Next Article in Journal
Non-Intrusive Load Disaggregation by Convolutional Neural Network and Multilabel Classification
Previous Article in Journal
Multisensory Plucked Instrument Modeling in Unity3D: From Keytar to Accurate String Prototyping
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic and Hydrophobic Composite Polyurethane Sponge for Oil–Water Separation

1
Key Laboratory for Green Chemical Process of Ministry of Education, School of Chemical Engineering & Pharmacy, Wuhan Institute of Technology, Wuhan 430073, China
2
School of Civil and Environmental Engineering, University of Technology Sydney, Sydney NSW 2007, Australia
*
Authors to whom correspondence should be addressed.
Submission received: 19 January 2020 / Revised: 14 February 2020 / Accepted: 17 February 2020 / Published: 21 February 2020

Abstract

:
Crude oil spills from offshore oil fields will cause serious pollution to the marine ecological environment. Many 3D porous materials have been used for oil–water separation, but they cannot be widely used due to complex preparation processes and expensive preparation costs. Here, a facile and cheap approach to disperse expanded graphite (EG), stearic acid, and Fe3O4 magnetic nanoparticles on the skeleton surface of polyurethane (PU) sponge to prepare the magnetic and hydrophobic composite polyurethane sponge for oil–water separation. The results show that the composite PU sponge had a strong oil absorption capacity for various oils, the oil adsorption capacities has reached 32–40 g/g, and it has become more hydrophobic. The addition of Fe3O4 magnetic nanoparticles endowed the sponge with magnetic responsivity, and the composite PU sponge still had a strong oil adsorption capacity after several adsorbing-squeezing cycles. The magnetic and hydrophobic composite polyurethane sponge is a very promising material for practical oil adsorption and oil–water separation.

1. Introduction

With the rapid development of the petroleum industry, more and more oilfields are exploited, including many offshore oilfields. However, problems such as oil spills can occur in the process of oil extraction, especially oil spills in offshore oil fields can pollute the marine environment [1,2,3]. It is especially important to recover the leaked oil from the surface of seawater. Many materials have been used to adsorb oil from water, but their efficiency and selectivity in oil/water separation are not high enough [4,5,6,7,8,9]. Changing the wetting properties of the material surface changes the selective adsorption of water-oil. A variety of oil-adsorbing and hydrophobic materials for oil–water separation have been prepared and encouraging results were obtained [10,11,12]. In addition, there are many 3D porous materials with hydrophobic and oil adsorption capacity that have been successfully applied to oil–water separation, but their practical applications are still limited by complex preparation methods and high raw material costs [13,14,15,16,17].
Polyurethane (PU) sponge is a commercially obtained low-cost 3D porous material with strong adsorption capacity and reusability. Many studies have reported the preparation of functional polyurethane sponges for oil–water separation, but the preparation cost was high and the amount of oil adsorption was insufficient [18,19,20,21,22,23,24,25]. Yu et al. [26] bonded the reduced graphene oxide (rGO) onto the surface of polyurethane (PU) sponge using (3-aminopropyl)triethoxysilane and titanium(IV (triethanolaminato)isopropoxide and found that pump oil adsorption capacities of sponges are around 20–30 g/g. Pan et al. [27] coated the robust superhydrophobic polysiloxane layer onto the surface of 3D porous polyurethane sponges and the crude oil and bean oil adsorption capacities of sponges are around 24 g/g and 23 g/g, respectively. From the above findings, it can be seen that the preparation process of the above-mentioned hydrophobic oil-adsorbing polyurethane sponge for oil–water separation is complicated and the preparation cost is high, and it is extremely inconvenient to recover the sponge from the surface of seawater.
Therefore, it is necessary to prepare a PU sponge with low cost, simple preparation process, and magnetic properties, which can be applied to oil–water separation. Expanded graphite (EG) is a porous worm-shaped material formed by the expansion of natural graphite flakes at high temperatures, and it is extremely inexpensive and has a strong adsorption capacity [28,29,30,31,32]. Stearic acid is a water-insoluble fatty acid, which has good hydrophobic properties and is inexpensive [33,34,35,36]. Fe3O4 magnetic nanoparticles are a kind of spinel ferrite with stable magnetic permeability and superparamagnetism, and the preparation process of Fe3O4 magnetic nanoparticles is simple and inexpensive [37,38,39,40]. In this work, we attempted to disperse EG and Fe3O4 magnetic nanoparticles into the polyurethane sponge, and apply a stearic acid coating on the skeleton surface of the polyurethane sponge to form a hydrophobic and magnetic sponge for oil–water separation. The EG attached to the skeleton surface of the polyurethane sponge can enhance the oil absorption ability, and the stearic acid coating on the skeleton surface can give the sponge better hydrophobicity. By adding Fe3O4 magnetic nanoparticles to the PU sponge as a filler, a PU sponge with magnetism can be obtained, thereby facilitating the use of a magnet to retrieve the PU sponge from the surface of seawater. The magnetism, oil absorption capacity, hydrophobicity, and reusability of the composite PU sponge were tested.

