You are on page 1of 7

Journal of Colloid and Interface Science 283 (2005) 278284 www.elsevier.

com/locate/jcis

Letter to the Editor

The removal of humic acids from water by solvent sublation


Yujuan Lu a, , Jianhong Liu a , Jie Tang b , Bo Wei a , Xueli Zhang a
a Chemical and Biology Department, Normal College, Shenzhen University, Shenzhen 518060, Peoples Republic of China b Soils & Materials Engineering Corporation Limited, Hong Kong

Received 11 August 2004; accepted 16 September 2004 Available online 19 November 2004

Abstract The application of solvent sublation in the removal of humic acids was investigated in the present study. The humic acids (HA) were removed from an aqueous solution by solvent sublation of humic acidhexadecylpyridium chloride (HPC) complex (sublate) into isopentanol. Several parameters were examined towards the optimization of humic acid removal; the dosage of a surfactant was found to be the major one, controlling the overall efciency of the progress. The removal rate was somewhat enhanced by higher airow rate and almost independent of the volume of the organic solvent oating on the top of the aqueous column. The effects of electrolytes (e.g., NaCl), nonhydrophobic organics (e.g., ethanol), and pH of the solution upon the process were studied. Under the optimized condition, the treatment performance was found to be very efcient, reaching almost 100%, indicating that solvent sublation can serve as a possible alternative technology for the removal of humic acids. The solvent sublation process follows rst-order kinetics. A characteristic parameter, apparent activation energy of attachment of the sublate to bubbles, was estimated at a value of 9.48 kJ/mol. Furthermore, the simulation of a mathematical model with the experiments on the solvent sublation of humic acidHPC was described here. 2004 Elsevier Inc. All rights reserved.
Keywords: Solvent sublation; Humic acids; Kinetics; Thermodynamics; Simulation

1. Introduction Solvent sublation, originated by Sebba [1] for ionicsurfactant complexes, has shown a promise for removing hydrophobic molecular compounds and ion-pair complexes from aqueous systems. Solvent sublation is one among the several adsorptive bubble separation techniques, wherein a hydrophobic compound is levitated on a bubble surface to the top of an aqueous column where they encounter a solvent layer (e.g., mineral oil, lauryl alcohol) to which the material is transferred as the bubbles move through the solvent layer [2,3]. Solvent sublation process is different from solvent extraction. When ne gas bubbles (of very small radii in the range of 0.010.05 cm and in laminar ow) pass through an aque* Corresponding author. Fax: +86-755-26534605.

E-mail address: yjlv@21cn.com (Y. Lu). 0021-9797/$ see front matter 2004 Elsevier Inc. All rights reserved. doi:10.1016/j.jcis.2004.09.051

ous column (with an overlaying organic layer) containing hydrophobic compounds, because of their inherent tendency to concentrate at the airwater interface, these hydrophobic materials would be collected on the bubble surface by diffusion through the thin boundary layer surrounding the air bubble. As the bubble transits the aqueous column and moves through the organic solvent layer, the adsorbed phase gets stripped into the organic phase. Thus the solvent sublation process improves the efciency of a bubble aeration column. Moreover, the presence of the organic solvent also reduces the eventual redispersion of the material into the aqueous column upon bubble bursting which usually occurs in conventional bubble aeration columns [4,5]. The solvent sublation techniques have been studied by some workers. Most of these studies have been focused on the removal of organic pollutants from aqueous systems, such as alkyl phthalate, volatile chlorinate organics, dichlorobenzenes, nitrophenols, polynuclear aromatics, and chlorinated pesticides [611], and the emission of

