Applicability of New Acrylic, Weakly Basic Anion Exchanger Purolite A

Apr 23, 2012 - According to the manufacture description Purolite A-830 is a sorbent of very high capacity (2.75 eq/dm3), which was the main reason for...
0 downloads 0 Views 4MB Size
Article pubs.acs.org/IECR

Applicability of New Acrylic, Weakly Basic Anion Exchanger Purolite A-830 of Very High Capacity in Removal of Palladium(II) Chlorocomplexes Anna Wołowicz* and Zbigniew Hubicki Department of Inorganic Chemistry, Faculty of Chemistry, Maria Curie-Skłodowska University, Maria Curie-Skłodowska Square 2, 20-031 Lublin, Poland S Supporting Information *

ABSTRACT: According to the manufacture description Purolite A-830 is a sorbent of very high capacity (2.75 eq/dm3), which was the main reason for examining its applicability in the removal of palladium(II) chloro-complexes from acidic solutions and comparison of its experimental capacity with other weakly basic anion exchangers. The influence of several parameters such as hydrochloric and nitric acids concentrations, competitive anions (effect of AlCl3 addition) and temperature increase were studied. Sorption isotherms were obtained and modeled using the Langmuir and the Freundlich model. The influences of agitation speed, palladium concentration, and particle size on sorption kinetics were considered. Kinetic curves were modeled using the pseudofirst- and pseudo-second-order kinetic equations. According to the evaluation using the Langmuir equation, the maximum sorption capacity obtained from the Langmuir isotherm was equal to 356.88 mg Pd(II)/g. The reduction of Pd(II) uptake was observed with the increasing concentration of chloride ions, temperature, and AlCl3 addition. Amberlite IRA-35 by elution with 0.05 dm3 of thiourea−ethyl alcohol solution was obtained. On the basis of the above-mentioned examples and literature data, among others, it can be stated that the anion exchange resins are characterized by higher selectivity than the cationic ones, but owing to strong precious metals binding, their elution is difficult and sometimes not quantitative. In our previous studies4−12 only two examples of acrylic, weakly basic anion exchange resin applications in Pd(II) removal can be found. These are Varion ADAM (macroporous) and Amberlyst A-24 (gel), but the capacities of both resins as given by the manufacturer were much lower that that for Purolite A-830. On the basis of the above-mentioned facts, this paper describes the application of the new, acrylic, weakly basic anion exchange resin Purolite A-830 in palladium(II) chloro-complexes removal from the chloride solutions (0.1−6.0 M HCl−100 mg Pd(II)/dm3; 0.1−0.9 M HCl−0.9−0.1 M HNO3−100 mg Pd(II)/dm3). The adsorption and desorption behaviors of Pd(II) on/from the WBA resin were surveyed in the column, batch systems, and only in the batch system, respectively. The effects of the following factors on Pd(II) sorption were considered: hydrochloric and nitric acids concentrations, palladium concentrations, agitation speed, beads size, AlCl3 addition, and temperature.

1. INTRODUCTION Palladium natural resources are still being depleted, therefore the recovery of palladium from different scrap materials is becoming more and more important. Hydrometallurgical methods are successfully used for this purpose. After the leaching process of scraps, precious metals usually occur in the form of anionic complexes, and their recovery using anion exchange resins of different basicity appears very promising. However, only a few examples concerning applicability of acrylic, weakly basic anion exchange resins can be found in literature.1−3 Kononova et al.2 applied acrylic, macroporous resin Purolite S 985 of the polyamine functional groups (pKa = 0.74) for palladium(II) removal from model nitrate solutions and from solutions obtained after the leaching of spent catalysts. This resin is characterized by good sorption properties; the recovery factor of Pd(II) from the initial Pd(II) solution (0.0005 M, pH 2) using Purolite S 985 is 94%. This fact was also confirmed in the case of real solutions where the the Pd(II) recovery factor was about 88%. Moreover, the authors assumed that Pd(II) recovery proceeds according not only to the anion exchange mechanism but also probably to the mixed one that is anion exchange and coordination mechanism. The elution of Pd(II) from Purolite S 985 using 1 M thiourea in 0.01 M HNO3 was not satisfactory (model solutions, 78%; solutions of spent catalyst, not exceeding 25%). Similar research was carried out by Matsubara et al.,3 concerning trace amounts of gold(III), palladium(II), and platinum(IV) metals separating from the base metals and their recovery using Amberlite IRA-35, also acrylic but of the gel structure. This resin seems to be more efficient for Pt(IV) (Kd = 5.2 × 104, 0.05 M HCl) than for Au(III) (Kd = 2.2 × 104, 0.05 M HCl) and Pd(II) (Kd = 2.7 × 104, 0.05 M HCl) removal. Quantitative recovery of each noble metal ion on © 2012 American Chemical Society

