2023-CHIBANV Et Al

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

Contents lists available at ScienceDirect

Case Studies in Chemical and Environmental Engineering


journal homepage: www.sciencedirect.com/journal/case-studies-in-chemical-
and-environmental-engineering

Case Report

Performance assessment of a phengite clay-based flat membrane for


microfiltration of real-wastewater from clothes washing: Characterization,
cost estimation, and regeneration
Salek Lagdali a, b, **, Youssef Miyah c, Mohamed El-Habacha a, b, Guellaa Mahmoudy a,
Mohammed Benjelloun c, Soulaiman Iaich a, b, *, Mohamed Zerbet a, Mohamed Chiban a,
Fouad Sinan a
a
Laboratory of Applied Chemistry and Environment, Faculty of Sciences, Ibnou Zohr University, Agadir, Morocco
b
Research Team of Energy and Sustainable Development, Higher School of Technology, Guelmim, Ibnou Zohr University, Agadir, Morocco
c
Laboratory of Materials, Processes, Catalysis, and Environment, Higher School of Technology, University Sidi Mohamed Ben Abdellah, Fez, Morocco

A R T I C L E I N F O A B S T R A C T

Keywords: The challenge of the scientific community is to synthesize innovative, low-cost, and environmentally friendly
Cost estimation membrane materials for the treatment of industrial wastewater. The objective of this work is the elaboration and
Microfiltration membrane characterization of a new flat ceramic membrane based on a natural Moroccan phengite clay by the paste casting
Phengite clay
method for microfiltration applications. The ceramic membrane was sintered from 850 ◦ C to 1150 ◦ C for 2 hours.
Real-wastewater
The optimal membrane sintered at 1050 ◦ C has a porosity of 34.5%, an average pore diameter of 3.9 μm, water
Regeneration
permeability of 43.50 L/h m2 bar, mechanical strength of 26.7 MPa, and excellent chemical corrosion resistance
in acidic and basic media. The performance of the optimal membrane was evaluated by frontal microfiltration of
the pre-treated real wastewater (RWW3) from a local clothes washing. The obtained results show that the
removal percentage of electrical conductivity, total dissolved solids, and suspended matter is 66.2%, 71.8%, and
100% respectively. The cost of preparing the ceramic membrane was estimated at 3.5 $/m2, which is cheaper
compared to those commercially available. The high regeneration efficiency showed that demineralized water
was able to adequately clean the fouled microfiltration membrane by 82%. The obtained filtration results are
very promising and could allow the use of the membrane prepared from a locally available material as an
alternative process in the treatment of various sources of industrial wastewater.

1. Introduction employed in a broad range of industries, including food processing,


biotechnology, chemistry, seawater desalination, and the treatment of
Water pollution is a major environmental issue, potentially threat­ industrial liquid effluent [7,8]. The performance of microfiltration is due
ening ecosystems and public health. This situation implies the imple­ to the use of polymer or ceramic-based membranes. However, polymeric
mentation of strict regulations to reduce the level of pollution due to membranes have some disadvantages, in particular low mechanical and
human and industrial activities [1,2]. As a result, many researchers have thermal resistance, and short life span. Ceramic membranes are more
developed numerous pollutant removal techniques including photo­ favorable and preferred than polymeric membranes for wastewater
catalysis, ozonation, coagulation-flocculation, ion exchange, and treatment because of their low operating costs and low energy con­
adsorption [3–5]. However, these techniques represent an exorbitant sumption, and high resistance [9–12]. In addition, MF membranes have
cost and produce quantities of secondary waste that are difficult to treat shown excellent results in improving water quality through the removal
[6]. of micro-organisms, colloidal elements, and suspended matter [13,14].
Microfiltration (MF), ultrafiltration (UF), nanofiltration (NF), Compared to UF membranes, MF membranes also have excellent resis­
reverse osmosis (RO), and other membrane technologies are often tance to fouling [15].

* Corresponding author. Laboratory of Applied Chemistry and Environment, Faculty of Sciences, Ibnou Zohr University, Agadir, Morocco.
** Corresponding author. aboratory of Applied Chemistry and Environment, Faculty of Sciences, Ibnou Zohr University, Agadir, Morocco.
E-mail addresses: salek.lagdali@edu.uiz.ac.ma (S. Lagdali), s.iaich@uiz.ac.ma (S. Iaich).

https://doi.org/10.1016/j.cscee.2023.100388
Received 29 April 2023; Received in revised form 30 May 2023; Accepted 30 May 2023
Available online 1 June 2023
2666-0164/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

Table 1 magnesite to treat industrial textile wastewater. Khemakhem et al. [29]


Physicochemical characteristics of RWW1, RWW2, and RWW3. synthesized a tubular ceramic microfiltration and ultrafiltration mem­
Physicochemical Parameters RWW1 RWW2 RWW3 branes from natural Tunisian illite clay to treat a dextran polymer so­
lution of different molecular weights.
Temperature ( C) ◦
25 25 25
pH 7.3 8.6 8.4 This paper describes the preparation and characterization of a new
Electrical Conductivity (μS/cm) 4300 3933 3616 low-cost flat ceramic microfiltration membrane, based on natural
TDS (ppm) 2580 2360 2170 phengite clay collected in South-East Morocco, whose purpose is to treat
Suspended matter (mg/L) 7160 160 40 real wastewater. This clay was chosen for its natural abundance, higher
porosity, mechanical stability, low sintering temperature, and good
chemical resistance, especially in alkaline environments. The membrane
was prepared by the paste casting method, followed by a drying and
sintering process in a programmable electric furnace. The physico­
chemical and mechanical properties of the prepared ceramic membrane
such as shrinkage, porosity, bulk density, water absorption, membrane
surface morphology, pore size distribution, mechanical strength,
chemical resistance, and water permeability were investigated as a
function of sintering temperature (850 ◦ C–1150 ◦ C). The efficiency of
the optimal membrane was evaluated by filtration of pre-treated real
wastewater from a local clothes washing (RWW3) in terms of permeate
flux, pH, electrical conductivity, suspended matter, and total dissolved
solids (TDS). In addition, the regeneration, and the anti-fouling perfor­
mance analysis were evaluated. Finally, the cost of preparing our
membrane was examined in terms of the price of raw materials, the cost
of electrical power consumption, and the manpower cost.

2. Materials and methods

2.1. Raw materials and chemical products

The flat ceramic microfiltration membranes were prepared using


natural clay collected from the Zagora region in South-East Morocco
(Latitude: 30◦ 19′ 56′′ Nord, Longitude: 5◦ 50′ 18′′ Ouest, Altitude to sea
level: 733 m). The clay sample was ground in a ball mill type MM 500,
RETSCH®, and then sieved using a digital electromagnetic vibrating
analysis sieve type KTL 300. Fragments smaller than 125 μm are used for
the preparation of flat microfiltration membranes. Chemical resistance
Fig. 1. Manufacturing process of the flat ceramic membrane through paste
tests were performed using hydrochloric acid (HCl, 37 wt%), and so­
casting method.
dium hydroxide (NaOH, 98 wt%). Raw real wastewater from a local
clothes-washing workshop (RWW1; Guelmim, Morocco) was used for
Commercial membranes generally based on pure oxides of titanium,
the front-end microfiltration applications. The RWW1 is physically pre-
zirconia, alumina, and silica are not the best choice because they require
treated by decantation and filtration before the microfiltration experi­
higher costs for raw ingredients and greater energy consumption during
ment to remove large solid particles and minimize rapid clogging of our
the thermal treatment, it needs a sintering temperature higher than
ceramic membrane (Fig. S1). Table 1 shows the physicochemical char­
1200 ◦ C [16]. The scientific community has recently become interested
acteristics of the raw real wastewater (RWW1), decanted for 24 hours
in ceramic membranes developed from clay due to their low-cost and
(RWW2), and filtered through a 100 μm sieve (RWW3).
excellent additional properties. In addition, multiple studies have been
The raw real wastewater (RWW1) is characterized by an almost
performed to develop new types of ceramic membranes made from
neutral pH (7.3), while the decanted real wastewater (RWW2), filtered
natural materials such as kaolinitic clay [17], pyrophilitic clay [18],
on a 100 μm sieve (RWW3) is characterized by a basic pH (8.6–8.4). A
bentonite clay [19], phosphate [20], and dolomite [21]. Moreover, the
remarkable decrease in the electrical conductivity from 4300 μS/cm to
manufacturing of ceramic membranes through natural clays will in­
3933 μS/cm, and then to 3616 μS/cm for RWW1, RWW2, and RWW3
crease their economic value, and evolve the technology of membranes
respectively. The high values of electrical conductivity are due to the
[22,23].
existence of chemical salts in our real wastewater. An abatement of the
From a technical point of view, some researchers have developed
high concentrations of suspended matter and TDS which are respectively
low-cost ceramic membranes using natural materials. Mouiya et al. [20]
7160 mg/L and 2580 ppm for RWW1 to 40 mg/L and 2170 ppm for
produced a flat ceramic microfiltration membrane based on natural clay
RWW3. The presence of suspended matter and TDS is related to the
and Moroccan phosphate for desalination and industrial wastewater
organic and inorganic materials that can be caused by membrane fouling
treatment. Rekik et al. [24] fabricated a ceramic microfiltration mem­
and biofouling during microfiltration.
brane from kaolinitic clay to treat raw effluent from the seafood in­
dustry. Agarwalla et al. [25] developed a flat ceramic microfiltration
2.2. Elaboration of the circular flat ceramic membrane
membrane using natural kaolinitic clay collected in India for the treat­
ment of methylene blue dye. Achiou et al. [26] prepared a tubular
The flat membranes were prepared by the paste casting method (wet
ceramic microfiltration membrane based on pozzolan for raw seawater
process) using natural phengite clay with a particle size of less than 125
desalination. Saja et al. [27] manufactured a ceramic microfiltration
μm. The clay powder was mixed with distilled water (230 ml/Kg) to
membrane based on natural perlite for the treatment of two industrial
obtain the ceramic pastes, followed by aging in a closed box for 24 hours
wastewater discharges (food processing and tannery). Manni et al. [28]
under high humidity (4 ◦ C) to ensure the perfect distribution of water in
fabricated a tubular ceramic microfiltration membrane from natural
the prepared pastes, and to avoid the formation of air bubbles [1]. The

