You are on page 1of 10

Materials Research Express

PAPER

Facile preparation of porous carbon nanospheres via hydrothermal


method using chlorinated polypropylene as precursor
To cite this article: Huazhang Wei et al 2019 Mater. Res. Express 6 0950b8

View the article online for updates and enhancements.

This content was downloaded from IP address 137.111.162.20 on 03/03/2020 at 18:22


Mater. Res. Express 6 (2019) 0950b8 https://doi.org/10.1088/2053-1591/ab07dd

PAPER

Facile preparation of porous carbon nanospheres via hydrothermal


RECEIVED
15 November 2018
method using chlorinated polypropylene as precursor
REVISED
17 January 2019
ACCEPTED FOR PUBLICATION
Huazhang Wei1,2, Weiwei Liu1, Tengfei Liu1, Qian Li1,3 and Huayi Li1,3
18 February 2019 1
Beijing National Laboratory of Molecular Sciences, CAS Key Laboratory of Engineering Plastics, Institute of Chemistry, Chinese Academy
PUBLISHED of Sciences, Beijing 100190, People’s Republic of China
2
31 July 2019 University of Chinese Academy of Sciences, Beijing 100049, People’s Republic of China
3
Authors to whom any correspondence should be addressed.
E-mail: liqian2010@iccas.ac.cn and lihuayi@iccas.ac.cn

Keywords: carbon nanospheres, chlorinated polypropylene, hydrothermal method, porous materials

Abstract
A facile approach to prepare carbon nanomaterials with the spherically shaped morphology and high
BET surface area via the hydrothermal treatment combining high temperature carbonization using
chlorinated polypropylene as precursor was reported in this work. The polymer nanospheres (PNSs)
were firstly prepared by nanoprecipitation, and the effects of concentration, temperature and
pH value on the size and size distribution of the PNSs were studied. The results showed that the
particle size can be controlled by adjusting the experimental conditions, and PNSs with average
diameters of about 100 nm were finally obtained. Then the obtained PNSs were transformed into
carbon nanospheres (CNSs) by hydrothermal treatment combining carbonization, and the carboniza-
tion at different temperatures was investigated. Results showed that the hydrothermal treatment was
essential to firm the spherical shapes of the products, and the high temperature carbonization enabled
the products with high carbon contents. As a result, CNSs with good conductivity and porous
structures were obtained. The electrical resistivity value of the as-prepared CNSs can be down to
75.08 mΩ·cm because of the formation of graphite structures, and the BET surface area of CNS-800
is up to 1249 m2 g−1 with the narrow-distributed mesoporous structures of about 4 nm and the pore
volume of 0.91 cm3 g−1.

1. Introduction

As one kind of the most important carbon materials, carbon spheres perform various outstanding
performances, including low density, high surface area, thermal stability, unique electronic property and the
tailored architectures [1], and they are believed to have wide potential applications in Batteries [2], capacitors [3],
fuel cells [4, 5], catalyst carrier [6] and composite materials [7, 8]. Carbon spheres have been prepared by many
different approaches [9], such as chemical vapor deposition (CVD) [10], high pressure carbonization [11], arc-
discharge [12] and the reduction of carbon dioxide [13]. However, the experimental conditions of these
approaches are generally harsh and the instruments used are relatively complex. To overcome these
disadvantages, hydrothermal treatment, by which the operation process is simple, was introduced, and has been
proved to be one of the greenest approaches to prepare carbon spheres [14].
Various water-soluble organic substances and carbon enriched solid products have been obtained through
treating the organic substances, such as saccharides (glucose [15], sucrose [16] or starch [17]) and simple furfural
compounds [18], under the conditions of temperature ranging from 150 to 350 °C at autogenous pressure. For
example, Wang et al successfully prepared carbonaceous microspheres with controlled morphology and size (at
least 1.5 μm) through the hydrothermal treatment of sucrose for the first time [16]. Chen et al reported a kind of
monodisperse carbonaceous spheres with diameters ranging from 160 to 400 nm, and they were prepared via a
developed two-step hydrothermal method under the monomer concentration range of 0.1–0.4 M [19]. Besides,
through the hydrothermal carbonization of cellulose at 220 °C–250 °C, Sevilla et al prepared the carbon spheres

