Vous êtes sur la page 1sur 11

Hydrometallurgy 64 (2002) 89 99 www.elsevier.

com/locate/hydromet

Potential of protonated alginate beads for heavy metals uptake


n ez a,*, Yoshiaki Umetsu b Juan Patricio Iba
b a Department of Metallurgy, Arturo Prat University, Av. Arturo Prat 2120, Casilla 121, Iquique, Chile Institute for Advanced Materials Processing, Tohoku University, Katahira 2-1-1, Aoba-ku, Sendai 980-8577, Japan

Received 22 May 2001; received in revised form 11 January 2002; accepted 6 February 2002

Abstract The potential for uptake of several heavy metal ions by alginate in the form of protonated beads was investigated at 25 jC. The morphological characteristics of the beads and their behavior in aqueous solutions were examined as well. The ability of protonated alginate beads to remove heavy metal ions from dilute aqueous solutions was demonstrated. The uptake of trivalent chromium, copper, zinc, nickel and cobalt was found to be about 75, 77, 46, 43 and 35 mg per g of beads, respectively. The uptake increased with solution pH and acid concentration for protonation, and with decreasing ionic strength. The cross-linking agents, Ba and Ca, were not released during uptake, and protons were liberated instead. Therefore, the uptake was associated with ion exchange between protons of the free carboxylic functional groups of the alginate and metal ions from the solutions. EPMA-EDX analyses showed that heavy metals were uniformly distributed throughout the surface (external and internal) of the beads, indicating that protonated dry alginate beads can be considered as porous ion exchangers. Observation of protonated dry alginate beads by scanning electron microscopy (SEM) showed a corrugated surface having a uniform distribution of regular nodules and cavities. A mean diameter of about 1.0 mm was obtained for these beads. They were found to have chemical and structural stability in acidic and slightly acidic environments. Bead forming reaction with Ba or Ca makes it possible to use alginate as sorbent for heavy metal ions in extremely dilute aqueous solutions. D 2002 Elsevier Science B.V. All rights reserved.
Keywords: Alginite beads; Protonation; Heavy metal ions

1. Introduction Treatment of industrial effluents having less than about 0.1 g/L of dissolved metal ions by conventional processes presents several disadvantages, such as high operating cost (Jansson-Charrier et al., 1995), low efficiency, and generation of solid wastes often toxic, that may require safety disposal protocols or even further treatment. The capacity of biological

Corresponding author. Tel.: +56-57-394-415; fax: +56-57394-152. n ez). E-mail address: jpibanez@cec.unap.cl (J.P. Iba

entities to take up heavy metal ions from dilute aqueous solutions is well documented (Volesky, 1990), and it can be considered as an alternative to be used under conditions where conventional processes lose competitiveness. The immobilization of those entities in polymeric matrices such as alginate creates biosorbent materials with the right size, mechanical strength, rigidity, and porosity (Veglio and Beolchini, 1997). Furthermore, this yield beads or granules that can be stripped of metals, reactivated, and reused in a similar manner to ion exchange resins and activated carbons (Ashley and Roach, 1990). This implies a significant minimization of wastes (solid and liquid) to discharge.

0304-386X/02/$ - see front matter D 2002 Elsevier Science B.V. All rights reserved. PII: S 0 3 0 4 - 3 8 6 X ( 0 2 ) 0 0 0 1 2 - 9

90

n J.P. Iba ez, Y. Umetsu / Hydrometallurgy 64 (2002) 8999

Alginate is often used for immobilization of biological entities (Gilson and Thomas, 1995; Martinse et al., 1989; Cheethan et al., 1979). Cheethan et al. (1979) reported physical properties appropriate for columnar applications of calcium alginates when used as a cell immobilization support. Moreover, the important contribution of alginate of certain biological entities, such as algae, to the uptake of heavy metal ions has been reported (Schiewer et al., 1995; Fourest and Volesky, 1996; Holan et al., 1993; Kuyucak and Volesky, 1989a,b). Although these facts are known, there are few studies on the potentiality of the alginate per se as sorbent for the removal of heavy metals from low concentration aqueous solutions (Chen et al., 1997; Chen and Yiacoumi, 1997; Romero-Gonzalez et al., 2001). The aim of this work is to study the heavy metal sorption potentiality of the alginate in the form of beads. This paper will report the uptake of chromium, copper, zinc, nickel and cobalt from dilute solutions by using protonated alginate beads prepared with barium and calcium ions. The influence of acid concentration for protonation, ionic strength and pH on the uptake is reported as well. Morphological characteristics and behavior in aqueous solution of these beads are also discussed.