2
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

pastes were cast on a surface in the shape of circular discs using a ( )


msat − mdry
stainless steel ring with an inner diameter of 49 mm and a thickness of 3 Porosity = × 100 (3)
mm. After air drying for 24 hours, the membranes were transferred to a msat − mwet
muffle furnace (Nabertherm L(T)9/12/B410, Germany) for heat treat­ ( )
msat − mdry
ment from room temperature to sintering temperature. The preparation Water absorption = × 100 (4)
mdry
process of the membranes is exhibited in Fig. 1.
mdry
2.3. Thermal program Bulk density = (5)
msat − mwet

After the preparation of the ceramic membranes, a heat treatment The mechanical strength of flat ceramic membranes that were
step is essential in this process. Generally, heat treatment is divided into developed and sintered at different temperatures was evaluated using
two main parts: drying and sintering [26]. Fig. S2 shows the thermal the three-point flexural method. Rectangular samples with 60 mm × 20
program for the sintering of the produced flat membranes. The sintering mm × 5 mm dimensions were developed by the paste molding method
process has two steps: a step at 250 ◦ C and a step at the final sintering using the same formulation of pastes employed to manufacture our flat
temperature. Firstly, the temperature was increased by a firing rate of membranes. These specimens are sintered using the same thermal pro­
2 ◦ C/min from room temperature to 250 ◦ C and then maintained at this tocol mentioned above. Subsequently, the flexural strength σ (MPa) was
temperature for 1 hour, then the temperature was increased by 4 ◦ C/min evaluated using a laboratory scale tool conforming to ASTM C674-88,
to the desired final temperature. Finally the membranes were sintered at and then calculated according to the following formula, Eq. (6) [1,23].
different temperatures (850 ◦ C, 900 ◦ C, 950 ◦ C, 1000 ◦ C, 1050 ◦ C, 3.F.L
1100 ◦ C, 1150 ◦ C) for 2 hours. The sintering temperature is necessary for σ= (6)
2.l.e2
the final consolidation of the flat ceramic membranes [27].
where F is the force applied to the specimen in (N), L is the distance
2.4. Characterization techniques between the two specimens’ supports in (mm), l is the specimen width in
(mm) and e is the specimen thickness in (mm). The filtration experiment
The chemical composition of the clay was determined by X-ray was performed at the laboratory level using a frontal filtration pilot. The
fluorescence with a SIEMENS 300 instrument. The structural properties experimental setup is shown in Fig. S3. Experimental instruments made
of the raw and heat-treated clay were characterized by X-ray diffraction of transparent plastic were used for filtration at room temperature (25
(XRD) using a Philips X’Pert X-ray diffractometer operating with Cu Kα ± 1 ◦ C) and low pressure (0.1, 0.15, and 0.2 bar). The effective filtration
radiation (λ = 1.54056 Å). The surface morphology of the developed surface is 15.5 cm2 and the flat ceramic membrane is inserted hori­
ceramic membranes (top view) was observed through scanning electron zontally into the membrane module. The membrane was immersed in
microscopy (SEM; JEOL JSM IT-100 type, Japan). Image J software distilled water for 24 hours before each filtration test to obtain a stable
version 1.44e was used to estimate the average pore diameter and pore flux [38]. The variation of the water height h between the surface of the
size distribution (PSD) of the membranes through the SEM images [30]. flat membrane and the water level in the feed tank allows the trans­
The average pore diameter dave (μm) was calculated using Eq. (1) membrane pressure ΔP (bar) to vary. This pressure is calculated by the
considering that the membrane surface has a cylindrical porous texture following relationship, Eq. (7) [39].
[22,31,32].
ΔP = ρ.g.h (7)
⎡ ⎤0.5
3
∑n 2
with ρ being the density of water (g/cm ), h being the height of water
⎢ i=1 ni .di ⎥
dave = ⎢ ⎥ (1) (m), and g being the gravity field (N/Kg).
⎣ ∑ n ⎦
ni Distilled water was used to measure the water permeability of our
i=1
membrane. The permeate flux Jw is measured using the volume of
with n being the number of pores and di being the diameter of the ith permeate collected as a function of time. This magnitude can be calcu­
pore (μm). We measure the diameter of the flat ceramic membranes lated by using Eq. (8) [40]. The permeability (LP) is calculated by
before and after the sintering process with a digital caliper, the applying Darcy’s law, which is based on the variation of the permeate
shrinkage (%) may be measured by Eq. (2) [33]. flux Jw as a function of the transmembrane pressure ΔP, Eq. (9) [41].
( ) V
Shrinkage =
D0 − D1
× 100 (2) Jw = (8)
D0 S.Δt

where D0 (mm) and D1 (mm) are the diameter of the green and the Jw = Lp .ΔP (9)
sintered membrane respectively. The porosity (%), water absorption 2
with Jw is the permeate flux (L/h.m ), S is the membrane filtration
(%), and bulk density (g/cm3) of the clay-based ceramic membranes surface (m2), V is the permeate volume (L), Δt is the sampling time (h),
were obtained by boiling water immersion according to ASTM C373-88. LP is the permeability (L/h.m2.bar), and ΔP is the transmembrane
To further explain the experiment, the samples were dried in an oven at pressure (bar). Chemical resistance tests are performed in the acidic and
110 ◦ C for 24 hours to determine the dry mass mdry (g), then placed in a basic media at room temperature (25 ± 1 ◦ C). pH = 1 and pH = 13.5
distilled water bath and boiled for 5 hours, taking care to cover the solutions were prepared using hydrochloric acid (HCl) and sodium hy­
specimens with water, and using pins to separate the samples from the droxide (NaOH), respectively. The prepared membranes were pre-dried
bath’s bottom and sides. After boiling, the samples are placed in the bath at 80 ◦ C for 24 hours and then immersed in beakers containing acidic
for another 24 hours. Then the wet mass mwet (g) of each sample sus­ and basic solutions for 7 successive days. After washing with distilled
pended in water is determined. Finally, the saturated mass msat (g) of water and drying for 24 hours at 80 ◦ C the weight loss (%) is measured,
each sample is determined after softly wiping the surface with a lint-free which is indicated the chemical resistance of the membranes according
cloth or cotton towel to remove any excess water from the surface. Eqs. to the following formula, Eq. (10) [10,42].
(3)–(5) were used for calculating the porosity, water absorption, and ( )
bulk density of water, respectively [34–37]. Weight loss =
m0 − m1
× 100 (10)
m0