Y. Lu et al. / Journal of Colloid and Interface Science 283 (2005) 278284

279

volatile organic compounds (volatile chlorinated organics and toluene) to atmosphere in the solvent sublation [12,13]. Other studies are the solvent sublation of dyes in the aqueous solution, such as removal of bromophenol blue from water by solvent sublation with hexadecyl pyridium chloride into isopentanol [14,15], the separation of methyl orange from Rhodamine B [16], the solvent sublation of the methyl orangehexadecyltrimethylammonium ion pair [17], Magenta (a cationic dye) with sodium lauryl sulfate [18], and Direct Red and Acid Red (two anionic dyes) with hexadecyltrimethylammonium ion [19,20]. Recently both the mechanism and kinetics of solvent sublation were investigated by Palagyi et al. [21,22]. They have proposed an ionic associate formation mechanism and second-order kinetics in the separation of iodide from water by solvent sublation with CPC (N -cetylpyridinium chloride) into benzene. However, the studies of both kinetics and thermodynamics of the solvent sublation are few [2325]. Humic acids, alkali soluble but acid insoluble fraction of humic substances, are structurally complexly large macromolecules; contain several aromatic groups linked together by short saturated aliphatic chains. As we know that the removal of humic acids in aquatics systems is considered of great environmental concern [26]. The searches for a method, which can be equally efcient, but meanwhile more rapid and less expensive than traditional method, have resulted in an interest in the application of solvent sublation. In this paper, kinetics and thermodynamics of humic acids ion complexes, were carried out. Furthermore, a simulation is made between the experimental and the theoretical results by the mathematical model.

spectrophotometer (Unico Com., China) at the maximum peak of 254 nm. 2.3. Procedure Stock solutions of humic acids were prepared by dissolution of 5 g of humic acids in 1 L of 0.1 M NaOH solution, stirring for 48 h. The HPC stock solutions were made by dissolving 0.5 g HPC in 250 ml of water, in order to obtain the nal concentration of 2 g/L. Humic acids are known to be fully protonated, i.e., they are not ionized at pH value 3, thus negatively charged above this pH value [28]. As we know, HPC was a common cationic surfactant. Thus in common, the humic acids are expected to interact strongly with cationic HPC, formation the complex through the ion attraction. The concentration of humic acids were measured at 254 nm on a spectrophotometer (Unico PC2100 UV/vis), using quartz vesicles. A perfect linear relationship within a wide dynamic range between the concentration of humic acids and UV absorbance at 254 nm was found. However, no absorbance was observed for the surfactant HPC at 254 nm. Therefore the surfactant has no effect on the concentration determination of humic acids during the whole solvent sublation process. For the solvent sublation runs, HPC was rstly added to the 50 mg/L HA solution (500 ml) to form the humic acids surfactant complex, followed by co-solute, such as ethanol and NaCl. The HAHPC solution was then poured into the sublation column, to which 5 ml isopentanol was added. Immediately, the timer was started and the samples of aqueous solution were taken out for analysis at a certain time. In the lower pH value, the samples of aqueous solution pH were adjusted to alkali by adding 0.1 M NaOH, keeping the humic acids completely soluble.

2. Experiments 2.1. Reagents Reagent-grade hexadecylpyridium chloride (Shanghai Chemical Agents Factory, China) was used as collector without further purication. Reagent-grade humic acids, isopentanol (Shanghai Chemical Agents Factory, China), and the other agents were all analytical grade. 2.2. Apparatus The solvent sublation system was similar to that described earlier [27]. The glass column was 90 cm in length and 3.7 cm in inside diameter with three access ports, the upper access for solvent discharge, the middle one for sample analysis and the lower one for aqueous discharge. A titanium plate (pore size 510 m) was used to introduce air bubbles into the aqueous phase. Airow rate was measured with a soap ow meter. The pH of the solution was measured with a pHs3C (Shaihai Rex Industry, China). UVvisible spectra of the sample solutions were obtained on a Unico PC2100 UV/vis

3. Experimental results 3.1. The effect of pH The effect of pH on the removal of HAHPC in the process of solvent sublation is shown in Fig. 1. The removal rate and removal efciency of HA decreased with the increase of the value of pH, for the humic acids is the alkali soluble but acid insoluble fraction of substances. In the lower value of pH (acidic), organic protonation was increased, thus reducing the negative charge density of colloid organic matter and subsequently, the less demand of cationic collector for neutralizing the anionic humic acids. On the other hand, in higher pH values, the HA is soluble with larger negative charge density, the demand for collector was increased. For the convenience, the pH 6 was used in the followed experiments.

