Sie sind auf Seite 1von 6

Available online at www.sciencedirect.

com

Microporous and Mesoporous Materials 112 (2008) 419–424


www.elsevier.com/locate/micromeso

Adsorption behavior of methylene blue on halloysite nanotubes


Mingfei Zhao, Peng Liu *

State Key Laboratory of Applied Organic Chemistry and Institute of Polymer Science and Engineering,
College of Chemistry and Chemical Engineering, Lanzhou University, Lanzhou, Gansu 730000, China

Received 22 August 2007; received in revised form 9 October 2007; accepted 11 October 2007
Available online 18 October 2007

Abstract

The halloysite nanotubes (HNTs) were used as nano-adsorbents for the removal of the cationic dye, methylene blue (MB), from aque-
ous solutions. The dye adsorption experiments were carried out by using bath procedure. Experimental results have shown that the basic
pH, increasing initial dye concentration and lower temperature favored the adsorption. The dye adsorption equilibrium was rapidly
attained after 30 min of contact time. The factors controlling the adsorption process were also calculated and discussed. And a maximum
adsorption capacity of 84.32 mg/g of methylene blue was achieved. It was noted that the dye adsorbed halloysite nanotubes had poor
stability in aqueous suspension and deposited completely within 30 min while the original aqueous suspension of halloysite nanotubes
remained stable for months.
Ó 2007 Elsevier Inc. All rights reserved.

Keywords: Adsorption; Removal; Methylene blue; Halloysite nanotubes; Nano-adsorbent

1. Introduction Many kinds of adsorbents such as activated carbon [2],


silica [3], natural polymeric materials [4] and sewage sludge
Cationic dyes are widely used in industries such as [5], etc have been developed for various applications.
textiles, pulp mills, leather, dye synthesis, printing, food, However, their operating costs are high. This has led to
and plastics, etc. Since many organic dyestuffs are harmful searches for unconventional adsorbents as alternative
to human being and toxic to microorganisms, removal of adsorbents. The adsorbents with high surface areas, such
dyestuffs from wastewater has received considerable as mesoporous material [6–17], microporous materials
attention over the past decades. The dyes include a broad [18,19], carbon nanotubes [20,21] and titania nanotubes
spectrum of different chemical structures, primarily based [22], had been applied to decrease the dosage of adsorbent.
on the substituted aromatic groups. Due to the complex Most recently, the clay minerals were reported to be
chemical structure of these dyes, they are resistant to unconventional adsorbents for the removal of dyes from
breakdown by chemical, physical and biological treat- aqueous solutions due to their cheap and abundant
ments. Furthermore, any degradation by physical, chemi- resources, higher surface areas [23]. Furthermore, the regen-
cal or biological treatments may produce small amount eration of these low-cost substitutes is not necessary whereas
of toxic and carcinogenic products. Adsorption is known regeneration of activated carbon is essential because of the
to be a promising technique, which has great importance abundant resources. Clay materials with sheet-like struc-
due to the ease of operation and comparable low cost of tures [24–27] and needle-like structure [28–31] have been
application in the decoloration process [1]. increasingly gaining attention because they are cheaper than
activated carbons and their also provide highly specific sur-
face area.
*
Corresponding author. Tel.: +86 931 8912516; fax: +86 931 8912582. Halloysite nanotubes (HNTs) is a kind of aluminosili-
E-mail address: pliu@lzu.edu.cn (P. Liu). cate clays with hollow nanotubular structure mined from

