Interlock or Chemical Bond: Investigation on the Interface of Graphene

May 23, 2019 - ... to produce composites with excellent properties through chemical ways such as covalent bonding(5) and hydrogen-bonding interactions...
0 downloads 0 Views 5MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 9120−9128

http://pubs.acs.org/journal/acsodf

Interlock or Chemical Bond: Investigation on the Interface of Graphene Oxide and Styrenic Block Copolymers as Layer-by-Layer Films Kun Lei,†,+ Chidao Zhang,†,+ Xinling Wang,† Yunlong Sun,† Haijun Xiao,*,‡ and Zhen Zheng*,† †

School of Chemistry and Chemical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China Department of Orthopedics, Central Hospital of Fengxian District, Sixth People’s. Hospital of Shanghai, Shanghai 201400, China



Downloaded by 91.243.93.94 at 06:24:58:100 on May 24, 2019 from https://pubs.acs.org/doi/10.1021/acsomega.9b00515.

S Supporting Information *

ABSTRACT: In the paper, graphene oxide (GO) and two kinds of styrenic resins, poly[styrene-b-(ethylene-co-butylene)-b-styrene] (SEBS) and maleic anhydride (MA) grafted SEBS (MA-g-SEBS), were utilized to explore the interfacial interaction of carbon-based materials and block copolymers as layer-by-layer (LBL) assembly films. The details of the interlayer interaction of the two kinds of composite films were investigated through the analysis of the mechanical properties and internal structure of the composites. For the SEBS/GO composite film, the “interlock” structure tended to form between the GO sheets and SEBS resin, and the physical “interlocking effect” could make full use of the excellent mechanical properties of GO nanosheets. As a result, both failure strength and elongation at break of the SEBS/GO composite film were enhanced by 50 and 25%, respectively. On the other hand, some different structures were found in the MA-g-SEBS/ GO composite film, where the GO sheets stacked onto the resin closely because of the chemical interaction between them and no obvious “interlocks” was found within the interface, and the chemical interface interaction was strong enough to prevent the slide of GO nanosheets under tension after the graphene sheets were highly oxidized, so the mechanical properties of the MA-gSEBS/GO composite film could be also enhanced. Based on an overall consideration of the research results of these LBL assembled composites, choosing more perfect materials and structures is needed, which should use physical and chemical interfacial interactions more efficiently, to obtain better mechanical properties of inorganic carbon−organic resin composites.

1. INTRODUCTION Composites have been designed to obtain stronger and tougher materials in various fields since the very early time.1,2 Many studies have been done on inorganic fillers and organic resin matrix.3,4 Among them, the interfaces between the filler and matrix is still the key to improvement of the properties of the composites. In most cases, the stronger the interfacial interaction is, the better some properties of the composite are. Therefore, enhancing the interfacial interaction of composites is the common strategy to produce composites with excellent properties through chemical ways such as covalent bonding5 and hydrogen-bonding interactions.6 Zhao et al.7 fabricated ultrathin PVA/GO composite films in which the existence of oxygen functional groups located on both edges and basal plane of GO and the hydroxyl groups in PVA chains makes it possible to apply LBL self-assembly by hydrogen-bonding interaction. Furthermore, the covalent bonding interaction was found in graphene-based nanocomposites.8−10 On the other hand, physical interfacial interaction has considerable effect on the composite properties, which is, however, usually overlooked by many researchers.11−13 Yang et al.14 reported the presence of interlocks between platelets of nacre, and the physical interaction was © 2019 American Chemical Society

proved to be the key mechanism for the high toughness and strength of nacre. In this report, we first fabricated LBL assembled GO-based hybrid films using two kinds of styrenic block copolymer (SBC) resins, aiming to study the effects of both chemical and physical interfacial interactions, and to find out the key factor of improving the properties of composites with the sandwich structure. In recent years, graphene, an atom-thick sheet carbon material, has attracted tremendous scientific interests due to its unique structure and novel electronic, mechanical, and thermal properties.15−21 With extremely high value of Young’s modulus, fracture strength, and elastic modulus, graphene has great potential in the field of composites.18,19,22 However, its high specific surface area and strong π−π stacking have made it an intractable trouble to disperse homogenously in the composite matrix. Besides, with no functional groups on the planar, graphene can hardly form interfacial interaction with the polymer matrix, which is undoubtedly disadvantageous to the properties of composites. In order to achieve homogenous Received: February 24, 2019 Accepted: May 10, 2019 Published: May 23, 2019 9120

DOI: 10.1021/acsomega.9b00515 ACS Omega 2019, 4, 9120−9128

ACS Omega

Article

Figure 1. Schematic diagrams of (a) GO-based composite fabricated by the blending method and (b) GO-based hybrid composite fabricated by LBL assembly.

