Zirong Luo*a,
Shenbo Huangb,
Na Kong
c,
Jizhen Zhang
d,
Jinlong Tao*a,
Jihua Lie and
Shuang Lia
aHainan Provincial Key Laboratory of Natural Rubber Processing, Agricultural Products Processing Research Institute, Chinese Academy of Tropical Agricultural Sciences, Zhanjiang 524001, P. R. China. E-mail: luozirong3@163.com; jinlongt1983@163.com
bGuangdong Engineering & Technology Research Centre of Graphene-like Materials and Products, Department of Chemistry, College of Chemistry and Materials Science, Jinan University, Guangzhou 510632, P. R. China
cSchool of Life and Environmental Science, Centre for Sustainable Bioproducts, Deakin University, Geelong, Victoria 3216, Australia
dInstitute for Frontier Materials, Deakin University, Geelong, Victoria 3216, Australia
eChinese Academy of Tropical Agricultural Sciences, Haikou 571101, P. R. China
First published on 3rd January 2025
Addressing the environmental challenges posed by oil spills and industrial wastewater is critical for sustainable development. Graphene aerogels demonstrate significant potential as highly efficient adsorbents due to their high specific surface area, excellent structural tunability and outstanding chemical stability. Among available fabrication methods, the hydrothermal self-assembly technique stands out for its low cost, high tunability and good scalability. However, brittleness caused by stacking and agglomeration of graphene layers during self-assembly remains a significant challenge. In this study, we present a green and efficient self-assembly strategy combining a one-step hydrothermal process with a solution immersion method to fabricate a PDMS-coated epoxidized natural rubber–graphene composite aerogel (P@EGA). The resulting aerogel exhibits a high specific surface area (482.362 m2 g−1), hierarchical pore distribution from microporous to macroporous, ultra-low density (0.0104 g cm−3) and excellent hydrophobicity (contact angle = 147.6°). Remarkably, it retains 97.54% of its compressive stress after 50 compression-release cycles at 80% strain and quickly recovers its shape under a 500 g load. The P@EGA aerogel demonstrates outstanding adsorption capacities (65.37–132.75 g g−1) for various oils and organic solvents, complete oil absorption in 0.4 seconds, and effortless regeneration through simple squeezing. Furthermore, its dual functionality in gravity-driven and powered water–oil separation systems underscores its broad application potential in environmental remediation.
To mitigate this brittleness, flexible skeletal materials such as nanomaterials,15,16 polymer matrices,17–19 and biomass20 can be incorporated between graphene sheets. These materials inhibit sheet stacking and enhance the mechanical properties of the resulting composites. Polymers, in particular, show promise for fabricating elastic graphene composite aerogels. For instance, Huang et al. introduced a polyimide layer into graphene aerogels using a vacuum infiltration curing method, achieving a robust 3D network structure with a compressive strength of 175 kPa.21 Polydimethylsiloxane (PDMS) is a biocompatible and highly flexible polymer commonly utilized as a hydrophobic or elastomeric material.22 Wu et al. developed PDMS-coated 3D MXene aerogels that maintained a high electromagnetic interference (EMI) shielding efficiency of 48.2 dB after 500 compression-release cycles, demonstrating exceptional compressibility and durability.23 While the application of polymers on the surface of graphene aerogels can improve flexibility, achieving full penetration into the internal 3D network remains challenging, limiting the extent of toughness enhancement. Furthermore, polymers can also act as flexible supporting skeletons, blending with graphene to construct stable 3D porous networks hydrogen bonding and other interactions, addressing agglomeration, limited adsorption sites, and poor mechanical properties. For example, Liu et al. synthesized graphene/polypyrrole composite aerogels using a one-step hydrothermal method.24 The addition of polypyrrole nanorods (PNRs) not only prevented the aggregation of graphene sheets and enhanced mechanical strength but also effectively modified the dielectric constant of the aerogels. Similarly, Lu et al. developed graphene composite aerogels with high compression resistance and durability by introducing polyacrylic acid (PAA) during freeze casting.25 Notably, at approximately 30 wt% PAA, the aerogel's strength increased by 200–300%. Natural rubber is a natural polymer material characterized by excellent elasticity, abrasion resistance, and mechanical strength. Epoxidized natural rubber is produced by introducing epoxy groups via an epoxidation reaction, which not only preserves the high elasticity and strength of natural rubber but also enhances its polarity. These polar oxygenated groups can form stronger chemical or hydrogen bonds with the surface groups of graphene aerogel.26,27 This improved interfacial bonding facilitates better dispersion of graphene, thereby enhancing the mechanical properties of the composites.
