Almahdi A. Alhwaige‡
*ab,
Saeed M. Alhassanc,
Marios S. Katsiotisc,
Hatsuo Ishida*b and
Syed Qutubuddin*ab
aDepartment of Chemical and Biomolecular Engineering, Case Western Reserve University, Cleveland, Ohio 44106-7217, USA. E-mail: aaa148@case.edu; sxq@case.edu
bDepartment of Macromolecular Science and Engineering, Case Western Reserve University, Cleveland, OH 44106-7202, USA. E-mail: hxi3@case.edu
cDepartment of Chemical Engineering, The Petroleum Institute, United Arab Emiratis
First published on 20th October 2015
Graphene oxide (GO)-reinforced nanocomposite aerogels of polybenzoxazine prepared via freeze-drying of GO suspensions in benzoxazine precursor solutions have been studied. The synthesis of GO is confirmed using Raman and Fourier transform infrared (FT-IR) spectroscopy. The benzoxazine monomer (SLTB(4HBA-t403)) has been synthesized using 4-hydroxybenzaldehyde as a phenolic component, paraformaldehyde, and tri-functional polyetheramine (Jeffamine T-403) as an amine source. The chemical structure of the benzoxazine monomer is confirmed by nuclear magnetic resonance (1H-NMR) spectroscopy and FT-IR. The interactions of GO and SLTB(4HBA-t403) have been investigated using FT-IR. The morphological and thermal stability of nanocomposite aerogels are examined and compared with the neat polybenzoxazine aerogel. The structures of the aerogels and the effect of GO on the morphology of the aerogels are studied using X-ray diffraction (XRD), scanning electron microscopy (SEM), and transmission electron microscopy (TEM). The effect of GO on the ring-opening polymerization of benzoxazine is also evaluated using differential scanning calorimetry (DSC) whereas the thermal stability of the nanocomposite aerogels is characterized by thermogravimetric analysis (TGA).
Among the cross-linking polymers, polybenzoxazines have gained much attention as a versatile class of attractive polymeric materials.7–11 Although the first synthesis of small molecular weight benzoxazine was reported in 1940s,12 the synthesis and properties of polybenzoxazine derived from monomeric type benzoxazine resins was reported in 1994.13 More recently, main-chain type cross-linkable polybenzoxazines where the oxazine ring is placed in the main chain have been developed.14 The development of polybenzoxazine aerogels was first reported in 2009.15 Meanwhile, several other studies on polybenzoxazine-based polymeric and carbon aerogels containing different monomers have been reported.6,7,16–20 The advantages of benzoxazine-based nanostructured materials include near-zero shrinkage upon polymerization, excellent thermal stability, and high char yields.17,21
Nanocomposite materials are regarded one of the fastest growing fields in the polymer industry.22 Specific properties of polymeric materials can be enhanced by a uniform dispersion of reinforcement in the polymer matrix, in particular using a small quantity of nanofiller.23–27 Many different types of fillers, including clay, carbon black, graphene, carbon nanotubes, carbon fibers, and glass fibers have been investigated for producing materials with improved properties. Growing literature about the still-developing field of polybenzoxazine nanocomposites presents major advances that are worthy of consideration. Polybenzoxazine based nanocomposites have attracted the interest of researchers in film form; however, research in development of polybenzoxazine-based nanocomposite aerogels is limited. Very recently, graphene-reinforced benzoxazine films have been investigated.28–30 The current study, involves the incorporation of GO into aldehyde-terminal benzoxazine matrix in the form of nanocomposite aerogel, which has not yet been reported.
