You are on page 1of 37

Reaction Kinetics, Mechanisms and Catalysis

Preparation of powerful exchanged Faujasite zeolite materials used as effective


heterogeneous catalyst for photo-Fenton oxidation of methyl orange (MO)
--Manuscript Draft--

Manuscript Number: REAC-D-23-00417R1

Full Title: Preparation of powerful exchanged Faujasite zeolite materials used as effective
heterogeneous catalyst for photo-Fenton oxidation of methyl orange (MO)

Article Type: Full Paper

Corresponding Author: Olfa Ouled Ltaief


Ecole Nationale d'Ingenieurs de Sfax
TUNISIA

Corresponding Author Secondary


Information:

Corresponding Author's Institution: Ecole Nationale d'Ingenieurs de Sfax

Corresponding Author's Secondary


Institution:

First Author: Olfa Ouled Ltaief

First Author Secondary Information:

Order of Authors: Olfa Ouled Ltaief

Sophie Fourmentin

Stéphane Siffert

Mourad Benzina

Order of Authors Secondary Information:

Funding Information:

Abstract: Novel heterogeneous photo-Fenton catalyst, iron coated Faujasite zeolite (Fe-FAU)
was synthesized by exchange-calcination method. Iron was exchanged in zeolite
support based on Faujasite type Zeolites prepared in laboratory from instratified illitic-
kaolinite raw Tunisian clay. The studied materials were characterized by XRD, BET,
SEM-EDX and FTIR methods. X-ray diffraction (XRD) analysis showed that the zeolite
structure was not affected by the catalyst preparation method. BET results revealed
that Fe-FAU possess a trimodal structure (micro,meso and macropores) with a porosity
of 0.28 cm3/g. The catalytic performance of the Fe-FAU was studied for the oxidation
of methyl orange (MO) in aqueous solution in the presence of hydrogen peroxide. The
effects of different operating parameters on the degradation efficiency were studied at
atmospheric pressure. The results indicated that the Fe doped zeolite exhibited good
degradation efficiency. Measurements of the mineralization of MO by TOC analysis
showed that the maximum degradation of MO (94%) was detected with 50 mg of
adsorbent dosage after 20 min of contact time and under UVC irradiation. Moreover, in
order to predict the empirical variables significance, the variances analysis “ANOVA”
was used. Response Surface Methodology (RSM) incorporating Central Composite
Design (CCD) of experiments was applied to optimize oxidation parameters. High
regression (R2 = 0.95) and the low probability (p value <0.0001) values signify the
validity of the quadratic model to predict the removal (%) of MO.

Response to Reviewers: Response to Reviewers' Comments

Manuscript ID: REAC-D-23-00417


Title: Preparation of Powerful exchanged Faujasite zeolite materials used as effective
heterogeneous catalyst for Photo-Fenton oxidation of Methyl Orange (MO)
Authors: Olfa Ouled Ltaief*, Sophie Fourmentin, Stéphane Siffert, Mourad Benzina

Dear Editor, First of all, the authors would like to express deep respect and thanks to
the editor and the reviewer for their efforts in reviewing our manuscript “Preparation of

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Powerful exchanged Faujasite zeolite materials used as effective heterogeneous
catalyst for Photo-Fenton oxidation of Methyl Orange (MO).»
We have carefully considered all comments and here is a point-by-point response to
the reviewers’ comments.

Reviewer #1:
Overall Recommendation: Major Revision

In this work, the author has reported a methodology to synthesis Fe doped zeolite,
which exhibits an excellent capability in the degradation of methyl orange. Meanwhile,
the author proposed a statistical method to optimize the parameters of the degradation.
However, this manuscript has several issues to be addressed, and the reviewer agrees
with the publication of this manuscript in Reaction Kinetics, Mechanisms and Catalysis
after improvement of the important concerns:
1-Please make sure the format of this manuscript was further modified, include the
upper and lower case of the words, position of the sub-headings, and the spaces in the
manuscript et al.
Author reply:
has been changed as requested

2.In section 2.3, the author has mentioned that "the adsorption of organic contaminant
plays a significant role in the kinetics of the degradation", and "the solution was
equilibrated during 30 min to analyze the adsorption of MO by the catalyst". However,
the related result has not been presented in this manuscript and the supplementary
data, and just "a small decrease of MO was detected after 5min that describes the
adsorption equilibrium" was mentioned in section 3.3.1.
Author reply:
We displayed the section related to the adsorption on the same degradation graphs
since, in reality, the adsorption measurement revealed that the true or actual
adsorption occurs during the final five minutes of the overall solution equilibrium time
prior to the injection of H2O2. Furthermore, to describe the effect correctly, the
sentence on the manuscript «the adsorption of organic contaminant plays a significant
role in the kinetics of the degradation » was replaced by «the adsorption of an organic
contaminant on the surface of photo-catalyst have an impact in the kinetics of
degradation.»

3-The reviewer was confused about the statement of the XRD characterization of the
as-prepared samples. The XRD of the Fe-FAU is insufficient to prove the insertion of
the metal ions. Besides, the statement "These facts confirmed that the framework and
crystallinity of zeolite were not destroyed" is in opposition to "It can be concluded that
the insertion of metal ions into the Faujasite structure affects the crystallinity but not the
zeolite structure".
Author reply:
The author wants to clarify that the structure of the faujaiste zeolite has not been
destroyed and that the same crystal network with the same characteristic pics has
been conserved after catalyst preparation. The only change noted is the decrease in
pic height that occurs during the incorporation of ferric ions into the zeolitic structure;
thus, the crystallinity has not been completely destroyed but has only been slightly
impacted. Our results and interpretations are similar to previous work (reference 26 on
the manuscript).

4-The symbols in figure 1 should be aligned with the descriptions.


Author reply:
has been changed and replaced.

5-The reviewer expects to receive an explanation from the author about the XRD
patterns in which the characteristic diffraction peaks of Fe-FAU are shifted to different
distances and in different directions, when it was compared to the result of FAU.
Author reply:
The prepared material (Fe-FAU)exhibit similar peaks to those of the pure FAU zeolite.
So, no new crystalline pattern appears; the only change is the decrease of the intensity
of characterstic peaks.In fact, the exchange-calcination method has not affected the d-
spacing on the structure of Faujasite to any considerable degree. This demonstrates

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
that exchange process of ferric ions has taken place only on the surface of the zeolite
and doesn’t change the crystal structure of the zeolite material. Our results are similar
to other reported in literature (references; 24 and 26 on the manuscript) .

6-The order of the positions for Al and Si in Fig.2 (a) and Fig. 2(b) is opposite, please
exhibit the corrected SEM images, and the SEM results of Fe-FAU is needed if
possible.
Author reply:
It was a typing error for the SEM image of the FAU that has been corrected on the
figure. However, we do not have a SEM image for the Fe-FAU.

7-Generally, the specific surface area and pore volume distribution of the catalyst are
related to the N2 isothermal adsorption and desorption profiles of the samples. The
description in the main text and Table 1are inconsistent with Fig. 3.
Author reply:
The results obtained for the textual analysis of samples and their interpretations have
been published in previous works (reference [23] on the manuscript), all interpretations
were drawn from the literature.

8-Multiple errors have been noticed in this manuscript in the serial number of the
images, which can not match the descriptions. Such as, the "Fig. 4" mentioned in Page
7 Line 16, which was descripted as the nitrogen adsorption-desorption isotherms of
prepared catalyst, however, Fig .4 was the FTIR results of the samples.
Author reply:
Figure numbers and tables have been checked and corrected when needed.

9-Please give more information in Figure S1 to show the reviewer, which is the raw
clay, FAU and Fe-FAU, respectively.
Author reply:
Well done .

10-The author has mentioned that degradation was more efficient than color removal
during the catalytic process, but there are no graphic results about the variations of the
TOC.

Author reply:
Because we are limited on the number of figures and tables that can be included in the
manuscript due to space constraints so we have chosen to show the figures
corresponding to the MO concentration variation. However, we include here a figure
that show the variation of the TOC with Fe-FAU catalyst dose.

11-The author is expected to give an explanation for the phenomenon that further
increases in the dose of catalyst more than 50 mg did not result in catalytic
performance improvement in section 3.3.1, which is inconsistent with the results of
section 3.5.
Author reply:
For the catalytic experiments the catalytic system reaches a saturation point beyond 50
mg of catalyst. This means that after the threshold, additional catalyst dose does not
contribute significantly to the improve of degradation rate of MO as the active sites may
already be saturated or the reaction may be limited by other factors. While for statistical
analysis with ANOVA, it is true that for the study of the the mutual effect of catalyst
dose (mg), MO initial concentration (mg/L) and H2O2 dosage (µL) on MO removal (%)
with 3D surface plots, the effect of variation of catalyst dose is not so clear. However,
the latter appears clearly in Figure S6. which presents the maximum desirability ramp
for the optimum combination of zeolite dose , H2O2 dosage and MO concentrations to
obtain maximum degradation efficiency. For which, the optimum oxidation efficiency of
MO was obtained with an optimal catalyst dose equal to 69.1403 mg of Fe-FAU, .

12-Which are the blank control experiments in the degradation of MO, were the
degradation of MO by UVC (λ=254 nm) without photocatalyst in 60 min taken in
consideration? Please show the related results.
Author reply:
When preliminary experiments were carried out without photocatalyst, in presence of

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
MO and H2O2 under UVC irradiation, the degradation rate of MO was about 35% by
the addition of H2O2 to the reaction without Fe-FAU catalyst. This should be assigned
to the amount of OH radicals released by H2O2 through direct photolysis reactions
when the mixture of MO and hydroperoxide was stirred for 60 minutes. It is worth
noting that the combination of hydroperoxide and UV irradiation enhanced the removal
of MO. This should be assigned to the synergistic effect of the UV photon energy on
increasing the rate of H2O2 decomposition to produce free radical OH•,which in turn
improve the reaction rate of MO oxidation.

