Вы находитесь на странице: 1из 8

Journal of Colloid and Interface Science 295 (2006) 71–78

www.elsevier.com/locate/jcis

Application of zeolite MCM-22 for basic dye removal from wastewater


Shaobin Wang a,∗ , Huiting Li a , Longya Xu b
a Department of Chemical Engineering, Curtin University of Technology, G.P.O. Box U1987, Perth, WA 6845, Australia
b Dalian Institute of Chemical Physics, Chinese Academy of Science, Dalian, People’s Republic of China

Received 30 June 2005; accepted 3 August 2005


Available online 6 September 2005

Abstract
MCM-22 was employed as an effective adsorbent for removal of basic dyes including methylene blue, crystal violet, and rhodamine B from
aqueous solution. The adsorption kinetics and isotherms were investigated. The adsorption capacity of MCM-22 for three dyes follows an order of
MB > CV ∼ RB. Kinetic studies indicate that the adsorption follows the pseudo second-order kinetics and the adsorption is a two-step diffusion
process with film diffusion dominating the process. The adsorption isotherm can be well fitted by both the Langmuir and the Freundlich models.
Thermodynamic calculations suggest that the adsorption of basic dyes on MCM-22 is an endothermic reaction.
 2005 Elsevier Inc. All rights reserved.

Keywords: MCM-22; Adsorption kinetics; Isotherm; Basic dyes; Wastewater treatment

1. Introduction zeolite composites for methylene blue adsorption and found


that the adsorption capacity of zeolite decreased when it was
Dye effluents, discharged from the dyestuff manufacturing, embedded in composites. Meshko et al. [5] studied the adsorp-
dyeing, printing, and textile industries, may contain chemicals tion of basic dyes from aqueous solutions onto granular acti-
that exhibit toxic effects toward microbial populations and can vated carbon and natural zeolite. The equilibrium studies have
be toxic and/or carcinogenic to mammals. Conventional meth- shown that the granular activated carbon has a higher adsorp-
ods for the removal of dyes in effluents include physical, chem- tion capacity than the natural zeolite. Armagan and co-workers
ical, and biological processes [1]. Physical adsorption is gener- [6–8] examined the ability of natural and modified zeolites to
ally considered to be an effective method for quickly lowering remove reactive dyes from aqueous solutions. The adsorption
the concentration of dissolved dyes in an effluent, and acti- results indicated that the natural zeolite has a limited adsorp-
vated carbon is the most widely used adsorbent for dye removal. tion capacity for reactive dyes but is substantially improved
However, activated carbon suffers from high-cost production upon modification of its surfaces with quaternary amines. Qua-
and regeneration [2]. Therefore, other adsorbents such as ze- ternary amine surfactant is not only hydrophobic to the surface
olite, pillared clays with higher surface areas are alternatives. of the samples but also neutralizes the negative charge, which
Synthetic and natural zeolites have become increasingly im- makes the samples receptive to the negative charge of the dye
portant due to the wide range of their chemical and physical molecules by means of electrostatic attraction. The modified
properties and have been used as adsorbents, molecular sieves, zeolites are comparable to activated carbon. Metes et al. [3] in-
membranes, ion exchangers, and catalysts in the past decades. vestigated several synthetic zeolites for cleaning printing ink
However, the application of zeolites for dye removal from wastewater and found that adsorption is independent of pore
wastewater has rarely been previously reported [3]. Balkose structure. ZSM-5 and NH4 -Beta are effective while other zeo-
et al. [4] reported an investigation using poly(vinyl chloride)- lites studied showed a lower efficiency. We also tested a natural
zeolite for dye removal from aqueous solution [9]. The natural
* Corresponding author. zeolite exhibits a higher adsorption capacity than fly ash but
E-mail address: wangshao@vesta.curtin.edu.au (S. Wang). still lower than unburned carbon in fly ash.
0021-9797/$ – see front matter  2005 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcis.2005.08.006
72 S. Wang et al. / Journal of Colloid and Interface Science 295 (2006) 71–78