2. EXPERIMENTAL SECTION 2.1. Reagents Preparation. In all experiments in this study, all chemicals used were of analytical reagent grade. Received: Revised: Accepted: Published: 7223

December 7, 2011 March 21, 2012 April 22, 2012 April 23, 2012 dx.doi.org/10.1021/ie202859n | Ind. Eng. Chem. Res. 2012, 51, 7223−7230

Industrial & Engineering Chemistry Research

Article

Aqueous Stock and Working Solution Preparation. Aqueous palladium(II) stock solution (10 000 mg/dm3) was prepared from solid PdCl2 (>99%, POCh, Poland) by its dilution with concentrated hydrochloric acid, temperature adjustment, and microwaves. A proper amount of PdCl2 was weighed, then concentrated hydrochloric acid was added (to keep the H+ concentration at 1 M), and the mixture was left for 24 h. After this time the mixture was placed in ultrasonic washers with a thermoregulator set from 293 to 333 K (Inter Sonic, type IS-1 with a thermoregulator, Poland) and digested for 15 min at 333 K. This operation was repeated four times to obtain the total dissolution of Pd. Working solutions of 100 mg/dm3 were obtained by dilution of the stock solution with HCl and HNO3 to adjust the H+ concentration to the desired value. Concentrated HNO3 (65%) and HCl (37%) (POCh, Poland) were used throughout this study. In the case of the solution with the aluminum chloride addition (10% AlCl3−5% HCl−100 mg Pd(II)/dm3) a proper amount of AlCl3·6H2O was added to the solutions. Regeneration Solutions Preparation. Different eluting agents were prepared by dilution of concentrated hydrochloric acid (0.1−6.0 M HCl), nitric acid (0.1−4.0 M HNO3), and ammonia (0.5−2.0 M NH4OH). Standard Solutions Preparation. Standard solutions (up to 10 mg/dm3) were prepared by dilution of the standard stock solutions (1000 mg/dm3 in 0.1 M HNO3) with 2 M HCl, in flasks washed by 2 M HCl, distilled water, 2 M HNO3, and again distilled water. 2.2. Batch and Column Studies. Batch Studies. The effect of phase contact time (1−1440 min), initial concentration of Pd(II) solutions (50−300 mg/dm3), agitation speed (120−180 rpm), beads size, AlCl3 addition (10% HCl−5% AlCl3) and temperature (298−308 (±2) K) on the palladium(II) complexes removal as well as the efficiency of regeneration process of the anion exchanger and sorption−desorption cycles were investigated. In the experiments of kinetics, 0.5 g of the samples of Purolite A-830 was shaken mechanically in the conical flask (100 cm3) with 50 cm3 of Pd(II) solution at 298 ± 2 K using a thermostatic mechanical shaker, Elpin+, type 358S, (Lubawa, Poland). The procedure of equilibrium experiments is identical to that of kinetic tests, but the initial concentration of Pd(II) solution changes in the range from 100 to 4000 mg/dm3. In the sorption−desorption cycles the experimental conditions were similar to those in the kinetic studies. In all cases the Erlenmeyer flasks after Pd(II) sorption process were removed from the shaker, and the phases were separated using the filtration procedure. Then the Pd(II) concentrations in the liquid phase were determined. Column Studies. The breakthrough curves of Pd(II) were determined using 10 cm3 of the swollen anion exchangers packed in a column of internal diameter equal to 1 cm and a column length of 25 cm with a glass wool support at the end. The initial palladium(II) solution of 100 mg/dm3 concentration was passed through the column at a flow rate of 0.4 cm3/min. The concentration of Pd(II) in the effluent collected into the fractions of 100−500 cm3 volume was analyzed by AAS. On the basis of the above-described experimental techniques, the sorption, kinetic, and isotherm parameters were calculated, and they are presented in the Supporting Information, Table S1. 2.3. Apparatus and Analytical Procedure. The surface morphology of Purolite A-830 was studied using the scanning

