Enhancement of Reverse Osmosis Membranes For Groundwa 2023 Environmental Nan

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

Contents lists available at ScienceDirect

Environmental Nanotechnology, Monitoring & Management


journal homepage: www.elsevier.com/locate/enmm

Enhancement of reverse osmosis membranes for groundwater purification


using cellulose acetate incorporated with ultrathin graphitic carbon
nitride nanosheets
Eman S. Mansor a, *, H. Abdallah b, M.S. Shalaby b, A.M. Shaban a
a
Water Pollution Research Department, Environment and Climate Change Research Institute, National Research Centre, 33 El-Bohouth St. (Former El-Tahrir St.), Dokki,
Cairo, Egypt
b
Chemical Engineering Department, Engineering and Renewable Energy Research Institute, National Research Centre, 33 El-Bohouth St. (Former El-Tahrir St.), Dokki,
Cairo PO Box 12622, Egypt

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

Keywords: Modified reverse osmosis (RO) membranes are emerging in the fields of water desalination and purification as
Ultrathin graphitic carbon nitride next-generation separation techniques. Highly efficient graphitic carbon nitride (g-C3N4) RO membranes were
Cellulose acetate fabricated by incorporating g-C3N4 into polyvinyl alcohol-based cellulose acetate. X-ray diffraction, scanning
Polyvinyl alcohol
electron microscopy, Fourier transform infrared spectroscopy, and transmission electron microscopy were
Desalination
employed to determine the crystal structure and textural qualities of the ultrathin graphitic layers. The interlayer
spacing of the g-C3N4 in the modified membranes was relatively narrow, which significantly impacted membrane
performance. Phase inversion was employed to fabricate uniform and compact g-C3N4 hybrid membranes. The
impact of the g-C3N4 loading quantity in the polymeric solution on viscosity and membrane performance was
investigated. An optimized concentration of 0.5 wt% for the ultrathin g-C3N4 resulted in a distinguished dense
top layer. The fabricated RO membranes were evaluated using a synthetic saline solution (2000 ppm of NaCl),
resulting in a percentage rejection of 95% at a reasonable operating pressure of 15 bar to exhibit 10.5 LMH. The
fabricated membranes exhibited self-cleaning properties after soaking in malachite green dye (10 mgL− l); this
demonstrated the antifouling properties of the membrane. The g-C3N4 membrane was successfully applied in a
groundwater purification process and its efficiency was observed to reach approximately 97%. This membrane is
promising for groundwater treatment.

(Huang et al., 2014).


1. Introduction Melamine-based membranes have demonstrated exceptional thermal
and chlorine resistance, owing to the difference in the chemical struc­
Water scarcity is a critical issue that affects many people worldwide ture of the triazine ring. Additionally, they can reduce toxicity,
(Shannon et al., 2008). Furthermore, water scarcity and contamination compared with m-phenylenediamine (MPD) and trimesoylchloride
are becoming increasingly severe, culminating in freshwater scarcity (TMC), which are the monomers for the common method to fabricate
(Pendergast and Hoek, 2011). Many approaches, such as photocatalysis, high-flux reverse osmosis (RO) membranes in a complex process (Han,
adsorption, and membrane separation, have been used to alleviate water 2013).
problems. Many significant advances involve chemicals and necessitate G-C3N4, as a carbon material, is produced from the polymerization of
a significant influx of finance, thereby impeding their practical imple­ melamine and can be applied as a photosensitizer, in hydrogen storage,
mentation (Shannon et al., 2008). and as a drug nanocarrier for cancer imaging and therapy (Bu et al.,
Membrane separation is a promising physical strategy that is widely 2013; Lin et al., 2014; Wu et al., 2013; Zhang and Antonietti, 2010;
employed in water treatment and other sectors, including chemical, Zhang et al., 2013; Zhang et al., 2010).
biological, and medicinal fields (Jackson and Hillmyer, 2010; Marchetti RO membranes can be modified through functionalization by the
et al., 2014; Van Reis and Zydney, 2001). The limitations of polymeric carboxyl reaction, which enhances the desalination performance (Zhu
membranes are their ability to withstand high pressure and temperature et al., 2014). A reasonable modification depends on the mode of

* Corresponding author.
E-mail address: eman_mansor31@yahoo.com (E.S. Mansor).

https://doi.org/10.1016/j.enmm.2022.100760
Received 11 June 2022; Received in revised form 26 October 2022; Accepted 14 November 2022
Available online 21 November 2022
2215-1532/© 2022 Elsevier B.V. All rights reserved.
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

Nomenclature Jw1 Initial water flux


Jw2 Permeate flux after cleaning with water.
List of abbreviations and symbols g-C3N4 Graphitic carbon nitride
RO Reverse osmosis SEM Scanning electron microscopy
XRD X-ray diffraction TEM Transmission electron microscopy
FTIR Fourier transform infrared TMC Trimesoylchloride
MPD m-phenylenediamine MW Molecular weight
CA Cellulose acetate DMAC Dimethyl acetamide
PVA Polyvinyl alcohol BTCA 1,2,3,4-butanetetracarboxylic acid
BSA Bovine serum albumin GO Graphene oxide
DW Distilled water NPs Nanoparticles
NaCl Sodium chloride MG Malachite green
TFC Thin film composite σ br Stress at break
εbr Elongation at break ε Porosity
Ww Mass of wet membranes PH Water density
Wd Mass of dry membranes L Thickness (m)
A Membrane area (m2) A Film area (m2)
Q Amount of product water (L) CP Permeate concentration
T Collecting time (h) FRR Flux recovery ratio
CF Feed concentration Rir Irreversible fouling
Rr Reversible fouling DW Deionized water
Rt Total fouling ratio JP Permeate flux during the filtration process

reducing the interlayer spacing and swelling, such as through covalent


crosslinking (Jia et al., 2016; Yuan et al., 2017) and using materials with Table 1
narrow transport layers (Lia et al., 2018). Carbon-based membranes Mechanical properties of the used nonwoven fabric for casting the fabricated
exhibit electrostatic interaction on the membrane surface, and the membranes.
presence of charges enhances the desalination process. Property Nonwoven fabric
The introduction of various nanoparticles (NPs) and additives into
cellulose acetate (CA) membranes enhances certain properties. How­ fabric Area density g/m2 77.6
ever, challenges still hinder the attainment of maximum selectivity and Thickness Mm 0.1
permeability of the membranes. Polyvinyl alcohol (PVA) is an attractive Tensile strength (Machine direction) N/100 mm 194
polymer for synthesizing asymmetric porous membranes. It has low Tensile strength (Cross direction) N/25 mm 100
Elongation at break (Machine direction) % 1.6
fouling potential and is highly hydrophilic with high mechanical
Elongation at break (Cross direction) % 2
strength. In addition, it exhibits film-forming biocompatible properties
and a highly stable pH (Yin et al., 2016). PVA can be used as a secondary
polymer or low-molecular-weight additive to control membrane struc­
ture (Galiano et al., 2018). Table 2
Polymeric solution composition of the prepared unmodified and modified
Here, modified RO membranes were fabricated by blending CA as the
membranes.
main base polymer and PVA embedded with g-C3N4 to induce excellent
membrane selectivity and permeability. The surface of the fabricated g- Membrane Polymer composition Solvent Wt.% Non-solvent Wt.%
Wt.%
C3N4 RO membrane had negative charges. The fabricated membranes
were characterized and tested with saline water. The preparation CA PVA of NPs DMAC H2O
method was simple and environmentally friendly. CA-PVA 24 4 0 76 10
CA-PVA-0.1C3N4 24 4 0.1 75.9 10
2. Materials and methods CA-PVA − 0.3 C3N4 24 4 0.3 75.7 10
CA-PVA- 0.5 C3N4 24 4 0.5 75.5 10
CA-PVA − 0.7 C3N4 24 4 0.7 75.3 10
2.1. Materials