MCM-22 is a novel nanoporous zeolite, synthesized by Mo- 2.2. Sorption test


bil researchers in 1990 [10]. It consists of two independent
and noninterconnecting pore systems: one consists of two- Adsorption kinetics and isotherm experiments for all sam-
dimensional channels with 10-ring pore openings (0.6 nm in ples were undertaken in a batch reactor. The adsorption of
diameter); the second pore system consists of cavities with 12- dye was performed by shaking 0.02 g of solids in a 200 ml
ring openings exposed on the (001) crystals surface with egg- dye solution with varying concentrations at 100 rpm (Certomat
shaped supercages (1.4 nm in diameter). The unusual frame- R shaker from B. Braun) at different temperatures. The sam-
work topology, high thermal stability, large surface area and ple was collected by separation of solid from solution using a
good adsorption capacity render this zeolite very interesting for centrifuge. The determination of dye concentration was done
adsorption and catalysis [11–15]. In the past a few years, MCM- spectrophotometrically on a Spectronic 20 Genesis Spectropho-
22 has been used in gas adsorption and catalysis; few applica- tometer (USA) by measuring absorbance at λmax of 665, 590,
tions have been reported for MCM-22 in aqueous adsorption. and 556 nm for MB, CV, and RB, respectively. From the dy-
Dahl et al. [16] reported the application of MCM-22 to adsorb namic adsorption curve, the equilibrium time was determined
alcohols from water solutions. Enhanced adsorption was ob- for 10 days.
tained for alcohols with longer alkyl chains. To our knowledge, The data obtained from the adsorption tests were then used
no other investigations have been conducted for adsorption in to calculate the adsorption capacity, qt (mol/g), of the ad-
aqueous solution. In this paper, we report an investigation using sorbent by a mass–balance relationship, which represents the
MCM-22 as an effective adsorbent to remove some basic dyes, amount of adsorbed dye per the amount of dry adsorbent,
methylene blue, crystal violet, and rhodamine B, from aqueous (C0 − Ct )
solution. The kinetics and adsorption equilibrium of the basic qt = V, (1)
W
dyes on MCM-22 were investigated.
where C0 and Ct are the concentrations of dye in solution
(mol/dm3 ) at time t = 0 and t = t , respectively, V is the vol-
2. Experimental
ume of the solution (dm3 ), and W is the weight of the dry
adsorbent used (g).
2.1. Adsorbent and dyes
3. Results and discussion
The MCM-22 zeolite was synthesized by using an organic
mixture solution containing hexamethyleneimine and caprolac- 3.1. Adsorption of different dyes
tam as a template. The detailed synthesis procedure is described
in Ref. [17]. The MCM-22 has a Si/Al ratio of 30 with a surface Fig. 2 presents the dynamic adsorption of three different
area of 490 m2 /g. dyes on MCM-22. One can see that adsorption is fast before
Three basic dyes, methylene blue (MB), crystal violet (CV), 20 h and then it approaches equilibrium after 200 h. The time
and rhodamine B (RB), were selected for adsorption tests. They profile of dye adsorption is a single, smooth, and continuous
were obtained from AJAX Chemical. Their chemical structures curve leading to saturation, suggesting the possible monolayer
are displayed in Fig. 1. A stock solution with concentration at coverage of dye on the surface of the adsorbents. For the dif-
10−4 M was prepared and the solutions for adsorption tests ferent dyes, equilibrium adsorption is quite different. MCM-22
were prepared from the stock solution to the desired concen- exhibits much higher adsorption for MB while it shows simi-
trations (10−6 –10−5 M). lar adsorption for CV and RB. The equilibrium adsorption for
MB, CV, and RB will be around 1.7 × 10−4 , 1.2 × 10−4 , and

Fig. 1. Chemical structure of basic dyes. Fig. 2. Adsorption of basic dyes on MCM-22 at 30 ◦ C.
S. Wang et al. / Journal of Colloid and Interface Science 295 (2006) 71–78 73

Fig. 3. Adsorption isotherm of MB on MCM-22 at different temperatures. (a) Freundlich model, (b) Langmuir model.

Fig. 4. Adsorption isotherm of CV on MCM-22 at different temperatures. (a) Freundlich model, (b) Langmuir model.