2. Experimental

2.1. Materials

Polyurethane (C10H8N2O2·C6H14O3)n sponge (density: 0.03 g/cm−3) was obtained from Yonjia Sponge Products Co., Ltd, Ganzhou, Jiangxi, China. Stearic acid (CH3(CH2)16COOH, AR) was provided from TianJin Fuchen Chemical Reagents Factory, Tianjin, China. Expanded graphite was purchased from Kaiyu graphite powder Co., Ltd, Lianyungan, Jiangsu, China. Ferrous chloride (FeCl2, AR) was obtained from Beijing Hawk Science and Technology Co. Ltd, Jinan, Shandong, China. Ferric chloride (FeCl3, AR) was purchased from Shanghai Gongji Chemistry Co. Ltd, Shanghai, China. Absolute ethanol (AR) was provided from Zhengzhou Paini Technology Co., Ltd, Zhengzhou, Henan, China. Peanut oil was obtained from Shanghai Zhongxi Food Sales Co., Ltd, Shanghai, China. Lubricating oil was purchased from Luxianzi Co., Ltd, Shenzhen, Guangdong, China. Silicone oil was provided from Dongguan Tianyu Chemical Industry Co., Ltd, Dongguan, Guangdong, China. Ammonium hydroxide (NH3·H2O, AR) was obtained from INFO and Guangzhou Lixi Chemical Industry Co. Ltd, Guangzhou, Guangdong, China.

2.2. Preparation of Fe3O4 Nanoparticles

Magnetic ferroferric oxide (Fe3O4) nanoparticles were prepared by chemical co-precipitation method [41,42]. The reaction principle is: Fe2+ + 2Fe3+ + 8OH → Fe3O4 + 4H2O. Fifty milliliters of distilled water was added to the three-necked round bottom flask, and then FeCl2 and FeCl3 were quickly weighed in a mole ratio of 1:2 and added to the three-necked round bottom flask. A measuring cylinder was used to measure 50 mL ammonia water (25%~28%) and added to a three-necked round bottom flask. After vacuuming, nitrogen (N2) was introduced as a protective gas. The three-necked round bottom flask was placed on a magnetic stirrer at a temperature of 90 C and stirred at 800 r/min to 900 r/min for 24 h. After 24 h, the reaction product was magnetically separated with a strong magnet and washed three times with absolute ethanol. And the reaction product was dried in a vacuum oven for 12 h to finally obtain magnetic Fe3O4 nanoparticles.

2.3. Preparation of Composite Sponge Samples

Firstly, the commercial PU sponge was cut into many pieces of square shape (2 cm in width and length, 1 cm in height) and cleaned with absolute ethanol three times, and then square sponges were put into a drying oven to dry for 2 h at the temperature of 80 °C. Next, 0.05 g of magnetic Fe3O4 nanoparticles and 0.01 g expanded graphite were added into two beakers with 20 mL absolute ethanol and dispersed in ultrasonic cleaner for 1 h. Then the square sponges were put into two beakers and dispersed in ultrasonic cleaner for 1 h. The square sponges were taken from the beakers and put into a drying oven for 2 h at the temperature of 80 °C. Twenty milliliters of absolute ethanol was added to each of the two beakers, and then 1 g of stearic acid was dissolved in one of the beakers and no stearic acid was added to the other beaker. The sponges obtained from the above steps were immersed into two beakers for 2 h, respectively. Finally, sponges were taken from the two beakers and put into a drying oven for 2 h at a temperature of 80 °C. The composite sponge sample 2 modified by stearic acid and the composite sponge sample 1 not modified by stearic acid were obtained. The schematic preparation of the magnetic composite sponge is illustrated in Figure 1. The specific masses of magnetic Fe3O4 nanoparticles, expanded graphite, and stearic acid for different sponge samples are shown in the Table 1.

2.4. Characterization

The morphology of magnetic Fe3O4 nanoparticles was characterized by transmission electron microscope (TEM; F-30, FEI-Tecnai, 300 kV, Hillsboro, USA). Fe3O4 nanoparticle samples for TEM observations were prepared by dropping the ethanol solution of Fe3O4 nanoparticles samples on the copper grids and drying. The morphology of composite sponge samples with a thin layer of gold prior was evaluated by scanning electron microscope (GeminiSEM 300, Germanic Carl Zeiss Co., Ltd, Oberkochen, Germany.). The FT-IR data of magnetic Fe3O4 nanoparticles, expanded graphite, stearic acid and sponge samples were acquired from fourier transform Infrared Spectrometer within the wave number from 400 to 4000 cm−1 (NICOLET6700, American Thermo Fisher Co., Ltd, Waltham, USA.). The water contact angles of composite sponge samples were measured with purified water by contact angle measurement at room temperature (Theta Flex, Gothenburg, Sweden.). The weight of samples was obtained from an electronic analytical balance (HC2004, HOCHOICE Co., Ltd, Shanghai, China.).