electron and atomic force microscopes, BS: 301 (Telsa) and NanoScope V (Veeco, USA), respectively. The CHN analysis of the resin was carried out by using the Perkin-Elmer CHN 2400 analyzer and the Sartorius microbalance M2P. The concentration of palladium(II) ions after the sorption or desorption process was measured using the fast sequential atomic absorption spectrometer, Varian AA240FS, equipped with the appropriate hollow cathode lamps and SIPS autosampler. The analytical line used for atomic absorption measurements was 247.6 nm. An oxidizing air−acetylene flame was used for atomization of palladium(II), and the fuel was properly adjusted to maintain optimum atomization of the injected samples. The other parameters were as follows: lamp current, 10 mA; slight width, 0.2 nm; acetylene flow, 2 dm3/ min; air flow, 13.5 dm3/min.

3. RESULTS AND DISCUSSION 3.1. Purolite A-830 Characteristics. Purolite A-830 is a high capacity anion exchange resin with a polyacrylic matrix cross-linked with divinylbenzene. It has a macroporous structure with a total exchange capacity of 2.75 eq/dm3 (free base form, FB) and particle size range of 0.3−1.2 mm (1.2 mm max, 5%; uniformity coefficient max, 1.7). It is thermally stable up to 313 K (FB) or 373 K (Cl− form), and its operating pH limit is in the range from 0 to 9. Purolite A-830 occurs as white spherical beads. The moisture retention is equal to 47−53% (Cl− form). The high operating capacity of this resin is a result of the complex amine functionality. The percentage contents of carbon, hydrogen, and nitrogen are equal to 43.28%, 8.41%, and 16.88%, respectively. Moreover, Purolite A-830 matrix ensures excellent and fast removal of mineral and organic acids and the efficient desorption which occurs upon regeneration markedly reduces the risk of resin fouling. The tough resilient macroporous structure affords excellent mechanical strength and resistance to osmotic shock. Others properties of the anion exchanger Purolite A-830, for example, volume weight, water swelling coefficient, weak base exchange capacity, and chemical stability in aggressive media were determined.1 Supporting Information, Table S2 shows the values of the above-mentioned parameters. As can be seen, Purolite A-830 is characterized by good chemical stability in 1 and 5 M hydrochloric acid and sodium hydroxide solutions (no loss of the weak base exchange capacity is observed). On the other hand, hydrogen peroxide significantly decreases the volume and weight exchange capacities by about 51% or 68.5% in the case of 5% or 10% sodium hydroxide concentrations. The oxidizing agent can interact with the anion exchanger in a different way: (1) adsorption of H2O2 onto anion exchangers, (2) decomposition of the anion exchanger by hydrogen peroxide, (3) the catalytic decomposition of the hydrogen peroxide in contact with the anion exchanger. H2O2 adsorption results in water removal from the anion exchanger, whereas decomposition of exchanging groups of the anion exchanger is caused by H2O2 treatment or deamination. These effects are observed with the increasing concentration of sodium hydroxide. Sodium hypochlorite similar to hydrogen peroxide significantly diminishes the chemical stability of the anion exchanger. NaOCl and NaOH mixture has a high pH therefore the oxidizing agent cannot attack the macroreticular chain or ionogenic groups of the resin as in the aqueous solution containing only NaOCl and in consequence, the weak base anion exchange capacity losses are diminished. The presence of 7224

