You are on page 1of 7

Separation and Purication Technology 55 (2007) 2329

Adsorption thermodynamic and kinetic studies of Pb(II) removal from water onto a versatile Al2O3-supported iron oxide
Yao-Hui Huang , Chan-Li Hsueh, Chun-Ping Huang, Liang-Chih Su, Chuh-Yung Chen
Department of Chemical Engineering, National Cheng Kung University, Tainan City 701, Taiwan Received 4 July 2006; received in revised form 23 October 2006; accepted 23 October 2006

Abstract The use of versatile sorbents has been investigated for wastewater treatment. An investigation into the use of activated alumina-supported iron oxide (denoted as FeAA), which is a by-product of the wastewater treatment plant. In our previous study, the FeAA has successfully been as heterogeneous photoassisted Fenton catalyst for degradation azo dye at neutral pH 7.0. In this work, the test and use of FeAA as adsorbent for the removal of lead (Pb2+ ) from water are carried out. The highest Pb2+ adsorption capacity was determined as 0.14 mmol g1 for 0.8 mmol dm3 initial Pb2+ concentration at pH 5.0 and 318 K. Adsorption data were well described by the Langmuir model and the thermodynamic constants of the adsorption process: G , H and S were evaluated as 19.4 kJ mol1 (at 318 K), 25.73 kJ mol1 and 0.113 J mmol1 K1 , respectively. The pseudo-rst-order and pseudo-second-order kinetic models were applied to test the experimental data. The pseudo-second-order kinetic model provided the best correlation of the experimental data compared to the pseudo-rst-order model. 2006 Elsevier B.V. All rights reserved.
Keywords: Adsorption; Lead; Thermodynamic; Kinetic; Iron oxide

1. Introduction Heavy metals are generally recognized to be a threat toward humans and ecosystems because of their high potential toxicity. They could not be biologically decomposed into harmless materials and, to matters worse, were accumulated in the organisms [1]. Lead, as well as mercury, cadmium, chromium and arsenic, is in the group of serious hazardous heavy metals. Drinking those that contain Pb2+ ions for a long term, even if in a very low concentration, could lead to a wide range of spectrum health problems, such as renal failure, coma, nausea, cancer, convulsions and subtle effects on metabolism and intelligence [2,3]. Several methods have evolved over the years on the removal of heavy metals present in water and wastewater. These are chemical precipitations, conventional coagulation, reverse osmosis, ion exchange and adsorption. One of which, adsorption method, is simple and cost-effective, thus has been widely used [47]. Among various absorbents, adsorption onto activated carbon has proven to be one of the most effective and reliable physicochemical treatment methodologies [811]. However, among

Corresponding author. Tel.: +886 6 2757575x62636; fax: +886 6 2344496. E-mail address: yhhuang@ccmail.ncku.edu.tw (Y.-H. Huang).

commercially available activated carbons have two major shortcomings that considerably constrain their practical usefulness on a large scale. The rst is that the commercially available activated carbons are too expensive for practical use. Bailey et al. [12] have presented an interesting review, which focuses on the potential of a wide variety of low cost sorbents for heavy metals. According to these authors, a low cost sorbent can be assumed if it requires less prior processing, is abundant in nature, or is either a by-product or waste material from another industry. These materials could be alternatives for expensive treatment processes. The second is that the activated carbons have only single one adsorption function. They could not be used as the chemical reaction catalyst for organic compounds. Recently, there is growing interest in low-cost, highsurface-area materials, especially metal oxides and their unique applications, including not only adsorption but also chemical catalysis. Iron oxide has a relatively high surface area and charge; many researchers have applied iron oxide as adsorbent to treat heavy metals and organic compounds from wastewater [1317]. This study explores an activated alumina-supported iron oxide-composite material (FeAA), which is a by-product of the FBR-Fenton reaction [18,19], for use in the treatment of the bioefuent of tannery wastewater from a dyeing/nishing plant in Taiwan. The FeAA has successfully been as heterogeneous

1383-5866/$ see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.seppur.2006.10.023

