You are on page 1of 6

Materials Science and Engineering B 194 (2015) 62–67

Contents lists available at ScienceDirect

Materials Science and Engineering B


journal homepage: www.elsevier.com/locate/mseb

3D hierarchical porous graphene aerogels for highly improved


adsorption and recycled capacity
Caixia Chi a,b , Hongbo Xu a , Ke Zhang a,∗ , Yibo Wang a , Shuanghu Zhang b , Xusong Liu a ,
Xin Liu a , Jiupeng Zhao a , Yao Li c,∗
a
School of Chemical Engineering and Technology, Harbin Institute of Technology, Xida Zhi street 92, Harbin 150001, China
b
Food and Pharmaceutical Engineering College, Suihua University, Suihua 152061, China
c
Center for Composite Materials, Harbin Institute of Technology, Harbin 150001, China

a r t i c l e i n f o a b s t r a c t

Article history: This article demonstrates the fabrication of 3D hierarchical porous graphene aerogels (HGAs) with tun-
Received 30 September 2014 able porous architecture by a “fishing” process with polystyrene (PS) particles as sacrificial templates.
Received in revised form 2 December 2014 Structure and morphology were characterized by X-ray diffraction (XRD), Raman spectroscopy, and scan-
Accepted 23 December 2014
ning electron microscopy (SEM), respectively. The 3D HGAs exhibits low density, super hydrophobicity,
Available online 8 January 2015
high porosity and good thermal stability, which were used as adsorbents for removal of oils and organic
solvents, showing highly efficient adsorption. In addition, they could be regenerated by heat treatment
Keywords:
to release the adsorbates. After eight toluene-adsorbing and drying recycles, the adsorption capacity of
3D hierarchical porous graphene aerogel
Polystyrene particle
HGAs reaches 93% of the initial uptake.
Self-assembly © 2015 Elsevier B.V. All rights reserved.
Pore size tunable
Oil adsorption
Recycled capacity

1. Introduction Pd, Ir, Rh, or Pt, etc.) and treated hydrothermally at 120 ◦ C for
20 h. The samples had excellent mechanical properties, and had
Graphene, a single layer of graphite, has been attracted much been utilized as fixed-bed catalysts for a Heck reaction resulting in
attention due to its excellent mechanical [1,2], electrical [3,4], ther- both 100% selectivity and conversion. Yan et al. [23] demonstrated
mal [5,6] and optical properties [7,8]. Since the first discovery different kinds of reducing agents (NaHSO3 , Na2 S, HI, etc.) to be
of isolated graphene by mechanical exfoliation of graphite crys- effective in preparing 3D architectures of graphene which exhibited
tals [9], numerous approaches were developed for fabrication of high mechanical, thermal and electrical properties; furthermore, by
graphene, such as micromechanical cleavage [9], crystal epitaxial incorporating Fe3 O4 nanoparticles [22], the obtained composites
growth [10,11], chemical vapor deposition [12–14] and chemical showed a high performance as an anode for lithium-ion batter-
reduction of graphene oxide (GO) [15–23]. Relative to other tech- ies. Yu et al. [25] reported a one-step fabrication of graphene/iron
niques, chemical reduction of GO, is considered to be an efficient oxide hydrogels which is induced by ferrous ions, and revealed their
approach to produce processable graphene on a large scale [15], excellent capabilities for removal of oil and heavy metal ions in
in which it is usually prepared by oxidation of graphite powder water. Spongy graphene was made by reducing a suspension of
with strong oxidants such as a mixture of concentrated sulfuric acid GO at 180 ◦ C for 24 h in Ruoff’s paper [26]. The samples possessed
and potassium permanganate via Hummer’s method. The abundant highly efficient adsorption of fats and petroleum products, but also
oxygen-containing groups on GO endow it with excellent aqueous toxic solvents such as toluene and chloroform. Hu et al. [27] pre-
dispersion and make it easy to modify [17–24]. pared 3D structures of graphene via the chemical reduction of GO
Recently, efforts have been made to fabricate 3D graphene in the presence of a range of environmentally-friendly phenolic
through self-assembly of well-dispersed GO or functionalized acids at 95 ◦ C for 8 h without stirring. These products were used
graphene sheets [4,18–23]. The former was prepared by Wang as electrode materials for supercapacitors, as well as adsorbents
et al. [20], being promoted by a noble-metal nanocrystal (Au, Ag, for removal of oils, organic solvents and dyes from contaminated
water. In addition, an ice-templating method was used to prepare
the macroporous and hydrophobic prGO-PDMS structure through
∗ Corresponding authors. Tel.: +86 451 86403767; fax: +86 451 86403767. vapor deposition of PDMS on GO scaffolds, which exhibits a great
E-mail addresses: zhangke@hit.edu.cn (K. Zhang), yaoli@hit.edu.cn (Y. Li). improvement in both the adsorption capacity and reusability [28].

