Improvement of the resistance performance of carbon/cyanate ester composites during vacuum electron radiation by reduced graphene oxide modified TiO2

Dequn Penga, Wei Qin*b, Xiaohong Wu*a, Jinzhu wua and Yangyang Panc
aDepartment of Chemistry, Harbin, Heilongjiang 150001, P. R. China. E-mail: wuxiaohong@hit.edu.cn
bSchool of Materials Science and Engineering, Harbin Institute of Technology, Harbin, Heilongjiang 150001, P. R. China. E-mail: qinwei@hit.edu.cn
cShanghai Institute of Spacecraft Equipment, Shanghai, 200240, P. R. China

Received 11th June 2015 , Accepted 28th August 2015

First published on 28th August 2015


Abstract

Electron irradiation in outer space causes severe damage to the polymer materials of spacecrafts. An effective approach to prevent such damage is to incorporate nanoparticles into the polymeric materials. Herein, we fabricated modified cyanate ester (CE) and carbon/CE composites by the incorporation of reduced graphene oxide–TiO2 (rGO–TiO2) nanoparticles and studied their resistance performance to electronic radiation. Compared with the carbon/TiO2/CE composite, the interlayer shear strength of the resulting carbon/rGO–TiO2/CE composite increased by 10.4% and its mass loss reduced by 16.5%. Scanning electron microcopy (SEM) images showed that there are more cracks at the fiber and resin interfaces of carbon/CE than at the interfaces of carbon/rGO–TiO2/CE after irradiation. X-ray photoelectron spectroscopy (XPS) investigation showed that irradiation with 160 keV electrons could break the chemical bonds at the surface layer of the pristine CE resin, which is effectively prevented by the incorporation of rGO–TiO2 nanoparticles.


1. Introduction

Polymer materials have been applied widely in spacecrafts because of the advantages of high strength-to-weight ratios, near-zero coefficients of thermal expansion, chemical inertness, and thermo-optical characteristics.1,2 However, polymeric materials are usually damaged by many space environmental threats like UV radiation, atomic oxygen (AO), thermal cycling (TC) and charged particle radiations such as electrons and protons.3–10 Therefore, it is important to foresee these damages and protect polymeric matrix from severe working conditions.

Electron irradiation is one of the severe and inevitable space environmental factors, which is harmful to the durability of polymeric materials of spacecraft. The effects of electron irradiation on the polymer matrix composites have been reported. Gao et al.11 investigated the effect of irradiation with <200 keV electrons on AG-80 epoxy resin. It was found that electron irradiation could lead to the accumulation of electric charges. In the case where the charges are accumulated to a given extent, electric discharging occurs, which could result in ablation on the resin surface. Yu et al.12 studied the effects of electron irradiation on the surface of carbon/bismaleimide. The results indicated that electron irradiation in ultrahigh vacuum environment could change the surface molecular structure and chemical composition of the carbon/bismaleimide, resulting in damage effects such as mass loss and outgassing of the material skin, this may eventually result in the degradation of the material. Fig. 1(a) shows the interactions between the electron and the materials. The electron radiation is bremsstrahlung, in which the electrons interact with the spacecraft materials. Because the mass of electrons is very light, electrons are prone to deviate after colliding with atoms and their propagation path in the material is bent. Three pathways are possible: electrons are absorbed by the material, penetrated into the material, or are scattered backwards. Electrons interact with the spacecraft materials leading to ionization energy loss and radiation energy loss. As the bond energies of polymer molecules are considerably smaller than those of electrons, ionization and excitation of many molecules will occur to influence the performance of the materials.


image file: c5ra11113g-f1.tif
Fig. 1 A schematic showing the electron irradiation of CE (a) and rGO–TiO2/CE (b).