2.1. Preparation of the alginate beads Based on several series of experiments, the simplest preparation protocol of alginate having appropriate physical stability was established as follows: 0.1 L of a 2 mass % of viscous alginate solution was added dropwise into 0.5 L of 50 mmol/L Ba- or Cachloride solution under gentle stirring at 25 jC. Ba or Ca cross-linked alginate first was formed as a thin film on the surface of the alginate solution droplet, holding its spherical shape. Perfect spheres of gel alginate were obtained after 2 h. The excess of cross-linker solution was removed by washing several times. Then the uncross-linked functional groups of the alginate were protonated with 1.0 L of HNO3 1 mol/L. Finally, the protonated beads were repeatedly washed until neutrality. Protonated dry alginate beads were obtained by exposing them to the air at room temperature for several days. The lost of weight during drying was found to be 95.3%, which was accompanied with a diameter reduction of around 64.3%. The consumption of Ba and Ca during preparation of beads was 140.9 and 480.3 mg per g of alginate produced, respectively, i.e., around 3.5 mmol of alkali earth metal per gram of alginate. Although the dry alginate beads had rough surface and distorted shape resulting from the drying process, these beads were used in the main course of the experimental work. This selection was based on the advantages of dry beads over the gel beads; accurate measurement of the mass to be used, better handling nature, gain in mechanical strength and rigidity. Furthermore, dry alginate beads showed more favorable uptake of metal ions than the gel type. This may be in tight connection with generation of open pores and channels by removal of liquid from the gel beads, which facilitates the contact between H-form functional groups inside the beads and metal ions. 2.2. Behavior of alginate beads in aqueous solution Alginate beads were contacted with solutions of hydrochloric and nitric acids, and sodium hydroxide to study their behavior in aqueous solution. Twenty mg of Ba- and Ca-alginate beads were suspended in 0.1 L of those solutions at several pH values ranging from 1.6 to 8.7 for 3 h at 25.0 F 0.1 jC. The experimental set-up

2. Experimental Chemicals were of analytical reagent grade and were used without further purification. Stock and test solutions of chromium, zinc, copper, nickel and cobalt were prepared, respectively, from Cr(NO3)39H2O, ZnSO 4 7H 2 O, CuSO 4 5H 2 O, NiSO 4 6H 2 O and CoSO47H2O. Solutions were prepared with doubly distilled water. Sodium alginate with a viscosity range of 0.10 0.15 Ns/m2 for 1 w/v% was used. Solutions of the cross-linking agents, barium and calcium, were prepared from their chloride salts. Chemical analyses of solutions were made by an inductively coupled plasma atomic emission spectrometer, ICP-AES. Examination of the beads by scanning electron microscopy (SEM) was made after coating them with a thin layer of gold. The distribution of the target metal ions inside the metal-loaded beads was determined with an electron probe X-ray microanalyzer in the mode of energy-dispersive X-ray spectroscopy, EPMA-EDX.