2.5. Method for the Oil Adsorption Tests

All composite sponge samples were immersed in peanut oil at room temperature. The composite sponge samples were weighed firstly (M1), and then immersed into peanut oil until the samples adsorbed peanut oil to saturation. When the sponge was completely immersed in peanut oil and sunk to the bottom of the beaker, it was saturated. They were removed from the peanut oil and weighed (M2). The adsorption capacity Q is calculated according to the following equation:
Q = M 2 M 1 M 1
where M1 is the weight of initial sponge and M2 is the total weight of wet sponge with peanut oil. The adsorption capacities with adsorption cycles were also tested. The pre-weighed sponge samples were then immersed in the peanut oil and allowed for saturated adsorption. Once they were removed from the peanut oil, the saturated sponge samples masses were obtained. The sponge samples were squeezed thoroughly to remove most of the adsorbed peanut oil, and the absorbing–squeezing process was repeated for six cycles successively. In the beaker, the sponge was simply squeezed with a heavy object to perform a desorption test, and the peanut oil obtained was weighed for the adsorption capacity calculation. The squeezing was performed every time in the same way and squeezing tests were repeated on five sponges of the same kind. In our experimental process, to mimic the practical industrial application, the sponge samples were tested for the reusable adsorption capacity of the composite sponge within six adsorbing–squeezing cycles under pressure, successively. All the above processes were also carried out in lubricating oil and silicone oil.

3. Results and Discussion

The morphologies of magnetic Fe3O4 nanoparticles scanned by TEM are exhibited in Figure 2. The diameter of magnetic Fe3O4 nanoparticles is between 10–20 nm, and the particle size reaches the nanoscale. The magnetic Fe3O4 nanoparticles are spherical and dispersed. As shown in Figure 3, the magnetic Fe3O4 nanoparticles are firmly attracted to the glass wall by magnets, which shows that the magnetic Fe3O4 nanoparticles have excellent magnetism. The magnetic Fe3O4 nanoparticles which can be used in magnetic composite sponge were successfully prepared by chemical coprecipitation. Figure 4 compares the magnetic properties of the original sponge and the magnetic composite sponge by testing with magnets. Obviously, unlike the original PU sponge, the magnetic composite sponge can be easily attracted by magnets, indicating that the magnetic Fe3O4 nanoparticles are successfully filled into the original PU sponge.
The chemical composition of the coating on the sponge was identified by FT-IR, demonstrating the existence of magnetic Fe3O4 nanoparticles, expanded graphite and stearic acid in composite sponge. As showed in the Figure 5, the absorption bands at ~3297 cm−1, ~2867 cm−1, ~2275 cm−1, ~1721 cm−1, ~1106 cm−1, and ~1542 cm−1 of line iv in FT-IR spectra are attributed to the characteristic peak of polyurethane sponge [43]. The N−H peak at 3297 cm−1, the peak at 2928 cm−1 is attributed to the C−H stretching of −CH3 and −CH2, and the peak at 2275 cm−1 is due to the asymmetric stretching of −NCO. The peaks at 1721 cm−1 and 1106 cm−1 are associated with C=O stretching in the amide, urea and ether groups respectively, and the peak at 1542 cm−1 is attributed to the amide II band. As for the line v, the FT-IR spectra of magnetic composite sponge, the absorption bands at ~2919 cm−1, ~2849 cm−1, ~1703 cm−1, and ~1466 cm−1 are due to the characteristic peak of stearic acid [44]. The peaks at ~2919 cm−1 and ~2849 cm−1 are due to the anti-symmetry and symmetric stretching vibration of −CH2, the peak at ~1703 cm−1 is due to the stretching vibration of C=O and the peak at ~1466 cm−1 is ascribed to the scissoring vibration of −CH2. The absorption bands at ~565 cm−1 and ~3446 cm−1 are associated with the characteristic peak of magnetic Fe3O4 nanoparticles and expanded graphite, respectively [45,46,47]. The results of FT-IR spectra indicates that the expanded graphite, Fe3O4 magnetic nanoparticles and stearic acid were dispersed successfully into the original PU sponge and there is no new bond created on magnetic composite sponge surface.
