Journal Pre-proof Preparation of a fly ash-based geopolymer for removal of a cationic dye: Isothermal, kinetic and thermodynamic studies Ozkan Acisli, Ilker Acar, Alireza Khataee
PII:
S1226-086X(19)30603-3
DOI:
https://doi.org/10.1016/j.jiec.2019.11.012
Reference:
JIEC 4854
To appear in:
Journal of Industrial and Engineering Chemistry
Received Date:
4 October 2019
Revised Date:
7 November 2019
Accepted Date:
10 November 2019
Please cite this article as: Acisli O, Acar I, Khataee A, Preparation of a fly ash-based geopolymer for removal of a cationic dye: Isothermal, kinetic and thermodynamic studies, Journal of Industrial and Engineering Chemistry (2019), doi: https://doi.org/10.1016/j.jiec.2019.11.012
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Published by Elsevier.
Preparation of a fly ash-based geopolymer for removal of a cationic dye: Isothermal, kinetic and thermodynamic studies
Ozkan Acisli,a,* Ilker Acar, a Alireza Khataee, b,c,*
a
Department of Petroleum and Natural Gas Engineering, Faculty of Earth Sciences, Ataturk
b
ro of
University, 25400, Oltu, Erzurum, Turkey Research Laboratory of Advanced Water and Wastewater Treatment Processes, Department
of Applied Chemistry, Faculty of Chemistry, University of Tabriz, 51666-16471 Tabriz, Iran c
-p
Institute of Research and Development, Duy Tan University, Da Nang 550000, Vietnam
*
[email protected] (O. Acisli)
re
Corresponding authors:
Jo
ur
na
Graphical Abstract
lP
[email protected],
[email protected] (A. Khataee)
Highliths
Fly ash-based geopolymer was prepared for the removal of Basic Yellow 2. 1
The removal efficiency increased with the geopolymer dosage.
An increase in the initial dye concentration caused a decrease in the efficiency.
The Langmuir and Temkin models were well correlated with the obtained results.
The adsorption process occurs spontaneously and shows endothermic character.
Abstract A class F fly ash-based geopolymer was prepared for the removal of Basic Yellow 2 (BY2)
ro of
from aqueous solutions. The geopolymerization process transformed the spherical fly ash particles into a porous and amorphous polymeric structure, sodium-alumina-silicate hydrate gel. The removal efficiency increased with the geopolymer dosage while it decreased with BY2
-p
concentration. Specifically, the efficiency of 94.47% was achieved with the geopolymer dosage
re
of 2.0 g L-1. However, further increases resulted in a substantial reduction, down to 57.76% for 3.0 g L-1. At 293 K for 300 min, the efficiency decreased from 84.69 to 64.19% for BY2
lP
concentrations of 10 and 50 mg L-1, respectively. Based on the isothermal, kinetic and thermodynamic investigations of the adsorption process, the Langmuir and Temkin models well
na
correlated with the obtained results. The adsorption occurs spontaneously and shows endothermic character. The pseudo-second order and the intra-particle diffusion models are
ur
valid for 293 K while the adsorption results well suited with the pseudo-second order for 313 and 323 K. In conclusion, the prepared fly ash-based geopolymer provided favorable results for
Jo
the removal of Basic Yellow 2 from aqueous solutions.
Keywords: Fly ash; Geopolymer; Alkaline activation; Degradation; Cationic dye.
2
1. Introduction Clean water resources have been contaminated day to day by a variety of pollutants like heavy
ro of
metals, dyes, pharmaceuticals, pesticides and others. From these pollutants, especially organic dyes even in trace amounts can be very hazardous due to their toxic, carcinogenic and mutagenic effects. Though a large number of conventional methods attempted, adsorption is by
-p
far the most widely used process for the removal of dyes owing to its simplicity, low cost and high efficiency. Specifically, very high surface area, good stability and long durability
re
properties make activated carbon the most feasible adsorbent. However, its high synthesis cost
lP
and difficulties in the regeneration process have forced researchers to look for lower-cost alternatives [1-6].
na
The term, “Geopolymer”, first used by Davidovits in 1976, can be defined as a negatively charged three dimensions alumina-silicate framework in which Na+, K+ and Ca2+ ions function
ur
as charge-balancing cations [3, 7]. Geopolymers can generally be prepared by activating various silica and alumina rich materials with strongly alkaline solutions at ambient or slightly elevated
Jo
temperatures [8, 9]. In literature, various raw materials and waste products such as metakaolin [4, 7], coal fly ash [2, 10], blast furnace slag [11, 12], natural alumina-silicates [13, 14], red mud [8] and biomass fly ash [6] have been studied for this purpose. Depending on the activator and processing conditions used, the resulting structure is mostly amorphous alumina-silicates or semi-crystalline zeolites [15-18].
3
Coal fly ash is the major by-product generated in thermal power plants. Its annual worldwide production is in the range of 500-600 million tons, almost half of which is mainly used in cement and concrete industry. The remaining part is being disposed in especially landfills, and also ash ponds and lagoons [19-23]. The resulted environmental pollution can be reduced by turning this waste into a beneficial material. Since coal fly ash is mainly constituted by amorphous alumina-silicate framework, it can be considered as cheap and readily available raw material for the synthesis of geopolymers [10].
ro of
Fly ash-based geopolymers have some significant advantageous properties like high mechanical strength and temperature resistance alongside low permeability and shrinkage [15, 19]. In addition, less energy requirement and low CO2 emission in the manufacturing process
-p
make fly ash-based geopolymer a promising replacement material for ordinary Portland cement [8, 9, 24]. The extensive utilization of fly ash-based geopolymer seems to take place in
re
construction industry [16]. The other application areas involve adsorption of heavy metals [3,
lP
7, 10, 19] and dyes [2, 4, 6, 15], and immobilization of toxic metals [16, 25]. In this study, a class F fly ash-based geopolymer prepared by a combination of alkaline activation and thermal treatment was used for the removal of Basic Yellow 2 from aqueous
na
solutions. In terms of fly ash-based geopolymer usage, a considerable body of literature has been concentrated on cement replacement, and there are also several studies on heavy metal
ur
adsorption. However, the existing literature on dye removal is very limited to only a few studies.
Jo
In addition, there is no report in the literature as much detailed as this study which covers a combination of the detailed geopolymer characterization, the isothermal, kinetic and thermodynamic investigations as well as the effect of operational parameters.