dispersion of graphene and prevent the restacking of the flat sheets, surfactants can be used to stabilize the suspension,23−26 but the surfactants may give rise to other problems such as property decrease. Also, we can optimize the polymer matrix with graphene prior to the chemical reduction27 in risk of damaging the construction of the materials or the graphene sheets. Obviously, it will remain as a challenge to make full use of graphene’s properties in the field of composites. As a derivative of graphene, GO has emerged as a suitable alternative nanomaterial, bonded to oxygen in the form of hydroxyl and epoxy groups on the planar, as well as carboxyl groups on the edge.28−31 There has been extensive researches on GO-based composites that could be applied in various fields. Negatively charged functionalized graphene oxide layers were incorporated into polyelectrolyte multilayers (PEMs) to fabricate LBL assembled membranes, and micromechanical measurements showed enhancement of the elastic modulus by an order of magnitude, from 1.5 GPa for pure LBL PEM membranes to approximately 20 GPa for only 8.0 vol % graphene oxide encapsulated LBL membranes.32 Bao et al.33 used cross-linked poly(vinyl alcohol) (PVA) and GO to produce flexible, transparent high-barrier composite films, which exhibited excellent properties and prospect for applications in flexible electronics. Despite some decline of properties, the low-cost synthesis and superior solution processability have made GO more possible for applications in material science and composites. It is also important to choose an appropriate matrix to obtain composites with high mechanical performance. Styrenic block copolymer (SBC) is a good choice as a kind of thermoplastic elastomer with favorable processability and recyclability, extensively used in many fields, including protective materials.34 However, as the requirement for strength and modulus grows fast, modification of SBC is necessary to fabricate qualified material, and it is absolutely a promising choice to be incorporated with GO. There has been a few reports on GO/SBC nanocomposites. Cao et al.35 proposed a general and effective methodology to covalently functionalize graphene oxide sheets (GOSs) with SEBS by taking advantage of click chemistry, and the GOSs-c-SEBS could be applied as the filler to enhance polymer composites. Grigorescu et al.36 fabricated the composite with graphene and MA-g-SEBS, improving the compatibility of graphene and the polymer matrix. Esterification reaction took place by a “grafting to” method between maleic anhydride group of MA-g-SEBS and the residual hydroxyl of graphene, enhancing the tensile strength by 69%. Both researches effectively handled the problem of dispersion of the filler through chemical modification, but in order to prepare composites with highly protective properties, both effects of the chemical and physical interfacial interactions between GO and SBC should be made

into full utility to optimize the mechanical properties, which remains a challenging problem beyond the previous researches. Therefore, it is meaningful to determine which kind of interfacial interaction holds an advantage in improving the mechanical properties in different cases, and the answer should be of great importance in the future for the applications of the novel material GO in the field of composites. With keen interest, we started from investigating the relationship between the interface and mechanical properties within the GO/SBC LBL assembled composite. The two kinds of resins used were poly[styrene-b-(ethylene-co-butylene)-bstyrene] (SEBS) and maleic anhydride grafted poly[styrene-b(ethylene-co-butylene)-b-styrene] (MA-g-SEBS), the latter of which could form the chemical interfacial interaction with GO due to the grafted maleic anhydride. LBL assembly has been an extensively accepted method of fabricating organic−inorganic nanocomposites due to its simplicity for assembling functional building blocks with controlled type, morphology, internal organization, and molecular structure.37,38 As shown in Figure 1, it was applied to orient the GO nanoplatelets parallel to the surface of the resin to maximize the attractive energy within the interface.

2. RESULTS AND DISCUSSION SEM image of the cross section of the MA-g-SEBS/GO hybrid film shown in Figure 2a proves that the expected sandwich structure was fabricated. Samples for tensile tests are displayed in Figure 2b,c. As the number of the GO layer increases, the color of the hybrid films turns from colorless to dark brown.

Figure 2. (a) SEM image of the cross section of the MA-g-SEBS/GO hybrid film with three layers of GO. (b) Hybrid films containing 0/1/ 3/5/7 layers of GO. (c) Hybrid film with the size of 40 × 5 mm2 and a gauge length of 20 mm. 9121

DOI: 10.1021/acsomega.9b00515 ACS Omega 2019, 4, 9120−9128

ACS Omega

Article

Figure 3. (a) Stress−strain curves of the pure SEBS film and SEBS/GO hybrid film with four resin layers and three GO layers. (b) Stress−strain curves of the pure MA-g-SEBS film and MA-g-SEBS/GO hybrid film with four resin layers and three GO layers. (c) Tensile strength of SEBS/GO hybrid film samples with different layers of GO. (d) Tensile strength of MA-g-SEBS/GO hybrid film samples with different layers of GO.

