Blog

Polyethylene glycol-grafted graphene oxide nanosheets in tailoring the structure and reverse osmosis performance of thin film composite membrane | Scientific Reports

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

Scientific Reports volume  13, Article number: 16940 (2023 ) Cite this article peg 400

Introducing hydrophilic polymers such as polyethylene glycol (PEG) within the polyamide (PA) layer of thin film composite (TFC) membranes helps achieve high water desalination performance. Here, PEGs of different molecular weights (X: 1500, 6000, 16,000 g/mol) are effectively introduced into the PA layer of TFC membranes utilizing PEG-grafted graphene oxide (GOPX) nanosheets and their effects on the physicochemical properties and reverse osmosis (RO) performance of the thin film nanocomposite (TFN) membranes are investigated. Among the TFNs prepared the GOP16000/TFN exhibits the best performance with 68% improvement in water flux and almost constant salt rejection compared to those of the bare TFC. The influence of PEG molecular weight on the RO performance of the membranes is interpreted by different surface and bulk hydrophilicity as well as thickness and surface roughness of PA layers of GOPX/TFNs. Furthermore, TFNs with thinner and smoother PA layers and thus higher water flux are obtained by dispersing GOPXs in the aqueous phase of the PA interfacial polymerization reaction than by dispersing them in the organic phase of the reaction. Finally, the high antifouling potential of TFNs containing PEG-grafted GOs is demonstrated.

Polyamide thin-film nanocomposite (PA-TFN) membranes made by incorporating nanoparticles (NPs) into the PA thin layer are promising strategies for wastewater treatment and desalination1. The right choice of NPs with desired surface modification would ensure the fabricating of fouling-resistant TFNs possessing high water flux and salt rejection in different membrane processes, especially in reverse osmosis (RO)2,3. Among the recommended NPs for PA-TFN membranes preparation in the literature, graphene oxide (GO) nanosheets have shown remarkable efficacy in resolving the permeability-selectivity trade-off and membrane fouling issues4. In particular, the hydrophilic nature of GO allows water molecules to be embedded in its interlayer structure, thereby providing efficient transport pathways for water molecules across the polyamide (PA) layer5,6.

Furthermore, the inherent surface functional groups of GO such as hydroxyl and carboxylic acid provide suitable reaction sites for covalent modifications preventing nonspecific interactions among GOs, promoting their dispersion in the PA layer and introducing target functional groups to improve the TFN separation properties6,7. For instance, hydrophilic chitosan polysaccharide was grafted onto graphene oxide and the modified nanosheets were introduced into the top layer of a PA-TFC membrane. The resultant TFN membrane became more hydrophilic and smoother than the TFC membrane exhibiting high water flux and salt rejection5. Moreover, Zhang et al.6 prepared TFN membranes by incorporating nanofillers of p-aminophenol-modified graphene oxide (mGO) into the PA layer during the interfacial polymerization reaction. The addition of mGO into the active layer reduced the layer thickness and water contact angle by 79.1% and 30.7%, respectively, thereby displaying 24.5% and 99.7% increases in water flux and NaCl rejection, respectively. β-cyclodextrin functionalized graphene oxide nanosheets (β-CD-f-GO) were also used to fabricate TFN membranes possessing improved reverse osmosis and antifouling properties. These enhanced performances were ascribed to the role of the modified nanosheets in increasing the hydrophilicity and decreasing the surface roughness of the top PA layer7. Recently, Mahdavi and Rahimi8 synthesized TFN membranes containing GO nanoparticles modified with poly(2-(methacryloyloxy)ethyl dimethyl-(3-sulfopropyl)ammonium hydroxide) zwitterionic polymer in the top PA layer. The water flux and antifouling properties of the membranes were greatly improved while the salt rejection remained unchanged relative to the bare TFC membrane. In addition, Ma et al.9 incorporated GO functionalized with (poly(sulfobetaine methacrylate)) in the active PA layer of a TFC membrane and reported a twofold increase in the water flux while 12% and 5% compromise in the rejection of NaCl (monovalent ions) and MgSO4 (divalent ions), respectively.

Poly (ethylene glycol) (PEG) is a hydrophilic polymer widely used as an additive for morphology control and increasing the water flux and fouling resistance of single-layer water desalination membranes and support layers of TFCs10. In recent years, however, the grafting of PEG derivatives on the thin PA layer of TFC-RO membranes has been reported by some authors11,12,13,14,15,16. For example, Van Wegner et al.12 grafted poly (ethylene glycol) di-glycidyl ether (PEGDE) onto the surface of a RO membrane. Although the modified membranes with PEGDE generally exhibited promoted fouling resistance against charged surfactants and emulsions, their water flux was decreased due to this surface modification. Kang et al.13 utilized a chemical reaction between unreacted acyl halide groups on the PA surface of a TFC membrane and amino groups of aminopolyethylene glycol monomethyl ether (MPEG-NH2) to introduce poly(ethylene glycol) chains on the membrane active layer and the modified membrane sowed enhanced surface hydrophilicity and thus improved organic fouling resistance. The same group reported elsewhere that grafting PEGDE on the top surface of TFC-RO membranes renders fouling-resistant membranes against cationic surfactants and proteins14.

