Journal Search Engine
Search Advanced Search Adode Reader(link)
Download PDF Export Citaion korean bibliography PMC previewer
ISSN : 1226-0088(Print)
ISSN : 2288-7253(Online)
Membrane Journal Vol.30 No.2 pp.111-123
DOI : https://doi.org/10.14579/MEMBRANE_JOURNAL.2020.30.2.111

Two Dimensional (2D) Nanomaterials based Composite Membrane for Desalination

Yu Kyung Lee, Rajkumar Patel†
Energy and Environmental Science and Engineering (EESE), Integrated Science and Engineering Division (ISED), Underwood International College, Yonsei University, 85 Songdogwahak-ro, Yeonsu-gu, Incheon 03722, South Korea
Corresponding author(e-mail: rajkumar@yonsei.ac.kr)
April 13, 2020 ; April 25, 2020 ; April 26, 2020

Abstract


Growing industrialization and climate change lead to the huge demand for clean drinking water. Desalination of sea water by membrane separation process is one of the alternative and economically viable methods to fulfil the demand for water. In the membrane separation process, the presence of 2D materials enhances the performance of membrane by facilitating the water permeation, salt rejection, flux rate, and selectivity compared to the traditional reverse osmosis thin-film-composite membranes. In this review, composite membranes with different kinds of 2D materials are discussed on the basis of materials synthesis, characterization and desalination process.



2차원 나노재료 기반 복합막을 이용한 해수담수화

이 유 경, 라즈쿠마 파텔†
연세대학교 언더우드국제대학 융합과학공학부 에너지환경과학공학

초록


산업화와 기후 변화는 깨끗한 식수에 대한 수요 증가를 초래하였다. 막분리공정을 이용한 해수담수화는 물의 수요 에 대한 요구를 채울 수 있는 경제적으로 실현 가능한 대안 중 하나이다. 막분리공정에서 2차원 재료들은 기존 역삼투분리막 (reverse osmosis membrane) 기반의 폴리아마이드 박막복합막(TFC-PA)과 비교하였을 때 막의 강도를 높여주고 투수성을 용 이하게 하며 높은 염제거율 및 높은 선속률과 선택성을 보여준다. 이 리뷰 논문에서는 재료, 합성, 특징, 해수담수화 과정을 기반으로 다양한 2차원 재료로 구성된 복합막들을 소개하고 있다.



    1. Introduction

    Due to the worldwide water scarcity problem from climate change, removing salt from saline or brackish water to be used for fresh water, desalination in other words, has gained huge attention in recent years. Desalination is also an important process for wastewater treatment since toxicants in wastewater can easily leak into the environment and cause serious environmental damages that will eventually, negatively impact human beings[1-5]. To search for new desalination processes that are environmental-friendly, membrane filtration treatment has been proposed as the best option which include filtration methods such as nano filtration (NF) and reverse osmosis (RO). Among them, RO technology, specifically using 100 nm thick thin-film-composite polyamide (TFC-PA) membranes made up of polyester web, microporous interlayer, and thin surface barrier layer, has been widely-used in the past years. Adopting RO method has been useful so far as it has shown huge progress in membrane filtration technology in the removal of salt ions. However, current RO proc-esses using TFC-PA membranes have critical disadvantages in that they have insufficient water permeation rate and are prone to membrane fouling that lead to decrease in flux and salt rejection[6]. In order to resolve these drawbacks, research on two-dimensional (2D) nanomaterials has been ongoing which can be combined with RO membranes to generate nanocomposite membranes. Incorporation of 2D nanomaterials with RO membranes has been regarded as promising for the future since smooth surface, small pore sizes, conformal adhesion, and hydrophobicity can be expected from the grafting process which are some of the major requirements that need to be met for effective desalination process to occur[7]. Among various 2D nanomaterials such as graphene, graphene oxide, molybdenum disulfide (MoS2) are nanomaterials that are mainly presented in this review article.

    One of the notable 2D nanomaterials for the future is graphene since it has high chemical, mechanical, and thermal stability to withstand degradation in saline water. Also, graphene is flexible and has anti-fouling properties, allowing for fast water permeation and long lifetime of membranes[8]. Graphene can also be oxidized to form graphene oxide (GO) which is another 2D nanomaterial with extensive applications when making nanocomposite membranes. GO is composed of a single sheet of graphene rich with hydrophilic functional groups such as carboxyl, ether, and epoxy groups to be used as reaction sites to optimize transport properties[ 9]. Another major 2D nanomaterial is molybdenum disulfide (MoS2) monolayer which can be used for desalination due to the fact that vacancies can easily be introduced from chemical vapor deposition (CVD)[10]. Also, MoS2 is smaller than graphene and has intrinsically charged pore edges that are essential for high ion selectivity and water permeation[11]. Some other 2D nanomaterials include carbon nanotubes (CNTs), zeolite-derived materials such as silicoaluminophosphate (SAPO-34) and zeolitic imidazolate framworks (ZIFs-8), and covalent triazine organic frameworks (CTFs), all of which can expect membrane stability and high-water flux needed for selective desalination processes.