hybrid film with the maximum seven layers of GO, the content of GO is only less than 5 wt %. SEBS is a tri-block copolymer without functional groups on the molecular chains, and therefore, there is almost no chemical interfacial interaction within its GO-based nanocomposite. However, the mechanical properties of the SEBS/ GO hybrid film improve a lot compared with the pure SEBS film, as shown in Figure 3, which probably is due to the physical interaction within the interface, namely, the asmentioned interlock effect. Grafted maleic anhydride is the only difference between the two kinds of resins, which makes it possible to react with the oxygen-containing functional groups on the planar of GO. Esterification reaction can take place between maleic anhydride groups of the MA-g-SEBS resin and the hydroxyl groups of GO. In addition, the polar groups of the resin can improve the compatibility. Generally speaking, the chemical interaction would enhance the interface and thus improve the mechanical properties of the composite. However, it is interesting that we do not see that the properties of the MA-g-SEBS/GO hybrid film were enhanced as expected from Figure 3. The stress−strain curves of the pure SEBS film and SEBS/ GO hybrid film with four resin layers and three GO layers are shown in Figure 3a, from which we can see that compared with the pure SEBS film with the fracture stress of 37 MPa and the elongation of 626% at break, the two tensile properties indexes of SEBS/GO film are markedly improved, reaching 58 MPa and 796%, respectively. Tensile tests of SEBS/GO hybrid film

All the film samples for the tensile test were tailored into the shape of rectangle with the size of 40 × 5 mm2 and a gauge length of 20 mm. Filter paper was pasted onto both ends of the samples to prevent the films from slipping out the fixture during tensile tests. Due to the limited amount of GO in hybrid films prepared by the LBL assembly method and the fact that the oxygencontaining groups of GO will undergo pyrolysis as the temperature increases, it is difficult to obtain the mass fraction of GO in hybrid films by thermogravimetric analysis. In the preparation process of hybrid films, the surface density of the GO film did not change during the transfer process, and namely, the surface density of the GO film formed on the surface of deionized water was the same as that of the single layer of GO in the hybrid films. The culture dish with a diameter of 17 cm was used to measure the surface density of the GO film formed on the surface of deionized water. When the water surface is completely covered by the GO film, the required GO solution is 2.5 mL (5 mg/mL). The surface density of the single layer of GO in the hybrid films is calculated as follows: ρ = m/S = (2.5 × 5)/( π × 8.52 ) = 0.055 mg/cm 2

(1)

Every hybrid film sample with the size of 5 × 40 mm2 and a thickness of 100 μm was weighed 20 ± 0.5 mg, and the surface area was 2 cm2. Hence, it is concluded that in the hybrid films containing 0/1/3/5/7 layers of GO, the mass fractions of GO are 0, 0.53, 1.59, 2.65, and 3.71 wt %, respectively. For the 9122

DOI: 10.1021/acsomega.9b00515 ACS Omega 2019, 4, 9120−9128

ACS Omega

Article

Figure 4. (a) Surface roughness of SEBS and SEBS/GO films and (b) MA-g-SEBS and MA-g-SEBS/GO films measured by a profile meter.

Figure 5. AFM tapping mode for surface roughness of (a) SEBS, (b) SEBS/GO, (c) MA-g-SEBS, and (d) MA-g-SEBS/GO films.

samples with different layers of GO were carried out, and the results are exhibited in Figure 3c, from which we can see that with the increase of GO layers, the fracture stress of the SEBS/ GO hybrid film is enhanced first, and when the number of the GO layer reaches three, the value of the fracture stress tends to be constant. The stress−strain curves of the pure MA-g-SEBS film and MA-g-SEBS/GO hybrid film with four resin layers and three GO layers are shown in Figure 3b. Tensile strength of MA-g-SEBS/GO hybrid film samples with different layers of GO is shown in Figure 3d. From the tensile test results of the MA-g-SEBS/GO hybrid film in Figure 3b,d, we do not see the properties of the MA-g-SEBS/GO hybrid film were enhanced as expected. Besides, from Figure 3, it can be seen that in comparison with the MA-g-SEBS-based hybrid films with the fracture stress of less than 10 MPa, the fracture stress of the SEBS-based hybrid films is far higher than that of the MA-g-SEBS-based hybrid films with the fracture stress of more than 10 MPa.

Surface profiler and AFM were used to observe the morphology of the interface in order to find out the key to the 58% enhancement of fracture stress and 27% improvement of elongation at break for SEBS-based hybrid films and the constant fracture stress and elongation at break for MA-gSEBS-based hybrid films. Figure 4 shows the macroscopic surface morphology of SEBS, SEBS/GO, MA-g-SEBS, and MA-g-SEBS/GO in a string of line, emerging like successive peaks and valleys; therefore, the height of peaks (h, Å) and the distance between them (l, μm) can be used to represent the macroscopic roughness of the surface. From Figure 4a, we can see that the surface morphology of the SEBS changed little after incorporation of GO, with the average height of the peaks increases slightly from 17.0 to 18.4 nm, and the average distance decreases a bit from 2.50 to 2.27 μm. While the AFM image in Figure 5a,b exhibits a considerable increase of Ra from 0.781 to 1.458 nm. Both the alteration of surface roughness can be explained by the reactive inertness of SEBS polymer chains, and the mechanism 9123