1.1 × 10−4 mol/g. The difference in adsorption for three dyes is adsorption capacity, and KL is the constant related to the free
probably ascribed to the varying molecular size. As seen from energy of adsorption.
Fig. 1, MB is quite smaller than CV and RB, so that MB easily The Freundlich isotherm is an empirical equation employed
enters the inner pores of MCM-22. Vinod and Anirudhan [18] to describe heterogeneous systems. The Freundlich equation is
investigated the adsorption of three basic dyes, MB, CV, and
1/n
RB, on the humic acid (HA)-immobilized pillared clay (PILC) Qe = KCe , (3)
and found the adsorption follows the order of MB > CV > RB.
where K and n are Freundlich adsorption isotherm constants,
Graham et al. [19] studied methylene blue and crystal violet ad-
indicative of the extent of the adsorption and the degree of
sorption on activated carbon and reported similar results.
nonlinearity between solution concentration and adsorption, re-
spectively.
3.2. Adsorption isotherms
Figs. 3–5 present adsorption isotherms of three dyes on
The equilibrium adsorption isotherm is of importance in the MCM-22 at different temperatures and the fitted curves for
design of adsorption systems. Several isotherm equations are Langmuir and Freundlich models. The parameters obtained
available and two important isotherms are selected in this study, from the isotherms are given in Table 1. As seen the adsorp-
the Langmuir and Freundlich isotherms. tion isotherms at different temperatures for the same dye are
The Langmuir isotherm is based on an assumption that the very close and can be fitted very well using two isotherm mod-
adsorption occurs at specific homogeneous sites within the ad- els (evidenced from the correlation coefficients, >0.990). From
sorbent. The equation is Table 1, it is also seen that the obtained adsorption capacities
from the Langmuir isotherm show slightly higher adsorption
Q0 KL Ce for CV than RB on MCM-22.
Qe = , (2)
(1 + KL Ce ) The increase in dye adsorption with increasing temperature
where Qe is the adsorbed amount of the dye, Ce is the equilib- might also be due to the enhanced rate of external and intra-
rium concentration of the dye in solution, Q0 is the monolayer particle diffusion of the adsorbate, as diffusion is a dominant
74 S. Wang et al. / Journal of Colloid and Interface Science 295 (2006) 71–78

Fig. 5. Adsorption isotherm of RB on MCM-22 at different temperatures. (a) Freundlich model, (b) Langmuir model.

Table 1
Parameters of adsorption isotherms of basic dyes on MCM-22 at different temperatures
Temperature (◦ C) Dye Langmuir model Freundlich model
Q0 (mol/g) KL (dm3 /mol) R2 K [mol/g (L/mol)1/n ] 1/n R2
30 MB 1.73 × 10−4 4.03 × 107 0.999 3.11 × 10−4 0.0486 0.997
40 1.79 × 10−4 2.63 × 107 0.985 4.06 × 10−4 0.0673 0.996
50 1.76 × 10−4 3.53 × 107 0.999 2.86 × 10−4 0.0407 0.998
30 CV 1.20 × 10−4 8.97 × 106 0.994 2.06 × 10−4 0.0473 0.990
40 1.20 × 10−4 6.12 × 106 0.995 2.22 × 10−4 0.0543 0.991
50 1.26 × 10−4 8.70 × 106 0.995 2.27 × 10−4 0.0512 0.992
30 RB 1.05 × 10−4 8.89 × 106 0.989 3.10 × 10−4 0.0891 0.996
40 1.13 × 10−4 4.07 × 106 0.994 3.10 × 10−4 0.0891 0.996
50 1.11 × 10−4 7.87 × 106 0.999 3.35 × 10−4 0.0919 0.997

Table 2 Table 3
Comparison of adsorption capacity of various adsorbents Thermodynamic parameters for the adsorption of basic dyes on MCM-22
Dyes Adsorbent Adsorption capacity Reference Dye G0 (kJ/mol) H 0 (kJ/mol) S 0 (J/mol K)
(mol/g) 30◦ 40◦ 50◦
MB AC from coal 1.38 × 10−3 [21] MB −44.1 −44.5 −46.7 5.40 159.3
MB AC-aluminosilicate 3.66 × 10−4 [22] CV −40.3 −40.7 −43.0 1.24 133.9
MB AC from bamboo 4.48 × 10−4 [23] RB −40.3 −39.6 −42.6 4.96 142.3
MB AC from coconut shell 8.70 × 10−4 [23]
MB AC from coconut shell 5.24 × 10−5 [20]
MB AC from coconut husk 2.68 × 10−4 [19] MB, CV, and RB. MCM-22 shows a higher adsorption capacity
MB Natural zeolite 4.50 × 10−5 [9] than natural zeolite and comparable capacity to activated car-
CV AC from coconut husk 1.67 × 10−4 [19] bon.
CV AC from saw dust 8.37 × 10−4 [24]
Thermodynamic parameters, i.e., free energy (G0 ), en-
CV Humid acid-modified pillared clay 1.19 × 10−3 [18]
RB Humid acid-modified pillared clay 8.62 × 10−4 [18] thalpy (H 0 ), and entropy (S 0 ) changes, were also calculated
RB AC from rice husk (0.7–1.0) × 10−3 [25] based on the adsorption isotherms using Eqs. (4)–(6) [20] and
RB Activated charcoal 2.09 × 10−4 [26] are given in Table 3.
RB AC from fertilizer 1.67 × 10−4 [26]
1.80 × 10−4 G0 = −RT ln K1 , (4)
MB MCM-22 [This work]  
CV MCM-22 1.20 × 10−4 [This work] T2 T 1 K2
1.10 × 10−4
H 0 = −R ln , (5)
RB MCM-22 [This work] T 2 − T1 K1
H 0 − G0
S 0 = , (6)
process in the adsorption with an endothermic nature, which T
will be discussed in Section 3.3. where K1 and K2 are the Langmuir constants at T1 = 30 and
Table 2 compares the adsorption capacity of MCM-22 with T2 = 50 ◦ C.
activated carbons and other adsorbents reported before. As seen As seen G0 for all cases is negative, which indicates
activated carbon generally has a high adsorption capacity for the feasibility and spontaneous nature of basic dye adsorption
S. Wang et al. / Journal of Colloid and Interface Science 295 (2006) 71–78 75

