Hostname: page-component-7bb8b95d7b-dtkg6 Total loading time: 0 Render date: 2024-09-27T09:10:30.119Z Has data issue: false hasContentIssue false

Removal Efficiency of Basic Blue 41 by Three Zeolites Prepared from Natural Jordanian Kaolin

Published online by Cambridge University Press:  01 January 2024

Mousa Gougazeh*
Affiliation:
Geology Department, Taibah University, P.O. Box: 30002, 41447, Madinah, Saudi Arabia Natural Resources and Chemical Engineering, Tafila Technical University, P.O. Box 179, 66110, Tafila, Jordan
Fethi Kooli
Affiliation:
Community College, Taibah University- Al-Mahd Branch, 44112, Mahd Al-Dahb, Saudi Arabia
J.-Ch. Buhl
Affiliation:
Institute of Mineralogy, Leibniz University Hannover, Callinstr. 3, D-30167, Hannover, Germany
*
*E-mail address of corresponding author: dr_eng_mhag@yahoo.com
Rights & Permissions [Opens in a new window]

Abstract

The conventional method of zeolite synthesis involves an expensive hydrothermal step whereby a mixture of a metakaolinite, sodium hydroxide, and water is preactivated by thermal treatment between 400°C and 1000°C. The objective of the current study was to determine whether Jordanian kaolinite could be converted to zeolite materials without thermal pre-activation. The alkaline hydrothermal transformation of kaolinite into hydroxysodalite (HS) was achieved, then followed by a reaction with citric acid and solid sodium hydroxide to obtain Zeolite A, or by adding solid Na2SiO3 to prepare zeolite X. These materials were tested for their ability to serve as removal agents for Basic Blue 41 (BB-41) dye from artificially contaminated water, at concentrations ranging from 25 to 1000 mg/L. The maximum removal capacities were estimated using the Langmuir model, with a value of 39 mg/g for hydroxysodalite. Zeolite-X achieved the lowest value (19 mg/g). The feasibility of BB-41 removal was deduced from the Freundlich model for the zeolites studied. The reported low-cost method is proposed as an alternative way to reduce the cost of synthesizing zeolite, and the materials were shown to be potential candidates for the removal of BB-41 dye.

Type
Article
Copyright
Copyright © Clay Minerals Society 2019

Introduction

Synthetic zeolites have become increasingly important due to their wide range of chemical and physical properties; over recent decades they have been used as adsorbents, molecular sieves, membranes, ion exchangers, and catalysts (Rhodes Reference Rhodes2010; Zaarour et al. Reference Zaarour, Dong, Naydenova, Retoux and Mintova2014; Li et al. Reference Li, Li and Yu2017). Zeolite structures consist of three-dimensional frameworks of SiO4 and AlO4 tetrahedra. The aluminum ion (Al) is sufficiently small that it can occupy the position in the center of the tetrahedron of four oxygen atoms and the isomorphous replacement of Si4+ by Al3+ produces a negative charge in the lattice. The net negative charge is balanced by an exchangeable cation [sodium (Na+), potassium (K+), or calcium (Ca2+)]. These cations are exchangeable with certain other cations in solution (Barer, Reference Barer1987). Synthesized zeolites are used commercially due to their purity as crystalline products and the uniformity of particle size. The hydrothermal process is expensive, however, and fails to eliminate negative environmental impacts (Abdullahi et al. Reference Abdullahi, Harun and Othman2017). The possible use of natural resources and manufacturing wastes as raw materials for the synthesis of zeolite has been studied (Querol et al. Reference Querol, Moreno, Umana, Alastuey, Hernandez, Lopez-Soler and Plana2002; Hiraki et al. Reference Hiraki, Nosaka, Okinaka and Akiyama2009; Johnson & Arshad Reference Johnson and Arshad2014). The use of waste materials helps to reduce production costs, to reduce environmental problems, and turns waste materials into more useful and valuable products (Dubey et al. Reference Dubey, Goyal, Mishra, Mishra and Clark2013). A potential candidate is raw kaolinite. Kaolinite, a 1:1 clay mineral, has the general chemical formula (Al2Si2O5(OH)4); it can be mined in a relatively pure form from kaolin deposits (including some large ones in Jordan), making it an inexpensive source of alumina (Al2O3) and silica (SiO2) (Gougazeh & Buhl Reference Gougazeh and Buhl2010). Kaolinite has been used as the source of aluminum and silicon for the synthesis of several types of zeolite (Breck Reference Breck1974; Barer Reference Barer1982; Costa et al. Reference Costa, De Lucas, Uguina and Ruiz1988; Rees & Chandrasekhar Reference Rees and Chandrasekhar1993; Zhao et al. Reference Zhao, Deng, Harsh, Flury and Boyle2004; Heller-Kallai & Lapides Reference Heller-Kallai and Lapides2007; Ltaief et al. Reference Ltaief, Siffert, Poupin, Fourmentin and Benzina2015). Hydroxysodalite rather than other zeolites appears to be the primary product obtained after alkaline hydrothermal conversion of kaolinite, with or without intermediate phases (Madani et al. Reference Madani, Aznar, Sanz and Serratosa1990; Buhl et al. Reference Buhl, Hoffmann, Buckermann and Muller-Warmuth1997). During the synthesis of zeolites, the base material, kaolinite, must be heated at between 550 and 900°C before it is reacted with an alkali solution. In a more recent study, however (Gougazeh & Buhl Reference Gougazeh and Buhl2014), the raw kaolinite was shown to convert to a zeolite material (hydroxysodalite, HS) without the thermal pre-activation step. This method is considered to be an energy-saving, less expensive process, that does not generate solid or liquid residues that harm ecological systems.