2. Experimental 2.1. Materials 4
The coal fly ash sample from Çatalağzı Power Plant, Turkey was used as the feedstock. In the experimental stage, the as-received sample (CFA) was first subjected to a specific hydraulic classification process to recover its ultra-fine fraction (CUFA). The obtained CUFA was then used as the alumina-silicate source for the geopolymer (GEO) synthesis in which analytical grade NaOH pellets were used as the alkaline activator. Chemically pure grade Basic Yellow 2 (BY2) was supplied from Haining Deer Chemical Co.. BY2, which has a commercial name as Auramine-O, is a cationic dye. Its chemical formula, molecular weight and maximum
ro of
adsorption wavelength (λmax) are C17H22ClN3, 303.834 g mol-1 and 432 nm, respectively. The main chemical structure can be seen in the study conducted by Öztürk and Malkoc, 2014 [26].
-p
Distilled water was used throughout the experiments.
2.2. Methods
re
2.2.1. Hydraulic classification
lP
The main framework of the classification used is based on a lamella hydraulic classifier. The particle separation is accomplished by the differences in particle settling rates in water. In this
na
system, ultrafine particles (CUFA) are recovered from the top far end, overflow discharge area. Detailed information about the classifier can be obtained from Robl and Groppo, 2011 [27] and
ur
McCarthy et al., 2013 [28].
Jo
2.2.2. Preparation of the geopolymer A certain amount of CUFA was mixed with 10 M NaOH solution using a solid mass ratio of 1:1.2. After thoroughly stirred with a glass rod for 10 min, a homogenized geopolymer paste started to form. The top of the beaker was closed with a round-bottom glass plate and then treated at 350 oC for 1 h to obtain a fused form. After cooling to room temperature, it was mixed with distilled water using a mass ratio of 1:4, followed by curing for 24 h in a thermostatic 5
shaker at room temperature and 100 rpm. The slurry was then filtrated and washed with distilled water many times to remove excess NaOH. The obtained geopolymer paste (GEO) was finally dried at 105 oC overnight prior to its utilization.
2.2.3. Characterization of materials A Spectro IQ X-ray fluorescence (XRF) device was used to analyze the major chemical
ro of
inclusions. Loss on ignition (LOI) was determined based on ASTM C311. Particle size measurements were done by a Malvern Hydro 2000S instrument. The crystalline structures were examined by a PANalytical Empyrean X-ray diffractometer (XRD) operating with Cu-Kα
-p
radiation (λ = 1.54051 Å) at 40 kV and 30 mA over 10-80°. Microstructural characterization and point elemental content were determined using high-resolution Zeiss Sigma 300 and
re
Hitachi S-4800 scanning electron microscopes (SEM) equipped with an energy dispersive X-
lP
ray (EDX) analyzer. Fourier transform infrared (FT-IR) spectra of the samples were recorded by a Tensor 27, Bruker instrument using the KBr pellet technique in 4000-400 cm-1. The surface
na
area, total pore volume and pore size distribution were measured through N2 adsorption method at approximately -197 oC using a Micromeritics 3 Flex instrument. The BET surface areas of CUFA and its geopolymer were determined for the relative pressure range of 0.0-1.0. The total
ur
pore volume was determined for a single point P/P0 value of 0.950. The pore size distributions
Jo
of the samples were inferred from the isotherms by means of the Barrett-Joyner-Halenda (BJH) method. The measured samples were degassed for 15 h at 100 oC prior to the adsorption tests. The natural pH values of the samples were determined based on the method suggested by Li et al., 2006 [2].
2.2.4. Batch adsorption experiments 6
Batch adsorption tests were conducted in 100 mL glass flasks which were shaken at 100 rpm and a certain temperature in a thermostatic shaker. The volume of BY2 solution was kept constant as 100 mL throughout the experiments. At predetermined time intervals, the approximately 5-mL sample was withdrawn from the treated solution and then centrifuged two times for 3 min at 5000 rpm by a Hettich EBA 20 instrument. BY2 concentration in the solution was determined using an Optizen pop UV-Vis spectrophotometer at 432 nm (λmax for BY2). The adsorbed amount of BY2 (mg g-1) at equilibrium (qe) and a specific time (qt), and its
ro of
removal efficiency (RE, %) were calculated from the commonly-used well-known equations [6, 10, 17, 29].
-p
3. Results and discussion
re
3.1. Characterization of materials
3.1.1. Chemical composition and basic physical properties
lP
The main chemical compositions and basic physical properties of CFA, CUFA and CUFAbased geopolymer (GEO) are shown in Table 1. As seen from Table 1, Çatalağzı fly ash has a
na
typical chemical inclusion of a regular class F type. According to the ASTM C618 standard, both CFA and CUFA meet the chemical specifications of a class F fly ash owing to their high
ur
cumulative amounts of silica, alumina and iron oxide (SiO2+Al2O3+Fe2O3 ≥ 70) with low SO3 (≤ 5) and LOI (≤ 6) values [20, 21]. The mean sizes (d50) by volume were determined as 39.29,
Jo
5.25 and 0.70 µm for the respective CFA, CUFA and GEO. Accordingly, the BET surface area values of 0.83, 6.09 and 47.38 m2 g-1 were obtained for the sequential CFA, CUFA and GEO.
Table 1
7
The elemental inclusions in CUFA and GEO were also examined using EDX analysis, and the results were indicated with superscript a. According to the EDX analysis, O, Si and Al constitute the main structure of CUFA. The amount of these elements decreased and Na participated in the structure due to the alkaline activation with NaOH, resulting in the formation of sodiumalumina-silicate hydrate gel [30]. Fig. 1 illustrates the particle size distributions of CFA, CUFA and GEO. As seen from Fig. 1, CFA with a coarse mean size of 39.29 µm has a very large size distribution while its overflow
ro of
product, CUFA, with a very fine mean size of 5.25 µm has a relatively close distribution. After the geopolymerization, a very narrowly-sized distributed GEO product varying in the range of
-p
400-1100 nm was obtained with a mean size of approximately 700 nm.
re
Fig. 1
lP
3.1.2. Mineralogical analysis
The XRD patterns of CUFA and GEO are shown in Fig. 2. ICSD database was used to analyze
na
the diffraction peaks via a software program. As seen from Fig. 2, CUFA consists of mullite (Al4.68Si1.32O9.66, ICSD 98-008-0498) and quartz (SiO2, ICSD 98-005-7052) as the major
ur
crystalline inclusions, and also includes a considerable amount of hematite (Fe2O3, ICSD 98001-2749). In addition, the XRD analysis indicates that a vast majority of CUFA is constituted
Jo
by amorphous glass phase which substantially affects source material’s reactivity for the synthesis of geopolymers. In other words, higher amorphous content means better reactivity [10, 31].