Graphene nanosheets show unusual enhancement on the properties of nanocomposite materials.30–32 GO (Scheme 1a) is a layered material produced by the treatment of natural graphite using strong mineral acids and effective oxidizing agents, which introduce a variety of oxygen-based functional groups to the basal planes and edges of graphene layers.31–37 Graphite is the inexpensive natural source available for production of graphene and GO sheets.30–34,38 Polymer composites with exfoliated graphite have been reported since 1958.31 The forces of interaction between the polymers and GO are primarily dipole–dipole interaction and/or hydrogen bonding arising from the polar groups in the polymer and covalent bond formation with the functional groups of hydroxyl (–OH), epoxide (–COC–), and carboxyl (–COOH) on the surfaces of the GO. For example, graphene-based nanocomposite studies were reported using polymers, such as polybenzoxazine,29,30 polyaniline,32 polystyrene-polyacrylamide copolymer,33 poly(allylamine),39 poly(vinyl alcohol),40 and epoxy.41 The development of carbon-based aerogels from graphene and GO have been reported.42,43
![]() | ||
Scheme 1 Chemical structures of: (a) graphene oxide (GO) and (b) the star-like telechelic benzoxazine monomer (SLTB(4HBA-t403)) used in this work. |
Combination of benzoxazine and GO is a promising approach to develop the field of benzoxazine nanocomposite aerogels with properties not provided by neat polybenzoxazines. In the current study, a star-like telechelic benzoxazine (Scheme 1b) was used with various GO contents for synthesizing novel nanocomposite aerogels. The morphological and thermal stability of these aerogels have been studied in detail. The novelty of the current study involves our report on the first proposed mechanism of interaction between aldehydes-terminal benzoxazine and GO nanosheets.
Star-like telechelic benzoxazine was synthesized from 4-hydroxybenzaldehyde (4HBA, 98%, Aldrich Chemicals, USA), polyetheramine (Star-like Jeffamine T-403, Mn = 403 g mol−1, kindly supplied by Huntsman) and paraformaldehyde (96% w/w, Sigma-Aldrich Chemicals, USA) using 1,4-dioxane as solvent. The telechelic is hereinafter abbreviated as SLTB(4HBA-t403). Glacial acetic acid (CH3COOH), 1,4-dioxane, chloroform, and dimethylsulfoxide (DMSO) were purchased from Fisher Scientific (USA). All chemicals were used without further purification.
X-ray diffraction (XRD) was used to determine the increment of the d-spacing in the GO sheets through intercalation of the cationic organic. X-ray diffractograms of the aerogels were obtained by Cu-Kα radiation (λ = 0.15418 nm) with scanning rate of 0.2 degrees min−1 at room temperature. XRD is performed on dried thin film of aerogels. The aerogel samples were cut into thin disks prior to X-ray measurements. Bragg's equation, λ = 2dsin(θ) where d is the layer spacing and θ is the angle of diffraction, was used to compute the d-spacing for GO nano-sheets.
Fourier-transform infrared (FT-IR) spectroscopy was used to investigate the oxidation of graphene, polymerization of benzoxazine and the interactions between benzoxazine and GO. FT-IR spectra were obtained using a Bomem Michelson MB-100 FT-IR spectrometer with a dry air purging unit and a deuterated triglycine sulfate detector at a resolution of 4 cm−1.
Textural morphology of the aerogels was observed with Scanning Electron Microscopy (SEM) and Transmission Electron Microscopy (TEM). SEM analysis was performed with a FEI Quanta 250 (FEG), coupled with Energy Dispersion X-ray Spectroscopy (EDS). Regarding sample preparation, a thin slice (∼1.5 mm) cut on the vertical axis of each of the cylindrically shaped specimens was mounted on standard aluminum SEM stabs using conductive silver glue. TEM analysis was performed with a FEI Tecnai G20, coupled with EDS and a Gatan GIF 963 energy filtered camera. Samples were crushed using a mortar and pestle and then dispersed in high purity ethanol (99.99%) in an ultrasonic bath for ∼30 s. A drop of the dispersed phase would be deposited on Cu grid covered with a thin layer of amorphous carbon (lacey carbon), following which the grid would be immediately transferred to the TEM holder to avoid contamination.