Carbon/cyanate ester (CE) composites have become a complementary option to traditional carbon/epoxy composites in structural aerospace applications because of the high performance of CE resins such as high temperature-resistance, excellent dielectric properties, low water absorption rate, and good compatibility with carbon fibers.13,14 Nevertheless, radiation damage of CE is a major restriction for their practical applications. A promising approach is to incorporate nanoparticles into the polymeric matrix. Jiang et al.15 studied the resistance to charged particles for TiO2/epoxy nanocomposites. Experimental results show that under electron irradiation, the incorporation of TiO2 nanoparticles into the M40/EP648 reduced the mass loss and increased the interlayer shear strength. We recently reported that TiO2 was incorporated into cyanate ester resin to prepare TiO2/CE nanocomposites by casting and curing. Due to the shielding function of the TiO2 modifier, it effectively reduces irradiation damage of the TiO2/CE nanocomposites.16 The electrons incident to the internal of the materials will mostly reflect off the surfaces due to the blockage of TiO2 nanoparticles. Furthermore, some electrons incident to the material will lose their energy via absorption by TiO2 nanoparticles, thereby the degree of damage to the polymeric matrix by electron irradiation is reduced. In addition, recent studies showed that reduced graphene oxide (rGO) can be a reinforcing material in blends with polymers to enhance their electrical conductivities.17,18 Yang et al.19 added rGO to polyvinyl alcohol (PVA) and effectively improved the electrical conductivity of rGO/PVA composites. Therefore, reduced graphene oxide may be an alternative material to prevent the accumulation of electric charges in cyanate resin matrix. The electron irradiation resistance of reduced graphene oxide and TiO2 is shown in Fig. 1(b).

In this study, rGO–TiO2 nanoparticles were fabricated, and the resistance to the electron irradiation of the corresponding prepared carbon/rGO–TiO2/CE composites was investigated. To clarify the damaging effects of the electrons on the materials, the correlation of the mechanical properties and mass loss with the micro structures was analyzed before and after the electron irradiation. Our findings reveal the possibility of developing novel composites with superior radiation resistance to electrons for application in spacecrafts.

2. Experimental

Graphene oxide (GO) was prepared according to the modified Hummers method as described elsewhere.20,21 rGO–TiO2 was synthesized using a procedure described by Shen et al.22 50 mg of GO was added to 30 mL water. The mixture was sonicated for 1 h to obtain a clear brown dispersion of graphene oxide. Subsequently, 50 mg of glucose and 1 mL of ammonium hydroxide were added to the GO solution to obtain part A. 50 mg of tetrabutyl titanate was added to 3 mL of ethanol. The above mentioned mixture was slowly dropped into a mixture of 2 mg ammonium chloride and 2.5 mL of water to obtain part B. Subsequently, part A and part B were mixed. The mixture was placed into an autoclave and heated at 160 °C for 4 h. When the reduction reaction was completed, the as-synthesized product (rGO–TiO2) was isolated by centrifugation, washed with pure water and ethanol several times, and dried at 90 °C for 12 h.

The powder sample of reduced graphene oxide-loaded TiO2 was characterized by X-ray diffraction, Fourier transform infrared spectroscopy (FTIR), transmission electron microscopy (TEM) and X-ray photoelectron spectrometry (XPS) analysis. The crystal phases of the sample were evaluated using a D/max-rB type X-ray diffractometer (XRD) with Cu Kα X-rays in the 2θ range from 5° to 70°. Fourier-transform infrared (FT-IR) spectra of the compounds were recorded on a VECTOR22 type device using the attenuated total reflectance method. The measurements were conducted in the wavenumber range of 4000–500 cm−1, with a resolution of 4 cm−1. Transmission electron microscopy (TEM) images were obtained using an FEI Tecnai G2S-Twin microscope at an accelerating voltage of 200 kV.

A commercial bisphenol A dicyanate ester (BADCy) was used as the matrix material, which was supplied by the Shanghai Institute of Synthetic Resins in China. T700 carbon fibers were chosen as the reinforcement for the composites, and were supplied by TORAY Company in Japan. The prepreg tape for the T700/cyanate composite specimens was prepared to impregnate the carbon fiber directly into the cyanate matrix. Unidirectional T700/CE composites were manufactured using 16 prepreg tapes stacked in an orientation, laid onto the mold and placed in an autoclave for curing. The curing cycles were as follows: heating to 130 °C from room temperature at a heating rate of 1.5–2.0 °C min−1, holding for 40–60 min at 130 °C under a pressure of 0.6–0.7 MPa and then heating to 180 °C for 2 h for curing under a pressure of 0.6–0.7 MPa. Finally, the finished prepreg tape was also postcured for 3 h at 220 °C, and the temperature was allowed to decrease to room temperature thereafter. The volume fraction of T700 fibers in the manufactured composites was approximately 60%. The rGO–TiO2/CE (Fig. S1) and T700/rGO–TiO2/CE composites were also prepared by the same procedure.