© 2019 IOP Publishing Ltd


Mater. Res. Express 6 (2019) 0950b8 H Wei et al

with diameters being 2–10 μm [20]. It was found that the temperature of hydrothermal treatment determined
the types and morphologies of the products.
Although there are quite number of reports involving the carbon spheres prepared by hydrothermal
method, they are mainly concentrated on the study of carbon micron-grade spheres and the reports about
carbon nanoscale nanospheres are rare [21, 22]. Zhao et al prepared porous carbon spheres with particle sizes
between 80 and 400 nm via hydrothermal synthesis by using resorcin and formaldehyde as carbon sources and
surfactant F127 as templates [21]. Li et al used α-cyclodextrin as carbon source and surfactant F127 as template
to prepare hollow porous carbon spheres with controllable particle size through hydrothermal synthesis
combining high temperature carbonization [22]. However, the methods mentioned above needed to remove the
template, and the preparation process were relatively tedious. Actually, preparation of carbon nanospheres
through a facile and high efficient method is still a challenge.
Taking into account that mentioned above, we designed a facile route of hydrothermal treatment combining
high temperature carbonization to prepare the monodisperse carbon nanospheres with porous structures in this
work. Chlorinated polypropylene was specially selected as precursor because its high chlorine content is
beneficial to the formation of porous structures [23]. Firstly, chlorinated polypropylene polymer nanospheres
(PNSs) were obtained through precipitation method. Then hydrothermal treatment of the as-prepared PNSs
was conducted to obtained the primary carbonized products, and subsequently, the products were transformed
into carbon nanospheres (CNSs) through the high temperature carbonization. The chemical and structural
characteristics of the final products CNSs were studied.

2. Experimental

2.1. Materials
Chlorinated polypropylene, with the number-average molecular weight being Mn=7.1×103 g mol−1 and the
polydispersity being PDI=1.7, was supplied by Shandong Weifang Gaoxin Chemical Technology Co., Ltd and
used as received without further purification. The contents of carbon, hydrogen and chloride in the chlorinated
polypropylene were 27.04%, 4.51% and 68.45%, respectively. Tetrahydrofuran (THF, analytical reagent) and
ethanol (analytical reagent) were purchased from Beijing Chemical Works and used as received without further
purification. Graphite powder (D50<400 nm, 99.95% metals basis) was purchased from Shanghai Aladdin
biochemical technology co., LTD and used as received without further purification. Deionized water was used
for all experiments.

2.2. Preparation of chlorinated polypropylene PNSs as the precursors of CNSs


Chlorinated polypropylene PNSs were prepared through nanoprecipitation according to the study of Kurokawa
et al [24], and the specific procedure is described below. Chlorinated polypropylene (1 g) and THF (100 ml) were
added into a round-bottom flask and then the system was inserted in an oil bath at 60 °C with continuous
stirring for 2 h. Subsequently, the obtained solution of chlorinated polypropylene in THF was injected rapidly
into the distilled water (1000 ml) with the stirring speed 800 r min−1. Mixture of highly dispersed PNSs (PNS-1)
in water was obtained after the THF was removed by reduced pressure distillation. The obtained chlorinated
polypropylene PNSs are denoted as PNS-X, where X represents the dosage (gram) of chlorinated polypropylene
dissolved in 100 ml THF with the other experiment conditions unchanged.

2.3. Preparation of CNSs using hydrothermal method combining carbonization


Monodisperse porous CNSs were prepared via hydrothermal treatment combining carbonization using the as-
prepared PNS-1 as precursors, and the typical procedure is carried out as follows: the mixture of highly dispersed
PNSs in water was firstly concentrated to about 0.5 wt% by reduced pressure distillation, and then added into a
certain container to perform the hydrothermal treatment at 350 °C for 8 h. After the hydrothermal treatment,
the black precipitate on the container bottom, which was incompletely carbonized, was collected and then
flushed by anhydrous ethanol three times. Finally, the completely carbonized CNSs were obtained after the black
precipitate was further carbonized in a tube furnace under continuous nitrogen flow, heating rate of
10 °C min−1, and thermal retardation time of 3 h. Carbonization at four different temperatures of 600, 800,
1000 and 1200 °C, respectively, were conducted in this work, and the obtained CNSs will be abbreviated to CNS-
Y, where Y represents the carbonization temperature.