http://dx.doi.org/10.1016/j.mseb.2014.12.026
0921-5107/© 2015 Elsevier B.V. All rights reserved.
C. Chi et al. / Materials Science and Engineering B 194 (2015) 62–67 63

Moreover, Ahn et al. [29] reported a novel and practical technol- Successively, the as-prepared dispersion was heated at 80 ◦ C for
ogy to form 3D interconnected foam-like graphene layer and base 1.5 h to produce the PS/graphene hydrogels. Then, they were
graphene multi-layer on heater surface to effectively eliminate the immersed in tetrahydrofuran (THF) to remove PS particles to obtain
sudden and rapid increase of wall temperature at critical heat flux the hierarchical porous graphene hydrogels (HGHs). The porosity
(CHF). and pore size distribution of graphene can be controlled by vary-
Many literature reported integration of graphene sheets into ing the size of PS particles. Finally, the product was dialyzed with
3D structures by different strategies and investigated the advanced ultrapure water for three days, and followed by freeze-drying for
properties as aforementioned, nonetheless the 3D interconnected further use.
frameworks of graphene with a tunable hierarchically porous archi-
tecture has not been reported. Enlightened by the studies on 2.4. Characterization
preparation of a 3D graphene hydrogel [22–24], it is possible to
prepare composites of 3D graphene embedded nanoparticles if The morphology of samples was observed using a field emis-
some of the nanoparticles are dispersed homogeneously in a GO sion scanning electron microscope (SU8000). The X-ray diffraction
aqueous suspension during the reduction and self-assembly. In this (XRD) data were obtained on a Rigaku D/max-rb (12 KW) with Cu
paper, it is proposed that the capture of PS particles is similar to a K␣ radiation ( = 0.15418 nm). X-ray photoelectron spectroscopy
“fishing process with “fish net , where the “fish is the PS par- (XPS) data were obtained using a PHI 5700 ESCA system with an Al
ticle and the “fish net is composed of graphene/graphene oxide K 1486.6 eV as the X-ray source. Thermogravimetric analysis (TGA)
nanosheet, leading to a network during reduction via self-assembly. was carried out using a ZRY-2P thermal analysis station with a ramp
Polystyrene spheres are polymeric particles which are easily syn- rate of 5 ◦ C min−1 . Raman spectra were obtained with a Renishaw
thesized by different approaches such as emulsifier-free emulsion Raman system model 1000 spectrometer. The 514.5 nm radiation
polymerization [30] and dispersion polymerization [31]. Normally, from a 20 mW air-cooled argon ion laser was used as an exciting
PS particles as close-packed colloidal crystal templates (CCT) that source. The N2 adsorption–desorption isotherms of the samples
are used for preparation of three-dimensionally ordered macro- were measured at 77 K using specific surface porosity analyzer 3H-
porous (3DOM) nanostructures by chemically removed after the 2000PS1.
deposition of the desired material [30]. Here, the 3D self-assembly
architecture of PS/graphene hydrogels were fabricated owing to the 2.5. Oil-uptake experiments
electrostatic repulsions between the negative charged graphene
oxide and PS particles at pH = 5–7 in water during the reduction. Oil-uptake capacity of the HGAs was determined by measuring
PS particles are employed as sacrificial templates that yield porous the weight. The aerogel samples were put into various kinds of
structure when removed by organic solvent to obtain HGAs. The oils and organic solvents labeled with Sudan III dye, and taken out
microstructure (porosity and pore size distribution) of HGAs can be until they were fully adsorbed. Then the samples were weighed
controlled by varying the size of PS particles, and the macroscopic after removing the surface oils with filter papers. The oil-uptake
structure of HGAs (shape and size) can be dominated by changing performance was reflected by the saturated adsorption capacity
the shape and volume of vessel. Particularly, both the microscopic per unit mass of the aerogel samples, which is defined as follows:
and macroscopic structure of HGAs can be easily carried out in one
(Ws − Wi )
step. Q = (1)
Wi
2. Experimental Here, Q is the adsorption capacity at saturated state. Wi and Ws
are the initial and final weight (at adsorption saturation) of HGAs,
2.1. Preparation of GO respectively.