The surface morphological evolution of sponge sample 1 and 2 were investigated by SEM. Figure 6(a1–a4) and Figure 6(b1–b4) shows the morphology of sample 1 not modified by stearic acid and sample 2 modified by stearic acid, respectively, which demonstrate sponge samples both have a three dimensional porous structure with pore sizes in the range of 100–500 μm and the surface are rough and the skeleton coated with expanded graphite at the range of 10–30 μm. Obviously, the surface of sample 2 looks brighter than that of sample 1, indicating stearic acid coated at the surface skeleton of sponge sample 2 successfully. Compared with sponge sample 1, the surface of sponge sample 2 is rougher because of the coating of stearic acid.
The oil–water separation experiment was performed using the sponge sample 2 and the process is shown in the Figure 7. The peanut oil was used for the experiment. The sponge was moved above the oil–water mixture and then put on the oil surface carefully. The sponge sample 2 adsorbed peanut oil completely from the oil–water mixture and suspended on the water. Finally, there was only water left in the beaker after removing the sponge. Additionally, the photographs in Figure 7a–f exhibited the process of the sponge sample 2 adsorbing a thick layer of peanut oil on the water surface. Due to its excellent mechanical flexibility, the adsorbed oils can be easily removed from the composite sponge by a simple mechanical extrusion method [48]. Due to the good hydrophobicity of stearic acid on the surface of sponge skeleton, sponges can quickly adsorb oil and block water out.
In order to explore the adsorption capacity of sponge samples, we carried out more oil adsorption experiments. For comparison, the sponge unmodified by stearic acid was also tested. Sponge sample 1 and sample 2 were tested for the oil adsorption with peanut oil, lubricating oil, and silicone oil and the results are shown in Figure 8. Obviously, the two sponge samples could adsorb a wide range of oils, and exhibits high adsorption capacities ranging from around 32–40 times of its own weight as around 32–40 g/g. Moreover, the adsorption capacities of sponge samples in this work are comparable or even better than those of the other previously reported ones, which indicates the expanded graphite coated on the sponge samples skeleton surface increased the oils adsorption capacities. The hydrophobicity of sponge samples were also tested and the results are showed in Figure 8. Compared with sponge sample 1, the water contact angle of sponge sample 2 increases to around 45°, which demonstrates that the hydrophobicity of the sponge enhanced after the modification of stearic acid. A stearic acid hydrophobic layer was successfully formed on the skeleton surface of sponge sample 2. From Table 2, the oil adsorption capacities of different adsorbent materials are lower than that in this work. The adsorption capacity of pure PU sponge for lubricating oil is only 28.8 g/g. Even after the modification to FGN/polyurethane sponge, the adsorption capacity of lubricating oil is only 34.2 g/g, which is still lower than that of sponge sample 1 and sample 2 in our work [49]. The times taken for the composite sponge samples to adsorb different types of oil to reach full saturation are listed in Table 3. It can been seen that the composite sponge samples adsorbed different types of oil to reach full saturation within around 20 s.
The reusability is a crucial index to evaluate the property of composite sponge. In practical industrial application, the sponge will be used for adsorption over several cycles successively, so the sponge will be squeezed and for the next absorbing-squeezing cycle immediately. The recyclability of sponge sample 1 and 2 for oils adsorption for six absorbing-squeezing cycles is presented in Figure 9. As for the sponge sample 1, after an adsorbing-squeezing cycle, the oils adsorption capacities decreased, and with the increase of absorbing-squeezing cycles, the oils adsorption capacities turn to be stable. With the reuse of composite sponges, their adsorbent capacities slightly decreased due to the loss of adsorption capability as a result of the small amount of oil remaining on the sponge and some expanded graphite dropped off, causing the decrease of oils adsorption capacities probably [26]. In the contrast, with the increase of absorbing-squeezing cycles, the oils adsorption capacities of sponge sample 2 decrease slightly and turn to be stable. This indicates that the stearic acid hydrophobic layer formed on the skeleton surface of sponge sample 2 stopped the drop of expanded graphite in the process of adsorbing–squeezing cycles. The oil adsorption capacities of sponge sample 2 maintained stability in the process of adsorbing–squeezing cycles.