In this study, we designed a comprehensive strategy using epoxidized natural rubber (ENR) particles as a flexible support skeleton, combined with graphene oxide through hydrothermal reduction self-assembly. This approach effectively weakened the π–π conjugation between graphene sheets, thereby enhancing the strength and flexibility of formed aerogels. The resulting Epoxidized Natural Rubber/Graphene Aerogels (EGA) featured a stable 3D porous network structure. To further reinforce the EGA polydimethylsiloxane (PDMS) was introduced as a secondary phase, providing additional strength and imparting hydrophobic properties. These enhancements enabled effective and continuous adsorption and separation of various oils and organic solvents.
Organic solvents were purchased from Guangzhou Chemical Reagent Factory. Pump oil (density is 0.85 g cm−3, viscosity is 90.15 mm2 s−1) was obtained from Mitsubishi Chemical Corporation, Japan. Cooking oil (density is 0.93 g cm−3, viscosity is 10.25 mm2 s−1) was purchased from Wal-Mart. Diesel oil (density is 0.83 g cm−3, viscosity is 4.36 mm2 s−1) was supplied by Zhongke Refining & Petrochemical Co.
Q = (ms − m0)/m0 | (1) |
The recycling performance of the aerogels was evaluated using squeezing and extraction method. After reaching adsorption equilibrium, the aerogels were subjected to external force to squeeze out the adsorbed liquid or treated with n-hexane to extract all adsorbates. After this treatment, the aerogels were weighed again. This process was repeated over a total of 10 cycles. Each experiment was independently replicated three times, and the average results were reported to ensure reliability.
![]() | ||
Fig. 1 (a) Schematic diagram of preparation process of P@EGA, SEM images of (b) ENR particles, (c) GA, (d–f) EGA, (g) P@EGA. |
Subsequently, a polydimethylsiloxane (PDMS) layer was successfully coated between the sheets using the solution immersion method, resulting in a lightweight P@EGA with excellent hydrophobic properties. To examine the microstructure (Fig. 1b), ENR particles were fixed with osmium acid staining, revealing particle diameters ranging from 0.2 and 1.2 μm. As shown in Fig. 1c, the 3D structure of the pure GA consists of graphene sheets with significant stacking between layers. In contrast, Fig. 1d–f clearly illustrate that the graphene sheets in the EGA are covered with bumps corresponding to the ENR particles, confirming their successful embedding. This arrangement generates layered porous folds and numerous channels. A comparative analysis of aerogels with varying ENR additions revealed that EGA40% exhibits a more pronounced opening geometry and a superior pore structure compared to EGA20%. However, with increased ENR proportions, the intermolecular forces among ENR particles exceed the hydrogen bonding forces between graphene sheets, leading to a more disordered structure. At an 80% ENR-to-GO ratio, stacking and fusion of ENR particles were observed. Graphene sheets were sporadically stacked on the particle surfaces, leading to a significant reduction in the channels within the aerogel (ESI, Fig. S1†).