13.Fig. 6 has shown the remained H2O2 in the solution, however, the method to detect
the H2O2 has not been proposed in the manuscript.
Author reply:
The method used is added to the characterization techniques section.

14-There was a conflict in Page 11 Line 48, where you have declared the duration time
of each test was 2 h, while the description in Page 12 Line 3 was 60 min for the fifth
cycle.
Author reply:
The whole irradiation exposure time was two hours, although maximum degradation
rate was detected within sixty minutes.

15.The author has to give a relevant specification for the different sampling intervals in
Fig S4.
Author reply:
the description of total irradiation times and the steps used have been added to section
3.2.4.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Manuscript Click here to access/download;Manuscript;REAC-D-23-
00417.doc
Click here to view linked References

1
2
3
4
5
Preparation of powerful exchanged Faujasite zeolite materials used as
6 effective heterogeneous catalyst for photo-Fenton oxidation of methyl
7 orange (MO)
8
9 Olfa Ouled Ltaief1,2, Sophie Fourmentin², Stéphane Siffert², Mourad Benzina1
10
11 1
Water, Energy and Environment Laboratory, Code AD 10-02 ENIS, B.P. 3038, Sfax, Tunisia
12
13 2ULCO, UCEIV, Unit of Environmental Chemistry and Interactions with the Living Organisms,
14
15 E.A.4492, Dunkerque, France
16
17
18
19 E-mail: olfa.ol@live.fr
20
21
22
23
Abstract Novel heterogeneous photo-Fenton catalyst, iron coated Faujasite zeolite (Fe-FAU) was
24 synthesized by exchange-calcination method. Iron was exchanged in zeolite support based on
25 Faujasite type Zeolites prepared in laboratory from instratified illitic-kaolinite raw Tunisian clay.
26
27 The studied materials were characterized by XRD, BET, SEM-EDX and FTIR methods. X-ray
28 diffraction (XRD) analysis showed that the zeolite structure was not affected by the catalyst
29
30 preparation method. BET results revealed that Fe-FAU possess a trimodal structure (micro,meso and
31 macropores) with a porosity of 0.28 cm3/g. The catalytic performance of the Fe-FAU was studied
32
33 for the oxidation of methyl orange (MO) in aqueous solution in the presence of hydrogen peroxide.
34 The effects of different operating parameters on the degradation efficiency were studied at
35
atmospheric pressure. The results indicated that the Fe doped zeolite exhibited good degradation
36
37 efficiency. Measurements of the mineralization of MO by TOC analysis showed that the maximum
38 degradation of MO (94%) was detected with 50 mg of adsorbent dosage after 20 min of contact time
39
40 and under UVC irradiation. Moreover, in order to predict the empirical variables significance, the
41 variances analysis “ANOVA” was used. Response Surface Methodology (RSM) incorporating
42
43 Central Composite Design (CCD) of experiments was applied to optimize oxidation parameters.
44 High regression (R2 = 0.95) and the low probability (p value <0.0001) values signify the validity of
45
46 the quadratic model to predict the removal (%) of MO.
47
48
49
Keywords: Fe-FAU, Methyl Orange, oxidation, Photo-Fenton, ANOVA.
50
51
52 1.Introduction
53
54
Today, international environmental standards for water pollution are becoming more stringent,
55
56 therefore many researchers are working to reduce wastewater and have focused their attention on
57 developing technologies for removal of pollutants arising from various industries. In fact, Liquid
58
59 discharges from these manufacturing cause wastewater in large volume including hazardous
60 inorganic (metals) or organic (microbes, dyes/pigments) species. One of the most notable pollution
61
1
62
63
64
65
which is caused by paper, paint, textile, plastic industries is dye contaminants. Among dye
1 pollutants, aromatic structures such as methyl orange (MO) which is one of the commonest cationic
2
3 dyes used is a stable azo compound with biological toxicity [1]. The releasing of this Colored
4 organic compound into the environment is toxic to living beings and can have dangerous effects on
5
6 the human and aquatic life by generating serious undesirable effects [2]. In fact, in anaerobic
7 condition, they are reduced to potentially carcinogenic aromatic amines and long-term exposure to
8
9 MO can cause health problems such as anemia, vomiting, nausea, and hypertension [3]. Therefore,
10 several conventional treatments are widely used, such as coagulation [4], adsorption on activated
11
carbon [5] and membrane filtration [6]. Though these treatment processes are efficient in dye
12
13 removal they have many drawbacks like as partial ion elimination, great-energy consumption and
14 generation of waste which further causes a secondary pollution. Recently, among all available
15
16 processes, a big attention of many research groups has been fixed on the advanced oxidation
17 processes (AOPs) mostly photocatalysis, an environmentally friendly methods let treatment of dyes
18
19 in wastewater to the permissible limit at reasonable cost. These technologies have shown their
20 potential in treating toxic and organic pollutants because they convert contaminants, to a large
21
22 extent, into stable mineral products, such as carbon dioxide and water or biologically degradable
23 molecules [7]. These processes are based on the in situ formation of hydroxyl radicals (HO .) which
24
25
have a high oxidation power. The radicals (HO.) can incompletely or entirely mineralize most
26 organic compounds. Fenton process have been among the most common AOPs oxidation systems
27 used. The advantage of Fenton’s reaction in comparison with other advanced oxidation methods is
28
29 that this process offers a rich reaction medium of HO. radicals and it is simple to manipulate and
30 maintain. However, the homogeneous Fenton catalysis present the inconvenient of difficulty of
31
32 catalyst recovery and re-use. Thus, several efforts have been focused to evaluate the heterogeneous
33 photo-Fenton process due to the whole mineralization of the dissolved organic contaminants and the
34
35
easy recovery of used catalyst from the treated effluent. The best catalyst exhibits the lower iron
36 leaching during the photo-Fenton reaction. However, the problem of leaching manifests itself in the
37 majority of heterogeneous catalysts involved in the Photo-Fenton process. Thus, the most important
38
39 step in the heterogeneous photo-Fenton process is to prepare a heterogeneous catalyst with high
40 catalytic activity, sustainable sources, good stability and low cost. Many materials as Zeolites [8-
41
42 10], carbon materials [11-12], mesoporous and mixed oxides [13-14] and clay [15-17] have been
43 used as supports in the preparation of heterogeneous catalytic systems. Among these catalysts,
44
45 exchanged zeolites are preferred for improved efficiency, high exchange capacity, their higher
46 porosity and surface [18]. Pradeep et al. [19] have used Co ion exchanged ZSM-5, zeolite A and
47
zeolite X as effective photo-catalysts for decomposition of phenol. Also, Fe [20], Cu [21] and Ag
48
49 exchanged zeolites have been reported to act as photo-catalysts [22]. In the other hand, a big interest
50 was focused on zeolite materials synthesized by conversion of raw clays due to their hierarchical
51
52 structure, high exchange capacity, environmental compatibility, and reasonable costs. In fact, clay is
53 one of the most abundant and cheap materials widely utilized for numerous applications including
54
55 the oxidation of water pollutants. Nevertheless, clay due to the low surface area, small porosity, and
56 poor ion exchange capacity, is rarely used in its natural form for efficient dye degradation. Hence,
57
58
different modification methods are employed to improve the oxidation potential of clay materials
59 such as conversion to zeolitic materials.
60
61
2
62
63
64
65
In the present work, Faujasite type zeolite (FAU) was synthesized in laboratory from raw illitic
1 Tunisian clay [23] and used as a new support for catalytic oxidation of a dye. The choice of this
2
3 material is based on the diversity of its textural properties (hierarchical porosity) and its low
4 synthesis cost given the abundance of clay materials in our country. The aim of current work is to
5
6 examine the applicability of iron coated Faujasite zeolite (Fe-FAU) as novel heterogeneous catalyst
7 for the degradation of methyl orange (MO) dye as a model contaminant at room temperature and
8
9 atmospheric pressure. The Fe- FAU particles prepared were identified by different instrumental
10 techniques. The influence of reaction parameters such as initial MO concentration, catalyst dose,
11
and hydrogen peroxide concentration was studied. The degradation of MO was followed by the
12
13 measurement of TOC. The experiments were reproduced using response surface methodology
14 (RSM) to optimize the MO degradation efficiency (%). In fact, RSM has been frequently applied for
15
16 design of experiments and optimization processes by combination of mathematical equations and
17 statistical methods. Response surface methodology (RSM) is a robust technique that deduces an
18
19 empirical model which predict the experimental data by comparing with statistical variables. The
20 main objective of RSM is to optimize the oxidation reaction to make sure that the maximum dye
21
22 entities has been degraded via the Photo-Fenton reaction. By this, the lowest cost of the process
23 could be estimated during scale-up.
24
25
26 2. Experimental
27
28 2.1. Materials
29
30
Raw clay materials used for zeolite preparation were supplied from a deposit quarry situated in
31
32 southern Tunisia (Hamma, Gabes). Ferrous sulfate heptahydrate (Fe2SO4,7H2O), hydrogen peroxide
33 (H2O2, (30%w/w),), sodium hydroxide (NaOH, 98%), and Methyl Orange (C14H14N3NaO3S) were of
34
35 analytic reagent grade and purchased from Sigma-Aldrich. Distilled water was used to prepare all
36 the experimental solutions.
37
38
39 2.2. Preparation of Fe-FAU catalyst
40
41 Faujasite zeolite was synthetized via the alkali fusion method which consisted in two steps. Firstly,
42
43 raw illitic-kaolinite clay was washed three times with distilled water to remove surface contaminants
44 then placed in ceramic crucible and heated at 70 °C for 2 h in the hot air oven. The dried clay
45
46 material was crushed with the mortar to obtain fine particles (<50µm) and then mixed with NaOH
47 powder. The obtained NaOH-clay mixture was then placed in a nickel crucible (specific crucible
48
used for NaOH mixture to avoid parasite reactions with the walls at the fusion stage) and fused in a
49
50 muffle furnace at a temperature of 500 °C for 2 h. The second stage is the hydrothermal reaction
51 which consist in cooling the fused mixture to room temperature in vacuum desiccators., then well
52
53 grinded with mortar and mixed with the appropriate amount of distillated water (10 g of fused
54 product/70 mL of water) before vigorous stirring at room temperature for 12h. Obtained gel was
55
56 then putted into a 65 mL stainless autoclave filled at two-thirds of its total volume (taking into
57 account that the elevation of temperature results in the elevation of pressure within the autoclave)
58
59
and crystallized at 60°C for 24 h. The obtained precipitate was washed with distilled water to