4. Conclusions

In summary, a facile approach for preparing magnetic and hydrophobic composite polyurethane sponge is viable. We demonstrated that the magnetic and hydrophobic composite polyurethane sponge has an excellent performance for oil–water separation and great oil adsorption capacities ranging from around 32–40 times of its own weight. The binding of Fe3O4 nanoparticles, expanded graphite, and stearic acid on the skeleton surface of polyurethane sponge not only endows the sponge with magnetic responsivity and hydrophobicity, but also improves the oil adsorption capacity. The recyclability of magnetic and hydrophobic composite polyurethane sponge for oil adsorption demonstrate the sponge could still have a strong oil adsorption capacity for several adsorbing-squeezing cycles. Due to the low prices of expanded graphite and stearic acid, it is feasible to prepare a large number of magnetic and hydrophobic composite polyurethane sponges for oil–water separation in practical industrial applications.

Author Contributions

Conceptualization, P.J. and K.L.; Methodology, P.J.; Formal Analysis, P.J.; Investigation, P.J. and R.D.; Resources, K.L. and X.C.; Data Curation, P.J.; Writing-Original Draft Preparation, P.J.; Writing-Review & Editing, P.J.; Supervision, Y.Y.; Project Administration, P.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Al-Majed, A.A.; Adebayo, A.R.; Hossain, M.E. A sustainable approach to controlling oil spills. J. Environ. Manag. 2012, 113, 213–227. [Google Scholar] [CrossRef] [PubMed]
  2. Wiens, J.A.; Crist, T.O.; Day, R.H.; Murphy, S.M.; Hayward, G.D. A canonical correspondence analysis of the effects of the Exxon Valdez oil spill on marine birds. Ecol. Appl. 2001, 11, 828–839. [Google Scholar] [CrossRef]
  3. Dalton, T.; Jin, D. Extent and frequency of vessel oil spills in US marine protected areas. Mar. Pollut. Bull. 2010, 60, 1939–1945. [Google Scholar] [CrossRef]
  4. Radetic, M.M.; Jocic, D.M.; Jovancic, P.M.; Petrovic, Z.L.; Thomas, H.F. Recycled wool-based nonwoven material as an oil sorbent. Environ. Sci. Technol. 2003, 37, 1008–1012. [Google Scholar] [CrossRef]
  5. Abdelwahab, N.A.; Shukry, N.; El-Kalyoubi, S.F. Preparation and characterization of polymer coated partially esterified sugarcane bagasse for separation of oil from seawater. Environ. Technol. 2017, 38, 1905–1914. [Google Scholar] [CrossRef] [PubMed]
  6. Cardona, D.S.; Debs, K.B.; Lemos, S.G.; Vitale, G.; Nassar, N.N.; Carrilho, E.N.V.M.; Semensatto, D.; Labuto, G. A comparison study of cleanup techniques for oil spill treatment using magnetic nanomaterials. J. Environ. Manag. 2019, 242, 362–371. [Google Scholar] [CrossRef] [PubMed]
  7. Ceylan, D.; Dogu, S.; Karacik, B.; Yakan, S.D.; Okay, O.S.; Okay, O. Evaluation of Butyl Rubber as Sorbent Material for the Removal of Oil and Polycyclic Aromatic Hydrocarbons from Seawater. Environ. Sci. Technol. 2009, 43, 3846–3852. [Google Scholar] [CrossRef]
  8. Cojocaru, C.; Macoveanu, M.; Cretescu, I. Peat-based sorbents for the removal of oil spills from water surface: Application of artificial neural network modeling. Colloids Surf. A Physicochem. Eng. Asp. 2011, 384, 675–684. [Google Scholar] [CrossRef]
  9. Yang, Y.; Tong, Z.; Ngai, T.; Wang, C. Nitrogen-Rich and Fire-Resistant Carbon Aerogels for the Removal of Oil Contaminants from Water. Acs Appl. Mater. Interfaces 2014, 6, 6351–6360. [Google Scholar] [CrossRef]
  10. Gao, H.; Sun, P.; Zhang, Y.; Zeng, X.; Wang, D.; Zhang, Y.; Wang, W.; Wu, J. A two-step hydrophobic fabrication of melamine sponge for oil absorption and oil/water separation. Surf. Coat. Technol. 2018, 339, 147–154. [Google Scholar] [CrossRef]
  11. Zhang, T.; Li, Z.; Lu, Y.; Liu, Y.; Yang, D.; Li, Q.; Qiu, F. Recent progress and future prospects of oil-absorbing materials. Chin. J. Chem. Eng. 2019, 27, 1282–1295. [Google Scholar] [CrossRef]
  12. Zhu, J.; Liu, B.; Li, L.; Zeng, Z.; Zhao, W.; Wang, G.; Guan, X. Simple and Green Fabrication of a Superhydrophobic Surface by One-Step Immersion for Continuous Oil/Water Separation. J. Phys. Chem. A 2016, 120, 5617–5623. [Google Scholar] [CrossRef] [PubMed]
  13. Shi, L.; Chen, K.; Du, R.; Bachmatiuk, A.; Ruemmeli, M.H.; Xie, K.; Huang, Y.; Zhang, Y.; Liu, Z. Scalable Seashell-Based Chemical Vapor Deposition Growth of Three-Dimensional Graphene Foams for Oil-Water Separation. J. Am. Chem. Soc. 2016, 138, 6360–6363. [Google Scholar] [CrossRef] [PubMed]