The apparent bulk density (ρb) of the aerogel was obtained from the aerogel mass (m) and its volume (V) as described in eqn (1)
![]() | (1) |
The values of skeletal density (ρs) were obtained by using helium in gas-displacement pycnometer. The apparent porosity percentage of the aerogels was calculated according to the following relation.44
![]() | (2) |
The thermal stability of the aerogels were evaluated by thermogravimetric analysis (TGA) and differential thermogravimetric analysis (DTGA) using a TA Instruments High Resolution 2950. A sample (5–10 mg) from each obtained aerogel was placed in a platinum pan and heated from room temperature to 900 °C at a temperature ramp rate of 10 °C min−1 under nitrogen atmosphere with a gas flow rate of 60 mL min−1.
FT-IR results confirm the introduction of oxygen-based functional groups on the basal plane and edge of graphene (ESI, Fig. S1†). Expectedly, there is no significant peak found in the natural graphite spectrum, while the spectrum of GO shows characteristic band at 3380 cm−1 (the OH stretching mode), 1720 cm−1 (the CO stretching mode of carboxylic acid groups), 1209 cm−1 (
C–OH stretching vibrations) and 1043 cm−1 (C–O stretching vibrations). These results are in agreement with reported literature.34–37
![]() | ||
Scheme 4 Proposed hydrogen bonding network formed between oxygen-functionality on GO and SLTB(4HBA-t403): (a) hydrogen bonding in polymerized aldehyde-functional benzoxazine [adapted from ref. 56] and (b) hydrogen bonding in polymerized SLTB(4HBA-t403)/GO. |
The structure of SLTB(4HBA-t403) benzoxazine monomer was confirmed with 1H-NMR and FT-IR as shown in the ESI Fig. S2 and S3,† respectively; detailed discussion can be found in S2.2.† FT-IR was used to study the polymerization behaviour of SLTB(4HBA-t403)/GO nanocomposite aerogels. A typical FT-IR profiles of neat SLTB(4HBA-t403) at 75 °C and SLTB(4HBA-t403)/GO-5% aerogel at various temperatures (60, 160, 200 and 900 °C) are shown in the ESI, Fig. S4.† FT-IR spectrum of pristine GO (ESI, Fig. S1†) indicates that the hydroxyl functional groups on the surface of graphene sheets have characteristic band at 3380 cm−1 (the OH stretching mode). FT-IR results of SLTB(4HBA-t403) (ESI, Fig. S4a†) reveals that there is no absorption bands in the range 3400–3000 cm−1. After dispersion of GO in SLTB(4HBA-t403) matrix, a broad band appearing at 3465 cm−1 indicates the presence of hydroxyl functional groups (–OH) of GO.34 The frequency shift from 3380 cm−1 to 3465 cm−1 is likely due to the weakening of OH hydrogen bonds caused by the more hydrophobic environment of GO surrounded by SLTB(4HBA-t403), which is consistent with the exfoliated GO. The samples heated at higher temperatures show more absorption intensities of this peak due to the newly created phenolic –OH groups in the polybenzoxazine chains. The phenolic hydroxyl group is also confirmed by the appearance of new absorption peak around 3300 cm−1.7
From the ESI Fig. S4,† it is obvious that the characteristic bands of the GO functional groups heavily overlap with this newly generated OH band and it makes difficult to study the interaction of GO and benzoxazine by FT-IR. The problem is compounded by the low concentration of GO compared to the amount of polybenzoxazine matrix in all studied samples. Therefore, further experiments were performed to confirm SLTB(4HBA-t403)/GO interactions. The investigation of interfacial interactions between SLTB(4HBA-t403) and GO were evaluated by FT-IR using sample with high concentration of GO (SLTB(4HBA-t403)/GO is 40/60 weight percent), which forms a thin-layer of SLTB(4HBA-t403) on GO nanosheets.