60 remove excess alkaline until the pH reached around 9 and oven-dried at 70°C overnight [23].
61
3
62
63
64
65
Then, the catalyst material was prepared according to the following method: 10 g of synthesized
1 zeolite and 10 mg of (Fe2SO4·7H2O) was put in a Pyrex glass beaker with addition of 100 mL of
2
3 distilled water under stirring. the resulting product was dried at 60°C for 48 h. After that, the
4 precipitate was washed with distilled water until the supernatant was clear and dried at 105°C
5
6 overnight. Finally, the dried solid was calcined at 200°C for 5 h in a nitrogen atmosphere, and the
7 iron coated zeolite was obtained and called as Fe-FAU catalyst.
8
9
10 2.3. Characterization Techniques
11
12 Several characterization techniques were used to study the textural, structural, and chemical
13
14 properties of raw clay material, prepared FAU zeolite and Fe-FAU catalyst. Fourier transform
15 infrared absorption spectra (FTIR) were recorded using the KBr pellet technique on the spectrometer
16
FTIR Equinox 55 Bruker in the range 4000–400 cm-1. X-ray diffraction (XRD) technique was
17
18 employed to determine the mineralogical structure and the crystalline range for the studied
19 materials. The XRD patterns were carried out using X-ray diffraction apparatus (Bruker AXS D8
20
21 Advance diffractometer equipped with a copper anode,  = 1.5406Å). Data was collected in the
22 range of 5-60° with a step size of 0.02. Phase identification was performed by searching the ICDD
23
24 powder diffraction file database, with the help of JCPDS (Joint Committee on Powder Diffraction
25 Standards) files. The specific surface area, pore volume and pore size distribution were determined
26
27 from Nitrogen adsorption–desorption isotherms carried out at 77 K on a Sorptomatic 1990 series
28 apparatus. The micropore volume, Vmicro, was evaluated by the t-plot method method. The mesopore
29
size distribution was determined from the desorption isotherm by the Barrett–Joyner–Halenda (BJH)
30
31 method. The specific surface areas of the raw material and products were calculated by the
32 Brunauer–Emmett–Teller (BET) method. The morphologies and the qualitative chemical
33
34 composition of the raw clay and FAU zeolite were determined using Scanning Electron Microscope,
35 SEM (MEB LEO 438VP), coupled to an Energy Dispersive X-ray Spectrometer EDX, (ISIS 300).
36
37 The total organic carbon (TOC) was measured by a Shimadzu TOC-VCSH analyzer and expressed
38 as the percentage of the removal TOC. A colorimetric method was used to determine the remaining
39
40
hydrogen peroxide amount in the presence of titanium tetrachloride. The metal contents (ferric ions)
41 in the final solutions were analyzed using an AA-6300 Shimadzu atomic absorption
42 spectrophotometer (AAS).
43
44
45 2.4. Heterogeneous photo-Fenton oxidation of MO
46
47 The degradation experiments were performed using a montage specifically designed for this study.
48
49 A double-walled Pyrex reactor is installed in an irradiation chamber with a ventilation-rotation
50 system. Eight UV lamps with an intensity of 15W and different wavelengths (254, 310 and 360 nm)
51
52 are positioned regularly in an aluminum conserved tube in vertical hexagonal arrangement on the
53 walls of the chamber to develop the radiation via reflection. Catalytic reactions were performed in
54
55 the Pyrex cylindrical batch reactor of 250 mL. Stirring was applied to disperse catalyst particles. The
56 reactor assembly was surrounded by the UV lamps. In a typical run, the reactor was loaded with 50
57
mL of MO solution of known concentration. For each test, the solution was equilibrated during 30
58
59 min and then analyzed to determine the adsorption of MO by Fe-FAU. In fact, it is well known that
60 the adsorption of an organic contaminant on the surface of photo-catalyst have an impact in the
61
4
62
63
64
65
kinetics of degradation. That's why we have considered this phenomenon and before irradiation, the
1 solution was equilibrated under dark conditions for 30 min and then analyzed to determine the
2
3 adsorption of MO by the catalysts. Then, the beginning of the reaction (t = 0) was considered when
4 the H2O2 was added. In order to study the influence of various parameters on the degradation of MO
5
6 with photo-Fenton processes, several runs were conducted under UVC irradiation in the laboratory
7 to evaluate the effect of varying the catalyst dose (20-100 mg), initial concentration of the dye (20
8
9 ppm- 100 mg/L), the H2O2 dosage (50 μL to 300 μL) and the UV irradiation type. Samples were
10 taken at regular intervals of time and put through the syringe filter to the sample vials containers to
11
separate catalyst from the suspension and to determine TOCr.
12
13 The degradation percentage of TOC by catalysts was determined by a Shimadzu TOC-VCSH
14 analyzer and expressed as follows:
15
16 TOC0 - TOCr
17 ×100
18 Degradation percentage of TOC (%) = TOC 0 (1)
19
20
21
22 Here TOC0 and TOCr are the concentrations of initial and residual Total Organic Carbon. After each
23 experiment, the catalysts were removed by filtration and the metals contents in the final solutions were
24
analyzed using an AA-6300 Shimadzu atomic absorption spectrophotometer.
25
26
27 2.5. Experimental design and statistical analysis by CCD
28
29 Optimizations of the oxidation variables using statistical approaches are important for saving time and
30 resources. Furthermore, the statistical analysis gives clear information about the oxidation parameters
31
32 and reduces the experimental errors. Herein, the CCD statistical approach coupled with RSM was used
33 to optimize the oxidation of MO with Fe-FAU catalyst. In order to determine the range of operation
34
35
for a process, proper design of experiments should be followed and optimize the parameters. Catalyst
36 dosage, initial concentration of MO and H2O2 concentration are the parameters chosen to be studied
37 and optimized. The aim of this study is to predict extreme responses, so limit points must be
38
39 considered in the design. As well, CCD has an embedded factorial design, appropriate for successive
40 experiments. It comprises a rationalized number of design points for analyzing the input variables
41
42 within a given range, and the lack of fit would be checked by the information obtained from the
43 curvature estimation method. Here, CCD with 3 parameters, a full factorial run was experimented. The
44
45 method included 13 experiments with 12 trials for non-center points and one for center points. The
46 inferior and superior limits were fixed then the software makes design inside the limits of the
47
experiment. The coded variables and their equivalent limits are A) catalyst dose (20-100 mg), B) H2O2
48
49 concentration (50-300µL), C) initial MO concentration (20–100 mg.L−1). The Response Surface
50 Methodology computations for this study were executed with the help of Design-Expert software.
51
52 Statistical quadratic models based on interaction and polynomial relations were established for the
53 given response variables. The polynomial model with the upper order was chosen with the sequential
54
55 p-value ≤ 0.05, the lack of fit p value > 0.05 and maximum possible (<0.2) values for adjusted R 2 and
56 predicted R2. Also, other terms were examined for significance and selected the model, which is not
57
58
aliased. The selected model was examined via analysis of variance (ANOVA). Therefore, the best fit
59 model was selected after examination for any necessity for transformations to fit statistical molds from
60
61
5
62
63
64
65
the Box-cox plot. Eventually, the interaction properties of the studied process parameters on the
1 responses were explicated using a regression equation for the statistical model and RSM contour plots.
2
3
4 3. Results and discussion
5
6 3.1. Clay material, FAU and Fe-FAU characterization
7
8 The X-ray diffraction patterns (XRD) of the raw clay, FAU and Fe-FAU are reported in Fig.1.
9 According to XRD analysis of the raw clay, the obtained diffractogram show that clay material is
10
11 principally composed of Illite (I) by the presence of peaks at 9.97, 4.97 and 2.57 Å. The peaks at
12 7.15 and 3.57 Å can be attributed to Kaolinite phase (K). The existence e of Quartz (Q) in the clay
13
14 structure is marked by the peaks at 3.34 and 4.25 Å. The above-mentioned data suggest that this clay
15 is an instratified illite–kaolinite clay mineral. According to the diffractogram of prepared zeolite
16
material, the typical diffraction peaks agree well with the characteristic peaks of Faujasite zeolite by
17
18 comparing the d-values of the products with information got from Simulated XRD Powder Patterns
19 collection for Zeolites. No additional peaks are observed, indicating the crystallization of only one
20
21 highly pure phase of a pure Faujasite type zeolite. In fact, illitic, kaolinite and quartz phases of the
22 raw clay material disappear during the fusion step that can easily dissolve those components for
23
24 FAU formation by thermal treatment later. The XRD pattern of Fe-FAU catalyst exhibit similar
25 peaks to those of the as-prepared zeolite FAU and all angles for the treated material are almost
26
27 similar to those of the pure FAU zeolite. So, no new crystalline pattern appears; except for a small
28 change in the intensity of the peaks. Some authors attribute this phenomenon to the effect of the
29
charge during the incorporation of ferric ions on the zeolite structure [24-25].
30
31
32 These facts confirmed that the framework and crystallinity of zeolite were not destroyed during the
33
34 preparation of catalyst which show that all XRD angles are almost similar to those of the pure FAU
35 material. In fact, the exchange-calcination method has not affected the d-spacing on the structure of
36
37 Faujasite to any considerable degree. This demonstrates that preparation of catalyst was efficient
38 only on the surface of the zeolite and doesn’t change the crystal structure of the zeolite material.
39
40
Thus, the zeolite structure has not been destroyed during the catalyst preparation. It can be
41 concluded that the insertion of metal ions into the Faujasite structure affects the crystallinity but not
42 the zeolite structure [26].
43
44 The morphologies of studied materials are presented in Fig. 2. As shown in Fig. 2a, the raw clay
45 sample presents a platy morphology with particles in the form of sheets, typical of clay minerals.
46
47 Fig. 2b presents SEM micrograph of prepared zeolite. This figure showed zeolitic crystals, which
48 exhibit well-defined hexagonal morphology. The EDX spectra of raw clay and the FAU zeolite, are
49
50 shown in Fig. 2. EDX analysis, as, shows the cumulative percentage weight of significant
51 aluminosilicate elements as 95% (Si =50.4%, O =18.7%, Al = 25.9%) while other elements gave a
52
percentage weight of 5%, which confirms the presence of zeolite compounds in comparison with
53
54 raw clay materials which present a fewer percentage of aluminosilicate.
55 The nitrogen adsorption-desorption isotherms of raw clay, FAU zeolite and Fe coated FAU catalyst
56
57 are shown in Fig. 3. The isotherms of clay material exhibit an H4 hysteresis loop, relative to
58 particles in the form of a sheet, representative of clay minerals [23]. Based on Brunauer’s
59