Fig. 2
8
After the geopolymerization process, all the diffraction peaks extensively decreased, and most of the mullite peaks disappeared. The relative mullite content in the total crystalline inclusions diminished from 61.4 to 29.0% probably due to its reactive alumina-silicate structure. The sharp peaks converted into the broadband peaks with much lower intensities, which are characteristics for amorphous materials. All of these results indicate that the crystalline alumina-silicate phase altered to the amorphous structure as a result of the geopolymerization, which is consistent with
ro of
the work of Al-Zboon et al., 2011 [10].
3.1.3. Microstructural analysis
SEM images provide detailed information in terms of a material’s morphology and surface
-p
texture. Fig. 3 exhibits the SEM micrographs of CUFA and GEO. Fig. 3-(a-d) show various CUFA images in different magnifications. As seen clearly from these images, CUFA mainly
re
consists of solid and hollow spherical particles with mostly smooth surfaces, and also low
lP
amounts of irregularly shaped particles [20, 21]. These SEM images also verify the results obtained from the laser size analysis. Fig. 3-(e) clearly exhibits the transformation of spherical
na
CUFA particles into amorphous structure after the geopolymerization process [32]. Fig. 3-(f) shows the geopolymeric structure in conjunction with the unreacted CUFA particles, indicating
ur
incomplete geopolymerization. These unreacted particles are the refractory parts which are unaffected by the process mainly because of their strong crystalline nature [33]. Considering
Jo
Fig. 3-(f) together with the XRD analyses, it can be said that most of these unreacted particles are constituted by the crystalline quartz phase. The negative impact of higher Si/Al ratio of starting material on degree of geopolymerization was also reported by Siyal et al., 2016 [33]. According to the researchers, this can be resulted from the rapid setting of the paste prior to the completion of a major portion of dissolution. This explanation clarifies the incomplete geopolymerization in this study in which dissolution of mostly mullite and partly quartz phases 9
takes place. In other words, quartz particles are mostly solidified before the complete dissolution due to the chemical inclusions (SiO2). As far as is known, there is no specific method available for determining the quantity of this refractory part in geopolymers [33]. Fig. 3 (g-h) clearly indicates the alumina-silicate gel structure which is also observed from the literature [15, 19]. Fig. 3
ro of
The elemental inclusions in CUFA and its geopolymer were examined using EDX analysis. According to the EDX spectra shown in Fig. 4, O, Si, Al and Fe constitute the main structure of CUFA. After geopolymerization, Na accompanied with these elements due to the formation
re
-p
of sodium-alumina-silicate hydrate gel [30].
3.1.4. FT-IR analysis
lP
Fig. 4
FT-IR analysis is a commonly used method to examine the changes and modifications in the
na
structure of a substance due to any chemical or physico-chemical treatment processes [34]. The FT-IR spectra of CUFA and its geopolymer, GEO, can be seen from Fig. 5. The main
ur
characteristic bands, their identified species and the changes with the geopolymerization can be
Jo
clarified as follows.
Fig. 5
The strong asymmetric band at 3433 cm-1 is the stretching vibration of the H-O-H groups due to adsorbed molecular water in the fly ash. The weak bands placed at 1636 cm-1 and 2300 cm-1 could be assigned to the respective stretching and deformation vibrations of the OH [32, 35] 10
and simultaneous stretching vibrations of structural OH and Si-O bonds [6]. The band located at ~980 cm-1 can be related to the stretching of asymmetric Al-O-Al/Si-O-Si bonds [6, 36]. Two bands at 720 and 590 cm-1 could be assigned to the stretching vibrations of the symmetric SiO-Si (Al) bridges, corresponding to cyclo-silicates vibrations [31]. The band located at ~1440 cm-1 can be attributed to sodium carbonate inclusion, resulted from the reaction of residual sodium with atmospheric carbon dioxide, indicating the geopolymer formation [6, 32]. The bands at around 2500 cm-1 could be associated with the
ro of
stretching and bending vibrations of the respective –OH and H–O–H bonds in the water, which is adsorbed on the surface of the geopolymer and in the large cavities of the polymeric network [6, 37]. The band at 472 cm-1 are due to symmetric bending vibrations of Si-O-Si or Al-O-Al
-p
bonds in the geopolymer while, similarly, the one located at 452 cm-1 could be assigned to SiO or Al-O bonds [33]. The band located at 880 cm-1 could be assigned to the symmetric
re
stretching in Al-O tetrahedral framework which also specifies the geopolymeric inclusion [32, 38]. The peak at 2160 cm-1 can be attributed to stretching bending vibrations of Si-H. Moreover,
lP
the weak peaks observed in 1660 and 1974 cm-1 can be caused by C≡C and C=O stretching
na
bending of unburned carbons in the fly ash.
ur
3.1.5. Surface area and pore properties Fig. 6-(a) displays the nitrogen adsorption/desorption isotherms of CUFA and GEO which were
Jo
analyzed with the BET model. Based on the IUPAC classification, GEO follows the H3 type of the hysteresis loops. There are two main characteristic features of this loop. Firstly, the adsorption branch looks like a Type II isotherm. The other one is that the desorption branch has a lower limit normally located at the P/P0 caused cavitation. The H3 type loop resulted from the aggregation of non-rigid stratified particles like certain clays if macro-pores, which are not fully
11
filled with pore condensate, constitute the main pore network [39]. On the other hand, CUFA did not exhibit this loop.
Fig. 6
Fig. 6-(b) shows the BJH pore size distributions of the materials. Pore volume of GEO was mainly distributed in the size range of 2-20 nm. However, much lower pore volume values were
ro of
observed for CUFA. In this case, the pore volume decreased linearly with pore width and their distribution mainly varied from 2 to 10 nm. As also seen from the table in Fig. 6, CUFA has a total pore volume of 0.008 cm3 g-1 and an average pore width of 4.970 nm. Surface area values
-p
of 6.090 and 5.717 m2 g-1 were obtained based on the BET and BJH methods, respectively. After the geopolymerization, total pore volume and average pore width increased to the
na
3.2. Isothermal investigations
lP
were also observed from this table.
re
respective 0.156 cm3 g-1 and 9.792 nm. Accordingly, extensive increases in the surface areas
Isotherm studies were carried out with the constant adsorbent dosage of 2.0 g, the equilibrium
ur
time of 300 min and the natural pH of about 9.85 for the three different temperature values, 293, 313 and 323 K. A total of six adsorption isotherms, namely the Langmuir, Freundlich,
Jo
Temkin, BET, Dubinin-Radukevisch and Harkins-Jura were used to determine the adsorption mechanism of BY2 on the produced geopolymer. Together with the equations used, Table 2 summarizes the parameters of these isotherms calculated from the obtained experimental results.