2.4. Characterization
The morphology of the as-prepared nanospheres was characterized by scanning electron microscopy (SEM,
Hitachi S-4800 operated at 10 kV) and transmission electron microscopy (TEM, JEOL JSM-2010 operated at
120 kV). Zeta potentials and diameters of the PNSs were measured by dynamic light scattering (DLS) using a

2
Mater. Res. Express 6 (2019) 0950b8 H Wei et al

Figure 1. Effect of the concentration of the chlorinated polypropylene solutions in THF on the DLS data of the aqueous dispersions of
PNSs: (a) SEM image of PNS-1.

Zetasizer (Nano-ZS, Malvern Instruments ZEN3600). Thermogravimetric analysis (TGA) was conducted with a
Perkin-Elmer Pyris1 TGA. N2 adsorption was studied at −196 °C using an automated gas sorption analyzer
(Quantachrome instruments, autosorb iQ), and the samples were degassed at 200 °C for 4 h under vacuum
before measurement. The specific surface area was determined according to the Brunauer–Emmett–Teller
(BET) method in the relative pressure range of 0.05–0.3. The pore size distribution (PSD) was calculated based
on the Barett-Joyner-Halenda (BJH) method using the adsorption isotherm branch. The x-ray diffraction
(XRD) measurements were recorded using a Japan Rigaku D/max-2500 diffractometer with Cu Kα radiation
(1.5406 Å) at an operating voltage of 40 kV. X-ray photoelectron spectroscopy (XPS) was carried out by means of
a USA VG ESCA LAB 220I-XL spectrometer, using Al Kα (1253.6 eV) radiation from a double anode at 50 w.
Binding energies for the high-resolution spectra were calibrated by setting C 1 s to 284.6 eV. The Raman
spectrum was recorded at ambient temperature on a cofocal laser Raman spectrometer (HORIBA LabRAM HR
Evolution) with an argon-ion at an excitation wavelength of 514.5 nm. A four-contact method was applied to
measure the powder resistivity of as-prepared CNSs. The ST-2722 Semiconductor resistivity of the powder tester
combined with a ST2255 High resistance Weak current Tester (Suzhou Jingge Electronic Co., Ltd) was used. The
powder sample was pressed to slice in a cylinder shaped mould (cross sectional area (S)=1 cm2) with two
stainless-steel plungers. The pressure was set at 10 MPa, the instrument could in situ measure the thickness of
the sample slice (h) under 10 MPa pressure and record the volume electrical resistance (R). Then the electrical
resistivity was calculated by the formula of ρ=RS/h.

2.5. Morphology and size of PNSs prepared by nanoprecipitation


Nanoprecipitation is generally employed to prepare nanospheres, especially the monodisperse ones. For
example, Wu et al have successfully prepared the aqueous dispersion of nanoparticles, including the self-
assembly aggregates of π-conjugated polymers [25, 26]. Herein, nanoprecipitation was performed to prepare the
chlorinated polypropylene PNSs as the precursors of CNSs. Besides, SEM image could show that these
nanoparticles possessed well-spherical morphologies with the average diameters of about 110 nm (figure 1(a)).
The obtained aqueous dispersion of PNS-1 (as-expressed) presents homogeneous white, which is similar to
milk, and can be stable for more than 6 months with no evidence of aggregations.
It is reported that particles size of the polymer nanospheres, which were formed during the
nanoprecipitation, can be controlled by adjusting the concentration and pH values of the polymer solutions, as
well as the precipitant temperatures [24, 27, 28]. In this work, effects of the polymer solutions concentration on
particles size of PNSs and its distribution were firstly studied by DLS, and the results were shown in figure 1. It
was found that the particle sizes of the aqueous dispersions of PNSs gradually increased and its distribution
became wider and wider as the initial concentration of the polymer solutions was increased. Therefore, the
particles size of PNSs could be controlled by adjusting the initial concentration of polymer in THF, and
monodisperse nano-sized polymer spheres could be obtained at a relative low polymer solution concentration.
Spherical nanoparticles were obtained by injecting the solution of chlorinated polypropylene polymers in
THF (1 g/100 ml) into water at different temperatures, and the DLS date and Zeta potential were shown in
figure 2(a). It was found that temperature had little effect on the size of PNS, and diameters of the obtained PNS
prepared at different temperatures were all in a range from 50 to 150 nm. However, the zeta potential of the
mixture of PNS in water changed obviously with the precipitant temperatures as shown in figure 2(a). The