CA (MW100000) and PVA (MW90000) were provided by Sigma 2.3. Fabrication of RO membranes
Aldrich. Dimethyl acetamide (DMAC) and isopropyl alcohol were pur­
chased from Roth Specialities India. Bovine serum albumin (BSA) and The membranes were fabricated by phase inversion. The g-C3N4
melamine were purchased from Sigma Aldrich. As a crosslinker, 1,2,3,4- nanosheets were dispersed in DMAC under sonication for 1 h at 50 kHz.
butanetetracarboxylic acid (BTCA) was obtained from Merck, Germany This mixture was added to a polymeric solution comprising the main
(purity > 99). polymer (CA, 24 wt%) and PVA (4 wt%). The composition of the poly­
meric solution is presented in Table 1. The mixture was stirred for 12 h
2.2. Preparation of ultrathin g-C3N4 nanosheets and cast on nonwoven polyester with a thickness of 200 μm. Thereafter,
the sheet was instantly dipped into a deionized water (DW) coagulating
Bulk C3N4 was prepared by the thermal decomposition of melamine bath to form the membrane with annealing treatment. The variation in
at 600 ◦ C for 2 h. The sample was cooled and ground then sonicated in the mechanical properties of the membrane due to the nearness of the
isopropyl alcohol at 50 kHz for 12 h. NPs was compared with the good quality of the nonwoven textures,
where it was considered irrelevant since it was extremely small. The
mechanical properties of the nonwoven textures are listed in Table 2.

2
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

2.4. Characterization

Scanning electron microscopy (SEM; QUANTA FEG-250) and high-


resolution transmission electron microscopy (HRTEM; JOEL) were
employed to study the membrane morphologies. X-ray diffraction (XRD;
Philips powder diffractometer) was performed to determine the NP
structure. Attenuated total reflectance-Fourier transform infrared (ATR-
FTIR) spectroscopy was performed using a Bruker VERTEX 80 (Ger­
many) combined with Platinum Diamond ATR, comprising a diamond
disk, similar to that of an internal reflector in the range of 4000–400
cm− 1 with a resolution of 4 cm− 1 and refractive index of 2.4. The vis­
cosity of the polymeric matrix was tested using a Brookfield Ametek
DV2T viscometer with spindle entry code No. 27.
Mechanical testing was performed on the prepared membranes,
where the stress at break (σ) and elongation at break (ε br) were recorded
on samples with a length and width of 100 and 25 mm, respectively,
using an H5KS universal tensile testing machine.
The membrane porosity (ε) was calculated using a gravimetric
method, as shown in Eq. (1):
Fig. 1. XRD pattern for the prepared graphitic carbon nitride g-C3N4.
( )
Ww− wd
ε(%) = 100 (1)
ALPH

where Ww and Wd are the masses of the wet and dry membranes (g),
respectively. PH is the water density (0.998 g/cm3), and A and L are the
membrane area (m2) and thickness (m), respectively.

2.5. Water flux and fouling resistance

Membrane productivity was examined using a dead-end setup, with


an active exposed film area of 192 cm2 under a pressure of 15 bars for
120 min. Next, the DW flux was measured using Eq. (2) (Jamil et al.,
2018):
Q
Flux = (2)
AT
Q, A, and T are the amounts of product water (L), film area (m2), and
collecting time (h), respectively.
Salt separation efficiency was obtained using Eq. (3):
( ( ))
CP
Removl(%) = 1 − 100 (3)
CF
CP and CF are the salty water concentrations (mgL− 1) in the product
and feed, respectively, which were specified using a 4510-Jenway set for
electrical conductivity.
The evaluation of the membrane fouling was performed using BSA as
a model foulant. First, 1000-ppm BSA was added to a 2000-ppm NaCl
feed solution. The test was run for 20 h, and the samples were retrieved
after each hour. Thereafter, the fouled membrane was cleaned for 15
min using DW. This process was repeated four times with the sequence
of flow of DW through the membrane first, to determine Jw1. Afterward,
the BSA solution was fed through the membrane, and the flux (JP) was
recorded. The surface was washed using DW, and the DW flux (Jw2) was
recorded.
The total fouling rate (Rt) and fouling recovery rate (FRR) were
studied using the following equation:
Jw2
FRR = (4) Fig. 2. The textural properties of the prepared g-C3N4 (a) SEM images, (b) TEM
Jw1
images for the prepared ultrathin layers.
The reversible (Rr), irreversible (Rir), and total fouling ratio (Rt) were
calculated using the following equations:
JP
Rt = 1 − × 100 (5)
Jw1

Jw2 − Jp
Rr = × 100 (6)
Jw1

3
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

Fig. 4. Effect of embedding ratio of g-C3N4 nanosheets on the viscosity of the


polymeric solutions.
Fig. 3. Chemical structure in terms of FT-IR spectra.

viscosity. The optimal concentration was determined through studies


Rir =
Jw1 − Jw2
× 100 (7) when the polymer chain entanglement was expected (Liu and Li, 2017;
Jw1 Wood and Sourirajan, 1993). The use of the ideal polymer is critical to
achieving the best polymer concentration in a polymeric combination.
3. Results and discussion As shown in Fig. 4, the addition of g-C3N4 to the CA/PVA matrix
considerably increased the viscosity of the CA/PVA casting solution.
3.1. Characterization of the prepared g-C3N4 Considering that a higher viscosity causes a slower demixing of the
casting solution, this can have a significant impact on modifying the
XRD was employed to determine the phase and composition of the g- porosity of the CA/PVA membrane.
C3N4 sample. As shown in Fig. 1, the stacking of the conjugated aromatic The two straight lines based on the slope of the viscosity curve
system or the pure g-C3N4 powder exhibited a peak at 27.5◦ , corre­ indicated high viscosity (Ismail et al., 2014; Peng et al., 2014). The
sponding to the (0 0 2) arising plane (Zhu et al., 2018). critical concentration was calculated using the intersection of the two
The SEM image of the synthesized g-C3N4 powder is illustrated in straight lines. Here the significant embedding of NPs was 0.5% (Fig. 4).
Fig. 2. The g-C3N4 powder exhibited ultrathin nanosheet layers, which Ghaseminezhad et al. (2019) reported that the addition of graphene
were considered distinguished layers (Yu et al., 2017). However, the oxide (GO) to a CA solution resulted in a 1- wt% increase in top layer
HRTEM images provided further insight into the internal structure of g- thickness from 2 to 6 nm. However, the high viscosity of CA/GO was due
C3N4 and confirmed the nano size of the thin layers.Fig. 3. to agglomeration because of the use of 2 wt% of GO.