Photo-induced graft polymerization and plasma polymerization have also been used to incorporate different PEG-based moieties on the active layer of TFC membranes to develop membranes with higher fouling resistance and separation properties15,16. Nevertheless, the direct grafting of PEG-based polymers onto the membrane surface would find limited industrial applications due to the need for multi-step reactions under harsh conditions. Furthermore, most of the TFC membranes having top surfaces grafted with polymers suffer from declined permeance as a result of the direct grafting approach17. On the other hand, the embedding of hydrophilic PEG chains, during the IP reaction, within the top PA layer of TFC membranes may facilitate water transport across the layer and improve the membrane flux. To achieve this goal, immobilizing PEG chains on a substrate, e.g. the GO substrate, reduces the risk of leaching these highly hydrophilic chains during the interfacial polymerization reaction and water purification process. Furthermore, utilizing this strategy will be useful to improve the dispersion of GO nanosheets in the top PA layer of the membranes.

Herein, we utilized GO nanosheets as substrates for PEG chains in the top layers of PA-TFC membranes to exploit both the advantages of nanosheets and hydrophilic polymer for water desalination purposes, simultaneously. Furthermore, to study the effect of the molecular weight (MW) of PEG on the RO performance of TFC membranes, PEG chains with different MWs, including 1500, 6000 and16000 g/mol have been grafted onto GO nanosheets and introduced to the top layers of the PA-TFC membranes through the aqueous or organic phases of the interfacial polymerization reaction during the preparation of the membranes. Thus, the modified nanoparticles were successfully concentrated on the top layer of TFN membranes during the interfacial polymerization process. The results reveal the determining role of the MW of the grafted PEG chains on the morphology and separation performance of TFN membranes in an RO process.

Graphite (99.55%, particle size < 50 µm), sulfuric acid (H2SO4, 98%), potassium permanganate (KMnO4), dimethylformamide (DMF, 99.8%), polyethylene glycols (PEG, Mw = 1500, 6000, 16,000 g/mol), tetrahydrofuran (THF), triethylamine (TEA), sodium carbonate (Na2CO3), sodium hydroxide (NaOH) and sodium dodecyl sulfate (SDS, > 99%) were purchased from Merck. Thionyl chloride (SOCl2) was purchased from CDH, India. Hydrogen peroxide (H2O2, 35%), hydrochloric acid (HCl, 37%) and n-hexane were obtained from Dr. Mojallali CO, Iran. Ethanol (C2H5OH, 96%) was obtained from Taghtirkhorasan CO., Iran. Polysulfone microporous sublayer reinforced with a polyester non-woven support with a molecular weight cut-off (MWCO) of 5400 Da, was supplied by the Sharif Membrane Technology Center (SMTC), Iran. Trimesoyl chloride (TMC), and m-phenylenediamine (MPD) were obtained from Sigma Aldrich.

GO nanosheets were synthesized from graphite powders by modified Hummers’ method as reported in our previous work18. For the GO chlorination, first, 200 mg of the nanosheets were added to 10 mL DMF under ultrasonication to obtain a stable DMF/GO suspension. Then, 40 mL of SOCl2 was added to the suspension followed by refluxing for 24 h at 70 °C. Finally, the mixture was washed five times by centrifuging to remove excess SOCl2 and the chlorinated GO nanosheets (GO-Cl) were collected. Subsequently, 2 g of PEG of a certain MW and 5 mL of TEA were added to the chlorinated GOs and the mixtures were stirred for 3 days at 120 °C19. The final suspensions were then poured into ethanol and dried at 60 °C in a vacuum oven to obtain various PEGylated GOs, designated as GOPX, where X: 1, 2 and 3 refer to PEGs with MWs of 1500, 6000 and 16,000 g/mol, respectively. Figure 1 represents the chemical reactions for the preparation of GOPXs.

Chemical routes to the synthesis of GOPXs.

TFN membranes were prepared by in situ IP reaction of 2 wt% MPD in water with 0.2 wt% TMC in n-hexane atop the polysulfone substrate. The aqueous phase containing 2 wt% Na2CO3, 0.5 wt% SDS and 0.004 wt% GO or each GOPX was ultrasonicated for 1 h to obtain a stable suspension before the IP reaction. The aqueous suspension was poured on the polysulfone substrate for 2 min and then the substrate was held vertically to remove the excess MPD. The MPD-impregnated substrate was contacted with the organic phase containing TMC for 2 min and a PA thin film was formed on the substrate. Finally, the membranes were annealed in an oven for 10 min at 80 °C to complete the IP reaction and kept in deionized water before testing20. To investigate the effect of dispersing the nanosheets in the organic phase, instead of the aqueous phase, on the morphology and final performance of the membranes, GOP3 was also added in the organic phase for the IP reaction as explained above. The prepared membranes are designated according to Table 1.

X-ray diffraction (XRD) patterns of the nanosheets were obtained by Philips diffractometer with Cu-Kα radiation (λ = 1. 45 Å) at 40 kV/30 mA in the scanning range 2θ of 5° to 60°. Thermogravimetric analysis (TGA) was performed on an STA504 instrument (Bahr, Germany) under an argon atmosphere in the temperature range of 50 °C to 650 °C with a heating rate of 10 \(^\circ \) C/min. Modified nanoparticles were characterized by Fourier-transform infrared spectroscopy (FTIR, Perkin Elmer spectrum version 10.03.06, USA). The polyamide layer of the membranes was studied by attenuated total reflection Fourier-transform infrared spectroscopy (ATR-FTIR, Perkin Elmer spectrum version 10.03.06, USA) over a wave number range of 800–4000 cm−1. The surface topography of the membranes was examined by an atomic force microscope (AFM, Agilent Technologies, Inc., Santa Clara, CA), in a tapping mode. A field-emission scanning electron microscope (FE-SEM, Philips XL-30) was utilized to investigate the membrane’s morphology. The hydrophilicity of the membrane surfaces was characterized by water contact angle measurements using a digital microscope (WCA, Dino-lite instrument) at room temperature.