dx.doi.org/10.1021/ie202859n | Ind. Eng. Chem. Res. 2012, 51, 7223−7230

Industrial & Engineering Chemistry Research

Article

trace amounts of metal ions such as copper, iron, etc. which act as a catalyst of NaClO decomposition leads to much higher reduction of the weak base anion exchange capacity values. Additionally, degradation of the polymer skeleton of the anion exchanger, for example, by treatment using oxidizing agents leads to a decrease in cross-linking and in consequence, in increased swelling by water. The water swelling coefficient values for Purolite A-830 are high (0.77; 0.88 (5% H2O2); 0.83 (10% H2O2); 1.54 (13% NaOCl); 1.1 (6.5% NaOCl); and 1.11 g of water per g of dry sample (6.5% NaOCl, 5 M NaOH) and indicate that the destruction of the chemical structures is caused mainly by the breakage of the polymer skeleton. The increase of the water swelling coefficients (%) up to −, 15.06, 8.00, 100, 42.20, 45.00, respectively, is another evidence for the polymer network destruction.1 3.2. Breakthrough Curves−Column Study. To determine the sorption capacity of Purolite A-830 under flow conditions, a series of column sorption experiments were performed in down-flow mode, as described in the Experimental Section. The results are presented in Figure 1 in the form of breakthrough curves by plotting a normalized concentration C/Co versus the volume of the solution passed through the column. C/Co is the ratio of an outlet concentration with the volume passed through the bed to the initial concentration in the Pd(II) feeding solution. At the beginning of the column sorption process, the feeding solution of 100 mg Pd(II)/dm3 concentration flowing into the column passes through a virgin Purolite A-830 bed and at the top layers of WBA resin presets the sorption zone. This part of the bed where the sorption occurs is saturated by palladium(II) complexes therefore the effluent leaving the column from the bottom, is completely liberated from Pd(II) so that in the first few issued samples the concentration of Pd(II) is below the detectable concentration. Then the Pd(II) sorption zone moves down through the column and after reaching the end of the column palladium(II) appears in the effluent. Based on the breakthrough curves the sorption capacities were as follows: 0.119 g/cm3 (0.1 M HCl), 0.039 g/cm3 (1.0 M HCl), 0.006 g/cm3 (3.0 M HCl), 0.002 g/cm3 (6.0 M HCl), 0.007 g/cm3 (0.1 M HCl−0.9 M HNO3), 0.008 g/cm3 (0.2 M HCl−0.8 M HNO3), 0.016 g/cm3 (0.5 M HCl−0.5 M HNO3), 0.023 g/cm3 (0.8 M HCl−0.2 M HNO3) and 0.032 g/cm3 (0.9 M HCl−0.1 M HNO3) for the chloride−nitrate(V) solutions. The comparison of the values of other sorption parameters is given in Figure 1 panels c and d. The analysis of the presented data (Figure 1) leads to the following conclusions: (i) The total concentration of the chloride anions strongly impact the Pd(II) sorption onto Purolite A-830. The chloride anions can act in two different way. First of all the chloride anions compete with Pd(II) anions for binding on protonated amine groups and have an impact on palladium speciation.13,14 Under experimental conditions selected in the present work, Pd(II) predominates under the form of chloro-anionic species such as [PdCl4]2− and the increasing concentration of the chloride anions in the range of 0.1−6.0 M does not change its form (see the distribution graph of Pd species, for example, in ref 15). Additionally, the increasing concentration of Cl− anions significantly impacted on the sorption parameters resulting in 83%, 94.2%, 94.2%, and 98.8% reduction of sorption capacities, weight, bed distribution coefficients, and number of theoretical plates. (ii) The concentrations of hydrochloric and nitric acids cause the sorption capacities to increase with the hydrochloric acid

Figure 1. Breakthrough curves of Pd(II) sorption onto Purolite A-830 for (a) chloride, (b) chloride-nitrate(V) solutions and (c,d) sorption parameters (initial Pd(II) concentration = 100 mg/dm3, flow rate = 0.4 cm3/min, volume of the anion exchanger in the column = 10 cm3).

concentration increase, and nitric acid decreases similar to the number of theoretical plates, whereas the weight and bed distribution coefficients changes have a different trend. At first these values decreased, and after achieving the minimum for the 7225