    In this review article, free-standing slit graphene membranes, ultrathin graphene nanofiltration membranes, and graphene/polypropylene (PP) or graphene/polyvinylidenedifluoride (PVDF) composite membranes are discussed. In case of membranes using graphene oxide, layer-by-layer (LBL) deposition and GO nanosheets were introduced for higher salt rejection and water flux. Graphene oxide laminates were also bonded with hydrophilic chitosan polymer for high sorption capacities and free standing ultrathin reduced GO membranes were utilized for desalination as well. For its unique properties, molybdenum disulfide (MoS2) based filters were integrated with RO semipermeable membranes.

    2. 2D based Composite Membrane

    2.1. Graphene

    Ang et al. reported 2D graphene-based membranes due to higher flux rate than traditionally used reverse osmosis (RO) membranes[12] [Fig. 1 (a)]. As for the improvement of salt rejection rate, researchers have used free-standing graphene slit membranes since they possess slits known to enhance salt ion filtration and freely deform under stress while their edges remain rooted to the microporous support, similar to other 2D materials. For the preparation steps, atoms were removed from a sheet of graphene in order to fabricate graphene membrane and the 2D membrane edges were fixed to the microporous support using current reported fabrication methods[13]. The effectiveness of the free-standing membrane was with frozen membrane in terms of three different circular pore sizes of A20 (20.33 Å2), A40 (39.87 Å2), and A64 (64.48 Å2). Through the analysis, it shows lower water permeability and higher salt rejection for free-standing membrane compared to frozen membrane, and it showed 100% salt rejection rate in case of A20 membrane. This phenomenon could possibly be explained by flexibility of the free-standing graphene which has been observed to take 18% more time for the salt ions to permeate through the membrane compared to the frozen membrane, contributing to 20% increase in salt rejection. Also, membranes with slits were compared to circular pore membranes to verify that slits work more effectively in salt ion rejection. Hence, critical slit length (Lc) of 2.28 Å was detected which had 100% salt rejection and was compared with A20 pore. Slit membrane has higher water permeability than A20 membrane. To facilitate effective water purification, ultrathin graphene nanofiltration membranes (uGNMs) of 22~53 nm thick were deposited onto the microporous substrate along with chemically converted graphene (CCG) with 2D nanochannels in between the graphene sheets[14]. For the preparation of uGNMs, dilute base-refluxing reduced graphene oxide (brGO) dispersion was filtered using mico-sized filtration membranes. In case of CCG, it was prepared by reducing GO chemically or thermally into brGO through Wilson’s method, resulting in the formation of holes which is efficient for high water flux[15]. XRD showed double peaks, indicating the presence of several functional groups still existing on brGO that allow for single-layer dispersing of brGO in water to form well-packed layer structures. water flux under different conditions were measured to determine uGNMS performances. Through the experiment, it could be known that water flux decreased as brGO loading increased under the low pressure of 1.0 bar. However, the constant pressure increment led to the increase in water permeation rate at constant brGO loading. uGNM has high flux rate due to hydrophobic carbon nanochannel that enable high water flux to take place. To further examine the performance of uGNMs, organic dye retention intensity was measured for MB and DI 81 solutions, resulting in the high retention intensity of 99.8 and 99.9% respectively due to physical sieving and electrostatic interaction between oppositely charged dyes. uGNMs is efficient at water purification, possessing both the advantages of graphene and polymer materials in terms of their high chemical resistance and mechanical properties as well as their high flexibility and thermal stability.

