You are on page 1of 6

Journal of the Society of Leather Technologists and Chemists, Vol. 90 p.

ADSORPTION OF SURFACTANTS ON CHROMIUM LEATHER WASTE


ZHANG MI-NA, LIAO XUE-PIN and SHI BI*
The Key Laboratory of Leather Chemistry and Engineering of Ministry of Education, Sichuan University, Chengdu, 610065, P. R. China
Summary We have studied the adsorptions of aqueous solutions of three kinds of surfactants, anionic (sodium dodecylbenzenesulfonate, SDBS), cationic (dodecyl trimethylammonium bromide, DTB) and a non-ionic (Triton X-100, TX-100), by chromium-containing leather waste. The results indicated that the anionic surfactant can be significantly adsorbed by the waste material adsorbent over the pH range 4.0-8.0. The adsorption capacities of the cationic surfactant and non-ionic surfactant on the adsorbent are limited. These facts imply that the predominant adsorption sites of such an adsorbent are amino groups and the Cr(III) combined with collagen. The adsorption capacity of SDBS increased with the rise of temperature when the initial surfactant concentration was 2000mg/L, the greatest adsorption of SDBS at 293K was 375mg/g, reaching 423mg/g at 313K, which indicates that the mechanism of the adsorption process may be chemical adsorption. Higher ionic strength also leads to higher adsorption capacity. In general, the adsorption isotherms of SDBS on the adsorbent can be fitted by the Langmuir model. The adsorption kinetics data can be well described by the pseudo-second-order rate model. The adsorption capacities calculated by the pseudo-second-order rate model are close to those determined by actual measurements (error <10 %). The studies of column adsorption kinetics show that the breakthrough point of SDBS is around 150 x bed volume in the experimental system, which indicates the adsorption column has a high availability for the adsorption of anionic surfactant.

Surfactants are widely used in industrial processes for their favourable physicochemical characteristics such as detergency, foaming, emulsification, dispersion and solubilization effects.1-4 Such extensive applications of surfactants have produced environmental pollution and have raised problems in wastewater treatment plants.5-6 The experimental data have shown that surfactants can kill microorganisms at very low concentrations (1-5mg/L) and harm them at even lower concentrations (0.5mg/L).7 In addition, surfactants can produce foams, which are a significant problem in sewage treatment. Therefore, the removal of the surfactants from wastewater is important in reducing their environmental impact. Many techniques have been used for removal of surfactants in aqueous solution. Among these, biological degradation, ozonation and extraction are often costly and may possibly create secondary pollution because of excessive use of chemicals.3, 8-9 In addition, a large number of surfactants used at present, like linear alkylbenzene sulphonates, have relatively low biodegradability.10 The adsorption process seems to be an effective method for the removal of surfactants, especially in the case of low concentrations of surfactant. A number of adsorption systems have been explored for the removal of surfactants. Numerous adsorbents, such as activated carbon, layered double hydroxides, silica, mineral oxides and natural biomasses, have been extensively investigated.11-13 Recently research has focused on low cost and easily available materials such as waste activated carbon and rubber granules.14-15 In our previous work, it was found that chromiumcontaining leather waste was able to effectively adsorb some dyes from wastewater due to its particular structure.16 This kind of waste material has a high content of functional groups that could bind with the hydrophobic and
*Corresponding author. E-mail address: sibitannin@vip.163.com

Introduction

hydrophilic parts of surfactants by electrostatic forces or hydrophobic bonds. Moreover, the chromium adhering to the collagen also has a tendency to bind surfactants.17-18 Therefore, chromium-containing leather waste has the potential to be used as a low-cost adsorbent for the removal of surfactants from wastewater. The purpose of this study is to investigate the adsorption behavior of this adsorbent for typical surfactants in water. Adsorption equilibrium, adsorption kinetics and column adsorption kinetics were studied on a laboratory scale using typical commercial surfactants. All of the experiments were replicated and the errors were controlled to be less than 5%.