24

Y.-H. Huang et al. / Separation and Purication Technology 55 (2007) 2329

photoassisted Fenton catalyst for degradation azo dye at neutral pH 7.0 in our previous study [20]. The properties of the FeAA are summarized in Table 1. In this work, the tests and uses of FeAA as adsorbent for the removal of Pb2+ from water are carried out. Furthermore, we will investigate the thermodynamic and kinetic of the Pb2+ absorption onto FeAA, in order to obtain the thermodynamic parameters, to establish the adsorption rate equation and to assess the FeAA as adsorbent to treatment of heavy metals from wastewater. 2. Material and methods 2.1. Adsorbent (FeAA) and solutions preparation A novel and low cost adsorbent, iron oxide on an activated alumina support, was developed in the the FBR-Fenton reaction [21]. Deionized and doubly distilled water were used throughout this investigation. By dissolving lead nitrate in deionized water, we obtained the lead containing stock solution (1000 mg dm3 ), which was further diluted to the required concentrations before being used. 2.2. Characterization of FeAA Physico-chemical characteristics of the FeAA were determined using standard procedures. Some properties of the FeAA are shown in Table 1. Furthermore, X-ray diffraction (XRD), Fourier transform infrared spectroscope (FTIR) and scanning electron microscope (SEM) analyses of FeAA were carried out in this study. XRD powder diffraction measurement of FeAA was performed on a powder diffractometer (Rigaku RX III) equipped with a Cu K radiation. The accelerating voltage and current were 40 kV and 20 mA. A small portion of FeAA was nely ground and then pressed (in vacumm) in the form of a disc using spectroscopically pure dry KBr. The FTIR spectrum was recorded at room temperature using a Bio-Rad FTS-40A. The morphology of the activated alumina grain support and FeAA were described using Hitachi S-4100 and Philips XL-40FEG SEM. The BrunauerEmmettTeller (BET) surface area and porosity of the iron oxide were obtained from the data of the isotherms. The surface area of the iron oxide was calculated using the BET equation.
Table 1 Properties of the FeAA (Hsueh et al. [20]) Parameter Total iron content of FeAA Bulk density (g cm3 ) Absolute (true) density (g cm3 ) Specic surface area (m2 g1 ) Total pore volume (cm3 g1 ) Cation exchange capacities (meq g1 ) pKa1 pKa2 pHpzc (point of zero charge)a
a

2.3. Batch experimental programme To evaluate the thermodynamic properties, we rst prepared various solutions with initial Pb2+ concentration ranging from 0.1 to 0.8 mM (100 ml, pH 5), and then added 0.2 g FeAA to each solution. These samples were then mounted on a shaker and shaken continuously for 48 h at 288, 308 and 318 K, respectively. The suspensions were ltered using a 0.25 mm membrane, and the ltrates were immediately measured using an atomic adsorption spectroscopy (Hitachi Z-6100). The differences between the initial and the equilibrium Pb2+ concentrations determine the amount that Pb2+ adsorbed by FeAA. The Adsorption kinetics experiments were performed on Jar Test at a constant speed of 150 revolutions per minute (rpm). The samples were prepared by adding 5.0 g of FeAA into 1000 ml solution (pH 5.0), and the Pb2+ concentrations were 0.1, 0.2, 0.4 and 0.8 mM, separately, at 300 1 K. Samples were withdrawn and ltered using a 0.25 mm membrane and then analyzed by an atomic adsorption spectroscopy. The amount of sorbed metal per gram of adsorbent qt at time t was calculated as follows [22]: qt = C0 Ct , mads (1)

where C0 and Ct are the Pb2+ concentration in liquid phase at the initial and any time t (mmol dm3 ), respectively, and mads is the adsorbent amount in the solution (g dm3 ). 3. Results and discussion 3.1. Characterization of the FeAA adsorbent Fig. 1 depicts the XRD patterns of the activated alumina grain support and FeAA. The XRD data were analyzed with

Value (g kg1 ) 304 1.43 2.66 170 0.12 0.47 6.03 8.67 7.35

pHpzc = (pKa1 + pKa2 )/2.

Fig. 1. X-ray diffraction pattern of the: (a) activated alumina grain support and (b) FeAA showing the intensities in the region 2 = 1070 .