Synthetic dyes or pigments are used widely in textile industries to color-certain products, creating environmentally hazardous waste. The discharge of dyes from these industries into natural streams and rivers poses severe problems, as dyes are toxic to aquatic life and damaging to the aesthetic nature of the environment (Baughman & Perenich Reference Baughman and Perenich1988; Chung & Cerniglia Reference Chung and Cerniglia1992; de Sousa et al. Reference de Sousa, de Moraes, Matos Lopes, Montagnolli, de Angelis and Bidoia2012; Song et al. Reference Song, Chen, Hu and Richards2009). Many methods have been used in dye removal, such as sedimentation, chemical treatment, oxidation, biological treatment, electrochemical methodology, and adsorption (Gupta & Suhas Reference Gupta and Suhas2009). The adsorption technique using many types of adsorbents is still the most favored method for the removal of contaminants from wastewaters due to its efficiency, potential for high removal capacity, low operational cost, and insensitivity to toxic substances (Nandi et al. Reference Nandi, Goswami and Purkait2009; Li et al. Reference Li, Dong, Wu, Peng and Kong2011; Patel, Reference Patel2012; Kyzas & Kostoglou, Reference Kyzas and Kostoglou2014; Yagub et al. Reference Yagub, Kanti Sen, Afroze and Ang2014). Activated carbon is the most popular adsorbent and has been used with great success. However, use of this method is often limited due to high cost, making the method less popular (Huling et al. Reference Huling, Jones and Lee2007; Robinson et al. Reference Robinson, McMullan, Marchant and Nigam2001). Natural zeolites have been observed to be highly effective in reducing dye concentrations, since they have large cation exchange capacities, large surface areas, and large contents of residual carbon (Dubey et al. Reference Dubey, Goyal, Mishra, Mishra and Clark2013; Karadag et al. Reference Karadag, Akgul, Tok, Erturk, Kaya and Turan2007; Ng et al. Reference Ng, Zou, Mintova and Suib2013).

Synthetic zeolites have exchangeable cations which are relatively innocuous (Na+, Ca2+, and K+) and tuneable pore sizes that make them particularly suitable for removing undesirable materials from industrial or agricultural effluent (Wang & Peng Reference Wang and Peng2010; Hernández-Montoya et al. Reference Hernández-Montoya, Pérez-Cruz, Mendoza-Castillo, Moreno-Virgen and Bonilla-Petriciolet2013; Yuna Reference Yuna2016). From these synthetic materials, zeolite A (ZA) and zeolite X (ZX) were selected for study. Their synthesis, however, is achieved generally from high-cost chemicals, thus limiting their commercial application (Schwanke et al. Reference Schwanke, Balzer, Pergher, Martínez, Kharissova and Khatisov2017). The purpose of the present study was to test the hypothesis that zeolites (hydroxysodalite, zeolite A, and zeolite X) could be synthesized using raw kaolinite without prior thermal treatment, and to determine the utility of the synthesized materials by measuring how effective they are at removing BB-41 from artificially contaminated water.

Experimental

Materials

Natural kaolinite was collected from the Jabal Al-Harad kaolin deposit (Batn El-Ghoul area). It is located in southern Jordan, 280 km south of Amman, ~70 km southeast of Ma’an city, and 4 km to the east of the Ma’an–Mudawwara road at latitude 29°37′52″N and longitude 35°20′57″E (Fig. 1; Gougazeh & Buhl Reference Gougazeh and Buhl2010). According to Rietveld refinement of X-ray diffraction (XRD) data, the material used in this study for synthesis reactions is ~80 wt.% kaolinite and ~16 wt.% quartz and has a chemical composition of 53.86 wt.% SiO2, 0.74 wt.% TiO2, 32.45 wt.% Al2O3, 0.65 wt.% Fe2O3, 0.08 wt.% MgO, 0.13 wt.% CaO, 0.06 wt.% Na2O, 0.54 wt.% K2O and 11.21% LOI (Gougazeh & Buhl Reference Gougazeh and Buhl2010).

Fig. 1 Location map of the Jabal Al-Harad, Jordan, kaolin deposit

Sodium hydroxide (NaOH, 99%) as pellets was purchased from Merck Chemical Company (Darmstadt, Germany), citric acid and sodium silicate solution (Na2Si2O5) were provided by Fluka Chemie (Buchs, Switzerland), and analytical-grade Basic Blue 41 (BB-41) was purchased from Sigma–Aldrich (Munich, Germany) and used without further purification. BB-41 has the molecular formula C20H26N4O6S2 (molecular weight = 482.57 g/mol) with a color index number 11105. The chemical structure of BB-41 is shown in Fig. 2 (Tajul Islam et al. Reference Tajul Islam, Aimone, Ferri and Rovero2015). Distilled water was used in the standard purification methods.

Fig. 2 Chemical structure of the dye Basic Blue 41

Preparation of Zeolites

The zeolite A was prepared following the method described by Gougazeh & Buhl (Reference Gougazeh and Buhl2014). 1.0 g of raw kaolinite (RK) was treated in 8 M NaOH at 120°C for 20 h. The product was washed three times using distilled water and dried overnight at 80°C. Then, 1 g of the product, hydroxysodalite, was mixed with diluted citric acid (20 mL 1 M) and heated to ~70°C for 2–4 h. To this solution, solid NaOH powder (3.2 g) was added with vigorous stirring at room temperature for homogenization. The resulting zeolite precursor gel was placed in an autoclave (50 mL) and kept at ~100 °C for 3 h. After hydrothermal treatment, the product was filtered and washed three times with distilled water to remove excess alkali, then oven-dried at 80 °C for 4 h.

The process to prepare zeolite X consisted, firstly, of the alkaline hydrothermal transformation of kaolinite into hydroxysodalite, as described above for zeolite A, followed by the same treatment with diluted citric acid (20 mL 1 M) and heating to ~70°C for 2–4 h. To this solution, 3.2 g (8.0 M) of solid NaOH powder and 0.5 g of solid Na2SiO3 powder were added with vigorous stirring at room temperature for homogenization; the pH of the aluminosilicate solution was adjusted to ~12 to obtain a gel; it was then placed in an autoclave (50 mL) and kept at ~80°C for 16 h. After hydrothermal treatment, the product was filtered and washed three times with distilled water to remove excess alkali. Next, the samples were oven-dried at 80°C for 4 h.

Batch Mode Adsorption Studies

The adsorption experiments were performed in a batch process (Kooli et al. Reference Kooli, Yan, Al-Faze and Al-Sehimi2015b). Preliminary experiments demonstrated that equilibrium was established in ~8 h, but the experiments were allowed to run for 18 h. To obtain adsorption isotherms, a fixed amount (0.100 g) of kaolinite or zeolite was mixed with 10 mL of dye solution of different concentrations from 25 to 1000 mg/L. The supernatant was collected by centrifugation for 10 min at 2380×g using a Labofuge 200 centrifuge. The equilibrium BB-41 concentrations were determined by using a Cary 100 Conc spectrophotometer from Varian (Sydney, Australia) at 610 nm.

