Yanliang Wenab,
Jie Liu*b,
Jiangfeng Song*a,
Jiang Gongb,
Hao Chenb and
Tao Tang*b
aDepartment of Chemistry, College of Science, North University of China, Taiyuan 030051, China. E-mail: jfsong0129@nuc.edu.cn
bState Key Laboratory of Polymer Physics and Chemistry, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, China. E-mail: liujie@ciac.ac.cn; ttang@ciac.ac.cn
First published on 8th December 2015
Conversion of waste polymer on a metal-free catalyst is a promising method for the preparation of nanocarbons. Herein, we synthesized porous carbon sheets and hollow carbon shells through the carbonization of polystyrene on magnesium oxide with different morphologies at 700 °C using a one-pot method. The morphologies, microstructure, phase structure, surface element composition, thermal stability, and textural properties of the obtained nanocarbons were analyzed by SEM, TEM, XRD, TGA, and Raman. The yield of the nanocarbons increased as the weight ratio of magnesium oxide to polystyrene increased. Magnesium oxide acted as a template for the shape-controlled growth of the carbon nanostructure. The surface area of the porous carbon sheets and hollow carbon shells reached 854 and 523 m2 g−1 without any activation, respectively. The porous carbon sheets were used as adsorbents to remove methylene blue from water and showed an adsorption capacity of 358.8 mg g−1. Product composition for the pyrolysis of polystyrene in the presence of magnesium oxide was analyzed using GC and GC-MS to elucidate the reaction mechanism. The yield of styrene in the liquid products reached 50% by the catalysis of polygonal magnesium oxide. This strategy provides a cheap and sustainable catalyst for converting polymer into high-value nanocarbons and useful chemicals.
In one-pot conversion, polymer degradation and the growth of carbon proceed simultaneously. The type and composition of polymer, reaction temperature, reactor type, and catalysts are the important factors affecting the yield and morphology of the obtained nanocarbon. Kong et al. synthesized straight and helical carbon nanotube and Fe3O4@C composite through catalytic decomposition of PE in an autoclave.12 Pol et al. converted waste PE into carbon nanotube and carbon spheres using an autoclave as a reactor under high pressure, which show high performances in lithium electrochemical cells.13 Our group found that the combination of solid acid with nickel based catalyst is efficient for high-yield conversion of plastics into carbon nanotube under atmospheric pressure.14 Solid acid promote the degradation of polymer into light hydrocarbons and aromatics and then the resultant organic compounds react on the surface of nickel based catalyst.15
Although nanocarbon with a diversity of morphologies have been synthesized from different polymers using combined catalysts, the high cost of transition metal prompted us to explore a new facile method to convert waste polymers to nanocarbon. Recently, we synthesized graphene flakes consisting of several to a dozen layers of graphene layers by catalytic carbonization of waste PP using organically modified montmorillonite as degradation catalyst and template at 700 °C.16 This approach provides a novel way to prepare graphene flakes on the metal-free catalyst using waste plastics as carbon sources. However, the purification of the obtained char involves removal of montmorillonite by hydrofluoric acid, which is dangerous, tedious and time consuming. The morphology of the nanocarbon also needs to be improved. The surface area of the obtained nanosheet is low and requires additional activation in potassium hydroxide at 700 °C to increase, which further increased the cost of the technology.
Magnesium oxide (MgO) is a basic alkali earth oxide and can be easily removed by diluted acid. MgO can also be used as a catalyst for the growth of nanocarbon by chemical vapor deposition.17 Porous graphene,18 few-layer graphene shells,19 and nitrogen-doped amorphous graphene20 have been synthesized on MgO substrate by chemical vapor deposition. However, no studies have investigated the conversion of polymer into nanocarbon on MgO substrate.
Herein, we synthesized nanocarbon using PS as carbon source on MgO at 700 °C. The effects of the type and the amount of MgO on the yield, morphology, structure, and textural properties of the obtained nanocarbons were studied. The product composition for pyrolysis of PS in the presence of MgO was analyzed using GC and GC-MS. The potential application of produced nanocarbons in adsorption of MB was also explored.
Fig. 1 Effect of the weight ratio of MgO to PS on the yield of nanocarbon through catalytic carbonization of PS at 700 °C. |
Fig. 3 shows the SEM images of the nanocarbon. In Fig. 3a, PCS stacked together, which may be due to that the carbon deposited on the both sides of sheet MgO. When sheet MgO was removed, space between two graphene layers appeared. Hollow carbon shell was obtained on polygonal MgO (Fig. 3b). However, most of carbon shells were broken, which may be due to that part of polygonal MgO agglomerated during carbonization, leading to that hydrocarbons could not fully contact with polygonal MgO.