Fig. 6. Comparison of kinetic models of MB adsorption on MCM-22.


Fig. 7. Comparison of kinetic models of CV adsorption on MCM-22.
(a) First-order kinetics, (b) second-order kinetics, (c) diffusion model.
(a) First-order kinetics, (b) second-order kinetics, (c) diffusion model.

on MCM-22. The change in enthalpy (H 0 ) for MB, CV, t 1 1


= 2
+ t, (8)
and RB was found to be positive. The positive values con- qt k 2 qe qe
firm the endothermic nature of adsorption. The positive values
where k1 is the rate constant of pseudo-first-order adsorption
of the entropy change show the increased randomness at the
(h−1 ), k2 (g mol−1 h) the rate constant of pseudo-second-order
solid/solution interface with some structural changes in the ad-
adsorption, and qe and qt are amount of dye adsorbed on adsor-
sorbate and adsorbent and an affinity of the adsorbent toward
bent (mol g−1 ) at equilibrium and at time t , respectively.
those basic dyes.
Equation (7) can be transformed into nonlinear forms, which
can be used to predict the adsorption equilibrium:
3.3. Adsorption kinetics
qt = qe (1 − e−k1 t ). (9)
A study of adsorption kinetics is desirable as it provides
information about the mechanism of adsorption, which is im- In addition, the intraparticle diffusion model is a commonly
portant for the efficiency of the process. Successful application used technique for identifying the mechanism involved in the
of the adsorption demands innovation of cheap, easily avail- sorption process, which is expressed as
able, and abundant adsorbents of known kinetic parameters and
qt = kd (t 1/2 ), (10)
sorption characteristics. Adsorption kinetics can be modeled by
several models, the pseudo-first-order Lagergren equation [27] where kd is the diffusion coefficient.
and pseudo-second-order rate equation [28] given below as (7) Figs. 6–8 present the plots for the adsorption of three dyes on
and (8), respectively. MCM-22 using the pseudo-first-order kinetics, pseudo-second-
k1 order kinetics, and diffusion model. As shown the first-order
log(qe − qt ) = log qe − t, (7) model does not seem to be good for modeling the kinetics of
2.303
76 S. Wang et al. / Journal of Colloid and Interface Science 295 (2006) 71–78

Table 4
Parameters of kinetic models of basic dye adsorption on MCM-22
Dye Pseudo-first-order model Pseudo-second-order model Diffusion
k1 (h−1 ) qe (mol/g) R2 k2 (g/mol h) qe (mol/g) R2 kd R2
MB 0.5038 1.52 × 10−4 0.854 1247.7 1.70 × 10−4 0.999 3.55 × 10−6 0.964
CV 0.4910 1.01 × 10−4 0.844 2555.1 1.12 × 10−4 0.999 2.42 × 10−6 0.938
RB 0.4449 1.04 × 10−4 0.858 2717.8 1.10 × 10−4 0.999 2.16 × 10−6 0.919

Fig. 9. Corelationship between Bt and t of various dye adsorptions on


MCM-22.

