Sie sind auf Seite 1von 6

Polymer 45 (2004) 6285–6290

www.elsevier.com/locate/polymer

Chitosan functionalized with 2[-bis-(pyridylmethyl)


aminomethyl]4-methyl-6-formyl-phenol: equilibrium and kinetics of
copper (II) adsorption
Karin C. Justia, Mauro C.M. Laranjeiraa, Ademir Nevesa, Antônio S. Mangrichb,
Valfredo T. Fáverea,*
a
Departamento de Quı́mica, Universidade Federal de Santa Catarina, Florianópolis, SC 88040-900, Brazil
b
Departamento de Quı́mica, Universidade Federal do Paraná, Curitiba, PR 81531-970, Brazil

Received 13 April 2004; received in revised form 2 July 2004; accepted 7 July 2004
Available online 24 July 2004

Abstract
The complexation agent 2[-bis-(pyridylmethyl) aminomethyl]-4-methyl-6-formyl-phenol (HL) was immobilized in chitosan in order to
obtain a new adsorbent material to be employed in studies on adsorption and pre-concentration of Cu(II). The chitosan modified by the
complexation agent was characterized by infrared spectroscopy, DSC and TGA. The studies were conducted as a function of the pH of the
medium and the mechanism of Cu(II) adsorption in the solid phase was analyzed utilizing several kinetic models. The parameters for
the adsorption of Cu(II) ions by chitosan–HL were determined with a Langmuir isotherm, the maximum saturation capacity of the monolayer
being 109.4 mg of Cu(II) per gram of polymer. Electron paramagnetic resonance spectroscopy revealed that CuII ions coordinate to the donor
atoms of the HL ligand anchored to the surface of the polymer forming a stable chelate complex in the solid state.
q 2004 Elsevier Ltd. All rights reserved.

Keywords: Chitosan; Adsorption; Copper (II)

1. Introduction based on copper [5]. Its toxicity to humans occurs even at


low concentration. According to the US Environmental
The increase in the concentration of toxic metals in Protection Agency the maximum permitted concentration of
aquatic environments has been attributed principally to copper in potable water is 1.3 mg lK1 [6].
industrial and agricultural activities. Many of these metal The solid phase extraction (SPE) technique, commonly
cations are toxic even at very low concentration and should known as adsorption, has been widely used for the
therefore be removed for the remediation of waters [1–5]. separation and pre-concentration of a variety of metal
The conventional techniques most often used for the ions, and to remove color, odor and organic pollutants. The
removal of toxic metals from liquid industrial effluents are great advantage of this technique over others is the lower
chemical precipitation, filtration, reverse osmosis, electro- generation of residues, easy recuperation of metals and the
chemistry and ion exchange. Most of these techniques have possibilities for the reutilization of the adsorbent [7,8]. This
a high cost and produce a residual sludge containing metals, technique is normally based on chelate matrices with a high
which is difficult to treat [4,5]. capacity for adsorption and selectivity. These materials form
The principal sources of copper contamination are metal chelates with the metal ions in solid state and are potentially
cleaning, galvanoplastic baths, fertilizers and insecticides more selective than conventional ion exchange resins [8].
Technologies using a biomass as bioadsorbents have
* Corresponding author. Tel.: C55-483-319-230; fax: C55-483-319-
attracted great interest from researchers for the removal of
711. metals from effluents due to their organic nature and
E-mail address: favere@qmc.ufsc.br (V.T. Fávere). biodegradability [9].
0032-3861/$ - see front matter q 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.polymer.2004.07.009
6286 K.C. Justi et al. / Polymer 45 (2004) 6285–6290