60 classification, the isotherms of the both materials: FAU zeolite and Fe-FAU catalyst show a typical
61
6
62
63
64
65
IUPAC type II with a hysteresis loop in the relative pressure range of 0.45–0.85 which strongly
1 suggests mesoporosity in the samples and signifies the filling of uniform slit-shaped intercrystal
2
3 mesopores, arising from the filling of zeolite nanocrystals [27]. The saturation is often not reached,
4 which suggests a degree of macroporosity in the materials, confirmed also by a strong uptake in
5
6 nitrogen at high relative pressures (0.85–1). The isotherms show also a steep nitrogen uptake at low
7 relative pressures (P/P0 = 0.02), which corresponding to the filling of micropores. The Porous
8
9 structure parameters of the studied materials were given in Table 1. The table show that The BET
10 surface area and pore volume of the raw clay sample, measured as 50 m2/g and 0.011 cm3/g with
11
essentially a micropore structure. The FAU zeolite shows an improve of specific surface area and
12
13 total pore volume to reach 360 m2/g and 0.33 cm3/g respectively as compared to raw clay which
14 explain the importance of synthesis of zeolite instead using natural clay material. The nitrogen
15
16 adsorption–desorption isotherms of prepared catalyst presented in Fig.3 showed that the catalyst has
17 practically the same structure of untreated zeolite with a trimodal porosity (micro, meso and
18
19 macropores). Nevertheless, after the exchange-calcination method the nitrogen uptake at low
20 relative pressures (P/P0) decreased noticeably compared to the pure zeolite, suggesting a reduction in
21
22 the available microporosity. Similar trend was reported for Cu and Ag exchanged zeolite X [28].
23 The variation of textural properties after preparation of Fe-FAU catalyst is presented in Table 1. We
24
can notice a decrease of the total specific surface area and a drop of the total pore volume in
25
26 comparison with the FAU pure zeolite. In the other hand, the examination of pore distribution
27 curves shows bimodal distribution in the microporous and mesoporous regions with a decrease of
28
29 the height of the micropores peaks in comparison with the raw zeolite. Therefore, the proprieties like
30 surface areas of catalysts and pore volume seem to be undoubtedly affected by the exchange
31
32 method. In the other hand, after the exchange process the surface of FAU zeolite had been covered
33 by the iron, thus the apparent color of raw zeolite changed from white to pale yellow, as shown in
34
35
Fig. S1 the color change is the first proof of the fixation of metal ions at the surface of the zeolite
36 support material [29]. Indeed, the change of the color might be due to the oxide particles fixation
37 over the surface of prepared material and confirm the decrease of crystallinity observed with the
38
39 XRD technique. Fig. 5 show also the change of color from green for the clay to white for the zeolite
40 during zeolite preparation.
41
42
43 In order to investigate the chemical surface characteristics of the raw clay, pure zeolite and the
44 prepared catalyst, FTIR analyses were carried out in the range 4000–400 cm−1 and the spectra of
45
46 different studied samples are provided in Fig. 4. Infrared spectroscopy allowed observing the
47 changes that occur in the structure of raw materials after zeolite and catalyst preparation. The FTIR
48
49
spectra of the raw clay show a notable band at around 3618 cm–1 which was identified in earlier
50 studies as representative of illite. Authors assigned this observed band to the deformation of Mg–
51 Al– OH and the vibration in the plane of Al–Si–O liaisons in illitic phase. The absorption bands at
52
53 3440 cm–1 and 1626 cm–1 has been attributed to OH frequencies of the water molecule adsorbed on
54 the clay surface [30]. The Si–O–Si vibrations towards 800 and 1030 cm–1 are characteristics of
55
56 Quartz phase [31]. In comparison of FTIR spectra of raw clay and FAU zeolite, it was found that
57 characteristic bands of Faujasite zeolite appears with completely destruction of clay structure. In
58
59 fact, the IR spectra of the prepared material evidently shows the characteristic bands of Faujasite
60 type zeolite. The characteristic strong band of zeolite FAU at 976 cm−1 can be attributed to
61
7
62
63
64
65
elongating vibration of Si–O and Al–O groups [32]. The peaks observed at 745 cm−1 and 670 cm−1
1
can be assigned to vibration modes of the zeolite framework [33]. The peak located at 560 cm−1 is
2
3 related to the bending vibration of Si-O or Al-O groups in the zeolite structure [34]. The bands
4 assigned to the H2O deformation mode and to the OH elongating of water molecules present in the
5
6 zeolite channel lie in the range of 3500-1600 cm-1. In the other hand, FTIR spectra of the prepared
7 catalyst is practically similar to this of pure FAU zeolite. While, it shows characteristic bands of the
8
9 Faujasite phase in the range of 1000-400cm-1 with sharp decrease in the intensities of the absorption
10 bands compared to the FAU syn spectrum. It can be concluded that exchange-calcination method
11
can lead little structural modification of the zeolite.
12
13
14 3.2. Catalytic performance of the prepared materials
15
In order to determine the contributions of the incorporated iron in the structure of the raw
16
17 FAU and to elucidate the impact of different mechanisms participating in the degradation of MO by
18 heterogeneous Photo-Fenton process, comparative experiments were undertaken by testing the effect
19
20 of catalyst dose, hydrogen peroxide concentration and MO initial concentration.
21 3.2.1. Effect of the catalyst dose
22
23 The effects of the catalysts dose are shown in Fig. 5 and S2, in which the reactions were performed
24 for 20, 50, 70 and 100 mg of FAU and Fe-FAU respectively with concentration of MO equal to 20
25
mg/L, a constant H2O2 dosage (150 μL) and under UVC irradiation. From Fig. 5, we can notice that
26
27 before the H2O2 addition, which is the step illustrating the adsorption stage, the elimination of MO
28 by the prepared materials was commonly low. For the FAU, a small decrease of MO (not more than
29
30 8% for all the studied cases) was detected after 5 min that describes the adsorption equilibrium
31 followed by a significant decrease related to the oxidation phenomena. In addition, the sorption step
32
33 is the consequence of the formation of attractive forces between the positive charged adsorbent
34 surface and the anionic molecules of the pigment in solution. After the H2O2 addition, it was
35
36 observed that with an increase in the amounts of FAU and Fe-FAU loading from 20 to 100 mg, the
37 degradation efficiency of MO increased and a complete color elimination was reached within 20
38
min for each material. However, degradation was more efficient than color removal during the
39
40 catalytic process. In fact, when the amount of raw FAU was increased from 20 mg to 50 mg/50 mL
41 of solution, TOC elimination also increased from 40% to t reach 60%. Whereas Fe-FAU catalyst
42
43 exhibit higher TOC removal than the raw FAU. In addition, it was observed that with an increase in
44 the amounts of Fe-FAU loading from 20 to 50 mg, the degradation of MO increased from 62% up to
45
46 94% during 60 min of oxidation. The high catalytic activity of Fe-FAU catalyst in comparison with
47 the raw FAU can be attributed to the well dispersed Fe2+ species (major species in the surface).
48
49 These species are responsible for effective generation of very reactive hydroxyl radicals (HO ·) from
50 H2O2. Also, higher dosage will give higher overall amounts of metals ions in the zeolite framework
51 as well as higher concentration of metals ions in the bulk, which in turn increase the number of
52
53 hydroxyl radicals significantly. Likewise, higher dosage of catalyst increased the power of the
54 system for MO degradation [35]. However, further increases in the dose of catalyst more than 50 mg
55
56 did not result in catalytic performance improvement. Similar trend for the influence of catalyst
57 concentration on MO oxidation was reported by Dükkancı et al [36] for degradation of Rhodamine
58
59 6G using CuFeZSM-5 zeolite. The best catalyst dose for the degradation of MO chosen to be used in
60 this study was 50 mg.
61
8
62
63
64
65
3.2.2. Effect of H2O2 Concentration
1
2 Before studying the effect of the hydrogen peroxide concentration on the photocatalytic activity, it’s
3
4 mandatory to understand the reactions mechanisms that occurs during Photo-Fenton reaction in
5 presence of Fe-FAU catalyst. A simple mechanism is proposed in Eqs. (2–4) where the symbol ʊ
6
7 represents the surface of Fe-FAU catalyst.
8
9
10  Fe 2  H 2O2   Fe3  HO   OH  (2)
11
12
13 Fe  FAU  MO  HO  Fe  FAU  int ermidiates (3)
14
15
16 Fe  FAU  int ermidiates  HO   CO2  H 2O  inorganicsalts (4)
17
18
19 The process of MO degradation was initiated by the generation of OH radicals from H 2O2 stimulated
20 essentially by the presence of Fe2+ in the reaction medium, while Fe2+ was oxidized into Fe3+ on
21
22 FAU zeolite (Fenton’s reaction). Then, Hydroxyl radicals might attack the MO adsorbed into the
23 surface of Fe-FAU catalyst, resulting in the discoloration of reaction intermediates. Finally, these
24
25
obtained intermediate compounds are transformed into CO2, H2O and inorganic salts [37]. Based on
26 experimental approaches and previous studies, the photocatalytic reactivity of the Fe-FAU catalyst
27 is summarized in Fig. S3.
28
29
30 So, in order to determine the optimum amount of hydrogen peroxide needed to MO degradation
31 under the selected conditions, the oxidation kinetics of MO were investigated for different initial
32
33 H2O2 concentration values (in the range 0–300 µL) and the results are shown in Fig. 6. As observed,
34 the rate of MO degradation was improved by increasing the initial H2O2 concentration and the higher
35
removal efficiency was reached with 150µL added in solution. This is due to the oxidative puissance
36
37 of photo-Fenton process, which was enhanced with increasing OH● radicals in solution by the
38 decomposition of H2O2. Though, doses even greater than 150µL of H2O2 loading, at the same
39
40 operation conditions, does not result in important improvement in the rate of MO degradation. This
41 is justified by the fact that part of H2O2 cannot be used by the catalyst effectively (H2O2
42
43 decomposition without generating OH● radicals or scavenging effect) thus a fixed amount of catalyst
44 puts an upper limit on the rate of generation of OH● radicals. Since the H2O2 is inactive by itself, the
45
46 presence of excess oxidant has no influence [19]. It has been reported in the literature that the
47 oxidation kinetics of dyes usually decreases in the presence of an excessive amount of hydrogen
48
49 peroxide. Also, this phenomenon could be attributed to the hydroperoxyl radicals ( HO2 ) generation
50
51 in the presence of a local excess of H2O2 as shown in the below given equation. Such reaction
52
53 reduces the probability of attack of organic molecules by hydroxyl radicals, and doesn’t affect the
54
55 degradation rate. Although other radicals ( HO2 ) are produced, its oxidation potential is much lower
56
57 
58 than that of OH [38]. So, the hydroperoxyl radicals are less reactive and do not lead to the
59
60
61
9
62
63
64
65