3.2. Chemical structure in terms of FTIR spectra


3.4. SEM images of the fabricated membranes

The stretching vibration mode (C–N) was observed at an absorption


Fig. 5 shows the field emission (FE)-SEM images for the nano­
band of 1616 cm− 1, and the stretching mode of the C–N heterocycle
composite membranes of 0.5- wt% and 0.7- wt% C3N4/CA/PVA. The
was observed at four absorption peaks at 1531, 1441, 1336, and 1260
fabricated membranes were cut using liquid nitrogen. The SEM images
cm− 1. In addition, the out-of-plane ring bending modes of the C–N
showed that the membranes exhibited an asymmetric structure. The
heterocycles were responsible for the absorption band at 809 cm− 1,
cross-section of the 0.5- wt% C3N4/CA/PVA membranes did not indicate
whereas the O–H and N–H stretching modes were responsible for the
agglomeration because of the uniform dispersion of C3N4. Increasing the
absorption band at 3185 cm− 1 (Arzhandi et al., 2018).
C3N4 content to 0.7 wt% induced macrovoids in the membrane. How­
The bands at 910 and 835 cm− 1 indicated the occurrence of PVA
ever, using an amount<0.7 wt% led to a reduction in the number and
skeletal vibrations, which have been reported (Pawde and Deshmukh,
size of macrovoids, as reported by Majeed et al. (2012). To investigate
2008; Pawde et al., 2008). Antisymmetric and symmetric vibrations
how the casting affected the membrane structure, the solidification of
produced two separate absorption bands at 2960 and 2910 cm− 1,
the membrane needed to be understood. The solidification began, owing
respectively. Certain bands at 1750 and 1660 cm− 1 were attributable to
to the low miscibility between the polymeric solutions (CA–PVA in the
CA absorption and were similar to the C– – O stretching vibration of the
organic solvent) and the nonsolvent (DW). This miscibility caused the
carbonyl group (Krimm et al., 1956). In addition to the peak at 808
exchange of the organic solvent and DW in the top layer and sublayer of
cm− 1, the absorption band at 1619 cm− 1 in the membrane conformed to
the casted membrane. Thus, the nuclei of the polymer-poor phase were
the typical stretching vibration mode (C–N), indicating the interaction
formed and continued to increase to reach the maximum polymer con­
between CA–PVA and g-C3N4 (Fig. 1c). A large absorption band centered
centration at the pore–solution interface (Saljoughi, 2009). A delay
within the range of 3050–3700 cm− 1 was observed and attributed to the
observed in the solvent/nonsolvent demixing due to the C3N4 interfer­
–OH. The hydrogen bonding between the OH of the PVA molecules
ence in the process explained the suppression of macrovoid formation.
caused this wide absorption band (Liu et al., 2011; Lu et al., 2015).
DW was responsible for the locally induced nucleation that extended the
macrovoids in a demixing process with DW in the cast solution, as re­
3.3. Impact of viscosity on the polymeric arrangement ported by Smolders et al. (1992).
We predicted that the NH groups in C3N4 would create hydrogen
The viscosity of a polymeric matrix is related to the amount of bonds with the nonsolvent in the cast solution and the nonsolvent in the
polymer present. An increase in the polymer concentration increases the coagulation bath when added to the mixture.

4
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

Fig. 5. Cross-sectional topographies for the modified membranes with g- C3N4 nanosheets loading amount 0.5 wt% and 0.7 wt%.

The nucleation process may have been delayed because of the trap­ 2013). The stress strength of the manufactured membrane with an
ping of the nonsolvent in the polymeric solution. Thus, the hydrogen embedding ratio of 0.5 wt% was 9.12 N/m2. Additionally, the elonga­
bonding with the former delayed the nucleation process, where the tion was enhanced to 18.7 mm (Fig. 6). The mechanical properties
nonsolvent was more restrained in the polymeric solution mixture, remained high at 0.7 wt%, but the ideal embedding ratio from other
owing to the hydrogen bonds between the C3N4 and DW molecules (El studies was 0.5 wt%. This was due to the good distribution of NPs, which
Badawi et al., 2014). A high magnification was employed to confirm the provided a convenient filling and reacted with PES (Amirilargani et al.,
presence of the dense layer. The addition of g-C3N4 reduced the skin 2010; Razmjou et al., 2011).
thickness, and a dense top layer was formed. This was attributed to the
interaction between the base polymers (CA–PVA) and g-C3N4 through 3.6. Hydrophilic properties of the fabricated membranes
hydrogen bonds (Shahabi et al., 2019). Additionally, the nonporous
structure of g-C3N4 provided the modified membrane with an extra Contact angle measurements were performed to study the surface
sieving effect (Coa et al., 2015). hydrophilicity of the membranes, and the results are shown in Fig. 7.
The blank membrane had a lower contact angle than those of the
3.5. Mechanical properties modified membranes. The formed hydrogen bonds between the OH
molecules in the polymerization matrix and triazine groups slightly
The mechanical test results of the membranes indicated an increased the contact angle when g-C3N4 was incorporated into the
enhancement in the mechanical properties with an increase in the ratio polymeric solution of CA/PVA membranes (Fig. 7) (Safarpour et al.,
of NPs from 0% to 0.7%.wt. This was attributed to the excellent elec­ 2015; Vatanpour et al., 2017). The efficiency of the separation processes
trostatic and hydrogen bonding interactions between the poly­ strongly depends on the surface properties of the membrane, such as
ethersulfone (PES) chains and NPs (Huang et al., 2015; Martinez et al., hydrophilicity. A membrane is classified hydrophilic if its contact angle

5
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

Fig. 5. (continued).