Fig. 4 presents HRTEM images of PCS-10, HCS-10, sheet MgO@PCS-10 and polygonal MgO@HCS-10. Stacked graphene layer coated on the surface of MgO. The number of graphene layers ranged from 12 to 16 for both samples, indicating the formation of graphite nanosheet.
Fig. 4 HRTEM micrographs for (a) PCS-10, (b) HCS-10, (c) sheet MgO@PCS-10 and (d) polygonal MgO@HCS-10. |
Based on the above results, it could be demonstrated that both sheet MgO and polygonal MgO acted as the template for the growth of graphite nanosheet during the carbonization of PS.
Raman spectroscopy was used to identify the presence of sp2 carbon. Fig. 6 presents Raman spectra of PCS-10, PCS-1, HCS-10, and HCS-1. All the spectra show typical spectral characteristics of graphitic carbon: the D-band (∼1350 cm−1) and the G-band (∼1580 cm−1). The D-band is associated with vibration of carbon atoms with dangling bonds in the plane terminations of disordered graphite or glassy carbons, indicating the structural defects and partially disordered structures of sp2 carbon. The G-band corresponds to the first-order scattering of an E2g vibrational mode of hexagonal graphite and is related to the vibration of sp2-bonded carbon atoms in a graphite layer, which can be used to explain the degree of graphitization.21,22 The intensity ratio of IG/ID for the four samples was about 0.9, which suggests the obtained nanocarbons are defective. This result may be due to the vacancies or the edges on the graphene layer. The weight of MgO to PS had little effect on the graphitization of the nanocarbons. The degree of graphitization of the porous carbon sheet was increased by increasing reaction temperature (Fig. S2†). The order of the carbon can also be improved by post-heat treatment of the carbon products at high temperatures (>2000 °C) in an inert atmosphere. In addition, incorporation of Fe, Co, and Ni based catalysts in MgO may increase the degree of graphitization. These ideas will be attempted in the following experiment to increase the graphitization of the carbon.
Fig. 7a shows the XPS spectra of PCS-10 and HCS-10. The composition of carbon and oxygen for both two samples was about 97% and 3%, respectively. C1s (284.8 eV) and O1s (532.6 eV) peaks in the survey scan spectra indicated that MgO is completely removed. The XPS spectra of C1s were curve-fitted into three individual peaks (Fig. 7b and c) to determine the chemical component and oxidation state of carbon: graphitic carbon (284.8 eV), C–OH (285.5–285.7 eV), and CO (286.8–287.1 eV).23 Compared with the HCS, the PCS possessed relatively less graphite carbon and more CO groups. The functional groups containing oxygen could contribute to remove heavy metallic ions24,25 or organic dyes26 when PCS or HCS were used as adsorbents in wastewater treatment.
Fig. 7 XPS spectra of the PCS-10 and HCS-10 (a), and deconvoluted C1s spectra of PCS-10 (b) and HCS-10 (c). |
Fig. 8 shows the TGA and DTG curves of PCS-10 and HCS-10 under air. The weight loss of PCS and HCS exhibited two stages. The weight loss from 100 to 300 °C was attributed to the release of chemisorbed water and the pyrolysis of oxygen containing functional groups on the surface of the resultant carbon materials.27 From 400 to 700 °C, a remarkable weight loss was ascribed to the oxidation of the carbon.28 The maximum oxidation temperatures of the two samples were centered at about 630 °C. The residues of them at 800 °C were calculated to be less than 2 wt%, indicating the high purities of both PCS and HCS.
Nitrogen adsorption/desorption experiments were carried out at 77 K to characterize the textural properties of PCS and HCS. The BET surface area (SBET), mesopore surface area (Smeso), total pore volume (Vtotal), mesopore volume (Vmeso), and average pore diameter (DAV) of PCS-10, PCS-1, HCS-10, and HCS-1 are summarized in Table 1. Although the surface area of sheet MgO was only 47 m2 g−1, the surface area of the produced carbon nanomaterial obtained on sheet MgO reached 854 m2 g−1. The weight ratio of PS to MgO had no effect on the specific surface area and average diameter of the pores for both PCS and HCS. The surface area of PCS-10 (854 m2 g−1) was far higher than that of HCS-10 (523 m2 g−1), which can be attributed to the higher surface area of sheet MgO (47 m2 g−1, and that of polygonal MgO was 13 m2 g−1). All the samples exhibited a hysteresis loop (Fig. 9a), which are characteristics of mesopores. The pore size distributions of PCS and HCS were calculated using the Barrett–Joyner–Halenda (BJH) model from the desorption branches of the isotherms (Fig. 9b). Interestingly, no matter what types of MgO or weight ratio of PS to MgO was used, the diameter of mesopores located in a narrow range of 2 to 6 nm (centered at about 3.8 nm). From the above results, we can conclude that the textural properties of the produced nanocarbons are independent of the volume and diameter of pores on MgO. The porosity may come from the assembling of the reactant.