As shown in Fig. 1g and ESI S2,† the graphene sheets within the P@EGA remained well-dispersed, with no adhesion observed between the sheets. This demonstrates that the PDMS coating adhered solely to the surface of the walls and did not influence its 3D structure. Energy dispersive spectroscopy (EDS) elemental mapping (Fig. 2a and b) confirms that the predominant elements on both the inner and outer surfaces of the P@EGA are carbon (C), oxygen (O), and silicon (Si), validating the successful PDMS coating on the composite aerogel. As illustrated in Fig. 2c and d, the distribution ratio of oxygen and silicon on the outer surface is slightly higher than on the inner surface. This indicates that the PDMS has penetrated the 3D network structure, although the penetration is incomplete and uneven. As shown in Fig. 2e and f, the adsorption curves of GA and P@EGA exhibited hysteresis loops characteristic of capillary condensation, aligning with type IV isotherms according to IUPAC classification.29 Notably, the BET specific surface area of P@EGA (482.362 m2 g−1) was significantly higher than that of GA (97.931 m2 g−1). Additionally, the adsorption curve of P@EGA showed a rapid increase in adsorption capacity at low p/p0 region, attributed to micropore volume filling, indicating the presence of both micropores and mesopores. Compared to GA, the composite aerogels exhibited a broad pore size distribution, ranging from several nanometers to hundreds of nanometers, retaining some larger pores. This hierarchical pore structure, enriched by the incorporation of ENR particles, formed a stable 3D network. The micropores present in the P@EGA contributed to a higher specific surface area and more active sites, while mesopores facilitated the rapid diffusion of adsorbates, enhancing the adsorption performance for oil spills and organic pollutants.
![]() | ||
Fig. 3 (a and b) FTIR spectra of GO, GA, ENR, EGA and P@EGA, (c) XRD patterns of GO, GA, ENR and EGA, (d) Raman spectra of GO, GA and EGA. |
The interaction between ENR and GO was further analyzed through X-ray diffraction (XRD) and Raman spectroscopy. In the XRD pattern (Fig. 3c), GO exhibits a characteristic diffraction peak at 2θ = 9.38°. Following hydrothermal treatment, reduction of oxygen-containing functional groups causes graphene sheets to restack, resulting in the GA peak shifting to 2θ = 26.56°.34 Due to its amorphous structure, ENR shows a broad diffraction peak at 2θ = 18.9°. EGA display graphene related peaks that shift leftward as ENR content increases, indicating reduced π–π interactions between graphene sheets and increased interlayer spacing. For EGA80%, the peak aligns with the broad diffraction feature of ENR, reflecting significant interaction between ENR and graphene sheets. Raman spectra (Fig. 3d) of GO, GA and EGA highlight two prominent peaks: the D peak at 1349 cm−1, representing disordered vibrations associated with defects in graphene, and the G peak at 1590 cm−1, corresponding to the in-plane C–C stretching of sp2 carbon atoms. The D to G intensity ratio of (ID/IG) reflects material disorder and defect density. After hydrothermal reduction, GA exhibits an increased ID/IG ratio (0.96 to 1.14), suggesting the formation of additional sp2 domains.35 In EGA, ID/IG ratios decreases progressively from 1.06 (EGA20%) to 1.02 (EGA40%) and 0.98 (EGA60%) as ENR content increases, likely due to hydrogen bond interactions between GO and ENR, improving structural order. However, the ID/IG ratio rises to 1.08 for EGA80%, indicating a transition toward disorder, consistent with SEM observations of excessive ENR stacking and fusion. This behavior aligns with the XRD findings, supporting the hypothesis of increased disorder at high ENR content.