dx.doi.org/10.1021/ie202859n | Ind. Eng. Chem. Res. 2012, 51, 7223−7230

Industrial & Engineering Chemistry Research

Article

sorption kinetics can be well described by this equation. Similar changes of the kinetic parameters were observed, for example, by Parodi et al.19 (palladium binding on the imidazol contaning resin). Moreover, Helfferich18 pointed out that when the agitation speed affected the ion-exchange rate the sorption process was controlled by resistance to film diffusion, on the other hand, the sorption rate is not influenced by agitation speed but by intraparticle diffusion. As Table S3 shows mass transfer is only slightly affected by agitation speed at the beginning of the Pd(II) sorption process. Initial Concentration of Palladium(II) Solutions. The effect of the initial concentration of Pd(II) solution of 50−300 mg/ dm3 on the sorption process by Purolite A-830 was considered. Values of the amount of Pd(II) sorbed at time t/sorption capacities are collected in Supporting Information, Table S3 and indicate that increasing the initial palladium concentration results in the increase in equilibrium sorption capacities. On the basis of the pseudo-second-order kinetic equations, it can be stated that the calculated and experimental sorption capacities values are the same for the chloride solutions or very close varying by less than 1% for the chloride-nitrate(V) solutions. On the other hand, the initial sorption rate significantly increased with the initial palladium concentration. Additionally, the kinetic constant also increased over this concentration range. Helfferich18 pointed out that when the film diffusion is a rate controlling step, the metal concentration in the solution affects only the ion-exchange rate, but while the system is governed by intraparticle diffusion the sorption rate is not influenced by metal ion concentration. In the case of the chloride solutions, 15−30 min is the contact time necessary for palladium(II) to reach equilibrium. The initial concentration of palladium(II) has little influence on the time of contact necessary to reach equilibrium. Higher differences in contact time needed for the systems to reach equilibrium are observed in the case of the chloride−nitrate(V) solutions; with increasing palladium initial concentration in the solution, the time required to reach equilibrium is longer, varying from 60 to 360 min. Beads Size Effect. Generally, larger particles have a longer transport distance in the particle pores for sorption of metal ions. Therefore, the particle size influences the sorption kinetics. In this study, the beads of Purolite A-830 of different sizes were divided into three fractions, f [mm]: f1, 1; unfavorable isotherm • RL = 1; linear isotherm • RL = 0; irreversible isotherm • 0 < RL < 1, favorable isotherm20−22 The data in Supporting Information, Table S5 (figures of fitting, experimental points) and the values of RL found to be in the range of 0−1 indicate that palladium(II) sorption on the WBA resin was favorable. Favorable palladium sorption was also observed by other researchers, for example, in refs 20 and 21. 3.5. SEM and AFM Studies. SEM images (Figure 2e,f) represent the WBA resin beads at 300× magnification. Most of the resin beads are spherical, noncracked, and, as can be seen, possess different sizes. The bead surfaces are not exactly uniform; they seem to be cut. On their surfaces a large number of irregular lines can be observed which correspond to a large number of cavities, pores. After the palladium sorption process the irregular lines are less distinct. Atomic force microscopy was performed to examine the surface topography and to measure roughness values of the weakly basic anion exchange resin Purolite A-830 before (Figure 2a,c) and after (Figure 2b,d) the sorption process. Figure 2 presents the phase and topographical scans of Purolite A-830 that illustrate the difference in the topography of this resin. The Purolite A-830 surface before and after the palladium(II) sorption as shown by two-dimensional AFM images was rough with large size peaks and valleys. A large numbers of cavities (porous structure) are presented which facilitates the sorption of palladium(II) ions. The surface after the palladium(II) sorption seems to be similar to the surface before the sorption process, numerous similar size peaks are present. The alteration of surface topography was supported by changes in the measured values of the root-mean-square (rms) roughness; it increased from 26.07 to 27.91. Similar changes of roughness were also observed by other researchers.29−31 As mentioned by Deng et al.,29 the increase of roughness may be attributed to the formation of metal complexes which in the case of palladium(II) sorption seems to be possible. 7228

dx.doi.org/10.1021/ie202859n | Ind. Eng. Chem. Res. 2012, 51, 7223−7230

Industrial & Engineering Chemistry Research

Article

(4) Purolite A-830 is characterized by good kinetic properties which are very important from economical and time-consuming points of view. The pseudo-second-order kinetic equation describes well the palladium(II) sorption process on Purolite A830 (R2 > 0.999); the values of sorption capacities predicted by the PSO equation are in good agreement with the experimental ones (varying less than 7%). (5) Sorption equilibrium data correlated with the Langmuir and Freundlich equations showed that the Langmuir isotherm much better described the palladium(II) sorption process onto Purolite A-830. Its maximum sorption capacity for palladium(II) was equal to 356 mg/g and was close to the calculated value of 356.88 mg/g. Additionally, its maximum capacities are higher than those of many other ion exchangers, for example, Varion ADAM (121.5 mg/g), Amberlyst A-23 (137.4 mg/g), Dowex M-4195 (342.3 mg/g). (6) The reduction of capacity was observed in the case of aluminum chloride addition and increasing temperature, as well as with the hydrochloric acid increase and nitric acid decrease. (7) Purolite A-830 can be regenerated and reused but desorption studies presented here must be continued to obtain a more satisfactory eluting agent. Probable quantitative recovery can be obtained by using acidic thiourea solutions or by ion-exchange resin burning. Moreover, WBA resin capacity does not drop or just slightly drops after 3 cycle sorption−desorption. (8) Purolite A-830 can be recommended for palladium(II) recovery from the chloride and chloride−nitrate(V) solutions originating from the processing of spent catalytic convertors, used (petro)chemical catalysts, wastewaters, and anodic slime, etc.