Table 2 12
As seen clearly from Table 2, the experimental results were better correlated with the Langmuir and Temkin isotherms due to their much higher R2 values than those of the other ones. As it is well known, the Langmuir model assumes the adsorption as monolayer and homogenous. This type of adsorption can be resulted from not only physical interactions but also very strong electrostatic attractions. The first situation occurs when the pore radius is large enough for attaching adsorbate in a single layer, while the second one happens when there is an electrostatic
ro of
attraction between adsorbent and adsorbate [29, 40]. The favorability of the Langmuir isotherm can be expressed by a dimensionless constant, RL, defined by Eq. 1 [41, 42]:
1
𝑅𝐿 = (1+𝐾𝐶 )
(1)
re
-p
0
in which K and Co are the Langmuir constant and the initial dye concentration, respectively.
lP
According to this approach, the Langmuir isotherm is irreversible for the dimensionless constant equal to 0 (RL=0), favorable for the values between 0 and 1 (0< RL<1), linear for 1 and
na
not favorable for the values higher than 1 (RL>1) [41, 42]. Since the obtained RL values in Table 2 vary between 0.091 and 0.151, it can be said that the Langmuir isotherm is favorable for the
ur
adsorption of BY2 on the geopolymer in all the tested temperatures. The Freundlich model is used to examine the heterogeneity of an adsorption process, which is
Jo
determined by the heterogeneity factor, n. In this model, if n>1, adsorption occurs via physical interaction while the condition of n<1 points out chemical interaction between adsorbent and adsorbate. In addition, adsorption is a linear process if n equals to 1.0 [40]. As seen from Table 2, the low R2 values show that the adsorption process is unsuited with the Freundlich isotherm so the calculated n values have no meaning in this case. On the other hand, the Temkin model is better correlated with the obtained results compared to the other models tested especially for 13
the trial performed at 293 K. The R2 value of 0.987, the highest one obtained from the whole isotherm studies, was attained at this point. These results indicated that the Temkin model wellcharacterized the interactions between the geopolymer and the dye [43]. The heat of adsorption of the all molecules in the layer reduced linearly with coverage layer, and the energy is uniformly distributed [44]. In the Temkin equation, bT and aT are the constants for the respective heat of adsorption and maximum binding energy. The parameters calculated from the Temkin equation emphasized that the adsorption of BY2 at 293 K was characterized by a homogeneous
ro of
distribution binding energies up to the maximum point [40]. The low regression coefficients obtained from the BET, Dubinin-Radukevisch and Harkins-Jura isotherms clearly pointed out that the adsorption process was not compatible with these isotherms.
-p
The determination of the appropriate isotherm type has vital importance in the design of adsorption systems. Since adsorption isotherm clarifies mechanism between adsorbent and
re
adsorbate, it is critical in optimizing the appropriate adsorbent dosage [45]. Fig. 7 illustrates the
equilibrium time of 300 min.
na
Fig. 7
lP
adsorption isotherm patterns at three different temperatures, 293, 313 and 323 K, for the
ur
As seen from Fig. 7, the efficiency increased with temperature, indicating endothermic
Jo
character of the adsorption process. Similarly, the amount of adsorbed BY2 also increased with temperature. The equilibrium adsorption isotherm resembles the Type-I. In particular, the isotherm curves obtained at 313 and 323 K are also similar to the Type-I isotherm. This type of isotherm can be characterized with the mono-molecular chemical adsorption occurring as monolayer. On the other hand, the sharp increases take place close to the isotherm curves of micro-porous solids. When the surfaces of micro-pores with high adsorption potential are
14
coated mono-molecularly, the adsorption proceeds until the pores are completely filled. Adsorbed amount in the Type-I isotherm converges a limiting value mainly ruled by the volume of effective micro pore instead of the internal surface area [46].
3.3. Kinetic study The three kinetic models, namely pseudo-first-order, pseudo-second-order and intra-particle
ro of
diffusion were used to evaluate the kinetic investigation of BY2 adsorption on the geopolymer surface. Kinetic studies were conducted at 293, 313 and 323 K with the varying dye concentrations (10-50 mg L-1) for the constant adsorbent dosage of 2.0 g and the natural pH of
-p
9.85. The equations of these models are sequentially given by Eq. 2-4 [45, 47-51].
=
1 𝑘2 𝑞𝑒2
+
𝑡 𝑞𝑒
lP
𝑡 𝑞𝑡
re
𝑙𝑛(𝑞𝑒 − 𝑞𝑡 ) = 𝑙𝑛𝑞𝑒 − 𝑘1 𝑡
(3) (4)
na
𝑞𝑡 = 𝑘𝑖 √𝑡 + 𝐶
(2)
where qe and qt are the adsorbed BY2 amounts at the equilibrium and a specific time, k1 (s-1)
ur
and k2 (g mg-1 s-1) express the rate constants for the respective pseudo-first and pseudo-secondorder models, ki (mg g-1 s-1/2) can be defined as the rate constant of the intra-particle diffusion
Jo
model and C (mg g-1) represents the boundary layer thickness [47-49]. The obtained values of the kinetic constants are summarized in Table 3.
Table 3
15
According to Table 3, the pseudo-first order model did not fit to the experimental results. The highest correlation efficiency obtained based on this model was only 0.7173. On the other hand, the pseudo-second order model extremely well correlated with the results especially for the temperatures of 313 and 323 K. It was also understood that the obtained R2 values were independent of the initial dye concentrations. However, the rate constants (k2) generally decreased with the increasing dye concentrations except for 293 K. This result can be explained by more competition for the surface active sites at higher concentrations [29, 40]. The Weber
ro of
intra-particle diffusion model was also used to describe the adsorption process. Unlike the pseudo-second order model, this time the intra-particle diffusion was compatible with the results especially for 293 K (R2>0.9326) and did not fit to the ones obtained at 323 K. The
-p
growing C values with the increasing dye concentrations showed that thickness of the boundary layer has an effective role in the adsorption process for 293 K, indicating a larger boundary
re
layer diffusion effect [29, 41, 52]. However, the boundary layer most probably begins to
lP
disappear with increasing temperature, inhibiting the diffusion effect. Overall results suggest that the pseudo-second order and the intra-particle diffusion models both occur at the same time
na
for the low temperature.
ur
3.4. Thermodynamic investigation
The adsorption process was also applied to thermodynamic investigation in which the changes
Jo
of enthalpy (∆Hads), entropy (∆Sads) and Gibbs free energy (ΔGo) were calculated by Eq. 5-7 [53]:
𝑑(𝑙𝑛𝐶)
= 𝑑(1/𝑇)
∆𝐻𝑎𝑑𝑠
𝑑(𝑙𝑛𝐶)
∆𝑆𝑎𝑑𝑠
𝑑(𝑙𝑛𝑇)
=
(5)
𝑅
(6)
𝑅
16
∆𝐺 𝑜 = −𝑅𝑇𝑙𝑛𝐾
(7)
where T is temperature (K) and R is the gas constant (8.314x10-3 kJ mol-1 K-1). According to Eq. 5, when (ln C) is plotted against (1/T), the slope of the straight line will yield a value of (∆H/R). Similarly, based on Eq. 6, when (ln C) is plotted against (ln T), this time the slope will be (∆S/R). Thus, ∆H and ∆S values can be determined from simple calculations using the above
ro of
approximations. The obtained results are summarized in Table 4.