Properties | SBETa (m2 g−1) | Smesob (m2 g−1) | Vtotalc (cm3 g−1) | Vmesod (cm3 g−1) | DAVe (nm) |
---|---|---|---|---|---|
a The total specific surface area.b The specific surface area of mesopores.c The total volume.d The volume of mesopores.e The average diameter of pores. | |||||
Sheet MgO | 47 | 47 | 0.341 | 0.341 | 47.7 |
Polygonal MgO | 13 | 10 | 0.062 | 0.058 | 33.3 |
PCS-10 | 854 | 854 | 3.326 | 3.326 | 3.8 |
PCS-1 | 854 | 854 | 3.672 | 3.672 | 3.8 |
HCS-10 | 523 | 523 | 3.337 | 3.337 | 3.8 |
HCS-1 | 537 | 537 | 3.131 | 3.131 | 3.8 |
PS | PS-sheet MgO two-stagea | PS-polygonal MgO two-stagea | PS-sheet MgO one-potb | PS-polygonal MgO one-potb | |
---|---|---|---|---|---|
a 5 g of PS and 25 g of MgO, separately.b Mixture of 5 g of PS and 25 g of MgO.c Calculated by the volume of the gas first divided by 22.4 L mol−1, then multiply the volume percentage and the molar mass of different gases and finally add them together.d Calculated by the displacement of water.e Calculated by the volume of the gas divided by the total volume of gas products. | |||||
Carbon (wt%) | 0.0 | 0.5 | 0.4 | 12.2 | 9.2 |
Liquid (wt%) | 99.5 | 97.9 | 98.0 | 87.1 | 90.2 |
Gas (wt%)c | 0.2 | 0.3 | 0.3 | 0.4 | 0.4 |
Gas (L/100 g PS)d | 0.3 | 0.6 | 0.4 | 1.3 | 1.1 |
Gas composition (vol%)e | |||||
H2 | 36.7 | 63.5 | 45.3 | 89.1 | 75.8 |
CH4 | 17.5 | 21.6 | 22.9 | 7.2 | 12.7 |
C2H4 | 39.2 | 12.1 | 25.0 | 2.7 | 9.5 |
C3–C5 | 6.6 | 2.8 | 6.9 | 1.1 | 2.0 |
Table 2 also displays the composition of the gas from pyrolysis of PS and PS/MgO. The gaseous products mainly consisted of hydrogen, methane, ethane, ethylene, propane, propylene, i-butene, and pentene. When PS was pyrolyzed and passed through MgO, the yields of ethylene and C3–C5 gases decreased, whereas the yield of hydrogen and methane increased. MgO promote the conversion of C2–C5 gases into hydrogen and methane. Nevertheless, when MgO and PS were decomposed via one-pot method, the yield of hydrogen further increased, whereas the yield of methane decreased, implying that MgO could catalyze the decomposition of methane into carbon and hydrogen.
Fig. 10 shows the GC profiles for the liquid fraction of pyrolyzed products from PS and PS/MgO (sheet and polygonal). The assignment and the area percentages of the peaks are given in Table S2.† These components were grouped into different classes by the number of benzene rings (Table 3). The area percentages of three main peaks are also shown in Table 3. 2,4-Diphenyl-1-butene and 2-phenyl-1,2,3,4-tetrahydronaphthalene are dimmers of styrene. Styrene and its dimers were identified as major products in the liquid. When the degradation products of PS passed through MgO (two-stage), the yield of dimers of styrene decreased, whereas the yield of styrene increased. Considering there was almost no carbon produced, we can conclude that the compounds with two benzene rings cracked into compounds with one benzene ring (including styrene). The compounds with one or two benzene rings were not the carbon source for the growth of nanocarbons. The one-pot method can be considered as a combination of MgO catalyzed degradation of PS and carbonization of degradation products on MgO. The yield of styrene from both one-pot and two-stage process was close. The yield of dimers of styrene decreased, whereas the yield of compounds with three benzene rings increased. These results indicate that MgO promote the degradation of PS into aromatics with three benzene rings or more, which are the main carbon source for the growth of nanocarbon. The increasing trend is more evident on sheet MgO and more carbon is produced at the same time, which verifies the above conclusion.