Experimental procedure
Materials Chromium-containing leather shavings were obtained from a local tannery, and were washed with de-ionized water, dried at 60C for 6hrs and then ground into granules with particle size 0.1-0.2mm. The surfactants used in this study were: anionic surfactant (sodium dodecylbenzenesulfonate, SDBS); cationic surfactant (dodecyl trimethylammonium bromide, DTB) and non-ionic surfactant (Triton X-100, TX-100) with purity >99%, purchased from the Sigma Company. All the other chemicals were of analytical reagent grade. All the solutions were prepared using de-ionized water. Adsorption isotherms A series of batch experiments were carried out to determine the adsorption isotherms of surfactants on the adsorbent. 0.200g adsorbent was put into 100ml surfactant solutions at given concentration. The adsorption experiments were conducted by constant shaking at 293K for 24hrs. After adsorption, the surfactant solutions were isolated by centrifugation (GL-20G, Shanghai, China). 1

Then, the contents of SDBS and TX-100 in the residual solutions were determined by UV spectrophotometry (Model UV-2510PC, Japan) at max=224nm for SDBS and max=223nm for TX-100. The concentration of DTB in the residual solution was determined using gas chromatography (GC). The adsorption capacity of surfactant on the adsorbent, qe (mg/g), was calculated by a mass balance relation (1):

v qe = (c0 - ce) w

(1)

where c0 and ce are the initial and equilibrium concentrations of surfactant respectively (mg/L), v is the volume of solution (L), and w is the weight of the adsorbent used (g). The effects of temperature, pH value, adsorbent dose and electrolyte on the adsorption capacity of the adsorbent to SDBS were also investigated. 0.1N HCl or 0.1N NaOH was used to adjust the pH value of solutions. Adsorption kinetics 0.200g adsorbent was put into 100ml SDBS solutions at concentrations of 300mg/L, 400mg/L, 500mg/L, respectively. The adsorption kinetics experiments were carried out at 303K and the concentrations of SDBS in the solution during adsorption process were analyzed at a regular intervals until equilibrium was achieved. Column adsorption kinetics 2.00g adsorbent was soaked in de-ionized water for 24h. Then the adsorbent was filled into a glass column with inner diameter 11mm, bed height 270mm. 100mg/L SDBS solution was pumped into column at a constant volume velocity of 1.77BV/h (BV=bed volume). Column effluent was collected using an automatic collector (BSZ-100, Shanghai, China) and the concentration of SDBS in the effluent was determined spectrophotometrically.

cationic surfactant, and the adsorption mainly depends on its association with the limited residual carboxyl groups on collagen. Moreover, electrostatic repulsion may exist between positively charged adsorbent and surfactant. As a result, the equilibrium adsorption capacity of DTB is much lower than that of SDBS. Figure 1 shows that the equilibrium adsorption capacity of DTB has a maximum value. A similar type of isotherm has been found by several researchers for the adsorption of anionic or cationic surfactants on carbons, metals and fibres.19 The explanations for this are mainly focused on two aspects: (1) with the increase of the surfactant concentration, the ionic strength is increased and the depth of the surface double layer is decreased, which results in the transformation of a surface micelle, and (2) the swelling of the solid materials in surfactant solution causes the increase of available surface of adsorbent. Unfortunately, there is no definite conclusion that can be drawn at present. The lowest adsorption capacity which is shown by TX-100 on chromium-containing leather waste is due to the fact that the hydrophobic and hydrogen bond interactions between surfactants and the adsorbent are weak. As the anionic surfactant SDBS has the highest adsorption capacity on the adsorbent, subsequent experiments focused on the study of the adsorption of SDBS.