The electron irradiation test was performed in a simulator within the energy range of 30–200 keV. Specimens were placed in the vacuum chamber at a pressure of 1.2 × 10−5 Pa. The electron beam was perpendicular to the specimen surface. The electron energy was chosen as 160 keV and the flux as 5.0 × 1011 e s−1 cm−2. The fluence was altered within the range of 1.0 × 1015 to 2.0 × 1016 e cm−2. A liquid-nitrogen screen was placed inside the test chamber, maintaining the environment temperature at 173–193 K.

The interlayer shear strength was measured on an MTS-810 type tester. Interlaminar shear strength tests were conducted in accordance with the GB3357-82 standard, and each of the resulting data point is an average of five separate measurements.

The mass loss of the specimens before and after the irradiation was measured using a high precision microbalance with a sensitivity of 10−6 g. The mass loss ratio was calculated using the following equation:

image file: c5ra11113g-t1.tif
where m0 is the mass of non-irradiated specimens and mt is the mass of the specimens exposed to electron irradiation with different fluence.

The change in the surface morphology of the irradiated specimens was examined using a Phenom G2 pro type optical microscope. The interface analysis was performed with an SU8000 type scanning electron microscope. XPS analyses were performed using a Perkin-Elmer PHI-5700 ESCA system with Al(Kα) X-ray irradiation (energy = 1486.6 eV). A small amount of standard graphite was coated on the surface center of the specimens to calibrate the negative charge effect.

3. Results and discussion

3.1 Characterization of reduced graphene oxide–TiO2 (rGO–TiO2)

FTIR spectroscopy is commonly used to identify interactions between graphene and TiO2 nanoparticles. Fig. 2 shows FTIR transmittance spectra of GO and rGO–TiO2. As shown in the FTIR spectrum of GO, all the representative absorption peaks, including those at 3402 cm−1 (O–H stretching vibration), 1618 cm−1 (C[double bond, length as m-dash]C stretching of aromatic), 1728 cm−1 (C[double bond, length as m-dash]O stretching of COOH groups), 1225 cm−1 (C–O stretching of epoxide groups) and 1050 cm−1 (C–O stretching vibration of alkoxy/COOH), were clearly observed.23–25 As can be seen in the FTIR spectrum of rGO–TiO2, the oxygen-containing functional groups (C[double bond, length as m-dash]O and C–O) of GO decreased dramatically, while the skeletal vibration of the graphene sheets corresponding to the absorption peak at 1553 cm−1 appeared, indicating that the GO sheets were deoxidized to graphene. In addition, a typical peak of Ti–O–C appeared around 800 cm−1,26 indicating that chemical bonds are formed between graphene and TiO2 nanoparticles.
image file: c5ra11113g-f2.tif
Fig. 2 FTIR transmittance spectra of GO and rGO–TiO2.

The formation of GO and rGO–TiO2 was further characterized by XPS. Fig. 3(a) shows the XPS spectrum of GO with the symmetric C 1s peak at 284.4 eV for C–C bonds, while the oxygen-containing carbonaceous bands (C–O) appeared at higher binding energies. For rGO–TiO2 (Fig. 3(b)), the intensities of all C 1s peaks for carbon bound to oxygen decreased dramatically or disappeared, which indicates that GO was efficiently reduced into graphene during the hydrothermal reaction process. In addition, there is a relatively weak peak located at 288.9 eV, attributed to carboxyl carbon (O[double bond, length as m-dash]C–O). This surface functional group indicated that the –OH groups on TiO2 possibly react with the –COOH groups on the GO surface through esterification to form O[double bond, length as m-dash]C–O–Ti bonds,27 which is in accordance with the FT-IR observation discussed above. Fig. 3 (c) and (d) show the XPS of TiO2 and rGO–TiO2 nanocomposites in the Ti 2p3/2, Ti 2p1/2, and O 1s binding energy regions. For pure TiO2, the determined binding energies of Ti 2p3/2 and Ti 2p1/2 are 458.6 and 464.3 eV, respectively, and the O 1s peak at 529.9 eV for TiO2 is mainly attributed to the oxygen in the TiO2 crystal lattice, which is in good agreement with a previously reported data.28 After the formation of rGO–TiO2 nanocomposites, a shift in the Ti 2p and O 1s binding energies was observed in the XPS spectrum. Normally, this type of shift is induced by a change of the chemical state or coordination environment.