Chitin, a component of the exoskeleton of crustaceans temperature range of 25–500 8C, with a heating rate of
such as crabs, lobsters, shrimps, etc. is a biomass used in the 10 8C minK1, under a nitrogen atmosphere.
synthesis of chitosan [10]. This biopolymer has excellent Electron paramagnetic resonance (EPR) spectra were
properties for the adsorption of transition metals, principally recorded at X-band on a Bruker ESP300E spectrometer at
due to the presence of amino groups in the polymer matrix. room and liquid N2 temperatures.
However, its capacity to adsorb is dependent on the degree
of deacetylation, the nature of the metal ion and the pH of
the solution [4,5,11–13]. The presence of a high percentage 2.3. Chitosan–HL synthesis
of reactive primary amino groups distributed in the polymer
matrix, offers innumerous chemical modifications, such as The chitosan was dispersed in methanol in a ratio of
the immobilization of complexation agents. The modifi- 1.5 mol of HL ligand per mol of glucosamine and the system
cation of the polymer surface through the introduction of was maintained under agitation for 16 h and in reflux for
new complexation groups may result in the formation of 18 h. The desired product was obtained after the reaction
different chelates, so increasing its adsorption capacity and had an intense yellow color. The chitosan–HL was initially
selectivity toward metal ions in solution [9,14–16]. washed with methanol to remove the excess ligand,
The aim of this study was to prepare a new adsorbent followed by distilled water and then dried in a vacuum.
material, covalently immobilizing in the chitosan, via the Fig. 1 shows the structure of the Schiff base formed by the
Schiff’s base reaction, the complexation agent 2[-bis- chitosan and the HL ligand as an intermediate stage in the
(pyridylmethyl) aminomethyl]-4-methyl-6-formyl-phenol synthesis of the adsorbent used in this study. To increase
(HL), and to study the adsorption of Cu(II) ions in an the resistance of the adsorbent to acid solutions, the
aqueous medium. chitosan–HL was cross-linked with a 2.5% glutaraldehyde
solution for 24 h at room temperature [20,21]. The cross-
linked material was filtered and thoroughly washed with
2. Experimental distilled water. After reticulation, the imine groups were
reduced with a solution of sodium borohydride [22].
2.1. Materials
2.4. Adsorption experiments
The chitosan with a 90% degree of deacetylation was
obtained from Purifarma (Brazil). The 2[-bis-(pyridyl-
All batch experiments were carried out in a thermostated
methyl) aminomethyl]-4-methyl-6-formyl-phenol (HL)
bath at 25 8C, with a shaking of 200 rpm.
ligand was prepared through the reaction of 2-chloro-
The study of the pH dependency of the adsorption of
methyl-4-methyl-6-(formyl)phenol with bis-(2-pyridyl-
Cu(II) by chitosan–HL was carried out utilizing buffer
methyl)amine [17,18].
solutions: KCl/HCl (pHZ3), acetic acid/sodium acetate
A 1000 mg lK1 copper (II) stock solution was prepared
(pHZ4, 5 and 6), tris(hydroxymethyl) aminomethane
by dissolving the appropriate quantity of copper sulphate
(pHZ8) and ammonia/ammonium chloride (pHZ9 and
analytical grade and standardized with a standard solution of
0.01 mol lK1 EDTA [19]. Working standard solutions of the 10). Aliquots (50.0 ml) of the 100 mg lK1 Cu(II) solutions
Cu(II) metal ion were prepared through the dilution of the buffered at different pH values, were placed in contact
1000 mg lK1 stock solution with distilled water. with 100 mg samples of chitosan–HL for a period of
All other reagents utilized were of analytical grade. 2 h. The quantity of Cu(II) adsorbed was obtained from
the difference between the initial and the final
2.2. Instrumentation concentration of metal in solution and the mass of
adsorbent used.
An atomic absorption spectrophotometer, (FAAS) Spec- The adsorption kinetics were determined in a closed flask
tra AA 50 Varian, equipped with a flame atomizer and a containing 100 mg of chitosan–HL and 100.0 ml of Cu(II)
metal hollow cathode lamp, was used for the copper solution buffered at pH 5.5. After pre-determined time
determination. periods, 200 ml aliquots were removed and the concen-
The batch adsorption experiments were carried out in a tration of copper ions in solution was measured.
thermostated Lab-Line Shaker Bath with controlled In the adsorption equilibrium experiments, 100 mg
agitation. samples of chitosan–HL and 50.0 ml of solutions buffered
Infrared spectra were obtained from KBr pellets, using a at pH 5.5 containing several concentrations of Cu(II) were
spectrophotometer with Fourier Transform System, 2000 used. The system was maintained under shaking at 25 8C
FT-IR-Perkin Elmer. until adsorption equilibrium was reached. Aliquots (200 ml)
DSC and TG analyses were carried out with a DSC-50 were removed and diluted in volumetric flasks and the metal
Shimadzu and a TG-50 Shimadzu, respectively. In the case concentration was determined by atomic absorption spec-
of the DSC, the scanning was carried out within a trophotometry (FAAS) at 324.8 nm.
K.C. Justi et al. / Polymer 45 (2004) 6285–6290 6287