Results and Discussion


Adsorption isotherm of surfactants on chromiumcontaining leather waste The adsorption isotherms of the adsorbent to surfactants at 293K are shown in Figure 1. It can be seen that the three surfactants with different structures present varying adsorption isotherms. The adsorption isotherm of SDBS is a typical L type and the equilibrium adsorption capacity of SDBS on the adsorbent has the greatest value. The equilibrium adsorption capacity of DTB is considerably lower than that of SDBS and its adsorption capacity deceases rapidly after a plateau amount is reached. Compared with SDBS and DTB, TX-100 is little adsorbed by chromium-containing leather waste. The adsorption of an ionic surfactant at solid-liquid interface is strongly influenced by the charged groups on the surface. The active sites of collagen are mainly carboxyl and amino groups, but most of carboxyl groups on collagen had been blocked by Cr(III) through the complex reaction after tanning. Therefore, the predominant active sites of the chromium-containing leather waste should be amino groups and the Cr(III) combined with collagen. Both amino groups and Cr(III) have the potential to associate with anionic groups through electrostatic interaction. So the adsorption capacity of SDBS which contains sulfonic acid groups is significantly high. However, DTB is a 2
Figure 1. Adsorption isotherms of surfactants (293K)

Effect of temperature on the adsorption isotherm of SDBS The adsorption isotherms of SDBS at different temperatures are presented in Fig. 2(a). It can be seen that all the adsorption isotherms are of the typical Langmuir type. It is well known that this type of adsorption is valid in theory only under the following conditions:20 1. The adsorption is homogeneous, 2. both solute and solvent have equal molar surface area; 3. both surface and bulk phase exhibit ideal behavior (e.g., no solute-solute or solute-solvent interaction in either phase) and 4. the adsorption film is monomolecular. Obviously, in the case of this surfactant, these assumptions are not totally valid in our experimental system. The fact that the adsorption isotherms correspond to the Langmuir type may be due to the mutual compensation of several factors which affect the shape of isotherm.When adsorption isotherms follow the Langmuir type, the equilibrium data could be processed by employing the Langmuir equation:

ce ce 1 = + qe qmax b qmax

(2)

where ce and qe are respectively, the concentration (mg/L) and amount absorbed (mg/g) at equilibrium, b is the Langmuir constant, qmax is the maximum adsorption capacity (mg/g).

surfactant increased with the increase of adsorbent dose and attained 97%-98% when the adsorbent dose was 0.4g. This increase of removal should be due to the increase of adsorption sites available when more adsorbent is used. It was also found in the experiments, that the extent of removal of SDBS could not be further increased when the offer of adsorbent exceeded 0.400g. This might be due to the particle interaction, such as aggregation, resulting from high adsorbent concentration.21 Such aggregation would lead to a decrease in the total surface area of adsorbent and an increase in diffused path length.

s
(a) Adsorption isotherm of SDBS

Figure 3. Effect of adsorbent dose on the removal of SDBS.

(b) Langmuir equation fitting for adsorption isotherms of SDBS

Figure 2. Adsorption isotherms of SDBS on chromium containing waste.

Fig. 2(b) shows that the adsorption equilibrium data of the adsorbent towards SDBS can be well fitted by the Langmuir model. Table I shows the parameters obtained from equation (2). The correlation coefficients (R2) for the linear equation fittings at different temperatures were greater than 0.99. Further, the adsorption capacity increased with the rise of temperature. At 293K, the maximum adsorption of SDBS was 386mg/g and, as the temperature increased to 313K, the maximum amount rose to 438mg/g. These facts imply that the adsorption of SDBS on chromium-containing leather waste adsorbent is favored at high temperature and therefore the adsorption process is seen to be a chemical process.
TABLE I Adsorption isotherm parameters of SDBS at different temperatures