image file: c5ra11113g-f3.tif
Fig. 3 High-resolution XPS spectra of C 1s for GO (a) and rGO–TiO2 (b). The Ti 2p XPS spectra of TiO2 and rGO–TiO2 (c). The O 1s XPS spectra of TiO2 and rGO–TiO2 (d).

Fig. 4 shows the XRD patterns of GO and rGO–TiO2 nanocomposite, respectively. A characteristic (002) peak of GO appeared at a 2θ value of 10.5°. Compared to GO, the corresponding (002) peak of the rGO–TiO2 nanocomposite disappeared, further confirming the conversion of GO to reduced graphene in the final composite (Fig. 4(b)). In addition, the peaks located at 25.4°, 37.8°, 48.0°, 53.9° and 62.7° were indexed to the (101), (004), (200), (105) and (204) crystal planes of anatase TiO2. No typical diffraction peaks of graphene were observed for rGO–TiO2 nanocomposites. This is because the main characteristic peak of graphene at 25° overlaps with the (101) peak of anatase TiO2.29 This finding indicates that graphene as a carrier does not affect the crystalline structure of TiO2.


image file: c5ra11113g-f4.tif
Fig. 4 XRD patterns of GO (a) and rGO–TiO2 nanocomposite (b).

To explore the nature of TiO2 on the graphene sheets, the microstructure of rGO–TiO2 was investigated by TEM. Fig. 5 clearly shows that graphene is a solid support for the TiO2 nanoparticles. Moreover, the TiO2 nanoparticles prevent the graphene sheets from agglomeration during the reduction process. During the formation of the nanocomposite, due to the interactions between the hydrophilic functional groups (e.g., –OH, –COOH) of GO plane and the hydroxyl groups of TiO2, TiO2 nanoparticles could chemically adhere onto the surfaces of GO before its reduction to graphene.27 Fig. 5 shows that the TiO2 nanoparticles are well dispersed on the transparent graphene sheets. Based on the TEM images, the diameters of the TiO2 nanoparticles were estimated to be in the range from 10 to 20 nm.


image file: c5ra11113g-f5.tif
Fig. 5 TEM image of rGO–TiO2.

3.2 Alteration of the interlayer shear strength

Fig. 6 shows the effects of the rGO–TiO2 nanoparticles on the interlayer shear strength (ILSS) of the T700/rGO–TiO2/CE composite after electron irradiation. As can be seen from Fig. 6, the ILSS of the T700/CE composite decreased with an increase of the irradiation fluence. This could be due to the fact that electron irradiation may trigger the degradation reactions of the polymer, resulting in decreased interfacial adhesion between the carbon fibers and resin matrix. The incorporation of the rGO–TiO2 nanoparticles significantly enhanced the ILSS of the resulting T700/rGO–TiO2/CE composite after electron irradiation. This could be due to the absorption and reflection of radiations by the nanoparticles. As a result, degradation of the composite is prevented. T700/rGO–TiO2/CE with 3 wt% rGO–TiO2 gave the maximum ILSS value even after intense irradiation. The reason is that the nanoparticles added may disperse the stresses around the cracks and thus prevent their further development. However, excessive nanoparticles can spontaneously agglomerate in the matrix, resulting in reduced mechanical strength of the composites. Compared with the 3% T700/TiO2/CE, the ILSS of T700/rGO–TiO2/CE increased by 10.4%, indicating that the reduced graphene oxide has strengthening effects on the T700/rGO–TiO2/CE composite.
image file: c5ra11113g-f6.tif
Fig. 6 Interlayer shear strength of T700/rGO–TiO2/CE composites with different rGO–TiO2 contents vs. irradiation fluence.