The removal capacity of dye was calculated as follows:

(1) q e = C i C e × V W

and the removal percentage of dye was calculated as

(2) R % = C i C e C i × 100

where q e (mg/g) is the amount of dye adsorbed, C i (mg/L) is the initial dye concentration, C e is the dye concentration at equilibrium, W (g) is the mass of adsorbent added to the BB-41 solution, and V (L) is the volume of the dye solution.

Regeneration Process

The spent materials were regenerated as reported previously (Kooli et al. Reference Kooli, Yan, Al-Faze and Al-Sehimi2015b). The ZX, ZA, or HS samples were added separately to 10 mL of BB-41 solution (200 ppm) overnight. Each solid was collected by centrifugation and was then dispersed into 10 mL of Co(NO3)2 ·6H2O aqueous solution. The Co2+cations served as a homogeneous catalyst. The oxidant, 10 mg of oxone (2KHSO5 ·KHSO4 ·K2SO4) was added to the mixture to degrade the adsorbed BB-41, via sulfate radical oxidation. The mixture was stirred for 30 min, separated by centrifugation, washed four or five times with deionized water, and then recycled for the next run. The same oxone and Co(NO3)2 aqueous solution was used throughout the recycle runs.

Characterization

The identification of mineralogical phases comprising the raw kaolinite and the zeolite products was achieved using XRD patterns from a Bruker AXS D4 ENDEVOR diffractometer (Hannover, Germany) with Ni-filtered CuKα radiation at 40 kV and 40 mA. The powder samples were mounted on the sample holder with very little pressure so as to minimize preferred orientation of particles in the sample. The measurements were carried out with a step width of 0.03°2θ and a measuring time of 1 s per step. The powder data were evaluated with the Stoe WinXPOW software package. Fourier-transform infrared spectroscopy (FTIR) measurements were performed using a Bruker IFS66v FTIR spectrometer in the 4000–370 cm−1 region with samples as KBr pellets. A JEOL JSM-6390A scanning electron microscope (SEM) equipped with a field electron gun (20 kV) was used to study the morphology of the samples. The textural characteristics, such as specific surface area (S BET), total pore volume (TPV), and average pore diameter of the different samples, were determined from nitrogen adsorption isotherms using a Micromeritics (Norcross, Georgia, USA) ASAP 2040. All samples were degassed at 150°C overnight prior to measurement. The concentrations of the BB-41 dye after treatment with the kaolinite and zeolites were measured by UV-Visible spectroscopy at 610 nm.

Results and Discussion

X-ray Diffraction

The powder XRD data indicated that the kaolinite exhibited characteristic reflections at 12.35°2θ (001) and 24.64°2θ (002) with additional ones related to quartz (Fig. 3a), and the H-S material showed several common reflections at 14.10 (110), 19.98 (200), 24.54 (211), 31.85 (310), and 34.99°2θ (222) (Fig. 3b), as reported in previous studies (Lee et al. Reference Lee, Han, Park, Park and Choy2003; Gougazeh & Buhl Reference Gougazeh and Buhl2014). The products synthesized products (ZA and ZX) exhibited high crystallinity and purity, as no impurities were detected. ZX exhibited several common reflections detected at 2θ values of 6.10o (111), 9.99o (220), 11.74° (311), 15.45° (331), 18.44° (511), 20.10° (440), 23.32° (537), 26.70° (642), 30.47° (660), 31.15° (751), 32.03° (840), and 33.62° (864) (Fig. 3c). ZA exhibited characteristic reflections located at 7.14° (100), 10.10° (110), 12.38° (111), 16.38° (210), 21.58° (300), 23.92° (311), 27.00° (321), 29.82° (410), and 34.08° (332) °2θ (Fig. 3d; Gougazeh & Buhl Reference Gougazeh and Buhl2014).

Fig. 3 Powder XRD patterns of the samples, including (a) Jordanian kaolinite (JK), (b) synthesized hydroxysodalite (HS), (c) synthesized zeolite X (ZX), and (d) synthesized zeolite A (ZA)

FTIR Spectroscopy

The structural changes that occurred during synthesis were monitored by FTIR spectroscopy (Fig. 4). The IR spectra show the prominent peaks of raw kaolinite (Fig. 3a): 3620–3700 cm−1 for -OH stretching, 1000–1120 cm−1 for Si–O stretching, 910–940 cm−1 for -OH deformation modes, and 400–550 cm−1 for Si–O–Si bending vibrations (Farmer Reference Farmer, Olphen and Fripiat1979; Gougazeh & Buhl Reference Gougazeh and Buhl2010). The doublet between 3694 and 3619 cm−1 is indicative of a well ordered kaolinite structure (Farmer Reference Farmer, Olphen and Fripiat1979; Russell Reference Russell and Wilson1987; Gougazeh & Buhl Reference Gougazeh and Buhl2010; Pentrak et al. Reference Pentrak, Madejová and Komadel2009, Reference Pentrak, Madejová, Andrejkovičová, Uhlík and Komadel2012). For the synthesized products (Fig. 4b, c, d), the bands with maxima at 3435 and 1655 cm−1 were characteristic of vibrations of functional groups of the OH type and were ascribed to water molecules in the zeolitic materials. Similar behavior was observed in previous studies on the synthesis of crystalline zeolites HS, X, and A (Flanigen et al. Reference Flanigen, Khatami and Szymanski1971; Baren et al. 1999a; Alkan et al. Reference Alkan, Hopa, Yilmaz and Guler2005; Garrido-Pedrosa et al. Reference Garrido Pedrosa, Souza, Melo and Araújo2006), while obtaining the highly crystalline zeolites HS, X, and A.