3
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

Table 2 using pre-treated real wastewater RWW3. The initial permeate flux of
Chemical composition of raw phengite clay distilled water (Jw0, L/h.m2) was recorded for 180 minutes at 0.2 bar.
by XRF. RWW3 was passed through the membrane with similar experimental
Oxides wt.% conditions to produce permeate flux (Jp, L/h.m2). After 1 hour of
SiO2 40.330
washing the membrane with demineralized water, it was repassed with
Al2O3 13.537 distilled water (Jw1, L/h.m2). Additionally, Eqs. 14–17 were used to
Fe2O3 6.044 calculate the flux recovery rate (FRR (%)), total flux decline rate (TFR
CaO 11.676 (%); total fouling), reversible flux decline rate (RFR (%); reversible
MgO 4.488
fouling), and irreversible flux decline rate (IFR (%); irreversible fouling)
SO3 0.318
Na2O 0.650 [9,44].
K2O 2.948 ( )
Jw1
TiO2 0.690 Flux recovery rate = × 100 (14)
P2O5 0.139 Jw0
ZnO 0.012
( )
Mn2O3 0.155 Jp
Cr2O3 0.013 Total flux decline rate = 1 − × 100 (15)
Jw0
SrO 0.026
L.O.Ia 18.620 ( )
Total 99.646 Jw1 − Jp
Reversible flux decline rate = × 100 (16)
a
Jw0
Loss on ignition at 1000 ◦ C.
( )
Jw0 − Jw1
where m0 (g) and m1 (g) are the masses of the membrane before and Irreversible flux decline rate = × 100 (17)
Jw0
after the chemical contact respectively.
3. Results and discussion
2.5. Application of frontal microfiltration
3.1. Characterization of raw phengite clay powder
The elaborated optimal ceramic membrane’s performance was
assessed using real wastewater filtered through a 100 μm sieve (RWW3). 3.1.1. XRF analysis
The microfiltration tests were performed at a constant pressure of 0.2 The chemical composition of the raw phengite clay in mass per­
bar for 3 hours. The feed tank was refreshed every 30 minutes to reduce centages (wt.%) of the oxides is presented in Table 2. The loss on igni­
and adjust for the concentration effect. The physicochemical parameters tion (L.O.I) is calculated at 1000 ◦ C for the raw clay. The clay is mainly
of the permeate such as flux, pH, electrical conductivity, suspended composed of a large amount of silica oxide (40.330%), alumina oxide
matter, and total dissolved solids were measured every 30 min of (13.537%), calcium oxide (11.676%), and iron oxide (6.044%). These
filtration. A multiparameter instrument (HACH HQ40d, U.S.A) was used oxides were found to be the main constituents of the raw clay. This
for measuring the pH and electrical conductivity of the feed and group of oxides represents about 71,587% of the total, which explains
permeate solutions. The total dissolved solids were measured with a the physical and mechanical properties of this natural clay. The SiO2/
mobile TDS meter type TDS-3 Korea. The suspended matter was detec­ Al2O3 mass ratio is about 3, indicating that the phengite clay is an
ted by filtering using a 0.45 μm pore size membrane. The suspended aluminosilicate [41]. A very small amount of other oxides such as MgO
matter concentration in mg/L is calculated using the differential mass (4.488%), SO3 (0.318%), Na2O (0.650%), K2O (2.948%), TiO2
ratio of the filtered water volume, Eq. (11) [6]. (0.690%), P2O5 (0. 139%), ZnO (0.012%), Mn2O3 (0.155%), Cr2O3
(M1 − M0 ) (0.013%), and SrO (0.026%) as impurities were also detected.
Suspended matter = (11)
Vw
3.1.2. XRD analysis
where M0 is the mass of empty membrane (mg), M1 is the mass of To identify the influence of thermal treatment on the mineralogical
membrane and residue after oven drying at 105 ◦ C/24 h (mg) and Vw is composition of phengite clay, XRD analysis was used on both the raw
the volume of filtered water (L). The rejection factor (%) for all exper­ sample and samples treated at different temperatures. The obtained XRD
imental parameters was calculated using the following relationship, Eq. diagrams are presented in Fig. 2. It should be noted that the XRD
(12) [2,35]. analysis of our raw clay was carried out in another research work [45],
( ) and the results obtained reveal the existence of several crystalline pha­
XFeed − XPermeate
Rejection = × 100 (12) ses, such as quartz (Q), calcite (C), phengite (Ph), dolomite (D),
XFeed
vermiculite (V) and hematite (H). From the diffractograms of
850–900 ◦ C, we note the total decomposition of calcite and the disap­
where XFeed and XPermeate are the experimental parameters of the feed
pearance of reflections of vermiculite and hematite. On the diffracto­
and permeate respectively.
gram of 950 ◦ C we observed the dehydroxylation of phengite and the
appearance of the albite phase (A; NaAlSi3O8). As the temperature rises
2.6. Membrane regeneration and anti-fouling performance analysis
to 1000–1050 ◦ C, the diffractograms show the disappearance of dolo­
mite and the start of crystallization of the new ceramic phase of mullite
After a filtration process, the membrane is completely washed with
(M; 3Al2O3,2SiO2). When the temperature increases to 1100–1150 ◦ C,
demineralized water or any appropriate cleaning solution (e.g., alkali or
the diffractograms indicate the elimination of albite caused by fusion at
acid solution) to eliminate the foulant and restore the original perme­
about 1100 ◦ C [46], and the predominance of the mullite phase.
ability. The cleaning solution chosen is determined by the foulant va­
riety as well as the nature of the membrane. The cleaning efficiency (%)
can be calculated by Eq. (13) [43]. 3.2. Characterization of the phengite clay membrane
( )
Water permeability after cleaning
Cleaning efficiency = × 100 (13) 3.2.1. Visual appearance of flat membranes
Water permeability before filtration
Fig. 3 illustrates the sintered ceramic membranes at different tem­
The anti-fouling behavior of the optimal membrane was evaluated peratures from 850 ◦ C to 1150 ◦ C. The membranes’ surfaces are

4
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

Fig. 2. XRD of phengite clay at different sintering temperatures (A: Albite, D: Dolomite, M: Mullite, Ph: Phengite, and Q: Quartz).

Fig. 3. Views of sintered flat membranes at different temperatures.

membranes becomes relatively dark brown at 1150 ◦ C due to the


dominance of mullite phases.

3.2.2. Shrinkage
The shrinkage of the membranes was calculated using the diameter
dimensions before and after the sintering process. During thermal
treatment, shrinkage is often related to the weight loss and densification
of the ceramic membranes [14]. Fig. S4 presents the percentage of
shrinkage of flat ceramic membranes sintered at various temperatures
(850 ◦ C–1150 ◦ C). It is observed that sintering temperature has a sig­
nificant effect on shrinkage. As seen in the figure, when the sintering
temperature increases from 850 ◦ C to 1150 ◦ C, the shrinkage of the
membranes also increases from 1.3% to 8.6%. The densification process
(the internal rearrangement of the membrane) explains the increase in
shrinkage, which is confirmed by the SEM images. Several studies have
shown the same increasing shrinkage percentage evolution with sin­
tering temperature [47,48]. The deposition of UF and NF layers becomes
more perfect when the shrinkage of the membrane supports is lower to
avoid cracking and deformation of the filter layer [28].
Fig. 4. Variation of porosity and mechanical strength with sintering
3.2.3. Porosity and mechanical strength
temperature.
Fig. 4 shows the variation of porosity and mechanical strength as a
function of sintering temperature. For the porosity, Fig. 4 illustrates two
homogenous without any sort of macro defects or ruptures. The mem­
steps, the first one is in which the porosity increases from 30.3% to
branes keep the same cylindrical structure and vary in color from brick
34.5% by the increase of the sintering temperature from 850 ◦ C to
red to dark yellow when the sintering temperature increases from 850 ◦ C
1050 ◦ C. This evolution is explained by the release of CO2 from the
to 1100 ◦ C. In other words, the more the sintering temperature is
decomposition of calcium carbonates, and the combustion of organic
increased, the darker the color of the membrane will become. The
matter that exists naturally at high temperatures [49], then we observe
change in color of the flat ceramic membranes after the heat treatment is
in the second stage a decrease from 34.5% to 17.1% when the
caused by iron oxide oxidation [35]. In addition, the color of the

5
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

Fig. 5. SEM micrographs of the surface of sintered membranes at (a) 850 ◦ C, (b) 900 ◦ C, (c) 950 ◦ C, (d) 1000 ◦ C, (e) 1050 ◦ C, (f) 1100 ◦ C, and (g) 1150 ◦ C.