pattern [29,30]. Allen et al. investigated a basic dye adsorp-


tion on peat and attributed the diffusion in the three regions
to macropore diffusion, transitional pore diffusion, and microp-
ore diffusion, respectively [31]. MCM-22 possess two types of
pores and thus presents a two-phase intraparticle diffusion. The
results from Table 4 also show that MB has the highest diffu-
sion coefficient followed by CV and RB, resulting in the highest
adsorption capacity for MB.
In order to determine the actual rate-controlling step in-
volved in the dye sorption process, the sorption data were fur-
ther analyzed using the kinetic expression given by Boyd et al.
[30,32]
6
F =1− exp(−Bt), (11)
π2
Fig. 8. Comparison of kinetic models of RB adsorption on MCM-22. where F is the fraction of solute adsorbed at different times t
(a) First-order kinetics, (b) second-order kinetics, (c) diffusion model. and Bt is a mathematical function of F and given by
qt
F= , (12)
the whole adsorption process and the regression coefficients are qe
less than those obtained from the second-order kinetics and dif- where qt and qe represent the amount adsorbed (mol/g) at any
fusion model (Table 4). From the results, it is also seen that the time t and at infinite time. In this work, we take qe from the
equilibrium adsorption from the pseudo-second-order model is second-order kinetic model. Substituting Eq. (12) into Eq. (11),
close to the experimental data, suggesting the better application the kinetic expression becomes
of the second-order kinetics. For the diffusion model plots, one  
qt
can see that the process involves two phases. The two-phase Bt = −0.4977 − ln 1 − . (13)
plot suggests that the adsorption process proceeds by surface qe
sorption and intraparticle diffusion. The initial curved portion Thus the value of Bt can be calculated for each value of F us-
of the plot indicated a boundary-layer effect while the sec- ing Eq. (13). The calculated Bt values were plotted against time
ond linear portion is due to intraparticle or pore diffusion [29]. as shown in Fig. 9. The linearity of this plot will provide use-
Several previous investigations have reported a similar type of ful information for distinguishing between external-transport-
S. Wang et al. / Journal of Colloid and Interface Science 295 (2006) 71–78 77

Table 5
Parameters of kinetic models of basic dye adsorption on MCM-22
Temperature (◦ C) Dye Pseudo-second-order model
k2 (g/mol h) qe (mol/g) R2
30 MB 1247.7 1.70 × 10−4 0.999
40 1278.8 1.63 × 10−4 0.999
50 1461.8 1.74 × 10−4 0.999
30 CV 2555.1 1.12 × 10−4 0.999
40 2223.9 1.13 × 10−4 0.998
50 2558.9 1.17 × 10−4 0.998
30 RB 2717.8 1.10 × 10−4 0.999
40 6157.65 1.08 × 10−4 0.999
50 4056.0 1.09 × 10−4 0.999

Fig. 10 shows the dynamic adsorption of MB, CV, and RB


on MCM-22 at varying temperatures. As seen the amount of
adsorption for MCM-22 at different temperatures is very close,
which suggests that temperature does not play a significant role
in adsorption, as for the adsorption isotherm. Table 5 presents
the kinetic parameters obtained from the pseudo-second-order
model and it is seen that the equilibrium adsorption capacity
shows a slight increase with the increasing temperature. The
kinetic constants are similar at three temperatures.

4. Conclusion

MCM-22 is an effective adsorbent for basic dye removal


from aqueous solutions. The adsorption capacities for MB,
CV, and RB can reach 1.8 × 10−4 , 1.2 × 10−4 , and 1.1 ×
10−4 mol/g, respectively. The adsorption kinetics follows the
pseudo-second-order model and the external diffusion is the
controlling process. The adsorption isotherms can be well fitted
with Langmuir and Freundlich models. Thermodynamic calcu-
lations indicate that the adsorption of basic dyes on MCM-22
is spontaneous and an endothermic reaction. The H 0 for MB,
CV, and RB is 5.4, 1.2, and 5.0 kJ/mol, respectively.