3.3 Alteration of the mass loss ratio

The effects of the nanoparticles on the mass loss of the composites under electron irradiation are shown in Table S1 and Fig. 7. The mass loss ratios of the three experimental composites obviously increased with increasing fluence up to 1.0 × 1016 e cm−2 and then leveled off. Within the exposure range of 1.0 × 1015 to 2.0 × 1016 e cm−2, T700/CE showed the largest mass loss, followed by T700/TiO2/CE and T700/rGO–TiO2/CE. The mass loss of T700/rGO–TiO2/CE was about 16.5% less than that of T700/TiO2/CE due to the addition of reduced graphene oxide.
image file: c5ra11113g-f7.tif
Fig. 7 The mass loss of the composites as a function of the electron irradiation fluence: (a) T700/CE, (b) T700/TiO2/CE, and (c) T700/rGO–TiO2/CE.

At the early stage of electron irradiation with a fluence less than 1.0 × 1016 e cm−2, the mass loss ratio was remarkably increased. This might be explained as follows: (i) volatilization of the absorbed water and residual organic species in CE under high vacuum during the curing process and (ii) outgassing of small volatile molecules in CE induced by electron irradiation. However, as the irradiation fluence is further increased, a carbon-rich layer is formed on the surface of the materials, preventing the ejection of gaseous molecules from the surface layer of the materials.

3.4 Alteration of the surface morphology

Spacecrafts are vulnerable to energetically charged particles in the working space environment. Because the polymer materials used for spacecrafts are insulators, electrons penetrating the thin skin can be trapped in the dielectric materials to cause accumulation of electric charges. When the charges are accumulated to a certain extent, electric discharging occurs, which results in ablation of the resin surface30 and malfunction of electronic equipments inside the spacecraft.

Fig. 8 shows the surface discharge of pristine CE, TiO2/CE and rGO–TiO2/CE after electron irradiation with a fluence of 2.0 × 1016 e cm−2. There are some distinct discharge ablation phenomena on the surface of pristine CE, as shown in Fig. 8(a). However, with the addition of TiO2 nanoparticles, the amounts of discharge stripe were significantly reduced on the surface of TiO2/CE. Because of the faster interfacial charge-transfer rate and larger surface area of graphene, rGO–TiO2/CE exhibited the least discharge stripes on its surface relative to the other samples. This finding indicates that the rGO–TiO2 nanoparticles added can effectively reduce charge accumulation on the resin surface of the spacecrafts, which is required for their safe use.


image file: c5ra11113g-f8.tif
Fig. 8 Optical microscopy images of pure CE (a), TiO2/CE (b) and rGO–TiO2/CE (c) exposed to the electron irradiation with the fluence of 2.0 × 1016 e cm−2.

3.5 Alteration of the interface morphology

It is well known that the interface plays a decisive role in the performance of the composites. Therefore, we studied the damaging effects of the electron irradiation on the interfaces of the T700/CE, T700/TiO2/CE and T700/rGO–TiO2/CE composites by SEM.

Fig. 9 shows the morphologies of the interfaces of the T700/TiO2/CE and T700/rGO–TiO2/CE composites before and after the electron irradiation. Fig. 9(a) and (b) present the SEM images of the T700/CE composite before and after the electron irradiation (30[thin space (1/6-em)]000×). Compared with the original samples, obvious cracks appeared in the interfaces between the fibers and resin matrix after irradiation at the fluence of 2.0 × 1016 e cm−2. This finding indicates that the bonding strength of the interface is reduced by the electron irradiation. The adhesion of the fiber and resin matrix after irradiation was improved with the addition of TiO2 nanoparticles into the CE resin, as shown in Fig. 9(c) and (d). This shows that TiO2 effectively prevents irradiation damage of the composites. Fig. 9(e) and (f) show that the fiber still closely bonds with the resin matrix before and after electron irradiation with the addition of rGO–TiO2 nanoparticles. This could be due to the quick charge dispersion onto the surface of reduced graphene oxide, which effectively relieves and eliminates the electrons to prevent further damage to the composites.


image file: c5ra11113g-f9.tif
Fig. 9 Interface morphologies of T700/CE, T700/TiO2/CE and T700/rGO–TiO2/CE composites before and after electron irradiation at the fluence of 2.0 × 1016 e cm−2. Magnification is at 30[thin space (1/6-em)]000×. (a) T700/CE composite before irradiation, (b) T700/CE composite after irradiation, (c) T700/TiO2/CE composite before irradiation, (d) T700/TiO2/CE composite after irradiation, (e) T700/rGO–TiO2/CE composite before irradiation, and (f) T700/rGO–TiO2/CE composite after irradiation.