The sample was prepared by blend of a concentration of benzoxazine (40 wt%) and GO (60 wt%) and thoroughly mixed for one min. The solution was casted on KBr pellet and dried at room temperature for FTIR measurements. Then, the casted film was subjected to a ring-opening polymerization of benzoxazine monomer schedule of 2 h at each 100, 125, 175 °C.
Fig. 3 shows FT-IR spectra of neat GO, neat SLTB(4HBA-t403), and treated SLTB(4HBA-t403)/GO nanocomposite at various temperatures. The FT-IR characteristics of GO are separately discussed in detail (ESI, Fig. S1†). The important characteristic infrared absorptions of the neat SLTB(4HBA-t403) structure are clearly observed as discussed (ESI, S2.2 and Fig. S3†). The FT-IR spectrum of adsorbed layer of SLTB(4HBA-t403) on GO sheets (Fig. 3c) shows a characteristic absorption band at 1728 cm−1, which is attributed to the CO stretching band of carboxylic acid-functional groups of GO.34,53 This peak slightly changed with heat treatment due to functional groups attached to the carbonyl with a benzene ring or double bond in conjugation with carbonyl.54 The results in Fig. 3c reveal that mixing of GO to SLTB(4HBA-t403) significantly shifts the absorption band at 1114 cm−1 of polyether chain to lower wavenumber due to hydrogen bonding with the adsorbed water on the GO-functional groups. After thermal treatment from 100 °C up to 175 °C, this band returned to the same position (Fig. 3b) due to the evaporation of moisture.
Moreover, the ester-linkage formation in benzoxazine/epoxy resin blend system has been reported.49 The characteristic FT-IR bands of epoxy band (around 950–900 cm−1) overlap with the absorption bands of benzoxazine, making it difficult to follow the disappearance of epoxy bands upon polymerization. However, the observation of the relative increase in the band at 1728 cm−1 is due to ester group formation. More recently, chemical interaction of GO with phenolic resin has been reported.55 Zhou et al.55 documented that esterification reaction occurred between hydroxyl groups (–OH) on resole and carboxyl groups (–COOH) on GO, as confirmed by the absorption band at 1722 cm−1. Therefore, herein, the absorption band at 1728 cm−1 is attributed to the esterification reaction between phenolic –OH and carboxylic groups of GO after heat treatment. In addition, the absorption bands at 2333 and 2364 cm−1 are assigned to CO2 as a result of the oxidation of aldehyde groups.56 In aldehyde-functional benzoxazines, oxidation of aldehyde groups to form carboxylic groups has been reported.56 Therefore, ester formation resulting from carboxylic groups-epoxy reaction possibly occurred as seen in Scheme 3a. In this case of oxidation of aldehyde groups, the released CO2 would present a great advantage to partially contribute to the porous structure of the aerogels, further investigation will be discussed in details elsewhere.57
Furthermore, during heat treatment of ring-opening polymerization of benzoxazine, the produced phenolic hydroxyl is proposed to react with epoxy groups of GO (etherification). The polyether band at 1114 cm−1 became a broader peak and overlapped with 1161 cm−1, which attributed to the chemical attachment of polybenzoxazine with GO via phenolic hydroxyl-epoxy reaction as proposed in Scheme 3b. The resulting absorption band increased with increasing heat treatment due to production of phenolic hydroxyl groups (–OH). The absorption bands in the range 1608–1583 cm−1 in the FT-IR spectra of SLTB(4HBA-t403)/GO nanocomposites become overlapped bands due to newly appeared CC band from thermal treatment of GO.30 Heat treatment of GO partially decomposes the functional groups resulting in aromatic conjugated bonds of graphene.