Fig. 1. Diagram showing the reaction of chitosan with the HL ligand.

3. Results and discussion During the degradation of chitosan–HL a 27.0% loss of


material mass at 334.11 8C occurs, this degradation
3.1. Chitosan–HL characterization temperature being higher than that for pure chitosan
which indicates that the modification rendered the polymer
The characterization of the new adsorbent material was thermally stable. This increased thermal stability of the
carried out through infrared, DSC, TGA and EPR spectra. chitosan polymer containing the HL ligand can be attributed
In the IR spectra for the chitosan there were specific to the presence of an imine group along with the phenolic
vibrational bands at 1660 cmK1 which can be assigned to aromatic ring, resulting in a rigid structure. Similar thermal
the CaO absorption of the –NH–CaO structure, 1599 cmK1 stability was also observed for the aromatic polymer
to the NH2 deformation, 1075 cm K1 to the –C–O fluorine-containing polybenzoxazoles in a rigid structure
stretching of primary alcohols and 3373 cmK1 to the –OH when compared to the polymer containing aliphatic
stretching absorption band. Through comparison of the hexafluoroisopropylidene group [24].
spectra, the appearance of bands related to the HL ligand
could be seen in the spectra for the modified chitosan. The
bands located at 763 and 862 correspond to the angular
deformation of C–H, that at 1269 cmK1 is attributed to C–O
stretching of the phenol of the ligand and those at 1434,
1469 and 1595 cmK1 are attributed to the CaC stretching
and are features of the aromatic pyridine rings; and the band
which appears at 1151 cmK1 is associated with the
stretching of the C–N bond of the tertiary aliphatic amine.
The absence of a band at 1680 cmK1, attributed to the
carbonyl of the HL ligand, along with the new intense band
at 1631 cmK1 relating to the stretching of the imine CaN
bond formed, proves the chemical modification of chitosan
[23].
In Fig. 2 the DSC and TGA thermograms for the
chitosan, the HL ligand and the chitosan–HL can be
observed. The DSC thermogram for chitosan shows a
wide endothermic peak at around 103 8C and an exothermic
peak at 305.40 8C. The endothermic peak originates from
the water contained in the chitosan, that is, it indicates a
process of dehydration, while the exothermic peak can be
attributed to the thermal decomposition of the polymer. In
the thermogram for HL an sharp endothermic peak can be
seen at 133.95 8C, which is probably related to the fusion of
the material. The chitosan–HL shows a wide exothermic
peak around 306.57 8C. A small change in the thermal
behavior of the polymer was observed when the ligand is
immobilized in the matrix [4,24].
The TGA curve for chitosan shows thermal degradation
at 326.20 8C, with a 66.7% loss in polymer mass. The HL
ligand undergoes a mass loss in two stages, at 279.42 and
417.35 8C, with losses of 41.6 and 34.9%, respectively. Fig. 2. DSC (a) and TG (b) thermograms for chitosan, HL and chitosan–HL.
6288 K.C. Justi et al. / Polymer 45 (2004) 6285–6290