3
Mater. Res. Express 6 (2019) 0950b8 H Wei et al

Figure 2. (a) DLS data and Zeta potential of the aqueous dispersion of PNS-1 prepared at different precipitant temperatures, and (b)
DLS data of the aqueous dispersion of PNS-1 prepared at different pH values at 30 °C.

Figure 3. (a) TGA curve of the raw material chlorinated polypropylene in nitrogen atmosphere, and (b) XPS spectra of CNSs at
different treatment conditions.

absolute value of the zeta potential reached the maximum (−37 mV) when the temperature of water was 30 °C,
indicating that the mixture of PNSs in water prepared at 30 °C was the stablest.
By changing the water pH values ranging from 1 to 13, PNSs with different average diameters were obtained
while keeping the water temperature at 30 °C and the concentration of the chlorinated polypropylene/THF
solution being 1 g/100 ml. As shown in figure 2(b), the DLS results demonstrated that the average diameters of
PNSs were about 2 μm under acidic precipitant conditions, that’s pH values were less than 7. And the particle
sizes of the PNSs decreased sharply when the pH values were equal to or higher than 7, and the average diameters
were about 100 nm. It is clear that when the pH value of water is 10, the particle size of PNSs is the smallest.
Preparation of very small PNSs was the target by precipitation because the purpose of this study was to
obtain carbon nanospheres. The experiment results above showed that the diameters of PNSs prepared under
lower concentration and alkaline conditions were smaller. However, the dosage of organic solvent (THF) was
high and the PNSs output was low when the concentration was 0.5%. Besides, the alkali would damage the
morphology of the carbon nanospheres in the hydrothermal carbonization process. Therefore, PNS-1, which
was prepared under concentration of 1% in distilled water at 30 °C, was specifically selected as the precursor to
prepare CNSs.

2.6. Carbonization study of PNSs at different treatment conditions


TGA analysis was conducted to study the carbonization of chlorinated polypropylene by evaluating its weight
loss during the temperature-rise period. The test was carried out under nitrogen flow (40 cm3 min−1) with a
heating rate of 10 °C min−1 from 30 °C up to 700 °C, and the result was shown in figure 3(a). Chlorine radicals
were generated on the active sites of the chlorinated polypropylene molecule chains once the temperature
reached a point, resulting in the chain reaction of dehydrochlorination as the temperature increased [29].
Therefore, weight loss of the polymer sample occurred, and the initial decomposition temperature was about

4
Mater. Res. Express 6 (2019) 0950b8 H Wei et al

Figure 4. (a) SEM image, (b) TEM image and (c) HRTEM image of the incompletely CNSs after hydrothermal treatment, (d) SEM
image, (e) TEM image and (f) HRTEM image of the as-prepared CNS-800.

200 °C as shown in figure 3(a). Then a lot of HCl was generated and escaped as the temperature increased,
leading to the sharp weight loss of the test sample. The temperature of the maximum weight loss rate was
between 300 and 350 °C, and the carbon yield ratio was 26.5% at 700 °C according to the TGA curve. Based on
the TGA analysis results, temperature of the hydrothermally treatment was set at 350 °C to obtain the primary
carbonized CNSs, and further carbonization were executed at 600 °C–1200 °C to obtain the completely
carbonized CNSs.
XPS spectra were recorded to study the carbonization degree of the as-prepared CNSs at different treatment
conditions. As shown in figure 3(b), all the XPS spectra demonstrated only two elements, that is C (≈285 eV) and
O (≈533 eV). The carbon content of the primary carbonized CNSs after hydrothermal treatment was 79.89%
without Cl residues, and further extending the treatment time has no use for raising the carbon content. It’s
known from literature that the hydroxyl, carboxyl and ketone groups on the nanospheres could be removed by
high temperature carbonization [30]. Therefore, carbonization in a high-temperature furnace was further
conducted to prepare CNSs with high carbon content. It was found that the carbon content of CNSs was
observably raised after the carbonization, and increased as the carbonization temperature increased. The carbon
contents of CNS-600, CNS-800, CNS-1000 and CNS-1200 are 94.23%, 93.09%, 96.44% and 97.77%,
respectively. Whereas the oxygen content was decreased as the carbonization temperature increased, and the
oxygen content of CNS-1200 was only 1.9%. It is clear that the CNSs with high carbon contents are obtained.

