

488

(29) Deng, J.; Li, M.; Wang, Y. Biomass-derived carbon: synthesis and applications in energy

489

storage and conversion. Green Chem. 2016, 18 (18), 4824-4854, DOI: 10.1039/c6gc01172a.

490

(30) Gao, Z.; Zhang, Y.; Song, N.; Li, X. Biomass-derived renewable carbon materials for

491

electrochemical

492

10.1080/21663831.2016.1250834.

493

(31) Zhao, Q.; Wang, X.; Wu, C.; Liu, J.; Wang, H.; Gao, J.; Zhang, Y.; Shu, H. Supercapacitive

494

performance of hierarchical porous carbon microspheres prepared by simple one-pot method. J.

495

Power Sources 2014, 254, 10-17, DOI: 10.1016/j.jpowsour.2013.12.091.

496

(32) Ganguly, R.; Aswal, V.; Hassan, P.; Gopalakrishnan, I.; Kulshreshtha, S. Effect of SDS on

497

the self-assembly behavior of the PEO-PPO-PEO triblock copolymer (EO)20 (PO)70 (EO)20. J.

498

Phys. Chem. B 2006, 110 (20), 9843-9849, DOI: 10.1021/jp0607061.

499

(33) Zhang, Z.; Qin, M.; Jia, B.; Zhang, H.; Wu, H.; Qu, X. Facile synthesis of novel bowl-like

500

hollow carbon spheres by the combination of hydrothermal carbonization and soft templating.

501

Chem. Commun. 2017, 53 (20), 2922-2925, DOI: 10.1039/c7cc00219j.

502

(34) Wu, L. M.; Tong, D. S.; Li, C. S.; Ji, S. F.; Lin, C. X.; Yang, H. M.; Zhong, Z. K.; Xu, C. Y.;

503

Yu, W. H.; Zhou, C. H. Insight into formation of montmorillonite-hydrochar nanocomposite

504

under

505

10.1016/j.clay.2015.06.015.

506

(35) Lian, P.; Wang, J.; Cai, D.; Ding, L.; Jia, Q.; Wang, H. Porous SnO2@C/graphene

507

nanocomposite with 3D carbon conductive network as a superior anode material for lithium-ion

508

batteries. Electrochim. Acta 2014, 116, 103-110, DOI: 10.1016/j.electacta.2013.11.007.

energy

hydrothermal

storage.

conditions.

Mater.

Appl.

Res.

Clay

Lett.

2016,

Sci.

26

ACS Paragon Plus Environment

2016,

5

(2),

119,

69-88,

116-125,

DOI:

DOI:

Page 27 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

509

(36) Sun, X.; Li, Y. Hollow carbonaceous capsules from glucose solution. J. Colloid Interface Sci.

510

2005, 291 (1), 7-12, DOI: 10.1016/j.jcis.2005.04.101.

511

(37) Möller, M.; Harnisch, F.; Schröder, U. Hydrothermal liquefaction of cellulose in subcritical

512

water—the role of crystallinity on the cellulose reactivity. RSC Adv. 2013, 3 (27), 11035-11044,

513

DOI: 10.1039/c3ra41582a.

514

(38) Sevilla, M.; Fuertes, A. B. The production of carbon materials by hydrothermal

515

carbonization of cellulose. Carbon 2009, 47 (9), 2281-2289, DOI: 10.1016/j.carbon.2009.04.026.

516

(39) Liu, X.; Pang, J.; Xu, F.; Zhang, X. Simple approach to synthesize amino-functionalized

517

carbon dots by carbonization of chitosan. Sci. Rep. 2016, 6, 31100, DOI: 10.1038/srep31100.

518

(40) Cao, X.; Ro, K. S.; Chappell, M.; Li, Y.; Mao, J. Chemical Structures of Swine-Manure

519

Chars Produced under Different Carbonization Conditions Investigated by Advanced Solid-State

520

13

521

10.1021/ef101342v.

522

(41) Falco, C.; Perez Caballero, F.; Babonneau, F.; Gervais, C.; Laurent, G.; Titirici, M. M.;

523

Baccile, N. Hydrothermal carbon from biomass: structural differences between hydrothermal and

524

pyrolyzed carbons via

525

10.1021/la202361p.