1387-1811/$ - see front matter Ó 2007 Elsevier Inc. All rights reserved.
doi:10.1016/j.micromeso.2007.10.018
420 M. Zhao, P. Liu / Microporous and Mesoporous Materials 112 (2008) 419–424

natural deposits in countries such as China, America, Bra- 319.86; maximum wavelength: 662 nm) supplied by
zil, France and so on. It possesses a regular nanotubular Merck, was not purified prior to use.
morphology, bulk structure and rich mesopores and
nanopores [32,33]. Recently, it was used as adsorbents 2.2. Methylene blue absorption
[34] and nanotemplates or nanoscale reaction vessels
instead of carbon nanotubes or boron nitride nanotubes All of the methylene blue solution was prepared with
[35–37]. distilled water. The pH of the solution was adjusted with
In the present study, halloysite nanotubes (HNTs) were 0.1 N HCl or 0.1 N NaOH by using a Model 3C Digital
used as adsorbents for the removal of cationic dye from an pH-meter with a combined pH electrode. The pH-meter
aqueous solution. Methylene blue was selected as a model was standardized with NBS buffers before every
compound in order to evaluate the capacity of HNTs for measurement.
the removal of dyes from aqueous solutions. The aim of The effect of contact time on the amount of dye
this study was to investigate the adsorption of methylene adsorbed was investigated at 159.93 mg/l (0.50 m mol/l)
blue onto HNTs, which is a low-cost adsorbent for the initial concentration of dye and at different temperatures
removal of dyes. (20, 45 and 70 °C). Fifteen milliliter HNTs suspension
(including 0.235 g HNTs) was mixed with 85 ml MB dye
2. Experimental and methods solution.
For the other adsorption experiments, 15 ml HNTs sus-
2.1. Raw materials pension (including 0.235 g HNTs) was added into 85 ml of
dye solution with known initial concentration at desired
Halloysite clay was obtained from Hebei Province, pH and temperature and stirred with a rate of 150 rpm
China. It was pretreated by the followed procedure: for 180 min. A constant bath was used to keep the temper-
250 g raw halloysite clay and 500 ml water were mixed ature constant. At the end of the adsorption period, the
and milled with SF400 type wit sand-grinding dispersing solution was centrifuged for 10 min at 15,000 rpm. After
machine at 4000 rpm for 2 h. The halloysite nanotubes centrifugation, the dye concentration in the supernatant
(HNTs) suspension was centrifugated (3000 rpm for solution was analyzed using a UV spectrophotometer
3 min) to throw away the deposit. The stable suspension (Shimadzu UV-260) by monitoring the absorbance
was used for further experiments. Its TEM image was changes at a wavelength of maximum absorbance
shown in Fig. 1 (The morphology of the HNTs was char- (662 nm). The samples were pipetted from the medium
acterized with a JEM-1200 EX/S transmission electron reaction by the aid of a very thin point micropipette,
microscope (TEM). The bare HNTs suspended colloid which prevented the transition to the solution of the
solution and was deposited on a copper grid covered with HNTs samples. Preliminary experiments showed that the
a perforated carbon film). Methylene blue (MB) (CI: effect of the separation time on the amount of adsorbed
52015; chemical formula: C16H18ClN3S; molecular weight: dye was negligible.
The amounts of dye adsorbed on HNTs at any time, t,
were calculated from the concentrations in solutions before
and after adsorption. At any times, the amount of MB
adsorbed (mol/g) (qt), onto HNTs was calculated from
the mass balance equation as follows:
qt ¼ V ðC 0 –C e Þ=W ð1Þ

where qt is the amount of adsorbed dye on HNTs at any


time (m mol/g); C0 and Ce are the initial and equilibrium
liquid-phase concentrations of MB (m mol/l), respectively;
V is the volume of MB solution, and W is the mass of
HNTs sample used (g) [29,38].

2.3. Sedimentation time

The sedimentation time of the methylene blue adsorbed


halloysite nanotubes suspensions were investigated by the
following method: 15 ml HNTs suspension sample (includ-
ing 0.235 g HNTs) was mixed with 35 ml water or 35 ml
159.93 mg/l MB aqueous solution in a 50 ml measuring cyl-
inder, respectively. Then the transparence of the suspen-
Fig. 1. TEM image of halloysite nanotubes. sions was observed.
M. Zhao, P. Liu / Microporous and Mesoporous Materials 112 (2008) 419–424 421