Fig. 7. Water contact angle for the neat membrane and the modified with
Fig. 6. Mechanical properties of the prepared membranes with different g-c3N4.
loading amount of g-C3N4.
angle decreased and eventually reached 60◦ at a high g-C3N4 concen­
is below 90◦ and hydrophobic if its contact angle exceeds 90◦ (Amir­ tration of 0.7 wt% (Wang et al., 2017). The hydrophilic nanosheets of g-
uddin et al., 2022). Here, CA–PVA had the lowest contact angle (53◦ ) C3N4 on the surface with a concentration exceeding 0.7 wt% stimulated
because of their high hydrophilicity, while the modified membrane had the formation of hydrogen bonds with the water molecules, which
a contact angle of 67◦ . The hybrid membranes exhibited contact angles strengthened the affinity between the water and membrane surface and
slightly higher than that of the neat membrane, but significantly, the reduced the water contact angle (Oskoui et al., 2019).
modified membrane was hydrophilic because of the hydrophilic groups
in g-C3N4. However, as the g-C3N4 concentration increased, the contact

6
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

3.7. Impact of NP loading amount on separation performance

NaCl rejection at room temperature was employed to assess the in­


fluence of changing the NP embedding ratio in the polymeric matrix
from 0.0 to 0.7 wt% during membrane filtration for 5 h (Fig. 8). The
results indicated that using a low NP loading of 0.1 wt% increased the
permeate flux from 23 to 28 LMH. However, with an increase in the
embedding to 0.5 wt%, the permeate flux gradually decreased to 4.5
LMH. The salt rejection gradually increased with the loading (0.0–0.5 wt
%) of embedded C3N4 NPs in the polymeric matrix. The salt rejection
exceeded 90% for the CA/PVA membrane with 0.5- wt% C3N4. There
was a reduction in the salt separation to below 60% at a high C3N4
loading of 0.7 wt%. These fluctuations were most probably due to
changes in the polymeric structure and potential voids between the NPs
and polymeric solution. Furthermore, the variation in the NP aggrega­
tion was responsible for the varying membrane performance (Cohen,
2013). As shown in Fig. 8, the salt rejection of the CA/PVA membrane
modified with 0.5- wt% g-C3N4 reached 95% and was three times that of
the pristine CA/PVA membrane. Its flux slightly reduced, compared with
that of the pristine CA/PVA membrane. Therefore, introducing g-C3N4
in the RO membrane had a positive effect of GO on the salt rejection
more than that on the flux (Shahabi et al., 2019).

3.8. Desalination performance

Fig. 9 shows the flux of pure water for the blank and altered mem­
branes after 1 h. With the addition of g-C3N4 nanosheets to the poly­
meric layer, the NaCl salt separation improved and reached its
maximum (99.7%) in the membranes containing 0.5- wt% nanosheets.
This was due to the discord impact between the negative charges on
the film surface, owing to the presence of the g-C3N4 nanosheets and salt
particles. The zeta potential at a pH of 7 for the g-C3N4 nanosheets was
− 21.7 mV (Wen et al., 2017; Zhang et al., 2012). With an increase in the
g-C3N4 loading to 0.7 wt%, the NaCl dismissal of the alter­
ed layers diminished. This was due to the strong binding of the g-C3N4
nanosheets, which probably caused abandons on the membrane surface
and decayed the water penetration and salt dismissal.
The dense layer on the surface is a critical parameter in determining
Fig. 8. (a) Salt rejection of the membranes modified with different concen­ the salt separation of membranes. As shown in the SEM images, the
trations of g-C3N4. (b) Permeate water flux of the membranes modified with modified membranes had a thicker layer than those of the pristine
different concentrations of g-C3N4, membranes, indicating that the modified membranes could have more
selectivity and, as a result, higher salt separation. The water flux of the
modified membranes decreased as the loading of nanosheets increased,
which can be credited to the highly inflexible structure with cohered
nanosheets (Zarrabi et al., 2016).
Furthermore, owing to the hydrophilic character of the surface
amino groups, the flux of water particles penetrated the inside holes of
the g-C3N4 nanosheets with a loading quantity of 0.1 wt% (Dong et al.,
2014). The ideal pore size and thin thickness of the implanted g-C3N4
nanosheets might have improved water molecule permeability (Kim and
Nair, 2013). The water flux was also affected by the number of pores of
the selective layer of the membranes, and the flux slightly reduced with
a loading of 0.5 wt%. However, using 0.7- wt% g-C3N4 in the polymeric
matrix increased the membrane porosity, thereby increasing the water
flux (Wang et al., 2015).
The antifouling properties of the bare and C3N4-modified RO mem­
branes were tested for 120 min in a solution containing 2000-ppm NaCl
and 100-ppm BSA (Fig. 9). According to the osmotic pressure, the salty
water permeate flux reduced, compared with that of pure water.
Owing to the nature of fouling for the membranes exposed to the
protein solution, the BSA solution flux was expected to be lower than
Fig. 9. The bsa/nacl solution flux and rejection using the membranes modified that for clean water.
with different concentrations of g- c3N4. Using 0.5- wt% C3N4 for the modified membranes resulted in better
antifouling due to the enhancement in membrane hydrophilicity and the
negatively charged surface of the membranes. Moreover, the generated
nanopore-structure C3N4 nanosheets on the surface layer prevented the

7
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

Fig. 11. Water flux of 0.5% C3N4-CA/PVA membrane at different pH.

Fig. 10. The fouling resistance for 0.5% C3N4-CA/PVA membranes, (b) the
ratio of irreversible fouling (Rir) and reversible fouling (Rr) to the total fouling
(Rt) of the 0.5% C3N4-CA/PVA membranes.
Fig. 12. Permeate flux behavior of 0.5% C3N4-CA/PVA membranes as a func­
fouling agents from entering the membrane pores. The BSA molecules tion of time and fouling evaluation of BSA.
had a negative charge, which repelled charges on the membrane surface
comprising the negative C3N4 nanosheet (Chen et al., 2016). Using high modified membranes was significantly higher than that of neat mem­
concentrations of C3N4 nanosheets led to a drop in BSA rejection for branes, indicating that with the intercalation of C3N4, Rr increasingly
membranes with C3N4 nanosheets of 0.7- wt% loading. became the major fouling factor. This demonstrated the increased
The antifouling mechanisms of RO membranes can be roughly clas­ antifouling performance of the C3N4-based membranes, with the 0.5- wt
sified into inert and active antifouling mechanisms. The first is divided % C3N4-CA/PVA exhibiting the best antifouling performance.
into fouling removal and fouling-resistant mechanisms, whereas the The separation of solutes using RO membranes is dependent on the
second includes release and contact antifouling mechanisms. In general, pore size. Here, we had a nonporous membrane with a dense layer, as
the antifouling mechanism is widely used and investigated based on the observed through SEM, with a loading amount of 0.5 wt%. Many model
introduced antifouling principles (electroneutrality, hydrophilicity, theories have been developed to illustrate the RO separation mecha­
hydrogen-bond receptor, and nonhydrogen bond donor) (Ostuni et al., nism. Among these models, the dissolution–diffusion model has been
2001). Here, the antifouling mechanism depended on the hydration applied for dense and nonporous surfaces. This model was proposed by
layer effect and steric hindrance effect. The steric hindrance effect places Lonsdale (1972) and was suitable for the present study. In this model,
the diffusion process in a state of thermodynamic entropy reduction; the transmission rate depends on passing the solute and the solvent
therefore, the adsorption of solutes is inhibited (Shao et al., 2020). through the membrane under applied pressure in the form of molecular
The Rr and Rir were calculated to determine the easiest freely rein­ diffusion. The diffusion of the solute and solvent relies on the Fick law.
forced matter on the surface and the precipitation of matter inside the Therefore, the permeability of the tested solution depends on its solu­
pores, respectively. Fig. 10a indicates a higher FRR as a better anti­ bility on the membrane surface. Here, the water molecules passed faster
fouling ability for 0.5- wt% C3N4-CA/PVA (Shi et al., 2019; Xu et al., than the solute molecules through the membrane, owing to the higher
2016; Zhang et al., 2017). The 0.5- wt% C3N4-CA/PVA membrane solubility diffusion coefficient of water.
exhibited a high FRR of 98%. Fig. 10b shows that a high Rr indicated
reversible fouling, which can be easily removed by hydraulic cleaning.
This membrane also exhibited the lowest irreversible fouling (Rir). The 3.9. Membrane stability and recyclability
neat membranes exhibited close Rir and Rr values, indicating that irre­
versible and reversible fouling occurred. The Rr value of the membrane The stability of the 0.5- wt% C3N4-CA/PVA membranes was assessed
increased after C3N4 inclusion, but the Rir declined. The Rr of the C3N4- using water flux at diverse pH levels (3–11). As shown in Fig. 11, the
surface of the modified cellulosic membranes became positively charged

