Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Journal of Industrial and Engineering Chemistry 21 (2015) 1191–1196

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


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

Adsorption isotherms and kinetics of cationic and anionic dyes


on three-dimensional reduced graphene oxide macrostructure
Han Kim a,1, Sung-Oong Kang a,1, Sungyoul Park b,**, Ho Seok Park a,*
a
Department of Chemical Engineering, College of Engineering, Kyung Hee University, 1 Seochon-dong, Giheung-gu, Youngin-si, Gyeonggi-do 446-701,
Republic of Korea
b
Greenhouse Gas Department, Korea Institute of Energy Research, 152 Gajeong-ro, Yuseong-go, Daejeon 305-343, Republic of Korea

A R T I C L E I N F O A B S T R A C T

Article history: We demonstrate adsorption behaviors of acid red 1 (AC1) and methylene blue (MB) dyes on three-
Received 14 April 2014 dimensional (3D) reduced graphene oxide (rGO) macrostructure. The distinct interactions of two dyes
Accepted 25 May 2014 with the active sites result in the different adsorption behaviors and thus, represent the notable disparity
Available online 2 June 2014
in the adsorption capacities and kinetics. The equilibrium data for MB are fitted to Langmuir isotherm
model, while Freundlich model is suitable for the equilibrium isotherm of AC1. The adsorption rates of
Keywords: both dyes are found to follow the pseudo-second order kinetics. The 3D rGO macrostructures are more
Adsorption
favorable for the adsorption of cationic dyes rather than the anionic dyes due to strong specific
Water purification
Graphene
interactions.
Kinetics ß 2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
Organic dye reserved.

Introduction Recently, carbon-based porous materials assembled by two-


dimensional (2D) building nanoblocks of graphene oxide (GO)
Organic dyes discharged from industrial coloring and printing and reduced GO (rGO) have attracted a great attention as
processes have been one of serious water-contaminating sources effective pollutant adsorbents owing to their high adsorption
because the dyestuffs in wastewater contain toxic organic capacities, rapid adsorption rates, and tunable functionality of
compounds and substances [1,2]. A wide range of physicochemical surface properties along with the intriguing mechanical, thermal
techniques, such as ozonation, flocculation, photo-degradation, and electrical properties of constitutive materials [14–17]. In
adsorption, and membrane filtration, have been employed to refine particular, chemical approaches to construct 3D macroporous
and decolorize the wastewater [3–7]. Among these various structures opened feasible ways to cast the 3D rGO macroscopic
purification treatments, a liquid-phase adsorption on a stationary adsorbents with desired shapes, dimensions, and properties in a
phase is believed to be the most effective way to sequester and mass-production scale [18–20]. In a regime of dye adsorption,
remove organic dyestuffs from the wastewater. Activated carbons rGO has several active sites effectively interacting with the
and silica-based mesoporous materials have been widely and dye molecules, including surface oxygen-containing groups for
commercially used in the liquid-phase purification mainly due to hydrogen bonding and electrostatic interaction, and conjugated
their high degree of porosity and extensive surface area [8–13]. p electrons in sp2 carbon populations [21–23]. Nonetheless,
However, these conventional porous dye adsorbents have been an in-depth study of adsorption isotherms and kinetics of
distressed from the problematic terms of low adsorption organic dyes on the 3D macrostructured rGO platform with the
capability, slow adsorption kinetics, inefficient dye extraction, adsorptive performance, which is comparable to those of
and/or poor recyclability. commercial and conventional adsorbents, has yet to be
addressed [24,25].
Here, we report the adsorption isotherms and kinetics of
* Corresponding author at: Kyung Hee University, Chemical Engineering, cationic and anionic dyes such as acid red 1 (AC1) and methylene
1 Seochon, Giheung, Youngin, Gyeonggi 446-701, Republic of Korea. blue (MB) on the 3D rGO macrostructures. The adsorption MB and
Tel.: +82 31 2013327; fax: +82 31 2048114. AC1 dyestuffs were found to follow the Langmuir and Freundlich
** Corresponding author at: Greenhouse Gas Department, Korea Institute of
Energy Research, 152 Gajeong-ro, Yuseong-go, Daejeon 305-343, Republic of Korea.
isotherm models, respectively, indicating the different adsorption
Tel.: +82 42 8603046; fax: +82 42 8603134. mechanism on the surface of 3D rGO adsorbent. As derived from
E-mail addresses: redsoil@kier.re.kr (S. Park), phs0727@khu.ac.kr (H.S. Park). the pseudo second-order rate equation of the adsorption kinetics
1
First two authors are equally contributed to this work.