Fig. 4 FTIR spectra of the samples, including (a) Jordanian kaolinite, (b) synthesized zeolite X (ZX), (c) synthesized hydroxysodalite (HS), and (d) synthesized zeolite A (ZA)

The strong signal at ~1000 cm−1 characterized the asymmetrical T–O–T valence vibrations (T–Si, Al) of the aluminosilicate zeolite A framework, and the band at ~550 cm−1 represented the double ring vibration of the structure (Flanigen et al. Reference Flanigen, Khatami and Szymanski1971). The bands at ~664 cm−1 (symmetric stretching mode) and 460 cm−1 (T–O bending mode) corresponded to the internal vibrations of the TO4 tetrahedra (Fig. 3b, c, d). The broader signals at ~550 cm−1 and 1000 cm−1 of the reaction products characterized silica gel in the mixture with zeolite A, and were another indication of the formation of an amorphous material from the quartz content of the kaolinite (see the XRD paragraph). The additional vibrations in the 720–660 cm−1 range were in good agreement with the symmetric stretching modes of hydroxysodalite (729, 701, and 660 cm−1) reported by Flanigen et al. (Reference Flanigen, Khatami and Szymanski1971) (Fig. 3c, d). In the same manner, the band near 420 cm−1 and the intensive mode at 464 cm−1 also indicated the presence of hydroxysodalite in this sample (Fig. 3c); these findings were in good agreement with the reported values of the T–O bending modes of the hydroxysodalite framework structure (Flanigen et al. Reference Flanigen, Khatami and Szymanski1971). The detailed FTIR assignments for hydroxysodalite were summarized by Barnes et al. (Reference Barnes, Addai-Mensah and Gerson1999) and later by Zhao et al. (Reference Zhao, Deng, Harsh, Flury and Boyle2004). The broad bands at ~3435 cm−1 and 1655 cm−1 were attributed to zeolite water (Fig. 3c, d). The synthesized zeolite X showed adsorption bands at 972, 560, and 460 cm−1 related to external linkage of the zeolite structure (Somerset et al. Reference Somerset, Petrik, White, Klink, Key and Iwuoha2005). Furthermore, a strong medium band at ~1650 cm−1was attributed to the H2O deformation mode due to incomplete dehydration of the zeolite samples (Fig. 4b). Moreover, the observed single strong band at 3435 cm−1 was ascribed to the presence of hydroxyl groups in the faujasite (FAU) supercages (Garrido Pedrosa et al. Reference Garrido Pedrosa, Souza, Melo and Araújo2006). Ultimately, the intensities of these adsorption bands were proportionate to the purity of the samples, and were in agreement with the interpretation of the XRD results.

The band at 458 cm−1 was close to the band at 456 cm−1 (bending vibrations of the TO4) of FAU zeolite. The bands at 560 and 972 cm−1 corresponded to the 6-membered double-ring vibration: to the symmetric stretching and to the asymmetric stretching, respectively. The band at 557 cm−1 could represent the beginning of the crystallization of a zeolite with double rings (Gougazeh & Buhl Reference Gougazeh and Buhl2014).

Scanning Electron Microscopy

SEM photomicrographs were utilized to determine the microscale structure and morphology of the raw kaolinite and zeolite samples. The progress of the zeolitization reactions was observed through changes in the morphology of the starting material and synthesized products. The SEM photomicrograph of raw kaolinite consisted predominantly of the plate-like morphology and pseudo-hexagonal crystals characteristic of kaolinite (Gougazeh & Buhl Reference Gougazeh and Buhl2014; Fig. 5a). The occurrences of hydroxysodalite (HS) crystals growing at the surface of zeolite A and spherical morphologies corresponding to HS associated with cubic crystals of zeolite A can be clearly identified (Gougazeh & Buhl Reference Gougazeh and Buhl2014; Fig. 5b). The morphology of the synthesized samples can be observed in the images of zeolite A, with characteristic cubic morphology (Fig. 5c). In contrast, zeolite X exhibited well defined crystals (Fig. 5d).

Fig. 5 SEM images of: (a) of pseudohexagonal crystals of Jordanian kaolinite; (b) spherical (spheroidal) agglomerates of hydroxysodalite grown on the surface of well developed cubes of zeolite A; (c) very well defined cubes of zeolite A; and (d) well defined crystals of zeolite X

Adsorption Isotherms

The capacity of raw kaolinite to remove BB-41 was very small and did not exceed 6 mg/g. However, the zeolitic materials exhibited larger removal percentages and capacities (40.9 mg/g for HS, 27.3 mg/g for ZA, and 16.9 mg/g for ZX). For all the samples, the removal percentage of BB-41 decreased as the initial concentration of BB-41 increased. However, the removal capacity of the different zeolites increased as the initial concentrations did (Fig. 6). Similar results were reported using different adsorbents (Adeyemo et al. Reference Adeyemo, Adeoye and Bello2015).

Fig. 6 Basic Blue 41 adsorption isotherms for Jordanian kaolinite, hydroxysodalite, zeolite A, and zeolite X

The experimental adsorption data were subjected to Langmuir and Freundlich adsorption isotherm models (Freundlich Reference Freundlich1906; Langmuir Reference Langmuir1916) to determine the adsorption capacity of the zeolitic materials, and to develop an equation that could be used for design purposes (Ismadji & Bhatia Reference Ismadji and Bhatia2000). The Langmuir isotherm represents the equilibrium distribution of dye molecules between the solid and liquid phases. The linearized form allows the calculation of the maximum adsorption capacity (q max) and the Langmuir constant (KL), and is given by eq. 3:

(3) C e q e = 1 q max . K L + C e q max

where q e is the amount adsorbed at equilibrium (mg/g), C e is equilibrium concentration of the adsorbate (mg/L), q max is the maximum amount of dye removed (mg/g), and KL is related to the energy of adsorption (Langmuir constant, L/mg). These parameters can be calculated from the intercept and slope of the linear plot for the experimental data of C e/q e vs. C e. The isotherms examined fitted well with this model, with R2 values close to 1 for all of the samples, except for the raw kaolinite. The largest calculated values for q max (Table 1) (the maximum amount or monolayer capacity of BB-41 removed) was for hydroxysodalite (close to 39 mg/g), the smallest value was for the ZX sample (17.69 mg/g), and for ZA the value was intermediate at 26.9 mg/g. The KL values indicate the affinity of BB-41 molecules for the zeolite surfaces, and it increased as the q max values increased.

Table 1 Langmuir parameters for the removal of BB-41 by the samples studied

RK, Raw kaolinite; HS, Hydroxysodalite; ZA, Zeolite A; ZX, Zeolite X

The Freundlich isotherm is an empirical equation which assumes that the adsorption process occurs on heterogeneous surfaces, with a non-uniform distribution of heat of adsorption over the surface (Freundlich Reference Freundlich1906). The amount of BB-41 adsorbed (q e) increased curvilinearly with increasing equilibrium concentration (C e) but the data did not fit the linear form very well, expressed as

(4) ln q e = ln K F + 1 n ln C e

where KF (mg/g) and n are the Freundlich constants related to the adsorption capacity, and adsorption intensity of adsorbents, respectively. Values of n in the range 0 < 1/n < 1 represent favorable adsorption conditions. If a plot of lnq e vs. lnC e is linear, values of the intercept KF and the slope 1/n can be obtained. Because the experimental data did not fit the Freundlich equation well for ZA and ZX samples, the R2 values obtained were far from 1 (Table 2).