    Graphene monolayer was deposited on copper foil by chemical vapor deposition (CVD) and transferred onto polypropylene (PP) and polyvinylidenedifluoride (PVDF) commercial microfiltration membranes[16]. For the final process of attaching the graphene onto the target membrane, Cu was etched away using ammonium persulfate (APS) etchant. For the removal of defects in graphene layers formed during the transfer process, Nylon 6, 6, was used to fill in the defects using Franz cell for interfacial polymerization (IP). In order to determine the effectiveness of graphene transfer, PP and PVDF substrates were analyzed in terms of surface roughness, wettability, and porosity using SEM, AFM, and contact angle measurement. Using the root mean square (rms) values, surface roughness was estimated to be 42.4 and 23.8 nm for PP and PVDF membranes respectively, indicating the presence of a relatively smooth surface needed to contact with graphene. Also, average contact angle for PP and PVDF membranes showed the values of 115 and 90° respectively, supporting the fact that these two membranes are hydrophobic enough to avoid de-attachment graphene layer. Determining the ionic transport properties before interfacial polymerization, was analyzed for PP and PVDF membranes by evaluating the transport of 0.5 M KCl solution. The results showing the increase in the conductivity proportional to time implied 57 and 40% ion transport blockage for PP and PVDF substrate respectively, showing the existence of defects on the transferred graphene layer. Likewise, ion transport flux of PP and PVDF membranes revealed higher flux for PVDF membrane, revealing more suffering of defects for PVDF membranes than for PP membranes. In order to fill in the defects, Nylon 6, 6 was applied to the PP and PVDF substrates which was also investigated using transport measurements of KCl. Looking at the percentage of ion blockage, both graphene/PP and graphene/PVDF membranes showed an increment up to 67%. Using Solution B (HDMA), ion blockage could be enhanced up to 76% at 75 mM HDMA, indicating its effectiveness in defect sealing. Also, maximum ion blockage of 84% was seen when interfacial polymerization process reaction time reached 1 minute. In this regard, monolayer graphene transferred onto PP and PVDF substrates through CVD process, and then sealing the defects using Nylon 6, 6 by IP process proved to be effective in desalination process.