Substance | PS | PS-sheet MgO two-stagea | PS-polygonal MgO two-stagea | PS-sheet MgO one-potb | PS-polygonal MgO one-potb |
---|---|---|---|---|---|
a 5 g of PS and 25 g of MgO, separately.b Mixture of 5 g of PS and 25 g of MgO. | |||||
27.2 | 42.6 | 51.1 | 36.2 | 50.2 | |
12.3 | 11.5 | 9.5 | 2.3 | 3.3 | |
15.9 | 12.6 | 13.5 | 3.7 | 7.5 | |
1 | 35.0 | 50.3 | 60.3 | 51.8 | 72.3 |
2 | 59.0 | 46.3 | 38.1 | 32.4 | 24.1 |
3 | 4.7 | 2.3 | 0.7 | 14.8 | 3.3 |
4 | 0.6 | 0.5 | 0.4 | 0.8 | 0.2 |
A possible mechanism was proposed to explain the formation of porous carbon sheet and hollow carbon shell through catalytic carbonization of PS by MgO based on the above results. Degradation of PS began with chain scission promoted by heat and basic sites on MgO. From GC-MS results, the compounds with three benzene rings or more were the main carbon source for the growth of nanocarbon on MgO. Styrene and its dimers were identified as major products in the liquid. Polycyclic aromatic hydrocarbons may be generated from styrene via hydrogen abstraction acetylene addition.29 The traditional vapor–liquid–solid mechanism is not suitable for this case because carbon dissolution in MgO is impossible. The reaction probably proceeds through a bottom-up mechanism.30 The polycyclic aromatic hydrocarbons absorbed and assembled on the surface of MgO to form graphene layer. Further carbon deposition led to the stacked graphene layer and thicker shell. Reilly et al. considered that the carbon coating on the catalyst can decompose hydrocarbons but not poison catalyst.31 Thus, this is the reason why many pores existed on the obtained nanocarbon. The shape of the obtained nanocarbon is dependent on the initial morphology of MgO, which verifies the template-shaped role of MgO.
The equilibrium isotherm was introduced to clarify the interaction between adsorbent and adsorbate. The correlation of the experimental results to the adsorption model helps to understand the adsorption mechanism. The Langmuir model was employed to analyze the experimental data for MB adsorption, which is represented by the following equation:
qe = qmKLCe/(1 + KLCe) | (1) |
qe = (C0 − Ce)/W | (2) |
Fig. 11 displays the equilibrium adsorption isotherms of MB on the PCS-10 and HCS-10. The isotherms were characteristic of the Langmuir isotherm, which was belonged to type I curve. The amount of adsorbed MB increased at a lower final solution concentration, which suggests a high affinity between MB molecules and the surface of PCS and HCS. The adsorbed amount then reached a plateau at a higher equilibrium solution concentration, reflecting the saturated adsorption. The qm of the PCS-10 for MB was as high as 358.8 mg g−1, which was larger than that of HCS-10 (238.6 mg g−1). Compared to other adsorbents (Table 4) including CNS,8 CNTs,32 graphene oxide,33 graphene nanosheet,34 activated CS-CNT,2 magnetic Ni/C nonmaterial,35 BN hollow sphere,36 montmorillonite,37 alkali-activated CNTs,38 and activated carbon,39 PCS and HCS showed an excellent adsorption performance of MB without any activation. We believe that both PCS and HCS are the promising materials for many applications to environment management.
Adsorbent | Maximum adsorption capacity (mg g−1) | Reference |
---|---|---|
CNS | 30.3 | 8 |
CNTs | 35.4–64.7 | 32 |
Graphene oxide | 64.23 | 33 |
Graphene nanosheet | 111.62 | 34 |
Activated CS-CNT | 172.4 | 2 |
Magnetic Ni/C nanomaterials | 165.5–175.2 | 35 |
BN hollow sphere | 191.7 | 36 |
Hollow carbon shell (HCS) | 238.6 | Present work |
Montmorillonite | 300.3 | 37 |
Porous carbon sheet (PCS) | 358.8 | Present work |
Alkali-activated CNTs | 400.0 | 38 |
Activated carbon | 452.2 | 39 |
Footnote |
† Electronic supplementary information (ESI) available: Yields of the different fractions for the pyrolysis of PS/sheet MgO at different temperatures (Table S1), TEM images and Raman spectra of nanocarbon obtained on sheet MgO (Fig. S1 and S2), the assignment and area percentage of the compounds in Fig. 10 (Table S2). See DOI: 10.1039/c5ra18505j |
This journal is © The Royal Society of Chemistry 2015 |