3. Results and discussion 66.8


66.4
3.1. Adsorption rate
66.0
pH=4
3.1.1. Effect of contact time and initial dye concentration 65.6 pH=7
The effect of concentration on contact time was also pH=10

qt (mg/g)
65.2
investigated as a function of initial dye concentration. The
64.8
effect of initial dye concentration and contact time on the
removal rate of MB by HNTs is shown in Fig. 2. As shown, 64.4
the adsorption increases with increasing initial dye concen- 64.0
tration. The results show that dye uptake is rapid for the 63.6
first 15 min and finally attains saturation within about
63.2
30 min. The equilibrium was attained at 30 min.
0 30 60 90 120 150 180
The amount of MB adsorbed at equilibrium increases
t (min)
from 0.1276 to 0.2588 m mol/g (40.81–82.78 mg/g) by
increasing the initial MB concentration from 0.30 to Fig. 3. Effect of contact time and initial pH on the removal rate of MB
0.70 m mol/l with the adsorption condition of initial pH onto HNTs from aqueous solutions (20 °C, C0: 159.93 mg/l).
7 and 20 °C. It is near to the sepiolite and much higher than
those other clay minerals [28]. that the adsorption equilibrium was attained faster with the
higher initial pH condition.
3.1.2. Effect of pH
Effect of pH on the removal rate of MB by HNTs is 3.1.3. Effect of temperature
shown in Fig. 3. As the pH increased, the removal rate Fig. 4 exhibits contact time versus adsorbed amount
increased. The pH value of the dye solution plays an graph at different temperatures. The initial pH was selected
important role in the whole adsorption process and partic- as 7 in order to avoid the impossible breakage of HNTs at
ularly on the adsorption capacity. The zeta-potential higher temperatures because SiO2 is the main component.
behavior of halloysite nanotubes is mostly negative at pH The equilibrium adsorption capacity of MB onto HNTs
2–12 due to the surface potential of SiO2 with a small con- was found to decrease with increasing temperature, decreas-
tribution from the positive Al2O3 inner surface [36,39,40] ing from 0.2081 m mol/g at 20 °C to 0.2058 m mol/g at 70 °C
(the chemical properties on the halloysite nanotubes outer- indicating that the dye adsorption on the adsorbent was
most surface are similar to the properties of SiO2 while the favored at lower temperatures. And the adsorption equilib-
properties of the inner cylinder core could be associated rium was attained slowly with the higher temperature.
with Al2O3). And it exhibits a strongly pH-dependent sur-
face charge. With the increasing of pH, the surface of 3.2. Adsorption kinetics
HNTs becomes more negatively charged, thereby increas-
ing electrostatic attractions between positively charged 3.2.1. The pseudo-first-order kinetic model
dye anions and negatively charged adsorption sites and The pseudo-first-order kinetic model has been widely
causing an increase in the dye adsorption. It was also found used to predict dye adsorption kinetics. A linear form of
pseudo-first-order model was described by Lagergren [41]:

90
66.9
84
78 66.6

72 66.3
66
qt (mg/g)

66.0
qt (mg/g)

60
95.96mg/L
65.7
54 159.93mg/L
223.90mg/L 65.4
48 o
20 C
42 65.1
o
45 C
o
36 70 C
64.8
30
0 30 60 90 120 150 180 0 30 60 90 120 150 180
t (min ) t (min)
Fig. 2. Effect of contact time and initial dye concentration on adsorption Fig. 4. Effect of contact time and temperature on the removal rate of MB
of MB onto HNTs from aqueous solutions (20 °C, pH 7). onto HNTs from aqueous solutions (C0: 159.93 mg/l, pH 7).
422 M. Zhao, P. Liu / Microporous and Mesoporous Materials 112 (2008) 419–424

logðqeq  qt Þ ¼ log qeq  k pf t=2:303 ð2Þ 5


where qt is the amount adsorbed at time t (mg/g), and kpf is
the equilibrium rate constant of pseudo-first-order adsorp- 95.96mg/L
4
159.93mg/L
tion (min1). The values of log(qeqqt) were calculated 223.90mg/L
from the kinetic data (Fig. 4).

t/qt (min g/mg)