2.7. Morphology and crystal structure of CNSs


The morphologies of CNSs were studied by SEM and TEM, and the images are shown in figure 4. At the
temperature of 350 °C, which was set in the hydrothermal treatment, the dehydrochlorination reaction, which
occurred intra- and inter-molecular chains of the chlorinated polypropylene, resulted in a massive release of the
HCl molecules and a large number of tiny pores were formed. At the same time, the aromatization occurring
between the macromolecule chains and the inter molecule bridging would make the chlorinated polypropylene
keep its original morphologies. As shown in figures 4(a) and (b), compared with the PNSs, the spherical
structures of CNSs maintained well after the hydrothermal treatment, and the nanosphere surfaces became
rough because of the formation of porous structures.
PNSs were highly carbonized after hydrothermal carbonization at 350 °C, and the carbon content the
primary carbonized CNSs was up to 79.89% as mentioned above. It was believed that the morphology of the
carbon nanospheres was formed fundamentally. Thus there were no further changes of the morphologies of
CNSs after the high temperature carbonization, and the average diameters of CNSs were still about 100 nm as
shown in figures 4(d) and (e).
The HRTEM images were also recorded to demonstrate the microstructures of CNSs. As shown in
figure 4(c), the structures of the obtained CNSs remained amorphous after the hydrothermal treatment. And the
ordered lattice structures of CNS-800 were presented with the lattice spacing of about 0.33 nm as shown in

5
Mater. Res. Express 6 (2019) 0950b8 H Wei et al

Figure 5. (a) XRD spectra and (b) Raman spectra of CNSs carbonized at different temperatures.

figure 4(f), suggesting that the graphite structures were formed locally during the high temperature
carbonization.
XRD and Raman analysis were carried out in order to further study the graphite structures of the as-prepared
CNSs. According to the index JCPDS No.41-1487, the peaks appeared at 26.4° (2θ), 42.2° (2θ) and 44.5° (2θ)
were assigned to (002), (100) and (101) crystal planes of the hexagonal graphitic carbon, respectively, as shown in
figure 5(a). The presence of high intense reflections corresponds to the highly crystalline nature of the nano-
porous carbon [31–33]. These peaks became stronger as the carbonization temperature increased, indicating
that the degree of graphitization was getting higher and higher [34]. The graphitization degree of the as-prepared
CNSs can be calculated according to Merring and Maire formula [35],
0.3440 - d 002
g= (1)
0.3440 - 0.3354
where d002 represents the interplanar spacing.
The interplanar spacing d002 can be calculated by Bragg equation,
2d 002 sin q = l (2)
where θ represents the angle between the incident x-ray and the corresponding crystal plane, and λ represents
the incident wavelength.
By calculation, the graphitization degrees of CNS-600, CNS-800, CNS-1000 and CNS-1200 were 0.40, 0.61,
0.65 and 0.90, respectively.
In addition, Raman spectra were also recorded to give more information about the structures of the as-
prepared CNSs. As shown in figure 5(b), the peaks appeared at 1342 and 1580 cm−1 are generally called D-band
and G-band, respectively. The intensity ratio of D-band and G-band (ID/IG) is another crucial parameter which
has been used to analysis the degree of graphitization [36]. In this work, the intensity ratio of CNS-600, CNS-
800, CNS-1000 and CNS-1200 are 0.77, 0.76, 1.09 and 0.91, respectively, suggesting that the amorphous carbon
and graphite carbon coexisted in the CNSs. Moreover, as the carbonization temperature increased, the
appearance of the 2D-band at 2685 cm−1 suggested that the graphitic phase had been formed by the ordered
graphene layers.