8
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

times of 2 h for total time of 20 h. This membrane has a steady-state


permeate water flux, which was attributed to the weakest attachment
for temporary blockage generated by the protein.
The four-stage filtering experiment using 0.5- wt% C3N4-CA/PVA
membranes was conducted for three cycles (Fig. 13). The results of the
rejection of BSA solution and permeate flux indicated enhancement in
the membrane performance by physical cleaning, demonstrating that
these membranes were recyclable. The results showed that the permeate
flux of the 0.5- wt% C3N4-CA/PVA membranes during the filtration
process were fixed after each cycle and cleaning because of the elimi­
nation of pore clogging.
The 0.5- wt% C3N4-CA/PVA membranes exhibited good stability,
great separation ability, and antifouling qualities after three cycles of
filtration and are promising candidates for the efficient treatment of
saline water.
It is important to restore the original phase of the membranes, and
this was applicable for the fabricated membranes using visible (VIS)
light because of the embedding of the prepared C3N4 in the polymeric
matrix. This performance is known as self-cleaning (Fig. 14). The
Fig. 13. The possibility of recycling ability of 0.5% C3N4-CA/PVA membrane in fabricated membrane was soaked in malachite green (MG) dye solution
terms of the flux and rejection after 3 cycles. (10 mgL− 1) for 60 min and exposed to VIS light irradiation for 30 min.
This cycle was repeated four times to ensure the self-cleaning ability. As
at a lower pH (<5) and more negatively charged at a higher pH (>7). shown in Fig. 14a, the fabricated membranes restored their original
Such amphoteric behavior was ascribed to the ester functional group in phase after the photodegradation of the MG molecules from the mem­
CA containing at least one member selected from a group comprising an brane surface. For the four cycles, the fabricated membrane achieved the
acid (Nakanishi et al., 2006). The flux slightly increased not>16.5 LMH required self-cleaning surface, resulting in a low change with an increase
in acidic media and 16.8 LMH in basic media. The modified RO mem­ in cycles.
brane had a slightly variable permeation flux, indicating that it was The VIS spectrum for the remaining concentration of the MG dye
extremely stable. after soaking the fabricated membrane for 60 min was measured and is
The BSA fouling evaluation and permeate water flux behavior of 0.5- shown in Fig. 14b. The remaining concentration for each cycle was
wt% C3N4-CA/PVA membranes as a function of time are shown in slightly lower, but the membrane still exhibited the same performance
Fig. 12. After exposure to a filtering BSA solution for 20 h, the flux was for dye adsorption, and its surface was close to the original state. Thus,
invariant versus time with a slightly steady-state value of 10.2 < Flux < the presence of g-C3N4 nanosheets as a photocatalyst in the fabricated
10.5 LMH at a working pH (6.1) and under operating conditions in batch membranes can regenerate the surface by generating hydroxyl radicals;

Fig. 14. (a) Photocatalytic effect of the 0.5% C3N4-CA/PVA membrane on malachite green dye after Four cycles under visible-light irradiation (b) the vis – spectrum
for the MG dye at 0 min and for each cycle after 60 min visible-light irradiation.

9
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

Table 3
Comparison of the C3N4 impact on the membrane performance in this study and others previous studies.
Polymer composition NPs Flux Solute Rejection% Ref.
Rejected
(LMH)

Polyacrylonitrile g- C3N4 11.7 Methyl blue(200 99.8 Li et al., 2019


mgL -1)
polyphenylsulfone CuO-C3N4 131 protein 96.0 Arumugham et al., 2019
Polyvinylidene Fluoride g- C3N4 320–350 Rhodamine B 94.0 Kolesnyk et al. (2020)
Polysulfone- PA layer g- C3N4 0.67 NaCl 98.0 Aziza et al., 2020
Polysulfone- PA layer HNTs- C3N4 18.88 NaCl 93.0 Dasht Arzhandia et al., 2018
Polysulfone- PA layer g- C3N4 60 NaCl 99.7 Shahabi et al. (2019)
Polysulfone- PA layer g- C3N4 10.05 NaCl 94.5 Rezaei- DashtArzhandi et al. (2018)
Polysulfone- PA layer g- C3N4 27.5 NaCl 99.2 Ge et al. (2021
cellulose acetate GO 65 NaCl 90.0 Ghaseminezhad et al., 2019
cellulose acetate r GO/g -C3N4 2700 RhB >60.0 Zhao et al. (2016)
cellulose acetate ZnO@ g- C3N4 51.3 C.I. Reactive 93.7 Vatanpour et al., 2021
cellulose acetate silica nanoparticles 1.6 NaCl 91.0 Amira et al., 2022
cellulose acetate/Polyvinyl alcohol g- C3N4 10.5 NaCl 95.0 This work