http://dx.doi.org/10.1016/j.jiec.2014.05.033
1226-086X/ß 2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
1192 H. Kim et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 1191–1196

for both dyestuffs, the kinetic adsorption parameters represent the and after adsorption. The porosity characteristics of 3D rGO
notable disparity in the adsorption capacities and the equilibrium macrostructures were measured by N2 adsorption–desorption
time, which mainly originate from the specific interactions experiment using a surface area and pore size analyzer
between the dye molecules and the surface of 3D rGO adsorbent (BELSORP mini2, Japan), where the specific surface area (SSA)
in different modes. and the pore size distribution could be determined from the
linear part of the Brunaure–Emmet–Teller (BET) equation and
Experimental the adsorption isotherm based on the Barrett–Joyner–Halenda
(BHJ) method.
Chemicals
Batch equilibrium studies
GO dispersion with a concentration of 2 wt% was purchased
from Angstron Materials (USA). Chemical reagents of hypophos- Adsorption behavior was investigated by a batch method using
phrous acid (H3PO2), iodine (I2), and dyes (acid red 1 and a series of aqueous dye-solutions with the concentration ranging
methylene blue) were obtained from Sigma-Aldrich Chemical from 50 to 300 mg/L. In typical, a certain mass of 3D rGO
Co., Germany. All chemicals were used without further purifica- adsorbents were immersed in each dye-solution and agitated for
tion. 48 h. The pH of solutions was adjusted to 7 by adding 0.1 mol/L HCl
or 0.1 mol/L NaOH solutions. After the adsorption for 48 h, the
Preparation of 3D rGO adsorbents concentrations in the residual dye-solutions were measured using
a spectrophotometer at the maximum absorbance wavelengths of
The 3D rGO adsorbents were produced via the chemical lmax = 664 nm for MB and lmax = 506 nm for AC1, respectively. The
reduction of GO dispersion using a hypophosphorous acid UV spectrophotometer was calibrated against a standard solution
(H3PO2)-iodine (I2) mixture [20]. In brief, the GO suspension before each set of measurement. The amount of the adsorbed
with a concentration of 2 mg/mL prepared from a commercial dyestuffs (qe) at equilibrium on the 3D rGO adsorbents were
GO dispersion was mixed with a certain amount of H3PO2 and I2, determined by the mass balanced equation:
where the weight ratio of precursors was controlled to be
1:100:10. The reduction process was conducted at 80 8C for 12 h ðC o  C e Þ  V
and the product was rinsed with EtOH and distilled water until qe ¼ (1)
M
the pH reaches 7 and then freeze-dried to obtain the 3D rGO
xerogels. where Co and Ce are the initial and equilibrium liquid-phase
concentrations of dyes (mg/g), M is the mass of adsorbent (g) and V
Characterization of 3D rGO macrostructures is the volume of the solutions (L).

Morphologies of chemically reduced 3D rGO were observed Batch kinetic studies


using scanning electron microscope (SEM, LEO SUPRA 55, ZEISS,
Germany) and their chemical compositions were analyzed X-ray For contact time experiments, the residual dye-solutions with
photoelectron spectrometry (XPS, AXIS Ultra DLD, Kratos. Inc.). initial concentrations of 300 mg/L for MB and 50 mg/L for AC1 were
Double-beam UV spectrometer (JASCO V570, Japan) determined collected at a predetermined time interval ranged from 10 to
the concentrations of dyes in the supernatant solution before 90 min. The amount of adsorption at time t, qt (mg/g), was
[(Fig._1)TD$IG]

Fig. 1. (a) Low- and (b) high-magnification SEM images of the chemically reduced 3D rGO macrostructure. Deconvolution of XPS spectra of (c) C1s of GO and (d) C1s of rGO.
[(Fig._2)TD$IG] H. Kim et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 1191–1196 1193

Fig. 2. (a) Equilibrium adsorption isotherms of (a) MB and (b) AC 1 on the 3D rGO macrostructures. The insets show (a) Langmuir isotherm linear plot for the sorption of MB
and (b) Freundlich isotherm linear plot for AC 1, respectively.