The water desalination performance of the membranes in the RO process was measured by a crossflow filtration device equipped with a circular cell with an effective area of 12.56 cm2. The measurements were performed using 2000 ppm NaCl solution at 16 bar pressure. The feed flow rate was adjusted to 70 L/h. The retentate was returned to the feed reservoir by a circulation system during the separation tests.

The permeate flux (F) was obtained by Eq. (1):

where V is the permeate volume, t is the sample collection time and A is the effective membrane area.

Furthermore, the salt rejection (R) was calculated by Eq. (2):

where Cf and Cp are the concentrations of NaCl in the feed and permeate, respectively20. The Cf and Cp values were determined by a digital conductivity meter (Oakton CON 110).

The XRD patterns of GO and PEGylated GOs are shown in Fig. 2. In the pristine GO pattern, the appearance of a peak at 2\(\theta =10.8^\circ \) with a d-spacing of 0.83 nm is attributed to the reflection of the (001) plane21. and points to the oxygenated functional groups of GO. The peak at \(2\theta \) = 10.8\(^\circ \) is disappeared in the XRD patterns of GOPXs and the modified nanosheets show additional diffuse peaks at 2\(\theta =\) 24\(^\circ \) . The disappearance of the XRD peak at 2θ = 10.8\(^\circ \) can be ascribed to the disordering and increase in interlayer spacing between GO nanosheets due to the grafting of PEG chains. Furthermore, the weak and broad diffraction peaks at 2\(\theta =\) 24\(^\circ \) in the GOPxs patterns are related to the amorphous structures of PEG chains22 because the XRD pattern of the crystalline PEG is characterized by two sharp and strong 2θ peaks at 19.1° and 23.6°, which are assigned to the typical planes of (120) and (112) of PEG23,24. These observations confirm the amorphous disordered state and random direction of the graphene oxide layers as a result of the grafting process25.

XRD patterns of GO and GOPXs.

Figure 3a exhibits FT-IR spectra of GO, GO-Cl and different GOPX nanosheets. The FT-IR spectra of PEGs with different molecular weights are also shown in Fig. 3b for comparison. The peak associated with the C=O of GO at 1740 cm-1 is blue shifted to 1754 cm−1 in the spectrum of the GO-Cl representing the acyl chloride formation after the chlorination reaction26. The GO-Cl also shows stretching bands at 540 cm−1, 1032 cm−1 and 1341 cm−1 corresponding to C–Cl formation in GO26. GOPXs show a peak at 1750 cm−1 corresponding to the stretching vibrations of C=O bands of their ester groups due to the reaction between hydroxyl groups of PEGs and CO–Cl of GO-Cls27. In addition, the peaks at 2923 cm−1 and 2853 cm−1 can be ascribed to C−H stretching vibrations of CH2 groups of PEG. Absorption bands at 1399 cm−1 and 1385 cm−1 in the spectra of GOPXs are representatives of the bending vibrations of the C−H of methylene groups of PEGs grafted on GOs27.

FT-IR spectra of (a) GO, GO-Cl and GOPXs and (b) PEGs with different molecular weights.

Figure 4 shows the TGA curves of the prepared nanosheets. The pristine GO exhibits a significant weight loss (43.2%) in the range of 150−210 °C, due to the decomposition of adsorbed water and unstable oxygen-containing functional groups28. For GOPXs, with a slower mass loss rate, such a weight loss is not observed at about 200 °C, which may be ascribed to the grafted PEG molecules that cause a reduction in the thermal transport in the nanosheets. This also results in much higher thermal stability of the PEGylated GOs, in comparison with pristine GO, which retains 60–70% of their weight at 650 °C. The thermal stability of the modified GOs follows the order of GOP1 > GOP2 > GOP3, at temperatures lower than 400 °C. However, the reverse holds at temperatures above 400 °C, due to the more thermal stability of PEGs with higher MWs at higher temperatures. Actually, the polar C-O bond is more unstable and breaks easily compared to the nonpolar C–C bond29. Below 400 ℃, the weaker C-O bond of the PEG chains starts to degrade. Thus, PEG with a higher MW (16,000 g/mol), which has more C-O bonds than PEG with a lower MW (1500 g/mol) shows lower thermal stability below 400 ℃. On the other hand, above 400 ℃, the higher number of C–C bonds of the higher MW PEG makes it more thermally stable. All the results above reveal that PEGs of different MWs have been successfully grafted onto GOs via ester bonds utilizing SOCl2 as an intermediate reagent30.

TGA thermograms of GO and GOPXs.