280

Y. Lu et al. / Journal of Colloid and Interface Science 283 (2005) 278284

Fig. 1. Effect of pH on solvent sublation, [HA]0 = 50 mg/L, [HPC]0 = 40 mg/L, Vw = 500 ml, V0 = 5 ml, rate of airow = 75 mL/min, duration of airow = 15 min.

Fig. 3. Effect of the dosage of HPC under different pH value on solvent sublation, [HA]0 = 50 mg/L, Vw = 500 ml, V0 = 5 ml, rate of airow = 75 ml/min, duration of airow = 15 min.

3.3. Effect of the dosage of HPC under different pH value on solvent sublation The effect of the surfactant dosage was also investigated in relation with the variation of pH value, by keeping constant the initial humic acids concentration at 50 mg/L (Fig. 3). The difference in pH had a direct effect on the required surfactant dosage for efcient humic acids removal. When the solution pH was adjusted to 4, efcient removal was achieved at surfactant concentration of 4060 mg/L for [HA]0 of 50 mg/L and was decreased greatly for values above 60 mg/L. When the pH of solution was increased to pH value of 6, the removal efciency was over 95% at surfactant concentration of 6075 mg/L for [HA]0 of 50 mg/L and decreased slightly for values above 75 mg/L. On the contrary, when the pH of solution was increased to pH value of 10, the sufcient dosage of surfactant was found to be 100120 mg/L. However, the higher surfactant dosage caused the emulsication of the isopentanol (which was observed during the solvent sublation process with a larger excess of surfactant) and a large of forms were produced. The experiment results indicated that the signicant effect of pH on the efciency on the solvent sublation as a treatment technology for removing humic acids. At lower pH values, the required dosage of collector for achieving sufcient removal of this humic acids concentration was lower than the amount required at higher pH values. At pH value of 6, the surfactant dosage had small effect on the removal efcient when the dosage reached 60 mg/L. Thus in the experiment, pH 6 was used. 3.4. Effect of the dosage of HPC in relation to the initial humic acids concentration on solvent sublation The effect of the dosage of HPC in relation to the initial humic acids concentration was presented in Fig. 4. It can be noticed that the optimum dosage of surfactant, in order

Fig. 2. Effect of the dosage of HPC on solvent sublation, [HA]0 = 50 mg/L, Vw = 500 ml, V0 = 5 ml, pH 6, rate of airow = 75 ml/min, duration of airow = 15 min.

3.2. Effect of the dosage of HPC on solvent sublation The effect of HPC concentration on the solvent sublation of HA was shown in Fig. 2. It was found that a 75 mg/L HPC for [HA]0 50 mg/L gave the fastest rate of separation and the lowest residual HA concentration, with over 99% of HA being removed in 15 min. At smaller concentration of surfactant, the rate of removal was slower and the level of the residual HA was greater, presumably due to incomplete formation of a HAsurfactant complex. But when surfactant dosage was higher than 75 mg/L, the rate of solvent sublation was smaller and the removal efciency was slower, presumably due to the competition of the bubble surface by the excess surfactant ion with the HAsurfactant complex. It was observed that large forms were produced when the surfactant dosage was over 80 mg/L. The excess surfactant also caused the emulsication of the isopentanol (which was observed during the solvent sublation process with a larger excess of surfactant), such that the HAsurfactant complex in the isopentanol was constantly dispersed back into the solution and the separation efciency decreased.