calculated by: Table 1


Langmuir and Freundlich isotherm parameters for the adsorption of MB and AC 1
ðC o  C t Þ  V dyes.
qt ¼ (2)
M Dye Langmuir model Freundlich model
2
where Co and Ct (mg/g) are the liquid-phase concentrations of dyes qmax (mg/g) KL r KF 1/n r2
at initial and any time t, M is the mass of adsorbent (g) and V is the MB 302.11 2.190 0.999 173.35 0.128 0.892
volume of the solutions (L). AC 1 277.01 0.031 0.985 28.51 0.364 0.997

Results

Characteristics of porous 3D rGO adsorbents describes how the solutes interact with the adsorbent and assesses
the distribution of solute between the solid and the liquid phase by
The as-prepared 3D rGO macrostructures were characterized measuring the distribution coefficient. As shown in Fig. 2(a), the
using SEM and XPS as shown in Fig. 1. As depicted in the SEM adsorption isotherm of MB onto the 3D rGO adsorbents follows the
images, the 3D macroporous interconnected networks are assem- Langmuir isotherm model, which assumes the monolayer adsorp-
bled by the randomly oriented and wrinkled rGO nanosheets. The tion on structurally and energetically homogeneous active sites
reduction of GO to the 3D rGO structures could be mainly and predicts the monolayer coverage at the outer surface of
confirmed by the C1s XPS spectra, which presents four carbon adsorbent [27]. The linear form of Langmuir isotherm equation is:
atomic components and a considerable decrease in the degree of  
Ce 1 1
oxidation compared with the GO precursor: C–C (284.5 eV), C–O ¼ þ Ce (3)
qe qmax K L qmax
(286.3 eV), C5
5O (289.1 eV) and O–C5 5O (289.1 eV) [26]. The carbon
to oxygen ratio (C/O ratio) significantly increases up to 8.46 from where Ce (mg/L) is the equilibrium concentration of adsorbate, qe
2.38 of the GO precursor. The specific surface area of 3D rGO (mg/g) is the amount of adsorbate adsorbed per unit mass of
macrostructure was 303 m2/g calculated by a BET model (see Fig. adsorbent, and qmax (mg/g) and KL (L/mg) are the Langmuir
S1 and Table S1). As shown in a nitrogen gas adsorption– constants related to maximum monolayer adsorption capacity and
desorption isotherm, type IV with hysteresis is indicative of the energy change in adsorption, respectively.
existence of mesopores arising from the intersheet voids. Based on On the other hand, the equilibrium data of AC1 is best fitted to the
the BET measurement and the SEM observations, the 3D rGO Freundlich isotherm model, which assumes the adsorption process
adsorbents used in this work were found to have a bimodal pore onto the heterogeneous adsorption sites with the different affinities
size distribution consisting of meso- and macropores. of binding sites on the surface of adsorbent and with the distinct
interactions with the dye molecules (Fig. 2(b)) [28]. The logarithmic
Adsorption isotherms form of Freundlich model is given by the following equation:
 
The adsorption thermodynamics of dyestuffs was evaluated as a 1
log qe ¼ log K F þ log C e (4)
[(Fig._3)TD$IG]
function of equilibrium concentration. The adsorption isotherm n

Fig. 3. The equilibrium isotherm data plotted with (a) the Freundlich equation for MB and (b) the Langmuir equation for AC1.
[(Fig._4)TD$IG]
1194 H. Kim et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 1191–1196

Fig. 4. Adsorption of (a) MB and (b) AC1 as a function of time.