DOI: 10.1021/acsomega.9b00515 ACS Omega 2019, 4, 9120−9128

ACS Omega

Article

without making full and best use of the ultrahigh mechanical properties of GO. Even if there exist chemical bonds within the interface, the lack of interlocks plays a leading role to the disadvantage of the mechanical properties of MA-g-SEBS/GO hybrid films. Schematic image of the mechanism is shown in Figure 6. Of course, after the graphene sheets were highly oxidized, the chemical interface interaction was strong enough to prevent the slide of GO nanosheets under tension, leading to a slight increase of the mechanical property of the MA-gSEBS/GO film (Table S1). Besides, dynamic thermomechanical analysis (DMA) is indirectly utilized to illustrate the effect of the super mechanical properties of GO on the improvement of the hybrid SEBS/GO films, as shown in Figure 7. It is obvious that the introduction of the GO endows the SEBS films with the improved mechanical properties in Figure 7b (the E′ of the SEBS/GO films is higher than the E′ of the SEBS films, especially in the temperature range below 0 °C). From the results of the dissipation factor (tan δ) characterization in Figure 7a, a distinct difference can be seen with the greater tan δ value of the SEBS/GO films in comparison with that of the SEBS films, indicating that there is better load dissipation capacity for the SEBS/GO films. This better load dissipation capacity is very useful for the loaded materials, effectively dispersing energy and preventing the target materials from breaking. That is to say that compared with the SEBS films, the SEBS/GO films show better mechanical performance, and this result is in accordance with Figure 3. The tan δ−T curves of the SEBS/GO hybrid films with different GO layers are all shown in Figure 8. The −50 °C is the Tg of the EB segment in the copolymer SEBS, and after the introduction of GO, there is no apparent change for the Tg of the EB segment in the copolymer SEBS with different GO layers. It is attributed to the absence of chemical reaction or hydrogen bonding effect between the SEBS resin and GO sheet. Hence, the intermolecular force of the EB segment in the copolymer SEBS is not enhanced, leading to the constant Tg of the EB segment. This also illustrates the absence of the chemical interfacial effect between the SEBS resin and GO sheet. Similarly, dynamic thermomechanical analysis (DMA) is also utilized to demonstrate the effect of the super mechanical properties of GO on the improvement of the hybrid MA-gSEBS/GO films, as shown in Figure 9. As can be seen from the

is briefly explicated in Figure 6. Due to the lack of functional groups, not enough interfacial interaction exists to make the

Figure 6. Schematic images of interlocks between (a) SEBS/GO and (b) MA-g-SEBS/GO films.

GO platelets stack closely to the SEBS resin, so that the platelets distribute onto the resin surface randomly, and only few of them parallels to the surface. On one hand, the general morphology changed little. On the other hand, the loose stack of wrinkled GO platelets gives rise to the formation of interlocks between the platelets and resin, which is in favor of taking advantage of the super mechanical properties of GO to make the hybrid film stronger and tougher. For the constant stress of the MA-g-SEBS/GO films, strangely but interestingly, the phenomenon also can be explained reasonably through the careful observation of the interface of the composites. Figures 4b and 5c,d exhibit macroscopic and microscopic surface roughness alteration, respectively. After incorporation of GO, the surface roughness decreases obviously. As to macroscopic surface roughness, the average height of the peaks goes down from 17.7 to 14.0 nm, and the distance between them narrows down from 2.91 to 1.83 μm. This may be attributed to the chemical interfacial interaction between the resin and GO platelets, which induces the GO platelets to align parallel to the resin surface. Therefore, the valleys are likely to be filled by the GO platelets, resulting in the decrease of the macroscopic surface roughness. In the meanwhile, close stack between the GO platelets and resin makes the interface smoother, so interlocks within the interface are unlikely to form. As a result, the GO platelets and the resin tend to slide apart under tensile force,

Figure 7. Hybrid films with four resin layers and three GO layers for the DMA test. (a) tan δ of SEBS and SEBS/GO films; (b) E′ of SEBS and SEBS/GO films. 9124

DOI: 10.1021/acsomega.9b00515 ACS Omega 2019, 4, 9120−9128

ACS Omega

Article

Figure 8. Effects of temperature on tan δ of SEBS/GO hybrid film samples with different layers of GO: (a) −100−120 °C; (b) −70 to −25 °C.

Figure 9. Hybrid films with four resin layers and three GO layers for the DMA test. (a) tan δ of MA-g-SEBS and MA-g-SEBS/GO films; (b) E′ of MA-g-SEBS and MA-g-SEBS/GO films.