Table 4
-p
As seen from Table 4, ∆Hads has a positive value while ∆Sads takes a negative one. The growing amount of the adsorbed dye (q) with increasing temperature implies the endothermic character,
re
also verifying the positive enthalpy change. On the other hand, the negative entropy indicates
lP
an increased order at the interface. Mobility of the dye ions increases with increasing temperature, causing the ions to escape from the solid surface to the liquid phase. This results in a decrease in the amount of dyes adsorbed. However, negative ∆G values specify the
na
spontaneous nature of the adsorption for all the temperatures and dye concentrations tested. After all, it can be said that these results can be used to characterize adsorption processes with
Jo
ur
similar adsorbents and adsorbates.
3.5. The effects of operational parameters on the adsorption 3.5.1. The adsorbent dosage Determination of appropriate adsorbent dosage is vital in terms of the capacity for a given initial concentration of the adsorbed material [54]. In this stage, the effect of GEO dosage varying from 0.5 to 3.0 g L-1 was examined as a function of processing time for the constant initial BY2 17
concentration of 30 mg L-1, the temperature of 293 K and natural pH of 9.85. The obtained results were illustrated in Fig. 8.
Fig. 8
As seen in Fig. 8-(a), the removal efficiency of BY2 increased with processing time independent of the dosages tested. However, much higher increasing ratios were obtained with 2.5 and
ro of
especially 2.0 g L-1. When the time was kept constant at 900 min, from Fig. 8-(b), the efficiency substantially increased from 13.28 up to 94.47% for the adsorbent dosages of 0.5 and 2.0 g L1
, respectively. The observed improvement can be attributed to the increase in surface area and
-p
availability of more active adsorption sites due to the increasing dosage [55]. On the other hand, further increase reduced the efficiency down to 57.76%, which was obtained for 3.0 g L-1. Since
re
the highest efficiency value, 94.47%, was achieved with the geopolymer dosage of 2.0 g L-1,
lP
this value was used to examine the effect of initial dye concentration.
na
3.5.2. The effect of initial dye concentration
The removal of pollutants in high concentrations is one of the success parameters used to evaluate the efficiency of a treatment process. Therefore, in this stage, BY2 concentrations
ur
varying from 10 to 50 mg L-1 were studied at three different temperatures, 293, 313 and 323 K,
Jo
for the constant GEO dosage of 2.0 g L-1 and the natural pH of 9.85. Fig. 9 exhibits the obtained efficiency values as a function of reaction time.
Fig. 9
18
According to Fig. 9, the efficiency decreased with the increasing dye concentrations for all of the three-temperature-set tested. This reduction can be attributed to the insufficient surfaceactive cites as a result of the increasing BY2 concentrations. However, for a specific point of the residence time, the differences between the maximum and minimum efficiency values decreased with the increasing temperature. For the constant reaction time of 300 min, the differences in the efficiencies are 20.50, 16.22 and 5.38% for the respective temperatures of 293, 313 and 323 K. Fig. 9 also showed that the equilibrium time of the adsorption reduced
ro of
with the increasing temperature. The reaction equilibrium was reached after about 700 min for 293 K and 300 min for both 313 and 323 K. In particular, as seen explicitly from Fig. 9-(c), the reaction almost attained to its equilibrium in 60 min at 323 K for most of the dye concentrations
-p
tested.
The obtained results were figured for the fixed BY2 concentration of 50 mg L-1, which can be
re
seen from Fig. 10. Similarly, the efficiency increased with temperature, and the differences in
lP
the efficiencies for a certain time decreased with the increasing reaction time. According to Fig. 10, for the reaction time of 300 min, the efficiency values of 64.19, 76.71 and 88.10% were obtained at the respective temperatures of 293, 313 and 323 K. The differences between the
na
maximum and minimum efficiency values were 46.06, 23.91 and 6.11% for the constant
Jo
Fig. 10
ur
residence time values of 120, 300 and 720 min, respectively.
4. Conclusion
In this study, a class F fly ash was used as alumina-silicate source for the preparation of a geopolymeric material which was then tested as an adsorbent for the removal of Basic Yellow 2 from aqueous solutions. The characterization results showed that the fly ash particles with 19
mostly smooth spherical shape turned into a porous and amorphous polymeric structure, sodium-alumina-silicate hydrate gel, as a result of the geopolymerization process. Specifically, the BET surface area and total pore volume values increased from 6.090 m 2 g-1 and 0.008 cm3 g-1 to the respective 47.380 m2 g-1 and 0.156 cm3 g-1. The experimental results indicated that the removal efficiency of BY2 extensively increased from 13.28 to 94.47% for the geopolymer dosages of 0.5 and 2.0 g L-1, respectively. However, the efficiency dramatically decreased down to 57.76% for 3.0 g L-1. On the other hand, efficiency reduced with increasing BY2
ro of
concentrations. At 293 K for 300 min, the efficiency values of 84.69 and 64.19% was obtained for the respective BY2 concentrations of 10 and 50 mg L-1. Based on the isothermal and thermodynamic investigations, the experimental results well suited with the Langmuir and
-p
Temkin models, and the adsorption can be considered as a spontaneous and endothermic process. In addition, the equilibrium adsorption isotherms resemble the Type-I, which is
re
characterized with the monolayer chemical adsorption. Kinetic studies exhibited that the
lP
pseudo-second order and the intra-particle diffusion models both occur simultaneously for 293 K while the pseudo-second order model is valid for 313 and 323 K. In conclusion, the prepared geopolymer gave promising results for the removal of cationic Basic Yellow 2 from aqueous
na
solutions. However, further research is required to achieve a shorter and more applicable
ur
equilibrium time for the adsorption process.
Jo
Declaration of Interest Statement We would like to confirm that there is no known conflict of interest associated with this publication.