where Ce (mg/L) is the equilibrium concentration of the adsorbate, MB rapidly reaches the equilibrium concentration within first
qe (mg/g) is the amount of adsorbate adsorbed per unit mass of 10 min, in sharp contrast, the adsorption of AC1 as a function of time
adsorbent, and KF (mg11/n/L1/n/g) and n are Freundlich constants, is obviously slower by 80 times than the adsorption rate of MB.
KF is the adsorption capacity of adsorbent and n is an indicator for The adsorption kinetics of dyes was plotted using a pseudo first-
favorable or unfavorable adsorption process. order equation [29]:
The value of 1/n calculated to be 0.364 stands for the favorable
logðqe  qt Þ ¼ logqe  k1 t (5)
adsorption of AC1 onto the 3D rGO adsorbent. The isotherm
parameters of dyes are listed in Table 1 and the reliability of the where qe and qt is the amount of dye (mg/g) adsorbed at
isotherm equations is confirmed by the correlation coefficients, equilibrium and at time t, k1 is the overall rate constant of pseudo-
referred to R2. The equilibrium isotherm data plotted with the first order adsorption (min1) and h is the initial adsorption rate.
Freundlich equation for MB and the Langmuir for AC1 are shown in Values of k1 were calculated from the linear plots of ln(qeqt) versus t.
Fig. 3, respectively. Consequently, the equilibrium data for MB are The kinetic isotherm data for both dyes was also fitted to a
fitted to Langmuir isotherm model, while the equilibrium isotherm pseudo second-order equation that is expressed as [30]:
of AC1 dye is fitted to Freundlich model.  
t 1 1
¼ 2
þ t (6)
qt k2 qe qt
Adsorption kinetics
where k2 (g/mg/min) is the rate constant of second-order
Adsorption kinetics experiments were conducted to investigate adsorption. The adsorption kinetics treated with pseudo first-
the effect of contact time and to obtain the resulting kinetic and second-order models are shown in Fig. 5.
parameters. Fig. 4 presents the variation in the adsorption of both Kinetic parameters obtained from the pseudo first- and second-
dyes as a function of contact time. It is recognized that the uptake of order equations are listed in Table 2. The large discrepancy
[(Fig._5)TD$IG]

Fig. 5. Pseudo first-order kinetics for the adsorption of (a) MB and (b) AC1, and the pseudo second-order kinetics for the adsorption of (c) MB and (d) AC1, resepctively.
H. Kim et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 1191–1196 1195

Table 2
Comparison of the pseudo first- and second-order kinteic parameters for the adsorption of MB and AC 1.