Y. Lu et al. / Journal of Colloid and Interface Science 283 (2005) 278284

281

Fig. 5. Effect of amount of ethanol on solvent sublation, [HA]0 = 50 mg/L, [HPC]0 = 40 mg/L, Vw = 500 ml, V0 = 5 ml, pH 6, rate of airow = 75 ml/min, duration of airow = 15 min. Fig. 4. Effect of the dosage of HPC in relation to the initial humic acids concentration on solvent sublation, [HA]0 = 50 mg/L, [HA]0 = 100 mg/L, [HA]0 = 200 mg/L, Vw = 500 ml, V0 = 5 ml, pH 6, rate of airow = 75 ml/min, duration of airow = 15 min.

to obtain efcient humic acids removal varying with initial humic acids concentration and the higher the humic acids concentration, the higher the required dosage of surfactant. This could be attributed to the fact the required dosage of solvent sublation reagent is closely related to the overall charge carried by the sublate. To get the efcient removal of humic acids, their charge has to be neutralized [26]. Therefore, as the amount of humic acids was increased, the charge that had to be neutralized was also increased and this required higher dosages of surfactant. The results presented in Fig. 4 showed that the optimum surfactant dosages were 5075 mg/L for [HA]0 50 mg/L, 120190 mg/L for [HA]0 100 mg/L, 260 360 mg/L for [HA]0 200 mg/L. The highest removal efcient (over 99%) was obtained when the surfactant dosages were 75 mg/L for [HA]0 50 mg/L, 190 mg/L for [HA]0 100 mg/L, 360 mg/L for [HA]0 200 mg/L. Nevertheless, excessive amounts of surfactant should be avoided, not only due to higher cost induced, but also because of other negative effects, such as larger form, the emulsication of the organic solvent, possibility for micelle or hemimicelle formation and the potential toxicity of residual amounts of collector in the treated efuent. 3.5. The effects of the co-solute The inuence of various amounts of ethanol used as cosolute, ranging from 0 to 10 ml in 500 ml HAHPC solution, upon the removal rates of HA was shown in Fig. 5. The low amount (28 ml) of ethanol enhanced the removal efcient and rates, while high amount of ethanol (10 ml) decreased the removal efcient and rates. The enhancement in removal efcient and rates might be due to two factors. First, we noticed that addition of ethanol changed the bubble properties considerably; the number of very small bubbles was much larger than those in which

Fig. 6. Effect of NaCl concentration on solvent sublation, [HA]0 = 50 mg/L, [HPC]0 = 40 mg/L, Vw = 500 ml, V0 = 5 ml, pH 6, rate of airow = 75 ml/min, duration of irow = 15 min.

ethanol was absent. This is a well-known effect arising from the lowering of surface tension of water, which prevents the bubbles from growing to be large sizes. So these smaller bubbles provide a very large surface area per unit volume of air, which contributes to enhance mass transfer from the liquid phase to the bubbles. At the same time, smaller bubbles also have slower rise velocities [29]. However, the phenomenon of decreasing the removal rate with higher mole fraction of ethanol is due to the increase of the solubility of sublate. The effect of inorganic salt NaCl used as co-solute on the sublation of HAHPC complex was shown in Fig. 6. The increasing of NaCl concentration tended to decrease the removal rate and the removal efciency. This was attributed to an ion-pair equilibrium that existed in the aqueous solution between the HA and surfactant molecules. In aqueous solution, there existed the equilibrium HP+ A + Na+ Cl = HP+ Cl + Na+ A , where A represents HA anionic ion and HP+ represents hexadecylpyridiumchlorid cationic ion. It was apparent that the increase of the salt concentration (e.g., NaCl) drove the equilibrium toward a larger concentration of Na+ A , which