3.6 Alteration of the surface chemical composition

Fig. 10 shows the XPS C 1s spectra of the specimens before and after the electron irradiation. Four peaks with binding energies of 284.60, 285.70, 286.62, and 288.22 eV were fitted and ascribed to –C–C–, –C–N–, –C–O– and –C–O–C–, respectively. Their concentrations were calculated based on the relevant peak areas and are presented in Table 1.
image file: c5ra11113g-f10.tif
Fig. 10 XPS C 1s spectra of the composites before and after electron irradiation at the fluence of 2.0 × 1016 e cm−2: (a) T700/CE composite before irradiation, (b) T700/CE composite after irradiation, (c) T700/TiO2/CE composite before irradiation, (d) T700/TiO2/CE after irradiation, (e) T700/rGO–TiO2/CE before irradiation, and (f) T700/rGO–TiO2/CE after irradiation.
Table 1 Correlative functional groups of the composites before and after electron irradiation
Samples Fluence (e cm−2) The concentration of correlative functional groups (%)
–C–C– –C–N– –C–O– –C–O–C–
T700/CE 0 75.0 18.4 5.1 1.5
2 × 1016 81.1 13.6 3.9 1.4
T700/TiO2/CE 0 73.2 13.6 8.9 4.3
2 × 1016 78.3 12.3 6.2 3.1
T700/rGO–TiO2/CE 0 77.4 16.9 4.1 1.5
2 × 1016 80.5 15.2 2.8 1.5


Electron irradiation, especially at high irradiation fluence, induces breakdown of the chemical bonds of the polymer surface. The massive intensity of the C–C peak was related to the carbon on the surface layer of the specimens. Under exposure, C–C and C–H bonds are inclined to break to form volatile gaseous radiolytic products, as evidenced by the increased intensity of C–C peaks (Table 1). However, with the addition of reduced graphene oxide, the wrinkled exterior surface of reduced graphene oxide could provide nanoscale space limitations, which enhances the bonding between the polymer chains. Therefore, T700/rGO–TiO2/CE exhibits better resistance under irradiation than other samples.

4. Conclusions

In this study, T700/rGO–TiO2/CE composites were prepared by incorporating the as-synthesized rGO–TiO2 nanoparticles into T700/CE composites. In this designed structure, the electrons could be absorbed and reflected by TiO2 nanoparticles and transferred quickly in the interface by the reduced graphene oxide with larger surface areas. As a result, the small molecules generated from the broken –CH2–CH2– bonds in the surface of the cyanate ester resin during vacuum electron radiation were released. In other words, the damage to the composite caused by the electrons could be effectively relieved and eliminated by rGO–TiO2. Under a fluence of 2.0 × 1016 e cm−2, the mass loss caused by the electrons is reduced by 16.5% after rGO–TiO2 modification. More importantly, the ILSS of the composites was also increased after modification. Our findings demonstrate that the fabricated T700/rGO–TiO2/CE composite has promising radiation resistance properties, which are required for space applications.

Acknowledgements

This study was financially supported by the National Natural Science Foundation of China (No. 51078101, 51173033), the Fundamental Research Funds for the Central Universities (NO. HIT. BRETIII. 201224 and 20132), and Shanghai key Laboratory of Deep Space Exploration Technology (No. 13dz2260100).