  14. Liang, W.; Guo, Z. Stable superhydrophobic and superoleophilic soft porous materials for oil/water separation. Rsc Adv. 2013, 3, 16469–16474. [Google Scholar] [CrossRef]
  15. Halake, K.; Bae, S.; Lee, J.; Cho, Y.; Jo, H.; Heo, J.; Park, K.; Kim, H.; Ju, H.; Kim, Y.; et al. Strategies for Fabrication of Hydrophobic Porous Materials Based on Polydimethylsiloxane for Oil-Water Separation. Macromol. Res. 2019, 27, 109–114. [Google Scholar] [CrossRef]
  16. Zhang, J.; Ji, K.; Chen, J.; Ding, Y.; Dai, Z. A three-dimensional porous metal foam with selective-wettability for oil-water separation. J. Mater. Sci. 2015, 50, 5371–5377. [Google Scholar] [CrossRef]
  17. Zhang, W.; Liu, N.; Cao, Y.; Lin, X.; Liu, Y.; Feng, L. Superwetting Porous Materials for Wastewater Treatment: From Immiscible Oil/Water Mixture to Emulsion Separation. Adv. Mater. Interfaces 2017, 4, 1600029. [Google Scholar] [CrossRef]
  18. Zhao, P.; Wang, L.; Ren, R.; Han, L.; Bi, F.; Zhang, Z.; Han, K.; Gu, W. Facile fabrication of asphaltene-derived graphene-polyurethane sponges for efficient and selective oil-water separation. J. Dispers. Sci. Technol. 2018, 39, 977–981. [Google Scholar] [CrossRef]
  19. Chen, X.-Q.; Zhang, B.; Xie, L.; Wang, F. MWCNTs polyurethane sponges with enhanced super-hydrophobicity for selective oil-water separation. Surf. Eng. 2020, 1–9. [Google Scholar] [CrossRef]
  20. Zhang, L.; Xu, L.; Sun, Y.; Yang, N. Robust and Durable Superhydrophobic Polyurethane Sponge for Oil/Water Separation. Ind. Eng. Chem. Res. 2016, 55, 11260–11268. [Google Scholar] [CrossRef]
  21. Cao, N.; Guo, J.; Boukherroub, R.; Shao, Q.; Zang, X.; Li, J.; Lin, X.; Ju, H.; Liu, E.; Zhou, C.; et al. Robust superhydrophobic polyurethane sponge functionalized with perfluorinated graphene oxide for efficient immiscible oil/water mixture, stable emulsion separation and crude oil dehydration. Sci. China Technol. Sci. 2019, 62, 1585–1595. [Google Scholar] [CrossRef]
  22. Li, M.; Yin, X.; Li, J. Robust superhydrophobic/superoleophilic sponge for efficient removal of oils from corrosive aqueous solutions. J. Adhes. Sci. Technol. 2019, 33, 1426–1437. [Google Scholar] [CrossRef]
  23. Liu, C.; Fang, Y.; Miao, X.; Pei, Y.; Yan, Y.; Xiao, W.; Wu, L. Facile fabrication of superhydrophobic polyurethane sponge towards oil water separation with exceptional flame-retardant performance. Sep. Purif. Technol. 2019, 229, 115801. [Google Scholar] [CrossRef]
  24. Yu, T.; Halouane, F.; Mathias, D.; Barras, A.; Wang, Z.; Lv, A.; Lu, S.; Xu, W.; Meziane, D.; Tiercelin, N.; et al. Preparation of magnetic, superhydrophobic/superoleophilic polyurethane sponge: Separation of oil/water mixture and demulsification. Chem. Eng. J. 2020, 384, 123339. [Google Scholar] [CrossRef]
  25. Zhang, J.; Liu, X.; Chen, F.; Liu, J.; Chen, Y.; Zhang, F.; Guan, N. An environmentally friendly and cost-effective method to fabricate superhydrophobic PU sponge for oil/water separation. J. Dispers. Sci. Technol. 2019, 1–9. [Google Scholar] [CrossRef]
  26. Tjandra, R.; Lui, G.; Veilleux, A.; Broughton, J.; Chiu, G.; Yu, A. Introduction of an Enhanced Binding of Reduced Graphene Oxide to Polyurethane Sponge for Oil Absorption. Ind. Eng. Chem. Res. 2015, 54, 3657–3663. [Google Scholar] [CrossRef]
  27. Zhu, Q.; Chu, Y.; Wang, Z.; Chen, N.; Lin, L.; Liu, F.; Pan, Q. Robust superhydrophobic polyurethane sponge as a highly reusable oil-absorption material. J. Mater. Chem. A 2013, 1, 5386–5393. [Google Scholar] [CrossRef]
  28. Chang, Y.; Wang, B.; Luo, H.; Zhi, L. Adsorption of Methylene Blue onto Secondary Expanded Graphite. Adv. Build. Mater. 2011, 168, 2571–2574. [Google Scholar] [CrossRef]
  29. Tan, S.-C.; Shi, P.-H.; Su, R.-J.; Zhu, M.-C. Removal of methylene blue from aqueous solution by powdered expanded graphite: Adsorption isotherms and thermodynamics. Adv. Res. Eng. Mater. Energy Manag. Control 2012, 424, 1313–1317. [Google Scholar] [CrossRef]
  30. Zhao, M.; Liu, P. Adsorption of methylene blue from aqueous solutions by modified expanded graphite powder. Desalination 2009, 249, 331–336. [Google Scholar] [CrossRef]
  31. Vedenyapina, M.D.; Borisova, D.A.; Simakova, A.P.; Proshina, L.P.; Vedenyapin, A.A. Adsorption of diclofenac sodium from aqueous solutions on expanded graphite. Solid Fuel Chem. 2013, 47, 59–63. [Google Scholar] [CrossRef]