282

Y. Lu et al. / Journal of Colloid and Interface Science 283 (2005) 278284

was hydrophilic, and less HPCHA complex (i.e., HP+ A ) existed in the aqueous phase. As a result, the rate of removal decreased. 3.6. The effect of the volume ratio of organic to aqueous solvent for solvent sublation Sebba [1] has shown that in the case of the solvent sublation of ionsurfactant complexes from aqueous solution into 2-octanol, the removal efciency is independent of the amount of the volume of 2-octanol. Caballero has made a conclusion that the sublation efciency is independent on the organic solvent volume only until the saturation of the phase by the sublate occurs [30]. Here two different solvents were investigated. We achieved the same results with these experiments, which are shown in Fig. 7 for isopentanol and in Fig. 8 for 2-octanol. It was observed that the removal efciency was about 30% without the organic solvent and the removal efciency increased with the increase of the solvent when the amount was not more than 7.5 ml. However, no signicant improvement in removal efciency was observed when the volume of isopentanol or 2-octanol

exceeded 7.5 ml in the 500-ml solution. The results showed that sublation was independent of the volume of organic solvent until the saturation of the phase occurs by the sublate. Generally, mass transfer occurs from gas bubbles crossing the aqueoussolvent interface and not from diffusion of solute across this interface, the amount of material transferred should depend only on the amount of air crossing the interface and not on the organic volume. While in liquid liquid extraction the volume ratio of the two immiscible phases is a very important parameter, which is an important difference between the liquidliquid extraction and solvent sublation. But if the organic volume used in solvent sublation is too low, the oilwater interface will be drastically disrupted at a high airow rate and the process will lose its efciency. Hence the airow rates and solvent volume should be chosen to keep the minimal disruption of the interface. 3.7. The effect of different airow rates on solvent sublation The solvent sublation of HA into isopentanol was investigated at three different airow rates (25, 75, 100 ml/min). It was observed that the removal rates increased with the increase of airow rates, as shown in Fig. 9. However the increase in removal rate was out of proportional to the increase of airow rates, similar to the results by Valsaraj et al. [4]. This was probably explained by the nding that with increasing airow rates the mean bubble radius increased, thus the interfacial area per unit volume of air (which is given by 3/r) decreased, and the bubble residence time also reduced since larger bubbles had higher rise velocities. Furthermore, the axial dispersion certainly increased with the increase of the airow rates, which would impair the performance of the sublation process. It was observed that at higher ow rates the oilwater interface was drastically disrupted and some drops of the top layer could return to the solution. Although the increased air-

Fig. 7. Effect of solvent (isopentanol) volume on solvent sublation, [HA]0 = 50 mg/L, [HPC]0 = 40 mg/L, Vw = 500 ml, V0 = 5 ml, pH 6, rate of airow = 75 ml/min, duration of airow = 15 min.

Fig. 8. Effect of solvent (2-octanol) volume on solvent sublation, [HA]0 = 50 mg/L, [HPC]0 = 40 mg/L, Vw = 500 ml, V0 = 5 ml, pH 6, rate of airow = 75 ml/min, duration of airow = 15 min.

Fig. 9. Effect of airow rate on solvent sublation, [HA]0 = 50 mg/L, [HPC]0 = 40 mg/L, Vw = 500 ml, V0 = 5 ml, pH 6, duration of airow = 15 min.

Y. Lu et al. / Journal of Colloid and Interface Science 283 (2005) 278284

283

ow rate can improve the removal rate of solvent sublation, the removal efciency would decrease if the airow rate was quite high, for the air currents arise from the quite high airow rates would disrupt the oilaqueous interface. 3.8. The kinetics and thermodynamics of the solvent sublation In the chemical reaction, the rate is followed by dc/dt = kcn , (1)

where c is the reaction concentration, k is the apparent rate constant and n is the order of the chemical reaction. Upon the analysis of the experimental results, we found that the kinetics of the solvent sublation process was obedient to the equation dc/dt = kc. (2)

Fig. 11. Curve of ln A vs t; data from Fig. 10.