3.2. pH dependency of the Cu(II) adsorption by the


chitosan–HL

Fig. 3 shows the effect of pH on the metal ion adsorption


by chitosan–HL. The adsorption of the Cu(II) ions by the
chitosan–HL increases with the pH of the solution until a
maximum value of 6.0 and then decreases with an increase
in pH. At acidic pH a decrease in the adsorption is attributed
to the increase in ionic strength of the solution and not to the
protonation of complexation sites as was observed in
chitosan, because the protonation pH of pyridyl groups in
the chitosan–HL is very low (pKas/2); they are strong
acids, as in the ligand (HL) [5,18]. At alkaline pH, the
solutions were buffered with auxillary complexation agents
in order to avoid the precipitation of copper hydroxide,
Fig. 4. Kinetics of the adsorption of Cu(II) by chitosan–HL.
however, buffers can form complexes with the metal ion
inhibiting, as a consequence, the adsorption of the ion by the kinetics data was the pseudo second-order model:
adsorbent.
t 1 1
Z 2
C t
qt k 2 q e qe
3.3. Adsorption kinetics
where qt is the quantity adsorbed in time t, qe, the quantity
Fig. 4 shows the kinetics of the adsorption of Cu(II) by adsorbed at equilibrium and k2 is the adsorption rate
chitosan–HL. The kinetic curve showed that the adsorption constant.
is rapid in the first minutes and reached equilibrium after The parameters of linear regression calculated employ-
approximately 200 min and remained constant for 48 h. In ing several models of the adsorption kinetic and the linear
order to evaluate the mechanism of the adsorption kinetics curves obtained gave the following equations:
several models were tested to interpret the experimental (Pseudo first-order) YZ1.57517K0.0078X (RZ
data [12]. A good correlation of the kinetics data explains 0.98039)
the adsorption mechanism. Thus, the angular coefficient of (Pseudo second-order) YZ0.20255K0.01832X (RZ
each kinetics equation was chosen to determine the 0.99997)
mechanism of the metal ion adsorption in the solid phase. (Intraparticle diffusion) YZ18.95375K2.14935X (RZ
The pseudo first-order, pseudo second-order and intrapar- 0.89507)
ticle diffusion models were tested in this study. The validity
of these models can be interpreted by the linear plots of Kinetic studies on sorption of Cu(II) ions by chitosan by
(qeKqt) vs. T, (t/q) vs. t and qt vs. t1/2, respectively [25]. The using the pseudo second-order rate expression provided the
equation, which gave the best fit for the experimental best fitting kinetic model [2]. The mechanism indicated that
the rate of Cu(II) adsorption by chitosan–HL depends on the
quantity of ions at the adsorbent surface and the quantity of
these adsorbed at equilibrium. The rate constant determined
was 1.7!10K3 mg (g min)K1.
The pseudo first-order and intrarparticle diffusion models
gave poor slopes and were therefore not suitable for the
interpretation of the adsorption mechanism within the
concentration range under study.

3.4. Adsorption isotherms

In the batch experiments, the isotherms were used to


interpret the adsorption data. For the system under study, it
was the Langmuir model which gave the best fit for the
experimental data of the Cu(II) adsorption by chitosan–HL
at pH 5.5. Fig. 5 shows the quantity of Cu(II) ions adsorbed
at the surface of the adsorbent and the concentration of
Fig. 3. Effect of pH on Cu(II) adsorption by chitosan–HL. Cu(II) ions in the aqueous phase at equilibrium. This
K.C. Justi et al. / Polymer 45 (2004) 6285–6290 6289

of Cu(II) per gram of chitosan–HL (1.72 mmol gK1) and a


Langmuir constant (K) of 2.68!10K2 l mgK1. This result
was higher than that obtained by Nghah et al. for the
capacity of Cu(II) adsorption by chitosan cross-linked with
glutaraldehyde (59.67 mg gK1). Therefore, the result
obtained for chitosan–HL indicates that chemical modifi-
cation on chitosan improved the adsorption capacity for
copper ions of the cross-linked polymer.