Acknowledgement The technical support by East Anatolia High Technology Application and Research Center (DAYTAM) of Atatürk University is thankfully appreciated. 20
References
Jo
ur
na
lP
re
-p
ro of
[1] A.A. Siyal, M.R. Shamsuddin, M.I. Khan, N.E. Rabat, M. Zulfiqar, Z. Man, J. Siame, K.A. Azizli, Journal of Environmental Management, 224 (2018) 327-339. [2] L. Li, S. Wang, Z. Zhu, Journal of Colloid and Interface Science, 300 (2006) 52-59. [3] Y. Liu, C. Yan, X. Qiu, D. Li, H. Wang, A. Alshameri, Journal of the Taiwan Institute of Chemical Engineers, 59 (2016) 433-439. [4] R.M. Novais, G. Ascensão, D.M. Tobaldi, M.P. Seabra, J.A. Labrincha, Journal of Cleaner Production, 171 (2018) 783-794. [5] N.B. Singh, G. Nagpal, S. Agrawal, Rachna, Environmental Technology & Innovation, (2018). [6] R.M. Novais, J. Carvalheiras, D.M. Tobaldi, M.P. Seabra, R.C. Pullar, J.A. Labrincha, Journal of Cleaner Production, 207 (2019) 350-362. [7] R.M. Novais, L.H. Buruberri, M.P. Seabra, J.A. Labrincha, J Hazard Mater, 318 (2016) 631640. [8] M. Zhang, M. Zhao, G. Zhang, D. Mann, K. Lumsden, M. Tao, Constr Build Mater, 124 (2016) 373-382. [9] I. Niklioć, S. Marković, I. Janković – Častvan, V.V. Radmilović, L. Karanović, B. Babić, V.R. Radmilović, Materials Letters, 176 (2016) 301-305. [10] K. Al-Zboon, M.S. Al-Harahsheh, F.B. Hani, Journal of Hazardous Materials, 188 (2011) 414-421. [11] T. Luukkonen, H. Runtti, M. Niskanen, E.-T. Tolonen, M. Sarkkinen, K. Kemppainen, J. Rämö, U. Lassi, Journal of Environmental Management, 166 (2016) 579-588. [12] H. Runtti, T. Luukkonen, M. Niskanen, S. Tuomikoski, T. Kangas, P. Tynjälä, E.-T. Tolonen, M. Sarkkinen, K. Kemppainen, J. Rämö, U. Lassi, Journal of Hazardous Materials, 317 (2016) 373-384. [13] H. Douiri, S. Louati, S. Baklouti, M. Arous, Z. Fakhfakh, Applied Clay Science, 139 (2017) 40-44. [14] S. Selmani, A. Sdiri, S. Bouaziz, E. Joussein, S. Rossignol, Applied Clay Science, 146 (2017) 457-467. [15] Y. Zhang, L. Liu, Particuology, 11 (2013) 353-358. [16] X.Y. Zhuang, L. Chen, S. Komarneni, C.H. Zhou, D.S. Tong, H.M. Yang, W.H. Yu, H. Wang, Journal of Cleaner Production, 125 (2016) 253-267. [17] S.H. Asl, M. Ahmadi, M. Ghiasvand, A. Tardast, R. Katal, Journal of Industrial and Engineering Chemistry, 19 (2013) 1044-1055. [18] S. Zhang, S. Ravi, Y.-R. Lee, J.-W. Ahn, W.-S. Ahn, Journal of Industrial and Engineering Chemistry, 72 (2019) 241-249. [19] P. Duan, C. Yan, W. Zhou, D. Ren, Ceramics International, 42 (2016) 13507-13518. [20] I. Acar, M.U. Atalay, Fuel, 180 (2016) 97-105. [21] I. Acar, M.U. Atalay, Fuel, 106 (2013) 195-203. [22] J.M. Lim, J. Park, J.T. Park, S. Bae, Journal of Industrial and Engineering Chemistry, 71 (2019) 378-386. [23] N.A. Mazumder, R. Rano, Journal of Industrial and Engineering Chemistry, 29 (2015) 359-365. [24] S. Zeng, J. Wang, Constr Build Mater, 121 (2016) 386-399. [25] V. Nikolić, M. Komljenović, N. Marjanović, Z. Baščarević, R. Petrović, Ceramics International, 40 (2014) 8479-8488. [26] A. Öztürk, E. Malkoc, Applied Surface Science, 299 (2014) 105-115.
21
Jo
ur
na
lP
re
-p
ro of
[27] T.L. Robl, J. Groppo, in Method for Hydraulically Separating Carbon and Classifying Coal Combustion Ash, U.S. Patent Ed., University of Kentucky Research Foundation, United States (2011). [28] M.J. McCarthy, M.R. Jones, L. Zheng, T.L. Robl, J.G. Groppo, Fuel, 111 (2013) 430-441. [29] J.L. Chang, J.C. Ma, Q.L. Ma, D.D. Zhang, N.N. Qiao, M.X. Hu, H.Z. Ma, Applied Clay Science, 119 (2016) 132-140. [30] M.B. Ogundiran, S. Kumar, Constr Build Mater, 125 (2016) 450-457. [31] I. Lecomte, C. Henrist, M. Liegeois, F. Maseri, A. Rulmont, R. Cloots, Journal of the European Ceramic Society, 26 (2006) 3789-3797. [32] Y. Liu, C. Yan, Z. Zhang, H. Wang, S. Zhou, W. Zhou, Fuel, 185 (2016) 181-189. [33] A.A. Siyal, K.A. Azizli, Z. Man, L. Ismail, M.I. Khan, Ceramics International, 42 (2016) 15575-15584. [34] I. Acar, O. Acisli, Applied Surface Science, 457 (2018) 208-213. [35] D.M. Tobaldi, A. Tucci, A.S. Škapin, L. Esposito, J. Eur. Ceram. Soc., 30 (2010) 24812490. [36] S. Kumar, F. Kristály, G. Mucsi, Adv. Powder Technol., 26 (2015) 24-30. [37] D. Panias, I.P. Giannopoulou, T. Perraki, Colloids Surf. Physicochem. Eng. Aspects, 301 (2007) 246-254. [38] S. Andini, R. Cioffi, F. Colangelo, T. Grieco, F. Montagnaro, L. Santoro, Waste Manage., 28 (2008) 416-423. [39] M. Thommes, K. Kaneko, A.V. Neimark, J.P. Olivier, F. Rodriguez-Reinoso, J. Rouquerol, K.S. Sing, Pure and Applied Chemistry, 87 (2015) 1051-1069. [40] A. Gurses, A. Hassani, M. Kiransan, O. Acisli, S. Karaca, J Water Process Eng, 2 (2014) 10-21. [41] L. Eskandarian, M. Arami, E. Pajootan, J Chem Eng Data, 59 (2014) 444-454. [42] C. Namasivayam, D. Kavitha, Dyes Pigments, 54 (2002) 47-58. [43] P.V. Temkin M.J., Acta Physico-Chimica Sinica, 12 (1940) 217-222 (URSS). [44] M.U. C. Aharoni, J. Chem. Soc. Faraday Trans. 73 (1977) 456–464. [45] Y. Bulut, H. Aydin, Desalination, 194 (2006) 259-267. [46] S. Lowell, J.E. Shields, M.A. Thomas, M. Thommes, Characterization of porous solids and powders: surface area, pore size and density, Springer Science & Business Media, (2012). [47] Y.S. Ho, G. McKay, Process Biochem, 34 (1999) 451-465. [48] S. Lagergren, Handlingar, 24 (1898) 1-39. [49] W.J. Weber, J.C. Morris, Journal of the Sanitary Engineering Division, 89 (1963) 31-60. [50] V. Basavarao, S. Rammohanrao, Chemical Engineering Journal, 116 (2006) 77-84. [51] D. Kavitha, C. Namasivayam, Bioresour Technol, 98 (2007) 14-21. [52] M. Dogan, Y. Ozdemir, M. Alkan, Dyes Pigments, 75 (2007) 701-713. [53] A. Misra, R. Gupta, R. Gupta, in Utilization of marble slurry in construction materials, (2008). [54] K. Al-Zboon, M.S. Al-Harahsheh, F.B. Hani, J. Hazard. Mater., 188 (2011) 414-421. [55] Ö. Açışlı, Preparation of Fe3O4/organo-montmorillonite nanocomposite and its use as an adsorbent for removal of an anionic dye, (2019).