6
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

Fig. 6. (a) The average pore diameter of the membranes as a function of sintering temperature, and (b) pore size distribution of the optimal membrane sintered at
1050 ◦ C/2h.

Fig. 7. Evolution of water flux, (a) as a function of time at different pressures, and (b) as a function of applied pressure.

temperature increases from 1050 ◦ C to 1150 ◦ C. This is due to the [33]. Then the water absorption decreases from 19.3% to 7.8% when the
densification of the porous structure and the agglomeration of the grains temperature varies from 1050 ◦ C to 1150 ◦ C. This decrease is explained
with the increase of the sintering temperature [50,51]. The sintering by the densification phenomenon of the ceramic material [28]. This
temperature can also be stated to have a considerable influence on the allows the formation of a high-density ceramic membrane and improves
porosity of the membranes. The optimal sintering temperature chosen to its mechanical properties. The presence of alkaline and alkaline-earth
continue our study is 1050 ◦ C, due to the highest porosity (34.5%). The oxides such as MgO, Na2O, and K2O favors the formation of glassy
optimal ceramic membrane must have good mechanical strength and be phases strongly responsible for the reduction of water absorption by
strong enough to resist all the hydraulic pressures applied during the filling the pore volume of the membrane materials [35]. On the other
filtration process. The mechanical strength of the ceramic membranes hand, an inverse behavior was recorded for the bulk density, which
increases from 14.9 MPa to 29.5 MPa when the sintering temperature decreased from 1.81 g/cm3 to 1.77 g/cm3, when the temperature in­
increases from 850 ◦ C to 1150 ◦ C with the optimal membrane exhibiting creases from 850 ◦ C to 1050 ◦ C. Then increases from 1.77 g/cm3 to 2.2
a value of 26.7 MPa. The increase is primarily attributable to the g/cm3 when the temperature increased from 1050 ◦ C to 1150 ◦ C which
densification and vitrification of some phases, namely amorphous silica, was due to the formation of a very viscous silica phase that facilitated
and albite, as well as the crystallization of more resistant phases, such as the crystallization of mullite [52]. In addition, it also indicates the for­
mullite [14,35]. A similar progression was identified for the membranes mation of a solid ceramic body through the densification of the material.
fabricated from natural Moroccan clay [50]. These results are in perfect Bulk density and porosity are inversely proportional, therefore the re­
agreement with the SEM images. sults of density confirm the results of porosity [28].

3.2.4. Water absorption and bulk density 3.2.5. Microstructural analysis


Fig. S5 displays the variation of water absorption and bulk density The densification phenomenon and surface structure of the prepared
with sintering temperature. The water absorption increases with tem­ ceramic membranes are observed by SEM. Fig. 5 shows the surface
perature from 16.7% to 19.3% when the temperature increases from morphology of membranes sintered at 850 ◦ C, 900 ◦ C, 950 ◦ C, 1000 ◦ C,
850 ◦ C to 1050 ◦ C. The water absorption of the flat ceramic membranes 1050 ◦ C, 1100 ◦ C, and 1150 ◦ C for 2 hours. The images from the SEM
correlates with the porosity, i.e. the two parameters evolve similarly indicate that the sintering temperature effects on the membrane

7
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

Fig. 8. (a) Permeate flux of RWW3 as a function of filtration time, (b) pH of RWW3 permeate as a function of filtration time, and (c) rejection of suspended matter,
TDS, and electrical conductivity of the RWW3 as a function of filtration time, (Experimental conditions: ΔP = 0.2 bar, T = 25 ◦ C).

Fig. 9. (a) RWW1, RWW2, and RWW3 before and after MF treatment, and (b) RWW3 microfiltration mechanism using our optimal ceramic membrane, (Experi­
mental conditions: ΔP = 0.2 bar, T = 25 ◦ C).

morphology. It can be seen that the surface of the ceramic membranes 850 ◦ C, 900 ◦ C, and 950 ◦ C are weakly sintered and the grains of the clay
does not show any defects such as micro-cracks, which confirms that the powder are not well compacted with each other. This confirms the in­
membranes are of very good quality. The ceramic membranes treated at crease in porosity (from 30.3% to 31.7% respectively). At 1000 ◦ C,

8
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

Table 3 start of material densification. The same evolution has been observed in
Physicochemical characteristics of feed and permeate recovered after 3 hours of several works [54,55]. Fig. 6b represents the pore size distribution of the
microfiltration, (Experimental conditions: ΔP = 0.2 bar, T = 25 ◦ C). optimal sintered membrane at 1050 ◦ C. The figure shows that the
Physicochemical characteristics RWW3 membrane has a single peak, which represents a unimodal distribution
Feed Permeate Rejection (%)
of the pore size. The average diameter is about 3.9 μm. Note that 78.6%
of the pores have a diameter d < 3.9 μm.
pH 8.4 8.7 –
Electrical Conductivity (μS/cm) 3616 1224 66.2
TDS (ppm) 2170 612 71.8 3.2.7. Permeability of the optimal ceramic membrane
Suspended matter (mg/L) 40 0 100 Permeability is an essential property to verify the performance of the
prepared ceramic membrane. Fig. 7a illustrates the evolution of water
flux measured at three different pressures (0.1, 0.15, and 0.2 bar) during
sintering becomes increasingly significant due to the phenomenon of 180 min. According to Fig. 7a, when the pressure varies from 0.1 to 0.2
densification, but is still not satisfactory to obtain excellent cohesion bar the water flux varies from 4.77 to 8.52 L/h.m2. The two fluxes
between the clay powder grains. When the temperature rises to 1050 ◦ C, measured at 0.15 and 0.2 bar decrease slightly with time until they have
the membrane has a uniform porous structure and the pores between the established a stable flux after 60 minutes, and the flux measured at 0.1
grains are relatively large, which is explained by the better porosity bar remains stable during the filtration. Fig. 7b shows a linear evolution
(34.5%). At 1100 ◦ C we observed that the densification continues its of the water flux when the pressure increases from 0.1 to 0.2 bar. This
evolution due to the partial vitrification of the amorphous silica and confirms the verification of Darcy’s law [56]. Pressure is the only driving
some impurities present in the clay material [19,35]. Vitrification force for permeation. The water permeability of the sintered ceramic
already dominates at the final sintering temperature of 1150 ◦ C and the membrane at 1050 ◦ C/2h obtained through the slope of the line is 43.50
pores are almost partially closed, which explains the low porosity value L/h m2 bar.
(17.1%) [8,53].
3.2.8. Chemical resistance
3.2.6. Average pore diameter and pore size distribution The corrosion chemical resistance of the optimal ceramic membrane
The average pore diameter of the elaborated ceramic membranes sintered at 1050 ◦ C was shown in Fig. S6. Two media were chosen to
and the pore size distribution of the optimal membrane were evaluated evaluate the weight loss for 7 days, acidic (HCl; pH = 1) and basic
through the Image J software. Fig. 6a presents that the average pore (NaOH; pH = 13.5). The ceramic membrane lost 4.6% of its weight after
diameter increases from 2.1 μm to 12.3 μm when the sintering tem­ immersion in the acidic solution. Then in the basic medium, the ceramic
perature varies from 850 ◦ C to 1150 ◦ C. The opening of the pores with membrane has excellent chemical resistance, and the weight loss does
temperature is caused by the agglomeration of the clay grains and the not exceed 0.6% during 7 days. The same evolution of chemical

Table 4
Comparison of the properties of the manufactured membrane with those of various MF membranes.
Membrane material Sintering Porosity Pore Permeability (L/ Mechanical Feed (Concentration/ Applied Rejection Reference
temperature (%) size h.m2.bar) strength (MPa) Turbidity/Suspended pressure (%)
(◦ C) (μm) matter) (bar)