The ATR-IR spectra of the TFC, GO/TFN and GOPXs/TFN membranes are shown in Fig. 5. All the membranes demonstrate characteristic peaks of the top PA layer at 1660 cm−1 and 1547 cm−1 due to C=O stretching of -NCHO (amide I) and N–H (amide II), respectively31. In GOPXs/TFN membranes, two peaks at 2918 cm−1 and 2849 cm−1 can be attributed to the asymmetric and symmetric C–H stretching bands of PEGs, respectively. Moreover, weak broad bands at about 3300 cm−1 in the spectra of TFN and membranes are assigned to the stretching vibrations of hydroxyl groups of GO sheets (for GO/TFN) and the combined stretching vibrations of hydroxyl groups of both GO and PEG (for GOPXs/TFNs)30,31,32,33. The small peaks at 1710 cm−1 in the spectra of TFC and TFN membranes may be ascribed to carboxylic acid C=O stretch resulting from the hydrolysis of some of the acyl chlorides of TMC molecules20. This partial hydrolysis of acyl chlorides creates free carboxylic acid groups and impedes complete IP reaction between TMC and MPD monomers thus reducing the cross-linking density of the PA layer. Therefore, the degree of cross-linking of the PA layers can be estimated and compared based on the ATR-FTIR spectra, by calculating the ratio of absorption intensities of amide I (at 1660 cm−1) to carboxylic acid C=O stretch (at 1710 cm−1) groups: the higher the ratio the larger the cross-linking density34. The amide I/carboxylic values for the TFC and TFN membranes are listed in Table 2. As can be seen, the addition of GOPx nanosheets causes a reduction in the crosslinking density of the TFC and TFN membranes. This could be due to the role of the grafted nanosheets with PEG molecules as barriers, which decrease the reaction probability between MPD and TMC monomers and thus increase the possible hydrolysis of acyl chlorides during the IP reaction. In addition, one could expect that annealing of the membranes at 80 °C would result in the evaporation of the residual water and increasing dehydration of carboxylic groups leading to an increase in the PA layer crosslinking density. However, the hydrophilic natures of GO and especially GOPx nanosheets prevent efficient water evaporation, which eventually results in the further hydrolysis of acyl chloride groups of TMC and the reduction of the cross-linked portion of the PA layers of GO/TFN and GOPx/TFNs as compared with that of the PA layer of the TFC membrane.

ATR-FTIR spectra of TFC and TFN membranes.

Figure 6 represents the water contact angle (WCA) of the prepared membranes. By introducing the pristine GO, the WCA of the TFC membrane is reduced from 78° to 70°, indicating an improvement in the hydrophilicity of the TFC membrane due to the oxygen-containing functional groups of GO. Furthermore, GOPXs/TFN membranes exhibited lower contact angles than TFC and GO/TFN. The hydrophilicity of GOPx/TFNs varies in the order of GOP1/TFN > GOP2/TFN > GOP3,Org/TFN > GOP3/TFN.

Water contact angles (WCA) of the prepared membranes.

The incorporation of GOPX nanosheets increases the hydrophilic groups of the top layers of the TFNs, which form hydrogen bonds with water molecules and hydration layers on the surfaces of the membranes35. It is seen that decreasing the MWs of the grafted PEGs increases the hydrophilicity of the GOPXs containing TFNs. Dharmaratne et al.36 recently showed that PEGs of higher MWs exhibit higher degrees of hydrophilicity. Therefore, GOP3 would be more hydrophilic than GOP2 and GOP1, respectively. When dispersed in the aqueous phase, the more hydrophilic GOP3 nanosheets possess more tendency than the other two PEGylated GOs to be in the bulk aqueous phase, while GOP2 and particularly, GOP1 tend to move partially towards the phase surface. Consequently, as Fig. 7 shows after the interfacial polymerization reaction, more hydrophilic groups are oriented towards the top layer surface of GOP1/TFN and GOP2/TFN than the surface of GOP3/TFN. This would render GOP1/TFN and GOP2/TFN surfaces more hydrophilic than the GOP3/TFN surface.

Schematic illustration of how PEG MW affects the distribution of GOPX nanosheets in the PA layer of the GOPX/TFN membranes.

Interestingly, GOP3,Org/TFN shows a slightly lower WCA and thus higher surface hydrophilicity than that of the GOP3/TFN. This can be ascribed to the diffusion of hydrophilic GOP3 nanosheets toward the aqueous phase of IP, during the IP reaction, and their potential trapping on the top layer surface of the GOP3,Org/TFN when they have already been dispersed in the organic phase. Figure 8 schematically compares the distribution of GOP3 nanosheets in the IP reaction zones of GOP3/TFN and GOP3,Org/TFN membranes.

The distribution of GOP3 nanosheets in the IP reaction zones of (a) GOP3/TFN and (b) GOP3,Org/TFN membranes.

The top surface and cross-sectional SEM images of the fabricated membranes are shown in Figs. 9 and 10, respectively. The TFC membrane exhibits a typical ridge-valley morphology caused by the IP reaction37. Meanwhile, the surface morphologies and the top layer thicknesses of the membranes are affected by the incorporation of GO and GOPXs in the PA active layer.

Surface SEM images of the prepared membranes.

Cross-sectional SEM images of the prepared membranes.

The layer thicknesses were measured according to the cross-sectional SEM images (Fig. 10) and are reported in Table 3. Moreover, Fig. 11 shows two and three-dimensional AFM images of the TFC and TFN membranes. Table 3 also lists the average roughness (Ra) and the root means square roughness (Rq) values derived from the AFM images. Introducing GO nanosheets (modified or pristine) into the top PA layer decreases the layer thickness and surface roughness. As the IP reaction is diffusion-controlled in nature, the dispersion of GO and GOPXs in the aqueous phase could hinder the diffusion of the aqueous monomer (MPD) toward the organic phase during the reaction due to the creation of a tortuous path for the diffusion of MPD molecules deep into the organic phase. The nanoparticles also increase the viscosity of the aqueous phase. For these two reasons, the formation of ridge-valley structures on the PA surface as well as the PA layer thickness is reduced8,35. In addition, altering the IP reaction rate due to the possible interactions of oxygen-containing functional groups of GO and GOPXs with MPD could be another reason for the observed morphologies of the TFNs9,31,33. Moreover, grafting PEGs onto GOs improves the nanosheet hydrophilicity and their dispersion in the aqueous phase. This can further hinder the diffusion of MPD molecules toward the interface and cause more reductions in the surface roughness and layer thickness of the GOPXs/TFN membranes (Table 3).