  32. Vedenyapina, M.D.; Vedenyapin, A.A. Dynamic adsorption of drug preparations from aqueous solutions on thermally expanded graphite. Solid Fuel Chem. 2015, 49, 41–44. [Google Scholar] [CrossRef]
  33. Emken, E.A. Metabolism of dietary stearic acid relative to other fatty acids in human subjects. Am. J. Clin. Nutr. 1994, 60, 1023S–1028S. [Google Scholar] [CrossRef] [PubMed]
  34. Grundy, S.M. Influence of stearic acid on cholesterol metabolism relative to other long-chain fatty acids. Am. J. Clin. Nutr. 1994, 60, 986S–990S. [Google Scholar] [CrossRef]
  35. Livesey, G. The absorption of stearic acid from triacylglycerols: An inquiry and analysis. Nutr. Res. Rev. 2000, 13, 185–214. [Google Scholar] [CrossRef] [Green Version]
  36. Premphet, K.; Horanont, P. Influence of stearic acid treatment of filler particles on the structure and properties of ternary-phase polypropylene composites. J. Appl. Polym. Sci. 1999, 74, 3445–3454. [Google Scholar] [CrossRef]
  37. Kim, M.I.; Shim, J.; Li, T.; Lee, J.; Park, H.G. Fabrication of Nanoporous Nanocomposites Entrapping Fe3O4 Magnetic Nanoparticles and Oxidases for Colorimetric Biosensing. Chem. Eur. J. 2011, 17, 10700–10707. [Google Scholar] [CrossRef]
  38. Wei, H.; Wang, E. Fe3O4 magnetic nanoparticles as peroxidase mimetics and their applications in H2O2 and glucose detection. Anal. Chem. 2008, 80, 2250–2254. [Google Scholar] [CrossRef]
  39. Martin, M.; Salazar, P.; Villalonga, R.; Campuzano, S.; Manuel Pingarron, J.; Luis Gonzalez-Mora, J. Preparation of core-shell Fe3O4@poly(dopamine) magnetic nanoparticles for biosensor construction. J. Mater. Chem. B 2014, 2, 739–746. [Google Scholar] [CrossRef]
  40. Wang, N.; Zhu, L.; Wang, D.; Wang, M.; Lin, Z.; Tang, H. Sono-assisted preparation of highly-efficient peroxidase-like Fe3O4 magnetic nanoparticles for catalytic removal of organic pollutants with H2O2. Ultrason. Sonochem. 2010, 17, 526–533. [Google Scholar] [CrossRef]
  41. Shen, W.; Sun, A.; Zhai, F.; Wang, J.; Xu, W.; Zhang, Q.; Volinsky, A.A. Fe3O4 magnetic nanoparticles synthesis from tailings by ultrasonic chemical co-precipitation. Mater. Lett. 2011, 65, 1882–1884. [Google Scholar]
  42. Kamyar, S.; Mansor, A.; Khalantari, K.; Khandanlou, R. Synthesis of talc/Fe3O4 magnetic nanocomposites using chemical co-precipitation method. Int. J. Nanomed. 2013, 8, 1817. [Google Scholar]
  43. Hao, J.; Wang, Z.; Xiao, C.; Zhao, J.; Chen, L. In situ reduced graphene oxide-based polyurethane sponge hollow tube for continuous oil removal from water surface. Environ. Sci. Pollut. Res. 2018, 25, 4837–4845. [Google Scholar] [CrossRef] [PubMed]
  44. Huang, H.; Tian, M.; Yang, J.; Li, H.; Liang, W.; Zhang, L.; Li, X. Stearic acid surface modifying Mg(OH)2: Mechanism and its effect on properties of ethylene vinyl acetate/Mg(OH)2 composites. J. Appl. Polym. Sci. 2008, 107, 3325–3331. [Google Scholar] [CrossRef]
  45. Sun, J.; Zhou, S.; Hou, P.; Yang, Y.; Li, M. Synthesis and characterization of biocompatible Fe3O4 nanoparticles. J. Biomed. Mater. Res. Part A 2007, 80, 333–341. [Google Scholar] [CrossRef]
  46. Xu, C.; Wang, H.; Yang, W.; Ma, L.; Lin, A. Expanded Graphite Modified by CTAB-KBr/H3PO4 for Highly Efficient Adsorption of Dyes. J. Polym. Environ. 2018, 26, 1206–1217. [Google Scholar] [CrossRef]
  47. Yang, M.; Zhao, Y.H.; Zhang, Y.C.; Li, D.X. Preparation of Modified Expanded Graphite by KOH. Adv. Mater. Res. 2011, 347, 800–803. [Google Scholar] [CrossRef]
  48. Beshkar, F.; Khojasteh, H.; Salavati-Niasari, M. Recyclable magnetic superhydrophobic straw soot sponge for highly efficient oil/water separation. J. Colloid Interface Sci. 2017, 497, 57–65. [Google Scholar] [CrossRef] [Green Version]
  49. Zhou, S.; Hao, G.; Zhou, X.; Jiang, W.; Wang, T.; Zhang, N.; Yu, L. One-pot Synthesis of Robust Superhydrophobic, Functionalized Graphene/Polyurethane Sponge for Effective Continuous Oil-water Separation. Chem. Eng. J. 2016, 302, 155–162. [Google Scholar] [CrossRef]
  50. Lin, B.; Chen, J.; Li, Z.-T.; He, F.-A.; Li, D.-H. Superhydrophobic modification of polyurethane sponge for the oil-water separation. Surf. Coat. Technol. 2019, 359, 216–226. [Google Scholar] [CrossRef]
  51. Wang, H.; Wang, E.; Liu, Z.; Gao, D.; Yuan, R.; Sun, L.; Zhu, Y. A novel carbon nanotubes reinforced superhydrophobic and superoleophilic polyurethane sponge for selective oil-water separation through a chemical fabrication. J. Mater. Chem. A 2015, 3, 266–273. [Google Scholar] [CrossRef]