Natural Moroccan clay 1100 28.11 2.5 928 17.5 Tannery beam house 0.14 99.80 [20]
and phosphate effluent (760 NTU)
Raw seawater (120 99.62
NTU)
Synthetic solution of 99.86
aluminum chloride (at
high pH) (725 NTU)
Natural kaolin clay 1250 27 0.73 20 28 Cuttlefish effluent (335 3 99 [24]
NTU)
− 9
Natural kaolin clay 850 34.5 2.28 6.12 × 10 (m 7.1 Methylene 10 1.5 99.07 [25]
Pa− 1 s− 1) blue dye mg/L
100 93.45
mg/L
Natural Moroccan 950 52.11 1.70 1433.46 21.68 Agro-food effluent 0.12 97 [27]
perlite (224 NTU)
Tannery effluent (36 96
NTU)
Natural magnesite clay 1100 48.15 1.12 922 6.1 Industrial textile 0.12 99.9 [28]
wastewater (144.5 ±
0.6 NTU)
Pyrophyllite clay 1150 40.50 0.80 1000 30 Raw seawater (20 1 99.40 [41]
NTU)
Mineral coal fly-ash 800 51 0.25 475 19.5 Textile dye wastewater 1 98.72 [54]
(45.5 NTU)
Alumina membrane - - 0.2 1022 – Beamhouse effluent 1 98–99 [64]
(commercial (132 NTU; suspended
ceramic membrane) matter = 4.32 g/L)
Textile effluent (219 89–94
NTU; suspended
matter = 6.50 g/L)
Dicing effluent (132 100
NTU; suspended
matter = 2.74 g/L)
Natural Moroccan 1050 34.5 3.9 43.50 26.7 RWW3 (suspended 0.2 100 This work
phengite clay matter = 40 mg/L)

9
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

Fig. 10. (a) Permeability of the membrane after washing with demineralized water, (b) clogged and washed membrane, (c) Jw0, Jp, and Jw1 as a function of filtration
time (Experimental conditions: ΔP = 0.2 bar, T = 25 ◦ C), and (d) fouling parameters.

resistance as a function of time has been reported in the literature [20, cake. Furthermore, this cake gives additional hydraulic resistance to the
57,58]. The loss of mass in the concentrated NaOH solution is negligible, membrane, which finally leads to the fouling of the membrane [39]. This
demonstrating its potential use in the pretreatment of highly alkaline fouling behavior is commonly encountered in the microfiltration process
media. In addition, chemical cleaning of our membrane with an alkaline as reported by Iaich et al. [6]. Fig. 8b shows the evolution of pH as a
solution at ambient temperature is preferable to cleaning with an acidic function of filtration time. The pH values remain constant during the
solution. The results of chemical resistance obtained indicate, that the microfiltration process. It should be noted that the pH goes from 8.4 in
ceramic membrane elaborated at 1050 ◦ C is efficient and desirable for the feed to 8.7 in the permeate and thereafter we noticed stability. This
microfiltration applications of industrial or urban wastewater. evolution could be explained by the elimination of the bio-organic
matter present in RWW3, which generally has an acid character [28,38].
3.3. Microfiltration of RWW3 Fig. 8c presents the variation of TDS, suspended matter, and elec­
trical conductivity removal with filtration time. The TDS rejection rate
In the present study, we used the optimal membrane to conduct a increases from 26.3% to 71.8% after 3 hours of filtration. This evolution
practical application of frontal microfiltration of pre-treated real can be linked to the adsorption of TDS on the filtration surface. On the
wastewater RWW3. The objective was to evaluate the purification per­ other hand, the suspended matter was completely removed. This
formance of our membrane. This type of wastewater is saturated with removal attained 100% during the frontal microfiltration period. The
organic matter, suspended solid particles, and chemical components larger particle size of the RWW3 is the main reason for the removal of
coming from the leaching products of the used clothes. Some of the suspended matter by our ceramic membrane [61]. The removal of TDS
detergents can produce allergies and can be hazardous to human health. and the suspended matter is caused by the accumulation of rejected
Therefore, this water must be treated to protect the environment [28]. particles forming a deposit on the membrane’s surface, which partially
Fig. 8a plots the variation of permeate flux with time for the sintered reduces pore size. This phenomenon is generally observed in the
membrane at 1050 ◦ C under 0.2 bar pressure for 180 min. The permeate microfiltration process [2,62,63]. Concerning electrical conductivity,
flux decreases with time from 4.13 L/h.m2 to 2.58 L/h.m2, then stabi­ the rejection rate is significant as a function of time, it goes from 11.5%
lizes after 120 min. These results can be explained by the clogging of the to 66.2% after 3 hours of filtration. This evolution is mainly linked to the
ceramic membrane by particles finer than the pores [6,59,60]. The deposition and adsorption of dissolved salts on the cake that was created
reduction in permeate flux is caused by the adsorption of solid and by the accumulation of large solid particles on the filtering surface of our
colloidal particles or micro-organisms on the filtering surface of the ceramic membrane [24].
membrane and the subsequent formation of a deposit in the form of a Fig. 9a illustrates the complete removal of suspended matter

10
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

Table 5 by the best physicochemical, mechanical, and pollution removal


Cost estimation of fabricated flat ceramic membrane. properties.

3.4. Regeneration and anti-fouling analysis

After the microfiltration process, our membrane is clogged at the


filtering surface by the solid particles removed from RWW3. And to
restore the initial conditions of the membrane in terms of permeability,
we proceeded to regeneration by washing it with demineralized water.
Fig. 10a shows the evolution of the water flux of the regenerated
membrane as a function of pressure. It can be seen that the permeability
obtained after washing (35.69 L/h m2 bar) is extremely comparable to
the initial permeability (43.50 L/h m2 bar) with 82% as cleaning effi­
ciency. Fig. 10b presents the state of the clogged membrane after
microfiltration of RWW3 and washed. Through washing with deminer­
alized water, the flux of distilled water (Jw1) was better improved. It can
be deduced that demineralized water could mitigate the membrane
fouling and increase the membrane regeneration flux (Fig. 10c).
Furthermore, The fouling parameters FRR, TFR, RFR, and IFR are
summarized in Fig. 10d. The FRR shows that the membrane had the best
anti-fouling capability. The FRR value reached 84.9% after washing the
membrane with demineralized water, which means that after washing,
our membrane can restore 84.9% of its filtering performance. In addi­
tion, fouling was studied using TFR (total fouling), RFR (reversible
fouling), and IFR (irreversible fouling). The TFR for our membrane was
69.7%, while the RFR and IFR were 54.6% and 15.1% respectively.

3.5. Membrane cost estimation

The efficient implementation of membrane technology on an in­


dustrial scale is based on the manufacturing cost. Based on several re­
ported cost estimates in the scientific literature, commercial ceramic
membranes range from 500 $/m2 (alumina) to 3000 $/m2 (stainless
steel) [65]. On the other hand, the prices of polymer membranes cost
about 50 $/m2 to 200 $/m2 [66]. Due to their low operating costs,
ceramic membranes are more advantageous and preferred than polymer
membranes under harsh operating circumstances for wastewater treat­
ment. In terms of mechanical, chemical, biological, and thermal prop­
erties, ceramic membranes have lower operating costs and a longer life
duration than polymer membranes. In the current study, our membrane
was elaborated from natural phengite clay, which is cheaper, and it was
sintered at a lower temperature (1050 ◦ C) than commercial alumina
membrane (more than 1200 ◦ C) and membranes made from pure oxides
such as Al2O3, TiO2, and ZrO2. The total preparation cost of the flat
ceramic membrane was estimated based on the unit price of the raw
materials, the cost of electric power consumption, and the cost of
manpower during the whole preparation process. In the present work,
the raw material cost was estimated to be 0.313 $/m2. In addition, the
total preparation cost of the flat ceramic membrane was calculated as
3.5 $/m2. A detailed analysis of the cost estimate for the preparation of
the flat membrane is presented in Table 5. The cost of the developed
responsible for the turbidity and color of the raw real wastwater. Fig. 9b
membrane is significantly lower compared to the cost of conventional
shows the microfiltration mechanism of RWW3 through the optimal
ceramic membranes, due to the appropriate selection of low-cost raw
ceramic membrane. These results show that the microfiltration ceramic
materials (0.313 $/m2). The above cost estimates demonstrate that the
membrane elaborated based on phengite clay is very efficient to treat
ceramic membrane fabricated using phengite clay is very competitive
liquid discharges in terms of suspended matter.
with commercial membranes in terms of cost and is also well-qualified
The physicochemical characteristics of the feed and permeate after 3
for industrial applications.
hours of microfiltration using the optimal ceramic membrane sintered at
1050 ◦ C/2h are summarized in Table 3. The results obtained show that
4. Conclusion
the suspended matter was completely removed. TDS and electrical
conductivity decreased significantly. During the filtration test, the pH of
Flat ceramic microfiltration membranes based on phengite clay were
the permeate remained constant.
manufactured by the paste casting method and then sintered from
For comparison, Table 4 illustrates several properties of some
850 ◦ C to 1150 ◦ C. The optimal membrane sintered at 1050 ◦ C/2h ex­
developed membranes based on different geomaterials, and commercial
hibits satisfactory characteristics, namely, a porosity of 34.5%, an
ceramic membranes based on alumina oxide. We can say that the
average pore diameter of 3.9 μm, a permeability of 43.50 L/h m2 bar, a
optimal ceramic membrane manufactured in this study is characterized
mechanical strength of 26.7 MPa, and excellent chemical stability in