Figure 3. Desorption efficiency (%) of palladium from Purolite A-830.

ASSOCIATED CONTENT

S Supporting Information *

Tables of sorption, kinetic and isotherm parameters characteristics and their values with fitting the experimental and calculated data, properties and chemical stability of Purolite A-830, effect of experimental conditions on Pd(II) sorption and comparison of maximum adsorption capacity obtained to other reported in the literature as well as the comparison of Pd(II) desorption efficiency for different (bio)sorbent are included. This material is available free of charge via the Internet at http://pubs.acs.org.

In more cases (Table S7) these eluting agents make the palladium(II) desorption nearly quantitative or completely quantitative. Sometimes the thiourea eluting solutions application is not satisfactory enough, for example, desorption of Pd(II) from Purolite S-985 and Purolite A-500 (spent catalyst, nitrate solution) using 1 M TU−1 M NaOH, (metal desorption efficiency, −25%)2 or Pt(IV) elution from strongly basic anion exchanger (PtSnIn/Al2O3 or Pt/Al2O3 catalysts).42 Sometimes noble metal ions recovery is possible with the application of a destructive procedure including burning of the loaded ion-exchange resin2 or by a reduction of metals, for example, Pt(IV) to Pt(II) by thiosulphate ions.42



AUTHOR INFORMATION

Corresponding Author

*Tel.: +48 81 537 57 38. Fax: +48 81 533 33 48. E-mail: [email protected].

4. CONCLUSIONS Studies on the sorption of palladium(II) on the weakly basic anion exchange resin Purolite A-830 were carried out. Optimal sorption conditions were determined by batch experiments. The obtained results can be summarized as follows: (1) The effectiveness of palladium(II) sorption onto Purolite A-830 is satisfactory. (2) Purolite A-830 has excellent sorption properties, short contact time, good chemical stability, and high sorption capacity resulting from its polyamine functional groups (N content 16. 88%) toward palladium(II) complexes. (3) The agitation speed, phase contact time, initial palladium concentration, and particle size affect the sorption process. Sorption efficiency slightly or significantly increases with increasing values of the these parameters.

Notes

The authors declare no competing financial interest.



REFERENCES

(1) Neagu, V.; Bunia, I. Acrylic weak base anion exchangers and their chemical stability in aggressive media. Polym. Degrad. Stab. 2004, 83, 133. (2) Kononova, O. N.; Goryaeva, N. G.; Dychko, O. V. Ion exchange recovery of palladium(II) from nitrate weak acidic solutions. Nat. Sci. 2009, 1, 166. (3) Matsubara, I.; Takeda, Y.; Ishida, K. Improved recovery of trace amounts of gold(III), palladium(II), and platinum(IV) from large amounts of associated base metals using anion-exchange resins. Fresenius’ J. Anal. Chem. 2000, 366, 213.