Temp. (K) 293 303 313

qmax (mg/g) 386 395 438

b x 102 2.26 2.38 2.91

Effect of pH on the extent of removal of SDBS Figure 4 shows the effect of pH on the extent of removal of SDBS in solution using chromium-containing leather waste as the adsorbent. The adsorbent exhibited high adsorption efficiency for SDBS with removal higher than 95% within a pH range of 4.0-8.0. The extent of removal decreased dramatically at pH >8.0. At relatively lower pH, more protons are available and, as a result, more amino groups and the complexed Cr(III) in the chromiumcontaining leather waste are positively charged, which increases the amount of SDBS adsorbed due to the increase of adsorption sites for the anionic surfactant. On the other hand, the adsorption capacities at alkaline pH are lower due to the decrease of positively charged sites on adsorbent and the competition between OH- and anionic surfactant for the adsorption sites. Further, comparative experiments were performed using white hide powder and chromium-containing hide powder. It was found that, when the concentration of SDBS was 50mg/L and the adsorbent dose was 0.100g, the removal of SDBS was 94.8% using chromium-containing hide powder, and only 37.6% using white hide powder. This fact suggests that the chromium combined with collagen plays a critical role in the removal of SDBS. Effect of electrolyte on the adsorption of SDBS The effects of electrolyte on the adsorption of SDBS are presented in Figure 5. The concentration of NaCl has a significant effect on the adsorption capacity of SDBS. An increase of the electrolyte concentration leads to an increase of adsorption capacity. At the interface between surfactant and chromiumcontaining leather waste adsorbent, there is always an unequal distribution of electrical charges. This unequal distribution gives rise to a potential across the interface and 3

R2 0.999 0.998 0.999

Effect of adsorbent dose on the extent of removal of SDBS Figure 3 shows the effect of adsorbent dose on the extent of removal of SDBS in solutions with varying initial concentrations. It was noticed that the extent of removal of

dqt = k2(qe-qt)2 dt

(3)

where qe (mg/g) and qt (mg/g) are the amounts of surfactant adsorbed on adsorbent at equilibrium and at time t (min) respectively; k2 (g/mg.min) is the rate constant of pseudo-second-order model. Separating the variable in equation (3) and integrating gives:

t t 1 = + 2 qt qe k2qe

(4)

Figure 4. Effects of pH on SDBS removal by chrome-containing leather waste.

forms a so-called electrical double layer.23 With the increase of NaCl concentration, the electrical double layer on the surface of adsorbent is compressed and the electrostatic repulsion between the adsorbed surfactant species decreases, which results in the increase of adsorption capacity. Adsorption kinetics The experimental data of adsorption kinetics indicated that the adsorption process was very fast at the beginning and then slowed down as equilibrium approached, as shown in Figure 6. Both pseudo-first-order and pseudosecond-order rate models were used to fit the experimental data. It was found that the pseudo-second-order rate model gave perfect fittings to the experimental data. The pseudo-second-order rate model can be written as:22

The equilibrium adsorption capacity (qe) and the pseudo-second-order model rate constant (k2) can be experimentally determined from the slope and intercept of a plot of t/qt against t, respectively. The fittings of experimental data by using pseudosecond-order model are shown in Figure 6 (b), and the parameters of pseudo-second-order model are listed in Table II. It is shown that the pseudo-second-order model gives a very close fit to all of the experimental data. The equilibrium adsorption capacities calculated by pseudosecond-order model are close to those of actual measurements, and the errors are less than 10%.
TABLE II Adsorption kinetics parameters of SDBS on chromium-containing leather waste adsorbent (303K)

Initial conc. of SDBS(mg/L) 300 400 500

k2x104 qe (g/min) (mg/g) 1.309 0.789 0.936 149.0 193.4 229.4

qe (mg/g) 142.7 178.3 208.0

R2 0.9996 0.9999 0.9998

Error 4.3 7.8 9.3

(a) Adsorption kinetics of SDBS (a) Adsorption isotherm of SDBS

(b) Langmuir equation fitting for adsorption isotherms of SDBS

(b) Pseudo-second order model fitting for adsorption

Figure 5. Effect of ion strength on adsorption of SDBS by chrome-containing leather waste adsorbent

Figure 6. Adsorption kinetics of SDBS on chrome-containing leather waste.