References

  1. M. Lv, Q. H. Wang, T. M. Wang and Y. M. Liang, Effects of atomic oxygen exposure on the tribological performance of ZrO2-reinforced polyimide nanocomposites for low earth orbit space Applications, Composites, Part B, 2015, 77, 215–222 CrossRef CAS PubMed.
  2. R. L. Kiefer, R. A. Anderson, M. H. Y. Kim and S. A. Thibeault, Modified Polymeric Materials for Durability in the Atomic Oxygen Space Environment, Nucl. Instrum. Methods Phys. Res., Sect. B, 2003, 208, 300–302 CrossRef CAS.
  3. K. B. Shin, C. G. Kim, C. S. Hong and H. H. Lee, Prediction of Failure Thermal Cycles in Graphite/Epoxy Composite Materials under Simulated Low Earth Orbit Environments, Composites, Part B, 2000, 31, 223–235 CrossRef.
  4. Y. Gao, S. He, D. Z. Yang, Y. Liu and Z. J. Li, Effect of Vacuum Thermo-Cycling on Physical Properties of Unidirectional M40J/AG80 Composites, Composites, Part B, 2005, 36, 351–358 CrossRef PubMed.
  5. T. K. Minton, M. E. Wright, S. J. Tomczak, S. A. Marquez, L. H. Shen, A. L. Brunsvold, R. Cooper, J. M. Zhang, V. Vij, A. J. Guenthner and B. J. Petteys, Atomic Oxygen Effects on POSS Polyimides in Low Earth Orbit, ACS Appl. Mater. Interfaces, 2012, 4, 492–502 CAS.
  6. X. W. Zhang, H. Y. Ren, J. H. Wang, Y. Zhang and Y. Y. Shao, (3-Glycidoxypropyl)-Terminated Silsesquioxane Impact on Nanomechanical Properties of Polyimide Coatings Exposed to Atomic Oxygen, Mater. Lett., 2011, 65, 821–824 CrossRef CAS PubMed.
  7. X. H. Zhao, Z. G. Shen, Y. S. Xing and S. L. Ma, An Experimental Study of Low Earth Orbit Atomic Oxygen and Ultraviolet Radiation Effects on a Spacecraft Material– Polytetrafluoroethylene, Polym. Degrad. Stab., 2005, 88, 275–285 CrossRef CAS PubMed.
  8. Y. Gao, S. L. Jiang, D. Z. Yang, S. Y. He, J. D. Xiao and Z. J. Li, A Study on Radiation Effect of <200 keV Protons on M40J/Epoxy Composites, Nucl. Instrum. Methods Phys. Res., Sect. B, 2005, 229, 261–268 CrossRef CAS PubMed.
  9. Q. Yu, P. Chen, J. J. Mu, Y. Gao and D. Liu, Surface Molecular Degradation of High Performance Carbon/Bismaleimide Composites Induced by Proton Irradiation, Nucl. Instrum. Methods Phys. Res., Sect. B, 2011, 269, 318–323 CrossRef CAS PubMed.
  10. G. Pippin, Space Environments and Induced Damage Mechanisms in Materials, Prog. Org. Coat., 2003, 47, 424–431 CrossRef CAS PubMed.
  11. Y. Gao, S. L. Jiang, M. R. Sun, D. Z. Yang, S. Y. He and Z. J. Li, Effect of Irradiation with <200keV Electrons on AG-80 Resin, Radiat. Phys. Chem., 2005, 73, 348–354 CrossRef PubMed.
  12. Q. Yu, P. Chen, Y. Gao, D. Liu and J. J. Mu, Surface Analysis Of High Performance Carbon/Bismaleimide Composites Exposed to Electron Irradiation, Surf. Interface Anal., 2011, 43, 1610–1615 CrossRef CAS PubMed.
  13. J. W. Li, Z. X. Wu, C. J. Huang, H. M. Liu, R. J. Huang and L. F. Li, Mechanical Properties of Cyanate Ester/Epoxy Nanocomposites Modified with Plasma Functionalized MWCNTs, Compos. Sci. Technol., 2014, 90, 166–173 CrossRef CAS PubMed.
  14. C. P. R. Nair, D. Mathew and K. N. Ninan, Cyanate Ester Resins, Recent Developments, Adv. Polym. Sci., 2001, 155, 1–99 CrossRef CAS.
  15. L. X. Jiang, S. Y. He, C. D. Li, D. Z. Yang and S. He, A study on resistance to VUV and charged particles for TiO2/epoxy nanocomposites, Proceedings of the 9th International Symposium on Materials in a Space Environment, Noordwijk-Netherlands, June 16-20, 2003 Search PubMed.