Table 2 Freundlich parameters for the removal of BB-41 by the samples studied

RK, Raw kaolinite; HS, Hydroxysodalite; ZA, Zeolite A; ZX, Zeolite X

Adsorbents such as zeolites consist primarily of silicon and aluminum oxides, the hydroxylated surface of which developed negative charges in aqueous solution (Bertolini et al. Reference Bertolini, Izidoro, Magdalena and Fungaro2013). The removal of BB-41 occurred through interaction forces between the negatively charged sites of the adsorbent and the positively charged dye molecules. The removal might have been achieved by cation exchange, when the dye molecule size was appropriate for accessing the pores. The BB-41 molecule is considered a large molecule, however, and could not access the pores of the prepared zeolites (Selim et al. Reference Selim, El-Makkawi and Ibrahim2018). Indeed, the primary crystalline spherical cavities of a-cages in zeolite A and supercages in zeolite X are 11.4 and 12.6 Å, respectively (Job Reference Job and Fahrner2005). Meanwhile, zeolite materials generally display high affinity for cationic dyes with very small molecular size, such as methylene blue (Awala et al. Reference Awala, Leite, Saint-Marcel, Clet, Retoux, Naydenova and Mintova2016).

A comparison of the maximum removal capacity of BB-41 by various zeolites with similar aluminosilicate structures (Table 3) revealed that the hydroxysodalite prepared has a maximum removal capacity of 39 mg/g, larger than the reported value of 27 mg/g (Selim et al. Reference Selim, El-Makkawi and Ibrahim2018). This difference could be related to the accessibility of removal sites within the structures of the various zeolite materials or to the specific surface areas of the zeolites used.

Table 3 Comparison of BB-41 removal by different adsorbents

RK, Raw kaolinite; HS, Hydroxysodalite; ZA, Zeolite A; ZX, Zeolite X

The well packed structure of kaolinite makes it difficult for the particles to be broken down, and the layers are not easily separated, so the adsorption properties of kaolinite are likely determined by the edges of the layers (Miranda-Trevinol & Coles Reference Miranda-Trevinol and Coles2003).The CEC is confined primarily to the external surface, in contrast to smectites where most of the CEC belongs to interior sites (Miranda-Trevinol & Coles Reference Miranda-Trevinol and Coles2003). Kaolinite, consequently, removed less BB-41 than smectite. In addition, the values obtained were smaller than those reported for some raw clay minerals and acid-activated derivatives (Roulia & Vassiliadis Reference Roulia and Vassiliadis2005; Kooli et al. Reference Kooli, Yan, Al-Faze and Al-Sehimi2015a). This difference could be related to the specific surface area or to the accessibility of the adsorbent sites for the BB-41 dye molecules. Indeed, the textural properties of the starting raw kaolinite (RK) and the derived zeolites revealed that the zeolite X and zeolite A exhibited larger surface areas (160 m2/g and 120 m2/g, respectively) than hydroxysodalite (26 m2/g, Table 4). However, small amounts of BB-41 were removed. In this case, the small removal capacities (of BB-41) using both ZX and ZA were probably due to its large size (Selim et al. Reference Selim, El-Makkawi and Ibrahim2018).

Table 4 Textural properties of raw kaolinite and synthesized zeolites

PV: pore volume; APD: average pore diameter; RK: Raw kaolinite; HS: Hydroxysodalite; ZA: Zeolite A; ZX: Zeolite X

Regeneration Data

The feasibility of applying the adsorbent systems to large-scale operations is determined by either disposal costs of spent adsorbents or regeneration costs. The commonly reported regeneration methods are thermal treatment (Tamon & Okazaki Reference Tamon and Okazaki1997: Wang et al. Reference Wang, Li, Xie, Liu and Xu2006; Vimonses et al. Reference Vimonses, Jim, Chow and Saint2009), chemical and solvent regeneration (Martin & Ng Reference Martin and Ng1978), electrochemical regeneration (Narbaitz & Karimi-Jashni Reference Narbaitz and Karimi-Jashni2008), ultrasonic regeneration (Lim & Okada Reference Lim and Okada2005), and wet air oxidation (Shende & Mahajani Reference Shende and Mahajani2002). An economical method was used to regenerate the spent zeolites, consisting of treating the spent zeolites with a solution in which Co2+ cations served as an homogeneous catalyst and the oxidant oxone was used to degrade the BB-41 removed (Anipsitakis et al. Reference Anipsitakis, Dionysiou and Gonzalez2006; Kooli et al. Reference Kooli, Yan, Al-Faze and Al-Sehimi2015a).

The regeneration of the spent zeolites (Fig. 7) revealed that the removal efficiency was preserved, but decreased slightly as the number of cycles increased from 1 to 4. A reduction of 10% was observed compared to the fresh zeolite, for hydroxysodalite. The regeneration efficiencies decreased by 30% after five cycles, and reached a lower value of 60% after seven cycles. Meanwhile, the efficiency of spent ZX and ZA was reduced after three cycles by an average of 25%. The reduction of efficiency might be related to deactivation of removal sites, and indicates that the BB-41 could not be released easily from the spent materials after many cycles, especially for ZX and ZA zeolites (Kooli et al. Reference Kooli, Yan, Al-Faze and Al-Sehimi2015b).

Fig. 7 Regeneration efficiency for (a) zeolite A, (b) zeolite X, and (c) hydroxysodalite after successive regeneration cycles

Conclusions

Natural Jordanian kaolin is a promising source of silicon and aluminum for the synthesis of zeolites. The formation of zeolites was achieved successfully from a Jordanian kaolin without its pre-activation at higher temperatures. Generally, zeolite A and X products obtained through the synthesis process exhibited well developed crystals with uniform particle-size distribution, as shown by SEM. The transformation of kaolinite into various zeolites led to an improvement in the removal of BB-41, the amounts removed depending on the initial concentrations of the BB-41and the types of zeolite materials obtained. The data fitted well to the Langmuir isotherm, with maximum removal capacities of 17, 29, and 39 meq/g for zeolites ZX, ZA, and hydroxysodalite, respectively. The adsorption capacity was only slightly suppressed at lower equilibrium concentrations for zeolite X and at higher concentration for hydroxysodalite and zeolite A. Zeolites prepared from Jordanian kaolin, especially hydroxysodalite, are promising for the treatment of wastes polluted by dyes. The removal efficiency was preserved after four cycles of regeneration for the zeolites used, and it depended on the type of materials with an average reduction of 10 to 30% after four cycles.