Figure 10. tan δ of MA-g-SEBS/GO hybrid film samples with different layers of GO: (a) −100−120 °C; (b) −70 to −10 °C.

there exists chemical interfacial effect between the MA-g-SEBS resin and GO sheets.

Figure 9, we do not see the properties of the MA-g-SEBS/GO hybrid film were enhanced as expected, and this result is also consistent with the stress−strain curve of MA-g-SEBS/GO hybrid films in Figure 3. The tan δ−T curves of the MA-g-SEBS/GO hybrid films with different GO layers are shown in Figure 10. As can be seen from the figure, with the introduction of more GO layers, the Tg of the EB segment in the copolymer MA-g-SEBS/GO increases from −64 to −52 °C. This is ascribed to the esterification reaction between the modified MA-g-SEBS resin and GO sheets, leading to stronger intermolecular forces of the EB segment. The stronger intermolecular force makes the motion of the EB segment more difficult, giving rise to the increase of Tg of the EB segment. This also illustrates that

3. CONCLUSIONS In this work, we have studied the effect of both physical and chemical interfacial interactions on the static and dynamic mechanical properties (DMA) between the GO and two kinds of SBC resins through the systematic investigations of the morphology observation of the interfaces. After incorporation of GO, mechanical properties of the SEBS/GO hybrid film markedly improved, with the fracture stress reaching 58 MPa and elongation at break increasing to 796%, which was attributed to the interlocks formed between the resin and GO layers. However, it is interesting that the property of the MA-g9125

DOI: 10.1021/acsomega.9b00515 ACS Omega 2019, 4, 9120−9128

ACS Omega

Article

SEBS/GO hybrid film was not improved as expected because of the decrease of interface roughness and the weakening of the physical interaction, even though the esterification reaction could take place between maleic anhydride groups of the MAg-SEBS resin and the hydroxyl groups of GO. Therefore, we draw the conclusion that with regard to SBC/GO LBL assembled composites, the physical interfacial interaction plays a more important role than the chemical interfacial interaction in enhancing the mechanical properties. The synergistic effect of interlocks within the interface is significant for the improvement of composites, especially for those which expect dramatic enhancement from low content addition of fillers such as GO and graphene. Moreover, we believe that the conclusion is not only applicable to LBL assembled hybrid composites but also of considerable reference value for extensive composites incorporated with sheet-shaped or particle-shaped fillers.

1/3/5/7 layers of GO were fabricated with SEBS and MA-gSEBS resins. By tuning the resin layers, the thickness of each film was controlled almost the same, about 100 μm. 4.4. Characterization. The Fourier transformed infrared spectroscopy (FTIR) test of SEBS, MA-g-SEBS, and their nanocomposites were conducted on a Perkin-Elmer 1000 FTIR spectrometer. Raman spectra were measured on a ThermoFisher/DXR. The X-ray diffraction (XRD) measurement was carried out on a Rigaku D/Max 2550 (Cu Kα, λ = 1.5418 Å) with a 2θ scan configuration in the range of 5°−40°. X-ray photoelectron spectroscopy (XPS) experiment was performed on a Shimadzu-Kratos (AXIS Ultra). Typical tapping-mode atomic force microscope (AFM) measurements were taken on a SIINT NanoNavi E-Sweep, with which not only the single-layer GO platelets were observed but also the microscopic surface roughness of the hybrid films was measured. Macroscopic surface roughness of the films was obtained from a Bruker DektakXT Surface Profiler. The sandwich structure of the LBL assembled films was observed using a JEOL JSM-7401F scanning electron microscope at an acceleration voltage of 5 kV, and all the samples were fractured after immersion in liquid nitrogen for a few minutes. Dynamic mechanical analysis (DMA) was carried out using a NETZSCH DMA 242 C/1/G instrument. The tensile tests were conducted on a MTS Criterion Model 43 universal testing machine, with a crosshead speed of 500 mm/min at the temperature of 25 °C.