n J.P. Iba ez, Y. Umetsu / Hydrometallurgy 64 (2002) 8999

91

consisted in a cylindrical round-bottom glass reactor of 0.1 L capacity placed in a thermostatically controlled water bath. The stirring of the system was provided by a rotor having an overhead Teflon coated mixing wing. After the operating temperature was reached, alginate beads were placed, at time zero, into solutions having the desired value of pH. During these experiments, the pH of the solution was kept constant by the addition of sodium hydroxide solution with a pH-stat unit. After contact, beads were separated from the solution by filtrating with cellulose acetate membrane filters (0.45 Am of pore diameter). An aliquot of 10 mL of the filtrate was analyzed by the ICP-AES technique to determine the content of Ba or Ca released from the beads. Alginate beads were then washed several times with doubly distilled water and air dried for several days at room temperature and stored for SEM observations. 2.3. Uptake of heavy metal ions by alginate beads Twenty milligrams of alginate beads were contacted with 0.1 L of metal ion-bearing solutions of known concentrations at different pH values for a fixed period of time in Erlenmeyer flasks of 0.2 L capacity placed in a thermostatically controlled shaker water bath at 25 F 1 jC and 120 stroke/min. The pH of the solutions was adjusted to the desired value before the addition of alginate beads and was measured at the end of the experiments; this last value was recorded as the equilibrium pH. For the study of the effect of ionic strength on the uptake, potassium nitrate was used to adjust it to 10 and 100 mmol/L. At the end of each run, the metal-loaded beads were separated from the solutions by filtration. The filtrates were analyzed by the ICP-AES technique for the remaining metal ions and cross-linker released from the beads. The beads were washed and air-dried for several days at room temperature and stored for EPMA-EDX analysis.

shows that gel beads have a smooth surface in spherical shape. The circle observed at the center of the beads and the thin film surrounding them correspond to reflection of the incident light used for observation and to doubly distilled water used to avoid drying during observation, respectively. Fig. 1B shows a scanning electron micrograph of a protonated dry alginate bead, which indicates that beads have a corrugated surface. The mean diameter of gel and dried beads was found to be around 2.8 and 1.0 mm, respectively; both types of beads showed a narrow size distribution. No morphological differences between the beads prepared with Ba or Ca ions, in both gel and dried states, were found. This indicates that the ion used to cross-link the alginic acid chains contribute in the same manner to the formation of beads. It seems that the specific sites where the cross-linking ions are located allow accommodation of ions of different ionic radii. This is a direct consequence of the buckled twofold conformation of the alginate chains. Discussions on formation of alginate gel from a structural point of view can be found in Smidsrd and Haug (1968, 1972), Morris et al. (1978), Rees and Welsh (1977), Grant et al. (1973) and Atkins et al. (1973). A schematic representation of the alginates structure is given in Fig. 2, which shows the egg-box junction of this configuration. It should be noted that this particular structure of the alginates leaves free carboxylic functional groups in the H-form, which may interact with metal ions. Since there are no morphological differences between Ba or Ca beads, hereafter the term protonated alginate beads refers to those prepared using barium unless otherwise specified. Fig. 1C E shows the scanning electron micrographs of higher magnification of the protonated alginate bead surface shown in Fig. 1B. The entire surface is made up of homogeneously distributed and corrugated nodules having cavities between them. This regular distribution of nodules and cavities may be considered as a macroscopic manifestation of the egg-box junction structure mentioned above. 3.2. Behavior of alginate beads in aqueous solution

3. Results and discussion 3.1. Morphological characteristics of alginate beads An optical micrograph of the protonated gel alginate beads right before drying is presented in Fig. 1A; it When alginate beads were contacted with HCl solutions, there was a slight release of Ba and Ca of 0.3 and 1.4 Amol, respectively. This release did not affect

92

n J.P. Iba ez, Y. Umetsu / Hydrometallurgy 64 (2002) 8999

Fig. 1. Morphology of gel (A) and dry (B to E) alginate beads. For details, see text.

the structure of the beads at pH 3.0 and lower values. Nevertheless, at pH 4.5 the structure did collapse, generating a viscous phase of higher density. On the other hand, the use of nitric acid did not significantly affect the structure of the beads at the pH where hydrochloric

acid did. These results suggest that the presence of chloride ions, in slightly acidic environments, resulted in the breakdown of the rigid structure of the beads. By studying the solubility of alginate in acidic environments, Haug and Larsen (1963) reported that

n J.P. Iba ez, Y. Umetsu / Hydrometallurgy 64 (2002) 8999

93

Fig. 2. Schematic representation of the cross-linking of alginic acid chains by Ba ions. The section is showing details of the egg-box junction, oxygen atoms coordinated to Ba are shown as filled circles. The structure leaves free carboxylic functional groups after formation of the alginate beads (modified from Rees and Welsh, 1977; Grant et al., 1973).