Two- and three-dimensional AFM images of the prepared TFC and TFN membranes.

As expected, PEGs of higher MWs are more effective in increasing the distance between individual nanosheets and improving their dispersions in the aqueous phase of the IP reaction. Moreover, diffusion of MPD is further prevented by longer PEG chains. Consequently, the hydrophilicity and dispersion of the nanosheets would be more enhanced if the PEG MW was increased and therefore, the roughness of the PA surface and thickness of the PA layer of GOP3/TFN are lower than those of the GOP1/TFN. As Figs. 10 and 11 and Table 3 show dispersing GOP3 nanosheets in the organic phase of the IP reaction results in a TFN membrane (GOP3,Org/TFN) with higher surface roughness and PA layer thickness than those of the TFN prepared by the IP reaction between GOP3-containing aqueous phase and organic phase (GOP3/TFN). The hindrance effect of GOP3 nanosheets on the TMC diffusion in the organic phase is considered to decrease the TMC concentration at the interface and the IP reaction rate. This delays the formation of the diffusion barrier and thus more MPD molecules would diffuse deep into the organic phase, increasing the surface roughness and PA layer thickness30.

In addition, hydrophilic GOP3 nanosheets in the organic phase increase the miscibility of the organic and aqueous phases of the IP reaction, thereby increasing the reaction zone width, surface roughness and PA layer thickness. Moreover, as compared with the mostly embedded nanosheets inside the PA layer of the GOP3/TFN membrane, more GOP3 nanosheets would be close to the outer surface of the GOP3,Org/TFN membrane. The uneven distribution of the nanosheets on the surface of the GOP3,Org/TFN membrane can also lead to a rougher surface than the surface of the GOP3/TFN one38.

Figure 12 shows the water flux and NaCl rejection of the prepared membranes. The bare TFC membrane shows the lowest water flux among the membranes. The water flux of the TFC membrane, 28.5 LMH, is enhanced by 28.8%, 44%, 51%, 68.4% and 61.4% in GO/TFN, GOP1/TFN, GOP2/TFN, GOP3/TFN and GOP3,Org/TFN, respectively (Fig. 12 and Table 4). On the other hand, the salt rejection of the TFC membrane remained almost constant, more than 94%, with the addition of GO and GOPXs in the PA layer. This indicates that although the improvements in water flux of GO-containing membranes may be partly due to the decrease in their crosslinking density compared to that of the bare TFC (Table 2), the platelet nature of GO nanosheets and their role in creating tortuous paths for salt molecules retain the membrane's selectivity at a high level. Another possible reason for increasing the water flux by embedding the GOPx nanosheets in the PA layers could be the introduction of hydrophilic groups on the surface and in the bulk of the polyamide layer of the GOPx-containing TFN membranes as confirmed by contact angle measurements (Fig. 6). The GOPx nanosheets on the surface and in the bulk of the polyamide layer could help dissolve water molecules on the membrane surface and improve the diffusion rate of water molecules passing through the TFN membrane, respectively8.

Water flux and salt rejection of the prepared membranes.

Furthermore, as discussed in the previous section, the embedding of GO and GOPXs significantly reduced the thickness of the top layer of TFN membranes decreasing the resistance to mass transfer and increasing the water flux. Moreover, the presence of hydrophilic PEGs with different MWs on GOPXs nanosheets increases the d-spacing between GO nanosheets and creates water-attractive channels through which water can flow even more easily across the membranes33.

Figure 12 shows that the increase in the MW of PEGs grafted on the GOs enhances the water flux of the membranes, i.e., the flux of GOPX/TFNs varies in the order GOP3/TFN > GOP2/TFN > GOP1/TFN. As Table 3 shows, increasing the PEG MW was accompanied by decreases in the PA layer thicknesses of GOPX/TFNs membranes. This can be the main reason behind the observed increase in the water flux of TFNs containing GOs grafted with PEGs of higher MW. However, the above trend of the water flux of GOPX/TFNs membranes seems to be contradictory with the contact angle results (Fig. 6), as one expected that a TFN membrane with a lower contact angle and higher surface hydrophilicity should exhibit higher water flux. This discrepancy can be resolved by the fact that GOs grafted with PEGs of higher MW further enhance the bulk hydrophilicity, as discussed earlier, and provide preferential paths for water molecules across the membranes. Consequently, combined effects of lower PA layer thickness and higher PA bulk hydrophilicity contributed to the observed enhancements of the water fluxes of GOPX/TFNs membranes. It is also seen from Fig. 12 that the water flux of the GOP3,Org/TFN membrane is lower than that of the GOP3/TFN membrane. This can be ascribed to the lower PA bulk hydrophilicity and higher thickness of the PA selective layer of the former membrane, as deduced from Fig. 6 and Table 3, respectively, leading to higher resistance against water flow through the membrane.

Time dependency of water flux of membranes is an important indicator of the membranes' propensity to fouling33. According to Fig. 13, the fluxes of all the TFC and TFN membranes are decreased with the time passed from the filtration process which indicates the occurrence of varying degrees of fouling during the water filtration. As listed in Table 4, the absolute value of the trend line slope of the flux-time profiles of the membranes increases in the order of GOP3/TFN < GOP1/TFN ≅  GOP2/TFN < GOP3,Org/TFN < GO/TFN < TFC. In other words, bare TFC with the highest permeance reduction rate among the membranes would be more prone to fouling, compared to GOPx/TFN and GOP3,Org/TFN membranes, which show the lowest flux reduction rate and thus the greatest anti-fouling properties. These potential anti-fouling properties can be attributed to the high surface hydrophilicity and low surface roughness (Table 4) of GOPx/TFN and GO/TFN membranes, which would lead to the deposition of a lesser amount of hydrophobic foulants on the membrane’s surfaces.

Water flux-time profiles of the prepared membranes (dotted lines indicate trend lines).

Table 5 compares the water permeance and NaCl rejection of GOPX-containing membranes with those of other GO-containing TFNs in the literature. As can be seen, GOPX/TFN membranes outperform most of the lab-fabricated relevant TFN RO membranes in terms of water permeance, although the use of hybrid nanoparticles, e.g., TiO2-GO nanoparticles, led to a high water permeance of 6.2 LMH/bar44. Furthermore, GOPX/TFNs show acceptable NaCl rejection, which is higher than those of commercial membranes, e.g. SWHR (FilmTec Corp. Edina, MN) with a NaCl rejection of 92%45.

This study aimed at exploiting simultaneously the advantages of both GO nanosheets and PEG polymer in the PA layer of TFC RO membranes. GO nanosheets were grafted with PEGs of different MWs (GOPX) via a facile grafting approach and successfully incorporated in the thin PA layer of TFC membranes. GOPX nanosheets improved water flux while maintained high salt rejection of the bare TFC and pristine GO-containing TFN. Among GOPXs, GOP3 nanosheets endowed TFNs with the highest PA bulk hydrophilicity, the lowest PA layer thickness and surface roughness and thus the best performance in the RO process. Moreover, it was shown that the TFN prepared by the addition of GOP3 nanosheets to the aqueous phase of the IP reaction possessed higher water flux than the TFN prepared by adding the nanosheets to the organic phase. Finally, the high antifouling potential of GOPX/TFN membranes was confirmed by analyzing and comparing the slopes of the water flux-time profiles of the prepared membranes.

The datasets used and/or analyzed in the current study are available from the corresponding author on reasonable request.

Lau, W. J. et al. A review on polyamide thin film nanocomposite (TFN) membranes: History, applications, challenges and approaches. Water Res. 80, 306–324. https://doi.org/10.1016/j.watres.2015.04.037 (2015).

Article  CAS  PubMed  Google Scholar 

Zhang, L., Zhang, M., Lu, J., Tang, A. & Zhu, L. Highly permeable thin-film nanocomposite membranes embedded with PDA/PEG nanocapsules as water transport channels. J. Memb. Sci. 586, 115–121. https://doi.org/10.1016/j.memsci.2019.05.065 (2019).

Article  ADS  CAS  Google Scholar 

Esfahani, M. R. et al. Nanocomposite membranes for water separation and purification: Fabrication, modification, and applications. Sep. Purif. Technol 213, 465–499. https://doi.org/10.1016/j.seppur.2018.12.050 (2019).

Liao, Z., Zhu, J., Li, X. & Van der Bruggen, B. Regulating composition and structure of nanofillers in thin film nanocomposite (TFN) membranes for enhanced separation performance: A critical review. Sep. Purif. Technol 266, 118567. https://doi.org/10.1016/j.seppur.2021.118567 (2021).

Hegab, H. M., Wimalasiri, Y., Ginic-markovic, M. & Zou, L. Improving the fouling resistance of brackish water membranes via surface modification with graphene oxide functionalized chitosan. Desalination 365, 99–107. https://doi.org/10.1016/j.desal.2015.02.029 (2015).

Zhang, Y. et al. Thin-film nanocomposite reverse osmosis membranes with enhanced antibacterial resistance by incorporating p-aminophenol-modified graphene oxide. Sep. Purif. Technol. 234, 116017. https://doi.org/10.1016/j.seppur.2019.116017 (2020).

Matshetshe, K. et al. Antifouling and antibacterial β-cyclodextrin decorated graphene oxide/polyamide thin-film nanocomposite reverse osmosis membranes for desalination applications. Sep. Purif. Technol. 278, 119594. https://doi.org/10.1016/j.seppur.2021.119594 (2021).

Mahdavi, H. & Rahimi, A. Zwitterion functionalized graphene oxide/polyamide thin film nanocomposite membrane: Towards improved anti-fouling performance for reverse osmosis. Desalination 433, 94–107. https://doi.org/10.1016/j.desal.2018.01.031 (2018).

Ma, W. et al. Zwitterion-functionalized graphene oxide incorporated polyamide membranes with improved antifouling properties. Langmuir 35, 1513–1525. https://doi.org/10.1021/acs.langmuir.8b02044 (2019).

Article  CAS  PubMed  Google Scholar 

Wu, Y., Zhu, H., Feng, L. & Zhang, L. Effects of polyethylene glycol on the structure and filtration performance of thin-film PA-Psf composite forward osmosis membranes. Sep. Sci. Technol. 51, 862–873. https://doi.org/10.1080/01496395.2015.1119846 (2015).

Castrillón, S. R., Lu, X. & Shaffer, D. L. Amine enrichment and poly (ethylene glycol) (PEG) surface modi fi cation of thin- fi lm composite forward osmosis membranes for organic fouling control. J. Memb. Sci. 450, 331–339. https://doi.org/10.1016/j.memsci.2013.09.028 (2014).

Van Wagner, E. M., Sagle, A. C., Sharma, M. M., La, Y. H. & Freeman, B. D. Surface modification of commercial polyamide desalination membranes using poly(ethylene glycol) diglycidyl ether to enhance membrane fouling resistance. J. Memb. Sci. 367, 273–287. https://doi.org/10.1016/j.memsci.2010.11.001 (2011).

Kang, G., Liu, M., Lin, B., Cao, Y. & Yuan, Q. A novel method of surface modification on thin-film composite reverse osmosis membrane by grafting poly(ethylene glycol). Polymer (Guildf). 48, 1165–1170. https://doi.org/10.1016/j.polymer.2006.12.046 (2007).

Kang, G., Yu, H., Liu, Z. & Cao, Y. Surface modification of a commercial thin film composite polyamide reverse osmosis membrane by carbodiimide-induced grafting with poly(ethylene glycol) derivatives. Desalination 275, 252–259. https://doi.org/10.1016/j.desal.2011.03.007 (2011).

Hong Anh Ngo, T., Mori, S. & Thi Tran, D. Photo-induced grafting of poly(ethylene glycol) onto polyamide thin film composite membranes. J. Appl. Polym. Sci. 134, 1–9. https://doi.org/10.1002/app.45454 (2017).

Zou, L. et al. Surface hydrophilic modification of RO membranes by plasma polymerization for low organic fouling. J. Memb. Sci. 369, 420–428. https://doi.org/10.1016/j.memsci.2010.12.023 (2011).

Chae, H. R., Lee, J., Lee, C. H. I., Kim, C. & Park, P. K. Graphene oxide-embedded thin-film composite reverse osmosis membrane with high flux, anti-biofouling, and chlorine resistance. J. Memb. Sci. 483, 128–135. https://doi.org/10.1016/j.memsci.2015.02.045 (2015).

Article  ADS  CAS  Google Scholar 

Koolivand, H., Sharif, A. & Kashani, M. R. Functionalized graphene oxide/polyimide nanocomposites as highly CO2-selective membranes. J. Polym. Res. https://doi.org/10.1007/s10965-014-0599-9 (2014).

Kazempour, M., Namazi, H., Akbarzadeh, A. & Kabiri, R. Synthesis and characterization of PEG-functionalized graphene oxide as an effective pH-sensitive drug carrier. Artif. Cells Nanomed. Biotechnol. 47, 90–94. https://doi.org/10.1080/21691401.2018.1543196 (2019).

Article  CAS  PubMed  Google Scholar 

Azizi, J. & Sharif, A. Optimization of water flux and salt rejection properties of polyamide thin film composite membranes. J. Appl. Polym. Sci. 137, 1–12. https://doi.org/10.1002/app.48858 (2019).

Asghari, M., Saadatmandi, S. P. & Parnian, M. J. Polypyrrole-aided surface decoration of graphene oxide nanosheets as fillers for poly(ether-b-amid) mixed matrix membranes to enhance CO2 capture. Int. J. Energy Res. 45, 10843–10857. https://doi.org/10.1002/er.6567 (2021).

Tu, J. et al. Latent heat and thermal conductivity enhancements in polyethylene glycol/polyethylene glycol-grafted graphene oxide composites. Adv. Compos. Hybrid Mater. 2, 471–480. https://doi.org/10.1007/s42114-019-00083-x (2019).

Khanna, L. & Verma, N. K. PEG/CaFe2O4 nanocomposite: Structural, morphological, magnetic and thermal analyses. Phys. B Condens. Matter. 427, 68–75 (2013).

Article  ADS  CAS  Google Scholar 

Gao, H., Chen, Y., Ma, H., Zeng, J. & Li, G. Preparation and characterization of Hesperidin-PEG 6000 complex. J. Chem. Soc. Pak. 36, 5 (2014).

Nethravathi, C. & Rajamathi, M. Chemically modified graphene sheets produced by the solvothermal reduction of colloidal dispersions of graphite oxide. Carbon N. Y. 46, 1994–1998. https://doi.org/10.1016/j.carbon.2008.08.013 (2008).

Nonahal, M. et al. Design, preparation, and characterization of fast cure epoxy/amine-functionalized graphene oxide. Nanocomposites 3, 1–12 (2017).

Xu, J. Z. et al. Highly enhanced crystallization kinetics of poly(l-lactic acid) by poly(ethylene glycol) grafted graphene oxide simultaneously as heterogeneous nucleation agent and chain mobility promoter. Macromolecules 48, 4891–4900. https://doi.org/10.1021/acs.macromol.5b00462 (2015).

Article  ADS  CAS  Google Scholar 

Alkhouzaam, A., Qiblawey, H. & Khraisheh, M. Polydopamine functionalized graphene oxide as membrane nanofiller: Spectral and structural studies. Membranes 1, 86. https://doi.org/10.3390/membranes11020086 (2021).

Blanksby, S. J. & Ellison, G. B. Bond dissociation energies of organic molecules. Acc. Chem. Res. 36, 255–263. https://doi.org/10.1021/ar020230d (2003).

Article  CAS  PubMed  Google Scholar 

Zhang, S., Xiong, P., Yang, X. & Wang, X. Novel PEG functionalized graphene nanosheets: Enhancement of dispersibility and thermal stability. Nanoscale 3, 2169–2174. https://doi.org/10.1039/C0NR00923G (2011).

Article  ADS  CAS  PubMed  Google Scholar 

Abu Tarboush, B. J., Rana, D., Matsuura, T., Arafat, H. A. & Narbaitz, R. M. Preparation of thin-film-composite polyamide membranes for desalination using novel hydrophilic surface modifying macromolecules. J. Memb. Sci. 325, 166–175. https://doi.org/10.1016/j.memsci.2008.07.037 (2008).

He, L. et al. Promoted water transport across graphene oxide-poly(amide) thin film composite membranes and their antibacterial activity. Desalination 365, 126–135. https://doi.org/10.1016/j.desal.2015.02.032 (2015).

Xia, S., Yao, L., Zhao, Y., Li, N. & Zheng, Y. Preparation of graphene oxide modified polyamide thin film composite membranes with improved hydrophilicity for natural organic matter removal. Chem. Eng. J. 280, 720–727. https://doi.org/10.1016/j.cej.2015.06.063 (2015).

Kamali, M., Gharibi, F. & Sharif, A. A systematic study on the effects of synthesis conditions of polyamide selective layer on the CO2/N2 separation of thin film composite polyamide membranes. J. Appl. Polym. Sci. 138, 1–13. https://doi.org/10.1002/app.50927 (2021).

Shahkaramipour, N., Tran, T. N., Ramanan, S. & Lin, H. Membranes with surface-enhanced antifouling properties for water purification. Membranes (Basel). 7, 1–18. https://doi.org/10.3390/membranes7010013 (2017).

Dharmaratne, N. U., Jouaneh, T. M. M., Kiesewetter, M. K. & Mathers, R. T. Quantitative measurements of polymer hydrophobicity based on functional group identity and oligomer length. Macromolecules 51, 8461–8468. https://doi.org/10.1021/acs.macromol.8b01747 (2018).

Article  ADS  CAS  Google Scholar 

Samieirad, S., Mousavi, S. M. & Saljoughi, E. Novel chlorine resistant thin-film composite forward osmosis membrane: Preparation and performance evaluation in the regeneration of MEG aqueous solution. Chem. Eng. Res. Des. 177, 554–568. https://doi.org/10.1016/j.cherd.2021.11.013 (2022).

Kadhom, M., Yin, J. & Deng, B. A thin film nanocomposite membrane with MCM-41 silica nanoparticles for brackish water purification. Membranes (Basel) https://doi.org/10.3390/membranes6040050 (2016).

Ali, M. E. A., Wang, L., Wang, X. & Feng, X. Thin film composite membranes embedded with graphene oxide for water desalination. Desalination 386, 67–76. https://doi.org/10.1016/j.desal.2016.02.034 (2016).

Inurria, A. et al. Polyamide thin-film nanocomposite membranes with graphene oxide nanosheets: Balancing membrane performance and fouling propensity. Desalination 451, 139–147. https://doi.org/10.1016/j.desal.2018.07.004 (2019).

Yin, J., Zhu, G. & Deng, B. Graphene oxide (GO) enhanced polyamide (PA) thin-film nanocomposite (TFN) membrane for water purification. Desalination 379, 93–101. https://doi.org/10.1016/j.desal.2015.11.001 (2016).

Choi, W., Choi, J., Bang, J. & Lee, J. H. Layer-by-layer assembly of graphene oxide nanosheets on polyamide membranes for durable reverse-osmosis applications. ACS Appl. Mater. Interfaces 5, 12510–12519. https://doi.org/10.1021/am403790s (2013).

Article  CAS  PubMed  Google Scholar 

Song, X., Zhou, Q., Zhang, T., Xu, H. & Wang, Z. Pressure-assisted preparation of graphene oxide quantum dot-incorporated reverse osmosis membranes: Antifouling and chlorine resistance potentials. J. Mater. Chem. A 4, 16896–16905. https://doi.org/10.1039/C6TA06636D (2016).

Al-Gamal, A. Q., Falath, W. S. & Saleh, T. A. Enhanced efficiency of polyamide membranes by incorporating TiO2-graphene oxide for water purification. J. Mol. Liq. 323, 114922. https://doi.org/10.1016/j.molliq.2020 (2021).

Fathizadeh, M. et al. Polyamide/nitrogen-doped graphene oxide quantum dots (N-GOQD) thin film nanocomposite reverse osmosis membranes for high flux desalination. Desalination 451, 125–132. https://doi.org/10.1016/j.desal.2017.07.014 (2019).

Rajakumaran, R., Kumar, M. & Chetty, R. Morphological effect of ZnO nostructures on desalination performance and antibacterial activity of thin-film nanocomposite (TFN)membrane. Desalination 495, 114673. https://doi.org/10.1016/j.desal.2020.114673 (2020).

Qian, X., Wang, X., Gao, X., Cao, W. & Gao, C. Effects of GO@CS core-shell nanomaterials loading positions on the properties of thin film nanocomposite membranes. J. Memb. Sci. https://doi.org/10.1016/j.memsci.2021.119102 (2020).

The authors would like to thank Dr. Abbas Mousavi and the membrane center of Sharif University of Technology for providing the PSf sublayer.

Polymer Reaction Engineering Department, Faculty of Chemical Engineering, Tarbiat Modares University, P.O. Box 14155-143, Tehran, Iran

Zahra Sanei, Taranom Ghanbari & Alireza Sharif

You can also search for this author in PubMed  Google Scholar

You can also search for this author in PubMed  Google Scholar

You can also search for this author in PubMed  Google Scholar

Data curation; Formal analysis; Investigation; Methodology; Roles/Writing—original draft (Z.S. and T.G.). Conceptualization; Data curation; Funding acquisition; Project administration; Supervision; Validation; Writing—review & editing (A.S.).

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Sanei, Z., Ghanbari, T. & Sharif, A. Polyethylene glycol-grafted graphene oxide nanosheets in tailoring the structure and reverse osmosis performance of thin film composite membrane. Sci Rep 13, 16940 (2023). https://doi.org/10.1038/s41598-023-44129-z

DOI: https://doi.org/10.1038/s41598-023-44129-z

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Scientific Reports (Sci Rep) ISSN 2045-2322 (online)

tartaric acid food grade Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.