    2.2. Graphene oxide

    Choi et al. reported the synthesis of polyamide membranes coated with graphene oxide (GO) by layer- by-layer assembly for reverse osmosis application [17] (Fig. 2). Polyamide (PA) selective layer that is crosslinked and fully-aromatic was synthesized on polysulfone/ polyester (PSF/Polyester) non-woven support by interfacial polymerization in order for the preparation of polyamide-thin film composite (PA-TFC) membrane. Then, GO multi-layer prepared in the earlier stage was deposited on PA-TFC membrane via layer-by-layer (LbL) deposition method of oppositely-charged GO and aminated graphene oxide (AGO). Positively-charged AGO layer was first deposited on the surface which was favored by the negatively-charged carboxylic acid groups on the PA surface. Next, negatively-charged GO layer was depositied on top of the AGO layer by electrostatic interaction, forming one GO bilayer. The process was repeated until the desired number of bilayers were formed. Along with the electrostatic interaction, hydrogen bonding also enhanced the stability of GO multilayers. Surface morphology for pristine and GO10-coated PA was analyzed using SEM. Through SEM analysis, it was observed that pristine PA had rough surface, normally seen in full-aromatic Pas whereas GO10-coated PA showed the smooth membrane surface. Also, arithmetic average roughness (Ra) was estimated using AFM. In case of pristine PA, the roughness was ~46.5 nm while the roughness was greatly reduced in GO10-coated PA with the Ra value of ~21.5 nm. GO-coated membrane performance was evaluated by measuring the water flux and NaCl rejection as a function of the number of GO bilayers. The result showed that the performance of the membrane remained the same regardless of the number of GO bilayers. This can be explained by the unique property of GO nanosheets which enable ultrafast water transport due to high water permeability. Separation performance before and after chlorine exposure was measured in order to evaluate the chlorine stability of GO-coated membranes. This showed an increased water flux and reduced salt rejection after the exposure to chlorine for every tested membrane. However, the extent of the water flux increase was reduced with the increased number of GO bilayers. Also, as GO bilayers increased, the extent of NaCl rejection decrease progressively rather than drastically as was the case with pristine PA membrane. Stacked graphene oxide (GO) nanosheets were used to fabricate water separation membranes for desalination processes, achieved by controlling the sizes of nanopores and their functional groups[18] (Fig. 3). Stacked GO nanosheets are essential due to their high water flux and functional groups are important to create stable bondings between GO nanosheets to increase stability in water. To synthesize GO membrane, porous polysulfone support coated with polydopamine was soaked with GO solution and tri-mesoyl chloride (TMC) alternatively for a number of times to create stacked layers. Non-linear increase of water flux rate with increasing number of GO layers suggested that thickness of GO membranes do not have any impact on water resistance of GO coating. NaCl, Na2SO4, and organic dye rejection of GO membranes were also tested, and the result showed that NaCl rejection rate increased to 6~19% and Na2SO4 rejection rate increased to 26~46% as there were more number of GO layers. The similar case was observed for organic dyes (MB and R-WT) which showed the rejection rate of 46~66 and 93~95% respectively. The reason behind the increased rejection rate is predicted to be related to spaces in-between GO layers and GO nanosheets charges. This can be further supported by measuring the rejection rate of NaCl and Na2SO4 under different solution concentrations, which led to a conclusion that ionic strength increase results in rejection rate decrease. Through these observations, it can be concluded that stacked GO membranes are effective for water flux and salt rejections. Huang et al studied graphene oxide (GO) laminates coupled with hydrophilic bio-inspired membranes for water transport channel, allowing for high selective water permeation[19]. For the preparation of hydrophilic polymeric layer, chitosan (CS) was used as polymeric layer to coat pristine GO laminates for the fabrication of CS@GO membrane. AFM imaging was used to observe that the average roughness of CS@GO membranes decreased to 52.24 nm compared to pristine GO laminates with the height of 61.05 nm. Also, the shift in CS@GO membrane peaks for XPS analysis serves as a proof for the existence of interactions between CS and GO. For the investigation of separation performances of CS@GO membrane, separation factor for CS@GO membrane and pristine GO laminates were compared, showing 2,580 at 30°C for CS@GO membrane which was six times higher than pristine GO laminates. Meanwhile, water flux level was shown to maintain the same high level but increase in the operation temperature yielded increase in flux level and decrease in separation factor. Also, the permeate water concentrations of CS@GO membranes exceed 99% between the temperatures 30~ 70°C, making it feasible to apply bio-inspired layers. For the additional proof to support that CS@GO membranes exhibit outstanding performance, they were compared with other state-of-the-art membranes. Finally, performance comparisons between different phases of CS and GO membranes indicated that CS@GO membrane showed synergetic effect in permeate flux and separation factor. Liu et al. reported water purification using freestanding ultrathin reduced graphene oxide (rGO) that overcome the previous limitations detected in pristine GO and rGO membranes[20]. Filtration of diluted GO dispersion allowed the fabrication of a supported GO membranes on a hydrophilic mixed cellulose ester (MCE) filter, forming GO/MCE membrane. Then, GO/MCE membrane was submerged in hydroiodic acid (HI) solution and placed on a water surface, thereby forming rGO/MCE membrane. SEM was used to analyze the structures of GO and rGO membranes. The top-view SEM images showed firm attachments between GO and MCE filters compared to rGO and MCE filters which had the gap of 1~2 μm caused by the decrease in adhesion forces due to the loss of oxygen containing groups in GO during HI treatment. Furthermore, forward osmosis (FO) filtration was carried out to measure the water flux of rGO membrane and commercial cellulose triacetate (CTA) membrane. According to the result, water flux of rGO membrane exhibited linear proportionality to the salt concentration within 0.5~2.0 mol/L range, implying the near removal of inductively coupled plasma (ICP) in the freestanding ultrathin rGO membrane. In comparison, high salt concentrations had a self-limiting effect on CTA, leading to the restriction of the water flux. Lastly, rGO membranes were compared with GO and CTA membranes in terms of salt ion and water permeation rate. rGO membranes have higher salt rejection rate and water permeation rate compared to the other two membranes. In this sense, freestanding ultrathin rGO membranes is effective for water purification due to their high salt rejection rate and water permeation rate as indicated in the above experiments. In order to enhance the desalination ability of polyamide (PA) thin layer, thin-film nanocomposite (TFN) containing graphene oxide (GO) was prepared using in-situ interfacial polymerization (IP) process[21]. By submerging PSU support layer into 2.0% managed pressure drilling (MPD) - water solution and then soaking the layer into trimesoyl chloride (TMC) - hexane solution, PA-TFN was fabricated which allowed the incorporation of GO for the final product of GO-TFN membrane. First, surface roughness was analyzed by AFM through observing root mean square (RMS) value that shows the height of GO coated on PA-TFN. The analysis showed decrease in roughness, implying the interruption of the growth of GO nanosheets on PA-TFN’s leaf-like structure during IP process. However, formation of GO clusters caused surface roughness to increase proportionally to amount of loaded GO. Water permeability and salt rejection values were checked in order to consider the effectiveness of the TFN membrane performance upon incorporating GO. The experiment showed an increase in permeate water flux from 39.0 ± 1.6 L/m2 h to 59.4 ± 0.4 L/m2 h as GO concentration increased whereas NaCl and Na2SO4 rejection showed a slight decrease as comparable to other studies. Therefore, this proved the good stability of GO in terms of salt rejection and water permeability under 300 psi. From XRD it is observed that the presence of interlayer spaces around 0.83 nm between GO nanosheets along with TFN membrane which has internal mesoporous silica provide the structure for ultrafast water transport. Also, incorporation of GO onto the membrane surface increases surface hydrophilicity, thereby improving membrane water permeability. Presence of interlayer space in between GO nanosheets allow GO-TFN to be used effectively as water channels by increasing the rate of water permeability.