References

[1] H.S. Rai, M.S. Bhattacharyya, J. Singh, T.K. Bansal, P. Vats, U.C. Baner-
jee, Crit. Rev. Environ. Sci. Technol. 35 (2005) 219–238.
[2] S.B. Wang, Y. Boyjoo, A. Choueib, Z.H. Zhu, Water Res. 39 (2005) 129–
138.
[3] A. Metes, D. Kovacevic, D. Vujevic, S. Papic, Water Res. 38 (2004) 3373–
3381.
Fig. 10. Effect of temperature on dynamic adsorption of basic dyes on [4] D. Balkose, S. Ulutan, F. Ozkan, S. Ulku, U. Kokturk, Sep. Sci. Tech-
MCM-22. (a) MB, (b) CV, (c) RB. nol. 31 (1996) 1279–1289.
[5] V. Meshko, L. Markovska, M. Mincheva, A.E. Rodrigues, Water Res. 35
(2001) 3357–3366.
and intraparticle-transport-controlled rates of adsorption [30]. [6] B. Armagan, O. Ozdemir, M. Turan, M.S. Celik, J. Chem. Technol. Biot.
If a plot of Bt vs time (having slope B) is a straight line 78 (2003) 725–732.
passing through the origin, then adsorption is governed by a [7] B. Armagan, M. Turan, O. Ozdemir, M.S. Celik, J. Environ. Sci. Heal.
particle-diffusion mechanism, otherwise it is governed by film A 39 (2004) 1251–1261.
[8] O. Ozdemir, B. Armagan, M. Turan, M.S. Celik, Dyes Pigments 62 (2004)
diffusion. From Fig. 9, it was observed that the plots were not
49–60.
linear, indicating that external mass transport mainly governs [9] S.B. Wang, L. Li, H. Wu, Z.H. Zhu, J. Colloid Interface Sci. (2005), in
the rate-limiting process. Other reports also show that adsorp- press.
tion of dyes on adsorbents is film-diffusion controlled [29,30]. [10] M.K. Robin, P. Chu, US patent 4954325, 1990.
78 S. Wang et al. / Journal of Colloid and Interface Science 295 (2006) 71–78

[11] F. Eder, Y.J. He, G. Nivarthy, J.A. Lercher, Recueil Travaux Chim. Pays- [20] K.P. Singh, D. Mohan, S. Sinha, G.S. Tondon, D. Gosh, Ind. Eng. Chem.
Bas J. R. Neth. Chem. Soc. 115 (1996) 531. Res. 42 (2003) 1965–1976.
[12] A. Corma, V. Martinez-Soria, E. Schnoeveld, J. Catal. 192 (2000) 163– [21] Y. Zou, B.X. Han, Energy Fuel 15 (2001) 1383–1386.
173. [22] R.A. Shawabkeh, Micropor. Mesopor. Mater. 75 (2004) 107–114.
[13] G. Dahlhoff, U. Barsnick, W.F. Holderich, Appl. Catal. A 210 (2001) 83– [23] N. Kannan, M.M. Sundaram, Dyes Pigments 51 (2001) 25–40.
95. [24] S. Chakraborty, S. De, S. DasGupta, J.K. Basu, Chemosphere 58 (2005)
[14] N. Kumar, R. Byggningsbacka, M. Korpi, L.E. Lindfors, T. Salmi, Appl. 1079–1086.
Catal. A 227 (2002) 97–103. [25] Y.P. Guo, J.Z. Zhao, H. Zhang, S.F. Yang, J.R. Qi, Z.C. Wang, H.D. Xu,
[15] J. Cejka, A. Krejci, N. Zilkova, J. Kotrla, S. Ernst, A. Weber, Micropor. Dyes Pigments 66 (2005) 123–128.
Mesopor. Mater. 53 (2002) 121–133. [26] A. Bhatnagar, A.K. Jain, J. Colloid Interface Sci. 281 (2005) 49–55.
[16] I.M. Dahl, E. Myhrvold, A. Slagtern, M. Stocker, Adsorpt. Sci. Tech- [27] Y.S. Ho, C.C. Chiang, Adsorpt. J. Int. Adsorpt. Soc. 7 (2001) 139–147.
nol. 15 (1997) 289–299. [28] Y.S. Ho, G. McKay, Process Biochem. 34 (1999) 451–465.
[17] Q.X. Wang, S.J. Xie, L.Y. Xu, J. Bai, Z.H. Wu, L.Y. Zhang, CN 1113812C, [29] V. Vadivelan, K.V. Kumar, J. Colloid Interface Sci. 286 (2005) 90–100.
2003. [30] K.V. Kumar, V. Ramamurthi, S. Sivanesan, J. Colloid Interface Sci. 284
[18] V.P. Vinod, T.S. Anirudhan, Water Air Soil Pollut. 150 (2003) 193–217. (2005) 14–21.
[19] N. Graham, X.G. Chen, S. Jayaseelan, Water Sci. Technol. 43 (2001) 245– [31] S.J. Allen, G. McKay, K.Y.H. Khader, Environ. Pollut. 56 (1989) 39–50.
252. [32] V.K. Gupta, I. Ali, Water Res. 35 (2001) 33–40.

Вам также может понравиться