  16. W. Qin, D. Q. Peng, X. H. Wu and J. H. Liao, Study on the Resistance Performance Of TiO2/Cyanate Ester Nano-Composites Exposed to Electron Radiation, Nucl. Instrum. Methods Phys. Res., Sect. B, 2014, 325, 115–119 CrossRef CAS PubMed.
  17. X. L. Wang, H. Bai, Z. Y. Yao, A. R. Liu and G. Q. Shi, Electrically conductive and mechanically strong biomimetic chitosan/reduced graphene oxide composite films, J. Mater. Chem., 2010, 20, 9032–9036 RSC.
  18. S. Bose, T. Kuila, M. E. Uddin, N. H. Kim, A. K. T. Lau and J. H. Lee, In situ synthesis and characterization of electrically conductive polypyrrole/graphene nanocomposites, Polymer, 2010, 51, 5921–5928 CrossRef CAS PubMed.
  19. P. Avouris, Graphene: Electronic and Photonic Properties and Devices, Nano Lett., 2010, 10, 4285–4294 CrossRef CAS PubMed; J. H. Yang and Y. D. Lee, Highly electrically conductive rGO/PVA composites with a network dispersive nanostructure, J. Mater. Chem., 2012, 22, 8512–8517 RSC.
  20. H. K. Chae, D. Y. Siberio-Perez, J. Kim, Y. Go, M. Eddaoudi, A. J. Matzger, M. O'Keeffe and O. M. Yaghi, A Route to High Surface Area, Porosity and Inclusion of Large Molecules in Crystals, Nature, 2004, 427, 523–527 CrossRef CAS PubMed.
  21. W. S. Hummers and R. E. Offeman, Preparation of Graphitic Oxide, J. Am. Chem. Soc., 1958, 80, 1339 CrossRef CAS.
  22. J. F. Shen, M. Shi, N. Li, B. Yan, H. Ma, Y. Hu and M. Ye, Facile Synthesis and Application of Ag-Chemically Converted Graphene Nanocomposites, Nano Res., 2010, 3, 339–349 CrossRef CAS.
  23. J. F. Shen, B. Yan, M. Shi, H. W. Ma, N. Li and M. X. Ye, One Step Hydrothermal Synthesis of TiO2-Reduced Graphene Oxide Sheets, J. Mater. Chem., 2011, 21, 3415–3421 RSC.
  24. C. Z. Zhu, S. J. Guo, P. Wang, Li. Xing, Y. X. Fang, Y. M. Zhai and S. J. Dong, One-Pot, Water-Phase Approach to High-Quality Graphene/TiO2Composite Nanosheets, Chem. Commun., 2010, 46, 7148–7150 RSC.
  25. K. F. Zhou, Y. H. Zhu, X. L. Yang and X. Jiang, Li C. Z. Preparation of Graphene-TiO2Composites with Enhanced Photocatalytic Activity, New J. Chem., 2011, 35, 353–359 RSC.
  26. Z. Y. Wang, B. B. Huang, Y. Dai, Y. Y. Liu, X. Y. Zhang, X. Y. Qin, J. P. Wang, Z. K. Zheng and H. F. Cheng, Crystal facets controlled synthesis of graphene@TiO2 nanocomposites by a one-pot hydrothermal process, CrystEngComm, 2012, 14, 1687–1692 RSC.
  27. Q. J. Xiang, J. G. Yu and M. Jaroniec, Enhanced photocatalytic H2-production activity of graphene-modified titania nanosheets, Nanoscale, 2011, 3, 3670–3678 RSC.
  28. B. J. Jiang, C. G. Tian, Q. J. Pan, Z. Jiang, J. Q. Wang, W. S. Yan and H. G. Fu, Enhanced Photocatalytic Activity and Electron Transfer Mechanisms of Graphene/TiO2with Exposed {001} Facets, J. Phys. Chem. C, 2011, 115, 23718–23725 CAS.
  29. Y. H. Zhang, Z. R. Tang, X. Z. Fu and Y. J. Xu, Engineering the Unique 2D Mat of Graphene to Achieve Graphene–TiO2 Nanocomposite for Photocatalytic Selective Transformation: What Advantage does Graphene Have over Its Forebear Carbon Nanotube?, ACS Nano, 2011, 5, 7426–7435 CrossRef CAS PubMed.
  30. Y. Gao, S. L. Dong, D. Z. Yang, S. Y. He and Z. J. Li, Damage Effects of 120-keV Electron Radiation on AG-80 Resin, J. Polym. Sci., Part B: Polym. Phys., 2006, 44, 177–184 CrossRef PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra11113g

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.