Column adsorption kinetics Figure 7 is the breakthrough curve of SDBS on the adsorption column. It can be seen that the breakthrough point was around 150 x bed volume, and SDBS was almost completely adsorbed before the breakthrough point. This observation suggests that the chromium-containing leather waste has a high capacity for removing SDBS in aqueous solution.

(0.15mol/L) were used to desorb the surfactant adsorbed the chromium-containing leather waste. onto Unfortunately, none of these solutions was effective enough to regenerate the adsorbent. This issue needs to be further studied. The difficulty of desorption of SDBS may be related to the adsorption mechanism of surfactant on adsorbent. Six mechanisms are believed to occur during surfactant adsorption onto solid substrates from aqueous solution.11,23 They are: (1) ion exchange, (2) ion pairing, (3) acid-base interaction, (4) adsorption by polarization of pelectron, (5) adsorption by dispersion, and (6) hydrophobic bonding. Considering the structure of surfactants and the characteristics of chromium-containing leather waste, it could be inferred that the adsorption mechanisms of (1), (2), (5) and (6) may be involved in the adsorption of SDBS. The difficult desorption also indicates that interaction between the adsorbent and SDBS is very sophisticated.

Figure 7. Breakthrougth curve of SDBS on adsorption column (inlet conc: 100mg/L; T293 K; feed rate1.77 BV/h; bed height 270mm; column bore 11mm).

Conclusions

The column adsorption kinetic was further specified by determining the mass transfer coefficient according to the Adams-Bohart equation:23

1n(Ct/Ci) = k x Ci x t - k x qv(Z/ui)

(5)

where Ct is the concentration of SDBS in the effluent (mg/L) at time t (min), Ci is the initial or inlet SDBS concentration (mg/L), qV is the volume of SDBS solution uptaken by adsorbent (mg/L), Z is the height of column (m), ui is the linear flow rate of solution (m/min) and k is the mass transfer coefficient (L/mg.min). A straight line was attained for this system by plotting ln (Ct / Ci) against t, as illustrated in Fig. 8, which gives the value of k from the slope of the line. The mass transfer coefficient for the experimental system was 4.08 x 10-5 L/(mg.min).

The experimental results indicate that absorption on chromium-containing leather waste is an effective way of removing anionic surfactant from wastewater. The results of adsorption equilibrium and adsorption kinetics depend on initial concentration of surfactant, temperature, pH, electrolyte and adsorbent levels. Column adsorption experiments demonstrate the potential capability of chromium-containing leather waste to be used for removing anionic surfactants in aqueous solutions. But further studies will be required to discover an effective desorption method for practical application.

Acknowledgment

The authors thank National Science Fund for Distinguished Young Scholars (20325619) and National Science Foundation of China (20476062) for the financial support for this work. (Received August 2005)

t x 10-3 (min)

Figure 8. Determination of the mass transfer coefficient in column adsorption

Solutions of sodium bicarbonate (0.1mol/L), sodium carbonate (0.1mol/L), ethanol-water(1:1), acetonewater(1:1), carbamide (0.5mol/L) and ammonia 5

References 1. Lawrence, M. J. and Ress, G. D. Microemulsion-based media as novel drug delivery systems. Advanced Drug Delivery Reviews, 2000, 45(1), 89-121. 2. Czapla, C. and Bart, H. J. Characterization and modeling of the extraction kinetics of organic acids considering boundary layer charge effects. Chemical Engineering & Technology, 2000, 23(12), 1058-1062. 3. Lin, S. H., Lin, C. M. and Leu, H. G. Operating characteristics and kinetics studies of surfactant wasterwater treatment by fenton oxidation. Water Research, 1999, 33(7), 1735-1741. 4. Sabah, E., Turan, M. and Celik, M. S. Adsorption mechanism of cationic surfactants onto acid- and heat-activated sepiolites. Water Research, 2002, 36(16), 3957-3964. 5. Odokuma, L. O. and Okpokwasili, G. C. Seasonal influences of the organic pollution monitoring of the New Calaber river, Nigeria. Environmental Monitoring Assessment, 1997, 45(1), 43-57. 6. Daley, M. A., Mangun, C. L., Debarr, J. A. et al., Adsorption of SO2 onto oxidized and heat-treated activated carbon fibres (ACFS). Carbon, 1997, 35(3), 411-417. 7. Falbe, J., Surfactant in consumer products: theory, technology and application, 1986. Verwertungsgesellshaft Wor, Munich. Chap. 8-9.