and after filtration using the fabricated RO membrane are listed in


Table 4 Table 4.
Physico- chemical analysis for ground water before and after filtration.
The purification of groundwater using large volumes of well water
Parameters Well Well No.1 Well Well No.2 with different compositions did not show deformation of the elaborated
No.1 purified No.2 purified
membranes in contact with feed water.
pH 6.9 ± 6.6 ± 0.05 7.4 ± 6.5 ± 0.04 Turbidity was completely removed through the filtration process.
0.05 0.05 The ions responsible for the high conductivity (sodium, chloride, sulfate,
Turbidity, NTU 2.4 ± 0 ± 0.05 1.6 ± 0 ± 0.05
0.04 0.05
and calcium carbonates) were purified with results in the range of 96%–
Conductivity, µs/ 1970 ± 31 ± 0.1 366 0 ± 1 25 ± 0.4 100%. A maximum removal of iron and manganese ions was achieved
cm 0.6 (98%–100%). The purification results of the well water after filtration
TDS, mgL-1 1310 ± 1 22 ± 0.3 2100 ± 22 ± 0.1 showed that this membrane was suitable for groundwater treatment.
0.6
Ca 2+, mgL-1 144 ± 2 ± 1.1 205 ± 1 ± 1.2
1.3 1.6 4. Conclusion
2+ -1
Mg , mgL 50 ± 1.5 1 ± 2.1 90 ± 1.6 1 ± 1.3
Na+, mgL-1 250 ± 7 ± 0.4 900 ± 5 ± 0.6 Modified ultrathin RO g-C3N4-based CA/PVA membranes were suc­
0.4 0.6
cessfully prepared. The ultrathin graphitic powder was characterized by
K+, mgL-1 4.5 ± 0.6 0.1 ± 0.4 75 ± 0.6 0±7
Cl-, mgL-1 480 ± 8 ± 0.7 1250 ± 5 ± 0.6 XRD, SEM, FTIR, and TEM. Viscosity measurements revealed that the
0.4 0.6 viscosity of the polymeric solution was significantly increased by
SO2-
4 , mgL
-1
210 ± 1 ± 0.9 510 ± 0 ± 0.5 introducing C3N4 nanosheets. By increasing the C3N4 content to 0.5 wt
1.3 0.7
%, the SEM images showed a dense top layer responsible for enhancing
NO–3, mgL-1 15 ± 0.2 1 ± 0.2 0.8 ± 0.3 0 ± 0.3
Mn2+ -1
, mgL 0.5 ± 0.1 0.01 ± 0.1 0.02 ± 0 ± 0.1
selectivity. A great enhancement in the mechanical properties of the
0.7 membrane was achieved by embedding only 0.5- wt% C3N4 in the
Fe2+, mgL-1 3.4 ± 0.3 0.05 ± 0.1 0.82 ± 0 ± 0.2 polymer matrix, stress reached 9.12 N/m2, and elongation reached
0.4 approximately 18.7 mm, compared with those of the unmodified
membrane. RO experiments applied using a 2000-ppm NaCl aqueous
the membranes are self-cleaning. solution indicated that the addition of C3N4 nanosheets improved the
Table 3 reveals that introducing g-C3N4 by blending in the polymeric membrane RO performance; water permeability was 10.5 LMH, and salt
solution of CA/PVA had great potential for water desalination, rejection was approximately 95%. Modified RO membranes containing
compared with the fabricated thin film composite (TFC) membranes that 0.5- wt% g-C3N4 represented better antifouling behavior due to the
depended on the polyamide layer. Embedding these nanosheets in enhancement in membrane hydrophilicity. The negative charges gained
polyacrylonitrile, polyvinylidene fluoride, and polyphenylsulfone poly­ using the g-C3N4 nanosheets enhanced the membrane fouling indicators
mers produced a high removal for dyes and proteins. Thus, g-C3N4-CA/ to yield a higher FRR ratio of 98%. The RO membrane stability was
PVA had a comparable and high removal of NaCl relative to other CA studied for 0.5- wt% C3N4-CA/PVA, and reasonable stability was
membrane-modified carbon materials. Amira et al. (2022) modified CA observed with great separation performance. Antifouling behavior was
membranes with silica NPs to remove NaCl with 91% and a water flux of established after four cycles of filtration. The fabricated membranes
1.6 LMH. The fabricated g-C3N4-CA/PVA membrane exhibited the best were observed to be promising candidates for efficient water desalina­
sieving for salts with convenient flux. tion. The RO-modified membranes with g-C3N4 can be nominated as
self-cleaning membranes after the removal of MG dye. Brackish water
was used to evaluate the g-C3N4-modified RO membrane; a significant
3.10. Treatment of groundwater
removal of different solutes with a removal percentage exceeding 97%
was achieved.
Groundwater was collected from two wells in the Qalyubia for water
and wastewater in Egypt. The wells were located at depths of 120 and
170 m. They contain 1310 and 2100-mgL− 1 total dissolved solids where Declaration of Competing Interest
the predominant ions are chloride, sodium, sulfate, and calcium in well
1. Iron and manganese ions were detected at a high concentration in well The authors declare that they have no known competing financial
1. The feed water from the well was treated using a testing cell under a interests or personal relationships that could have appeared to influence
pressure of 15 bars. The water characteristics of the two wells prior to the work reported in this paper.