7229

dx.doi.org/10.1021/ie202859n | Ind. Eng. Chem. Res. 2012, 51, 7223−7230

Industrial & Engineering Chemistry Research

Article

(4) Hubicki., Z.; Wołowicz, A.; Leszczyńska, M. Studies of removal of palladium(II) ions from chloride solutions on weakly and strongly basic anion exchangers. J. Hazard. Mater. 2008, 159, 280. (5) Hubicki, Z.; Wawrzkiewicz, M.; Wołowicz, A. Application of ion exchange methods in recovery of Pd(II) ionsA review. Chem. Anal. 2008, 53, 759. (6) Hubicki, Z.; Wołowicz, A.; Wawrzkiewicz, M. Application of commercially available anion exchange resins for preconcentration of palladium(II) complexes from chloride-nitrate solutions. Chem. Eng. J. 2009, 150, 96. (7) Wołowicz, A.; Hubicki, Z. Palladium(II) complexes adsorption from the chloride solutions with macrocomponent addition using strongly basic anion exchange resins, type 1. Hydrometallurgy 2009, 98, 206. (8) Hubicki, Z.; Wołowicz, A. A comparative study of chelating and cationic ion exchange resins for the removal of palladium(II) complexes from acidic chloride media. J. Hazard. Mater. 2009, 164, 1414. (9) Wołowicz, A.; Hubicki, Z. Sorption of palladium(II) complexes onto the styrene−divinylbenzene anion exchange resins. Chem. Eng. J. 2009, 152, 72. (10) Hubicki, Z.; Wołowicz, A. Adsorption of palladium(II) from chloride solutions on Amberlyst A 29 and Amberlyst A 21 resins. Hydrometallurgy 2009, 96, 159. (11) Wołowicz, A.; Hubicki, Z. Selective adsorption of palladium(II) complexes onto the chelating ion exchange resin Dowex M 4195 Kinetic studies. Sol. Extr. Ion. Exch. 2010, 28, 124. (12) Wołowicz, A.; Hubicki, Z. Effect of matrix and structure types of ion exchangers on palladium(II) sorption from acidic medium. Chem. Eng. J. 2010, 160, 660. (13) Jermakowicz-Bartkowiak, B. A preliminary evaluation on the use of the cyclam functionalized resin for the noble metals sorption. React. Funct. Polym. 2007, 67, 1505. (14) Sicupira., D.; Campos, K.; Vincent, T.; Leao, V.; Guibal, E. Palladium and platinum sorption using chitosan-based hydrogels. Adsorption 2010, 16, 127. (15) Kim, Y. H.; Nakano, Y. Adsorption mechanism of palladium by redox within condensed-tannin gel. Water Res. 2005, 39, 1324. (16) Bhandari, V. M.; Yonemoto, T.; Juvekar, V. A. Investigating the differences in acid separation behavior on weak base ion exchange resins. Chem. Eng. Sci. 2000, 55, 6197. (17) Płaziński, W.; Rudziński, W. Theoretical models of sorption kinetics including a surface reaction mechanism: A review. Adv. Colloid. Interface Sci . 2009, 152, 2. (18) Helfferich, F. Ion Exchange; Dover Publications, Inc.: Mineola, NY, 1995. (19) Parodi, A.; Vincent, T.; Pilsniak, M.; Trochimczuk, A. W.; Guibal, E. Palladium and platinum binding on an imidazol containing resin. Hydrometallurgy 2008, 92, 1. (20) Ramesh, A.; Hasegawa, H.; Sugimoto, W.; Maki, T.; Ueda, K. Adsorption of gold(III), platinum(IV), and palladium(II) onto glycine modified crosslinked chitosan resin. Biores. Technol. 2008, 99, 3801. (21) Fujiwara, K.; Ramesh, A.; Maki, T.; Hasegawa, H.; Ueda, K. Adsorption of platinum(IV), palladium(II), and gold(III) from aqueous solutions on L-lysine modified crosslinked chitosan resin. J. Hazard. Mater. 2007, 146, 39. (22) Wawrzkiewicz, M. Sorption of sunset yellow dye by weak base anion exchangerKinetic and equilibrium studies. Environ. Technol. 2010, 32, 455. (23) Paduraru, C.; Bilba, D.; Sarghie, I.; Tofan, L. A sorption of Pd(II) on aminomethylphosphonic Purolite resin S-940. J. Serb. Chem. Soc. 2005, 70, 1205. (24) Muslu, N.; Gülfen, M. Selective separation and concentration of Pd(II) from Fe(III), Co(II), Ni(II), and Cu(II) ions using thioureaformaldehyde resin. J. Appl. Polym. Sci. 2011, 120, 3316. (25) Brinci, E.; Gülfen, M.; Aydin, A. O. Separation and recovery of palladium(II) from base metal ions by melamine−formaldehyde− thiourea (MFT) chelating resin. Hydrometallurgy 2009, 95, 15.