11
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

acidic (4.6%) and basic (0.6%) environments. As a result, the optimal peanut: experiments, equilibrium, thermodynamic, and regeneration studies,
Chem. Afr. 5 (2022) 375–393. https://doi:10.1007/s42250-022-00328-1.
microfiltration membrane demonstrated better-suspended matter
[6] S. Iaich, Y. Miyah, F. Elazhar, S. Lagdali, M. El-Habacha, Low-cost ceramic
removal efficiency (100%) for real wastewater RWW3, very good TDS microfiltration membranes made from Moroccan clay for domestic wastewater and
reduction (71.8%), and electrical conductivity (66.2%). The anti-fouling Congo Red dye treatment, Desalination Water Treat. 235 (2021) 251–271. https
performance was examined by measuring the fouling parameters of the ://doi:10.5004/dwt.2021.27618.
[7] M. Addich, N. El Baraka, A. Laknifli, N. Saffaj, A. Fatni, A. El Hammadi, A.
RWW3 filtration, such as FRR, TFR, RFR, and IFR. The membrane was A. Alrashdi, H. Lgaz, New low-cost tubular ceramic microfiltration membrane
regenerated by washing with demineralized water, and it recovered based on natural sand for tangential urban wastewater treatment, J. Saudi Chem.
84.9% of its flux. The preparation cost of the developed membrane is Soc. 26 (2022), 101512. https://doi:10.1016/j.jscs.2022.101512.
[8] J. Saikia, S. Sarmah, J.J. Bora, B. Das, R.L. Goswamee, Preparation and
approximately 3.5 $/m2 due to the use of low-cost raw materials, costing characterization of low cost flat ceramic membranes from easily available potters’
about 0.313 $/m2. The prepared low-cost membrane could be used as a clay for dye separation, Bull. Mater. Sci. 42 (2019) 104. https://doi:10.1007/s120
potential candidate in several industrial microfiltration applications. 34-019-1767-7.
[9] Y. Wang, B. Ma, M. Ulbricht, Y. Dong, X. Zhao, Progress in alumina ceramic
membranes for water purification: status and prospects, Water Res. 226 (2022),
Authors contributions 119173. https://doi:10.1016/j.watres.2022.119173.
[10] A. Tahiri, L. Messaoudi, N. Tijani, M.H. Zerrouk, M. Messaoudi, Manufacture and
characterization of flat membrane supports based on Moroccan Rif clay, Mater.
S. Lagdali: Writing – original draft, data analysis, review and edit­ Today Proc. 43 (2021) 209–215. https://doi:10.1016/j.matpr.2020.11.638.
ing, Y. Miyah: Review and editing, M. El-Habacha: Review and editing, [11] B. Jafari, M. Abbasi, S.A. Hashemifard, M. Sillanpää, Elaboration and
G. Mahmoudy: Review and editing, M. Benjelloun: Review and edit­ characterization of novel two-layer tubular ceramic membranes by coating natural
zeolite and activated carbon on mullite-alumina-zeolite support: application for
ing, S. Iaich: Co-supervision, conceptualization, data analysis, review, oily wastewater treatment, J. Asian Ceram. Soc. 8 (2020) 848–861. https://doi:1
editing, and validation, M. Zerbet: Supervision, review and editing, M. 0.1080/21870764.2020.1793475.
Chiban: Review and editing, F. Sinan: Review and editing. [12] D. Naseer, J.H. Ha, J. Lee, I.H. Song, Preparation of Al2O3 multichannel cylindrical-
tube-type microfiltration membrane with surface modification, Appl. Sci. 12
(2022) 7993. https://doi:10.3390/app12167993.
Funding [13] M. Zielińska, M. Galik, Use of ceramic membranes in a membrane filtration
supported by coagulation for the treatment of dairy wastewater, Water Air Soil
Pollut. 228 (2017) 173. https://doi:10.1007/s11270-017-3365-x.
This research received no external funding. [14] A. Belgada, B. Achiou, S.A. Younssi, F.Z. Charik, M. Ouammou, J.A. Cody,
R. Benhida, K. Khaless, Low-cost ceramic microfiltration membrane made from
natural phosphate for pretreatment of raw seawater for desalination, J. Eur. Ceram.
Declaration of competing interest Soc. 41 (2021) 1613–1621. https://doi:10.1016/j.jeurceramsoc.2020.09.064.
[15] H.T. Lay, J.W. Chew, Critical flux of colloidal foulant in microfiltration: effect of
organic solvent, J. Membr. Sci. 616 (2020), 118531. https://doi:10.1016/j.memsci
We declare; Salek Lagdali, Youssef Miyah, Mohamed El-Habacha, .2020.118531.
Guellaa Mahmoudy, Mohammed Benjelloun, Soulaiman Iaich, [16] D. Rashad, S.K. Amin, M.S. Mansour, H. Abdallah, Fabrication of low-cost
Mohamed Zerbet, Mohamed Chiban, Fouad Sinan, authors of the orig­ antibacterial microfiltration tubular ceramic membranes, Ceram. Int. 48 (2022)
11489–11501. https://doi:10.1016/j.ceramint.2022.01.005.
inal article entitled: “Performance assessment of a phengite clay-based [17] P. Singh, N.A. Manikandan, M. Purnima, K. Pakshirajan, G. Pugazhenthi, Recovery
flat membrane for microfiltration of real-wastewater from clothes of lignin from water and methanol using low-cost kaolin based tubular ceramic
washing: characterization, cost estimation, and regeneration”, that this membrane, J. Water Process Eng. 38 (2020), 101615. https://doi:10.1016/j.
jwpe.2020.101615.
article has not been published in any other journal nor is it currently [18] P.J. Sánchez-Soto, E. Garzón, L. Pérez-Villarejo, D. Eliche-Quesada, Sintering
under consideration for publication elsewhere. All authors agree with behaviour of a clay containing pyrophyllite, sericite and kaolinite as ceramic raw
the actual form of the manuscript. We have no conflits of interest to materials: looking for the optimum firing conditions, Bol. Soc. Esp. Cerámica Vidr.
62 (2023) 26–39. https://doi:10.1016/j.bsecv.2021.09.001.
disclose.
[19] R. Chihi, I. Blidi, M. Trabelsi-Ayadi, F. Ayari, Elaboration and characterization of a
low-cost porous ceramic support from natural Tunisian bentonite clay, Compt.
Data availability Rendus Chem. 22 (2019) 188–197. https://doi:10.1016/j.crci.2018.12.002.
[20] M. Mouiya, A. Abourriche, A. Bouazizi, A. Benhammou, Y. El Hafiane,
Y. Abouliatim, L. Nibou, M. Oumam, M. Ouammou, A. Smith, H. Hannache, Flat
No data was used for the research described in the article. ceramic microfiltration membrane based on natural clay and Moroccan phosphate
for desalination and industrial wastewater treatment, Desalination 427 (2018)
42–50. https://doi:10.1016/j.desal.2017.11.005.
Appendix A. Supplementary data
[21] A. Harrati, Y. Arkame, A. Manni, S. Aqdim, R. Zmemla, A. Chari, A. El Bouari, I.
E. El Hassani, A. Sdiri, F.O. Hassani, C. Sadik, Akermanite-based ceramics from
Supplementary data to this article can be found online at https://doi. Moroccan dolomite and perlite: characterization and in vitro bioactivity
assessment, Open Ceram. 10 (2022), 100276. https://doi:10.1016/j.oceram.20
org/10.1016/j.cscee.2023.100388.
22.100276.
[22] A. Dhivya, A. Keshav, Fabrication of ball clay based low-cost ceramic membrane
References supports and their characterization for microfiltration application, J. Indian Chem.
Soc. 99 (2022), 100557. https://doi:10.1016/j.jics.2022.100557.
[23] S.L.S. Rani, R.V. Kumar, Fabrication and characterization of ceramic membranes
[1] M. Mouiya, A. Bouazizi, A. Abourriche, A. Benhammou, Y. El Hafiane,
derived from inexpensive raw material fuller’s earth clay, Mater. Sci. Eng. B 284
M. Ouammou, Y. Abouliatim, S.A. Younssi, A. Smith, H. Hannache, Fabrication and
(2022), 115877. https://doi:10.1016/j.mseb.2022.115877.
characterization of a ceramic membrane from clay and banana peel powder:
[24] S.B. Rekik, J. Bouaziz, A. Deratani, S. Beklouti, Study of ceramic membrane from
application to industrial wastewater treatment, Mater. Chem. Phys. 227 (2019)
naturally occurring-kaolin clays for microfiltration applications, Period. Polytech. -
291–301. https://doi:10.1016/j.matchemphys.2019.02.011.
Chem. Eng. 61 (2017) 206. https://doi:10.3311/PPch.9679.
[2] M. Addich, N. El Barakaa, N. Saffaj, A. Laknifli, A. Karim, K. Sbihi, A. El Hammadi,
[25] A. Agarwalla, K. Mohanty, Comprehensive characterization, development, and
Elaboration of innovative ceramic microfiltration membrane from natural
application of natural/Assam Kaolin-based ceramic microfiltration membrane,
Moroccan sand for wastewater treatment, Desalination Water Treat. 260 (2022)
Mater. Today Chem. 23 (2022), 100649. https://doi: 10.1016/j.mtchem.2021.100
299–308. https://doi:10.5004/dwt.2022.28550.
649.
[3] Y. Miyah, M. Benjelloun, A. Lahrichi, F. Mejbar, S. Iaich, G. El Mouhri, R. Kachkoul,
[26] B. Achiou, H. Elomari, A. Bouazizi, A. Karim, M. Ouammou, A. Albizane,
F. Zerrouq, Highly-efficient treated oil shale ash adsorbent for toxic dyes removal:
J. Bennazha, S.A. Younssi, I.E. El Amrani, Manufacturing of tubular ceramic
kinetics, isotherms, regeneration, cost analysis and optimization by experimental
microfiltration membrane based on natural pozzolan for pretreatment of seawater
design, J. Environ. Chem. Eng. 9 (2021), 106694. https://doi:10.1016/j.jece.20
desalination, Desalination 419 (2017) 181–187. https://doi:10.1016/j.desal.20
21.106694.
17.06.014.
[4] Y. Miyah, M. Benjelloun, R. Salim, L. Nahali, F. Mejbar, A. Lahrichi, S. Iaich,
[27] S. Saja, A. Bouazizi, B. Achiou, M. Ouammou, A. Albizane, J. Bennazha, S.
F. Zerrouq, Experimental and DFT theoretical study for understanding the
A. Younssi, Elaboration and characterization of low-cost ceramic membrane made
adsorption mechanism of toxic dye onto innovative material Fb-HAp based on
from natural Moroccan perlite for treatment of industrial wastewater, J. Environ.
fishbone powder, J. Mol. Liq. 362 (2022), 119739. https://doi:10.1016/j.molliq.20
Chem. Eng. 6 (2018) 451–458. https://doi:10.1016/j.jece.2017.12.004.
22.119739.
[28] A. Manni, B. Achiou, A. Karim, A. Harrati, C. Sadik, M. Ouammou, S.A. Younssi,
[5] M. Benjelloun, Y. Miyah, R. Bouslamti, L. Nahali, F. Mejbar, S. Lairini, The fast-
A. El Bouari, New low-cost ceramic microfiltration membrane made from natural
efficient adsorption process of the toxic dye onto shells powders of walnut and