Hydrogen bonding is possible between the aldehyde-functional benzoxazines and GO formed. Polybenzoxazines contain extensive intermolecular and intramolecular hydrogen bonding between phenolic hydroxyl –OH and electronegative atoms such as O and N.7,56,58–60 Also, hydrogen bonding interaction between the aldehyde groups and polybenzoxazine has been reported as seen in Scheme 4a.56 In the polymerized aldehyde-functional benzoxazine monomer, the aldehyde groups contain electronegative O atoms, which can form intermolecular hydrogen bonding with H atoms of phenolic hydroxyl groups. Herein, the H atoms of hydroxyl groups of GO may show similar tendency (Scheme 4b). Fig. S5† shows the infrared spectra of the CO stretching (aldehyde groups) of SLTB(4HBA-t403) and SLTB(4HBA-t403)/GO samples. The neat SLTB(4HBA-t403) gives a carbonyl absorption peak at 1688 cm−1 (Fig. S5b†).56,58,59 After adding the GO into the SLTB(4HBA-t403) system (Fig. S5c†), the carbonyl stretching frequency is shifted into 1683 cm−1, corresponding to the hydrogen bonded aldehyde groups of SLTB(4HBA-t403) with functional groups on GO sheets as described in Scheme 4b. This new band (1683 cm−1) gradually shifts into the hydrogen bonded carbonyl (1633 cm−1) with increasing heat treatment of nanocomposite sample (SLTB(4HBA-t403)/GO).58 These results suggest that there interaction in the residual aldehyde groups of polybenzoxazine resins. These results are in agreement with literature.56 As seen in Fig. S5,† the existence of hydrogen bonding makes the frequency of the C
O band (1683 cm−1) decrease in the FT-IR spectrum. The hydrogen bonding results from the interaction of electronegative atoms in the residual aldehyde groups with H atoms of hydroxyl groups of both GO and polybenzoxazine. Also, the phenolic hydroxyl groups of polybenzoxazine may form hydrogen bonding with hydroxyl groups of GO. Further hydrogen bonding may occur between carboxylic groups of GO and H atoms of polybenzoxazine (–OH and –COH). According to these results, one can conclude that polymerized SLTB(4HBA-t403)/GO contains a large amount of intermolecular hydrogen bonding formed by residual aldehyde and phenolic hydroxyl groups with oxygen-functional groups of GO (Scheme 4). Adhikari61 et al. reported that functional groups (–OH, –COOH) are able to form hydrogen bonds with other molecules under appropriate conditions. As result of hydrogen bonds, polymers can be assembled with GO to form a hydrogel matrix.43,62
To conclude the FT-IR results, incorporation of neighboring group of aldehyde moieties in benzoxazine monomers play a role for providing more interaction sites during hybridization of polybenzoxazine with GO. These interactions contribute to the enhanced performance of the aerogels as a new class of nanocomposite aerogels. The strong interaction of benzoxazine chains with GO nanosheets leads to high performance nanocomposites-based materials.
The results show that the d-spacing of graphene layers was increased from 0.34 nm to 0.83 nm after the oxidation process due to the formation of oxygen-functional moieties which are acting as spacers. The challenges are to achieve molecule-level dispersion and maximum interfacial interaction between the nanofiller and the polymer matrix at low loading. The XRD pattern of neat SLTB(4HBA-t403) (Fig. 4c) shows a typical diffraction of an amorphous polymer with a broad peak at 2θ = 18°. In the case of SLTB(4HBA-t403)/GO aerogels, the effect of SLTB(4HBA-t403) on exfoliation of GO also has been evaluated. Up to 5 wt% of GO, no obvious diffraction peak is seen below 2θ = 10°. This means that GO may be fully exfoliated in polybenzoxazine matrix due to the strong interaction between GO and SLTB(4HBA-t403), although the concentration of GO is so low and verification of complete exfoliation only by XRD is difficult. However, the aerogel containing 10 wt% GO shows a broad diffraction peak at 2θ = 5.6°, indicating that polybenzoxazine has intercalated fraction of GO. Observation of intercalated fraction implies the trend to aggregate the nanoparticles starts around this concentration. As a result of polybenzoxazine intercalation, the d-spacing of GO increased from 0.826 nm to 1.58 nm.