1 oxidative degradation of organic compounds, which take place just by reaction with OH .
2
3 Therefore, the amount of hydroxyl radicals that can be generated in the reaction medium is directly
4 related to the initial H2O2 concentration.
5
6
7 H 2O2  OH   HO2  H 2O (5)
8
9
10 In order to more explain the effect of the initial H2O2 concentration on the oxidation process of MO
11
12 molecules, the consumption of hydrogen peroxide was studied for different values of its initial
13 concentration (Fig.9). For a [H2O2]0 = 50µL, the total hydrogen peroxide present in the reaction
14
15 medium was consumed after 20 min of reaction, and the dye degradation is not complete (42 % of
16 degradation rate). For the dose [H2O2]0 = 100 µL, the content of H2O2 has totally reacted after 25
17
18 min. Then, for a H2O2 dosage of 150 mg/L, better degradation rate was obtained and the whole H2O2
19 was consumed in 45 min. for H2O2 doses greater than 150 µL, the H2O2 concentration is restraining
20
and there is no oxidative species in the reaction medium.
21
22
23 It can be concluded that an initial H2O2 concentration higher than 150µL corresponds to an
24 unprofitable consumption of H2O2. So, the dose tested of 150µL was the value chosen to carry out
25
26 the following runs.
27
28
3.2.3. Influence of the initial concentration of MO
29
30 Photo-Fenton process also depends on initial concentration of organic compound present in the
31 reaction solution. So, it is also of practical interest to investigate the effect of the initial pollutant
32
33 concentration, as it is of importance in any process of wastewater treatment. To investigate the
34 optimum concentration of MO, experiments were executed at different MO initial concentrations
35
36 ranging from 20 to 10 0 mg/L. The found results are shown in Fig. 7. It is clearly observed that the
37 degradation of MO and TOC is favorable at low concentrations. Obviously, the initial MO
38
39 concentration showed an undesirable influence on its degradation, i.e., the higher the initial dye
40 concentration, the higher was the time required to degrade it completely. So, complete color removal
41
was obtained after 19 min of reaction time for an initial concentration of 20 mg/ L, 27 min for 50
42
43 mg/L, 45 min for 70 mg/L and 60 min for 100mg/L. As well as, TOC monitoring exhibited that the
44 best mineralization performance is obtained for 20 mg/L of MO. This may be due the non-
45
46 availability of sufficient hydroxyl radical [39]. In fact, in the heterogeneous photo-Fenton process,
47 the hydroxyl radicals generated at the active sites on the surface of Fe-FAU reacts with adsorbed
48
49 MO molecules. Definitely, for higher MO concentration, the number of active sites accessible
50 decreases because of competitive adsorption of the MO molecules on the catalyst surface [40].
51
52 Moreover, the intermediate compounds formed during MO degradation compete with the pigment
53 molecules for the available iron active sites. Thus, the lower oxidation efficiency for higher dye
54 concentration can be attributed to more pollutant and intermediates molecules for active sites on Fe-
55
56 FAU surface, which had to be degraded by the constant amount of the reactive radicals [41]. Then,
57 the solution becomes extremely colored, and smaller amount photons [42-43] reach the surface of
58
59 the Fe-FAU, reducing the photocatalytic activity. Indeed, the concentration of organic compound
60 plays a very important role in reactions in accordance with the collision concept of chemical
61
10
62
63
64
65
reactions. The same inhibiting effect of the initial dye concentration on the oxidation performance
1 was observed by earlier study [44].
2
3
4 3.2.4. Effect of UV wavelength
5
6 In order to optimize the experimental parameters influencing the degradation of MO dye, it is
7
8 considerable to plan the mode of irradiation of the oxidant. In fact, most of the work reported in the
9 Photo-Fenton reaction literature has used UV-C irradiation (254 nm) because of its high absorption
10
11 energy, resulting in efficient degradation of organic compounds. However, UV light covers a wide
12 range of wavelengths, UV-B (280-320 nm) and UV-A (320-400 nm), and these next to UV-C are
13
14 also considered effective. Total irradiation times were 150, 250 and 300 min for tests conducted
15 under UVA, UVB and UVC respectively. Samples were taken in different steps depending on the
16
duration of the experiment. The results of MO degradation in the presence of Fe -FAU obtained by
17
18 varying the irradiation type and keeping constant the other parameters are shown in Fig. S4.
19 Looking at the degradation time a significantly faster degradation is achieved with UV-C irradiation.
20
21 As well as, the increase in the degradation rate was about 50% and 30 % for the UV-C in
22 comparison with UV-B and UV-A respectively. In fact, the shortest reaction time, obtained with the
23
24 UV-C irradiation with almost complete degradation (96%) was of 20 min. As shown in Fig. S4, a
25 negative impact in terms of abatement of the MO concentration was observed in the second and
26
27 third tests using UV-B and UV-A irradiations. Thus, total reaction time increases to reach 2 h and 4
28 h for UVB and UVA wavelengths. This contribution prevents the system to be economically
29
interesting regarding the high energy consumption. So, the key advantage of using UV-C irradiation
30
31 type is the improvement of catalytic performance by minimizing the reaction time and increasing the
32 degradation rate probably due to the higher absorbance of the Fe-FAU catalysts at 254 nm and
33
34 enhanced photocatalytic effect [45]. This trend is confirmed by earlier study [46].
35
36 3.3. Catalyst reusability and leaching test
37
38
39 Stability is an essential property for effective catalysts. In fact, one of the key advantages of
40 processing with heterogeneous more than homogeneous catalysts in the treatment of organic dyes is
41
42 the possibility for reusability of the used catalyst for several runs. To study the possible leaching of
43 metallic ions into the solution and the stability of the catalysts, 50 mg of Fe-Fau was used for
44
45
repeated experiments with initial concentration of MO kept constant (50 mg. L−1) for 2 h irradiation
46 under UVC. After each experiment, the catalyst was removed by filtration and then washed by
47 distilled water for several times and dried at 70 °C for 12 h. The final solutions were analyzed by
48
49 AAS. The obtained results for the Fe2+ presented in Fig. 8 show undetectable ions in the final
50 solutions. In fact, the leached amounts were not enough to considerably have an effect on the
51
52 catalytic activity in five successive runs. It was shown that the concentration of iron ions in the
53 solution after reuse of the catalyst for five repeated cycles were 0.056, 0.0770, 0.0810, 0.0964 and
54
55 0.0766 mg. L−1. The low amounts of metals loss in the solution which is below the permissible level
56 as reported in other studies [47 et 13] prove the good stability of the catalyst and showed that active
57
sites remained on the catalyst. This indicates that the photocatalytic activity is mostly owing to
58
59 metallic ions present on the catalyst surface rather than the trace amounts of leached ones. In the
60 other hand, we can notice that there was no obvious variation of oxidizing capacity in UV/H2O2/Fe-
61
11
62
63
64
65
FAU process. The degradation rate of MO within 60 min was over 81% after five recycles. All these
1 results all confirmed possibility of reuse and the reasonable stability of prepared catalyst for
2
3 photocatalytic oxidation of organic compounds.
4
5 3.4. Experimental design and statistical analysis
6
7
8 3.4.1. statistical analysis
9
10 Optimization of the oxidation variables via statistical approaches is important for saving time and
11
12 resources. Furthermore, the statistical study gives good information about the oxidation parameters
13 and reduces the experimental errors. Here, the BBD (Box–Behnken experimental design) statistical
14
15 approach coupled with RSM was employed to improve the oxidation of MO on Fe-FAU. Then,