The changes in atomic composition and chemical structure of composite aerogels were further analyzed using X-ray photoelectron spectroscopy (XPS). The wide scan XPS spectrum in Fig. 4a revealed O 1s and C 1s peaks at 534.1 eV and 285.1 eV, respectively, for all materials. The introduction of the PDMS layer in the P@EGA was evidenced by the appearance of a new Si peak at 103.1 eV. The C 1s spectra of GO, GA and P@EGA presented in Fig. 4b–d demonstrated significant differences. The spectrum of GO deconvolved into four peaks at 284.5 eV, 285.3 eV, 286.9 eV, and 289.4 eV, corresponding to CC, C–OH, C–O–C, and C
O, respectively, reflecting the abundance oxygen-containing functional groups. In GA, the proportions of C–OH and C–O–C decreased by 16.4% and 12.68%, respectively, due to the hydrothermal reduction that removed epoxy and hydroxyl groups from GO. In the P@EGA the C–O–C content further decreased, accompanied by a 22.64% increase in C
C. This shift indicates ring-opening of epoxy groups in GO and ENR, forming robust hydrogen bonds that enhance interfacial compatibility and structural integrity.31,36
![]() | ||
Fig. 4 (a) XPS spectra of GO, GA, ENR, EGA and P@EGA, high resolution C 1s core level XPS spectra of (b) GO, (c) GA, and (d) P@EGA. |
The P@EGA demonstrates exceptional hydrophobicity and oleophilicity. It retains water droplets on its surface and within its structure, maintaining an ultralight density of 0.0104 g cm−3, even after PDMS coating, allowing it to rest effortlessly on down cotton (Fig. 6a and b). The water contact angle for the EGA is 95.5°, whereas the standard aerogel achieves a remarkable 147.6° (Fig. 6c and d). Diesel fuel takes 4 s to diffuse into the EGA, compared to only 0.4 s for complete adsorption in the standard aerogel (Fig. 6e, f and Movie S2†). The PDMS layer further enhances the aerogel's ability to adsorb diesel fuel entirely, underscoring its hydrophobic and lipophilic properties.
To investigate the effect of ENR introduction on the adsorption capacity of composite aerogels, EGA with varying ENR ratios were employed for the adsorption of pump oil, diesel fuel, tetrachloromethane, and petroleum ether. The results (Fig. 7a) indicated that the adsorption capacity of EGA increased initially with the higher ENR content but declined gradually when the ENR proportion exceeded 60%. This decrease can be attributed to the increased density and aggregation of ENR particles at higher concentrations, which reduces the internal channels of the aerogel. As shown in Fig. 7b, the adsorption capacity of EGA ranged from 64.43 to 130.47 g g−1, significantly surpassing that of GA (21.44–43.59 g g−1). This improvement can be attributed to the introduction of ENR particles between the GO sheets, which significantly extends the specific surface area of the 3D network structure and provides more adsorption sites. The P@EGA exhibited a marginally higher adsorption capacity (65.37–132.75 g g−1) compared to EGA, indicating that hydrophobic modification enhanced water repellency without significantly altering its lipophilicity. Interestingly, the adsorption capacity of P@EGA for low-density petroleum ether was only 65.37 g g−1, whereas it reached 132.75 g g−1 for carbon tetrachloride. This suggests that the adsorption capacity of the samples is roughly proportional to the density of the adsorbate, implying that pore distribution plays a dominant role in influencing adsorption behavior.37
The P@EGA was immersed in three representative adsorbates, and the relationship between adsorption capacity and time was recorded and plotted (Fig. 7c). To further investigate the adsorption behavior of P@EGA, the experimental data were fitted using the pseudo-first-order kinetic model (2) and the pseudo-second-order kinetic model (3):
ln(Qe − Qt) = Qe − k1t | (2) |
t/Qt = 1/k2Qe + t/Qe | (3) |
The analysis of the adsorption process reveals that the P@EGA demonstrates a distinct kinetic behavior across the three tested adsorbates: pump oil, edible oil, and tetrachloromethane. During the initial phase, the adsorption capacity rises rapidly, reflecting the high availability of active sites. This is followed by a gradual approach to saturation, as the aerogel's sites become increasingly occupied. The saturation times—75 s for pump oil, 45 s for edible oil, and 30 s for tetrachloromethane—underscore the influence of adsorbate properties on diffusion rates. Low-viscosity substances like tetrachloromethane diffuse more efficiently into the aerogel's porous network, achieving saturation quickly. High-viscosity substances like pump oil exhibit slower diffusion but adhere strongly to the aerogel's rough surfaces, leading to enhanced adsorption at equilibrium. The kinetic fitting results (Table 1) suggest that the adsorption process is better described by the pseudo-first-order model, with determination coefficient R2 values supporting this conclusion. This indicates a predominance of physical adsorption, driven by surface interactions and diffusion, rather than chemical bonding mechanisms. Overall, these results highlight the aerogel's versatility in adsorbing both high- and low-viscosity organics efficiently. In addition, EGA and P@EGA were immersed in two different adsorbates (pump oil and tetrachloromethane) and their adsorption capacities at various time points were recorded (ESI Fig. S3†). In the adsorption of pump oil, the adsorption rate of EGA was slightly higher than that of P@EGA, but the final adsorption capacity of both materials was comparable. There was no significant difference in the performance of the two materials during the adsorption of carbon tetrachloride. This suggests that although the impregnation of the PDMS layer may affect the rate of adsorption of highly viscous organic compounds by the aerogel, the hindrance caused is not significant.