Acknowledgments

The authors are grateful to the Tafila Technical University (TTU), Jordan for technical support of this research work. Special thanks are due to the Institute of Mineralogy, Leibniz University, Hannover, Germany, for allowing the use of research facilities. Thanks also to Prof. Dr. C. Ruscher, Dr. Lars Robben, and Dipl. Geow. Valeriy Petrov for assistance with the acquisition of FTIR and SEM data.

References

Abdullahi, T., Harun, Z., & Othman, M. H. D. (2017). A review on sustainable synthesis of zeolite from kaolinite resources via hydrothermal process. Advanced Powder Technology, 28, 18271840.CrossRefGoogle Scholar
Adeyemo, A. A., Adeoye, I. O., & Bello, O. S. (2015). Adsorption of dyes using different types of clay: A review. Applied Water Science, 7, 126.Google Scholar
Alkan, M., Hopa, C., Yilmaz, Z., & Guler, H. (2005). The effect of alkali concentration and solid/liquid ratio on the hydrothermal synthesis of zeolite NaA from natural kaolin. Microporous Mesoporous Materials, 86, 176184.CrossRefGoogle Scholar
Anipsitakis, G. P., Dionysiou, D. D., & Gonzalez, M. A. (2006). Cobalt-mediated activation of Peroxymonosulfate and sulfate radical attack on phenolic compounds. Implications of chloride ions. Environmental Science and Technology, 40, 10001007.CrossRefGoogle ScholarPubMed
Awala, H., Leite, E., Saint-Marcel, L., Clet, G., Retoux, R., Naydenova, I., & Mintova, S. (2016). Properties of methylene blue in the presence of zeolite nanoparticles. New Journal of Chemistry, 40, 42774284.CrossRefGoogle Scholar
Barer, R. M. (1982). Hydrothermal chemistry of zeolites. London: Academic Press.Google Scholar
Barer, R. M. (1987). Zeolites and clay minerals as sorbent and molecular sieves. New York: Academic Press.Google Scholar
Barnes, M. C., Addai-Mensah, J., & Gerson, A. R. (1999a). The mechanism of the sodalite-to- cancrinite phase transformation in synthetic spent Bayer liquor. Microporous Mesoporous Materials, 31, 287302.CrossRefGoogle Scholar
Baughman, G. L., & Perenich, T. A. (1988). Fate of dyes in aquatic systems: I. Solubility and partitioning of some hydrophobic dyes and related compounds. Environmental Toxicology Chemistry, 7, 183199.CrossRefGoogle Scholar
Bertolini, T. C. R., Izidoro, J. C., Magdalena, C. P., & Fungaro, D. A. (2013). Adsorption of crystal violet dye from aqueous solution onto zeolites from coal fly and bottom ashes. Orbital Electron Journal Chemistry, 5, 179191.Google Scholar
Breck, D. W. (1974). Zeolite molecular sieves: Structure, chemistry and use. New York: John Wiley & Sons Inc.Google Scholar
Buhl, J. C., Hoffmann, W., Buckermann, W. A., & Muller-Warmuth, W. (1997). The crystallization kinetics of sodalites grown by the hydrothermal transformation of kaolinite studied by 29Si MAS NMR. Solid State Nuclear Magnetic Resonance, 9, 121128.CrossRefGoogle ScholarPubMed
Chung, K. T., & Cerniglia, C. E. (1992). Mutagenicity of azo dyes: Structure-activity relationships. Mutation Research, 277, 201220.CrossRefGoogle ScholarPubMed
Costa, E., De Lucas, A., Uguina, M. A., & Ruiz, J. C. (1988). Synthesis of 4A zeolite from calcined kaolins for use in detergents. Industrial and Engineering Chemistry Research, 27, 12911296.CrossRefGoogle Scholar
Dubey, A., Goyal, D., & Mishra, A. (2013) Zeolites in wastewater treatment. Pp. 82104 in: Green materials for sustainable water remediation and treatment (Mishra, A. & Clark, J.H., editors). RSC Green Chemistry Series, The Royal Society of Chemistry, Cambridge, UK.CrossRefGoogle Scholar
Farmer, V.C. (1979) Infrared spectroscopy. Pp. 285337 in: Data handbook for clay minerals and other non-metallic minerals (Olphen, H. van & Fripiat, J.J., editors). Pergamon Press, New York, USA.Google Scholar
Flanigen, E., Khatami, M., & Szymanski, H. A. (1971). Infrared structural studies of zeolite frameworks. Advances in Chemistry, 101, 201229.CrossRefGoogle Scholar
Freundlich, U. (1906). Dye adsorption in lusungen. Journal of Physical Chemistry, 57, 385470.Google Scholar
Garrido Pedrosa, A. M., Souza, M. J. B., Melo, D. M. A., & Araújo, A. S. (2006). Cobalt and nickel supported on HY zeolite: Synthesis, characterization and catalytic properties. Materials Research Bulletin, 41, 11051111.CrossRefGoogle Scholar
Gougazeh, M., & Buhl, J. C. (2010). Geochemical and mineralogical characterization of the Jabal Al-Harad kaolin deposit, southern Jordan, for its possible utilization. Clay Minerals, 45, 281294.CrossRefGoogle Scholar
Gougazeh, M., & Buhl, J. C. (2014). Conversion of natural Jordanian kaolin into zeolite a without thermal pre-activation. Zeitschrift für Anorganische und Allgemeine Chemie, 640, 16751679.CrossRefGoogle Scholar
Gupta, V. K., & Suhas, (2009). Application of low-cost adsorbents for dye removal-a review. Journal of Environmental Management, 90, 23132342.CrossRefGoogle ScholarPubMed
Heller-Kallai, L., & Lapides, I. (2007). Reactions of kaolinites and metakaolinites with NaOH-comparison of different samples (part 1). Applied Clay Science, 35, 99107.CrossRefGoogle Scholar
Hernández-Montoya, V., Pérez-Cruz, M. A., Mendoza-Castillo, D. I., Moreno-Virgen, M. R., & Bonilla-Petriciolet, A. (2013). Competitive adsorption of dyes and heavy metals on zeolitic structures. Journal of Environmental Management, 116, 213221.CrossRefGoogle ScholarPubMed
Hiraki, A., Nosaka, A., Okinaka, N., & Akiyama, T. (2009). Synthesis of zeolite-X from waste metals. Journal of the Iron and Steel Institute of Japan International, 49, 16441649.CrossRefGoogle Scholar
Huling, S. G., Jones, P. K., & Lee, T. R. (2007). Iron optimization for Fenton-driven chemical oxidation of MTBE-spent granular activated carbon. Environmental Science and Technology, 41, 40904096.CrossRefGoogle ScholarPubMed
Humelnicu, I., Baiceanu, A., Ignat, M. E., & Dulman, V. (2017). The removal of basic blue 41 textile dye from aqueous solution by adsorption onto natural zeolitic tuff: Kinetics and thermodynamics. Process Safety and Environmental Protection, 105, 274287.CrossRefGoogle Scholar
Ismadji, S., & Bhatia, S. K. (2000). Adsorption of flavor esters on granular activated carbon. Canadian Journal of Chemical Engineering, 78, 892901.CrossRefGoogle Scholar
Job, R. (2005) Zeolites and nanoclusters in zeolite host lattices. Pp. 127142 in: Nanotechnology and nanoelectronics: materials, devices, measurement techniques (Fahrner, W.R. editor). Springer, Berlin.Google Scholar
Johnson, E. B. G., & Arshad, S.E. (2014). Hydrothermally synthesized zeolites based on kaolinite: A review. Applied Clay Science, 97-98, 215221.CrossRefGoogle Scholar
Karadag, D., Akgul, E., Tok, S., Erturk, F., Kaya, M. A., & Turan, M. (2007). Basic and reactive dye removal using natural and modified zeolites. Journal of Chemical & Engineering Data, 52, 24362441.CrossRefGoogle Scholar
Kooli, F., Yan, L., Al-Faze, R., & Al-Sehimi, A. (2015a). Effect of acid activation of Saudi local clay mineral on removal properties of basic blue 41 from an aqueous solution. Applied Clay Science, 116–117, 2330.CrossRefGoogle Scholar
Kooli, F., Yan, L., Al-Faze, R., & Al-Sehimi, A. (2015b). Removal enhancement of basic blue 41 by waste brick from an aqueous solution. Arabian Journal of Chemistry, 8, 333342.CrossRefGoogle Scholar
Kyzas, G., & Kostoglou, M. (2014). Green adsorbents for wastewaters: A Critical Review. Materials, 7, 333364.Google ScholarPubMed