16 Parameters influencing the MO oxidation process were optimized via the BBD method which is a
17
18
second-order design based on three-level factorial designs, 13 experiments were carried out and the
19 complete BBD matrix for the three important factors [zeolite dose (A), H2O2 dosage (B), and MO
20 initial concentration (C)] and response [degradation rate (%) for both the coded and real values] are
21
22 presented in Table 2. And then, the degradation percentages of TOC were calculated as the response
23 variable, as given in Table 2. ANOVA for the regressions was executed to evaluate the significance
24
25 under a confidence interval of 95% (Table 3).
26
27 Statistical quadratic models involving the interaction and polynomial terms were established for all
28
29 the given response variables. The cause-effect can be denoted as presented in Equation (6).
30
31 n n n 1 n
32
33
Y  b0   bi X i   bii X i2    bij X i X j (6)
i 1 i 1 i 1 j i 1
34
35
36
37
where Y is the predicted response (degradation of TOC (%) of MO), b is a constant while, bi, bii, and
38 bij refers to the linear, quadratic and interaction coefficients, respectively. X i and Xj are Xi2 are the
39 levels of the process variables.
40
41 The empirical relationship between oxidation parameters and MO degradation (%) rate generated
42
43 from the BBD was found to be as it is shown in Eq.7 :
44
45 MO degradation(%)= 93+10.87A+6.25B+8.38C-4.5AB-6.24 AC-3.00 BC-3.63 A²-11.87 B²-6.13 C²
46 (7)
47
48 Where the positive coefficients of factors show their contribution in improving the degradation rate
49
50 (%) of MO and vice versa.
51
52 Then, the significance of each variable was evaluated after examining the p-value and F value. In
53
54 fact, the corresponding variables being more significant is shown by a higher F-value also called F-
55 statistic and lower P value also called significance probability. Statistical models corresponding to
56
the responses for all the experiences were significant, with all sequential p values < 0.05 and few of
57
58 fit p value > 0.05. The insignificant lack of fit values shows that the model is valid for the given
59 experimental parameters and the regression model was found to be statistically significant at a 95%
60
61
12
62
63
64
65
confidence level. The model F values (86.31) further confirms the significance of the quadratic
1 model to suitably describe the oxidation of MO and there is small chance that noise causes this
2
3 large F value to occur. Moreover, the linear (A, B, and C) and quadratic (AB, AC, A2, B2, and C2)
4 terms of the quadratic equation (Eq. 9) were significant (p value<0.05) in the 95% confidence level.
5
6 P values greater than 0.1000 indicate the model terms are not significant, and therefore BC
7 interaction terms are not significant.
8
9
10 Model validation further examined by means of regression (R2) coefficient. The predicted and
11 experimental values were plotted and associated for all the experiments in table 4. The Predicted R 2
12
13
(0.9499) is in judicious agreement with the Adjusted R2 (0.9785) because their difference is less
14 than 0.2. Moreover, values close to unit of these coefficients indicates that the model conforms with
15 acceptance standards. Seeing all the other regression variables, the quadratic model can fit the
16
17 response data for MO degradation and is adequate to explain interaction between factors.
18
19 Fig. S5 presents the diagnostic curves that give additional information about the validation of model.
20
21 The comparison between the plot of predicted and actual degradation (%) of MO (Fig. S5a) shows a
22 good fit to the linear regression. That’s the model reproduces fairly the experimental series . The plot
23
24 of normal % probability with externally studentized residuals is shown in Fig. S5b. The studentized
25 residuals ae used to detect any outliers. The graphs show almost linear arrangement of points which
26
27 indicates the adequacy of the model and the normal distribution of errors crosswise the runs . Fig.
28 S5c demonstrates a plot of externally studentized residual against the run number is presented in
29
Fig. S5c. The points falling outdoor the red limits are considered as outliers. Hence, all the presented
30
31 points in Fig. S5 are inside the range of the model assumptions and there are no outliers detected.
32 Fig. S5d show the perturbation plot which present the comparative influence of all the adsorption
33
34 factors at a specific point in the design space. In addition, the perturbation curve helps to detect the
35 factor that most affects the degradation (%) of MO dye. The steepest slope for factor A (slope =
36
37 10.87) than B (6.25B) and C (8.38) shows that MO degradation (%) is depending on zeolite dose.
38
39 3.4.2. Effect of experimental parameters on MO degradation
40
41
42 The 3D surface plot of the oxidation process lets to visualize the interaction between factors for the
43 degradation (%) of MO. Fig. 9 shows the 3D surface plots of MO removal (%) with studying the
44
45
mutual effect of catalyst dose (mg), MO initial concentration (mg/L) and H2O2 dosage (µL) under
46 UVC irradiation. Regarding the combined effect of catalyst dose and H2O2 initial concentration, at
47 constant MO initial concentration, as shown in Fig. 9a. the MO removal (%) increased with an
48
49 increase in both catalyst dose and H2O2 dosage until these later rich about 200µL; from there, it
50 decreased slightly. This result confirms the fact that the presence of an excess of hydrogen peroxide
51
52 in the reaction medium has a negative effect on dye oxidation of MO. Fig. 9b shows, the mutual
53 effect of MO initial concentration and zeolite dose. The obtained 3D plot surface shows a positive
54
55 effect of zeolite dose in the removal (%) of MO while the maximum degradation efficiency is
56 obtained for an MO initial concentration raging between 40 and 70 mg/L beyond that removal rate
57
slightly decreases. Fig. 9c illustrate the mutual effect of MO initial concentration (mg/L) and H2O2
58
59 dosage. It confirms the above conclusions that the maximum degradation efficiency is obtained for a
60 H2O2 dosage and MO concentration ranging between 150-200 µL and 40-70mg/L respectively.
61
13
62
63
64
65
Usually, statistical analysis aims to reach some goals for example augment, diminish, the range and
1 aim for optimizing the responses and factors [48, 49]. Here, statistical optimization was anticipate to
2
3 improve removal (%) of MO and factor values in the range. Fig. S6 presents the maximum
4 desirability ramp for the optimum combination of zeolite dose (20–100 mg), H2O2 dosage (50–300
5
6 µL) and MO concentrations (20–100 mg/L) to obtain maximum degradation efficiency. Hence, the
7 optimum oxidation efficiency of MO was found to be 95.294% at the best combination of 60 mg/L
8
9 of MO, 69.1403 mg Fe-FAU, and 175 µL of H2O2.
10
11 Fig. S7 displays the individual effect of each influencing factor. It is clear that catalyst dosage have
12
13
a positive impact on the MO degradation. This result is an accordance with results presented in
14 table 3. As it shown, the catalyst dose have the maximum effective degree of the individual factor on
15 MO degradation (F = 92.2 and p =< 0.0001). Thus, in the case of catalyst dose, a clear rise in MO
16
17 removal was obtained when the Fe-FAU dosage pass from 20 to 100 mg. Whereas, from Fig. S7, we
18 can notice that H2O2 dosage and MO initial concentration have a limited positive impact on MO
19
20 degradation in limited range around 60 mg/L and 175 µL for respectively the MO and H2O2
21 concentrations. Beyond those limits the MO removal (%) decreased.
22
23
24 4. Conclusion
25 In the present study, highly crystalline and pure phase zeolite FAU with a hexagonal morphology
26
27 was prepared from Tunisian illitic clay via alkali-hydrothermal treatment. Then, Fe-FAU, a
28 heterogeneous Photo-Fenton catalyst was synthesized by a simple impregnation-calcination method.
29
The characterization results investigated using X-ray diffraction (XRD), Fourier transform infrared
30
31 spectroscopy (FTIR) and the specific surface areas of solids showed that the zeolite structure was