4. EXPERIMENTAL SECTION 4.1. Materials. Graphite with an average particle size of 800 mesh and a purity of >99% were supplied from Aladdin Industrial Corporation (Shanghai). SEBS (FG1650M) and MA-g-SEBS (FG1901X) with 1.0−1.7 wt % MA were purchased from Kraton Performance Polymer Inc. Concentrated sulfuric acid (H2SO4, 98%), potassium permanganate (KMnO4), hydrochloric acid (HCl, 37%), hydrogen peroxide (H2O2, 30%), tetrahydrofuran (THF) and N,N-dimethylformamide (DMF) were purchased from Sinopharm Chemical Reagent Co., Ltd. 4.2. Preparation of GO. Graphene oxide was prepared with Hummers’ method.39 Natural graphite (3 g) and concentrated H2SO4 (69 mL) were added into a three-neck flask and stirred in an ice bath until the temperature dropped to 0−4 °C, then KMnO4 (18 g) was slowly added into the flask to keep the temperature below 20 °C. After stirring for 2 h, the solution was heated to 35 ± 3 °C and maintained for 45 min. Then, deionized water (145 mL) was added into the flask very slowly with the temperature below 70 °C, followed by heating the solution to 95 ± 3 °C and maintaining for 15 min. The reaction was terminated by adding deionized water (350 mL) and H2O2 (20 mL). The mixture was washed successively with 10% HCl aqueous solution and deionized water until the pH was nearly neutral. At last, GO was obtained after vacuum drying. 4.3. Fabrication of LBL Assembled Hybrid Film. SEBS and MA-g-SEBS solution used for LBL assembly were prepared by dissolving 20 g resin into 100 mL of THF. GO was dispersed into DMF with the concentration of 5 mg/mL, emerging yellow-brown suspension after sonication. First, the resin was coated onto a glass slide using the wet coating technique that has been widely applied in industry and dried in an oven to form a thin layer of resin. Second, the GO suspension was spread over the surface of deionized water drop by drop gradually until the GO platelets covered the entire surface, followed by sonication to form an ultrathin layer of GO, which was adsorbed at air−water interface with highly oriented two-dimensional (2D) structure.40 Then, the GO layer was transferred onto the glass slide with a resin layer by dip-coating. After slight rinse with deionized water, the glass slide was dried in an oven to finish the second step. Alternate repetition of the two steps above was conducted to fabricate hybrid films with the structure of sandwich, the top and bottom of which were resin layers. Hybrid films containing 0/



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b00515. Raman, XRD, XPS, AFM spectra, TEM image characterization of GO; SEM images of hybrid films; IR spectra of the SEBS resin and MA-g-SEBS resin; failure tensile strength of GO hybrid films with different GO oxidation degrees (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (H.X.). *E-mail: [email protected] (Z.Z.). ORCID

Xinling Wang: 0000-0001-7158-5737 Zhen Zheng: 0000-0001-6853-5181 Author Contributions +

K.L. and C.Z. contributed equally, and all authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the Innovation Fund (IFPM2016B005) from the Joint Research Center for Precision Medicine set up by the Shanghai Jiao Tong University and Affiliated Sixth People’s Hospital South Campus (Fengxian Central Hospital). 9126