2.8. Electrical conductivity of CNSs


As is known, materials with graphite structures would have excellent conductivity, and the materials
conductivity was increased as the graphitization degree increased. In this work, the electrical resistivity of the as-
prepared CNSs was measured by a resistivity tester using the four-contact method [37], and the results were
shown in figure 6. As mentioned in the section of characterization, the powder sample CNS was pressed to slice
under the pressure of 10 MPa when the resistivity testing was conducted. Therefore, in all cases, the electrical
resistivity and conductivity referred to the performance of the bulky samples prepared from the as-prepared
CNSs in this paper. It was found that the electrical resistivity of the as-prepared CNSs was decreased as the
carbonization temperature increased, that is to say, the conductivity was increased as the carbonization
temperature increased. In this work, the dosage of CNSs in every single measurement was the same, and the
particle sizes of each CNSs test sample was almost the same. Thus, it would not make a big difference of the
arrangement mode and the thickness of the slices, which were formed under the same test conditions. There is
reason to believe that the difference of the resistivity between each CNSs test sample was only caused by the
difference of the graphitization degree of CNSs. The electrical resistivity of the as-prepared CNSs was decreased
as the graphitization degree of CNSs increased, and this result was consistent with the XRD results. As illustrated,

6
Mater. Res. Express 6 (2019) 0950b8 H Wei et al

Figure 6. The electrical resistivity of CNSs carbonized at different temperatures.

Figure 7. (a) Nitrogen adsorption isotherms, (b) pore size distribution and (c) specific surface area and pore volume of the as-prepared
CNSs.

the electrical resistivity value of CNS-600 was up to 9340 mΩ·cm, and a remarkable decrease was observed with
the carbonization temperature increased, indicating that the conductivity of the obtained CNSs became better.
The electrical resistivity values of CNS-800, CNS-1000 and CNS-1200 are 171.4 mΩ·cm, 81.48 mΩ·cm and
75.08 mΩ·cm, respectively. The electrical resistivity of the graphite powder was also measured under the same
test conditions, and the test value was 39 mΩ·cm. It is well known that graphite is a kind of very good electrical
conductors [38], and the nearby electrical resistivity values of the CNSs carbonized at high temperature
indicated that they would have good electrical conductivity.

2.9. Pore structure analysis of CNSs


Nitrogen sorption isotherms of the as-prepared CNSs were activated at different temperatures and the results
were shown in figure 7(a). Hysteresis loops exhibited on the isotherms suggested that the as-prepared CNSs
possessed high surface area and mesoporous features. Besides, the pore sizes of the CNSs were measured and the
results were shown in figure 7(b). It was found that the pore sizes of all the CNSs were around 4 nm, and the pore
size distribution of CNS-800 was relatively narrow. The specific surface area and pore volume of the as-prepared
CNSs, which were calculated from the desorption by DFT method, changed obviously with different
carbonization temperatures as shown in figure 7(c). The specific surface area of CNS-600 was 615 m2 g−1, and
the pore volume was 0.32 cm3 g−1. The remarkable increase was observed when the sample was carbonized at
temperature 800 °C, and the as-prepared CNS-800 were with specific surface area of 1249 m2 g−1 and pore
volume of 0.91 cm3 g−1. In addition, both the specific surface area and pore volume of the CNSs started to
decrease when the carbonization temperatures were higher than 800 °C, indicating that the pore collapse would
occur at higher temperatures.

7
Mater. Res. Express 6 (2019) 0950b8 H Wei et al

3. Conclusions

In this contribution, the feasibility of fabricating high-quality CNSs using the conventional polymers of
chlorinated polypropylene as precursors was demonstrated. Carbon nanoparticles with well-spherical
morphologies and porous structures have been prepared through the facile and efficient hydrothermal method
combining high temperature carbonization. The as-prepared CNSs with average diameters about 100 nm
possess high carbon contents. Besides, the CNSs products have good conductivities because of the formation of
graphite structures, and the electrical resistivity value can be down to 75.08 mΩ·cm when the carbonization
temperature is 1000 °C. What’s more, the as-prepared CNSs have mesoporous structures with the average size
about 4 nm, and the BET surface area is up to 1249 m2 g−1. Our work was to simplify the search for an extremely
facile and reliable approach for preparing high-quality carbon nanospheres, and the approach would show great
potential for commercial applications.