Pseudo 1st kinetic model Pseudo 2nd kinetic model


2
Dye qe (exp.) (mg/g) k1 (min 1
) qe (cal.) (mg/g) r k2 (g/mg/min) qe (cal.) (mg/g) r2

MB 300.60 0.048 69.02 0.721 0.0050 301.20 0.999


AC 1 50.61 0.013 121.94 0.948 0.0004 51.97 0.998

between the experimental and the calculated qe indicates that the the AC1 molecules hamper the adsorption of dye molecules and
adsorption of both dyes does not follow the pseudo first-order thus slow down the adsorption rate onto the 3D rGO adsorbent. In
kinetic model. Instead, the linear plots of t/q versus t represent the consequence, the chemically reduced 3D rGO scaffolds can be the
best agreement between the experimental and the calculated qe efficient platform for the adsorption of positively charged MB with
values with the high R2 values that have been generally indexed for the value of qe of 300 mg/g and the fast adsorption rate compared
the adsorption kinetics of carbon-based porous structures [9,31– with the activated carbons [9,33], however, may be unsuitable for
33]. As a result, the pseudo second-order kinetic model is efficient adsorption of AC1 negatively charged from the dissoci-
applicable to the adsorption kinetics of MB and AC1 dyes onto ated SO3 groups in the aqueous solution [12,13]. Nevertheless, it
the 3D rGO adsorbents, in which the adsorption process is the rate- is noteworthy that the saturated adsorption capacity of MB on the
limiting step involving the surface and intra-particle adsorption 3D rGO adsorbent per unit surface area (approximately 1 mg/m2)
[34]. is considerably higher than those of the conventional adsorbents.
The fast kinetics in the MB dye adsorption also suggests the 3D
Discussion rGO macrostructures to surpass the adsorption performance of
the traditional porous materials. By extending the surface area
The resultant kinetic parameters can be explained by the and the surface-functionalization of graphene, the large adsorp-
different adsorption mechanisms of cationic (MB) and anionic tion capacity per unit area and the rapid adsorption rate for the
(AC1) dyes onto 3D rGO adsorbent. As aforementioned, the 3D rGO specific dye species strongly proposes a potential of this
adsorbent has several active sites interacting with the dye graphene-based structures to be versatile, light-weight and
molecules: (i) negatively charged surface-functional groups of cost-effective dye-pollutant adsorbents with ultrahigh adsorptive
hydroxyl (–OH) and carboxyl (–COOH) groups in an aqueous capability and fast adsorption kinetics [35–37]. Besides, the
solution; (ii) delocalized p electrons within sp2 carbon grains synthetic feasibility in casting the rGO macrostructures into the
interacting with the lone electron pairs of atoms and the free desired dimensions, sizes and shapes is greatly beneficial in the
electrons in the aromatic rings of the dye molecules (p–p commercial and mass-scaled applications, which is inherently
interaction); (iii) oxygen-containing groups available to form the different from the use of fine particles of activated carbons and
hydrogen bonds with the dye species [21,35]. Considering the silica-based materials [14,15,17,20].
chemical structures of dye molecules (see Fig. S2), the large [(Fig._6)TD$IG]
discrepancy in the adsorption capacities and the equilibrium time
of dyes obtained from the kinetic studies (Table 2) originates from
the different interaction between the rGO-based adsorbent and the
dye molecules. For the adsorption of MB, two types of interactions
are expectable, such as the electrostatic attraction between the
positively charged amino groups and the negatively charged
oxygen-containing surface groups, and the p–p interaction
between the localized p electrons in the conjugated aromatic
rings of the adsorbent and the dye molecules. However, hydrogen
bonding between the oxygen-containing functional groups on the
rGO surface and the MB dye species is not dominant. On the other
hand, the negatively charged sulfonate (SO3) groups of AC1
molecules counteract the electrostatic interaction with the
oxygen-containing groups on the rGO surface. Instead, the p–p
interaction and the hydrogen bonding with the rGO adsorbent may
play dominant roles in the dye adsorption of AC1. In particular, the
intermolecular hydrogen bonding between the AC1 dye molecules
induces the multilayer formation of AC1 dye molecules in the
aqueous solution, which plausibly explains the Freundlich model
describing the multilayer adsorption of AC1 dye. As a result, such
different interacting modes of anionic and cationic dyes reasonably
results in the notable disparity between the adsorption capacities
and the equilibrium time for the adsorption onto the 3D rGO
adsorbent. Schematic illustrations of interactions between the dye
molecules and the 3D rGO adsorbent are displayed in Fig. 6.
The adsorption process can be suggested as follow: the solute
diffuses from the bulk solution through the macropores to the film
surrounding the adsorbent and transport from the film onto the
surface of adsorbent, and finally the adsorbate is adsorbed onto
the active sites. The final adsorption step is the slowest rate-
limiting step in the entire adsorption process. That is, the Fig. 6. Schemes of interactions between the surface of rGO-based adsorbent and (a)
repulsive surface charges between the surface of adsorbent and MB dye molcules and (b) AC1 dye molecules.
1196 H. Kim et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 1191–1196

Conclusions [4] J. Panswed, S. Wongchaisuwan, Water Sci. Technol. 18 (1986) 139.