32 not affected by the catalyst preparation method. The oxidation efficiency of Fe-FAU catalyst
33
34 towards MO dye in aqueous solution via Photo Fenton process was studied with varying, the
35 catalyst dose zeolite (mg), H2O2 dosage (µL) and initial concentration of MO (mg/L). Complete
36
37 color removal and 94 % TOC degradation was achieved within 20 min of reaction time with 150 μL
38 of H2O2 and 50 mg of the Fe-FAU syn. Statistical optimization of oxidation factors such as zeolite
39
40
dose, H2O2 dosage (µL) and initial concentration of MO at fixed temperature (25°C) was found to be
41 a well-fitted BBD model. The low probability (p values) and high regression(R2>0.94) demonstrated
42 the validity of the quadratic model to predict the degradation rate (%) of MO. The oxidation
43
44 efficiency of MO was found to be 95.294% at the optimum combination of 60 mg/L of MO,
45 69.1403 mg Fe-FAU, and 175 µL of H2O2. Thus, it is evident from the above study that the zeolite
46
47 based catalyst can successfully be employed for the oxidative removal of dyes from wastewaters.
48
49
50
51 Acknowledgements The authors are grateful to PHC Maghreb n°27959PD for financial support
52
and to the scientific partnership between ENIS, Sfax (Tunisia) and the University of Littoral Côte
53
54 d’Opale, Dunkerque (France) for the financial support of Olfa Ouled Ltaief’s PhD thesis.
55
56
57
58
59
60
61
14
62
63
64
65
References
1
2 [1] Aishah AJ, Sugeng T, Adam SH, Rahim ND, Aziz MAA, Hairom NHH, Razali NAM, Abidin
3 MAZ., Mohamadiah MKA (2010) Adsorption of methyl orange from aqueous solution onto
4
5 calcined volcanic mud. J Hazard Mater 181: 755–762.
6
7 [2] Phillips D (1996) Environmentally friendly, productive and reliable: priorities for cotton dyes
8 and processes. J. Soc. Dye. Colour. 112: 183–186.
9
10 [3] Brown D, Hamberger B (1987) Degradation of dye stuffs. Part III. Investigation of their ultimate
11
12 degradability. Chemosphere 16: 1539–1553.
13
14 [4] Wang CC, Lee CK, Lyu MD, Juang LC (2008) Photocatalytic degradation of C.I. basicviolet 10
15 using TiO2 catalysts supported by Y zeolite: an investigation of the effects of operational
16
parameters. Dyes Pigm 76: 817–824.
17
18
[5] Malik PK (2004) Dye removal from wastewater using activated carbon developed from sawdust:
19
20 adsorption equilibrium and kinetics. J Hazard Mater B113: 81–88.
21
22 [6] Lu XJ, Yang B, Chen JH, Sui R (2009) Treatment of wastewater containing azo dye reactive
23 brilliant Red X-3B using sequential ozonation and up flow biological aerated filter process. J Hazard
24
25 Mater 161: 234–241.
26
27 [7] Kondru AK., Kumar P, Chand Sh (2009) Catalytic wet peroxide oxidation of azo dye (Congo
28 red) using modified Y zeolite as catalyst. J Hazard Mater 166: 342–347.
29
30 [8] Neamtu M, Zaharia C, Catrinescu C, Yediler A, Macoveanu M, Kettrup A (2004) Fe-exchanged
31
32 Y zeolite as catalyst for wet peroxide oxidation of reactive azo dye Procion Marine H-EXL. Appl
33 Catal B 48: 287–294.
34
35 [9] Duarte F, Madeira LM, (2010) Fenton- and Photo-Fenton-Like Degradation of a Textile Dye by
36 Heterogeneous Processes with Fe/ZSM-5 Zeolite. Sep Purif Technol 45: 1512–1520.
37
38 [10] Amat AM, Arques A, Bossmann SH, Braun AM, Miranda MA, Vercher RF (2005) Synthesis,
39
40 loading control and preliminary tests of 2, 4, 6-triphenylpyrylium supported onto Y-zeolite as solar
41 photocatalyst. Catal Today 101: 383–388.
42
43 [11] Ramirez JH, Maldonado-Hódar FJ, Pérez-Cadenas AF, Moreno-Castilla C, Costa CA, Madeira
44
45 LM (2007) Fenton-like oxidation of Orange II solutions using heterogeneous catalysts based on
46 saponite clay. Appl Catal B 75: 312–323.
47
48 [12] Shukla PR, Wang Sh, Sun H, Ming Ang H, Tadé M (2010) Activated carbon supported cobalt
49
50
catalysts for advanced oxidation of organic contaminants in aqueous solution. Appl Catal B 100:
51 529–53.
52
53 [13] Panda N, Sahoo H, Mohapatra S (2011) Decolorization of Methyl Orange using Fenton-like
54 mesoporous Fe2O3–SiO2 composite. J Hazard Mater 185: 359–365
55
56 [14] Chong MN, Vimonses V, Lei Sh., Jin B, Chowd Ch, Saint Ch (2009) Synthesis and
57
58 characterization of novel titania impregnated kaolinite nano-photocatalyst.
59 Microporous Mesoporous Mater 117: 233–242.
60
61
15
62
63
64
65
[15] Sum OSN, Feng J, Hu X, Yue PL (2004) Pillared laponite clay-based Fe nanocomposites as
1 heterogeneous catalysts for photo-Fenton degradation of acid black 1. Chem Eng Sci 59: 5269–
2
3 5275.
4
5 [16] Carriazo J, Guélou E, Barrault J, Tatibouët JM, Molina R, Moreno S (2005) Synthesis of
6 pillared clays containing Al, Al-Fe or Al-Ce-Fe from a bentonite: Characterization and catalytic
7
8 activity. Catal Today 107–108: 126–132.
9
10 [17] Barrault J, Abdellaoui M, Bouchoule C, Mejeste A, Tatibouët JM, Louloudi A, Papayannakos
11 N, Gangas NH (2000) Catalytic wet peroxide oxidation over mixed (Al–Fe) pillared clays. Appl
12
13
Catal B 27: L225–L230.
14
15
[18] Hailu SL, Nair BU, Redi-Abshiro M., Aravindhan R., Diaz I, Tessema M, Devi LG, Kumar
16 SG, Reddy KM, Munikrishnappa C (2009) Photo degradation of methyl orange an azo dye by
17 advanced fenton process using zero valent metallic iron: influence of various reaction parameters
18
19 and its degradation mechanism. J Hazard Mater 164: 459–467.
20
21 [19] Shukla PR, Wang Sh., Singh K, Anga H, Tadé MO (2010) Cobalt exchanged zeolites for
22 heterogeneous catalytic oxidation of phenol in the presence of peroxymonosulphate. Appl Catal B
23
24 99: 163–169.
25
26 [20] Hassan H, Hameed BH (2011) Oxidative decolorization of Acid Red 1 solutions by Fe– zeolite
27 Y type catalyst. Desalination 276: 45–52.
28
29 [21] Murcia-López S, Bacariza MC, Villa K, Lopes JM, Henriques C, Morante JR, Andreu T (2017)
30
Controlled Photocatalytic Oxidation of Methane to Methanol through Surface Modification of Beta
31
32 Zeolites. ACS Catal 7 : 2878–2885
33
34 [22] Zainal Abidin A, Abu Bakar NH.H, Ng E.P, Tan WL (2017) Rapid Degradation of Methyl
35 Orange by Ag Doped Zeolite X in the Presence of Borohydride. J Taibah Univ Sci 11: 1070–1079.
36
37 [23] Ltaief OO, Siffert S, Poupin C, Fourmentin S, Benzina M (2015) Optimal Synthesis of
38
39 Faujasite-Type Zeolites with a Hierarchical Porosity from Natural Clay. Eur J Inorg Chem 28: 4658-
40 4665.
41
42 [24] Ruda TA, Dutta PK (2005) Fenton chemistry of FeIII – exchanged zeolitic minerals treated
43
44 with antioxidants. Environ Sci Technol 39: 6147–6152.
45
46 [25] Fach E, Waldman WJ, Williams M, Long J, Meister RK, Dutta PK (2002) Analysis of the
47 biological and chemical reactivity of zeolite-based aluminosilicate fibers and particulates.
48
Environ Health Perspect 110: 1087–1096.
49
50 [26] Salavati-Niasari M, Salimi Z, Bazarganipour M, Davar F (2009) Synthesis characterization and
51
52 catalytic oxidation of cyclohexane using a novel host (zeoliteY)/guest (binuclear transition metal
53 complexes) nanocomposite materials. Inorganica Chim Acta 362: 3715-3724.
54
55 [27] Huang Y, Wang K, Dong D, Li D, Hill MR, Hill AJ, Wang H (2010) Synthesis of hierarchical
56
57 porous zeolite NaY particles with controllable particle sizes. Microporous Mesoporous Mater. 127:
58 167–175.
59
60
61
16
62
63
64
65
[28] Benaliouche F, Boucheffa Y, Ayrault P, Mignard S, Magnoux P (2008) NH3-TPD and FTIR
1 spectroscopy of pyridine adsorption studies for characterization of Ag- and Cu-exchange zeolite X
2
3 zeolites. Microporous Mesoporous Mater. 111: 80–88.
4
5 [29] Nahar MSh, Hasegawa K, Kagaya S, Kuroda S (2009) Adsorption and aggregation of Fe (III)–
6 hydroxy complexes during the photodegradation of phenol using the iron-added-TiO2 combined
7
8 system. J Hazard Mater 162: 351-355.
9
10 [30] Perraki Th, Orfanoudaki A, (2004) Mineralogical study of zeolites from Pentalofos area,
11 Thrace, Greece. Appl. Clay Sci. 25: 9–16.
12
13 [31] Jiang T, Li G, Qiu G, Fan X, Huang Z (2008) Thermal activation and alkali dissolution of
14
15
silicon from illite. Appl. Clay Sci. 40: 81–89.
16
[32] Murali RS, Surya Murali R, Ismail AF (2014) Mixed matrix membranes of Pebax-1657 loaded
17
18 with 4A zeolite for gaseous separations. Sep Purif Technol 129:1–8.