10
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

Data availability Lonsdale, H.P.H., 1972. Reverse Osmosis Membrane Research. Plenum, New York,
pp. 43–57.
Lu, Y.-C., Chen, J., Wang, A.-J., Bao, N., Feng, J.-J., Wang, W., Shao, L., 2015. Facile
Data will be made available on request. synthesis of oxygen and sulfur co-doped graphitic carbon nitride fluorescent
quantum dots and their application for mercury(II) detection and bioimaging.
References J. Mater. Chem. C. 3, 73–78.
Majeed, S., Fierro, D., Buhr, K., Wind, J., Du, B., Boschetti-de-Fierro, A., Abetz, V., 2012.
Multiwalled carbon nanotubes (MWCNTs) mixed polyacrylonitrile (PAN)
Amira, S., Ali, Mohammed, Soliman, Moataz M., Kandil, Sherif H., Ebrahim, Shaker, ultrafiltration membranes. J. Membr. Sci. 403, 101–109.
Khalil, Marwa, 2022. Tailoring nanocomposite membranes of cellulose, of acetate/ Marchetti, P., Solomon, M.F.J., Szekely, G., Livingston, A.G., 2014. Molecular separation
silica nanoparticles for desalination. J. Materiom. with organic solvent nanofiltration: a critical review. Chem. Rev. 114, 10735–10806.
Amirilargani, M., Sadrzadeh, M., Mohammadi, T., 2010. Synthesis and characterization Martinez, L.M.P., Torres, S.M., Kontos, A.G., Moustakas, N.G., Faria, J.L., Rodriguez, J.M.
of polyether sulfone membranes. J. Polym. Res. 17, 363–377. D., Falaras, P., Silva, A.M.T., 2013. surface modified TiO2 and grapheneoxide–TiO2
Amiruddin, A.A.B.M., Chan, M.K., Ng, S., 2022. Development of contact angle prediction photocatalysts for degradation of water pollutants under near UV/Vis and visible
for cellulosic membrane: intelligent computing & optimization, 2022, 207–216 light. J. Chem. Eng. 224, 17–23.
applications. J. Membr. Sci. 564 (2018), 562–586. Nakanishi Y, H., Taniguchi, K., Ueda, in Cellulose Acetate and Dope Containing the
Arumugham, T., Amimodu, R.G., Kaleekkal, N.J., Rana, D., 2019. Nano CuO/g-C3N4 Same, U.S. Patent Ed., Daicel Chemical Industries, Ltd., Osaka, JP (2006).
sheets-based ultrafiltration membrane with enhanced interfacial affinity, antifouling Oskoui, S.A., Vatanpour, V., Khataee, A., 2019. Effect of different additives on the
and protein separation performances for water treatment application. J. Environ. Sci. physicochemical properties and performance of NLDH/PVDF nanocomposite,
82, 57–69. membrane. Sep. Purif. Technol. 209, 921–935.
Arzhandia, M.R.D., Sarrafzadeh, M.H., Goh, P.S., Lau, W.J., Ismail, A.F., Mohamed, M.A., Pawde, S.M., Deshmukh, K., Parab, S., 2008. Preparation and characterization of poly
2018. Development of novel thin film nanocomposite forward osmosis membranes (vinyl alcohol) and gelatin blend films. J. Appl. Polym. Sci. 109, 1328.
containing halloysite/graphitic carbon nitride nanoparticles towards enhanced Pawde, S.M., Deshmukh, K., 2008. Characterization of Polyvinyl Alcohol/Gelatin Blend
desalination performance. Desalination 447, 18–28. Hydrogel Films for Biomedical Applications. J. Appl. Polym. Sci. 109, 3431.
Aziza, A.A., Wong, K.C., Goha, P.S., Ismail, A.F., Azelee, I.W., 2020. Tailoring the surface Pendergast, M.M., Hoek, E.M.V., 2011. A review of water treatment membrane
properties of carbon nitride incorporated thin film nanocomposite membrane for nanotechnologies. Energy Environ. Sci. 4, 1946–1971.
forward osmosis desalination. J. Water Process Eng. 33, 101005. Peng, N., Chung, T.-S., Wang, K.Y., 2014. Macrovoid Evolution and Critical Factors to
Bu, Y., Chen, Z., Yu, J., Li, W., 2013. A novel application of g-C3N4 thin film in Form Macrovoid-free Hollow Fiber Membranes. J. Memb. Sci. 318, 363–372.
photoelectrochemical anticorrosion. Electrochim. Acta. 88, 294. Razmjou, A., Mansouri, J., Chen, V., 2011. The effects of mechanical and chemical
Chen, J., Li, Z., Wang, C., Wu, H., Liu, G., 2016. Synthesis and characterization of g-C3N4 modification of TiO2 nanoparticleson the surface chemistry, structure and fouling
nanosheet modified polyamide nanofiltration membranes with good permeation and performance of PES ultrafiltration membranes. J. Membr. Sci. 378, 73–84.
antifouling properties. RSC Adv. 6, 112148–112157. Safarpour, M., Vatanpour, V., Khataee, A., Esmaeili, M., 2015. Development of a novel
Cohen, S.T., Araque, J.C., Hoek, E.M.V., Escobedo, F.A., 2013. Molecular dynamics of high flux and fouling-resistant thin film composite nanofiltration membrane by
equilibrium and pressure-driven transport properties of water through LTA-type embedding reduced graphene oxide/TiO2. Sep. Purif. Technol. 154, 96–107.
zeolites. Langmuir 29, 12389–12399, 2018. Development of novel thin film Saljoughi, E., Sadrzadeh, Mohammadi T., 2009. Effect of preparation variables on
nanocomposite forward osmosis membranes containing halloysite/graphitic carbon morphology and pure water permeation flux through asymmetric cellulose acetate
nitride nanoparticles towards enhanced desalination performance. desalination. 447, membranes. J. Membr. Sci. 326, 627–634.
18-28. Shahabi, S.S., Azizia, N., Vatanpourb, V., 2019. Synthesis and characterization of novel
Dong, F., Wang, Z., Li, Y., Ho, W.-K., Lee, S., 2014. Immobilization of polymeric g-C3N4 g-C3N4 modified thin film nanocomposite reverse osmosis membranes to enhance
on structured ceramic foam for efficient visible light photocatalytic air purification desalination performance and fouling resistance. Sep. Purif. Technol. 215, 430–440.
with real indoor illumination. Environ. Sci. Technol. 48, 10345–10353. Shannon, M.A., Bohn, P.W., Elimelech, M., Georgiadis, J.G., Marinas, B.J., Mayes, A.M.,
El Badawi, N., Ramadan, A.R., Esawi, A.M., El-Morsi, M., 2014. Novel carbon nanotube– 2008. Science and technology for water purification in the coming decades. Nature
cellulose acetate nanocomposite membranes for water filtration applications. 452, 301–310.
Desalination. 344, 79–85. Shi, Y., Huang, J., Zeng, G., Cheng, W., Hu, J., Shi, L., Yi, K., 2019. Evaluation of self-
Galiano, F., et al., 2018. Advances in biopolymer-based membrane preparation and cleaning performance of the modified g-C3N4 and GO based PVDF membrane
applications. J. Membr. Sci. 564, 562–586. toward oil-in-water separation under visible-light. Chemosphere 230, 40–50.
Ge, M., Wang, X., Wu, S., Long, Y., Yang, Y., Zhang, J., 2021. Highlyantifouling and Smolders, C.A., Reuvers, A.J., Boom, R.M., Wienk, I.M., 1992. Microstructures in
chlorine resistance polyamide reverseosmosis membranes with g-C3N4 nanosheets phaseinversion membranes. Part 1. Formation of macrovoids. J. Membr. Sci. 73,
as nanofiller. Sep. Purif. Technol. 258, 117980 https://doi.org/10.1016/j. 259–275.
seppur.2020.117980. Van Reis, R., Zydney, A., 2001. Membrane separations in biotechnology. Curr. Opin.
Ghaseminezhad, S.M., Barikani, M., Salehirad, M., 2019. Development of graphene Biotechnol. 12, 208–211.
oxide-cellulose acetate nanocomposite reverse osmosis membrane for seawater Vatanpour, V., Safarpour, M., Khataee, A., Zarrabi, H., Yekavalangi, M.E., Kavian, M.,
desalination. Comp. Part B: Eng. 161, 320–327. 2017. A thin film nanocomposite reverse osmosis membrane containing amine-
Han, R., 2013. Formation and characterization of (melamine–TMC) based thin film functionalized carbon nanotubes. Sep. Purif. Technol. 184, 135–143.
composite NF membranes for improved thermal and chlorine resistances. J. Membr. Vatanpour, V., Faghani, S., Keyikoglu, R., Khataee, A., 2021. Enhancing the permeability
Sci. 425–426, 176–181. and antifouling properties of cellulose acetate ultrafiltration membrane by
Huang, H.B., Ying, Y.L., Peng, X.S., 2014. Graphene oxide nanosheet: an emerging star incorporation of ZnO@graphitic carbon nitride nanocomposite. Carbohydr.
material for novel separation membranes. J. Mater. Chem. A. 2, 13772–13782. Polymers 256, 117413.
Huang, X., Zhang, J., Wang, W., Liu, Y., Zhang, Z., Li, L., Fan, W., 2015. Effects of PVDF/ Wang, J., Li, Meisheng, Zhou, Shouyong, Xue, Ailian, Zhang, Yan, Zhao, Yijiang, 2017.
SiO2 hybrid ultrafiltration membranes by sol–gel method for the concentration of Graphitic carbon nitride nanosheets embedded in poly(vinyl alcohol) nanocomposite
fennel oil in herbal water extract. RSC Adv. 5, 18258–18266. membranes for ethanol dehydration via pervaporation. Sep. Purif. Technol.
Jackson, E.A., Hillmyer, M.A., 2010. Nanoporous membranes derived from block Wang, Y., Ou, R., Wang, H., Xu, T., 2015. Graphene oxide modified graphitic carbon
copolymers: from drug delivery to water filtration. ACS Nano 4, 3548–3553. nitride as a modifier for thin film composite forward osmosis membrane. J. Membr.
Jamil, T.S., Abdallah, H., Shaban, A.M., Mansor, E.S., Souaya, E.R., 2018. The influence Sci. 475, 281–289.
of the Polyacrylonitrile proportion on the fabricated UF blend membranes Wen, J., Xie, J., Chen, X., Li, X., 2017. A review on g-C3N4-based photocatalysts. Appl.
performance for humic acid removal. J. Polym. Eng. 38, 129–136. Surf. Sci. 391, 72–123.
Jia, Z.Q., Shi, W.X., Wang, Y., Wang, J.L., 2016. Dicarboxylic acids crosslinked grapheme Wood, J.W.H., Sourirajan, S., 1993. The effect of polyethersulfone concentration on flat
oxide membranes for salt solution permeation. Colloids Surf. A: Physicochem. Eng. and hollow fiber membr\ane performance. Sep. Sci. Technol. 28, 2297–2317.
Aspects. 494, 101–107. Wu, M., Wang, Q., Sun, Q., Jena, P., 2013. Functionalized graphitic carbon nitride for
Kim, W.-G., Nair, S., 2013. Membranes from nanoporous 1D and 2D materials: A review efficient energy storage. J. Phys. Chem. C. 117, 6055.
of opportunities, developments, and challenges. Chem. Eng. Sci. 104, 908–924. Xu, Z., Wu, T., Shi, J., Teng, K., Wang, W., Ma, M., Li, J., Qian, X., Li, C., Fan, J., 2016.
Kolesnyk, I., Kujawa, J., Bubela, H., Konovalova, V., Burban, A., Cyganiuk, A., Photocatalytic antifouling PVDF ultrafiltration membranes based on synergy of
Kujawski, W., 2020. Photocatalytic properties ofPVDF membranes modified with g- graphene oxide and TiO 2 for water treatment. J. Membr. Sci. 520, 281–293.
C3N4 in the process ofRhodamines decomposition. Sep. Purif. Technol. 117231. Yin, J., Fan, H., Zhou, J., 2016. Cellulose acetate/poly (vinyl alcohol) and cellulose
Krimm, S., Liang, C.Y., Sutherland, G.B.B.M., 1956. Infrared spectra of high polymers. V. acetate/crosslinked poly (vinyl alcohol) blend membranes: preparation, and
Polyvinyl alcohol. J. Polym. Sci. 22, 227. antifouling properties. Desalin. Water Treat. 57 (23), 10572–10584.
Lin, S., Cong, Z.X., Li, J., Ke, K.M., Guo, S.S., Yang, H.H., Chen, G.N., 2014. Graphitic- Yu, S., Webster, R.D., Zhou, Y., Yan, X., 2017. novel ultrathin g-C3N4 nanosheet with
phase C3N4 nanosheets as efficient photosensitizers and pH-responsive drug hexagonal CuS nanoplates composite under solar light irradiation for H2 production.
nanocarriers for cancer imaging and therapy. J. Mater. Chem. B. 2, 1031. Catal. Scie. Technol. 7, 10.
Liu, H.L.X., Li, P., 2017. Effect of polymer dope concentration on the morphology and Yuan, Y.Q., Gao, X.L., Wei, Y., Wang, X.Y., Wang, J., Zhang, Y.S., 2017. Enhanced
performance of PES/PDMS hollow fiber composite membrane for gas separation. desalination performance of carboxyl functionalized graphene oxide nanofiltration
J. Material Sci. 1, 1–5. membranes. Desalination 405, 29–39.
Liu, J., Zhang, T., Wang, Z., Dawson, G., Chen, W., 2011. Simple pyrolysis of urea into Zarrabi, H., Yekavalangi, M.E., Vatanpour, V., Shockravi, A., Safarpour, M., 2016.
graphitic carbon nitride with recyclable adsorption and photocatalytic activity. Improvement in desalination performance of thin film nanocomposite nanofiltration
J. Mater. Chem. 21, 14398–14401. membrane using amine-functionalized multiwalled carbon nanotube. Desalination
394, 83–90.