3.5. Mechanism of Cu(II) complexation in the solid phase

In order to interpret the mechanism of modified


chitosan–HL complexation with CuII, EPR studies on the
resulting material were carried out. In the solid state, at
room temperature temperature, the association complex
with CuII ions exhibit an axial EPR spectrum (Fig. 7) with
Fig. 5. Adsorption isotherm of Cu(II) on chitosan–HL. gsOgtO2.076 and AsZ185!10K4 cmK1, which is
suggestive of a tetragonally distorted octahedral, square-
relationship showed that the adsorption capacity increases
pyramidal or square-planar geometry having a d2x K d2y
with the equilibrium concentration of the metal ion in the
ground state (similar parameters gsZ2.18, gtZ2.076,
aqueous phase, progressively reaching Cu(II) saturation at
AtZ25!104 and AsZ185!10K4 cmK1 were obtained at
higher concentrations. The isotherm revealed that the new
liquid nitrogen temperature, 77 K) [26,27]. A comparison of
material has a good capacity to adsorb Cu(II). From the
these parameters with those observed in the spectrum of an
analysis of the isotherm, the values for the quantities
ethanol frozen solution of [CuIIHLCl2] not anchored to
adsorbed, the equilibrium concentration, the mathematical
chitosan, (gsZ2.26, gtZ2.065 and AsZ180!10K4 cmK1),
equation of isotherm was placed in a linear form more
reveals that when [CuIIHL] is bound to chitosan the gs value
suitable of determination of the adsorption parameters:
decreases while the corresponding As value increases and such
maximum adsorption capacity (qmax.) and Langmuir con-
a result is most probably associated with a tetrahedral
stant (K). The linearization of the equation is also shown in
distortion of a square planar geometry as well as with distinct
Fig. 6. The linear curve obtained resulted in the following
equatorial coordination environments around the CuII centers
equation:
[28]. It is important to note that an octahedral structure with a
Y Z 0:34061 C 0:00914ðR Z 0:99356Þ strong Jahn–Teller distortion in the Cl–Cu–Ophenol bond
direction has recently been determined for [CuIIHLCl2] [29]
This equation suggests that the Langmuir model interprets and that in this structure, the chloro ligands can be easily
the experimental adsorption data reasonably well. The replaced by stronger ligands like H2O, OH and amine
angular coefficient of this equation corresponds to the nitrogens [29]. Therefore, the EPR data clearly demonstrate
1/qmax. and the intercept represents the 1/qmax.K. The value that there is a significant interaction between chitosan–HL
for the maximum adsorption capacity (qmax.) was 109.4 mg and CuII ions and that the coordination geometry and/or
coordination environment of the CuII in the [CuHL]
complex is at least to some extent affected when the
complex is bound to the polymer chitosan.

4. Conclusions

The characterization studies showed that the HL ligand


was immobilized at the chitosan surface, resulting in a new
material. The results show that the CuII ion adsorption
process is dependent on pH and that the kinetics mechanism
follows a pseudo second-order model, where a chemical
reaction at the surface of the solid phase occurs. The
isotherm which best fits the adsorption data was the
Langmuir monolayer model and EPR studies in the solid
state revealed that coordination of the tripodal HL ligand,
anchored to the surface of the polymer, to CuII ions takes
Fig. 6. Linearization of adsorption isotherm. place with the formation of a stable chelate complex.
6290 K.C. Justi et al. / Polymer 45 (2004) 6285–6290

Fig. 7. X-band EPR spectrum for chitosan–HL with CuII in the solid state. (—) Experimental spectrum and (.) simulated spectrum using the Winepr SimFonia
Program. The Hamiltonian parameters are gsZ2.26, gtZ2.076, AsZ175!10K4 cmK1 and AtZ20 cmK1.