19
20 [33] Zavareh S, Farrokhzad Z, Darvishi F (2018) Modification of zeolite 4A for use as an adsorbent
21 for glyphosate and as an antibacterial agent for water. Ecotoxicol Environ Saf 155:1–8.
22
23 [34] Zou W, Bai H, Zhao L, Li K, Han R (2011) Characterization and properties of zeolite as
24
25 adsorbent for removal of uranium(VI) from solution in fixed bed column. J Radioanal Nucl Chem
26 288:779–788.
27
28 [35] Lucas MS, Peres JA (2006) Decolorization of azo dye Reactive Black 5 by Fenton and photo
29
30 Fenton oxidation. Dyes Pigm 71: 236–244.
31
32 [36] Dükkancı M, Gündüz G, Yılmaz S, Prihod’ko RV(2010) Heterogeneous Fenton-like
33 degradation of Rhodamine 6G in water using CuFeZSM-5 zeolite catalyst prepared by hydrothermal
34 synthesis. J Hazard Mater 181: 343–350.
35
36 [37] Xu H, Wang Y, Shi T, Zhao H, Tan Q, Zhao B, He X, Qi S (2018) Heterogeneous Fenton-like
37
38 discoloration of methyl orange using Fe3O4/MWCNTs as catalyst: kinetics and Fenton-like
39 mechanism, Front. Mater. Sci. 12: 34–44.
40
41 [38] Neamtu M, Catrinescu C, Kettrup A (2004) Effect of dealumination of iron (III)- exchanged Y
42
43 zeolites on oxidation of Reactive Yellow 84 azo dye in the presence of hydrogen peroxide. Appl
44 Catal B 51: 149–157.
45
46 [39] Devi LG, Kumar SG, Reddy KM, Munikrishnappa C (2009) Photo degradation of methyl
47
48
orange an azo dye by advanced fenton process using zero valent metallic iron: influence of various
49 reaction parameters and its degradation mechanism. J Hazard Mater 164: 459–467.
50
51 [40] Zhan Y, Li H, Chen Y, Copper hydroxyphosphate as catalyst for the wet hydrogen peroxide
52 oxidation of azo dyes. J Hazard Mater. 2010, 180: 481–485.
53
54 [41] Abdel-Aziz R, Ahmed MA, Abdel-Messih MF (2020) A novel UV and visible light driven
55
56 photocatalyst AgIO4/ZnO nanoparticles with highly enhanced photocatalytic performance for
57 removal of rhodamine B and indigo carmine dyes. J. Photochem. Photobiol. A Chem. 389: 112245.
58
59
60
61
17
62
63
64
65
[42] Khadhri N, El Khames Saad M, ben Mosbah M, Moussaoui Y (2019) Batch and continuous
1 column adsorption of indigo carmine onto activated carbon derived from date palm petiole, J.
2
3 Environ. Chem. Eng. 7: 102775.
4
5 [43] Oppong SOB, Anku WW, Opoku F, Shukla SK, Govender PP (2018) Photodegradation of
6 eosin yellow dye in water under simulated solar light irradiation using La–doped ZnO nanostructure
7
8 decorated on graphene oxide as an advanced photocatalyst. Chemistry Select 3: 1180–1188
9
10 [44] Pintar A, Levec J (1992) Catalytic oxidation of organics in aqueous solutions: I. Kinetics of
11 phenol oxidation. J Catal 135: 345–357.
12
13 [45] Vijayabalan A, Selvam K, Velmurugan R, Swaminathan M (2009) Photocatalytic activity of
14
15
surface fluorinated TiO2-P25 in the degradation of Reactive Orange 4. J Hazard Mater. 172: 914–
16 921.
17
18 [46] Wang SB, (2008) A Comparative study of Fenton and Fenton-like reaction kinetics in
19 decolourisation of wastewater. Dyes Pigm 76: 714-720.
20
21 [47] Noorjahan M, Kumari VD, Subrahmanyam M, Panda L (2005) Immobilized Fe(III)-HY: an
22
23 efficient and stable photo-Fenton catalyst. Appl Catal B 57: 291–298.
24
25 [48] Trinh H, Yusup S, Uemura Y (2018) Optimization and kinetic study of ultrasonic assisted
26 esterification process from rubber seed oil. Bioresour Technol 247:51–57 59.
27
28 [49] Belachew N, Bekele G (2020) Synergy of magnetite intercalated Bentonite for enhanced
29
30 adsorption of Congo red dye. Silicon,12: 603–612.
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
18
62
63
64
65
Figure captions
1
2 Fig. 1. XRD patterns of the raw clay, FAU zeolite and the Fe-FAU catalyst.
3
4 Fig. 2. SEM images and EDX analysis of the raw clay and the FAU zeolite.
5
6 Fig. 3. Nitrogen adsorption-desorption isotherms and pore size distribution of raw clay, FAU zeolite
7 and the Fe-FAU catalyst.
8
9 Fig.e 4. FTIR spectra of raw clay material, FAU zeolite and the Fe-FAU catalyst.
10
11 Fig.e 5. Effect of Fe-FAU catalyst dose on MO degradation (Experimental parameters: [MO]0 = 20
12
13 mg/L, [H2O2]0 =150 μL, T=20°C, UVC irradiation)
14
15 Fig. 6. Effect of initial H2O2 concentration on MO degradation (Experimental parameters: [MO]0 =
16 20 mg/L, catalyst dosage=50 mg, T=20°C, UVC irradiation) and Evolution of the ratio
17
18
[H2O2]t/[H2O2]0 over time for different initial H2O2 concentrations.
19
20
Fig. 7. Influence of initial MO concentration on the degradation efficiency (Experimental
21 parameters: catalyst dosage=50 mg, [H2O2]0 =150 μL, T=20°C, UVC irradiation)
22
23 Fig. 8. The photodegradation efficiency and Iron leaching through five consecutive photocatalysis
24 experiments versus cycle number
25
26 Fig. 9. The 3D surface plots for the degradation of MO at different factors (a) the initial MO
27
28 concentration and H2O2 dosage, (b) the initial MO concentration and the catalyst dose, and (c) the
29 the H2O2 dosage and the catalyst dose.
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
19
62
63
64
65
Fig. 1
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
20
62
63
64
65
Fig. 2
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
21
62
63
64
65
1
2 Fig. 3
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
22
62
63
64
65
1
2 Fig. 4
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
23
62
63
64
65
Fig. 5
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
24
62
63
64
65
1
2
3
4 Fig. 6
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
25
62
63
64
65
Fig. 7
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
26
62
63
64
65
Fig. 8
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
27
62
63
64
65
Fig. 9
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27 (a) (b)
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51 (c)
52
53
54
55
56
57
58
59
60
61
28
62
63
64
65
1
2 Table 1. Properties of raw clay and the prepared FAU material
3
4
Materials
5 SBET [m2 /g] Vtotal [cm3 /g] Vmicro [cm3 /g] Smesopores [m2 /g] Average pore size [nm]
6 Micropore Mesopore
7 clay
Raw 50 0.15 0.007 35 - 12
8
9
Fau 360 0.33 0.0927 126.4 1.8 23.0
10
11
Fe-FAU
12 290 0.28 0.0701 120.5 1.3 21.0
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
29
62
63
64
65
Table 2: The RSM-BBD design matrix of the three variables with coded and real values; and
1 Response with actual and predicted
2
3
4
5 Response (MO
Coded Values Real Values
Run
6 degradation)
Order
7 MO MO
Catalyst dose H2O2 dosage Catalyst dose H2O2 dosage Actual Predicted
8 concentration concentration
(mg) (µL) (mg) (µL) Value Value
9 (mg/L) (mg/L)
101 +1 0 -1 100 175 20 91.95 91.96
11
122 +1 -1 0 100 50 60 89.00 86.62
13
3 -1 0 +1 20 175 100 87.00 86.99
14
154 -1 +1 0 20 300 60 75.00 77.38
16
175 0 -1 -1 60 50 20 55.00 57.37
18
196 -1 0 -1 20 175 20 59.00 57.75
20
217 0 +1 +1 60 300 100 89.00 86.63
228 0 0 0 60 175 60 93.00 93.00
23
249 0 -1 +1 60 50 100 79.00 80.13
25
10
26 0 +1 -1 60 300 20 77.00 75.87
27
11
28 +1 0 +1 100 175 100 95.00 96.25
29
12 +1 +1 0 100 300 60 89.00 90.12
30
13
31 -1 -1 0 20 50 60 57.00 55.88
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
30
62
63
64
65
Table 3 Results of ANOVA test for the experimental design data.
1
2
3
4 Degree of
5 Source Sum of Squares Mean Square F-value p-value
freedom
6
7 Model 4426.79 9 369.64 86.31 <0.0001
8
9 A 945.04 1 945.04 92.20 <0.0001
10
11 B 312.50 1 312.50 30.49 <0.0001
12
13 C 561.96 1 561.96 54.82 <0.0001
14
15 AB 81.00 1 81.00 7.90 0.0472
16
AC 155.63 1 155.63 15.18 0.0300
17
18 BC 36.00 1 36.00 3.51 0.1576
19
20 A² 30.14 1 30.14 2.94 0.011849
21
22 B² 321.98 1 321.98 31.41 0.0112
23
24 C² 85.93 1 85.93 8.38 0.0627
25
26 Residual 30.75 3 10.25
27
28 Cor Total 2457.54 12
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
31
62
63
64
65
Table 4 Summary of the main statistical specifications of the response surface quadratic correlation.
1
2
3
4
5
Std. Dev. 3.20 R² 0.9875
6 Coefficient of variance (C.V. %) 4.02 Adjusted R² 0.9785
7
8 Adequate Precision 14.376 Predicted R² 0.9499
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
32
62
63
64
65
Supplementary Material

Click here to access/download


Supplementary Material
REAC-D-23-00417Supplementary.docx

You might also like