526

(42) Zhou, T.; Zhou, Y.; Ma, R.; Zhou, Z.; Liu, G.; Liu, Q.; Zhu, Y.; Wang, J. In situ formation of

527

nitrogen-doped carbon nanoparticles on hollow carbon spheres as efficient oxygen reduction

528

electrocatalysts. Nanoscale 2016, 8 (42), 18134-18142, DOI: 10.1039/c6nr06716f.

529

(43) Chen, C.; Wang, H.; Han, C.; Deng, J.; Wang, J.; Li, M.; Tang, M.; Jin, H.; Wang, Y.

530

Asymmetric Flasklike Hollow Carbonaceous Nanoparticles Fabricated by the Synergistic

C Nuclear Magnetic Resonance (NMR) Spectroscopy. Energy Fuels 2011, 25 (1), 388-397, DOI:

13

C solid state NMR. Langmuir

2011, 27 (23), 14460-14471, DOI:

27

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 31

531

Interaction between Soft Template and Biomass. J. Am. Chem. Soc. 2017, 139 (7), 2657-2663,

532

DOI: 10.1021/jacs.6b10841.

533

(44) Wei, J.; Zhou, D.; Sun, Z.; Deng, Y.; Xia, Y.; Zhao, D. A Controllable Synthesis of Rich

534

Nitrogen-Doped Ordered Mesoporous Carbon for CO2 Capture and Supercapacitors. Adv. Funct.

535

Mater. 2013, 23 (18), 2322-2328, DOI: 10.1002/adfm.201202764.

536

(45) Yao, C.; Shin, Y.; Wang, L. Q.; Windisch, C. F.; Samuels, W. D.; Arey, B. W.; Wang, C.;

537

Risen, W. M.; Exarhos, G. J. Hydrothermal dehydration of aqueous fructose solutions in a closed

538

system. J. Phys. Chem. C 2007, 111 (42), 15141-15145, DOI: 10.1021/jp0741881.

539

(46) Wang, S.; Shang, L.; Li, L.; Yu, Y.; Chi, C.; Wang, K.; Zhang, J.; Shi, R.; Shen, H.;

540

Waterhouse, G. I. Metalang, K.; Z W. D.; Arey, B. W.; Wang, C.; Risen, W. M.; Exarhos, G. J.

541

Hydrothermal dehydration of ar Conformal Phototherapy. Adv. Mater. 2016, 28 (38), 8379-8387,

542

DOI: 10.1002/adma.201602197.

543

(47) Chen, D.; Wang, C.; Jiang, F.; Liu, Z.; Shu, C.; Wan, L. J. In vitro and in vivo

544

photothermally enhanced chemotherapy by single-walled carbon nanohorns as a drug delivery

545

system. J. Mater. Chem. B 2014, 2 (29), 4726-4732, DOI: 10.1039/c4tb00249k.

546

(48) Chen, Y.; Xu, P.; Wu, M.; Meng, Q.; Chen, H.; Shu, Z.; Wang, J.; Zhang, L.; Li, Y.; Shi, J.

547

Colloidal RBC-Shaped, Hydrophilic, and Hollow Mesoporous Carbon Nanocapsules for Highly

548

Efficient

549

10.1002/adma.201400303.

550

(49) Luo, H.; Yang, Y.; Mu, B.; Chen, Y.; Zhang, J.; Zhao, X. Facile synthesis of microporous

551

carbon for supercapacitors with a LiNO3 electrolyte. Carbon 2016, 100, 214-222, DOI:

552

10.1016/j.carbon.2016.01.004.

Biomedical

Engineering.

Adv.

Mater.

2014,

28

ACS Paragon Plus Environment

26

(25),

4294-4301,

DOI:

Page 29 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

553

(50) Raymundo-Piñero, E.; Kierzek, K.; Machnikowski, J.; Béguin, F. Relationship between the

554

nanoporous texture of activated carbons and their capacitance properties in different electrolytes.

555

Carbon 2006, 44 (12), 2498-2507, DOI: 10.1016/j.carbon.2006.05.022.