Y.-H. Huang et al. / Separation and Purication Technology 55 (2007) 2329

25

the activated alumina grain support and FeAA of iron and aluminum using Joint Committee on Powder Diffraction Standards (JCPDS) diffraction les. The JCPDS data of oxide and oxyhydroxides of aluminum and iron were compared. The main diffraction peaks of the activated alumina grain support (Fig. 1a) at 2 = 37.7 , 45.8 and 66.8 were carefully compared with the standard for aluminum oxide (Al2 O3 -le number 77-0396). Accordingly, the peaks of activated alumina grain support could be identied with the phase Al2 O3 . The main diffraction peaks of the FeAA (Fig. 1b) at 2 = 21.2 , 36.6 and 53.2 were carefully compared with the standard for goethite ( -FeOOH le number 81-0464). Therefore, the peaks of FeAA could be identied with the phase -FeOOH. However, it was noticed that the XRD pattern of FeAA appears from the very weak diffraction intensities obtained in the region of 2 = 1070 that there should be very small amount of crystalline goethite present in the FeAA. The FTIR spectrum obtained from FeAA sample is provided in Fig. 2. From the FTIR spectrum, it is clear that the peaks that occur in the region around 795, 890, 1079, 1362, 1508, 1629 and 3392 cm1 . When these peaks are eliminated from analysis in the FeAA FTIR spectrum, only certain other signicant peaks require to be identied. It has been pointed out by them that the absorption band at higher wave number region is due to OH stretching and at lower wave number is because of FeO lattice vibration [23]. The broad band seen at 30003500 cm1 is due to hydration of the FeAA. It is known that OH stretching leads to a strong peak between 3000 and 3700 cm1 , while

Fig. 2. FTIR spectrum from the FeAA.

OH bending to a medium band between 1200 and 1500 cm1 [24,25]. Therefore, this implies hydration in the FeAA. Furthermore, it can be noted from Fig. 2 that the presence of -FeOOH is conrmed by the appearance of peaks at 890 cm1 [26] and 795 cm1 [23]. Additionally, Fig. 3 depicts the morphology of the original activated alumina grain support (Fig. 3a and b) and FeAA (Fig. 3c and d). Compare Fig. 3a with Fig. 3c [20], and

Fig. 3. Scanning electron micrographs of original activated alumina grain support: (a) 100; (b) 3500 and FeAA: (c) 100; (d) 12,000.

26

Y.-H. Huang et al. / Separation and Purication Technology 55 (2007) 2329

Fig. 4. Adsorption isotherms of Pb2+ onto FeAA at different temperatures.

it was smoother after the reaction in the FBR had proceeded for 3 months. Moreover, Fig. 3b and d shows details of the original activated alumina grain support with a typical sheet-like structure and the surface of the FeAA with ne granular structure. The specic surface areas of the original activated alumina grain support and FeAA are 127 and 170 m2 g1 , respectively. 3.2. Adsorption isotherm Fig. 4 shows the adsorption isotherms of Pb2+ on FeAA at 288, 308 and 318 K. Equilibrium data can be analyzed using commonly known adsorption isotherms, which provide the basis for the design of adsorption systems. The most widely used isotherm equation for modelling of the adsorption data is the Langmuir equation, which is valid for monolayer sorption onto a surface with a nite number of identical sites and is given by Eq. (2). qe = KL qm Ce , 1 + K L Ce (2)

Fig. 5. Linearized: (a) Langmuir and (b) Freundlich isotherm models for Pb2+ adsorption by the FeAA at different temperatures.

where KL is the adsorption equilibrium constant including the afnity of binding sites (mM1 ), qm the maximum adsorption capacity (mmol g1 ) and qe is the amount of sorbed Pb2+ at equilibrium (mmol g1 ). It represents a practical limiting adsorption capacity when the surface is fully covered with Pb2+ . qm and KL can be determined from the linear plot of Ce /qe versus Ce [27,28]. The Freundlich model is an empirical equation based on sorption on a heterogenous surface. It is given as:
1/n qe = KF Ce ,

linear regression. According to Fig. 5b, the values of KF and n are obtained similarly. The results are presented in Table 2 with the correlation coefcients (R2 ). It can be seen from Table 2 that the Langmuir model yields a better t than the Freundlich model by comparing the results of correlation coefcient values at different temperatures. The values of Freundlich constants increased with increasing temperature and the highest KF value was reported as 11.15 at 318 K. All n values were found high enough for adsorption (>1.0). Values of qm and KL at different temperatures are also tabulated in Table 2. The maximum capacity, qm dened the total capacity of FeAA for Pb2+ adsorption
Table 2 Parameters of Langmuir and Freundlich adsorption isotherm models for Pb2+ on FeAA T (K) Langmuir qm 288 308 318 0.096 0.11 0.14 KL 17.5 35.45 48.13 R2 0.986 0.997 0.996 Freundlich KF 6.15 8.37 11.15 n 4.49 4.76 4.83 R2 0.983 0.981 0.971