Adsorbate | Pseudo-first-order kinetic model | Pseudo-second-order kinetic model | ||||
---|---|---|---|---|---|---|
Qe (g g−1) | k1 (s−1) | R2 | Qe (g g−1) | k2 (g g−1 s−1) | R2 | |
Pump oil | 116.5976 | 0.03894 | 0.99272 | 147.5131 | 0.00026 | 0.97583 |
Edible oil | 92.2327 | 0.08772 | 0.99304 | 103.7898 | 0.00118 | 0.95743 |
Tetrachloromethane | 136.9812 | 0.08442 | 0.99063 | 155.0107 | 0.00075 | 0.95666 |
The regeneration ability of P@EGA has been thoroughly evaluated using three organic matter models (pump oil, edible oil, and cyclohexane) and two regeneration methods (extrusion and extraction). As shown in Fig. 7d, the adsorption capacity of the composite adsorbent decreased during the second adsorption cycle for both pump oil and edible oil when regenerated using the extrusion method. This reduction can be attributed to residual adsorbates trapped within aerogel structure after extrusion. Notably, higher viscosity organic matter resulted in greater residue retention. Despite this, the extrusion-regenerated composite adsorbent achieved recovery rates of 93.1% and 95.4% for pump oil and edible oil, respectively, after 10 adsorption cycles. In contrast, the extraction method exhibited superior performance, with almost no loss of adsorption capacity for either pump oil or edible oil across 10 cycles. Furthermore, the recovery rate for cyclohexane exceeded 99% after 10 cycles of adsorption regeneration, regardless of the method employed. This exceptional recovery is primarily due to the volatile nature of cyclohexane, which prevents significant retention within the aerogel structure. Although the extraction method is more effective for desorbing organic matter, it incurs higher costs and requires more time compared to the extrusion method. Based on these findings, we recommend prioritizing the low-cost and high-efficiency extrusion regeneration method. The P@EGA demonstrated excellent reusability through simple extrusion. SEM images of the P@EGA after the tenth adsorption–desorption cycle (ESI Fig. S4†) revealed well-preserved layered pores and ENR particles dispersed within the sheets, indicating the recycled aerogel retained its original structure and shape. As shown in Fig. 7e, the measured contact angle of 138° after 30 cycles confirmed minimal loss of the PDMS layer. This further demonstrated that the dual enhancement effect of ENR and PDMS on the composite aerogels remained effective, ensuring good stability even after multiple regeneration cycles. In addition, HCl and NaOH solutions with different pH values were prepared as test liquids to evaluate the hydrophobic stability of P@EGA (Fig. 7f).38 The results indicate that the water contact angle of P@EGA remains relatively stable across a broad spectrum of pH values, demonstrating high hydrophobicity (CA = 140.5°) even at the lowest pH level (pH = 1). This characteristic renders it suitable for use in diverse aqueous environments. Table 2 provides a comparative analysis of various adsorbents in terms of adsorption capacity, adsorption rate, and recycling performance. Notably, the P@EGA in this study outperformed many other adsorbents, showcasing distinct advantages for water–oil separation applications.