[5] L. Lucarelli, V. Nadtochenko, J. Kiwi, Langmuir 16 (2000) 1102.
[6] F.-C. Wu, R.-L. Tseng, J. Hazard. Mater. 152 (2008) 1256.
In summary, the adsorption of MB and AC1 dyes on the chemically [7] E. Eren, B. Afsin, Dyes Pigm. 76 (2008) 220.
reduced 3D rGO macrostructure was studied. It was found that the [8] P.C.C. Faria, J.J.M. Órfão, M.F.R. Pereira, Water Res. 38 (2004) 2043.
[9] B.H. Hameed, A.T.M. Din, A.L. Ahmad, J. Hazard. Mater. 141 (2007) 819.
surface charges and the chemical structures of dye molecules bring [10] A. Sayari, S. Hamoudi, Y. Yang, Chem. Mater. 17 (2005) 212.
about the notable disparity in the adsorptive capacities and the [11] A.M. Donia, A.A. Atia, W.A. Al-amrani, A.M. El-Nahas, J. Hazard. Mater. 161 (2009)
equilibrium adsorption time based on the different adsorption 1544.
[12] M. Anbia, S. Salehi, Dyes Pigm. 94 (2012) 1.
mechanisms. The resulting isotherm and the kinetic parameters [13] Z. Yan, G. Li, L. Mu, S. Tao, J. Mater. Chem. 16 (2006) 1717.
indicate that the chemically reduced 3D rGO macrostructures can be [14] Z. Niu, J. Chen, H.H. Hng, J. Ma, X. Chen, Adv. Mater. 24 (2012) 4144.
effective for the adsorption of cationic dyes, on the other hand, may [15] S.J. Yang, J.H. Kang, H. Jung, T. Kim, C.R. Park, J. Mater. Chem. A 1 (2013)
9427.
be unsuitable for the anionic dyes. By the extension of surface area
[16] Y. Zhao, C. Hu, Y. Hu, H. Cheng, G. Shi, L. Qu, A. Versatile, Angew. Chem. Int. Ed. 51
and the surface-tailoring of graphene, we believe that such (2012) 11371.
graphene-based structures can significantly contribute to develop- [17] H. Bi, X. Xie, K. Yin, Y. Zhou, S. Wan, L. He, F. Xu, F. Banhart, L. Sun, R.S. Ruoff, Adv.
ment of the next-generation dye adsorbents to be efficiently Funct. Mater. 22 (2012) 4421.
[18] S. Park, R.S. Ruoff, Nat. Nanotechnol. 4 (2009) 217.
implemented and utilized in the water-purification field. [19] I.K. Moon, J. Lee, R.S. Ruoff, H. Lee, Nat. Commun. 1 (2010) 73.
[20] H.D. Pham, V.H. Pham, T.V. Cuong, T.-D. Nguyen-Phan, J.S. Chung, E.W. Shi, S. Kim,
Chem. Commun. 47 (2011) 9672.
Acknowledgments
[21] S. Pei, H.-M. Cheng, Carbon 50 (2012) 3210.
[22] D.R. Dreyer, S. Park, C.W. Bielawski, R.S. Ruoff, Chem. Soc. Rev. 39 (2010) 228.
This work was supported by National Agenda Project (NAP) [23] K.P. Loh, Q. Bao, P.K. Ang, J. Yang, J. Mater. Chem. 20 (2010) 2277.
[24] Z. Geng, Y. Lin, X. Yu, Q. Shen, L. Ma, Z. Li, N. Pan, X. Wang, J. Mater. Chem. 22
through National Fusion Research Institute (NFRI), the research
(2012) 3527.
program of Korea Institute of Energy Research (Project no. B4- [25] X. Yu, H. Cai, W. Zhang, X. Li, N. Pan, Y. Luo, X. Wang, J.G. Hou, ACS Nano 5 (2011)
2434-03), and the Fundamental R&D Program for Core Technology 952.
of Materials funded by the Ministry of Trade, Industry & Energy. [26] D. Yang, A. Velamakanni, G. Bozoklu, S. Park, M. Stoller, R.D. Piner, S. Stankovich, I.
Jung, D.A. Field, C.A. Ventrice Jr., R.S. Rouff, Carbon 47 (2009) 145.
[27] I. Langmuir, J. Am. Chem. Soc. 40 (1918) 1361.
Appendix A. Supplementary data [28] H.M.F. Freundlich, Z. Phys. Chem. 57A (1906) 385.
[29] Y.S. Ho, J.C.Y. Ng, G. McKay, Sep. Purif. Methods 29 (2000) 189.
[30] G. Mckay, Y.S. Ho, Process Biochem. 34 (1999) 451.
Supplementary data associated with this article can be found, in [31] P.K. Malik, J. Hazard. Mater. B113 (2004) 81.
the online version, at doi:10.1016/j.jiec.2014.05.033. [32] M. Zhao, P. Liu, Desalination 249 (2009) 331.
[33] B.H. Hameed, A.L. Ahmad, K.N.A. Latiff, Dyes Pigm. 75 (2007) 143.
[34] A. Sharma, K.G. Bhattacharyya, Adsorption 10 (2004) 327.
References [35] D.W. Lee, T.-K. Hong, D. Kang, J. Lee, M. Heo, J.Y. Kim, B.S. Kim, H.S. Shin, J. Mater.
Chem. 21 (2011) 3438.
[36] Y. Zhu, S. Murali, M.D. Stoller, K.J. Ganesh, W. Cai, P.J. Ferreira, A. Pirkle, R.M.
[1] Y. Wong, J. Yu, Water Res. 33 (1999) 3512.
Wallace, K.A. Cychosz, M. Thommes, D. Su, E.A. Stach, R.S. Ruoff, Science 332
[2] K.R. Ramakrishna, T. Viraraghavan, Water Sci. Technol. 36 (2–3) (1997) 189.
(2011) 1537.
[3] M. Koch, A. Yedlier, D. Lienert, G. Insel, A. Kettrup, Chemosphere 46 (2002)
[37] M.F. El-Kady, V. Strong, S. Dubin, R.B. Kaner, Science 335 (2012) 1326.
109.

You might also like