    Hegab reported the improvement in the desalination of brackish water (BW) using functionalized graphene oxide (GO)-based membranes that can give hydrophilicity and smoothness to the membrane surface needed for antifouling properties[22]. GO nanosheets were functionalized with chitosan (GO/f-CS) that allowed carboxylic groups in GO and amino groups in chitosan to form amide bonds for the membrane hydrophilicity. Also, GO/f-CS were coated onto the BW thin film composite (TFC) membrane surface for negative polyamide (PA) layer, positive CS, and negative GO nanosheets to form stronger electrostatic interactions needed for antifouling and better membrane performances. GO/f-CS solution was deposited onto the PA substrate. For membrane comparisons, low GO content (GO1) and high GO content (GO2) membrane samples were also prepared. In order to compare GO and GO/f-Cs solutions, UV-visible spectrometer was used for the samples GO1, GO1/f-CS, GO2, and GO2/f-CS. By observing the graph at zero-time, one can interpret that there is a strong interaction between water, GO nanosheets, and protonated chitosan known as chitosanium. The dispersion stayed the same after 24 hours, indicating that there was no precipitation. Also, the large difference between the concentrations of GO2 and GO1 indicate that GO2 has higher absorbance than GO1 and the different features for GO/f-CS imply the presence of newly formed amide bonds between GO and CS. AFM analysis was performed, conforming that modified membranes of GO/f-CS had much smoother surface morphology. Looking at the water flux rate, GO/f-CS membranes first experience a decline because of the thickness of the surface but showed a net increase of water permeation by 9.8% for the higher GO content membranes as the surface became more hydrophilic to allow higher water permeation. Also, modified membranes had higher salt rejection rate than unmodified membranes, reaching up to 95.56% for the membranes with higher GO content. Finally, membrane fouling performance was tested using BSA as an alternative for the organic foulant. Unmodified membranes experienced steeper decline in the flux as it rapidly adsorbs onto the surface of the unmodified membranes unlike GO/f-CS. GO/f-CS membranes can work effectively for the desalination of BW and enhancement of membrane fouling performances. Perreault et al. reported inactivation of bacteria cells by using graphene oxide (GO) nanosheets covalently bounded to thin film composite (TFC) polyamide (PA) active layer through surface functionalization[23] (Fig. 4). For the preparation of TFC membranes, interfacial polymerization of PA onto the polysulfone support layer was performed. GO was functionalized onto the active layer of TFC membrane using 1-ethyl-3-[3-(dimethylamino) propyl] carbodiimide hydrochloride (EDC) and N-hydroxysuccinimide (NHS), forming amide bonds in the process. SEM showed the smooth surface of the GO-TFC membrane, indicating the successful GO functionalization onto the TFC membrane. Membrane properties were also evaluated through the transport properties measurement Furthermore, salt rejection values for Ctrl-TFC and GO-TFC were 98.2 ± 1.2 and 97.8 ± 0.8% respectively, indicating the maintenance of membrane performances even after surface modification took place. Finally, comparisons of antimicrobial activities were analyzed between Ctrl-TFC membrane and GO-TFC membrane. The result showed 64.5% reduction of the number of active E. coli in GO-TFC, serving as a proof for the antimicrobial function of GO which is known to damage cell membranes of bacteria. It is seen in SEM that E. coli attached to the surface of Ctrl-TFC whereas it is shrunken due to the presence of GO which is present on the GO-TFC membrane. GO is effective at reducing microbial activities when functionalized onto the surface of PA active layer.

    2.3. Molybdenum disulfide

    Nanoporous molybdenum disulfide (MoS2) filters have been utilized for water desalination through the application of tensile strain to open and close the filters [24]. Water flux was measured under different strain conditions. Specifically, there was no water passage via MoS2 when 3% strain was applied while water flux increased directly proportionally to the applied pressure under 6, 9, and 12% strain. Through this, it could be known that MoS2 filter is closed at 3%, but changes into an open state when the strain reaches 6%. Water permeability of MoS2 filter is about 100 g/m2⋅s⋅atm compared to the commercial reverse osmosis (RO) membrane filters with water permeability of 1.18 g/m2 ⋅s⋅atm, implying that MoS2 filters are more efficient than commercial RO membranes and are capable of tuning when mechanical stretching is applied. Nanopores get larger as tensile strain increases which facilitates the water flux to increase. Finally, nonequilibrium potentials of mean force (NPMFs) for water molecules were analyzed in order to find the physical force of MoS2 filter. In case of water NPMFs, the graph was evenly spread out, indicating that water pass through the filters easily as there are no energy barriers. In comparison, sodium ions showed high energetic barriers of ~20 kJ that meant that sodium ions could not pass through the nanopores, which could be effectively explained by the Coulombic barrier existing between the positively charged molybdenum atoms and sodium cations. For chloride ions, two energy wells that are distributed symmetrically are observed through the NPMF, meaning that the accumulation of chloride ions will occur around the nanopore entrance, forming plateau. Therefore, opening and closing of MoS2 nanpores through the application of strain allow the efficient selective filtration to take place.