It showed that the kinetics mechanism of solvent sublation process was consistent with the rst-order kinetics equation. In the solvent sublation process, rst the HAHPC molecules and the bubbles attracted, and then the repulsive forces increased, which would result in the boundary layer compressed and become thinner and then eclipse nally. The change of energy of this process was similar to that of the chemical reaction process [4]. So we could use the Arrhenius equation to describe the process, ln k = Es /RT + B, (3)

solvent sublation process experiences. With the same concentration of HAHPC solution, the experiment temperature was set at 298, 318, 338 K. The kinetics of HA at different temperature is shown in Fig. 10. It was observed that the rate of solvent sublation was decreased with the increase of temperature. The relationship of ln k vs 1/T is linear, which was shown in Fig. 11. The value of apparent active energy was calculated as 9.48 kJ/mol by the slope of Fig. 11. The value of the HAHPC system was higher than that of the methyl violetdodecylbenzulfonate system (8.19 kJ/mol) [15], so the solvent sublation of the front was slower than the latter.

4. Comparison of theory and experiment The mathematical model of dyesurfactant for the solvent sublation has been described in our earlier reports [25]. In the experiment condition, the HA was anionic style in the solution. So the mathematical model can be applied to the HAHPC complex system. The experimental parameters such as column radius, airow rate, concentration of HA and HPC, organic volume, and aqueous phase volume are as the real experimental data in the model. By adjusting the value of other microparameters, the simulation results can be obtained. The simulation of experimental results for the solvent sublation of HAHPC was shown in Fig. 12. The symbols can be referred in the literature [25]. The results showed that the mathematical can describe somewhat the solvent sublation of HAsurfactant. The discrepancy was existed when the humic acids were many substances and the microparameters were not sole.

where k is the apparent rate constant, R is the common gas constant, B is the integrate constant, Es is the solvent sublation apparent activation energy. The relationship of ln k and 1/T is linear, and the value of the apparent active energy can be calculated from the slope of the line, which can be used as a characteristic parameter to describe the solvent sublation process. The higher of the value is, the slower of the

5. Conclusions Solvent sublation was found to be an efcient treatment option as a posttreatment stage for removing humic acids with a cationic surfactant, HPC, from aqueous system. Under optimized conditions, the removal of humic acids was

Fig. 10. Kinetics of solvent sublation with different temperature, [HA]0 = 50 mg/L, [HPC]0 = 40 mg/L, Vw = 500 ml, V0 = 5 ml, pH 6, rate of airow = 75 ml/min, duration of airow = 15 min.

284

Y. Lu et al. / Journal of Colloid and Interface Science 283 (2005) 278284

greatly decrease the removal rate of sublation, because it moved the ion equation toward the hydrophilic product direction. The kinetics of the solvent sublation was followed by rst-order kinetics. The apparent active energy was put forward as a parameter of the solvent sublation and was calculated as 9.48 kJ/mol. And the mathematical model can describe the experimental results with a small discrepancy.