Sample code | Bulk density (g cm−3) | Porosity (%) |
---|---|---|
SLTB(4HBA-t403)/GO-0% | 0.1550 | 84.50 |
SLTB(4HBA-t403)/GO-1% | 0.1475 | 85.38 |
SLTB(4HBA-t403)/GO-3% | 0.1477 | 85.65 |
SLTB(4HBA-t403)/GO-5% | 0.1478 | 85.91 |
SLTB(4HBA-t403)/GO-10% | 0.1482 | 86.52 |
It is clear that the pore structure of nanocomposite aerogels is different than that of the organic aerogel as the pore sizes in polymeric aerogel (Fig. 5b) appear larger than those of the GO reinforced aerogel (Fig. 5d). The structure of pores is dependent on the size of the ice crystals during the fabrication of the aerogels. During the freezing, as the ice crystals grow, micro-ribbons form on the walls of the aerogel.1,45 The small size of ice crystal growth leads to narrow pore, resulting in an increase in the surface area. In addition, the GO reinforced aerogels exhibit a more layered structure than the polymeric aerogel as shown by SEM images of lateral sections (side view) of the 5% GO reinforced aerogels (Fig. 5e–g). Furthermore, it is clearly seen that the surface of the layers exhibits increased roughness.1 In addition, the layers extend through the length of the aerogel and connected together via strands to create irregular pores. Fig. 5g illustrates the morphology of the fractured structure of edge between two aerogel pores at high magnification (×40000 and scale of 3 μm). The inner surface of the fracture is very different from that of the outer surface due to the presence of the nanofiller. It is interesting to note that the center area shows rough appearance with what appears as a hierarchical structure, while the GO is distributed on the interior surface of the layer without any aggregation.
Carbon aerogels possess large surface area, which have advantages in catalysis and adsorption applications64 as well as electrodes for supercapacitors.16 Morphological behavior of carbon aerogels has been extensively studied using SEM.16,42,43 Polybenzoxazine-based carbon aerogels and xerogels have been investigated.16,17 Herein, morphology of carbon neat polybenzoxazine and polybenzoxazine/GO nanocomposites aerogels studied. Fig. 6 contains the SEM micrographs of carbonized SLTB(4HBA-t403)/GO-5% aerogel. Fig. 6b reveals that the surface layers become coarser after carbonization. The high magnification SEM images (Fig. 6c–e) show surface topology becomes irregularly wrinkled upon carbonization.
![]() | ||
Fig. 6 SEM micrographs of poly(SLTB(4HBA-t403)/GO-5%) carbon aerogels at different microscope magnifications. |
The possible contribution to these wrinkles may be the crumpling of graphene nanosheets that result from the reduced GO and after the decomposition of soft segments of polybenzoxazine. Therefore, random distribution of GO sheets can be observed (Fig. 6f). The observations made with SEM are in good agreement with XRD results which show that GO is highly exfoliated in the polymeric matrix.