556

(51) Sharma, P.; Bhatti, T. S. A review on electrochemical double-layer capacitors. Energy

557

Convers. Manage. 2010, 51 (12), 2901-2912, DOI: 10.1016/j.enconman.2010.06.031.

558

(52) Zhang, L.; You, T.; Zhou, T.; Zhou, X.; Xu, F. Interconnected hierarchical porous carbon

559

from lignin-derived byproducts of bioethanol production for ultra-high performance

560

supercapacitors.

561

10.1021/acsami.6602774.

562

(53) Hao, P.; Zhao, Z.; Tian, J.; Li, H.; Sang, Y.; Yu, G.; Cai, H.; Liu, H.; Wong, C.; Umar, A.

563

Hierarchical porous carbon aerogel derived from bagasse for high performance supercapacitor

564

electrode. Nanoscale 2014, 6 (20), 12120-12129, DOI: 10.1039/c4nr03574g.

565

(54) Zhang, W.; Jiang, X.; Zhao, Y.; Carne-Sanchez, A.; Malgras, V.; Kim, J.; Kim, J. H.; Wang,

566

S.; Liu, J.; Jiang, J. S.; Yamauchi, Y.; Hu, M. Hollow carbon nanobubbles: monocrystalline MOF

567

nanobubbles and their pyrolysis. Chem. Sci. 2017, 8 (5), 3538-3546, DOI: 10.1039/c6sc04903f.

568

(55) Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J. W.; Potts, J. R.; Ruoff, R. S. Graphene and

569

graphene oxide: synthesis, properties, and applications. Adv. Mater. 2010, 22 (35), 3906-3924,

570

DOI: 10.1002/adma.201001068.

571

(56) Li, Z.; Xu, Z.; Tan, X.; Wang, H.; Holt, C. M.; Stephenson, T.; Olsen, B. C.; Mitlin, D.

572

Mesoporous nitrogen-rich carbons derived from protein for ultra-high capacity battery anodes

573

and supercapacitors. Energy Environ. Sci. 2013, 6 (3), 871-878, DOI: 10.1039/c2ee23599d.

574

(57) Yun, Y. S.; Park, M. H.; Hong, S. J.; Lee, M. E.; Park, Y. W.; Jin, H. J. Hierarchically porous

ACS

Appl.

Mater.

Interfaces

2016,

29

ACS Paragon Plus Environment

8

(22),

13918-13925,

DOI:

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

575

carbon nanosheets from waste coffee grounds for supercapacitors. ACS Appl. Mater. Interfaces

576

2015, 7 (6), 3684-3690, DOI: 10.1021/am5081919.

577

(58) Tang, J.; Liu, J.; Salunkhe, R. R.; Wang, T.; Yamauchi, Y. Nitrogen-doped hollow carbon

578

spheres with large mesoporous shells engineered from diblock copolymer micelles. Chem.

579

Commun. 2016, 52 (3), 505-508, DOI: 10.1039/c5cc07610b.

580

(59) Zhu, Y.; Murali, S.; Stoller, M. D.; Ganesh, K.; Cai, W.; Ferreira, P. J.; Pirkle, A.; Wallace, R.

581

M.; Cychosz, K. A.; Thommes, M. Carbon-based supercapacitors produced by activation of

582

graphene. science 2011, 332 (6037), 1537-1541, DOI: 10.1126/science.1200770.

583

(60) Xu, X.; Zhou, J.; Nagaraju, D. H.; Jiang, L.; Marinov, V. R.; Lubineau, G.; Alshareef, H. N.;

584

Oh, M. Flexible, Highly Graphitized Carbon Aerogels Based on Bacterial Cellulose/Lignin:

585

Catalyst-Free Synthesis and its Application in Energy Storage Devices. Adv. Funct. Mater. 2015,

586

25 (21), 3193-3202, DOI: 10.1002/adfm.201500538.

30

ACS Paragon Plus Environment

Page 30 of 31

Page 31 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

587

ACS Sustainable Chemistry & Engineering

TOC graph “For Table of Contents Use Only”

588

589 590 591

Synopsis: The dynamic formation process from bowl to sphere’s shape and mechanism of hollow

592

carbon spheres were revealed.

31

ACS Paragon Plus Environment