12
S. Lagdali et al. Case Studies in Chemical and Environmental Engineering 8 (2023) 100388

magnesite for industrial wastewater treatment, J. Environ. Chem. Eng. 8 (2020), [46] J.P. Greenwood, P.C. Hess, Congruent melting kinetics of albite: theory and
103906. https://doi:10.1016/j.jece.2020.103906. experiment, J. Geophys. Res. Solid Earth 103 (1998) 29815–29828. https://do
[29] S. Khemakhem, R. Ben Amar, A. Larbot, Synthesis and characterization of a new i:10.1029/98JB02300.
inorganic ultrafiltration membrane composed entirely of Tunisian natural illite [47] D. Vasanth, R. Uppaluri, G. Pugazhenthi, Influence of sintering temperature on the
clay, Desalination 206 (2007) 210–214. https://doi:10.1016/j.desal.2006.03.567. properties of porous ceramic support prepared by uniaxial dry compaction method
[30] N. Ahmed, F.Q. Mir, Fabrication of a cost effective ceramic microfiltration using low-cost raw materials for membrane applications, Separ. Sci. Technol. 46
membrane by utilizing local kashmir clay, Trans. Indian Ceram. Soc. 80 (2021) (2011) 1241–1249. https://doi:10.1080/01496395.2011.556097.
41–46. https://doi:10.1080/0371750X.2020.1864663. [48] A. Elgamouz, N. Tijani, I. Shehadi, K. Hasan, M.A.F. Kawam, Characterization of
[31] M.F. Twibi, M.H.D. Othman, S.K. Hubadillah, S.A. Alftessi, M.R.B. Adam, A. the firing behaviour of an illite-kaolinite clay mineral and its potential use as
F. Ismail, M.A. Rahman, J. Jaafar, Y.O. Raji, M.H.A. Aziz, M.N.B.M. Sokri, membrane support, Heliyon 5 (2019), e02281. https://doi:10.1016/j.heliyon.20
H. Abdullah, R. Naim, Hydrophobic mullite ceramic hollow fibre membrane (Hy- 19.e02281.
MHFM) for seawater desalination via direct contact membrane distillation [49] H. Ouallal, M. Azrour, M. Messaoudi, H. Moussout, L. Messaoudi, N. Tijani,
(DCMD), J. Eur. Ceram. Soc. 41 (2021) 6578–6585. https://doi:10.1016/j.jeurc Incorporation effect of olive pomace on the properties of tubular membranes,
eramsoc.2021.06.024. J. Environ. Chem. Eng. 8 (2020), 103668. https://doi:10.1016/j.jece.2020.10366
[32] B. Hatimi, J. Mouldar, A. Loudiki, H. Hafdi, M. Joudi, E.M. Daoudi, H. Nasrellah, I. 8.
T. Lançar, M.A. El Mhammedi, M. Bakasse, Low cost pyrrhotite ash/clay-based [50] N. Saffaj, M. Persin, S.A. Younsi, A. Albizane, M. Cretin, A. Larbot, Elaboration and
inorganic membrane for industrial wastewaters treatment, J. Environ. Chem. Eng. characterization of microfiltration and ultrafiltration membranes deposited on raw
8 (2020), 103646. https://doi:10.1016/j.jece.2019.103646. support prepared from natural Moroccan clay: application to filtration of solution
[33] T.M. Zewdie, I. Prihatiningtyas, A. Dutta, N.G. Habtu, B.V.D. Bruggen, containing dyes and salts, Appl. Clay Sci. 31 (2006) 110–119. https://doi:10.1016
Characterization and beneficiation of Ethiopian kaolin for use in fabrication of /j.clay.2005.07.002.
ceramic membrane, Mater. Res. Express 8 (2021), 115201. https://doi:10.1088/ [51] B.K. Nandi, R. Uppaluri, M.K. Purkait, Preparation and characterization of low cost
2053-1591/ac2f75. ceramic membranes for micro-filtration applications, Appl. Clay Sci. 42 (2008)
[34] M. Mouafon, G.L. Lecomte-Nana, N. Tessier-Doyen, A. Njoya, D. Njoya, 102–110. https://doi:10.1016/j.clay.2007.12.001.
D. Njopwouo, Processing and characterization of low-thermal conductivity, clay- [52] C. Sadik, I.E.E. Amrani, A. Albizane, Processing and characterization of
based, ceramic membranes for filtering drinking water, Clay Clay Miner. 69 (2021) alumina–mullite ceramics, J. Asian Ceram. Soc. 2 (2014) 310–316. https://doi:10
339–353. https://doi:10.1007/s42860-021-00131-y. .1016/j.jascer.2014.07.006.
[35] R. Mouratib, B. Achiou, M.E. Krati, S.A. Younssi, S. Tahiri, Low-cost ceramic [53] M.M. Bazin, N. Ahmad, Y. Nakamura, Preparation of porous ceramic membranes
membrane made from alumina- and silica-rich water treatment sludge and its from Sayong ball clay, J. Asian Ceram. Soc. 7 (2019) 417–425. https://doi:10.1
application to wastewater filtration, J. Eur. Ceram. Soc. 40 (2020), 5942-5950, 080/21870764.2019.1658339.
https://doi:10.1016/j.jeurceramsoc.2020.07.050. [54] I. Jedidi, S. Khemakhem, S. Saïdi, A. Larbot, N. Elloumi-Ammar, A. Fourati,
[36] K.V.V. Satyannarayana, S.L.S. Rani, S. Baranidharan, R.V. Kumar, Indigenous A. Charfi, A. Ben Salah, R. Ben Amar, Preparation of a new ceramic microfiltration
bentonite based tubular ceramic microfiltration membrane: elaboration, membrane from mineral coal fly ash: application to the treatment of the textile
characterization, and evaluation of environmental impacts using life cycle dying effluents, Powder Technol. 208 (2011) 427–432. https://doi:10.1016/j.
techniques, Ceram. Int. 48 (2022) 28843–28855. https://doi:10.1016/j.ceramint powtec.2010.08.039.
.2022.03.156. [55] A. Harabi, A. Guechi, S. Condom, Production of supports and filtration membranes
[37] A. Lahnafi, A. Elgamouz, N. Tijani, L. Jaber, A.N. Kawde, Hydrothermal synthesis from Algerian kaolin and limestone, Procedia Eng. 33 (2012) 220–224. https:
and electrochemical characterization of novel zeolite membranes supported on flat //doi:10.1016/j.proeng.2012.01.1197.