GO was prepared by an improved Hummer’s method as 3. Results and discussion


described in Ref. [32]. The as-prepared GO was dispersed in water
(2 mg mL−1 ) and sonicated for 30 min in an ultrasonic bath. The The proposed mechanism for fabrication of HGAs is shown in
resulting brown dispersion was subjected to centrifugation for Fig. 1. PS particles are well dispersed in GO solution (Fig. 2(a) – left)
30 min at 5000 rpm to remove un-exfoliated GO. under controlled pH between 5 and 7 due to electrostatic repul-
sion, where zeta potential values are measured to be −42 ± 1.5 mV
2.2. Preparation of PS particles for GO and −50 ± 5.2 mV for PS. At early-stage reduction, partially
reduced GO sheets come up, and graphene/GO nanosheets that
The nano-sized PS particles were synthesized by an emulsifier- can be self-assembled together to form network. Then PS particles
free emulsion polymerization [30]. Firstly, styrene monomer and are spontaneously captured into the network in a simple one-step
initiator K2 S2 O8 were dispersed in aqueous solution under a rapid process. Some sheets of reduced GO assemble together in parallel,
stirring of 400 rpm and it was initiated at 70 ◦ C after removing but others may assemble at any angle during the reduction. At the
oxygen by nitrogen purging. The PS particles with different size end, much water was expelled out from the aggregates to obtain
(256 nm, 356 nm, 410 nm and 598 nm) were obtained by increase of the compact 3D architectures (Fig. 2(a) – right, 3(a)).When PS are
monomer content. On the other hand, the micro-sized PS particles removed, a certain hydrogel with hierarchical porous structure is
with average diameters of 1000 nm, 1200 nm and 1580 nm were formed (Fig. 3(b)). The as-prepared graphene hydrogels still con-
prepared by dispersion polymerization and then they are treated tain plenty of water (above 97 wt%) and the corresponding aerogels
in sulfuric acid to obtain the sulfoated PS particles [31]. are obtained by removing the water via a freeze-drying method
(Figs. 2(b) and 3(c)). Moreover, HGAs can be molded into any shape
2.3. Preparation of HGHs and HGAs (Fig. 2(b)), thus it could be placed directly in polluted areas and
dredged intact afterwards, leading to simpler recovery.
GO aqueous dispersion (∼2 mg mL−1 ) was mixed with PS par- Fig. 4 shows XRD patterns of pristine graphite, GO and HGAs. The
ticle solution (1.44 wt%), and 30 mg ascorbic acid was added XRD pattern of pristine graphite exhibits a basal reflection (0 0 2)
as reducing agent. The homogeneous red-brown dispersion was peak at 2 = 26.6◦ (d-spacing = 0.34 nm) [33]. For GO, the 0 0 2 reflec-
obtained after sonication for 15 min to dissolve the ascorbic acid. tion peak shifts to the lower angle of 8.66◦ which corresponds to
64 C. Chi et al. / Materials Science and Engineering B 194 (2015) 62–67

Fig. 1. Illustration of the process for preparation of HGAs.

Fig. 2. Photographs of (a) an aqueous mixture of GO, PS (dPS = 410 nm) and ascorbic acid before (left) and after (right) chemical reduction at 80 ◦ C (b) the HGAs with different
shape and size.