Langmuir, I. (1916). The constitution and fundamental properties of solids and liquids. Journal of the American Chemical Society, 38, 22212295.CrossRefGoogle Scholar
Lee, S. R., Han, Y. S., Park, M., Park, G. S., & Choy, J. H. (2003). Nanocrystalline sodalite from Al2O3 pillared clay by solid-solid transformation. Chemistry of Materials, 15, 48414845.CrossRefGoogle Scholar
Li, C., Dong, Y., Wu, D., Peng, L., & Kong, H. (2011). Surfactant modified zeolite as adsorbent for removal of humic acid from water. Applied Clay Science, 52, 353357.CrossRefGoogle Scholar
Li, Y., Li, L., & Yu, J. (2017). Applications of zeolites in sustainable chemistry. Chem, 3, 928949.CrossRefGoogle Scholar
Lim, J. L., & Okada, M. (2005). Regeneration of granular activated carbon using ultrasound. Ultrasonics Sonochemistry, 12, 277282.CrossRefGoogle ScholarPubMed
Ltaief, O. O., Siffert, S., Poupin, C., Fourmentin, S., & Benzina, M. (2015). Optimal synthesis of Faujasite-type zeolites with a hierarchical porosity from natural clay. European Journal of Inorganic Chemistry, 28, 46584665.CrossRefGoogle Scholar
Madani, A., Aznar, A., Sanz, J., & Serratosa, J. M. (1990). Silicon-29 and aluminum-27 NMR study of zeolite formation from alkalileached kaolinites: influence of thermal pre-activation. Journal of Physical Chemistry, 94, 760765.CrossRefGoogle Scholar
Martin, R. J., & Ng, W. J. (1978). The repeated exhaustion and chemical regeneration of activated carbon. Water Research, 21, 961965.CrossRefGoogle Scholar
Miranda-Trevinol, J. C., & Coles, C. A. (2003). Kaolinite properties, structure and influence of metal retention on pH. Applied Clay Science, 23, 133139.CrossRefGoogle Scholar
Nandi, B. K., Goswami, A., & Purkait, M. K. (2009). Removal of cationic dyes from aqueous solutions by kaolin: Kinetic and equilibrium studies. Applied Clay Science, 42, 583590.CrossRefGoogle Scholar
Narbaitz, R. M., & Karimi-Jashni, A. (2008). Electrochemical regeneration of granular activated carbons loaded with phenol and natural organic matter. Environmental Technology, 30, 2736.CrossRefGoogle Scholar
Ng, E.P., Zou, X., and Mintova, S. (2013) Environmental synthesis concerns of zeolites. Pp. 289310 in: New and future developments in catalysis: Hybrid materials, composites, and organocatalyts (Suib, L.S., editor). Elsevier, Amsterdam.CrossRefGoogle Scholar
Patel, S. (2012). Potential of fruit and vegetable wastes as novel biosorbents: Summarizing the recent studies. Reviews in Environmental Science and Biotechnology, 4, 365380.CrossRefGoogle Scholar
Pentrak, M., Madejová, J., & Komadel, P. (2009). Acid and alkali treatments of kaolins. Clay Minerals, 44, 511523.CrossRefGoogle Scholar
Pentrak, M., Madejová, J., Andrejkovičová, S., Uhlík, P., & Komadel, P. (2012). Stability of kaolin sand from Vyšný Petrovec deposit (South Slovakia) in acid environment. Geologica Carpathica, 63, 503512.Google Scholar
Querol, X., Moreno, N., Umana, J. C., Alastuey, A., Hernandez, E., Lopez-Soler, A., & Plana, F. (2002). Synthesis of zeolites from coal fly ash: An overview. International Journal of Coal Geology, 50, 413423.CrossRefGoogle Scholar
Rees, L., & Chandrasekhar, S. (1993). Hydrothermal reaction of kaolin in presence of fluoride ions at pH less than 10. Zeolites, (13), 534541.CrossRefGoogle Scholar
Rhodes, C. J. (2010). Properties and applications of zeolites. Science Progress, 93, 223284.CrossRefGoogle Scholar
Robinson, T., McMullan, G., Marchant, R., & Nigam, P. (2001). Remediation of dyes in textile effluent: A critical review on current treatment technologies with a proposed alternative. Bioresources Technology, 77, 247255.CrossRefGoogle ScholarPubMed
Roulia, M., & Vassiliadis, A. A. (2005). Interactions between C.I. Basic blue 41 and aluminosilicate sorbents. Journal of Colloid and Interface Science, 291, 3744.CrossRefGoogle ScholarPubMed
Russell, J.D. (1987) Infrared methods. Pp. 133173 in: A Handbook of Determinative Methods in Clay Mineralogy (Wilson, M.J., editor). Chapman & Hall, London.Google Scholar
Schwanke, A.J., Balzer, R., and Pergher, S. (2017) Microporous and mesoporous materials from natural and inexpensive sources. Pp. 122 Chapter 1. in: Handbook of Ecomaterials, (Martínez, L.M.T., Kharissova, O.V., and Khatisov, B.I., editors), Springer, Basel.Google Scholar
Selim, M.M., El-Makkawi, D. M., & Ibrahim, F. A. (2018). Innovative synthesis of black zeolites-based kaolin and their adsorption behavior in the removal of methylene blue from water. Journal of Materials Science, 53, 33233331.CrossRefGoogle Scholar
Shende, R. V., & Mahajani, V. V. (2002). Wet oxidative regeneration of activated carbon loaded with reactive dye. Waste Management, 22, 7383.CrossRefGoogle ScholarPubMed
Somerset, V. S., Petrik, L. F., White, R. A., Klink, M. J., Key, D., & Iwuoha, E. I. (2005). Alkaline hydrothermal zeolites synthesized from high SiO2 and Al2O3 co-disposal fly ash filtrates. Fuel, 84, 23242329.CrossRefGoogle Scholar
Song, L., Chen, F., Hu, J. C., & Richards, R. (2009). NiO (111) nanosheets as efficient and recyclable adsorbents for dye pollutant removal from wastewater. Nanotechnology, 20, 275707275716.CrossRefGoogle ScholarPubMed
de Sousa, M. L., de Moraes, P. B., Matos Lopes, P. R., Montagnolli, R. N., de Angelis, D., & Bidoia, E. D. (2012). Contamination by Remazol red brilliant dye and its impact in aquatic photosynthetic microbiota. Environmental Management and Sustainable Development, 1, 129138.Google Scholar
Tajul Islam, M., Aimone, F., Ferri, A., & Rovero, G. (2015). Use of N-methylformanilide as swelling agent for meta-aramid fibers dying: Kinetics and equilibrium adsorption of basic blue 41. Dyes and Pigments, 113, 554561.CrossRefGoogle Scholar
Tamon, Z. H., & Okazaki, M. (1997). Influence of surface oxides on ethanol regeneration of spent carbonaceous adsorbents. Journal of Colloid and Interface Science, 196, 120122.CrossRefGoogle ScholarPubMed
Vimonses, V., Jim, B., Chow, C. K. W., & Saint, C. (2009). Enhancing removal efficiency of anionic dye by combination and calcination of clay materials and calcium hydroxide. Journal of Hazardous Materials, 171, 941947.CrossRefGoogle ScholarPubMed
Wang, S., & Peng, Y. (2010). Natural zeolites as effective adsorbents in water and wastewater treatment. Chemical Engineering Journal, 156, 1124.CrossRefGoogle Scholar
Wang, S., Li, H., Xie, S., Liu, S., & Xu, L. (2006). Physical and chemical regeneration of zeolitic adsorbents for dye removal in wastewater treatment. Chemosphere, 65, 8287.CrossRefGoogle ScholarPubMed
Yagub, M. T., Kanti Sen, T., Afroze, S., & Ang, H. M. (2014). Dye and its removal from aqueous solution by adsorption: A review. Advances in Colloid and Interface Science, 209, 172184.CrossRefGoogle ScholarPubMed
Yuna, Z. (2016). Review of the natural, modified, and synthetic zeolites for heavy metals removal from wastewater. Environmental Engineering Science, 33, 443454.CrossRefGoogle Scholar
Zaarour, M., Dong, B., Naydenova, I., Retoux, R., & Mintova, S. (2014). Progress in zeolite synthesis promotes advanced applications. Microporous and Mesoporous Materials, 189, 1121.CrossRefGoogle Scholar
Zhao, H., Deng, Y., Harsh, J., Flury, M., & Boyle, J. (2004). Alteration of kaolinite to cancrinite and sodalite by simulated Hanford waste and its impact on cesium retention. Clays Clay Minerals, 52, 113.CrossRefGoogle Scholar
Figure 0