    In order to enhance the performance of the membrane Li et al. introduced reduced nanoporous graphene oxide into the membrane[25] (Figs. 5, 6). The permeability of the membrane enhanced by 26 times due to the presence of ~3 nm pore without costing salt rejection. Porous structure reduce the path length of water permeation. At the same time rejection of the salt is dictated by exclusion, based on interspace size and Donnan effect. Yang et al. recently review beautifully the mechanism of transport in graphene oxide membrane[ 26]. Structural, mechanical and filtration ability of GO membrane is hugely affected in aqueous solution. Ye et al. modified GO by in situ polymerization of ionic liquid which enhanced mechanical strength drastically along with other properties[27] (Figs. 7, 8).

    2.4. Others

    In recent study, carbon nanotube (CNT) has gained attention in desalination processes due to its ability of being able to increase the water flux without changing selectivity and fouling resistance of the membrane[28] [Fig. 1 (b)]. Functionalized carbon nanotube (fCNT) blended support layer showed higher water flux and fouling resistance. For the functionalization of CNT, chemical oxidation method was used. Polyethersulfone (PES) and fCNT blended PES (fPES) were prepared by phase inversion method. fTFC membrane were prepared by interfacial polymerization. For the determination of the hydrophilicity of the fTFC support layer, contact angles between fTFC membrane and TFC membrane support was compared. The result showed the decrease in contact angle for fTFC membrane support from 54.76 to 51.03°, implying that fTFC membrane support layers have more hydrophilic functional groups attached through the coating of fCNT layer. Also, increase of the average porosity of fPES com-pared to PES membranes from 79.69 to 82.05% along with the increase of average pore width and total pore area indicated that the increased pore size as well as ultrafast water channel of CNT enhance membrane permeability after being incorporated as fCNT support layer. fTCF membrane showed higher water flux than TFC membrane at the beginning due to its porous structures. Also, it took only 50 minutes for TFC membrane to reach the completion of the organic fouling while it took 150 minutes for fTFC, proving that CNT incorporation reduced the fouling. CNT functionalized onto TFC support layer membrane enhanced water flux and salt rejection, allowing CNT to be an appropriate material for effective desalination process. Duke et al. reported desalination process using zeolite-derived materials, zeolite silicoaluminophosphate (SAPO-34) and zeolitic imidazolate frameworks (ZIFs-8). SAPO-34 contains micropores of 0.38 nm[29]. Both materials are known to permeate water while blocking out other ions such sodium and chloride ions from permeation. For the preparation of SAPO-34, aluminum isopropoxide, phosphoric acid was mixed in deionized water, followed by the addition of Ludox AS-40 colloidal silica, tetraethylammonium hydroxide, and dipropylamine. The solution was treated at 220°C for 24 hydrothermally to prepare SAPO-34 crystals. In case of ZIFs-8 crystals, zinc nitrate hyxahydrate, 2-methylimidazole, and methanol were mixed together and stirred at room to temperature. Membrane of these two materials are prepared using porous support. Through the experiment, it could be known that SAPO-34 and ZIF-8 membranes absorb monovalent ions into their structure and release Al ions and Zn ions respectively when treated with sea water. Performance of the membrane was tested by desalination test. It showed that both membranes were unsuitable for reverse osmosis (RO) but were suitable for sorption processes as they were able to absorb ions and maintained stability in distilled and sea water.

    In another report, covalent organic frameworks (COFs) which has ordered crystalline structures and high pore density necessary for the increase in salt rejection and water permeability was studied[30]. Specifically, chemically stable 2D covalent triazine organic frameworks (CTFs) for desalination processes have gained attention since CTFs can overcome challenges that COFs face regarding its unstability in water. For the preparation steps, dynamic trimerization reaction was performed to synthesize and measure 2D CTFs. In order to determine the effectiveness of desalination, salt rejection percentage was analyzed for CTFs of different pore sizes. For CTF-0, salt rejection was 100% but the pore size was substantially small to enable water molecules to pass through the membrane. In comparison, CTF-1 had approximately 91% salt rejection rate that would allow desalination of brackish water. CTF-1 has high pore density, allowing water permeation. Moreover, due to the advantage that CTFs possess regarding structural tenability, results show an increase in salt rejection rate for the functionalized CTF1 frameworks although decrease in water permeability is also observed compared to the CTF1 membranes.

    3. Conclusions

    There are several types of 2D nanomaterials such as graphene, graphene oxide, MoS2, and other promising nanomaterial frameworks are incorporated in RO membranes for effective desalination of sea water or brackish water. These novel materials are single-atom thickness for ultrafast water transport, abundant functional groups for property modifications, small pores to ensure large surface areas and selective permeability, and conformal adhesion needed for their attachment onto membranes. Although further research still need to be conducted in order for these 2D nanomaterials to be commercialized, there is no doubt that these materials have provided a steppingstone for desalination technology that can ultimately resolve water scarcity problems. In this review different type of composite membrane consisting of 2D materials are discussed in detail.