Fig. 3. SEM images of (a) HGHs with PS (b) HGHs without PS (c) HGAs.

increase of the interlayer spacing (1.02 nm) during the oxidation. Raman spectra of the graphite, GO and HGAs are shown in
This peak completely disappears after GO was reduced to HGAs; Fig. 5. Raman spectrum is highly sensitive to the electronic structure
instead, a new peak appears at 19.4◦ , which attributes to a decrease and has proven to be an essential tool for the characterization of
interlayer spacing of 0.44 nm. In addition, the broad peak of HGAs carbon-based materials (especially for C C bands) [34]. The Raman
is different from the narrow peak of pristine graphite, suggesting spectrum of graphite, displays a prominent G peak as the feature at
that the disordered stacked graphene sheets in the microscopic 1574.7 cm−1 , corresponding to the first-order scattering of the E2g
scale. mode [35]. In the Raman spectrum of GO, the G band is broadened

Fig. 4. XRD pattern of (a) graphite, (b) GO and (c) HGAs. Fig. 5. Raman spectra of (a) graphite, (b) GO and (c) HGAs.
C. Chi et al. / Materials Science and Engineering B 194 (2015) 62–67 65

Fig. 6. XPS spectra of C1s of (a) GO and (b) HGAs.

and shifted to 1585 cm−1 . In addition, the D band at 1353.8 cm−1


becomes prominent, indicating the reduction in size of the in-plane
sp2 domains, possibly due to the extensive oxidation. After reduc-
tion, the G band appears at 1574.7 cm−1 and with an increased
ID /IG intensity ratio of 1.06 compared to that in GO (0.80). This
change suggests a degree of disorder in the carbon arrangement
upon reduction of the exfoliated GO [36].
Fig. 6 shows the C1s XPS spectra of GO and HGAs. For GO
(Fig. 6(a)), four different peaks centered at 284.6 eV, 286.5 eV,
288.2 eV, and 290.0 eV are detected, which correspond to C C/C–C
in aromatic rings, C–O (epoxy and alkoxy), C O, and COOH groups, Fig. 8. Image of water droplet on HGAs surface.
respectively. After the reduction, the intensities of all C1s peaks of
the carbons binding to oxygen decrease dramatically, especially the
starting mass loss temperature (∼350 ◦ C) of the HGAs is roughly
peak of C–O (epoxy and alkoxy). These results reveal that most oxy-
250 ◦ C higher than that of GO. This is attributed to the removal of
gen containing functional groups were removed and the majority
oxygen functionalities from the surface of GO during the reduction
of the conjugated graphene networks have been restored after the
process. On the other hand, the adsorbed liquids can be readily
reduction.
recycled by heating in air without destroying the porous structures
To confirm the reduction effect, TGA was carried out from
due to the fine thermal stability of HGAs.
25 ◦ C to 700 ◦ C under an air atmosphere, as shown in Fig. 7. GO
The HGAs (9.22∼23.05 mg cm−3 ) is of highly hydrophobic (the
is thermally unstable and starts to lose mass (10%) upon heating,
contact angle is 139 ± 0.5◦ , as shown in Fig. 8), which has a
even below 100 ◦ C, which is attributed to the rapid evaporation
maximum surface area of 237.47 m2 g−1 with tunable pore size
of the intercalated water. The main mass loss (80%) takes place
(N2 adsorption–desorption isotherms, as shown in Fig. 9). The
at around 180 ◦ C [15] due to the decomposition of the oxygen
isotherms in Fig. 9 are characteristic of type III isotherms, which
functional groups, yielding CO, CO2 , and steam. The gases produced
suggest the multilayer adsorption on a macroporous solid at the
by thermal paralysis’ of the oxygen-containing functional groups
relatively high pressure region (P/P0 of 0.9–0.95) [38,39]. The dis-
accelerate GO expansion and delimitation [37]. In contrast, the
tribution of pore size computed using the BJH method (Fig. 9 insert),
thermal stability of HGAs is much higher than that of GO. The

Fig. 9. N2 adsorption–desorption isotherms and pore size distribution (insert) of


Fig. 7. TGA plot of (a) GO and (b) HGAs. HGAs (dPS = 410 nm).
66 C. Chi et al. / Materials Science and Engineering B 194 (2015) 62–67