ORCID iDs

Huayi Li https://orcid.org/0000-0001-7147-4914

References
[1] Deshmukh A A, Mhlanga S D and Coville N J 2010 Carbon spheres Materials Science and Engineering: R: Reports 70 1–28
[2] Hou H S, Craig E B, Jing M I, Zhang Y and Ji X B 2015 Carbon quantum dots and their derivative 3D porous carbon frameworks for
sodium-ion batteries with ultralong cycle life Adv. Mater. 27 7861–6
[3] Jayaraman S, Madhavi S and Aravindan V 2018 High energy Li-ion capacitor and battery using graphitic carbon spheres as an insertion
host from cooking oil J. Mater. Chem. A 6 3242–8
[4] Kim J H, Fang B, Kim M and Yu J S 2009 Hollow spherical carbon with mesoporous shell as a superb anode catalyst support in proton
exchange membrane fuel cell Catal Today 146 25–30
[5] Fang B Z, Kim M, Kim J H and Yu J S 2008 Controllable synthesis of hierarchical nanostructured hollow core/mesopore shell carbon
for electrochemical hydrogen storage Langmuir 24 12068–72
[6] Zeng J, Francia C, Gerbaldi C, Dumitrescu M A, Specchia S and Spinelli P 2012 Smart synthesis of hollow core mesoporous shell
carbons (HCMSC) as effective catalyst supports for methanol oxidation and oxygen reduction reactions Journal of Solid State
Electrochemistry 16 3087–96
[7] Wang Z, Wu B, Gong Q M, Song H H and Liang J 2008 In situ fabrication of carbon nanotube/mesocarbon microbead composites
from coal tar pitch Mater. Lett. 62 3585–7
[8] Yang Q, Wang L, Xiang W, Zhou J F and Deng L 2006 Preparation of core–shell carbon black nanoparticles and their crystallization-
induced orientation Eur. Polym. J. 43 1718–23
[9] Xu L Q, Zhang W Q, Yang Q, Ding Y W, Yu W C and Qian Y T 2005 A novel route to hollow and solid carbon spheres Carbon 43
1090–100
[10] Wu H C, Hong C T, Chiu H T and Li Y Y 2009 Continuous synthesis of carbon spheres by a non-catalyst vertical chemical vapor
deposition Diamond Relat. Mater. 18 601–5
[11] Pol V G, Pol S V, Moreno J M C and Gedanken A 2006 High yield one-step synthesis of carbon spheres produced by dissociating
individual hydrocarbons at their autogenic pressure at low temperatures Carbon 44 3285–92
[12] Schur D V, Dubovoy A G, Zaginaichenko S Y, Adejev V M, Kotko A V, Bogolepov V A, Savenko A F and Zolotarenko A D 2007
Production of carbon nanostructures by arc synthesis in the liquid phase Carbon 45 1322–9
[13] Lou Z, Chen C, Zhao D, Luo S and Li Z 2006 Large-scale synthesis of carbon spheres by reduction of supercritical CO2 with metallic
calcium Chem. Phys. Lett. 421 584–8
[14] Yuan D S and Chen J X 2008 Preparation of monodisperse carbon nanospheres for electrochemical capacitors Electrochemistry
Communications 10 1067–70
[15] Sun X M and Li Y D 2004 Colloidal carbon spheres and their shell/core structures with noble- metal nanoparticles Angewandte
Chemie-Intemational Edition 43 597–601
[16] Wang Q, Li H, Chen L Q and Huang X J 2001 Monodisperse hard carbon spherules with uniform nanopores Carbon 39 2211–4
[17] Zheng M T, Liu Y L, Xiao Y, Zhu Y, Guan Q, Yuan D S and Zhang J X 2009 An easy catalyst-free hydrothermal method to prepare
monodisperse carbon microspheres on a large scale J. Phys. Chem. C 113 8455–9
[18] Liang X, Zeng M and Qi C 2010 One-step synthesis of carbon functionalized with sulfonic acid groups using hydrothermal
carbonization Carbon 48 1844–8
[19] Chen C, Sun X, Jiang X, Niu D, Yu A, Liu Z and Li J G 2009 A two-step hydrothermal synthesis approach to monodisperse colloidal
carbon spheres Nanoscale Res. Lett. 4 971–6
[20] Sevilla M and Fuertes A B 2009 The production of carbon materials by hydrothermal carbonization of cellulose Carbon 47 2281–9
[21] Liu J, Yang T Y, Wang D W, Lu G Q M, Zhao D Y and Qiao S Z 2013 A facile soft-template synthesis of mesoporous polymeric and
carbonaceous nanospheres Nat. Commun. 4 2780–98
[22] Yang Z C, Zhang Y, Wang J, Kong J H, Wong S Y and Li X 2013 Hollow carbon nanoparticles of tunable size and wall thickness by
hydrothermal treatment of α-Cyclodextrin templated by f127 block copolymers Chem. Mater. 25 704–10
[23] Zhang G X, Dai L M and Sun X M 2016 Unconventional carbon: alkaline dehalogenation of polymers yields n-doped carbon electrode
for high-performance capacitive energy storage Adv. Funct. Mater. 26 3340–8
[24] Kurokawa N, Yoshikawa H and Hirota N 2004 Size-dependent spectroscopic properties and thermochromic behavior in poly
(substituted thiophene) nanoparticles ChemPhysChem 5 1609–15
[25] Wu C F, Szymanski C and McNeill J 2006 Preparation and encapsulation of highly fluorescent conjugated polymer nanoparticles
Langmuir 22 2956–60