    Figures

    MEMBRANE_JOURNAL-30-2-111_F1.gif

    Schematic diagram of (a) 2D materials based composite membrane and (b) carbon nanotube based composite membrane.

    MEMBRANE_JOURNAL-30-2-111_F2.gif

    Top-down SEM images of (a) the pristine, uncoated polyamide (PA) and (b) the GO10-coated PA membranes (scale bar = 1 μm). AFM height images of (c) the pristine PA and (d) the GO10-coated PA membranes (Reproduced with permission from Choi et al., 17, Copyright 2013, American Chemical Society).

    MEMBRANE_JOURNAL-30-2-111_F3.gif

    Schematic illustration of (a) a step-by-step procedure to synthesize the GO membrane, (b) the mechanism of reactions between polydopamine and TMC, and (c) the mechanism of reactions between GO and TMC (Reproduced with permission from Hu et al., 18, Copyright 2013, American Chemical Society).

    MEMBRANE_JOURNAL-30-2-111_F4.gif

    (A) Colony-forming units after E. coli cells had been in contact with the membranes for 1 h at room temperature. (B) SEM micrograph showing normal E. coli cells at the surface of the Ctrl-TFC membrane. (C and D) SEM micrographs showing compromised E. coli cells at the surface of the GO-TFC membrane active layer. Compromised cells are denoted with an arrow in panel C. (Reproduced with permission from Perreault et al., 23, Copyright 2013, American Chemical Society).

    MEMBRANE_JOURNAL-30-2-111_F5.gif

    Schematic illustration of the fabrication process of reduced nanoporous graphene oxide (rNPGO) membrane (Reproduced with permission from Li et al., 23 Copyright 2019, American Chemical Society).

    MEMBRANE_JOURNAL-30-2-111_F6.gif

    Water transport and salt rejection mechanism in rGO and rNPGO membranes (Reproduced with permission from Li et al., 23 Copyright 2019, American Chemical Society).

    MEMBRANE_JOURNAL-30-2-111_F7.gif

    GO membrane and GO hybrid membrane in aqueous solution (Reproduced with permission from Ye et al., 27, Copyright 2019, American Chemical Society).

    MEMBRANE_JOURNAL-30-2-111_F8.gif

    (a) Digital pictures of the i-AVG membrane before and after peeling off from the CA membrane. (b) SEM image of the cross-section of the i-AVG membrane, and (c) the FT-IR spectra of the GO membrane, [AVIM]Cl, and i-AVG membrane. The preparation conditions of the i-AVG membrane here: mass ratio of [AVIM]Cl/GO = 2, absorbed dose 100 kGy, and dose rate 0.625 kGy⋅s-1 (Reproduced with permission from Ye et al., 27, Copyright 2019, American Chemical Society).