Fig. 10. Adsorption of soya oil in HGAs at intervals of 30 s.

organic solvents, but also to the pore size, specific surface area and
pore volume of adsorbent. Its adsorption efficiency is much higher
than that of previously reported adsorptive materials, e.g., macro-
scopic multifunctional graphene-based aerogels [25]. In addition,
although the adsorption capacity is close to spongy graphene [26],
the preparation process of HGAs is much simpler, in which HGAs
could be obtained without long-time treatment. In comparison
with commercial adsorbents, the generated HGAs exhibit better
performance than that of sisal, coir fiber, sponge-gourd [40] and
polypropylene nonwoven web [41].
As shown in Fig. 11, the adsorption capacities of HGAs for identi-
cal oils and organic solvents decrease after first increasement with
the increase of particle size, i.e. there is the optimum particle size
existed for the adsorption. Microporous zeolites and polymers have
been used to remove organics and oil from water due to their
large specific surface area and hydrophobicity [42,43]. In liquid
chromatography, the adsorption value of organics will generally
increase with the increase of molecular weight unless the molec-
ular is too big to enter the pore. The ideal condition is that the
Fig. 11. Adsorption efficiency of HGAs for different oils and organic solvents. adsorbent has large numbers of pores with the suitable size for
adsorbate to enter. Too small pores will block the entry of adsor-
finding that HGAs possess hierarchical porous structure. The max- bate; too large pores will cause low specific area. Normally, bigger
imum pore size is about 120 nm. The progress of adsorption for BET will bring stronger adsorption affinity, but this theory has its
soybean oil is shown in a series of photos in Fig. 10. To further limitation in practical applications. The surface area, pore size and
reveal the adsorption performance of HGAs, it was tested in various pore volume were analyzed for the HGAs with addition of PS par-
kinds of oils and organic solvents on the HGAs with different density ticles (Fig. 12). The pore size as opposed to the surface area shows
and pore diameters. As shown in Fig. 11, the maximum adsorption an increasing trend with increasing particle size (Fig. 12a). In addi-
capacities are obtained to be from 21 to 77 times as heavy as the tion, regarding the effect of particle size on the pore volume, it is
weight of the aerogel samples for different oils and organic solvents, found that the pore volume reaches the maximum value when the
which may be related to the density of the oils and organic solvents. added PS particles is 410 nm, revealing that accumulative total pore
The adsorption capacities are different on the HGAs with various volume of HGAs does not increase with increasing pore size. From
PS size as template for identical oils and organic solvents, which this, it is confirmed that the pore volume may be the principal factor
may be related not only to the molecular dimension of the oils and determining the oil adsorption in HGAs. Therefore, it can be used

Fig. 12. Surface area, pore size and pore volume of the HAGs samples as function of the added PS partciels.
C. Chi et al. / Materials Science and Engineering B 194 (2015) 62–67 67

Research Innovation Foundation in Harbin Institute of Technol-


ogy (XWQQ5750013113) and Heilongjiang Post-doctoral Science
Foundation (AUGA4110029613).