(26) Ma, H.; Liao, X.; Liu, X.; Shi, B. Recovery of platinum(IV) and palladium(II) by bayberry tannin immobilized collagen fiber membrane from water solution. J. Membr. Sci. 2006, 278, 373. (27) Sari, A.; Mendil, D.; Tuzen, M.; Soylak, M. Biosorption of palladium(II) from aqueous solution by moss (Racomitrium lanuginosum) biomass: Equilibrium, kinetic and thermodynamic studies. J. Hazard. Mater. 2009, 162, 874. (28) Park, J; Won, S. W.; Mao, J.; Kwak, I. S.; Yun, Y.-S. Recovery of Pd(II) from hydrochloride solution using polyallylamine hydrochloride-modified Escherichia coli biomass. J. Hazard. Mater. 2010, 181, 794. (29) Neghlani, P. K.; Rafizadeh, M.; Taromi, F. A. Preparation of aminated-polyacrylonitrile nanofiber membranes for the adsorption of metal ions: Comparison with microfibers. J . Hazard. Mater. 2011, 186, 182. (30) Deng, S.; Bai, R.; Chem, J. P. Aminated polyacrylonitrile fibers for lead and copper removal. Langmuir 2003, 19, 5058. (31) Das, S. K.; Das, A. R.; Guha, A. K. Adsorption behavior of mercury on functionalized aspergillus versicolor mycelia: Atomic force microscopic study. Langmuir 2009, 25, 360. (32) Barakat, M. A.; EL-Mahdy, G. A.; Hegazy, M.; Zahran, F. Hydrometallurgical recovery of Nano-palladium from spent catalyst. Open Miner. Process. J. 2009, 2, 31. (33) Hubicki, Z.; Leszczyńska, M.; Łodyga, B.; Łodyga, A. Palladium(II) removal from chloride and chloride−nitrate solutions by chelating ion-exchangers containing N-donor atoms. Miner. Eng. 2006, 19, 1341. (34) Kim, K.-R.; Lee, M-S; Ahn, D-H; Yim, S-P; Chung, H Preparation of functional anion-exchange resin and its selective adsorption of palladium in nitric acid medium. J. Ind. Eng. Chem. 2002, 5, 472. (35) Topp, K.-D.; Grote, M. Synthesis and characterization of a 1,2,4,5-tetrazine-modified ion-exchange resin. React. Funct. Polym. 1996, 31, 117. (36) Moawed, E. A. Preparation of novel ion exchange polyurethane foam and its application for separation and determination of palladium in environmental samples. Anal. Chim. Acta 2006, 580, 263. (37) Leśniewska, B. A.; Godlewska-Ż yłkiewicz, B.; Ruszyńska, A.; Bulska, E.; Hulanicki, A. Elimination of interferences in determination of platinum and palladium in environmental samples by inductively coupled plasma mass spectrometry. Anal. Chim. Acta 2006, 564, 236. (38) Matsubara, I.; Takeda, Y.; Ishida, K. Improved recovery of trace amounts of gold (III), palladium (II), and platinum (IV) from large amounts of associated base metals using anion-exchange resins. Fresenius’ J. Anal. Chem. 2000, 366, 213. (39) Li, W.; Coughlin, M.; Albright, R. L.; Fish, R. H. Polymer pendant ligand chemistry. 4. Recovery of precious metal ions from strongly acidic solution with a polymer-supported o-phenylenediamine hydrochloride ligand. React. Funct. Polym. 1995, 28, 89. (40) Hassanien, M. M.; Abou-El-Sherbini, K. S. Synthesis and characterisation of morin-functionalised silica gel for the enrichment of some precious metal ions. Talanta 2006, 68, 1550. (41) Chakrapani, G.; Mahanta, P. L.; Murty, D. S. R.; Gromathy, B. Preconcentration of traces of gold, silver and palladium on activated carbon and its determination in geological samples by flame AAS after wet ashing. Talanta 2001, 53, 2139. (42) Marinho, R. S.; da Silva, C. N.; Afonso, J. C.; da Cunha, J. W. S. Recovery of platinum, tin and indium from spent catalysts in chloride medium using strong basic anion exchange resins. J. Hazard. Mater. 2011, 192, 1155.

7230

dx.doi.org/10.1021/ie202859n | Ind. Eng. Chem. Res. 2012, 51, 7223−7230