3
The calculated qeq, kpf, and the corresponding linear
regression correlation coefficient r21 values are shown in
Table 1. It was observed that the rate constant kpf 2
increased first with an increase in temperature and then
decreased. It was also observed that correlation coeffi-
1
cients were lower for all temperatures. This shows no
applicability of the pseudo-first-order model in predicting
the kinetics of the methylene blue adsorption onto HNTs 0
particles. 0 30 60 90 120 150 180
t (min)
3.2.2. The pseudo-second-order kinetic model Fig. 5. Ho pseudo-second-order kinetics for MB onto HNTs (20 °C, pH
The kinetic data were further analyzed using Ho’s 7).
pseudo-second-order kinetics, represented by [42]
3.2.3. Intraparticle diffusion
t=qt ¼ 1=ðk ps q2eq Þ þ t=qeq ð3Þ
The adsorbate species are most probably transported
where kps is the rate constant of second-order adsorption from the bulk of the solution into the solid-phase through
(g/mol min). The values were calculated from the kinetic an intraparticle diffusion process, which is often the rate-
data (Fig. 2). A plot between t/q versus t gives the value limiting step in many adsorption processes. The possiblity
of the constants k2 (g/mg h) and also qeq (mg/g) can be cal- of intraparticle diffusion was explored by using the intra-
culated. If the second-order kinetics is applicable, then the particle diffusion model [43]
plot of t/qt versus t should show a linear relationship. qt ¼ k id t1=2 þ C ð4Þ
The curves of the plots of t/qt versus t were given in
Fig. 5 and the calculated qeq, kps, and the corresponding where C is the intercept and kid is the intraparticle diffusion
linear regression correlation coefficient r22 values are sum- rate constant (mol/g min1/2). The values of qt were found to
marized in Table 1. The linear plots of t/qt versus t show be linearly correlated with the values of t (Fig. 2). Plots be-
good agreement between experimental and calculated qeq tween t/q versus t1/2 were given in Fig. 6. The values kid, C,
values. The correlation coefficients for the second-order and the corresponding linear regression correlation coeffi-
kinetics model (r22 ) are greater than 0.999, indicating the cient r23 values are given in Table 1.
applicability of this kinetics equation and the second-order The intraparticle rate constants calculated from Fig. 6
nature of the adsorption process of methylene blue onto are 0.037, 0.064, and 0.095 mg/g min1/2 at 20, 45, and
HNTs. 70 °C, respectively. From Table 1, it is observed that kid
increased with increasing temperature.

Table 1
Adsorption kinetic parameters of methylene blue onto HNTs 66.6
Pseudo-first-order T (°C)
66.3
20 45 70
kpf, min1 0.011 0.023 0.017 66.0
qeb, mg/g 66.55 66.46 65.82
r21 0.7912 0.8753 0.9976
q t (mg/g)

65.7
Pseudo-second-order C0 (mg/l)
95.96 159.93 223.90 65.4 o
20 C
qea, mg/g 40.82 66.53 84.32 o
45 C
kps, g/mg min 3.97 0.195 0.019 o
qeb, mg/g 40.82 66.55 86.35 65.1 70 C

r22 0.9999 0.9999 0.9999


64.8
Intraparticle diffusion T (°C)
kid, mg/g min1/2 0.037 0.063 0.095 0 3 6 9 12 15
C 66.09 65.77 64.62 1/2 1/2
r23 0.8424 0.8404 0.9914 t (min )
a
Calculated. Fig. 6. Intraparticle diffusion plots for MB onto HNTs (C0: 159.93 mg/l,
b
Experimental. pH 7).
M. Zhao, P. Liu / Microporous and Mesoporous Materials 112 (2008) 419–424 423

and temperature of the solution. Adsorption kinetic fol-


lows pseudo-second-order kinetic model confirming the
chemosorption of methylene blue (MB) on HNTs. The
methylene blue adsorbed halloysite nanotubes aggregated
and deposited completely within 30 min. It is expected that
the halloysite nanotubes could be used as low-cost uncon-
ventional nano-adsorbent for the removal of dyes from
aqueous solutions.