References

[1] M.A. Rafiee, J. Rafiee, Z. Wang, H. Song, Z.Z. Yu, N. Koratka, ACS Nano 3 (2009)
3884–3890.
[2] R.A. Barton, B. Ilic, A.M. van der Zande, W.S. Whitney, P.L. McEuen, J.M. Parpia,
H.G. Craighead, Nano Lett. 11 (2011) 1232–1236.
[3] Y.B. Zhang, T.T. Tang, C. Girit, Z. Hao, M.C. Martin, A. Zettl, M.F. Crommie, Y.R.
Shen, F. Wang, Nature 459 (2009) 820–823.
[4] X.C. Dong, H. Xu, X.W. Wang, Y.X. Huang, M.B. Chan-Park, H. Zhang, L.H. Wang,
W. Huang, P. Chen, ACS Nano 6 (2012) 3206–3213.
[5] A.A. Balandin, S. Ghosh, W.Z. Bao, I. Calizo, D. Teweldebrhan, F. Miao, C.N. Lau,
Nano Lett. 8 (2008) 902–907.
[6] A.A. Balandin, Nat. Mater. 10 (2011) 569–581.
[7] T. Mueller, F.N.A. Xia, P. Avouris, Nat. Photon. 4 (2010) 297–301.
[8] C.C. Lee, S. Suzuki, W. Xie, T.R. Schibli, Opt. Express 20 (2012) 5264–5269.
[9] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, Y. Zhang, S.V. Dubonos, I.V.
Grigorieva, A.A. Firsov, Science 306 (2004) 666.
[10] A.K. Geim, K.S. Novoselov, Nat. Mater. 6 (2007) 183–191.
Fig. 13. Recyclability of HGAs (dPS = 410 nm) repetitively adsorbed toluene and [11] W.P. Sutter, J. Flege, A.E. Sutter, Nat. Mater. 7 (2008) 406–411.
[12] A. Reina, X.T. Jia, J. Ho, D. Nezich, H.B. Son, V. Bulovic, M.S. Dresselhaus, J. Kong,
released its vapor under heat treatment (105 ◦ C) for 8 cycles.
Nano Lett. 9 (2009) 30–35.
[13] K.S. Kim, Y. Zhao, H. Jang, S.Y. Lee, J.M. Kim, K.S. Kim, J.H. Ahn, P. Kim, J.Y. Choi,
B.H. Hong, Nature 457 (2009) 706–710.
[14] D.C. Wei, B. Wu, Y.L. Guo, G. Yu, Y.Q. Liu, Acc. Chem. Res. 46 (2013) 106–115.
to explain why the adsorption capacity declines after reaching one
[15] S. Stankovich, D.A. Dikin, R.D. Piner, K.A. Kohlhaas, A. Kleinhammes, Y. Jia, Y.
peak (Fig. 11). Wu, S.T. Nguyen, R.S. Ruoff, Carbon 45 (2007) 1558–1565.
Furthermore, it is found that the HGAs had an excellent regener- [16] C.M. Chen, Q.H. Yang, Y.G. Yang, W. Lv, Y.F. Wen, P.X. Hou, M.Z. Wang, H.M.
Cheng, Adv. Mater. 21 (2009) 3007–3011.
ation capacity through releasing the adsorbed organics by heating.
[17] A.V. Talyzin, V.L. Solozhenko, O.O. Kurakevych, T. Szabo, I. Dekany, A. Kurnosov,
As shown in Fig. 13, the HGAs maintained a high adsorption capac- V. Dmitriev, Angew. Chem. Int. Ed. 47 (2008) 8268–8271.
ity (93% of the initial uptake) after eight toluene-adsorbing and [18] Y.X. Xu, H. Bai, G.W. Lu, C. Li, G.Q. Shi, J. Am. Chem. Soc. 130 (2008) 5856–5857.
drying recycles. [19] L.J. Cote, F. Kim, J.X. Huang, J. Am. Chem. Soc. 131 (2009) 1043–1049.
[20] Z.H. Tang, S.L. Shen, J. Zhuang, X. Wang, Angew. Chem. Int. Ed. 49 (2010)
4603–4607.
4. Conclusion [21] K.X. Sheng, Y.X. Xu, C. Li, G.Q. Shi, New Carbon Mater. 26 (2011) 9–15.
[22] W.F. Chen, S.R. Li, C.H. Chen, L.F. Yan, Adv. Mater. 23 (2011) 5679–5683.
[23] W.F. Chen, L.F. Yan, Nanoscale 3 (2011) 3132–3137.
In summary, 3D HGAs with tunable pore sizes has been suc- [24] B.G. Choi, M. Yang, W.H. Hong, J.W. Choi, Y.S. Huh, ACS Nano 6 (2012)
cessfully prepared by a wet chemical self-assembly technique, i.e. 4020–4028.
“fishing” process during the reduction of GO in aqueous suspension, [25] H.P. Cong, X.C. Ren, P. Wang, S.H. Yu, ACS Nano 6 (2012) 2693–2703.