![]() | ||
Fig. 7 TEM (a) and HRTEM (b) images of poly(SLTB(4HBA-t403)) aerogel; inset: SAED pattern (collected at area marked with red rectangle) reveals no crystalline structure. |
![]() | ||
Fig. 8 TEM images of SLTB(4HBA-t403)/GO-5% polymerized aerogel; (a) low magnification, (b) high magnification. |
Fig. 8 shows TEM micrographs of SLTB(4HBA-t403)/GO-5% polymerized aerogel. Concerning the polymerized SLTB(4HBA-t403)/GO-5% aerogel, its morphology is similar to the polybenzoxazine aerogel, exhibiting a wide distribution of pore size and shape. Thin carbon layered structures with no aggregation of GO were observed, as expected from previous work by Alhassan et al.30 Overall, the GO sheets were found to be homogeneously dispersed into the polybenzoxazine amorphous structure. Similar finds are observed for the carbon aerogel containing 5 wt% GO, although in this case the ordering of the GO sheets appeared to be higher; this was observed with HRTEM analysis at the edge of particles, where turbostratic and graphitic layers were observed (see Fig. 9b).65
A strong indication of the increased GO presence was provided with EELS analysis, shown in Fig. 10. Specifically, carbon hybridization was evaluated by acquiring the C K-edge for the three specimens mentioned above: poly(SLTB(4HBA-t403)) aerogel, SLTB(4HBA-t403)/GO-5% polymerized aerogel and SLTB(4HBA-t403)/GO-5% carbon aerogel. The characteristic peaks corresponding to sp2 hybridization (1s → π* at 284 eV and 1s → σ* at 291 eV) are more prevalent in the carbonized and polymerized aerogels, due to GO presence in the matrix. On the other hand, the polybenzoxazine aerogel exhibits significantly lower sp2 hybridization, which is mostly attributed to its amorphous nature.
The overall of microstructure analysis can be summarized as the GO sheets were found to be homogeneously distributed into the polybenzoxazine aerogels. In addition, GO enhances the microstructure and reduces the pore size of polybenzoxazine aerogels, while the surface topology of GO/polybenzoxazine aerogel becomes irregularly wrinkled upon carbonization due to the unique structure of GO.
It has been reported that both phenolic and carboxylic groups exist on the GO surface.66 Also, both phenolic and carboxylic acids are effective benzoxazine polymerization initiators and/or catalysts.51,67 As a result, increase in the GO contents decreases the onset polymerization temperature of the SLTB(4HBA-t403)/GO aerogels. The recent report by Gu et al.29 demonstrated that GO has a catalytic effect on the ring-opening polymerization of benzoxazine; however, their DSC investigation did not show de-oxygenation peak of GO. On the other hand, the exothermic peak of GO at 207 °C observed in this study is in agreement with reported results.60,68 Zhao et al.63 found that de-oxygenation of GO has a large single exothermic peak at 200 °C. Furthermore, Alhassan et al.30 reported two exothermic peaks for benzoxazine/GO nanocomposites. It is noteworthy to mention that presence of GO in all cases has catalytic effect on the ring-opening polymerization of the benzoxazine monomer. The results in Fig. 11 reveal that the maximum temperature of the polymerization exotherm on DSC and the area under the peak are significantly shifted to lower values. As an example, 5 wt% GO the onset temperature and maximum temperature of neat SLTB(4HBA-t403) decreased from 142 and 207 °C to 127 and 167 °C, respectively, indicating that GO has significant catalytic activity on benzoxazine polymerization.
DSC results also show that addition of GO to benzoxazine matrix has catalytic effect on reducing the polymerization enthalpies of aldehyde-terminal benzoxazine monomer. Neat aldehyde-terminal benzoxazine and GO show exotherm with enthalpies of 165 and 1105 J g−1, respectively. Accordingly, for instant, theoretical values of enthalpies for benzoxazine polymerization of samples containing 1% and 3% of GO should read 174 and 193 J g−1, respectively. However, the DSC results indicated that the polymerization enthalpies of same nanocomposite samples are 145 and 165 J g−1, respectively.