(3)

where KF and n are the Freundlich constants related to the adsorption capacity and adsorption intensity, respectively. The Freundlich equation can be linearized by taking logarithms and constants can be determined [28]. Linear plots of Ce /qe versus Ce and ln qe versus ln Ce are shown in Fig. 5a and b. For each isotherm in Fig. 5a, the values of qm and KL were calculated from experimental data through

Y.-H. Huang et al. / Separation and Purication Technology 55 (2007) 2329 Table 3 G values for adsorption of Pb2+ on FeAA at different temperatures Temperature (K) 288 308 318 KL (mM1 ) 17.50 35.45 48.13 G (kJ mol1 ) 6.85 9.14 10.24

27

and increased with increasing temperature. Its maximum value was determined as 0.14 mmol g1 at 318 K. The increase of sorption equilibrium constant with temperature showed that there was a chemical interaction between adsorbent and adsorbate. Furthermore, entropy and energy factors should be considered in order to determine what processes will occur spontaneously in engineering practice. The Gibbs free energy indicates the degree of spontaneity of the adsorption process and higher negative value reects a more energetically favorable adsorption. The Gibbs free energy change of adsorption is dened as G = RT ln KL , (8.314 J mol1 K1 ) (4)

Fig. 7. Effect of contact time on Pb2+ adsorption rate for different concentrations (pH 5.0, at 300 1 K).

where R is the universal gas constant and T is the absolute temperature in Kelvin [29]. The equilibrium constant may be expressed in terms of standard enthalpy change of adsorption ( H ) and entropy change of adsorption ( S ) as a function of temperature. The relationship between the KL and temperature is given by the vant Hoff equation: H S + , (5) RT R H and S can be obtained from the slope and intercept of the plot of ln KL versus 1/T [30]. The equilibrium constants obtained from the Langmuir model at 288, 308 and 318 K were used to determine the Gibbs free energy changes. Table 3 shows the Gibbs free energy values for the adsorption process. The H and S were determined as 25.73 kJ mol1 and 0.113 J mmol1 K1 from Fig. 6, respectively. A positive standard enthalpy change of 25.73 kJ mol1 obtained in this study indicates that the adsorption of Pb2+ by the FeAA adsorbent is endothermic, which fact is evidenced ln KL =

by the increase in the adsorption of Pb2+ with temperature. A negative change in adsorption standard free energy reveals that the adsorption reaction is a spontaneous process [29,30]. The positive standard entropy change may be caused the release of water molecules in the ion exchange reaction between the adsorbate and the functional groups on the surfaces of the FeAA adsorbent. These results were similar to previous literature [31] which indicated that although the standard enthalpy change for adsorption of very different adsorbate onto distinct adsorbent covers a wide range (85 to +160 kJ mol1 ), the standard free energy change at 30 C remain within 30 kJ mol1 . Restated, the enthalpy and entropy contributions for driving the adsorption process are largely compensated each other for very different adsorbate/adsorbent systems. This phenomenon needs further investigation since if it were correct, the reason why a universal correlation could exist between the corresponding enthalpy change and entropy change following adsorption remains unclear [31]. 3.3. Adsorption kinetic study Adsorption kinetics, demonstrating the solute uptake rate, is one of the most important characters which represent the adsorption efciency of the FeAA and therefore, determines their potential applications. According to Fig. 7, the Pb2+ adsorption rates increase dramatically in the rst 2 h for various initial concentrations, and they reach equilibrium gradually at about 4, 8, 12 and 36 h, corresponding to Pb2+ initial concentrations of 0.1, 0.2, 0.4 and 0.8 mM, respectively. To analyze the adsorption rate of Pb2+ onto the FeAA, the pseudo-rst-order equation of Lagergren (6) and the pseudo-second-order rate Eq. (7) were evaluated based on the experimental data [11,32], log q e qt k1 t , = qe 2.3 (6) (7)

Fig. 6. ln K vs. 1/T plot.