Figure 1. Schematic preparation of magnetic and hydrophobic composite polyurethane sponge.
Figure 1. Schematic preparation of magnetic and hydrophobic composite polyurethane sponge.
Applsci 10 01453 g001
Figure 2. TEM images of magnetic Fe3O4 nanoparticles.
Figure 2. TEM images of magnetic Fe3O4 nanoparticles.
Applsci 10 01453 g002
Figure 3. Digital photographs (a) Fe3O4 nanoparticles (b) the magnetic test by magnet for Fe3O4 nanoparticles.
Figure 3. Digital photographs (a) Fe3O4 nanoparticles (b) the magnetic test by magnet for Fe3O4 nanoparticles.
Applsci 10 01453 g003
Figure 4. Digital photographs of magnetic test by magnet for (a) the original polyurethane sponge and (b) the composite sponge.
Figure 4. Digital photographs of magnetic test by magnet for (a) the original polyurethane sponge and (b) the composite sponge.
Applsci 10 01453 g004
Figure 5. FT-IR spectra of (i) expanded graphite, (ii) stearic acid, (iii) Fe3O4 nanoparticles, (iv) original polyurethane sponge, and (v) composite sponge sample 2.
Figure 5. FT-IR spectra of (i) expanded graphite, (ii) stearic acid, (iii) Fe3O4 nanoparticles, (iv) original polyurethane sponge, and (v) composite sponge sample 2.
Applsci 10 01453 g005
Figure 6. Morphology of composite sponge sample 1 (a1–a4) and composite sponge sample 2 (b1–b4).
Figure 6. Morphology of composite sponge sample 1 (a1–a4) and composite sponge sample 2 (b1–b4).
Applsci 10 01453 g006
Figure 7. Digital photographs (af) show the oil-water separation process of composite sponge sample 2.
Figure 7. Digital photographs (af) show the oil-water separation process of composite sponge sample 2.
Applsci 10 01453 g007
Figure 8. The adsorption capacities and water contact angles of composite sponges sample 1 and sample 2.
Figure 8. The adsorption capacities and water contact angles of composite sponges sample 1 and sample 2.
Applsci 10 01453 g008
Figure 9. The adsorption capacities of composite sponge (a) sample 1 and (b) sample 2 for peanut oil, lubricating oil, and silicone oil over six cycles.
Figure 9. The adsorption capacities of composite sponge (a) sample 1 and (b) sample 2 for peanut oil, lubricating oil, and silicone oil over six cycles.
Applsci 10 01453 g009
Table 1. The specific masses of magnetic Fe3O4 nanoparticles, expanded graphite, and stearic acid for different sponge samples.
Table 1. The specific masses of magnetic Fe3O4 nanoparticles, expanded graphite, and stearic acid for different sponge samples.
SampleFe3O4 Nanoparticles (g)Expanded Graphite (g)Stearic Acid (g)
10.050.010
20.050.011
Table 2. Oil adsorption capacity of different adsorbent materials.
Table 2. Oil adsorption capacity of different adsorbent materials.
Adsorbent MaterialsTypes of OilOil Adsorption CapacityReference
Polyurethane spongeLubricating oil28.5 g/g[49]
FGN/polyurethane spongeLubricating oil34.2 g/g[49]
Superhydrophobic spongeLubricating oil24 g/g[27]
F-SiO2/polyurethane spongePeanut oil15 g/g[50]
U-SiO2/polyurethane spongePeanut oil28 g/g[50]
PU–CNT–PDA–ODA spongeLubricating oil26g/g[51]
PU–CNT–PDA–ODA spongeSilicone oil29 g/g[51]
IRGO/polyurethane spongePump oil26.5 g/g[43]
IRGO/polyurethane spongeSoybean oil31.2 g/g[43]
IRGO/polyurethane spongeOlive oil36.1 g/g[43]
Table 3. The times taken for the composite sponge samples to adsorb different types of oil to reach full saturation.
Table 3. The times taken for the composite sponge samples to adsorb different types of oil to reach full saturation.
SampleTypes of OilTime
1Peanut oil20.56 ± 0.26
Lubricating oil21.31 ± 0.75
Silicon oil21.76 ± 0.58
2Peanut oil19.91 ± 0.72
Lubricating oil20.96 ± 0.57
Silicon oil21.38 ± 0.63

Share and Cite

MDPI and ACS Style

Jiang, P.; Li, K.; Chen, X.; Dan, R.; Yu, Y. Magnetic and Hydrophobic Composite Polyurethane Sponge for Oil–Water Separation. Appl. Sci. 2020, 10, 1453. https://0-doi-org.brum.beds.ac.uk/10.3390/app10041453

AMA Style

Jiang P, Li K, Chen X, Dan R, Yu Y. Magnetic and Hydrophobic Composite Polyurethane Sponge for Oil–Water Separation. Applied Sciences. 2020; 10(4):1453. https://0-doi-org.brum.beds.ac.uk/10.3390/app10041453

Chicago/Turabian Style

Jiang, Peng, Kun Li, Xiquan Chen, Ruiqi Dan, and Yang Yu. 2020. "Magnetic and Hydrophobic Composite Polyurethane Sponge for Oil–Water Separation" Applied Sciences 10, no. 4: 1453. https://0-doi-org.brum.beds.ac.uk/10.3390/app10041453

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop