You are on page 1of 14

Applied Nanoscience

Influence of melamine and graphene oxide on the performance of polyamide reverse


osmosis membranes for desalination
--Manuscript Draft--

Manuscript Number:

Full Title: Influence of melamine and graphene oxide on the performance of polyamide reverse
osmosis membranes for desalination

Article Type: Original Article

Funding Information:

Abstract: HIGHLIGHTS

• Melamine functionalized with Graphene oxides to obtain MEL/GO


• MEL/GO blended in a polyamide layer lowers the water contact angle of the
membranes
• Over 15% higher tensile strength obtained for MEL/GO membrane
• 9-Fold higher flux with a significant alteration in salt rejection in MEL/GO membrane •
Good separation of the divalent ions using MEL/GO membrane

Corresponding Author: Fatma Taher


Al-Azhar University Faculty of Science for Girls in Cairo
EGYPT

Corresponding Author Secondary


Information:

Corresponding Author's Institution: Al-Azhar University Faculty of Science for Girls in Cairo

Corresponding Author's Secondary


Institution:

First Author: Esraa M. Elghonemy

First Author Secondary Information:

Order of Authors: Esraa M. Elghonemy

Gehad Hamdy

Heba Abdallah

Naglaa Saad

Fatma Taher

Order of Authors Secondary Information:

Author Comments:

Suggested Reviewers: Alireza Shakeri


University of Tehran College of Science
alireza.shakeri@ut.ac.ir
has an experience in the field

Syed Javaid Zaidi


Qatar University
smjavaidzaidi@gmail.com
has an experience in the field

Hui Ding
Tianjin University School of Environmental Sciences And Technology: Tianjin
University School of Environmental Science and Engineering
dinghui@tju.edu.cn
has an experience in the field

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Click here to access/download;Manuscript;Final esraa..pdf

Click here to view linked References

1 Influence of melamine and graphene oxide on the performance of polyamide reverse osmosis
2 membranes for desalination

4 Esraa M. Elghonemy1, Gehad Hamdy2,3, Heba Abdallah4, Naglaa Saad2, Fatma A. Taher2,3,
1
5 M.SC. student, Chemistry Department, Faculty of science, Al-Azhar University (Girls), Nasr City, 11754 Cairo, Egypt
2Chemistry
6 Department, Faculty of science, Al-Azhar University (Girls), Nasr City, 11754 Cairo, Egypt
3
7 Al-Azhar Technology Incubator (ATI), Al-Azhar University (Girls), Nasr City, Egypt
4
8 Chemical Engineering & Pilot Plant Department, Engineering and Renewable Energy Research Institute, National Research
9 Center, 33 El-Bohouth St. (Former El-Tahrir St.), Dokki, Giza, Egypt
10
11 HIGHLIGHTS
12 • Melamine functionalized with Graphene oxides to obtain MEL/GO
13 • MEL/GO blended in a polyamide layer lowers the water contact angle of the membranes
14 • Over 15% higher tensile strength obtained for MEL/GO membrane
15 • 9-Fold higher flux with a significant alteration in salt rejection in MEL/GO membrane
16 • Good separation of the divalent ions using MEL/GO membrane
17 ABSTRACT

18 Water resource pollution, extreme weather patterns, population expansion, growing agricultural needs, and fast industrialization
19 all put pressure on the creation of novel technology for the production of drinkable water. Hence, using membrane-based

1 20 desalination technology to treat produced water opens new avenues for water recovery. Here, the phase inversion approach was
2 21 efficiently utilized to create melamine (MEL) grafted by graphene oxide (GO) nanosheet-based polyamide (PA) reverse osmosis
3
4 22 (RO) membranes. Many aspects of membrane characteristics, like hydrophilicity, porosity, surface and cross-sectional
5
6 23 morphology, permeability, and nanofiltration performance, were examined in relation to this integration. Because the membrane
7
8 24 matrix included extremely hydrophilic MEL/GO filler, the resulting membranes had increased hydrophilicity, porosity, and
9
10 25 permeability. The constructed membranes' pure water flow rose from 10.01 LMH/bar for the bare M0 to 73.47 LMH/bar, 23.35
11
12
26 LMH/bar, and 88.21 LMH/bar for the optimal contents of MEL, GO, and MEL/GO, respectively, at an operating pressure of 20
13 27 bar and temperature of 25 °C. Salt rejection (%) increased with membrane having the optimal MEL/GO ratio (0.1/0.3 wt%),
14
15 28 from 71.74% for the bare M0 to 96.57%. The current study's findings ultimately resulted in the introduction of the small molecule
16
17 29 melamine into the GO layer for the first time. Melamine interacts with oxygen-containing groups to enrich the amine functional
18
19 30 group, which helps to promote water transportation and prepare very hydrophilic nanofillers.
20
21 31 Keywords: melamine – RO membranes – polyamines – GO nanosheets – polyamide membrane
22
32
23
24 33 GRAPHICAL ABSTRACT
25
26 34
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54 35
55
56 36 1. Introduction
57
58 37 The availability of pure water is a large issue facing humanity today. Large quantities of clean water resources are
59
60 38 needed nowadays for various uses, from domestic to agricultural irrigation, due to the rapid population growth (1, 2). According
61
62
63
64 1
65
39 to estimates, two-thirds of the global people is thought to experience acute water shortage at least once a year (3). Based on
40 the United Nations World Water Development Report, and according to research by the Water Resources Group, Water and
41 Climate Change claims that if present trends continue, by 2030, the globe will have around 60% of the water it needs (4).
42 Consequently, the value of water recycling and desalination—the process of turning the readily accessible saltwater into a
43 usable state—becomes crucial. Many people believe that seawater desalination and wastewater recycling are effective ways to
44 increase water supplies and reduce water shortages (5, 6). More than 97% of the world's water supplies are seawater. Hence,
45 seawater desalination has emerged as a significant and favorable water purification technique (7).

46 Desalination is the procedure by which salts, minerals, and impurities are removed from wastewater effluent, brackish
47 water, and seawater to generate pure water for domestic, agricultural, and industrial use. Using the desalination process, pure
48 water has been produced from 67% of seawater, 19% of brackish water, 8% of river water, and 6% of other effluent (8, 9). The
49 two types of techniques for desalination are phase-change, also referred to as the thermal technique (10), and membrane-based,
50 which is based on the mechanism of separation (11). For the former, a constant procedure of evaporation and condensation
51 removes salts from water (12). Regarding the letter, Reverse osmosis (RO), nanofiltration (NF), and electrodialysis (ED) are
52 three membrane-based desalination methods that employ a membrane to filter water while preserving salt and other minerals
53 (13, 14).

54 In comparison to other water purification techniques, RO is a leading technology and it is used in more than 50% of the
55 world's desalination facilities (13, 15). Moreover, it provides sustainably fresh water from saltwater or brackish water owing to
56 its simplicity, energy-saving, and easy scale up (16). In order to achieve equilibrium in solute concentration on the two membrane
1 57 sides, the solvent typically flows from the region of reduced to increased solute concentration by the mechanism of osmotic
2
3 58 pressure. As a result, the RO process requires applying external pressure in order to reverse the direction of the water flow and
4
5 59 counteract the osmotic pressure produced by the running water (17). The feed water flows through the membrane, leaving salt
6
7 60 and minerals behind, and produces clean water on the other side when the conducted pressure is more than the osmotic pressure
8
9 61 (18, 19).
10 62 The highly water-permselective thin-film composite (TFC) membrane composition is a vital factor in identifying the
11
12 63 effectiveness of RO procedures. It typically consists of three crucial layers named from top to bottom: selective layer (polyamide;
13
14 64 PA), porous layer, and nonwoven fabric (20, 21). Phase inversion technique is usually utilized for the preparation of porous layer
15
16 65 (middle layer) by dissolving polymer in a solvent and then casting knife is used to form a caste on the surface of glass support
17
18 66 layer. Directly, the casted membrane is dipped in water to enhance solvent elimination (22, 23). The mechanical strength of the
19
20 67 membrane during the filtering process is determined by these layers (24).
21
22
23 68 The uppermost selective layer, also referred to as the PA layer, is classically produced by employing the interfacial
24
69 polymerization method (IP). This approach involves the irreversible and quick polymerization of two intermediates at the
25
26 70 interface between the aqueous (liquid system) and organic (heterogeneous system) stages (25, 26). IP method made a
27
28 71 revolutionization in the production of dense polymeric membranes owing to its ease along with reduced production cost (27).
29
30 72 Acid chloride monomer (Trimesoyl chloride; TMC) and amine-functionalized monomer (m-phenylenediamine; MPD) are
31
32 73 usually utilized as intermediate monomers for production of PA layer whereas they are dissolved in organic and aqueous phases,
33
34 74 respectively (28). The importance of PA layer raising from its responsibility in regulating the main membrane properties
35 75 including salt selectivity, permeability, and resistances against chlorination and fouling (29). The successfully fabricated
36
37 76 membrane from polyamide TFC pioneered present RO processes because of its elevated water permeance and salt rejection (30).
38
39 77 Nevertheless, there is an urgent demand to decrease energy utilized during RO processes whereas this difficulty limit the
40
41 78 achievement of desalination owing to its strong trade-off and intrinsic association between water/salt water permeability and
42
43 79 selectivity (15).
44
45 80 One of the most studied approaches to address this restriction in polymeric water purification membranes is the
46 81 introduction of nanoporous materials, also known as nanofillers, into a polyamide matrix. This enables the fabrication of thin-
47
48 82 film nanocomposite (TFN) membranes, which leverage the benefits offered by both constituent phases. Nanoporous materials
49
50 83 (i.e., nanofillers) including carbon nanotubes, titanium dioxide, graphene oxide (GO), gold nanoparticle, and zeolite,
51
52 84 metal−organic frameworks are recently embedded in the polyamide layer to make TFN membranes to overcome TFC limitation
53
54 85 (31). This enhances the performance of membrane during RO process where additional channels are formed in the PA matrix
55
56
86 which transports water. The suitability of TFN membranes for the scale-up of production comes from the achieved improvement
57 87 in the RO performance compared with traditional TFC membranes (32, 33).
58
59
60
61
62
63
64 2
65
88 Among them, graphene was utilized as matrix during preparation of PA layer whereas surface modification is required
89 to avoid formation of hard matrix and enhance its dispersion in polymer matrices. This could be resulted from strong π–π
90 attraction along with van der Waals forces between the nanosheets (34, 35). One of the graphene derivatives that may be made
91 using inexpensive chemical functionalization methods is GO. The major advantage of GO compared with graphene is the high
92 hydrophilicity which arises from the existence of well-distributed oxygenated functional groups like carboxylic, carboxyl,
93 hydroxyl, and epoxy groups (36). Moreover, GO has additional advantages including excellent chemical stability, large surface
94 area, and mechanical stability (37). Therefore, GO has been utilized as the favored filler material during the preparation of
95 polymeric matrices of PA layer (38). It has been found that GO remarkably increases mechanical strength of the prepared filter
96 membranes (39, 40). For example, Ali et al. prepared TFC consisting of TMC, and MPD along with its counterpart containing
97 GO. It was found that plain TFC had low salt rejection capability at high applied pressure compared with GO embedded one.
98 This could be attributed to high elongation capability of TFC containing GO which make it able to withstands applied high
99 pressure. Moreover, water flux through GO embedded membrane was higher than plain membrane. However, this increase in
100 water flux capability was not increased to a notable degree (41).

101 The low water flux through filter membrane could be resolved in integration of monomers rich with hydrophilic groups
102 such as melamine (42). In addition, it has been reported that incorporation of melamine within prepared filter membranes increase
103 the desalination value (43). Therefore, melamine (MEL) was utilized during the preparation of PA as amine monomers solution
104 which increased cross-linking the prepared layer. In this context, GO and MEL were utilized to fabricate TFC able to withstand
105 high pressure along with its ability to increase waster flux through membrane with high ability to remove salts from the filtrate.

1
2 106 Therefore, this investigation aimed to create a new TFN RO polyamide membrane. The membrane was designed to
3
107 possess superior characteristics such as elevated permeation flux, increased separation performance, and elevated chemical
4
5 108 stability. This was achieved by modifying MEL through a chemical reaction with GO between amine groups of MEL and the
6
7 109 carboxylic acid moieties located at the boundaries of the GO layers. Furthermore, we investigated the impacts of incorporating
8
9 110 the adjusted GO in the TFC RO membrane performance. To further validate the findings, a range of membrane characteristics
10
11 111 were performed, including FTIR, FESEM, and contact angle.
12
13 112 2. Materials and methods
14
15 113 2.1. Materials
16
17 114 The German company BASF provided the polyethersulfone (PES, MW = 58,000 g/mol), N-methyl-2-pyrrolidone
18
19 115 (NMP), and non-woven polyester fabric (0.1 Micron, 90mm, 30/PK, Sterlitech, PET019030). The hydrophilic agent polyethylene
20
21 116 glycol 400 (PEG, Mw 400 g/mol, acquired from Merck) , m phenylenediamine (MPD, Aldrich, 78420), triethylamine (TEA,
22
23 117 Sigma-Aldrich, T0886), camphorsulfonic acid (CSA, Sigma-Aldrich,282146), % trimesoyl chloride (TMC, Aldrich, 147532),
24
118 N-hexane and melamine(Mel,MW= 126,12 g/mol) were obtained from Acros Organics (USA). GO was made by previously
25
26 119 established method (modified Hummer’ method) in our lab (44).
27
28 120
29
121 2.2. Preparation of PES support membrane
30
31
32 122 In order to produce the thin film composite (TFC) membrane, a membrane of support with ultrafiltration made of
33
34 123 PES was first fabricated using the phase inversion approach. The casting solution consisted of 18 wt.% PES, 1 wt.% PEG, and
35 124 79 wt.% NMP. PEG was utilized as an additive in the solution to enhance the membrane's porosity. The first step included
36
37 125 dissolving it in NMP, subsequent to which PES was added. The resulting mixture was then subjected to 4 h of mechanical stirring,
38
39 126 resulting in the formation of a homogeneous solution. The solution was thereafter left without stirring for one day to facilitate
40
41 127 the elimination of any air bubbles that may have existed in the casting solution. Subsequently, the solution was applied onto a
42
43 128 commercial polyester non-woven fabric employing a casting knife with a knife gap of 20 mm. After that, the material was
44
45
129 submerged in a coagulation bath, including a precipitant medium, which is clean water, that functions as a non-solvent in the
46 130 process of precipitation. The membrane is embedded in water for one day in order to facilitate the process of phase separation.
47
48 131 Following this, the membrane underwent a drying process by being placed at room temperature for 24 h between two filter paper
49
50 132 sheets .
51
52
53 133 2.3. Preparation of PA TFC and PA MEL RO membranes
54
55 134 The PA TFC membranes were made up utilizing the process of IP on the supporting membranes of PES. The membranes
56
57 135 supported by PES were immersed in an aqueous MPD solution (2.25 wt. % in water) with additives (1.7 wt % TEA and 1.7 wt
58
59 136 % CAS) and allowed to soak for 5 min. Remove any extra MPD solution and let the membrane undergo natural air drying. At
60
61
62
63
64 3
65
137 room temperature, excessive solution evaporated in the air for 5 min. Subsequently, the PES membranes that had been saturated
138 were submerged in (0.125 wt %) an organic solution of TMC in n-hexane for one minute. This process resulted in the creation
139 of a thin layer on the supporting membrane. Following this, the membrane was exposed to a 5-minute drying at 80 °C in an
140 oven. Subsequently, it was cleansed by a sodium carbonate solution (0.2 wt %) at a temperature of 50 °C. Subsequently, the
141 membranes underwent a rinsing process with distilled water, followed by placement in an airtight container for the purpose of
142 conducting performance evaluations and characterizing their properties.

143 The Copolyamide thin film composite (Co-TFC) membranes were produced utilizing a method identical to that of PA
144 TFC membranes but with different concentrations of MEL. MEL was introduced into an aqueous solution at several
145 concentrations (0.05, 0.1, and 0.2 wt.%) and underwent a 30-minute full dispersion employing sonication. The resulting solution
146 was utilized as an aqueous solution for the process of interfacial polymerization, whereby it was subjected to a reaction with an
147 organic solution. These manufactured Co-TFC RO membranes are indicated as Mm0.05, Mm0.1 and Mm0.2 respectively
148 assigned to their MEL concentration in the aqueous solution as represented in Table 1.

149 2.4. Preparation of PA TFC and PA TFN GO RO membranes

150 The thin film nanocomposite (TFN) GO membranes of PA were made up using identical processes as the PA TFC
151 membrane, with the exception of the aqueous solution. Various quantities of GO were introduced to the aqueous solution at
152 varying concentrations of (0.1 ,0.3 and 0.5 wt %) as represented in Table 1. well-dispersed solution. These syntheses PA TFN
153 GO RO membranes are denoted as MG0.1, MG0.3, and MG0.5, respectively, introduced to their GO levels in the aqueous
154 solution as represented in Table 1.
1
2 155 2.5. Preparation of Novel Copolyamide TFC and TFN GO RO membranes
3
4 156 The (Co-TFN) RO membranes were produced by including a combination of MEL and GO in a water-based solution at
5
6 157 various concentrations (0.1/0.3 wt. %). The mixture was thoroughly dispersed using sonication for a duration of 30 minutes. The
7
8 158 resulting solution was employed as an aqueous solution for the process of interfacial polymerization, whereby it was subjected
9
10 159 to reaction with an organic solution. The produced Co-TFN RO membranes are identified as Mm0.1/G0.3 respectively assigned
11
160 to their (MEL/GO) concentration in the aqueous solution as represented in Table1.
12
13
14 161 Table 1
15
162 The structure of the aqueous and organic solutions for preparation of RO membranes.
16
17 Membrane PES PEG NMP Aqueous phase Organic
18
19 (%) (%) (%) phase
20
21 MPD(%) TEA(%) CAS(%) GO(%) MEL(%) TMC(%)
22
23 2.25 1.7 1.7 --- --- 0.125
24 M0 18 1 79
25
26
27 Mm0.05 18 1 79 2.25 1.7 1.7 --- 0.05 0.125
28
29
Mm0.1 18 1 79 2.25 1.7 1.7 --- 0.1 0.125
30 Mm0.2 18 1 79 2.25 1.7 1.7 --- 0.2 0.125
31
32 MG0.1 18 1 79 2.25 1.7 1.7 0.1 --- 0.125
33
34 MG0.3 18 1 79 2.25 1.7 1.7 0.3 --- 0.125
35
36
37 MG0.5 18 1 79 2.25 1.7 1.7 0.5 --- 0.125
38
39
Mm0.1/G0.3 18 1 79 2.25 1.7 1.7 0.3 0.1 0.125
40 163
41
42
43 164 3. Membrane characterization
44
45 165 3.1. Morphology detection
46
47 166 The morphological analysis of the membranes was conducted by subjecting them to freezing in liquid nitrogen to get cross-
48
49
167 sectional pictures. Subsequently, the samples were coated with gold to enhance their electrical conduction property. QUANTA
50 168 FEG250 scanning electron microscope (SEM) was employed to capture the cross-sectional pictures of membranes.
51
52 169 3.2. Fourier transform infrared (FTIR) spectroscopy
53
54 170 The presence of functional groups on the surface of the produced membranes was verified utilizing FTIR analysis. The
55
56 171 analysis was conducted using a JASCO FTIR spectroscopy model: 6100, with a scanning rate of 16 scans per minute and a
57
58 172 resolution of 4 cm−1.
59
60
61
62
63
64 4
65
173 3.3 Mechanical characteristics
174 The mechanical characteristics at the point of breakage of the developed membranes were quantified by evaluating their
175 tensile stress and elongation. The H5KS universal tensile testing equipment was used to assess these values in the samples. The
176 starting length between grips was set at 180 mm, while the width of the examined membranes was 50 mm. The samples were
177 subjected to elongation at a consistent rate of 30 mm/min.
178 3.4. Contact angle and Membrane porosity
179 To examine the additives impact on the hydrophilicity of support layers that were manufactured before, the contact angle
180 of the membrane surfaces was evaluated. To accomplish the desired objective, the measurements were conducted employing the
181 sessile drop technique with the use of contact angle measuring equipment (OCA15 plus, 196 Data physics, Germany). The
182 obtained outcomes represent the average contact angle shown by droplets of DI water at three various sites on every specimen.
183 The investigation focused on the examination of porosity and water content as intrinsic characteristics of the membranes. This
184 was achieved by immersing the samples in distilled water, followed by drying at a temperature of 80°C utilizing an air circulating
185 oven for one day. The porosity and the equilibrium water content (EWC) of the prepared membrane were identified employing
186 Eqs. (1) and (2) (32, 33):

𝑤𝑒𝑡 𝑚𝑒𝑚𝑏𝑟𝑎𝑛𝑒 𝑤𝑒𝑖𝑔ℎ𝑡 (𝑔)−𝐷𝑟𝑦 𝑚𝑒𝑚𝑏𝑟𝑎𝑛𝑒 𝑤𝑒𝑖𝑔ℎ𝑡 (𝑔)


187 Porosity (%) = × 100 % (1)
𝑚𝑒𝑚𝑏𝑟𝑎𝑛𝑒 𝑎𝑟𝑒𝑎 ×𝑚𝑒𝑚𝑟𝑎𝑛𝑒 𝑡ℎ𝑖𝑐𝑘𝑛𝑒𝑠𝑠

𝑤𝑒𝑡 𝑚𝑒𝑚𝑏𝑟𝑎𝑛𝑒 𝑤𝑒𝑖𝑔ℎ𝑡 (𝑔)−𝐷𝑟𝑦 𝑚𝑒𝑚𝑏𝑟𝑎𝑛𝑒 𝑤𝑒𝑖𝑔ℎ𝑡 (𝑔)


188 W𝐻2𝑂 (%) = × 100 % (2)
𝐷𝑟𝑦 𝑚𝑒𝑚𝑏𝑟𝑎𝑛𝑒 𝑤𝑒𝑖𝑔ℎ𝑡 (𝑔)

189 where W𝐻2𝑂 is the EWC for the prepared membranes.


1
2
3 190 3.5 Membrane performance measurements
4
5 191 A laboratory desalination machine was employed to perform membrane performance trials (Figure 1). This system
6
7 192 includes a module of flat sheet membrane that consists of three distinct apertures, namely for feeding, concentrating, and
8
9
193 permeating. The feed was consistently supplied to the membrane module via a sealed feeding tank with a capacity of 50 liters,
10 194 employing a pump with increased pressure. Subsequently, the product was obtained from the downstream region of the
11
12 195 membrane module. The membranes that were made were placed into a stainless steel plate module with a diameter of 12 cm. In
13
14 196 all trials, a consistent supply of synthetic solutions with a concentration of 2000 mg/l of NaCl was continually introduced into
15
16 197 the membrane module. This process was performed at 25 °C and a pressure of 20 bar.
17
18
19
198 Long-term experiments were performed on the optimum membrane to identify the stability of the membrane. However, various
20 199 salts such as MgCl2, Na2SO4 and CalCl2 were used in a feed solution with concentration of 800 ppm, 1000 ppm and 500 ppm
21
22 200 respectively to study the efficiency of the optimum membrane at 25 °C and under the pressure of 20 bar.
23
24
25 201 Salt rejection and water flux were detected for all membranes in variuos trials (30–34).
26
27
28 202 Salt rejection R (%):
29 𝑐𝑝
30 203 𝑅 (%) = (1 − ) × 100%
31 𝑐𝑓
32
33 204 Cf: feed concentration, Cp: permeate concentration.
34
35 205 Water flux Jw:
36 m
37 206 Jw=
38 A. Δ t
39 207 where: m is the permeate water weight, A area of membrane and t time consumed.
40
41
42 208
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 5
65
209

210 Fig.1: Schematic drawing for lab desalination unit.


211 4. Results and discussion
212 4.1 Membrane characterization
213 4.1.1. Membrane morphology
214 Fig. 2 shows the captured SEM images of the prepared TFC membrane (M0) along with the other modified membranes

215 with MEL, GO, and a combination of MEL/GO. The membranes that were constructed underwent surface scanning to test the

216 impact of the included materials on the M0 membrane. It is obvious from Fig. 1a that the prepared TFC membrane has nodular
1 217 and leaf-like morphology which exhibits the effective development of the PA layer. Furthermore, a bridge-like construction was
2
3
4 218 detected on the surface of the prepared membrane indicated by brighter color which is known as the edge effect. This finding
5
6 219 indicated that the prepared membrane has a rough structure which is in agreement with a previously reported study (45). For
7
8
9 220 more investigation, the prepared membrane was cut into small slices, and a cross-section image was captured to visualize the
10
11 221 internal construction of the prepared PA layer. The picture obtained from the produced TFC (Fig. 2e) revealed that the M0
12
13
14 222 membranes exhibited a greater abundance of sponge-like structures in the lower layer and a thicker upper layer. This may be
15
16 223 attributed to the slower demixing rate of M0, which therefore delays the phase inversion process.
17
18
224 Regarding Mm0.1 and MG0.3 membranes, it was found that incorporation of MEL and GO resulted in the formation of
19
20
21 225 nodular and leaf like-morphology as indicated by surface images. In contrast to M0 membrane, the captured images of both
22
23 226 membranes showed that the surface has darker color with smoother surface. In addition, the leaf like-morphology was more
24
25
26 227 pronounced in MG0.3 membrane compared with Mm0.1 membrane. The cross-sectional image showed that both membranes
27
28 228 have finger- and spongy-like morphology at the top and bottom layers, respectively. Additionally, the PA layer was looser in
29
30
31 229 both membranes, and it was more prominent in the case of the Mm0.1 membrane.
32
33 230 The surface image showed that Mm0.1/G0.3 membrane has smooth surface with absence of leaf and nodular like-
34
35
231 morphology. The obtained results were in alignment with previously informed investigations that showed addition of MEL and
36
37
38 232 GO in PA layer increases smoothness of PA layer surface (46, 47). This resulted from the enhancement of heat release during
39
40 233 interfacial polymerization reaction (47). While the cross-section image of Mm0.1/G0.3 membrane revealed that the combination
41
42
43 234 of MEL and GO resulted in the construction of thinner PA layer compared with other membranes. The membrane, denoted as
44
45 235 Mm0.1/G0.3, was modified by including MEL coated with GO nanosheets. This modification resulted in the formation of finger-
46
47
48 236 like channels and an increased presence of macrovoids inside the sublayer of the membrane (48). This may be because of the
49
50 237 enhancement of pore diameter and porosity which leads to good diffusion rate for the prepared membrane.
51
52
53 238
54
55
56
57
58
59
60 (e) (f)
61
62
63
64 6
65
(a) (b)

(c) (d)

(e) (f)

1
2
3
4
5
6 (g)
7 (h)
8
9
10
11
12
13
14
15
16
17
18
19
20 239
21
22 240 Fig.2: surface and cross-section SEM images for the prepared membranes M0, Mm0.1, MG0.3 and Mm0.1/G0.3.
23
24 241 4.1.2. FTIR spectroscopy
25
26 242 The FTIR shows in Fig.3 the construction of polyamide thin film layer on PES membrane surface before and after GO
27
28 243 and MEL incorporation. In order to have a more comprehensive understanding of the peaks associated with the PA layer, the IR
29
30
31 244 spectrum of the PES substrate was further given. Fig. 3 (M0) displays the notable peaks seen at 1240 cm-1, 1486 cm-1, and 1578
32
33 245 cm-1, which correspond to the particular bonds of PES, namely Aryl-O-Aryl C-O stretching, C-C bond stretching, and C=C
34
35
246 aromatic ring stretching vibrations, respectively (25). In addition, the prepared TFC layer has characteristic function groups that
36
37
38 247 indicate successful polymerization reaction between the used reagents MPD and TMC. FTIR spectrum shows three characteristic
39
40 248 bands at around 1660, 1635 and 1076 cm−1 that are corresponding to C=O stretching of carboxylic, N-H stretching of amide, and
41
42
43 249 C-N stretching, respectively, which in alignment with a previously reported studies (49).
44
45 250 The membrane treated with melamine (Mm0.1) exhibits a specific peak in the FTIR spectrum at around 1550 cm -1, which
46
47
48 251 may be owing to the presence of the triazine ring within its structure. Furthermore, an enhanced peak at 1668 cm-1 was found,
49
50 252 indicating the effective integration of melamine into the structure of PA. Finally, Co-TFC shows characteristic broad peak at
51
52
53
253 3330 cm-1, which may be because of the stretching of the -NH group in MEL..
54
55 254 On other hand, MG0.3 membrane shows the bands at 3344 cm−1, 1734 cm−1 and 1072 cm−1 which correspond to hydroxyl,
56
57 255 carboxyl and epoxide functional groups of GO nanosheets, respectively. This outcome validates the effective integration
58
59
60 256 of GO into the polyamide layer, aligning with previously published research (44). Moreover, the free -OH and -COOH groups
61
62
63
64 7
65
257 could react with MPD present in the PA solution, resulting in the development of amide bonds. The incorporation of GO into

258 the membrane caused noticeable shifts and changes in band intensity, particularly within the range of 400 to 1600 cm −1, as

259 revealed in the FTIR spectra. These changes may be attributed to the interactions occurring between GO and the polyamide

260 active layer present in the membrane.

261 Finally, the FTIR spectrum of Mm0.1/G0.3 membrane shows some specific absorption peaks of GO and MEL at 1428,

262 1217, 1019 and 812 cm−1 along with some novel characteristic peaks. The presence of carboxylic OH groups in the GO is evident

263 from the specific wide absorption band seen at 2500-3300 cm−1. Similarly, the absorption peaks observed at 3500-3300 cm−1,

264 which may be owing to the stretching of NH2 groups in MEL, have transformed into two weak broad bands at 3459 and 3074

265 cm−1. The first absorption band may be attributed to the stretching of NH2 groups present in the grafted MEL molecules, whereas

266 the second absorption band is a result of unreacted carboxylic OH groups and intact alcoholic groups on the surface of the

267 graphene oxide. Furthermore, a novel peak at 1598 cm−1 has emerged, indicating the presence of NH bending in the recently

268 established amide groups. Furthermore, in contrast to the observed peaks resulting from the bending vibrations of the NH2

269 groups and the stretching of carboxylic carbonyl groups at 1719 and 1715 cm−1, respectively, a novel peak at 1600 cm−1 has

270 emerged. This newly observed peak can be owing to the stretching vibration of carbonyl groups that exist within the generated

271 amide groups. Therefore, it can be surmised that the process of grafting graphene oxide onto melamine was accomplished
1
2
3 272 effectively.
4
5 273
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40 274
41 275 Fig.3: FTIR spectra of M0, Mm0.1, MG0.3 and Mm0.1/G0.3 membranes.
42
43
44
276 4.1.3 Mechanical characteristics
45 277 The mechanical characteristics of the membranes that were created were examined to explore the impact of their
46
47 278 composition and identify the most suitable membrane that can withstand the mechanical strength during the process of
48
49 279 desalination. The tensile strength and elongation measurements of certain membranes were obtained by employing a mechanical
50
51 280 testing instrument at the site of membrane rupture, as seen in Fig. 4. Fig. 4 demonstrates the mechanical features of prepared
52
53 281 membranes, which reveals that the membrane Mm0.1/G0.3 has the highest tensile strength and elongation, which were 6.12
54
55
282 N/cm2 and 49.42% respectively. Results show that tensile strength of Mm0.1 and MG0.3 were close to each other and close to
56 283 tensile strength of M0. The sequence of the tensile strength and elongation is as follows M0>Mm0.1>MG0.3>Mm0.1/MG0.3.
57
58 284 According to the data shown in Table 1, it was obvious that the membranes exclusively derived from MEL or GO exhibit the
59
60 285 least mechanical characteristics. Conversely, the membranes resulting from the blending of MEL/GO (specifically denoted as
61
62
63
64 8
65
286 Mm0.1/G0.3) demonstrate the most superior mechanical features. Thus, the mechanical features of the MEL/GO mix was
287 enhanced as a result of the blending procedure, which may be because of the great surface area and strong interaction between
288 MEL/GO and PA.
60 7
Elongation %
50 Tensile Strength MPa 6

Tensile Strength [MPa]


5
40
Elongation % 4
30
3
20
2

10 1

0 0
M0 Mm 0.1 MG 0.3 Mm 0.1\G 0.3

Prepared Membranes

289
290 Fig.4: Mechanical properties for M0, Mm0.1, MG0.3 and Mm0.1/G0.3 membranes.

291 4.1.4 Contact angle and porosity


292 The contact angle measurements for the bare (M0), MEL blended (Mm0.1), GO blended (MG0.3), and MEL/GO blended

1 293 (Mm0.1/G0.3) membranes are illustrated in Table 2. The water contact angle of the PA membranes was around 88°, showing a
2 294 hydrophobic nature that poses a disadvantage for aqueous separations (50). The resistance to water infiltration inside the pores
3
4 295 is elevated, and the occurrence of fouling caused by proteins and organic contaminants is more prominent. Enhancing the
5
6 296 hydrophilic properties of the membrane matrix causes a diminution in fouling tendencies. The high hydrophilicity of GO and
7
8 297 MEL may be owing to the richness of hydroxyl, carbonyl, and amine functional groups existing in their chemical structures.
9
10 298 Therefore, the incorporation of GO and MEL into the PA membrane matrix could result in an augmentation of the membranes'
11
12 299 hydrophilicity. The measurement of static water contact angle, which serves as an indicator of the surface's hydrophilicity, has
13 300 been conducted and the outcomes are illustrated in Table 2. The contact angle of the PA membrane had the greatest value
14
15 301 compared to the other membranes investigated. The impact of additives, namely MEL, GO, and MEL/GO, on the water contact
16
17 302 angle of the membranes was investigated. Among these additives, the Mm/G membrane exhibited the lowest contact angle,
18
19 303 measuring 55.1°. The decrease in contact angle can be ascribed to the existence of amine groups covalently bonded to the surface
20
21 304 of GO. The enhancement of the hydrophilic properties of the additive resulted in a reduction of the contact angle and facilitated
22
23
305 the development of a finely porous construction, thus enhancing the permeability of the membrane.
24 306 The estimation of porosity (ε) was conducted by determining the weight and volume of the membrane sample in both dry
25
26 307 and completely hydrated states, as shown in Table 2. The introduction of MEL and GO to the membrane dope solution promoted
27
28 308 the porosity of the membrane because they form a coating on PA layer to decrease hydrophobic domains, while the Mm/G
29
30 309 membrane exhibited the highest porosity due to the increased dipolar character in Mm/G with hydroxyl, carbonyl and amine
31
32 310 group may have resulted in quicker phase separation yielding greatly porous membrane.
33
34
311 Table 2
35 312 Prepared membranes Contact angle and porosity
36 Membranes Contact angle Porosity %
37
38 M0 88 37.5
39 Mm 0.1m 75 42.5
40 MG0.3 g 70 49.5
41 Mm0.1/G0.3 55.1 54.7
42
43 313
44
45 314 4.1.5 Membrane performance measurements
46
47 315 The membranes' performance was assessed by determining the clear water flux and salt rejection of all the synthesized
48
49 316 membranes. This was achieved by subjecting a NaCl solution (2000 ppm) to filtration employing a cross-flow filtration system
50
51
317 under standard conditions of room temperature, 20 bar pressure, and a duration of 1 hour. The membrane's effective area was 17
52 318 cm2. The PA membrane (M0) had the lowest flux (10.01 LMH/bar), which is expected due to its reduced hydrophilicity and
53
54 319 porosity since it lacked any additives.
55
56 320 The introduction of additives into the membranes resulted in an augmentation of the membrane flow. The Mm0.1
57
58 321 membrane had a permeate flow that was more than 7-fold higher than that of the M0 membrane. Additionally, it demonstrated
59
60 322 an estimated 2.5-fold increase in permeate flux in comparison to the control MG0.3 membrane. The enhanced water flow seen
61
62
63
64 9
65
323 in MG0.3 membranes could be due to the presence of GO nanosheets, which operate as nanochannels inside the polymer matrix
324 or on the membrane's top surface. The use of GO nanosheets facilitates enhanced pathways for solvent transport. Furthermore,
325 the use of GO in RO membranes leads to an augmentation in both surface hydrophilicity and pore size, hence enhancing the flow
326 of clean water. The membranes that contain 0.3 wt. % of GO exhibit the highest efficiency. This may be attributed to the
327 phenomenon of GO agglomeration at higher GO concentrations, leading to the obstruction of membrane pores and subsequently
328 causing a decrease in water flow. On the contrary, the flow of Mm0.1 membrane is about 4.5 times greater than that of MG0.3
329 membranes. The hydrophilicity and porosity of Mm0.3 membranes are improved during phase inversion due to their elevated
330 nitrogen content and evenly distributed hydrophilic functions on their surface. This leads to increased membrane permeability
331 features.
332 The selection of MEL and GO at loading levels of 0.1 and 0.3wt.% for subsequent tests was based on the corresponding
333 maximum normalized recovered flux. As can be observed from Fig. 5, the addition of both MEL and GO led to raised flux, the
334 flux of Mm0.1/G0.3 membrane was almost 9 fold higher than M0 membrane and reached 88.21 LMH/bar. The membrane's
335 hydrophilicity and porosity throughout phase inversion are improved due to the high nitrogen content and evenly distributed
336 hydrophilic functions on the surface of GO. This leads to increased permeability characteristics.
337 Fig. 5 displays the salt rejection values for all produced membranes. Incorporating MEL and GO into the bare membrane
338 often caused a rise in salt rejection. The introduction of MEL and GO in the composite membranes leads to a higher degree of
339 structural compactness. This is attributed to the robust interaction between MEL and GO inside the PA layer, indicating the
340 development of a network-like structure. Thus, the elevated salt rejection may be due to this network construction. In
341 comparison, there's a small difference between Mm0.1 and MG0.3 membranes in salt rejection. The salt rejection of Mm0.1
1 342 raised by only 0.1 % from MG0.3 which can be due to the existence of nanochannels on the uppermost surface of the rejection
2
3 343 layer and increased hydrophilicity on both membranes.
4
5 344 Moreover, Mm0.1/MG0.3 membrane exhibits the superior slat rejection (96.5 %). The rise in the quantity of amino groups on
6
345 the side chains results in enhanced water flow and salt rejection. This suggests that the heightened dipolar nature of
7
8 346 Mm0.1/MG0.3 causes membranes possessing a greater quantity of smaller pores. The high porosity of Mm0.1/MG0.3
9
10 347 membranes further reinforces this conclusion. In addition, the affinity of both MEL and GO together with NMP in comparison
11
12 348 to to GO or MEL alone, which causes reduced cluster creation favoring uniform dispersion in polymer matrix, causing greater
13
14 349 quantity of smaller pores resulting in advanced permeate flux and greater salt rejection.
15
16 350
17
18 100
19 100 90
20
21 80
22 80
Permeate Flux [L/m2.h]

70
23
Salt Rejection %

24 60
25 60
26 50
27 40
28 40
29 30
30
31 20 20
32 10
33
34 0 0
35
36
37
38 Prepared Membranes
39
40
41 Salt Rejection (%) Water Flux (LMH)
42 351
43
44 352 Fig.5: Membrane performance results in terms of water flux and salt rejection.
45
46 353 For a long-term experiment, Fig.6 indicates using Mm0.1/G0.3 membrane for 360 min. The outcomes indicate that the
47
48 354 rejection approximately elevated slightly until reached to 99.6% after 360 min. While the flux decreased from 88.2 L/m2.h at the
49
50 355 beginning from the experiment to 80.5 L/m2.h after 360 min under pressure 20 bar using NaCl feed solution of concentration
51
52
356 2000 ppm. The stability of the rejection % means the chemical structure of the membrane surface undamaged under this high
53 357 pressure, while the slightly decrease in flux means no development of salt cake layer or adsorption of salts on the surface which
54
55 358 means the membranes can be easily cleaned and reused again under the same conditions.
56
57
58 359 For the rejection of divalent ions, Mm0.1/G0.3 membrane used to assess the ability of the divalent ions rejection. Fig.7
59
60 360 illustrates good separation of the divalent ions using Mm0.1/G0.3 membrane, where the removal of Mg ++ reached 95.6% and
61
62
63
64 10
65
361 Ca++ reached 89.8% , while SO4 – reached 98.8%, with permeate flux 50.8, 46.3 and76.5 L/m2.h respectively. The separation
362 behavior of divalent ions by Mm0.1/G0.3 membrane was due to the morphology and structural features of melamine and graphene that
363 promotes the diffusion of water molecules through the membrane pores exhibiting great features for selectivity and permeability.

364

365

89
100 88

87
80
86

Permeate Flux [L/m2.h]


85
Rejection %

60
84

83
40
82

20 Rejection % 81
Premeate flux 80

0 79
0 50 100 150 200 250 300 350 400
366 Time [min]

367 Fig.6: Long term experiments on the optimum prepared membrane Mm0.1/G0.3
1
2 368
3
4 369
5
6
7 90
8 100 Rejection % Flux [LMH]
9 80
10 70
Permeate Flux [L/m2.h]
11 80
12 60
13
Rejection %

14 60 50
15
16 40
17 40
30
18
19 20
20 20
21 10
22
23 0 0
24 Mg Ca SO4
25 Ions
26 370
27
28 371 Fig.7: Divalent ions separation using optimum membrane Mm0.1/G0.3
29
30
372 5. Conclusions
31
32
33 373 In this investigation, melamine (MEL), GO nanosheets and MEL grafted by graphene oxide nanosheets (MEL/GO) were
34
35 374 incorporated into the polyamide (PA) membrane to improve its permeability, hydrophilicity, and rejection performance. The
36 375 hydrophilicity of the membranes was elevated by the addition of MEL, GO, and MEL/GO, which may be because of the
37
38 376 abundance of hydrophilic NH2 and COOH groups. Consequently, the morphological characteristics, permeability, and salt
39
40 377 rejection of the membranes were enhanced by the optimal incorporation of MEL, GO, and MEL/GO. The addition of MEL, GO,
41
42 378 and GO/MDA resulted in a rise in the hydrophilicity and porosity of the membranes, which may be attributed to the higher
43
44 379 exchange rate between the solvent and non-solvent. Based on the findings from the SEM analysis, it was seen that all membranes
45
46
380 had a finger-like cross-sectional morphology. The permeability and salt rejection of the PA membrane exhibited a decline when
47 381 the concentrations of MEL and GO exceeded 0.1 wt. % and 0.3 wt. %, respectively. Hence, the optimal load of MEL/GO was
48
49 382 determined to be (0.1/0.3 wt. %) and designated as Mm0.1/G0.3. The pure water flow of the PA membrane exhibited a rise from
50
51 383 10.01 LMH/bar for the bare (M0) to 73.47 LMH/bar, 23.35 LMH/bar, and 88.21 LMH/bar for the optimal concentrations of
52
53 384 MEL, GO, and MEL/GO, respectively. The membrane with the optimal characteristics (Mm0.1/G0.3) also exhibited higher
54
55 385 performance in terms of salt separation. The findings of this work indicate that MEL, as a dendrimer ligand, is extremely suitable
56
386 for the synthesis of nanofillers with excellent hydrophilicity. These nanofillers may be effectively used in the creation of
57
58 387 nanocomposite membranes that exhibit increased performance in terms of permeability and rejection.
59
60 388 6. References
61
62
63
64 11
65
389 1. Panneerselvam B, Karuppannan S, Muniraj K. Evaluation of drinking and irrigation suitability of groundwater with special
390 emphasizing the health risk posed by nitrate contamination using nitrate pollution index (NPI) and human health risk assessment (HHRA).
391 Human and Ecological Risk Assessment: An International Journal. 2020;27(5):1324-48.
392 2. Pezeshki H, Hashemi M, Rajabi S. Removal of arsenic as a potentially toxic element from drinking water by filtration: A mini
393 review of nanofiltration and reverse osmosis techniques. Heliyon. 2023.
394 3. da Silva RI, de Souza Figueiredo KC. Incorporation of graphene oxide on thin film composite polysulfone/polyamide membranes.
395 Brazilian Journal of Chemical Engineering. 2022:1-7.
396 4. Nambi Krishnan J, Venkatachalam KR, Ghosh O, Jhaveri K, Palakodeti A, Nair N. Review of thin film nanocomposite membranes
397 and their applications in desalination. Frontiers in Chemistry. 2022;10:781372.
398 5. Elimelech M, Phillip WA. The future of seawater desalination: energy, technology, and the environment. science.
399 2011;333(6043):712-7.
400 6. Herrera-León S, Cruz C, Negrete M, Chacana J, Cisternas LA, Kraslawski A. Impact of seawater desalination and wastewater
401 treatment on water stress levels and greenhouse gas emissions: The case of Chile. Science of The Total Environment. 2022;818:151853.
402 7. Abdolahpour S, Mahdieh N, Jamali Z, Akbarzadeh A, Toliyat T, Paknejad M. Development of doxorubicin-loaded nanostructured
403 lipid carriers: preparation, characterization, and in vitro evaluation on MCF-7 cell line. BioNanoScience. 2017;7:32-9.
404 8. Hailemariam RH, Woo YC, Damtie MM, Kim BC, Park K-D, Choi J-S. Reverse osmosis membrane fabrication and modification
405 technologies and future trends: A review. Advances in colloid and interface science. 2020;276:102100.
406 9. Panagopoulos A. Brine management (saline water & wastewater effluents): Sustainable utilization and resource recovery strategy
407 through Minimal and Zero Liquid Discharge (MLD & ZLD) desalination systems. Chemical Engineering and Processing-Process
408 Intensification. 2022;176:108944.
409 10. Yaghoubi S, Babapoor A, Seyfaee A. Thermal energy optimization using salt-based phase change materials obtained from the
410 desalination of saline water. Renewable and Sustainable Energy Reviews. 2023;183:113463.
411 11. Valappil RSK, Ghasem N, Al-Marzouqi M. Current and future trends in polymer membrane-based gas separation technology: A
412 comprehensive review. Journal of Industrial and Engineering Chemistry. 2021;98:103-29.
413 12. Abdullah A, Alawee WH, Mohammed SA, Majdi A, Omara Z, Younes M. Utilizing a single slope solar still with copper heating
414 coil, external condenser, phase change material, along with internal and external reflectors—experimental study. Journal of Energy Storage.
415 2023;63:106899.
416 13. Asadollahi M, Bastani D, Musavi SA. Enhancement of surface properties and performance of reverse osmosis membranes after
417 surface modification: A review. Desalination. 2017;420:330-83.
418 14. Ju J, Choi Y, Lee S, Park C-g, Hwang T, Jung N. Comparison of pretreatment methods for salinity gradient power generation using
419 reverse electrodialysis (RED) systems. Membranes. 2022;12(4):372.
1 420 15. Lee TH, Roh JS, Yoo SY, Roh JM, Choi TH, Park HB. High-performance polyamide thin-film nanocomposite membranes
2 421 containing ZIF-8/CNT hybrid nanofillers for reverse osmosis desalination. Industrial & Engineering Chemistry Research.
3 422 2019;59(12):5324-32.
4 423 16. Rana D, Matsuura T, Kassim MA, Ismail A. Reverse osmosis membrane. Handbook of membrane separations: Chemical,
5 424 pharmaceutical, food, and biotechnological applications. 2015:35-52.
6 425 17. Ricci B, Skibinski B, Koch K, Mancel C, Celestino C, Cunha I, et al. Critical performance assessment of a submerged hybrid
7
8
426 forward osmosis-membrane distillation system. Desalination. 2019;468:114082.
9 427 18. Hafiz M, Hawari AH, Alfahel R, Hassan MK, Altaee A. Comparison of nanofiltration with reverse osmosis in reclaiming tertiary
10 428 treated municipal wastewater for irrigation purposes. Membranes. 2021;11(1):32.
11 429 19. Panagopoulos A. Process simulation and analysis of high‐pressure reverse osmosis (HPRO) in the treatment and utilization of
12 430 desalination brine (saline wastewater). International Journal of Energy Research. 2022;46(15):23083-94.
13 431 20. Halakoo E. Thin film composite membranes via layer-by-layer assembly for pervaporation separation. 2019.
14 432 21. Ismail MF, Islam MA, Khorshidi B, Tehrani-Bagha A, Sadrzadeh M. Surface characterization of thin-film composite membranes
15 433 using contact angle technique: Review of quantification strategies and applications. Advances in Colloid and Interface Science.
16
434 2022;299:102524.
17
18 435 22. Wibisono Y, Noviani V, Ramadhani AT, Devianto LA, Sulianto AA. Eco-friendly forward osmosis membrane manufacturing using
19 436 dihydrolevoglucosenone. Results in Engineering. 2022;16:100712.
20 437 23. Yousef S, Tuckute S, Tonkonogovas A, Stankevičius A, Mohamed A. Ultra-permeable CNTs/PES membranes with a very low
21 438 CNTs content and high H2/N2 and CH4/N2 selectivity for clean energy extraction applications. journal of materials research and technology.
22 439 2021;15:5114-27.
23 440 24. Marioryad H, Ghaedi AM, Emadzadeh D, Baneshi MM, Vafaei A, Lau WJ. A Thin Film Nanocomposite Reverse Osmosis
24 441 Membrane Incorporated with S‐Beta Zeolite Nanoparticles for Water Desalination. ChemistrySelect. 2020;5(6):1972-5.
25 442 25. Rastgar M, Shakeri A, Karkooti A, Asad A, Razavi R, Sadrzadeh M. Removal of trace organic contaminants by melamine-tuned
26
27
443 highly cross-linked polyamide TFC membranes. Chemosphere. 2020;238:124691.
28 444 26. Zheng D, Hua D, Cheng X, Pan J, Ibrahim AR, Hua H, et al. Polyamide composite membranes for enhanced organic solvent
29 445 nanofiltration performance by metal ions assisted interfacial polymerization method. AIChE Journal. 2023;69(2):e17896.
30 446 27. Tian J, Chang H, Gao S, Zhang R. How to fabricate a negatively charged NF membrane for heavy metal removal via the interfacial
31 447 polymerization between PIP and TMC? Desalination. 2020;491:114499.
32 448 28. Li S-L, Wu P, Wang J, Wang J, Hu Y. Fabrication of high performance polyamide reverse osmosis membrane from monomer 4-
33 449 morpholino-m-phenylenediamine and tailoring with zwitterions. Desalination. 2020;473:114169.
34 450 29. Lau W, Ismail A, Misdan N, Kassim M. A recent progress in thin film composite membrane: A review. Desalination. 2012;287:190-
35
451 9.
36
37 452 30. Wang L, Yang H, Li H, Lu P, Yu Y, Zhang X, et al. Diazotized polyamide membranes on commercial polyethylene textile with
38 453 simultaneously improved water permeance, salt rejections and anti-fouling. Desalination. 2023;549:116307.
39 454 31. Bodzek M, Konieczny K, Kwiecińska-Mydlak A. Nanotechnology in water and wastewater treatment. Graphene–the nanomaterial
40 455 for next generation of semipermeable membranes. Critical Reviews in Environmental Science and Technology. 2020;50(15):1515-79.
41 456 32. Aghili F, Ghoreyshi AA, Van der Bruggen B, Rahimpour A. Introducing gel-based UiO-66-NH2 into polyamide matrix for
42 457 preparation of new super hydrophilic membrane with superior performance in dyeing wastewater treatment. Journal of Environmental
43 458 Chemical Engineering. 2021;9(4):105484.
44 459 33. Zhang N, Song X, Jiang H, Tang CY. Advanced thin-film nanocomposite membranes embedded with organic-based nanomaterials
45
46
460 for water and organic solvent purification: A review. Separation and Purification Technology. 2021;269:118719.
47 461 34. de Oliveira TC, Ferreira FV, de Menezes BR, da Silva DM, dos Santos AS, Kawachi EY, et al. Engineering the surface of carbon-
48 462 based nanomaterials for dispersion control in organic solvents or polymer matrices. Surfaces and Interfaces. 2021;24:101121.
49 463 35. Kesavan Pillai S, Ray SS. Epoxy-based carbon nanotubes reinforced composites. IntechOpen; 2011.
50 464 36. Junaidi NFD, Othman NH, Fuzil NS, Shayuti MSM, Alias NH, Shahruddin MZ, et al. Recent development of graphene oxide-based
51 465 membranes for oil–water separation: A review. Separation and Purification Technology. 2021;258:118000.
52 466 37. Le TXH, Dumée LF, Lacour S, Rivallin M, Yi Z, Kong L, et al. Hybrid graphene-decorated metal hollow fibre membrane reactors
53 467 for efficient electro-Fenton-Filtration co-processes. Journal of Membrane Science. 2019;587:117182.
54
468 38. Matshetshe K, Sikhwivhilu K, Ndlovu G, Tetyana P, Moloto N, Tetana Z. Antifouling and antibacterial β-cyclodextrin decorated
55
56 469 graphene oxide/polyamide thin-film nanocomposite reverse osmosis membranes for desalination applications. Separation and Purification
57 470 Technology. 2021;278:119594.
58 471 39. Ahmad H, Zahid M, Rehan ZA, Rashid A, Akram S, Aljohani MM, et al. Preparation of polyvinylidene fluoride nano-filtration
59 472 membranes modified with functionalized graphene oxide for textile dye removal. Membranes. 2022;12(2):224.
60
61
62
63
64 12
65
473 40. Huang H, Ying Y, Peng X. Graphene oxide nanosheet: an emerging star material for novel separation membranes. Journal of
474 Materials Chemistry A. 2014;2(34):13772-82.
475 41. Ali ME, Wang L, Wang X, Feng X. Thin film composite membranes embedded with graphene oxide for water desalination.
476 Desalination. 2016;386:67-76.
477 42. Song X, Zhang Y, Wang Y, Huang M, Gul S, Jiang H. Nanocomposite membranes embedded with dopamine-melanin nanospheres
478 for enhanced interfacial compatibility and nanofiltration performance. Separation and Purification Technology. 2020;242:116816.
479 43. Zhu X, Xu D, Gan Z, Luo X, Tang X, Cheng X, et al. Improving chlorine resistance and separation performance of thin-film
480 composite nanofiltration membranes with in-situ grafted melamine. Desalination. 2020;489:114539.
481 44. Fathy M, Gomaa A, Taher FA, El-Fass MM, Kashyout AE-HB. Optimizing the preparation parameters of GO and rGO for large-
482 scale production. Journal of Materials Science. 2016;51:5664-75.
483 45. Chae H-R, Lee J, Lee C-H, Kim I-C, Park P-K. Graphene oxide-embedded thin-film composite reverse osmosis membrane with
484 high flux, anti-biofouling, and chlorine resistance. Journal of Membrane Science. 2015;483:128-35.
485 46. Bera A, Gol RM, Chatterjee S, Jewrajka SK. PEGylation and incorporation of triazine ring into thin film composite reverse osmosis
486 membranes for enhancement of anti-organic and anti-biofouling properties. Desalination. 2015;360:108-17.
487 47. Yu T, Wang X, Liu Z, Chen Z, Hong Z, Zhang M, et al. Structure-performance relationships between amino acid-functionalized
488 graphene quantum dots and self-cleaning nanofiltration membranes. Journal of Membrane Science. 2022;644:120068.
489 48. Safarpour M, Vatanpour V, Khataee A. Preparation and characterization of graphene oxide/TiO2 blended PES nanofiltration
490 membrane with improved antifouling and separation performance. Desalination. 2016;393:65-78.
491 49. Fathizadeh M, Tien HN, Khivantsev K, Song Z, Zhou F, Yu M. Polyamide/nitrogen-doped graphene oxide quantum dots (N-GOQD)
492 thin film nanocomposite reverse osmosis membranes for high flux desalination. Desalination. 2019;451:125-32.
493 50. Ganesh B, Isloor AM, Ismail AF. Enhanced hydrophilicity and salt rejection study of graphene oxide-polysulfone mixed matrix
494 membrane. Desalination. 2013;313:199-207.

495

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 13
65

You might also like