t 1 t = + , 2 qt 2k2 qe qe

28

Y.-H. Huang et al. / Separation and Purication Technology 55 (2007) 2329

oretical qe values were closed to the experimental qe values (Table 4). 4. Conclusion An Al2 O3 -supported iron oxide was developed for the removal of Pb2+ from aqueous solutions in this study. The surface of the FeAA adsorbent was characterized by XRD and FTIR and was identied as -FeOOH. Furthermore, the ability of FeAA to bind Pb2+ was tested using equilibrium, kinetic and thermodynamic aspects. The Langmuir and Freundlich adsorption models were used to express the sorption phenomenon of the sorbate. The equilibrium data were well described by the Langmuir model. In addition, the negative Gibbs standard free energy changes conrmed the spontaneous nature adsorption process. The positive standard enthalpy and the entropy changes showed the endothermic nature and irreversibility of Pb2+ adsorption, respectively. The pseudo-second-order rate model accurately describes the kinetics of adsorption. Acknowledgement The authors thank the National Science Council of the Republic of China for nancially supporting this NSC94-2211E-006-032. References
[1] C.O. Adewunmi, W. Becker, O. Kuehnast, F. Oluwole, G. D rer, Accuo mulation of copper, lead and cadmium in freshwater snails in southwestern Nigeria, Sci. Total Env. 193 (1996) 6973. [2] M.N. Rashed, Lead removal from contaminated water using mineral adsorbents, Environmentalist 21 (2001) 187195. [3] Y.H. Li, Z. Di, J. Ding, D. Wu, Z. Luan, Y. Zhu, Adsorption thermodynamic, kinetic and desorption studies of Pb2+ on carbon nanotubes, Water Res. 39 (2005) 605609. [4] D. Ghosh, K.G. Bhattacharyyra, Adsorption of methylene blue on kaolinite, J. Appl. Clay Sci. 20 (2002) 295300. [5] M.J.S. Yabe, E. Oliveira, Heavy metals removal in industrial efuents by sequential adsorbent treatment, Adv. Env. Res. 7 (2003) 263272. [6] M. Erdem, A. Ozverdi, Lead adsorption from aqueous solution onto siderite, Sep. Purif. Technol. 42 (2005) 259264. [7] S.R. Shukla, R.S. Pai, A.D. Shendarkar, Adsorption of Ni(II), Zn(II) and Fe(II) on modied coir bres, Sep. Purif. Technol. 47 (2006) 141 147. [8] M.O. Corapcioglu, C.P. Huang, The adsorption of heavy metals onto hydrous activated carbon, Water Res. 21 (1987) 10311044. [9] B. Gu, J. Schmitt, Z. Chen, L. Liang, J.F. McCarthy, Adsorption and desorption of nature organic matter on iron oxide: mechanisms and model, Env. Sci. Technol. 28 (1994) 3846. [10] S. Akhtar, R. Qadeer, Active carbon as an adsorbent for lead ions, Adsorp. Sci. Technol. 15 (1997) 815824. [11] Y.S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process Biochem. 34 (1999) 451465. [12] S.E. Bailey, T.J. Olin, R.M. Bricka, D.D. Adrian, Review of potentially low-cost sorbents for heavy metals, Water Res. 33 (1999) 24692479. [13] B.B. Johnson, Effect of pH, temperature, and concentration on the adsorption of cadmium on goethite, Env. Sci. Technol. 24 (1990) 112118. [14] A. Manceau, L. Charlet, The mechanisms of selanate adsorption on goethite and hydrous ferric oxide, J. Colloid Interface Sci. 168 (1994) 8793.

Fig. 8. Test of: (a) the pseudo-rst-order and (b) the pseudo-second-order kinetic equation for adsorption of different concentrations of Pb2+ by FeAA (pH 5.0, at 300 1 K).