Adsorbents | Density (mg cm−3) | Contact angle | Compressibility | Adsorbates | Adsorption capacity (g g−1) | Regeneration | Reference |
---|---|---|---|---|---|---|---|
Superhydrophobic cotton | — | 156° | — | Oils, organic solvents | 20–50 | Vacuum filtration | 39 |
NiCo@rGO aerogel microsphere | 7.2 | 130° | — | Oils, organic solvents | 107–270 | Solvent extraction | 9 |
PDMS sponge | — | 100°–143° | — | Oils, organic solvents | 1–8 | Squeezing | 40 |
Biomass carbon aerogel | 48 | 135° | — | Oils, organic solvents | 16–50 | Distillation | 41 |
Carbon aerogel | 0.16 | — | The stress retention rate at 50% strain was 88% after 1000 cycles | Organic solvents | 215–913 | Heating | 42 |
Reduced graphene aerogel | Low density | 150.51° | — | Oils, organic solvents | 19–26 | Burning/Distillation | 43 |
Nanocellulose/graphene aerogel | 18 | 130° | The stress retention rate at 90% strain was 98% after 100 cycles | Oils, organic solvents | 25–58 | Squeezing | 44 |
Magnetic MCS/tof aerogel | Ultralight | 153° | The stress retention rate at 70% strain was 89% after 10 cycles | Oils, organic solvents | 37.1–88.4 | Squeezing | 45 |
PDMS/CB@PU sponge | — | 155.4° | No significant change in strength at 50% strain after 500 cycles | Oils, organic solvents | 28.5–68.7 | Squeezing | 46 |
P@EGA | 10.4 | 147.6° | The stress retention rate at 80% strain was 97.5% after 50 cycles | Oils, organic solvents | 65.3–132.7 | Squeezing/Solvent extraction | This work |
Diesel oil and tetrachloromethane, both dyed with Sudan Red III, were utilized as model organics pollutants to assess the water–oil selectivity of the P@EGA. When the aerogel was immersed in a water–oil mixture (Fig. 8a, b and Movie S3†), it effectively and rapidly adsorbed diesel floating on the water surface and tetrachloromethane sinking to the bottom, leaving no observable residue. Upon submersion and subsequent pressing underwater, the aerogel's surface exhibited distinct phosphorescence, a phenomenon caused by an air layer trapped between the PDMS layer and the water. This is a characteristic feature of hydrophobic surfaces. Fig. 8c and Movie S4† illustrate the water–oil separation capability of the P@EGA under gravity-driven conditions. Due to its hydrophobic properties, water could not penetrate the aerogel, while tetrachloromethane readily passed through a glass tube. The upper water level remained constant throughout the experiment. The water–oil separation efficiency of P@EGA was determined to be as high as 98%, calculated as the ratio of the weight of tetrachloromethane collected to the weight of tetrachloromethane initially added to the mixture.47 To further simulate practical scenarios of oil-in-water contamination, an oil-in-water emulsion containing dyed diesel oil was prepared using high-speed stirring. The composite aerogel was employed for dynamic adsorption and separation (Fig. 8d and Movie S5†). After 112 s of stirring, the adsorption process was complete, leaving no visible residual red oil in the container. The dynamic collection of diesel oil floating on the surface of the water and tetrachloromethane sinking underwater was conducted separately using a peristaltic pump with P@EGA as the filter material (Fig. 9 and Movie S6†). The results showed that even when P@EGA was inserted underwater, it only selectively absorbed oil but not water. At the conclusion of the collection process, the water in the beaker on the left remained unchanged, whereas the beaker on the right contained only oil, devoid of any water admixture. This illustrates the exceptional water–oil separation capability of P@EGA.
![]() | ||
Fig. 9 Dynamic continuous adsorption of (a) diesel oil and (b) tetrachloromethane by P@EGA and peristaltic pump. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra06747a |
This journal is © The Royal Society of Chemistry 2025 |