solutions of hydrochloric acid at pH values higher than 3.0 decreased the viscosity of alginates from different sources. Decreasing the pH lower than 3.0 resulted in precipitation of alginates. Since solutions at pH lower than 3.0 will result in precipitation of alginate, the structure of the beads is favored at low pH. On the other hand, the contact of the beads with HCl solutions at pH values higher than 3.0 produced a decrease of the alginate viscosity that resulted in the deterioration of the structure of the beads. To clarify the effect of chloride ions on the stability of beads, uptake experiments at pH 4.5 were carried out with solutions containing 25 mg/L of Zn (as chloride). There was no release of barium from the

alginate beads and the residual concentration of zinc decreased to around 6.0 mg/L in 5 h; the structure of the beads was not affected. Therefore, the structure of the beads was not affected by the chloride ions in the presence of zinc ions. In the uptake, a divalent metal cation such as Zn2 + will require two unoccupied carboxylic groups to interact with. This interaction should generate bonds connecting different chains of alginate; the result is further reinforcement of the structure of the beads. It should be noted that metal ions do not occupy the positions where the crosslinker are located (there was not release of Ba). Hence, the metal ions only interact with the free carboxylic functional groups present in the structure

94

n J.P. Iba ez, Y. Umetsu / Hydrometallurgy 64 (2002) 8999

of the alginate after formation of beads. Therefore, formation of alginate beads and metal ion uptake are due to different mechanisms. When alginate beads were in contact with NaOH solutions, the bead structure collapsed and there was a significant consumption of OH ions, up to 181 Amol for Ca alginate. The destruction of the beads leads to the generation of a viscous solution. Although the viscosity was not measured, it was evident that this solution was less viscous than that obtained at lower pH (experiments with HCl at pH 4.5). The OH ion is known to be a nucleophilic anion, which may produce elimination of the hydrogen bound to the h-carbon. This nucleophilic attack will result in the destruction of the alginate structure and generation of H2O. Therefore, it should be accompanied with a consumption of OH ions as the experimental results showed. Working with alginate, Haug et al. (1963, 1967) reported significant degradation in concentrated alkaline solutions (pH > 10) occurring by an OH helimination reaction, which leads to the formation of unsaturated uronic acid derivatives. According to our results, this h-elimination reaction is also taking place in slightly alkaline solutions (pH around 8). The amount of Ba and Ca ions found in the filtrates was less than 1.0 and 1.5 Amol, respectively, at the pH values investigated and independent of the final state of the beads. This may support a consideration that the destroyed chains of alginate retain Ba and Ca in the polymeric structure. This view is supported by the fact that 20 mg of beads will contain around 70 Amol of the cross-linker, then if the primary structure of the alginate is destroyed, the amount of those ions in the filtrates should be several times higher than the values found. SEM observation of unaffected beads did not show any morphological change in the surface from those shown in Fig. 1B E. This indicates that there was no observable deterioration of the beads even at pH as low as 1.6. Therefore, protonated dry alginate beads are suitable to be used in acidic and slightly acidic aqueous solutions. 3.3. Uptake of heavy metal ions by alginate beads 3.3.1. Effect of protonation The effect of HNO3 concentration used for protonation of alginate beads on the uptake of trivalent chromium is presented in Fig. 3 at pH 5.0 for 24 h of

Fig. 3. Effect of nitric acid for protonation on the uptake of chromium.