Fig. 1 Location map of the Jabal Al-Harad, Jordan, kaolin deposit

Figure 1

Fig. 2 Chemical structure of the dye Basic Blue 41

Figure 2

Fig. 3 Powder XRD patterns of the samples, including (a) Jordanian kaolinite (JK), (b) synthesized hydroxysodalite (HS), (c) synthesized zeolite X (ZX), and (d) synthesized zeolite A (ZA)

Figure 3

Fig. 4 FTIR spectra of the samples, including (a) Jordanian kaolinite, (b) synthesized zeolite X (ZX), (c) synthesized hydroxysodalite (HS), and (d) synthesized zeolite A (ZA)

Figure 4

Fig. 5 SEM images of: (a) of pseudohexagonal crystals of Jordanian kaolinite; (b) spherical (spheroidal) agglomerates of hydroxysodalite grown on the surface of well developed cubes of zeolite A; (c) very well defined cubes of zeolite A; and (d) well defined crystals of zeolite X

Figure 5

Fig. 6 Basic Blue 41 adsorption isotherms for Jordanian kaolinite, hydroxysodalite, zeolite A, and zeolite X

Figure 6

Table 1 Langmuir parameters for the removal of BB-41 by the samples studied

Figure 7

Table 2 Freundlich parameters for the removal of BB-41 by the samples studied

Figure 8

Table 3 Comparison of BB-41 removal by different adsorbents

Figure 9

Table 4 Textural properties of raw kaolinite and synthesized zeolites

Figure 10

Fig. 7 Regeneration efficiency for (a) zeolite A, (b) zeolite X, and (c) hydroxysodalite after successive regeneration cycles