References

[1] E. Forgacs, T. Cserhati, G. Oros, Environ. Int. 30 (2004) 953–971.


[2] R.P. Manuel Fernando, F.S. Samanta, J.M.O. Jose, L.F. Jose,
Carbon 41 (2003) 811–821.
[3] S.K. Parida, S. Dash, S. Patel, B.M. Mishra, Adv. Colloid Interf. Sci.
121 (2006) 77–110.
[4] G. Crini, Prog. Polym. Sci. 30 (2005) 38–70.
[5] S. Rio, C. Faur-Brasquet, L. Le Coq, P. Le Cloirec, Environ. Sci.
Technol. 39 (2005) 4249–4257.
[6] H. Tamai, T. Yoshida, M. Sasaki, H. Yasuda, Carbon 37 (1999) 983–
989.
[7] Z. Hu, M.P. Srinivasan, Y. Ni, Carbon 39 (2001) 877–886.
[8] W. Tanthapanichakoon, P. Ariyadejwanich, P. Japthong, K. Nakag-
awa, S.R. Mukai, H. Tamon, Water Res. 39 (2005) 1347–1353.
[9] S.I. Kim, T. Yamamoto, A. Endo, T. Ohmori, M. Nakaiwa,
Micropor. Mesopor. Mater. 96 (2006) 191–196.
[10] K.Y. Ha, G. McKay, K.L. Yeung, Langmuir 19 (2003) 3019–
Fig. 7. The photos of the HNTs suspensions after 30 min: (a) in water and 3024.
(b) in methylene blue aqueous solution. [11] Z. Yan, S.Y. Tao, J.X. Yin, G.T. Li, J. Mater. Chem. 16 (2006) 2347–
2353.
[12] N. Andersson, P.C.A. Alberius, J.S. Pedersen, L. Bergstrom, Micro-
3.3. Sedimentation time por. Mesopor. Mater. 72 (2004) 175–183.
[13] S.B. Wang, H.T. Li, Micropor. Mesopor. Mater. 97 (2006) 21–
26.
The clays in nano-scale usually have excellent dispersi- [14] A. Tarafdar, P. Pranamik, Micropor. Mesopor. Mater. 91 (2006)
bility in aqueous-phase and the suspensions had good sta- 221–224.
bility because of the hydrophilic surface of the natural clay [15] D.A. Kron, B.T. Holland, R. Wipson, C. Maleke, A. Stein, Langmuir
minerals. The clay nano-adsorbents were separated from 15 (1999) 8300–8308.
the aqueous solutions via filtration or centrifugation pro- [16] L.C. Juang, C.C. Wang, C.K. Lee, Chemosphere 64 (2006) 1920–
1928.
cesses after the adsorption of dye in those reported works. [17] S.B. Wang, H.T. Li, L.Y. Xu, J. Colloid Interf. Sci. 295 (2006) 71–78.
In this work, it was interested that the methylene blue [18] F.C. Wu, R.L. Tseng, R.S. Juang, Sep. Purif. Technol. 47 (2005) 10–19.
adsorbed halloysite nanotubes aggregated together to mm [19] A.V. Maffei, P.M. Budd, N.B. McKeown, Langmuir 22 (2006) 4225–
scale particles and deposited completely within 30 min 4229.
while the initial HNTs suspension with the same solid con- [20] B. Fugestu, S. Satoh, T. Shiba, T. Mizutani, Y.B. Lin, N. Terui, Y.
Nodasaka, K. Sasa, K. Shimizu, T. Akasaka, M. Shindoh, K.I.
tent remains stable in months (Fig. 7). It might be due to Shibata, A. Yokoyama, M. Mori, K. Tanaka, Y. Sato, K. Tohji, S.
the hydrophobic surface of the methylene blue adsorbed Tanaka, N. Nishi, F. Watari, Environ. Sci. Technol. 38 (2004) 6890–
halloysite nanotubes. It is expected that the filtration or 6896.
centrifugation processes could be past over so the cost of [21] Y.M. Yan, M.N. Zhang, K.P. Gong, L. Su, Z.X. Guo, L.Q. Mao,
the separation could be saved. It is also an advantage of Chem. Mater. 17 (2005) 3457–3463.
[22] C.K. Lee, K.S. Lin, C.F. Wu, M.D. Lyu, C.C. Lo, J. Hazard. Mater.,
the application of the natural halloysite nanotubes as in press.
low-cost unconventional nano-adsorbent. [23] P. Liu, L.X. Zhang, Sep. Purif. Technol. 58 (2007) 32–39.
[24] F. Lopez Arbeloa, V. Martinez Martinez, J. Banuelos Prieto, I. Lopez
4. Conclusion Arbeloa, Langmuir 18 (2002) 2658–2664.
[25] J. Orthman, H.Y. Zhu, G.Q. Lu, Sep. Purif. Technol. 31 (2003) 53–59.
[26] H. Li, B.J. Teppen, C.T. Johnston, S.A. Boyd, Environ. Sci. Technol.
Hylloysite nanotubes (HNTs) have been proved to be an 38 (2004) 5433–5442.
effective nano-adsorbent for the removal of cationic dye via [27] S.S. Tahir, N. Rauf, Chemosphere 63 (2006) 1842–1848.
adsorption from aqueous solution. The equilibrium [28] Y. Ozdemir, M. Dogan, M. Alkan, Micropor. Mesopor. Mater. 96
(2006) 419–427.
adsorption was reached within 30 min. A maximum
[29] M. Alkan, O. Demirbas, M. Dogan, Micropor. Mesopor. Mater. 101
adsorption capacity of 84.32 mg/g of methylene blue was (2007) 388–396.
achieved. It is much higher than those other clay minerals. [30] P. Liu, J.S. Guo, Colloids Surf. A Physicochem. Eng. Asp. 282–283
The adsorption is highly dependent on concentration, pH (2006) 498–503.
424 M. Zhao, P. Liu / Microporous and Mesoporous Materials 112 (2008) 419–424