Finally, this new material shows significant capacity for [10] Kumar MNVR. React Funct Polym 2000;46:1–27.
CuII ion adsorption and can therefore be used as an [11] Hsien Y, Rorrer GL. Ind Eng Chem Res 1997;36:3631–8.
[12] Wu F-C, Tseng R-L, Juang R-S. J Hazard Mater 2000;B73:63–5.
alternative adsorbing agent in static and dynamic separation
[13] Kawamura Y, Mitsuhashi M, Tanibe H, Yoshida H. Ind Eng Chem
processes, pre-concentration and extraction of Cu(II) from Res 1993;32:386–91.
aqueous solutions. [14] Rodrigues CA, Laranjeira MCM, Fávere VT, Stadler E. Polymer
1998;39:5121–6.
[15] Inoue K, Yoshizuka K, Ohto K. Anal Chim Acta 1999;388:209–18.
Acknowledgements [16] Kondo K, Nakagawa S-I, Matsumoto M, Yamashita T, Furukawa I.
J Chem Eng Jpn 1997;30(846):851.
[17] Uozumi S, Furutachi H, Ohba M, Okawa H, Fenton DE, Shindo K,
The authors wish to thank CNPq for its financial support. Murata S, Kitko DJ. Inorg Chem 1998;37:6281–7.
This work was supported by grants from CNPq, PADCT and [18] Karsten P, Neves A, Bortoluzzi AJ, Lanznaster M, Drago V. Inorg
PRONEX. Chem 2002;41:4624–6.
[19] Schwarzenbach G, Flaschka H. Complexometric titrations. London:
Methuen and Co Ltd; 1969 p. 256–57.
References [20] Monteiro Jr OAC, Airoldi C. Int J Biol Macromol 1999;26:119–28.
[21] Hall LD, Yalpani M. Carbohydr Res 1980;83:C5–C7.
[22] Lu ZH, Bhongle N, Su X, Ribe S, Senanayake CH. Tetrahedron Lett
[1] Babel S, Kurniawan TA. J Hazard Mater 2003;B97:219–43.
2002;43:8617–20.
[2] Sag Y, Aktay Y. Biochem Eng J 2002;12:143–53.
[23] Silverstein RM, Bassler GC, Morrill TC. Spectrometric identification
[3] Akama Y, Yamada K, Itoth O. Anal Chim Acta 2003;485:19–24.
of organic compounds, 5th ed. New York: Wiley; 1991. Chapter 3.
[4] Ng JCY, Cheung WH, McKay G. J Colloid Interf Sci 2002;255:
[24] Lee JK, Kim JH, Kim YJ. Bull Kor Chem Soc 2003;24:1029–31.
64–74.
[5] Ngah WSW, Endud CS, Mayanar R. React Funct Polym 2002;50: [25] Wu F-C, Tseng R-L, Juang R-S. Water Res 2001;35:613–8.
181–90. [26] Hathaway BJ, Billing DE. Coord Chem Rev 1970;5:143–207.
[6] http://www.epa.gov/safewater/mcl.html, accessed at February 2004. [27] Orton JW. Electron paramagnetic resonance. London: Life Books
[7] Pesavento M, Baldini E. Anal Chim Acta 1999;389:59–68. LTD; 1968.
[8] Gurnani V, Singh AK, Venkataramani B. Anal Chim Acta 2003;485: [28] Sakaguchi U, Addison AW. J Chem Soc, Dalton Trans 1979;600–8.
221–32. [29] Oliveira MCB, Scarpellini M, Neves A, Terenzi H, Bortoluzzi, AJ,
[9] Lee S-T, Mi F-L, Shen Y-J, Shyu S-S. Polymer 2001;42:1879–92. Scarpellini M, Neves A. 2004. Personal communication.

Das könnte Ihnen auch gefallen