contacting time. The uptake was determined by the Cr mass balance in solution. This figure shows that the Cr uptake was significantly enhanced by the protonation treatment of the beads. By using beads treated with HNO3 1.0 mol/L, the uptake was almost doubled in comparison with the values obtained for beads without protonation. Furthermore, the concentration of the nitric acid solution used for protonation affected the performance of alginate on the uptake of metal ions as shown in Fig. 3. An increase of around 25 mg/ g in the uptake was observed by raising the concentration of nitric acid up to 0.01 mol/L. This was followed by a slight increase of the uptake up to around 60 mg/g when the concentration of acid was further raised up to 1 mol/L. The increase of the acid concentration for protonation of alginate beads will result in a more effective generation of carboxylic groups having labile protons suitable for exchange with metal ions. There was no further increase in the uptake by increasing the acid concentration from 0.1 to 1.0 mol/L. These results demonstrate the advantage of working with protonated alginate beads having the carboxylic functional group in the H-form when used for the uptake of metal ions. Therefore, this kind of bead was used in further experiments. It should be noted that the oxidizing properties of the nitric acid did not damage the alginate beads within the range studied. The carboxylic group is a highly oxidation resistant one, but the oxidizing agent (NO3 ) may attack the polymeric chains; if there is such an attack, the alginate beads will be seriously

n J.P. Iba ez, Y. Umetsu / Hydrometallurgy 64 (2002) 8999

95

damaged. Since there was no detectable deterioration of the beads, there was no effect of the oxidizing properties of the nitric acid during protonation. This suggests, then, that the complete structure of the alginate bead is oxidation resistant. 3.3.2. Effect of ionic strength Alginate is a highly charged polymer. The negative charge of the ionized functional group of the alginate ( COO ) will make the concentration of anions in the proximity of the polymer lower than the bulk solution. This difference should decrease by increasing the ionic strength of the solution. Then, the ionic strength may negatively affect the uptake process by compressing the electrical double layer surrounding the beads. A further consequence may be the competition between the cationic components of the salt used to adjust ionic strength and target metal ions by ionized functional groups, which may decrease the uptake. The effect of the solutions ionic strength on the uptake of chromic ions after 110 h is shown in Fig. 4; each experiment was conducted in duplicate. The ionic strength was about 1.0 mmol/L when the concentration of trivalent chromium was around 20 mg/L (0.35 mmol/L) in solutions of pH 4.5. Hence, KNO3 solution was not used in those experiments. In the other cases, KNO3 was used to adjust ionic strength to values of 10 and 100 mmol/L. This figure shows that increase in ionic strength decreased the uptake of Cr. This was coupled with an increase of the variation between equilibrium and initial pH, which suggests that protons of alginates free functional group are

Fig. 4. Effect of ionic strength on the uptake of chromium.

being exchanged by K + ions. Since the concentration of this ion in the bulk is high, the capacity of the formed-COOK to exchange K for Cr is lowered at higher ionic strengths. Accordingly, the uptake should decrease as shown in Fig. 4. Similar tendencies have been reported in the technical literature. Chen et al. (1997) and Chen and Yiacoumi (1997) reported that the increase of ionic strength, adjusted using NaClO4, of Cu-bearing solutions resulted in a significant decrease in the removal of copper with gel calcium alginate beads. Furthermore, the authors reported a release of Ca higher than 613 Amol/L when 2 mL of 1.5 wt.% gel beads were contacted with 0.1 L of 0.1 mmol/L Cu (ionic strength of 50 mmol/L) for 4.8 h at pH 4.5. Based on this work, they indicate that the decrease in the removal of copper at higher ionic strength was due to a competition for functional groups between copper and ionic strength components. However, there was no information concerning the final state of the beads. Experiments with dry Ba alginate showed a slight release of the structural Ba from beads, which was less than 0.4 Amol/L at ionic strength values of 10 and 100 mmol/L. The beads after more than 30 h of contact time, became swollen and lost robustness, and finally, after 110 h, the beads were degraded to soft fragments. These results demonstrate that high ionic strengths deteriorate the rigid structure of beads in the course of long time contacting. It has been demonstrated in previous sections that the oxidizing character of the NO3 anion does not affect the alginate beads. Hence, the high concentration of K + seems to produce the small release of Ba2 + observed. If monovalent cations are able to replace some divalent cations of the alginate structure, beads will be deteriorated. Since there was a slight release of barium, the alginate beads were partially degraded into the form of soft fragments. This indicates that the egg-box junction of the alginate was not completely degraded. Even though the beads were deteriorated, there was a significant uptake of trivalent chromium on the fragments of the alginate beads remaining. These results demonstrate the potentiality of the alginate as sorbent material of heavy metals even in adverse conditions. These results suggest that the decrease of the uptake observed by increasing the ionic strength is