[31] J.H. Huang, Y.F. Liu, Q.Z. Jin, X.G. Wang, J. Yang, J. Hazard. [38] M. Dogan, M. Alkan, Chemosphere 50 (2003) 517–528.
Mater. 143 (2007) 541–548. [39] G. Tari, I. Bobos, C.S. Gomes, J.M. Ferreira, J. Colloid Interf. Sci.
[32] S.R. Levis, P.B. Deasy, Int. J. Pharm. 243 (2002) 125–134. 210 (1999) 360–366.
[33] X.C. Liu, X.Y. Chuan, A.P. Wang, F.Y. Kang, Acta Geologica [40] S. Baral, S. Brandow, B.P. Gaber, Chem. Mater. 6 (1994) 1227–1232.
Sinica 80 (2006) 278–284. [41] V.C. Taty-Costodes, H. Fauduet, C. Porte, A. Delacroixs, J. Hazard
[34] M.T. Viseras, C. Aguzzi, P. Cerezo, C. Viseras, C. Valenzuela, Mater. 105 (2003) 121–142.
Micropor. Mesopor. Mater., in press, doi:10.1016/j.micromeso. [42] V. Vadivelan, K.V. Kumar, J. Colloids. Interf. Sci. 286 (2005) 90–100.
2007.03.033. [43] M. Dogan, M. Alkan, A. Turkyılmaz, Y. Ozdemir, J. Hazard Mater.
[35] S.J. Antill, Aust. J. Chem. 56 (2003) 723. B 109 (2004) 141–148.
[36] D.G. Shchukin, G.B. Sukhorukov, R.R. Price, Y.M. Lvov, Small 1
(2005) 510–513.
[37] A.P. Wang, F.Y. Kang, Z.H. Huang, Z.C. Guo, X.Y. Chuan,
Micropor. Mesopor. Mater., in press, doi:10.1016/j.micromeso.
2007.04.021.

Das könnte Ihnen auch gefallen