where k1 is the Lagergren rate constant of adsorption (h1 ) and k2 is the pseudo-second-order rate constant of adsorption (g mg1 h1 ). qe and qt are the amounts of metal ion sorbed (mg g1 ) at equilibrium and at time t, respectively. Plots of log(qe qt )/(qe ) versus t and t/(qt ) versus t are shown in Fig. 8a and b. Clearly, it can be seen from Fig. 8a and b that the pseudosecond-order kinetic model provided a good correlation for the adsorption of Pb2+ onto FeAA in contrast to the the pseudorst-order model. In addition, the correlation coefcients of the pseudo-second-order kinetic model are very high and the theTable 4 Kinetic parameters of Pb2+ adsorbed on FeAA at different initial Pb2+ concentrations Initial Pb2+ concentration (mM) 0.1 0.2 0.4 0.8 qe,exp (mg g1 ) 4.22 8.37 16.87 30.2 Pseudo-second-order kinetic model qe,cal (mg g1 ) 4.28 8.57 17.56 31.51 k2 (g mg1 h1 ) 0.1999 0.0557 0.0141 0.0061 R2 0.9999 0.9999 0.9992 0.9975

Y.-H. Huang et al. / Separation and Purication Technology 55 (2007) 2329 [15] S. Fendorf, M.J. Eick, P. Grossl, D.L. Sparks, Arsenate and chromate retention mechanisms on goethite. 1. Surface structure, Env. Sci. Technol. 31 (1997) 315319. [16] P. Venema, T. Hiemstra, P.G. Weidler, W.H. Van Riemsdijk, Intrinsic proton afnity of reactive surface groups of metal (hydr)oxides: application to iron (hydr)oxides, J. Colloid Interface Sci. 198 (1998) 252295. [17] S.G.J. Heijman, A.M. Paassen, W.G.J. Meer, R. Hopman, Adsorption removal of natural organic matter during water treatment, Water Sci. Technol. 40 (1999) 183190. [18] S. Chou, C. Huang, Y.H. Huang, Effect of Fe2+ on catalytic oxidation in a uidized bed reactor, Chemosphere 39 (1999) 19972006. [19] S. Chou, C. Huang, Y.H. Huang, Heterogeneous and homogeneous catalytic oxidation by supported -FeOOH in a uidized-bed reactor: kinetic approach, Env. Sci. Technol. 35 (2001) 12471250. [20] C.L. Hsueh, Y.H. Huang, C.C. Wang, C.Y. Chen, Photoassisted Fenton degradation of nonbiodegradable azo-dye (Reactive Black 5) over a novel supported iron oxide catalyst at neutral pH, J. Mol. Catal. A: Chem. 245 (2006) 7886. [21] Y.H. Huang, G.H. Huang, S. Chou, H.S. You, S.H. Perng, Process for chemically oxidizing wastewater with reduced sludge production, USA Patent, US6143182 (2000). [22] N. Chiron, R. Guilet, E. Deydier, Adsorption of Cu(II) and Pb(II) onto a grafted silica: isotherms and kinetic models, Water Res. 37 (2003) 30793086.

29

[23] R.M. Cornell, U. Schwertmann, The Iron Oxides, Wiley-VCH, New York, 2003. [24] R.A. Nyquist, R.A. Kagel (Eds.), IR Spectra of Inorganic Compounds, Academic Press, New York, 1971. [25] D.A. Skoog, J.J. Leary, Principles of Instrumental Analysis, fourth ed., Harcourt Brace College Publishers, New York, 1992, pp. 252 309. [26] T. Misawa, K. Asami, K. Hashimoto, S. Shimodaira, The mechanism of atmospheric rusting and the protective amorphous rust on low alloy steel, Corros. Sci. 14 (1974) 279289. [27] I. Langmuir, The constitution and fundamental properties of solids and liquids, J. Am. Chem. Soc. 38 (1916) 22212295. [28] J.M. Smith, Chemical Engineering Kinetics, third ed., McGraw-Hill, Singapore, 1981. [29] J.M. Smith, H.C. Van Ness, Introduction to Chemical Engineering Thermodynamics, fourth ed., Mcgraw-Hill, Singapore, 1987. [30] Z. Aksu, Determination of the equilibrium, kinetic and thermodynamic parameters of the batch biosorption of nickel(II) ions onto Chlorella vulgaris, Process Biochem. 38 (2002) 8999. [31] A. Ramesh, D.J. Lee, J.W.C. Wong, Adsorption equilibrium of heavy metals and dyes from wastewater with low-cost adsorbents: a review, J. Chin. Inst. Chem. Eng. 36 (2005) 203222. [32] S. Lagergren, Zur theorie der sogenannten adsorption geloster stoffe, Kungliga Svenska Vetenkapsakademiens, Handlingar 24 (1898) 139.

You might also like