96

n J.P. Iba ez, Y. Umetsu / Hydrometallurgy 64 (2002) 8999

occurring by two separated factors. The first one is a competition between the cationic component of the ionic strength reagent and the target metal ions, and a second factor concerns the physical stability of the sorbent material. Therefore, the combination of these factors will reduce the capacity of alginate beads to take up chromic species from low-concentration solutions. 3.3.3. Effect of pH The effect of equilibrium pH on the uptake of trivalent chromium after 110 h of contact time is presented in Fig. 5. The resulting ionic strengths in these Cr-bearing solutions were around 1 mmol/L. The uptake increased from about 41 to 70 mg/g when the equilibrium pH rose from 2.95 to 3.16. The uptake increased by around 6% when pH further rose up to 3.46, and then almost no increase ( < 2%) at higher pH values. The lower uptake observed at low equilibrium pH is due to the high concentration of H + in the bulk solution, which suppresses the dissociation of the free functional group present in the alginate beads structure. Furthermore, the predominant species of trivalent chromium in dilute solution is Cr3 + from pH values lower than around 3.0. This chromic ion requires three unoccupied carboxylic groups to be fixed by the alginate. Therefore, taking into account these two factors, the uptake of chromium ions decreased by lowering the pH as shown in Fig. 5. It should be indicated that Ba was not released from the beads and that a liberation of H + occurred during the uptake; equilibrium pH was always lower

Fig. 6. Uptake of divalent metal ions by protonated alginate beads at several values of pH. For details, see text.

than the initial one. This decrease in pH and the lack of Ba release strongly suggest that uptake may take place by an ion exchange process between protons of the free carboxylic functional groups of the alginate beads and chromic species of the solution. The present authors, by kinetic experiments under constant values of pH, demonstrated a stoichiometric ion exchange mechanism between H + of protonated alginate beads and the different species of trivalent chromium in n ez and Umetsu, 1997, solution during uptake (Iba 2002). The rigid structure of the beads was not affected during the uptake. SEM observations of the metal ionloaded beads showed the same corrugated surface presented in Fig. 1B E. 3.3.4. Uptake of divalent metal ions Fig. 6 summarizes the uptake of Cu, Zn, Ni and Co by protonated alginate beads from solutions having around 0.4 mmol/L metal ion at several values of pH after 24 h of contact time. The uptake of all metal ions increased with the equilibrium pH. The uptake of Cu was several times higher than that for the other metal ions, especially for equilibrium pH values over 3.1. Cu uptake increased by a factor of 2.4 when the pH rose from 2.7 to 3.1. A further increase of pH to 3.3 yielded around 6% increase of uptake. Similar behavior was observed for other metal ions. Uptake of Zn and Ni showed identical behavior with pH, an increase by a factor of around 3 in uptake was reached by increasing pH from 2.8 to 3.4. The Co uptake was the lowest, however the tendency and increase factor in uptake with pH was similar to that observed for Zn and Ni. As mentioned above, at low values of pH there is a high

Fig. 5. Effect of equilibrium pH on the uptake of chromium.

n J.P. Iba ez, Y. Umetsu / Hydrometallurgy 64 (2002) 8999

97

concentration of protons in solution, which suppress the uptake capacity of the alginate beads. Therefore, the heavy metal ion uptake decreases at lower values of pH as shown in Fig. 6, independent of the metal ion

considered. This figure also clearly shows an affinity between free carboxylic groups of protonated alginate beads and heavy metal ions. The order of affinity is CuHZn = Ni>Co, which agrees with the tendency

Fig. 7. EPMA-EDX micrographs of beads loaded with Cu (A), Zn (B), Ni (C) and Co (D). The distribution pattern of metals was measured along the lines crossing the beads.