11
E.S. Mansor et al. Environmental Nanotechnology, Monitoring & Management 19 (2023) 100760

Zhang, Y., Antonietti, M., 2010. Photocurrent generation by polymeric carbon nitride Zhang, X., Xie, X., Wang, H., Zhang, J., Pan, B., Xie, Y., 2013. Enhanced photoresponsive
solids: an initial step towards a novel photovoltaic system. J. Chem.–Asian. 5, 1307. ultrathin graphitic-phase C3N4 nanosheets for bioimaging. J. Am. Chem. Soc. 135,
Zhang, M., Liu, Z., Gao, Y., Shu, L., 2017. Ag modified g-C3N4 composite entrapped PES 18.
UF membrane with visible-light-driven photocatalytic antifouling performance. RSC Zhao, H., Chen, S., Quan, X., Yu, H., Zhao, H., 2016. Integration of microfiltration and
Adv. 7, 42919–42928. visible-light-drivenphotocatalysis on g-C3N4 nanosheet/reduced graphene
Zhang, Y., Mori, T., Ye, J., Antonietti, M., 2010. Phosphorus-doped carbon nitride solid: oxidemembrane for enhanced water treatment. Appl. Catal. BEnviron. 194,
enhanced electrical conductivity and photocurrent generation. J. Am. Chem. Soc. 134–140.
132, 6294. Zhu, X., Ji, H., Yi, J., Yang, J., She, X., Ding, P., Li, L., Deng, J., Qian, J., Xu, H., Li, H.,
Zhang, X., Xie, X., Wang, H., Zhang, J., Pan, B., Xie, Y., 2012. Enhanced photoresponsive 2018. A specifically exposed cobalt oxide/carbon nitride 2D heterostructure for
ultrathin graphitic-phase C3N4 nanosheets for bioimaging. J. Am. Chem. Soc. 135, carbon dioxide photo-reduction. Ind. Eng. Chem. Res. 57, 17394–17400.
18–21. Zhu, G., Lu, K., Sun, Q., Kawazoe, Y., Jena, P., 2014. Lithium-doped triazine-based
graphitic C3N4 sheet for hydrogen storage at ambient temperature. Comput. Mater.
Sci. 81, 275.

12

You might also like