8.

9.

10.

11. 12. 13. 14.

Rozzi, A., Antonelli, M., Angeretti, C. et al., Removal of non ionic surfactants used in the tannery by an adsorbent resin. In: Proceedings of the CIWEM Conference Wastewater Treatment: Standards and Technologies to Meet the Challenge of the 21st Century, Leeds, UK, 2002, 4-6. Wu, S. P. and Pendleton, P., Adsorption of anionic surfactants by activated carbon: effect of surface chemistry, ionic strength, and hydrophobicity. Journal of Colloid and Interface Science, 2001, 243(2), 306-315. Prats, D., Ruiz, F., Vazquez, B. et al., Removal of anionic and nonionic surfactants in a wastewater treatment plant with anaerobic digestion, a comparative study. Water Research, 1997, 31(8), 19251930. Pavan, P. C., Crepaldi, E. L. and Valim, J. B., Sorption of anionic surfactants on layered double hydroxides. Journal of Colloid and Interface Science, 2000, 229 (2), 346-352. Vanjara, A. K. and Dixit, S. G., Formation of mixed aggregates at the alumina. Aqueous surfactant solution interface. Langmuir, 1995, 11(7), 2504. Brown, W., and Zhao, J. X., Adsorption of sodium dodecylsulfate on polystyrene latex particle using dynamic light scattering and zeta potential measurements. Macromolecules, 1993, 26(11), 2711-2715. Sandeep, G., Anjali, P., Pranab, K. G. et al., Performance of waste activated carbon as a low-cost adsorbent for the removal of anionic surfactant from aquatic environment. Journal of Environmental Science and Health - Part A. Toxic/Hazardous Substance and Environmental Engineering, 2003, 38(2), 381-397.

15. Paritosh, D. P., Anjali, P. and Manas, B., Adsorption of anionic surfactant by a low- cost adsorbent. Journal of Environmental Science and Health - Part A Toxic/Hazardous Substances and Environmental Engineering, 2002, 37(5), 925-938. 16. Mina Zhang and Bi Shi., Adsorption of dyes from aqueous solution by chromium-containing leather waste, J. Soc. Leather Technol. Chem., 2004, 88(6), 236-241. 17. Tibor, C., Esther, F. and Gyula Oros., Biological activity and environmental impact of anionic surfactants, Environment International, 2002, 28(5), 337-348. 18. Krejei, J., Interaction of ionogenic surfactants with tanned collagen. Leather Science and Engineering, 2002, 12 (1), 9-10 (in Chinese). 19. Shinoda, K., Nakagawa, T., Tamamushi, B.I. et al., Colloidal surfactants. Academic Press, New York and London, (1963), Chap.3. 20. Rosen, M. J., Surfactants and interfacial phenomena. A WileyInterscience Publication, 1988, Chap.2. 21. Shukla, A., Zhang, Y. H., Dubey, P. et al., The role of sawdust in the removal of unwanted materials from water. Journal of Hazardous Materials, 2002, 95(1-2), 137-152. 22. Ho, Y.S. and Mckay, G., Sorption of dye from aqueous solution by peat. Chemical Engineering Journal, 1998, 70(2), 115-124. 23. Pethkar, A. V. and Paknikar, K. M., Recovery of gold from solution using Cladosporium cladosporioides biomass beads. Journal of Biotechnology, 1998, 63(2), 211-215.

You might also like