98

n J.P. Iba ez, Y. Umetsu / Hydrometallurgy 64 (2002) 8999

found when the properties of alginate solutions have been investigated (Haug and Smidsrd, 1965). The equilibrium pH of the solutions was in all experiments lower than the initial one, which demonstrated a release of protons during the uptake of these heavy metals. As in the case of trivalent chromium, there was no Ba in the final solutions. These results suggest, once more, that uptake may be occurring by an ion exchange mechanism between protons of the free carboxylic groups and heavy metal ions. This behavior was independent of the heavy metal considered. The rigid and corrugated surface of protonated alginate beads, as in the case of trivalent chromium, was not affected during the uptake process. SEM observations of the metal-loaded beads showed the same corrugated surface as presented in Fig. 1B E. EPMA-EDX analyses of the cross-section of the metal-loaded beads after experiments at the highest pH values are shown in Fig. 7A, B, C and D for Cu, Zn, Ni and Co, respectively. These micrographs show the distribution pattern of the metal ions taken along the lines crossing the beads. A uniform distribution of the metal ions throughout the beads was observed. This distribution demonstrates the formation of open pores and channels during the drying process of the alginate beads, which allow the penetration of solution into the inner part of the beads during experiments. Furthermore, the homogeneous distribution of the free carboxylic functional groups on the entire structure of the alginate beads is demonstrated as well. The EPMAEDX results, then, demonstrate that different metal ions are able to reach the functional groups of the entire structure of protonated alginate beads. The results presented in Figs. 3 7 clearly demonstrate that after formation of the alginate beads, the particular structure of this polymer has free carboxylic groups, i.e., nonstructural groups, where heavy metals such as Cr, Cu, Zn Ni and Co can be sorbed.

uptake. Ba- and Ca-alginate beads have identical morphology and similar behavior in aqueous solution. There was no release of the cross-linker ions from the beads during the uptake of heavy metals, which was furthermore coupled with a release of protons. This strongly suggests an ion exchange mechanism occurring between protons of the non-structural carboxylic functional groups of the alginate beads and metal ions from the solutions. Furthermore, these findings indicate that the mechanisms of beads formation and of metal uptake are different in nature.

Acknowledgements One of the authors, J.P.I., thanks the Ministry of Education, Science and Culture of Japan for the financial support provided through the Monbusho Scholarship Program.

References
Ashley, N.V., Roach, D.J., 1990. Review of biotechnology applications to nuclear waste treatment. J. Chem. Tech. Biotechnol. 49, 381 394. Atkins, E.D.T., Nieduszynski, I.A., Mackie, W., Parker, K.D., Smolko, E.E., 1973. Structural components of alginic acid: II. The crystalline structure of poly-a-L-guluronic acid. Results of Xray diffraction and polarized infrared studies. Biopolymers 12, 1879 1887. Cheethan, P.S.J., Blunt, K.W., Bucke, C., 1979. Physical studies on cell immobilization using calcium alginate gels. Biotechnol. Bioeng. 21, 2155 2168. Chen, J., Yiacoumi, S., 1997. Biosorption of metal ions from aqueous solution. Sep. Sci. Technol. 32, 51 69. Chen, J., Tendeyong, F., Yiacoumi, S., 1997. Equilibrium and kinetic studies of copper ion uptake by calcium alginate. Environ. Sci. Technol. 31, 1433 1439. Fourest, E., Volesky, B., 1996. Contribution of sulfonate groups and alginate to heavy metal biosorption by the dry biomass of sargassum fluitantas. Environ. Sci. Technol. 30, 277 282. Gilson, C.D., Thomas, A., 1995. Calcium alginate bead manufacture: with and without immobilized yeast. Drop formation at a two-fluid nozzle. J. Chem. Tech. Biotechnol. 62, 227 232. Grant, G.T., Morris, E.R., Rees, D.A., Smith, P.J.C., Thom, D., 1973. Biological interactions between polysaccharides and divalent cations: the egg-box model. FEEBS Lett. 32, 195 198. Haug, A., Smidsrd, O., 1965. The effect of divalent metals on the properties of alginate solutions. Acta Chem. Scand. 19, 341 351.