porous clay-based microfiltration system and its application of heavy metals [56] H. Elomari, B. Achiou, D. Beqqour, K. Khaless, R. Beniazza, M. Ouammou,
removal of synthetic wastewaters, Microporous Mesoporous Mater. 334 (2022), A. Aaddane, S.A. Younssi, R. Benhida, Preparation and characterization of low-cost
111778. https://doi:10.1016/j.micromeso.2022.111778. zirconia/clay membrane for removal of acid orange 74 dye, Mater. Today Proc. 51
[38] D. Beqqour, B. Achiou, A. Bouazizi, H. Ouaddari, H. Elomari, M. Ouammou, (2022) 1948–1956. https://doi:10.1016/j.matpr.2021.03.674.
J. Bennazha, S.A. Younssi, Enhancement of microfiltration performances of [57] M. Messaoudi, M. Douma, N. Tijani, L. Messaoudi, Study of the permeability of
pozzolan membrane by incorporation of micronized phosphate and its application tubular mineral membranes: application to wastewater treatment, Heliyon 7
for industrial wastewater treatment, J. Environ. Chem. Eng. 7 (2019), 102981. (2021), e06837. https://doi10.1016/j.heliyon.2021.e06837.
https://doi:10.1016/j.jece.2019.102981. [58] Y. Dong, X. Feng, D. Dong, S. Wang, J. Yang, J. Gao, X. Liu, G. Meng, Elaboration
[39] B. Achiou, H. Elomari, M. Ouammou, A. Albizane, J. Bennazha, A. Aaddane, S. and chemical corrosion resistance of tubular macro-porous cordierite ceramic
A. Younssi, I.E. El Hassani, Study of added starch on characteristics of flat ceramic membrane supports, J. Membr. Sci. 304 (2007) 65–75. https://doi:10.1016/j.
microfiltration membrane made from natural Moroccan pozzolan, J. Mater. memsci.2007.06.058.
Environ. Sci. 9 (2018) 1013–1021. https://doi:10.26872/jmes.2017.9.3.113. [59] W. Zhou, P. Wu, L. Zhang, D. Zhu, X. Zhao, Y. Cai, Fabrication and performance of
[40] S. Foorginezhad, M.M. Zerafat, Y. Mohammadi, M. Asadnia, Fabrication of tubular reticular ceramic fiber membranes by freeze casting using a gel network, J. Eur.
ceramic membranes as low-cost adsorbent using natural clay for heavy metals Ceram. Soc. 41 (2021) 6586–6595. https://doi:10.1016/j.jeurceramsoc.2021.06.0
removal, Clean. Eng. Technol. 10 (2022), 100550. https://doi:10.1016/j.clet.20 23.
22.100550. [60] S. Iaich, Y. Miyah, L. Messaoudi, Elaboration and characterization of low cost
[41] A. Talidi, A. Bouazizi, A. Essate, A. Karim, A. Aaddane, M. Ouammou, R. Saadani, tubular ceramic supports made of Moroccan clay for microfiltration and
S.A. Younssi, Manufacture and characterization of flat microfiltration membrane ultrafiltration membranes, Moroc. J. Chem. 9 (2021) 185–197. https://doi:
based on Moroccan pyrophyllite clay for pretreatment of raw seawater for 10.48317/IMIST.PRSM/morjchem-v9i2.22981.
desalination, Desalination Water Treat. 253 (2022) 24–38. https://doi:10. [61] K.P. Goswami, G. Pugazhenthi, Treatment of poultry slaughterhouse wastewater
5004/dwt.2022.28260. using tubular microfiltration membrane with fly ash as key precursor, J. Water
[42] N. El Qacimi, N. El Baraka, N. Saffaj, R. Mamouni, A. Laknifli, S.A. Younssi, Process Eng. 37 (2020), 101361. https://doi:10.1016/j.jwpe.2020.101361.
A. Faouzi, H. Zidouh, Preparation and characterization of flat membrane support [62] P. Mittal, S. Jana, K. Mohanty, Synthesis of low-cost hydrophilic
based on Sahara Moroccan clay: application to the filtration of textile effluents, ceramic–polymeric composite membrane for treatment of oily wastewater,
Desalination Water Treat. 143 (2019) 111–117. https://doi:10.5004/dwt.2019.2 Desalination 282 (2011) 54–62. https://doi:10.1016/j.desal.2011.06.071.
3516. [63] F. Fatima, H. Du, R.R. Kommalapati, Treatment of poultry slaughterhouse
[43] A. Nazir, K. Khan, A. Maan, R. Zia, L. Giorno, K. Schroën, Membrane separation wastewater with membrane technologies: a review, Water 13 (2021) 1905. https://
technology for the recovery of nutraceuticals from food industrial streams, Trends doi: 10.3390/w13141905.
Food Sci. Technol. 86 (2019) 426–438. https://doi:10.1016/j.tifs.2019.02.049. [64] A. Majouli, S. Tahiri, S.A. Younssi, H. Loukili, A. Albizane, Elaboration of new
[44] M. Azimifar, M. Ghorbani, M. Peyravi, Fabrication and evaluation of a tubular ceramic membrane from local Moroccan Perlite for microfiltration process.
photocatalytic membrane based on Sb2O3/CBO composite for improvement of dye Application to treatment of industrial wastewaters, Ceram. Int. 38 (2012)
removal efficiency, J. Mol. Struct. 1270 (2022), 133957. https://doi:10.1016/j. 4295–4303. https://doi:10.1016/j.ceramint.2012.02.010.
molstruc.2022.133957. [65] D. Vasanth, G. Pugazhenthi, R. Uppaluri, Biomass assisted microfiltration of
[45] M. EL–Habacha, A. Dabagh, S. Lagdali, Y. Miyah, G. Mahmoudy, F. Sinan, chromium(VI) using Baker’s yeast by ceramic membrane prepared from low cost
M. Chiban, S. Iaich, M. Zerbet, An efficient and adsorption of methylene blue dye raw materials, Desalination 285 (2012) 239–244. https://doi:10.1016/j.desal.20
on a natural clay surface: modeling and equilibrium studies, Environ. Sci. Pollut. 11.09.055.
Res. (2023). https://doi:10.1007/s11356-023-27413-3. [66] B.K. Nandi, B. Das, R. Uppaluri, M.K. Purkait, Microfiltration of mosambi juice
using low cost ceramic membrane, J. Food Eng. 95 (2009) 597–605. https://doi:10
.1016/j.jfoodeng.2009.06.024.

13

You might also like