Sample code | Tmax (°C) | Temperature of decomposition | ||||
---|---|---|---|---|---|---|
Td5 (°C) | Td10 (°C) | T20 (°C) | T50 (°C) | Char yield (%) | ||
Neat GO | 211 | 57 | 109 | 201 | 797 | 49 ± 1 |
0 wt% GO | 396 | 304 | 341 | 382 | 396 | 37 ± 2 |
3 wt% GO | 364 | 280 | 314 | 355 | 498 | 39 ± 2 |
5 wt% GO | 365 | 283 | 316 | 356 | 527 | 42 ± 3 |
10 wt% GO | 367 | 285 | 318 | 363 | 605 | 45 ± 2 |
The results in Table 2 indicated that the maximum degradation temperature of neat GO is 211 °C, which is much lower than that for the neat polybenzoxazine. The lower degradation temperature is attributed to the low thermal stability of soft segment of the functional groups of GO. This investigation is in agreement with literature. The previous studies reported that the thermal decomposition of instable oxygen-containing functional groups occurring at a temperature of 200 °C.34 The TGA results showed that stability of GO is shifted to about 400 °C after incorporation with benzoxazine, which ascribed to the strong interaction of benzoxazine structure with the functional groups of GO as early discussed.
As seen in Table 2, the maximum rate degradation temperature (Tmax) of GO is 211 °C. However, the 5, 10, 20 and 50% weight loss temperatures (Td5, Td10, Td20, and Td50) of poly(SLTB(4HBA-t403)/GO) aerogels increased with GO-loading. For example, Td5 and Td10 for poly(SLTB(4HBA-t403)/GO-3%) are 280 and 314 °C, respectively, whereas Td5 and Td10 for poly(SLTB(4HBA-t403)/GO-10%) are 285 and 318 °C. The improved properties of polymerized nanocomposite aerogels may be due to the strong interactions with GO as discussed earlier. Additionally, the GO reinforced aerogels exhibited much lower degradation behavior at higher temperatures (>400 °C) than the neat polybenzoxazine aerogel, resulting in increased char yields, which is important property for preparing carbon aerogels. As seen in Table 2, the char yield at 800 °C of polybenzoxazine aerogel increased from 37 to 45% by loading 10 wt% GO. The char yield of GO is high since it is mostly carbon. Thus, increased the char yield of the obtained samples is expected. Therefore, polybenzoxazine/GO aerogels not only exhibit improved morphology than the neat polybenzoxazine aerogel, but also show excellent precursor for carbon aerogels. The char yield of benzoxazine aerogels increased significantly with GO content, primarily due to their strong interaction. To show the synergistic effect of GO-loading on the char yield of polybenzoxazine aerogels, the theoretical char yield for each nanocomposite aerogels was calculated. The theoretical char yield was obtained based on the composition of the sample and the actual char yield for neat both GO and polybenzoxazine, poly(SLTB(4HBA-t403)). The theoretical char yield value was calculated according to the eqn (3):
Mtotal% = XPBZMPBZ% + (1 − XPBZ)MGO% | (3) |
In order to investigate whether interactions existed between aldehyde-functional benzoxazine and GO, the theoretical and experimental values of char yield was compared. Fig. 13 displays the expected and experimental char yields of the nanocomposite aerogels. The char yields of the nanocomposites are higher than the theoretical prediction, which is attributed to the synergism between polybenzoxazine and GO due to their strong interactions.
To summarize the TGA results, these nanocomposite aerogels derived from GO-reinforced polybenzoxazine exhibit more efficient for carbon aerogels than aerogels obtained from neat polybenzoxazine due to the increase in the char yield with increasing the GO content. Furthermore, the PBZ aerogels are mechanically stable due to the unique feature of PBZ for cross-linking ability.7 In addition, presence of aldehyde groups enhances PBZ/GO interactions as investigated earlier. As a result of the network structure, the cross-linking contributes stiffness to the aerogels. These advantages make PBZ/GO hybrid aerogels as one of the strong candidates of aerogels for many applications.
Footnotes |
† Electronic supplementary information (ESI) available: See ESI for experimental details of synthesis and characterizations of GO and SLTB(4HBA-t403), additional FT-IR spectra of SLTB(4HBA-t403)/GO-5% hybrid aerogel. See DOI: 10.1039/c5ra16188f |
‡ On leave from Al-Mergib University, Libya, E-mail: aaalhwaige@elmergib.edu.ly |
This journal is © The Royal Society of Chemistry 2015 |