22
Figures captions: Fig. 1. Particle size distribution of the materials. Fig. 2. Mineralogical analysis of the materials. Fig. 3. SEM images of CUFA (a-d) and GEO (e-h). Fig. 4. EDX spectra of the materials. Fig. 5. FT-IR spectra of CUFA and GEO. Fig. 6. Surface area and pore properties of the materials.
ro of
Fig. 7. The obtained isotherm patterns at three different temperatures for 300 min.
Fig. 8. The effect of adsorbent dosage on the removal efficiency of BY2 as a function of time (a) for the processing time of 900 min (b). Experimental conditions: Natural pH = 9.85,
-p
[BY2]0 = 30 mg L-1 and temperature = 293 K.
Fig. 9. The effect of initial BY2 concentration on the removal efficiency at 293 K (a), 313 K
re
(b) and 323 K (c). Experimental conditions: Natural pH = 9.85 and GEO dosage = 2 g L-1.
lP
Fig. 10. The effect of temperature on the removal efficiency. Experimental conditions: Natural
na
pH = 9.85, GEO dosage = 2 g L-1 and [BY2]0 = 50 mg L-1.
Jo
ur
Figures
23
30 GEO
CUFA
CFA
Intensity (%)
25 20
15 10
0 0.1
1
10
100
Jo
ur
na
lP
re
-p
Particle size (μm)
Fig. 1
24
ro of
5
1000
10 20 30 40
Position
Fig. 2
25
(o2Theta) 50
48.10-M 49.40-M-H 50.15-Q L
45.64-Q L
42.54-M-QL
40.83-M-H
39.23-M-QL -H
35.22-M 36.53-Q L 37.00-M
33.20-M-H
60
70
81.10-Q L
74.19-M 74.99-M-H
70.36-M
68.06-Q L
64.47-M
59.88-Q L 60.59-M
57.52-M-H
53.90-M-H
16.45-M
25.00-M 26.25-M
20.86-Q L
16.45-M
30.96-M
ro of
-p
re
lP
na
ur
Jo
68.12-Q L
59.99-Q L-M
50.15-Q L
36.55-Q L
26.63-Q L
20.89-Q L
Intensity (count)
26.67-Q L
QL: Quartz M: Mullite H: Hematite
GEO
Mullite; 29 Quartz; 71
Hematite; 8.9
CUFA
Quartz; 29.7
Mullite; 61.4
80
(a)
(b)
(c)
(d)
Solid sphere
ro of
Hollow sphere (Cenosphere)
-p
Unburned carbon
Irregularly-shaped particles
re
Irregularly-shaped particles
(f)
Unreacted CUFA
na
lP
(e)
ur
(h)
Jo
(g)
GEO
26
Fig. 3 CUFA
Element Weight % O 47.44 Mg 0.57 Al 16.35 Si 26.84 K 3.65 Fe 5.16
Si O
K
K
Mg Fe
Fe
K
0
2
4
Fe
6
8
Energy (keV) GEO
Element Weight % O 41.79 Na 11.90 Mg 0.30 Al 14.08 Si 24.08 K 1.17 Fe 6.66
Intensity
-p
Si
O
re
Al
Na KK
Mg
0
Fe
lP
K Fe
10
ro of
Intensity
Al
2
4
6
Jo
ur
na
Energy (keV) Fig. 4
27
Fe
8
10
4000 3600 3200
2800
2400
Fig. 5
28 2000
CUFA
1600
Wavenumber (cm-1)
1200
1200
720 590
1600
Si-O-Si/Al stre tching
2000
ro of
2400
Si-O-Si/Al stretching Si-O/Al stretching
880 Al-O stretching
1440 Na 2CO 3 / O-C-O stretching
1660 C=O stretching
1974 C≡C stretching
2160 Si-H Stretching
472 452
Transmittance (%)
~2500 H-O-H Vibration
980 Al-O-Al/Si-O-Si asym stretching
2800
1636 -OH Stretching
3200
-p
re
3600
2351 -OH /Si-H Stretching
3433 H-O-H Vibration
Transmittance (%) 4000
lP
na
ur
Jo GEO
800
800
400
400
8
CUFA 7
Parameter
CUFA
GEO
6
BET surface area (m2/g)
6.09
47.38
BJH cumulative surface area (m2/g)
5.717
51.63
Total pore volume (cm3/g) Average pore width (nm)
0.008 4.97
0.156 9.792
5 4
2
0.006
GEO
1
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
GEO
100 80
60 40 20
0 0.0
(a)
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
CUFA 0.005
0.004
0.003
0.002
0.001
0.000
1.0
1
Relative pressure (p/p0)
(b)
Jo
ur
na
lP
re
Fig. 6
29
10
Pore width (nm)
-p
0
ro of
0
Pore volume, dV/dlog(w) (cm³/g·nm)
Quantity adsorbed (cm³/g STP)
3
100
50
q (mg/g)
40
30
20 323 K 10
313 K
0 0
5
10
ro of
293 K 15
Ce (mg/L)
Jo
ur
na
lP
re
-p
Fig. 7
30
20
Removal efficiency (%)
100
(a)
80
60
40
0 0
200
400
600
Time (min) 1.0 g/L
1.5 g/L
Removal efficiency (%) 100
(b)
94.47
re
57.76
lP
29.60
40 13.28
ur
0.5
1.0
12.21
7.52
4.44
1.99
Jo
3.0 g/L
50.11
60
na
Uptake (%)
80
0
2.5 g/L
83.26
q (mg/g)
20
2.0 g/L
800
-p
0.5 g/L
ro of
20
1.5
2.0
12.49
2.5
Adsorbent dosage (g/L) Fig. 8