References
[1] F. Sebba, Ion Flotation, Elsevier, New York, 1962, pp. 112123. [2] R. Lemlich, Adsorptive Bubble Separation Techniques, Academic Press, New York, 1972. [3] T.F. Carleson, in: J.T. Scamehorn, J.H. Harwell (Eds.), SurfactantBased Separation Process, Dekker, New York, 1989. [4] K.T. Valsaraj, J.L. Portor, E.K. Liljenfeldt, C. Springer, Wat. Res. 20 (1986) 11611175. [5] Y.J. Lu, X.H. Zhu, Sep. Purif. Methods 13 (2001) 157189. [6] D.J. Wilson, D.E. Pearson, Report RU-83/6, Bureau of Reclamation, U.S. Department of Interior, Washington, DC, 1984. [7] K. Tamamushi, D.J. Wilson, Sep. Sci. Technol. 19 (1984) 10131023. [8] T. Lionel, D.J. Wilson, D.E. Pearson, Sep. Sci. Technol. 16 (1991) 907935. [9] D.J. Wilson, K.T. Valsaraj, Sep. Sci. Technol. 17 (1983) 13871396. [10] S.D. Huang, K.T. Valsaraj, D.J. Wilson, Sep. Sci. Technol. 18 (1983) 941968. [11] K.T. Valsaraj, D.J. Wilson, J. Colloids Surf. 8 (1983) 203224. [12] V. Ososkov, B. Kebbekus, C.C. Chou, Sep. Sci. Technol. 31 (1996) 213227. [13] V. Ososkov, B. Kebbekus, M. Chen, Sep. Sci. Technol. 31 (1996) 13771391. [14] Y.J. Lu, X.H. Zhu, Sep. Sci. Technol. 36 (2001) 37633776. [15] Y.J. Lu, X.H. Zhu, Sep. Sci. Technol. 38 (2003) 13851398. [16] A.B. Carger, B.L. Karger, Anal. Chem. 38 (1966) 652654. [17] B.L. Karger, T.A. Pinfold, Sep. Sci. 5 (1970) 603617. [18] G.L. Sheu, S.D. Huang, Sep. Sci. Technol. 22 (1987) 22532262. [19] J.Y. Huang, S.D. Huang, Sep. Sci. Technol. 26 (1991) 5971. [20] M.H. Cheng, S.D. Huang, J. Colloids Interface Sci. 126 (1998) 346 354. [21] S. Palagyi, Czech. J. Phys. 49 (1999) 739745. [22] S. Palagyi, Chem. Pap. Chem. Zvesti. 52 (1998) 671681. [23] Y.J. Lu, J.H. Liu, Y. Xiong, X.H. Zhu, J. Colloids Interface Sci. 263 (2003) 261269. [24] Y.J. Lu, X.H. Zhu, Talanta 57 (2002) 891898. [25] Y.J. Lu, X.H. Zhu, Anal. Bioanal. Chem. 374 (2002) 906914. [26] A.I. Zouboulis, J. Wu, I.A. Katsoyiannis, Colloids Surf. A 231 (2003) 181189. [27] Y.J. Lu, Y.S. Wang, Y. Xiong, X.H. Zhu, Fresen. J. Anal. Chem. 370 (2001) 10711076. [28] M. Bob, H.W. Walker, Colloids Surf. A 191 (2001) 17. [29] V.G. Levich, Physicochemical Hydrodynamics, PrenticeHall, Englewood Cliffs, NJ, 1962, p. 434. [30] M. Caballero, R. Cela, J.A. Perez-Bustamante, Talanta 37 (1990) 275 300.

Fig. 12. Simulation of experimental results for the solvent sublation of HAHP. The line represents the theoretical result; the circles represent the experimental result. [HA]0 = 50 mg/L, [HPC]0 = 40 mg/L, Ke = 0.22, Ka1 = 0.002, Ka2 = 0.046 s1 ; Qa = 75 ml/min, Kl1 = 0.001 cm/s, Kl2 = 0.0008, Kow = 800, Vw = 500 ml, V0 = 5 ml, rc = 3.7 cm, di = 0.0001 cm, a = 0.1 cm.

found to be very efcient for a wide range of initial humic acids concentration (50200 mg/L). The parameters, which control the removal efciency, were mainly the dosage of surfactant, the pH of solution, the airow rate, ionic strength and the solvent sublation time. The dosage of surfactant was the most important factor controlling the removal of humic acids and the optimum concentration was greatly dependent on the initial humic acids concentration and on the pH of solution; higher humic acids concentrations required the higher dosages of surfactant. The increased airow rates enhance the process of solvent sublation, in which the bubbles size was small. However, at higher airow rates, the generation of axial dispersion would compromise the efciency of the solvent sublation. The removal efciency increased with the increase of the solvent when the amount of solvent did not reach the saturation of the phase induced by the sublate. However, the solvent sublation was somewhat independent of the organic volume, provided that the organic volume was larger than a critical value, that is, disruption of organic-aqueous interface was minimal. A smaller fraction of ethanol (not more than 0.004) enhanced the efciency of the solvent sublation; by contrast, larger fraction of ethanol (0.005) decreased the efciency of removal. The increase of the NaCl concentration would

You might also like