DOI: 10.1021/acsomega.9b00515 ACS Omega 2019, 4, 9120−9128

ACS Omega



Article

Scientific and Technological Perspective. Adv. Mater. 2016, 28, 6232− 6238. (19) Kano, E.; Kvashnin, D. G.; Sakai, S.; Chernozatonskii, L. A.; Sorokin, P. B.; Hashimotoa, A.; Takeguchia, M. One-Atom-Thick 2D Copper Oxide Clusters on Graphene. Nanoscale 2017, 9, 3980−3985. (20) Liu, P.; Yan, T.; Shi, L.; Park, H.; Chen, X.; Zhao, Z.; Zhang, D. Graphene-based Materials for Capacitive Deionization. J. Mater. Chem. A 2017, 5, 13907−13943. (21) Chai, K.; Sun, Y. Y.; Wu, J. Y. The Effect of High Voltage Pulse Electric Field (HPEF) on Attachment of Diatom on Graphene Composite Coating and Carbon Fiber Composite Coating in Marine Environment. Prog. Org. Coat. 2018, 116, 35−42. (22) Wang, Z.; Wei, R.; Liu, X. Fluffy and Ordered Graphene Multilayer Films with Improved Electromagnetic Interference Shielding over X-Band. ACS Appl. Mater. Interfaces 2017, 9, 22408− 22419. (23) Yao, H.; Hawkins, S. A.; Sue, H. J. Preparation of Epoxy Nanocomposites Containing Well-Dispersed Graphene Nanosheets. Compos. Sci. Technol. 2017, 146, 161−168. (24) Liu, B.; Xia, Y.; Zhou, Y.; Nie, W.; Chen, P. Improved Dispersion of Attapulgite Nanorods in Polymer with Graphene Oxide Nanosheets. J. Mater. Sci. 2017, 52, 1369−1380. (25) Li, Y.; Jing, T.; Xu, G.; Tian, J.; Dong, M.; Shao, Q.; Wang, B.; Wang, Z.; Zheng, Y.; Yang, C. 3-D magnetic graphene oxidemagnetite poly(vinyl alcohol) nanocomposite substrates for immobilizing enzyme. Polymer 2018, 149, 13−22. (26) Wang, C.; Zhao, M.; Li, J.; Yu, J.; Sun, S.; Ge, S.; Guo, X.; Xie, F.; Jiang, B.; Wujcik, E. K.; Huang, Y. Silver nanoparticles/graphene oxide decorated carbon fiber synergistic reinforcement in epoxy-based composites. Polymer 2017, 131, 263−271. (27) Liu, Y.; Feng, J. An Attempt towards Fabricating Reduced Graphene Oxide Composites with Traditional Polymer Processing Techniques by Adding Chemical Reduction Agents. Compos. Sci. Technol. 2017, 140, 16−22. (28) Song, H.; Wang, B.; Zhou, Q.; Xiao, J.; Jia, X. Preparation and Tribological Properties of MoS2/Graphene Oxide Composites. Appl. Surf. Sci. 2017, 419, 24−34. (29) Yang, Q. D.; Li, J.; Cheng, Y.; Li, H. W.; Guan, Z.; Yu, B.; Tsang, S. W. Graphene Oxide as an Efficient Hole-Transporting Material for High-Performance Perovskite Solar Cells with Enhanced Stability. J. Mater. Chem. A 2017, 5, 9852−9858. (30) Gong, X.; Liu, Y.; Wang, Y.; Xie, Z.; Dong, Q.; Dong, M.; Liu, H.; Shao, Q.; Lu, N.; Murugadoss, V.; Ding, T. Amino graphene oxide/dopamine modified aramid fibers: Preparation, epoxy nanocomposites and property analysis. Polymer 2019, 168, 131−137. (31) Wang, X.; Liu, X.; Yuan, H.; Liu, H.; Liu, C.; Li, T.; Yan, C.; Yan, X.; Shen, C.; Guo, Z. Non-covalently functionalized graphene strengthened poly(vinyl alcohol). Mater. Des. 2018, 139, 372−379. (32) Kulkarni, D. D.; Choi, I.; Singamaneni, S. S.; Tsukruk, V. V. Graphene Oxide/Polyelectrolyte Nanomembranes. J. Mater. Chem. A 2010, 4, 4667−4676. (33) Ma, J.; Pan, J.; Yue, J.; Xu, Y.; Bao, J. High Performance of Poly(Dopamine)-Functionalized Graphene Oxide/Poly(Vinyl Alcohol) Nanocomposites. Appl. Surf. Sci. 2018, 427, 428−436. (34) Rabyk, M.; Destephen, A.; Lapp, A.; King, S.; Noirez, L.; Billon, L.; Hruby, M.; Borisov, O.; Stepanek, P.; Deniau, E. Interplay of Thermosensitivity and pH Sensitivity of Amphiphilic Block-Gradient Copolymers of Dimethylaminoethyl Acrylate and Styrene. Macromolecules 2018, 51, 5219−5233. (35) Cao, Y.; Lai, Z.; Feng, J.; Wu, P. Graphene Oxide Sheets Covalently Functionalized with Block Copolymers via Click Chemistry as Reinforcing Fillers. J. Mater. Chem. 2011, 21, 9271− 9278. (36) Grigorescu, R. M.; Ciuprina, F.; Ghioca, P.; Ghiurea, M.; Iancu, L.; Spurcaciu, B.; Panaitescu, D. M. Mechanical and Dielectric Properties of SEBS Modified by Graphite Inclusion and Composite Interface. J. Phys. Chem. Solids 2016, 89, 97−106. (37) de Campos Intema, R.; Wrobel, E. C.; Fujiwara, S. T.; Garcia, J. R.; Pessôa, C. A.; Wohnrath, K. The Effect of Surfactants in the