[26] H.C. Bi, X. Xie, K.B. Yin, Y.L. Zhou, S. Wan, L.B. He, F. Xu, F. Banhart, L.T. Sun, R.S.
PS was removed by subsequent treatment with THF. The prepared Ruoff, Adv. Funct. Mater. 22 (2012) 4421–4425.
HGAs exhibit low density, super hydrophobicity, high porosity and [27] J.L. Wang, Z.X. Shi, J.C. Fan, Y. Ge, J. Yin, G.X. Hu, J. Mater. Chem. 22 (2012)
good thermal stability, which could be promising candidates for 22459–22466.
[28] S. Park, S.-O. Kang, E. Jung, S. Park, H.S. Park, RSC Adv. 4 (2014) 899.
adsorption of oils and organic solvents with good adsorption and [29] H.S. Ahn, J.M. Kim, C. Park, J.-W. Jang, J.S. Lee, H. Kim, M. Kaviany, M.H. Kim, Sci.
regeneration capacity. This study suggests a facile, low cost and Rep. 3 (2013) 1960.
scalable route to prepare graphene based aerogels in application of [30] J.P. Zhao, Y. Li, Z. Cao, W.H. Xin, New J. Chem. 32 (2008) 1014–1019.
[31] W. Fan, C. Zhang, W.W. Tjiu, K.P. Pramoda, C. He, T.X. Liu, ACS Appl. Mater.
water purification.
Interfaces 5 (2013) 3382–3391.
[32] D.C. Marcano, D.V. Kosynkin, J.M. Berlin, A. Sinitskii, Z.Z. Sun, A. Slesarev, L.B.
Prime novelty statement Alemany, W. Lu, J.M. Tour, ACS Nano 4 (2010) 4806–4814.
[33] R. Bissessur, P.K.Y. Liu, S.F. Scully, Synth. Metals 156 (2006) 1023–1027.
[34] S. Niyogi, E. Bekyarova, M.E. Itkis, H. Zhang, K. Shepperd, J. Hicks, M. Sprinkle,
This work reports a facile, low cost and scalable route to prepare C. Berger, C.N. Lau, W.A. Deheer, E.H. Conrad, R.C. Haddon, Nano Lett. 10 (2010)
3D hierarchical porous graphene aerogels (HGAs) by using of 4061–4066.
[35] F. Tuinstra, J.L. Koenig, J. Chem. Phys. 53 (1970) 1126–1130.
polystyrene (PS) as sacrificial templates for the first time. It exhibits
[36] T. Muraliganth, A.V. Murugan, A. Manthiram, J. Mater. Chem. 18 (2008)
highly improved adsorption and recycled capacity for different oils 5661–5668.
and organic solvents than that of previously reported. [37] H.C. Schniepp, J.L. Li, M.J. McAllister, H. Sai, M. Herrera-Alonso, D.H. Adamson,
R.K. Prud’homme, R. Car, D.A. Saville, I.A. Aksay, J. Phys. Chem. B 110 (2006)
8535–8539.
Acknowledgements [38] M. Kruk, M. Jaroniec, Chem. Mater. 13 (2001) 3169–3183.
[39] L.L. Zhang, S. Li, J. Zhang, P. Guo, J. Zheng, X.S. Zhao, Chem. Mater. 3 (2010)
This research were supported by National Natural Science Foun- 1195–1202.
[40] T.R. Annunciado, T.H.D. Sydenstricker, S.C. Amico, Mar. Pollut. Bull. 50 (2005)
dation of China (Nos. 51010005, 90916020, 51174063, 21201128), 1340–1346.
the Program for New Century Excellent Talents in University [41] A. Bayat, S.F. Aghamiri, A. Moheb, G.R. Vakili-Nezhaad, Chem. Eng. Technol. 28
(NCET-08-0168), Natural Science Funds for Distinguished Young (2005) 1525–1528.
[42] M.O. Adebajo, R.L. Frost, J.T. Kloprogge, O. Carmody, S.J. Kokot, Porous Mat. 10
Scholar of Heilongjiang province, the project of International (2003) 159–170.
Cooperation supported by Ministry of Science and Technol- [43] A. Li, H.X. Sun, D.Z. Tan, W.J. Fan, S.H. Wen, X.J. Qing, G.X. Li, S.Y. Li, W.Q. Deng,
ogy of China (2013DFR10630), the project of Natural Scientific Energy Environ. Sci. 4 (2011) 2062–2065.

You might also like