31
8.66
3.0
100
(a)
Removal efficiency (%)
84.69
80
60
10 mg/L
64.19
20 mg/L 40 30 mg/L
40 mg/L
20
0
0
100
200
300
400
500
600
Time (min) 100
re
60
800
900
10 mg/L
20 mg/L 30 mg/L
lP
40
20
0
700
-p
80
na
Removal efficiency (%)
(b)
ro of
50 mg/L
100
200
300
400
50 mg/L 500
Time (min)
Jo
ur
0
40 mg/L
32
600
700
800
900
100
Removal efficiency (%)
(c) 80
60 10 mg/L
40
20 mg/L 30 mg/L 20
40 mg/L 50 mg/L
0 100
200
300
400
500
600
Time (min) Fig. 9
100
60
lP
64.19
40
20
0
100
Jo
ur
0
200
300
900
-p
76.71
800
re
80
na
Removal efficiency (%)
88.10
700
ro of
0
323 K 313 K
293 K 400
500
Time (min) Fig. 10
33
600
700
800
900
Table 1. Chemical composition and basic physical properties of the materials. Chemical composition (%)
CFA
Sia
57.09
55.34
26.84a
24.08a
Al2O3
Ala
27.46
29.39
16.35a
14.08a
Fe2O3
Fea
6.56
5.97
5.16a
6.66a
1.97
1.20
NIb
Mga
2.38
2.22
0.57a
0.30a
Na2O
Naa
0.31
0.26
NIb
11.90a
K2O
Ka
3.89
4.21
3.65a
1.17a
ro of
MgO
TiO2
1.13
1.23
NIb
SO3
<0.01
<0.01
NIb
Oa
NIb
LOI
1.68
Physical properties 39.29
BET surface area (m2 g-1)
0.83
re
Mean size (d50) by volume (μm)
NIb
47.44a
NIb
5.25
≈ 0.70
6.09
47.38
Jo
ur
na
lP
Determined by EDX elemental analysis (The average values of 10 measurements) Not included
Table 2. Parameters of the adsorption isotherms used.
34
41.79a
2.07
-p
-
b
GEO
SiO2
CaO
a
CUFA
Equation
Langmuir
C/q = 1/kqm+(1/qm)C
Freundlich
lnq = lnk+nlnC
Temkin
Parameters
293 K
313 K
323 K
qm k R2 RL n k R2 bT aT R2 qm k R2 K qDR R2 B A R2
36.364 0.282 0.917 0.151 1.006 2.496 0.788 0.293 1.161 0.987 -7.424 0.951 0.146 0.000 12.485 0.007 1.500 227.273 0.177
41.667 0.490 0.924 0.093 0.913 4.792 0.586 0.334 1.826 0.903 21.739 0.810 0.146 0.000 13.924 0.079 1.037 185.185 0.371
53.476 0.501 0.730 0.091 1.200 6.043 0.522 0.513 1.690 0.893 53.191 0.533 0.047 0.000 14.421 0.001 0.699 136.986 0.367
qe = (RT/bT)lnaT+(RT/bT)lnCe
BET
C/q(1-C) = 1/(qmk)+[(k-1/qmk)]C lnq = Kε2 + lnqDR
Harkins-Jura
1/q2 = (B/A)-(1/A)logC
Jo
ur
na
lP
re
-p
DubininRadukevisch
ro of
Isotherm
Table 3. Parameters of the kinetic models. Initial Temp. conc. (K) (mg L-1)
Pseudo-first order R2
Pseudo-second order k2
qe,exp 35
qe,cal
Intra-particle diffusion R2
ki1
C
R2
0.7390 0.9893 0.9931 0.9908 0.9920 0.9991 0.9966 0.9974 0.9970 0.9974 0.9951 0.9999 0.9998 0.9995 0.9997
re lP na ur 36
(mg g-1 s-1/2) 0.0301 0.0465 0.0697 0.0763 0.0789 0.0191 0.0379 0.0523 0.0634 0.0669 0.0144 0.0220 0.0310 0.0395 0.0445
0.3685 1.4366 2.0443 5.2287 8.9575 1.5660 2.4567 4.3200 6.9164 10.7960 2.3588 5.7297 8.8397 12.0790 16.0710
ro of
0.6259 0.6422 0.6628 0.6343 0.7173 0.2085 0.6447 0.5930 0.6718 0.6640 0.0113 0.1714 0.2878 0.3446 0.4028
(mg g-1) 6.0790 10.9649 16.1031 20.1613 24.4499 5.3996 10.3413 15.1286 20.0803 24.6305 5.1706 9.7847 14.6199 19.6078 24.5098
-p
293 293 293 293 293 313 313 313 313 313 323 323 323 323 323
(mg g-1) 3.9405 7.6216 11.4836 15.3974 19.4578 4.7940 8.9319 12.6819 17.1216 21.5181 4.6216 9.4491 14.0181 18.4233 23.3371
Jo
10 20 30 40 50 10 20 30 40 50 10 20 30 40 50
(g mg-1 s-1) 0.00001 0.00001 0.00001 0.00001 0.00001 0.00008 0.00003 0.00002 0.00002 0.00002 0.00010 0.00014 0.00010 0.00006 0.00006
0.9326 0.9572 0.9372 0.9456 0.9574 0.8507 0.9068 0.9102 0.9168 0.9199 0.5898 0.6053 0.6532 0.6768 0.6781
Table 4. Thermodynamic parameters of the geopolymer-BY2 adsorption system. T Ca (K) (mg L-1) 293 4.757 Geopolymer 313 2.136 (Co:20 mg/L) 323 1.102
q 1/T lnT lnC (mg g-1) 7.622 0.003 5.680 1.438 8.932 0.003 5.746 0.792 9.449 0.003 5.778 0.265
ΔGo (kJ mol-1) -27.672 -30.994 -32.047
ΔHads (kJ mol-1)
ΔSads (kJ mol-1)
29.752
-0.097
Jo
ur
na
lP
re
-p
ro of
Adsorbent
37