4. Summary The ability of protonated alginate beads to take up different heavy metal ions from extremely dilute aqueous solutions was clearly demonstrated. Slightly acidic environments, acid concentration of 1 mol/L for protonation, and ionic strengths lower than 10 mmol/ L were found to be the most favorable conditions for

n J.P. Iba ez, Y. Umetsu / Hydrometallurgy 64 (2002) 8999 Haug, A., Larsen, B., 1963. The solubility of alginate at low pH. Acta Chem. Scand. 17, 1653 1662. Haug, A., Larsen, B., Smidsrd, O., 1963. The degradation of alginates at different pH values. Acta Chem. Scand. 17, 1466 1468. Haug, A., Larsen, B., Smidsrd, O., 1967. Alkaline degradation of alginate. Acta Chem. Scand. 21, 2859 2870. Holan, Z.R., Volesky, B., Prasetyo, I., 1993. Biosorption of cadmium by biomass of marine algae. Biotechnol. Bioeng. 41, 819 825. n ez, J.P., Umetsu, Y., 1997. Uptake of Cr (III) from aqueous Iba solution by calcium alginate beads. Fall Meeting of MMIJ, Sapporo, Sept., B V-2, 103. n ez, J.P., Umetsu, Y., 2002. Hydrometallurgy, To be submitted to. Iba Jansson-Charrier, M., Guibal, E., Surjous, R., Le Cloiree, P., 1995. Continuous removal of uranium by biosorption onto chitosan. In: Jerez, C.A., Vargas, T., Toledo, H., Wiertz, J.V. (Eds.), Biohydrometallurgical Processing, vol. II. University of Chile, Chile, pp. 267 276. Kuyucak, N., Volesky, B., 1989a. Accumulation of cobalt by marine algae. Biotechnol. Bioeng. 33, 809 814. Kuyucak, N., Volesky, B., 1989b. The mechanism of cobalt biosorption. Biotechnol. Bioeng. 33, 823 831. k-Brk, A., Smidsrd, O., 1989. Alginate as Martinse, A., Skija immobilizer material: I. Correlation between chemical and physical properties of alginate gel beads. Biotechnol. Bioeng. 33, 79 89.

99

Morris, E.R., Rees, D.A., Thom, D., Boyd, J., 1978. Chiroptical and stoichiometric evidence of a specific primary dimerisation process in alginate gelation. Carbohydr. Res. 66, 145 154. Rees, D.A., Welsh, E.J., 1977. Secondary and tertiary structure of polysaccharides in solutions and gels. Angew. Chem., Int. Ed. Engl. 16, 214 224. Romero-Gonzalez, M.E., Williams, C.J., Gardiner, P.H.E., 2001. Study of the mechanism of cadmium biosorption by dealginated seaweed waste. Environ. Sci. Technol. 35, 3025 3030. Schiewer, S., Fourest, E., Chong, K.H., Volesky, B., 1995. Ion exchange in biosorption by dried seaweed: experiments and model predictions. In: Jerez, C.A., Vargas, T., Toledo, H., Wiertz, J.V. (Eds.), Biohydrometallurgical Processing, vol. II. University of Chile, Chile, pp. 219 228. Smidsrd, O., Haug, A., 1968. Dependence upon uronic acid composition of some ion exchange properties of alginates. Acta Chem. Scand. 22, 1968 1997. Smidsrd, O., Haug, A., 1972. Dependence upon the gel sol state of the ion exchange properties of alginates. Acta Chem. Scand. 26, 2063 2074. Veglio, F., Beolchini, F., 1997. Removal of metals by biosorption: a review. Hydrometallurgy 44, 301 316. Volesky, B., 1990. In: Volesky, B. (Ed.), Biosorption of Heavy Metals. CRC Press, USA, pp. 3 5.

Vous aimerez peut-être aussi