REFERENCES

(1) Ghaedizadeh, A.; Shen, J.; Ren, X.; Xie, Y. M. Designing Composites with Negative Linear Compressibility. Mater. Des. 2017, 131, 343−357. (2) Chu, K.; Wang, F.; Wang, X. H.; Li, Y. B.; Geng, Z. R.; Huang, D. J.; Zhang, H. Interface Design of Graphene/Copper Composites by Matrix Alloying with Titanium. Mater. Des. 2018, 144, 290−303. (3) Kuttner, C.; Tebbe, M.; Schlaad, H.; Burgert, I.; Fery, A. Photochemical Synthesis of Polymeric Fiber Coatings and Their Embedding in Matrix Material: Morphology and Nanomechanical Properties at the Fiber-Matrix Interface. ACS Appl. Mater. Interfaces 2012, 4, 3484−3492. (4) Habib, E.; Wang, R.; Wang, Y.; Zhu, M.; Zhu, X. X. Inorganic Fillers for Dental Resin Composites: Present and Future. ACS Biomater. Sci. Eng. 2015, 2, 1−11. (5) Wu, Q.; Zhao, R.; Ma, Q.; Zhu, J. Effects of Degree of Chemical Interaction between Carbon Fibers and Surface Sizing on Interfacial Properties of Epoxy Composites. Compos. Sci. Technol. 2018, 163, 34−40. (6) Qian, M.; Sun, Y.; Xu, X.; Liu, L.; Song, P.; Yu, Y.; Wang, H.; Qian, J. 2D-Alumina Platelets Enhance Mechanical and Abrasion Properties of PA612 via Interfacial Hydrogen-Bond Interactions. Chem. Eng. J. 2017, 308, 760−771. (7) Zhao, X.; Zhang, Q.; Hao, Y.; Li, Y.; Fang, Y.; Chen, D. Alternate Multilayer Films of Poly(vinyl alcohol) and Exfoliated Graphene Oxide Fabricated via a Facial Layer-by-Layer Assembly. Macromolecules 2010, 43, 9411−9416. (8) Hou, C.; Tai, Z.; Zhao, L.; Zhai, Y.; Hou, Y.; Fan, Y.; Dang, F.; Wang, J.; Liu, H. High performance MnO@C microcages with a hierarchical structure and tunable carbon shell for efficient and durable lithium storage. J. Mater. Chem. A 2018, 6, 9723−9736. (9) Papageorgiou, D. G.; Kinloch, I. A.; Young, R. J. Mechanical Properties of Graphene and Graphene-based Nanocomposites. Prog. Mater. Sci. 2017, 90, 75−127. (10) Ma, Y.; Bai, D.; Hu, X.; Ren, N.; Gao, W.; Chen, S.; Chen, H.; Lu, Y.; Li, J.; Bai, Y. Robust and Antibacterial Polymer/Mechanically Exfoliated Graphene Nanocomposite Fibers for Biomedical Applications. ACS Appl. Mater. Interfaces 2018, 10, 3002−3010. (11) Zhao, Z.; Bai, P.; Li, L.; Li, J.; Wu, L.; Huo, P.; Tan, L. The Reaction Thermodynamics during Plating Al on Graphene Process. Materials 2019, 12, 330−339. (12) Kirubasankar, B.; Murugadoss, V.; Lin, J.; Ding, T.; Dong, M.; Liu, H.; Zhang, J.; Li, T.; Wang, N.; Guo, Z.; Angaiah, S. In situ grown nickel selenide on graphene nanohybrid electrodes for high energy density asymmetric supercapacitors. Nanoscale 2018, 10, 20414− 20425. (13) Chen, Y.; Jamshidi, R.; Hong, W.; Hashemi, N. N.; Montazami, R. Study of Interfacial Interactions in Physically Transient Soft Layered Structures: A Step toward Understanding Interfacial Bonding and Failure in Soft Degradable Structures. Adv. Eng. Mater. 2017, 19, 1700139. (14) Gerhard, E. M.; Wang, W.; Li, C.; Guo, J.; Ozbolat, I. T.; Rahn, K. M.; Armstrong, A. D.; Xia, J.; Qian, G.; Yang, J. Design Strategies and Applications of Nacre-based Biomaterials. Acta Biomater. 2017, 54, 21−34. (15) Zhang, J.; Li, P.; Zhang, Z.; Wang, X.; Tang, J.; Liu, H.; Shao, Q.; Ding, T.; Umar, A.; Guo, Z. Solvent-free graphene liquids: Promising candidates for lubricants without the base oil. J. Colloid Interface Sci. 2019, 542, 159−167. (16) Wang, Z.; Wei, R.; Gu, J.; Liu, H.; Liu, C.; Luo, C.; Kong, J.; Shao, Q.; Wang, N.; Guo, Z.; Liu, X. Ultralight, highly compressible and fire-retardant graphene aerogel with self-adjustable electromagnetic wave absorption. Carbon 2018, 139, 1126−1135. (17) Zhang, Y.; Qian, L.; Zhao, W.; Li, X.; Huang, X.; Mai, X.; Wang, Z.; Shao, Q.; Yan, X.; Guo, Z. Highly Efficient Fe-N-C Nanoparticles Modified Porous Graphene Composites for Oxygen Reduction Reaction. J. Electrochem. Soc. 2018, 165, H510−516. (18) Palermo, V.; Kinloch, I. A.; Ligi, S.; Pugno, N. M. Nanoscale Mechanics of Graphene and Graphene Oxide in Composites: A 9127

DOI: 10.1021/acsomega.9b00515 ACS Omega 2019, 4, 9120−9128

ACS Omega

Article

Silsesquioxane Solution for LbL Films Assembly. J. Mater. Sci. 2017, 52, 7647−7663. (38) Kang, S.; Wang, L.; Zhang, J.; Du, J.; Li, M.; Chen, Q. Electroreductive Coupling Layer-by-Layer Assembly. ACS Appl. Mater. Interfaces 2017, 9, 32179−32183. (39) Hummers, W. S., Jr.; Offeman, R. E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339−1339. (40) Wang, L.; Zhao, J.; Sun, Y. Y.; Zhang, S. B. Characteristics of Raman Spectra for Graphene Oxide from ab initio simulations. J. Chem. Phys. 2011, 135, 184503.

9128

DOI: 10.1021/acsomega.9b00515 ACS Omega 2019, 4, 9120−9128