8
Mater. Res. Express 6 (2019) 0950b8 H Wei et al

[26] Schü tze F, Krumova M and Mecking S 2015 Size control of spherical and anisotropic fluorescent polymer nanoparticles via precise
rigid molecules Macromolecules 48 3900–6
[27] Landfester K, Montenegro R, Scherf U, Guntner R, Asawapirom U, Patil S, Neher D and Kietzke T 2002 Semiconducting polymer
nanospheres in aqueous dispersion prepared by a miniemulsion process Adv. Mater. 14 651–5
[28] Kietzke T, Neher D, Landfester K, Montenegro R, Guntner R and Scherf U 2003 Novel approaches to polymer blends based on polymer
nanoparticles Nat. Mater. 2 408–12
[29] Hsieh T H 1999 Effects of oxygen on thermal dehydrochlorination of poly(vinylidene chloride) Polymer Journal 31 948–54
[30] Gong Y T, Xie L, Li H R and Wang Y 2014 Sustainable and scalable production of monodisperse and highly uniform colloidal
carbonaceous spheres using sodium polyacrylate as the dispersant Chemical Communications 50 12633–6
[31] Kumari T S D, Jebaraj A J J, Raj T A, Jeyakumar D and Kumar T P 2016 A kish graphitic lithium-insertion anode material obtained from
non-biodegradable plastic waste Energy 95 483–93
[32] Tuinstra F and Koenig J L 1970 Raman spectrum of graphite J. Chem. Phys. 53 1126–30
[33] Lee Y H, Pan K C, Lin Y Y, Prem Kumar T and Fey G T K 2003 Lithium intercalation in graphites precipitated from pig iron melts
Mater. Chem. Phys. 82 750–7
[34] Li T Q and Yang X G 2009 Influence of testing position on XRD results of carbon material Aerospace Materials & Technology 4
1007–2330
[35] Inagaki M and Shiraishi M 1951 The evaluation of graphitization degree Carbon Tech. 5 165–75
[36] Ferrari A C and Robertson J 2000 Interpretation of Raman spectra of disordered and amorphous carbon Phys. Rev. B 61 14095–107
[37] Chen Z M, Gu M, Wang F, Gao C, Liu P, Ding Y F, Zhang S M and Yang M S 2019 Conductive TiO2 nanorods via surface coating by
antimony doped tin dioxide Mater. Chem. Phys. 225 181–6
[38] Chung D D L 2002 Review graphite J. Mater. Sci. 37 1475–89

You might also like