    Tables

    References

    1. N. Misdan, W. J. Lau, and A. F. Ismail, “Seawater reverse osmosis (SWRO) desalination by thin-film composite membrane - current development, challenges and future prospects”, Desalination, 287, 228 (2012).
    2. A. P. Rao, N. V. Desai, and R. Rangarajan, “Interfacially synthesized thin film composite RO membranes for seawater desalination”, J. Membr. Sci., 124, 263 (1997).
    3. D. W. Kim, “Review on graphene oxide-based nanofiltration membrane”, Membr. J., 29, 130 (2019).
    4. H. Oh, J. H. Lee, and R. Patel, “Removal of heavy metal ions from wastewater by polyacrylonitrile based fibers: A review”, Membr. J., 29, 123 (2019).
    5. B. W. Lee, S. Lee, and R. Patel, “Effect of antifouling composite membrane on membrane bioreactor: A review”, Membr. J., 30, 1 (2020).
    6. M. A. Shannon, P. W. Bohn, M. Elimelech, J. G. Georgiadis, B. J. Mariñas, and A. M. Mayes,“Science and technology for water purification in the coming decades”, Nature, 452, 301 (2008).
    7. S. C. O’Hern, C. A. Stewart, M. S. H. Boutilier, J.-C. Idrobo, S. Bhaviripudi, S. K. Das, J. Kong, T. Laoui, M. Atieh, and R. Karnik, “Selective molecular transport through intrinsic defects in a sin gle layer of CVD graphene”, ACS Nano, 6, 10130 (2012).
    8. E. N. Wang and R. Karnik, “Graphene cleans up water”, Nat. Nanotechnol., 7, 552 (2012).
    9. D. R. Dreyer, S. Park, C. W. Bielawski, and R. S. Ruoff, “The chemistry of graphene oxide”, Chem. Soc. Rev., 39, 228 (2010).
    10. Z. Yu, Y. Pan, Y. Shen, Z. Wang, Z.-Y. Ong, T. Xu, R. Xin, L. Pan, B. Wang, L. Sun, J. Wang, G. Zhang, Y. W. Zhang, Y. Shi, and X. Wang, “Towards intrinsic charge transport in monolayer molybdenum disulfide by defect and interface engineering”, Nat. Commun., 5, 5290 (2014).
    11. D. Cohen-Tanugi and J. C. Grossman, “Water desalination across nanoporous graphene”, Nano Letters, 12, 3602 (2012).
    12. E. Y. M. Ang, T. Y. Ng, J. Yeo, Z. Liu, and K. R. Geethalakshmi, “Free-standing graphene slit membrane for enhanced desalination”, Carbon, 110, 350 (2016).
    13. S. C. O’Hern, D. Jang, S. Bose, J.-C. Idrobo, Y. Song, T. Laoui, J. Kong, and R. Karnik, “Nanofiltration across defect-sealed nanoporous monolayer graphene”, Nano Letters, 15, 3254 (2015).
    14. Y. Han, Z. Xu, and C. Gao, “Ultrathin graphene nanofiltration membrane for water purification”, Adv. Funct. Mater., 23, 3693 (2013).
    15. J. P. Rourke, P. A. Pandey, J. J. Moore, M. Bates, I. A. Kinloch, R. J. Young, and N. R. Wilson, “The real graphene oxide revealed: Stripping the oxidative debris from the graphene-like sheets”, Angew. Chem. Int. Ed., 50, 3173 (2011).
    16. F. M. Kafiah, Z. Khan, A. Ibrahim, R. Karnik, M. Atieh, and T. Laoui, “Monolayer graphene transfer onto polypropylene and polyvinylidenedifluoride microfiltration membranes for water desalination”, Desalination, 388, 29 (2016).
    17. W. Choi, J. Choi, J. Bang, and J. H. Lee, “Layer-by-layer assembly of graphene oxide nanosheets on polyamide membranes for durable reverse- osmosis applications”, ACS Appl. Mater. Interfaces, 5, 12510 (2013).
    18. M. Hu and B. Mi, “Enabling graphene oxide nanosheets as water separation membranes”, Environ. Sci. Technol., 47, 3715 (2013).
    19. K. Huang, G. Liu, J. Shen, Z. Chu, H. Zhou, X. Gu, W. Jin, and N. Xu, “High-efficiency water- transport channels using the synergistic effect of a hydrophilic polymer and graphene oxide laminates”, Adv. Funct. Mater., 25, 5809 (2015).
    20. H. Liu, H. Wang, and X. Zhang, “Facile fabrication of freestanding ultrathin reduced graphene oxide membranes for water purification”, Adv. Mater., 27, 249 (2015).
    21. J. Yin, G. Zhu, and B. Deng, “Graphene oxide (GO) enhanced polyamide (PA) thin-film nanocomposite (TFN) membrane for water purification”, Desalination, 379, 93 (2016).
    22. H. M. Hegab, Y. Wimalasiri, M. Ginic-Markovic, and L. Zou, “Improving the fouling resistance of brackish water membranes via surface modification with graphene oxide functionalized chitosan”, Desalination, 365, 99 (2015).
    23. F. Perreault, M. E. Tousley, and M. Elimelech, “Thin-film composite polyamide membranes functionalized with biocidal graphene oxide nanosheets”, Environ. Sci. Techno. Lett., 1, 71 (2013).
    24. W. Li, Y. Yang, J. K. Weber, G. Zhang, and R. Zhou, “Tunable, strain-controlled nanoporous MoS2 filter for water desalination”, ACS Nano, 10, 1829 (2016).
    25. Y. Li, W. Zhao, M. Weyland, S. Yuan, Y. Xia, H. Y. Liu, M. P. Jian, J. D. Yang, C. D. Easton, C. Selomulya, and X. W. Zhang, “Thermally reduced nanoporous graphene oxide membrane for desalination”, Environ. Sci. Technol., 53, 8314 (2019).
    26. T. S. Yang, H. Lin, K. P. Loh, and B. H. Jia, “Fundamental transport mechanisms and advancements of graphene oxide membranes for molecular separation”, Chem. Mat., 31, 1829 (2019).
    27. J. Z. Ye, B. W. Zhang, Y. Gu, M. Yu, D. W. Wang, J. Y. Wu, and J. Y. Li, “Tailored graphene oxide membranes for the separation of ions and molecules”, Appl. Nano Mater., 2, 6611 (2019).
    28. M. Son, H. G. Choi, L. Liu, E. Celik, H. Park, and H. Choi, “Efficacy of carbon nanotube positioning in the polyethersulfone support layer on the performance of thin-film composite membrane for desalination”, Chem. Eng. J., 266, 376 (2015).
    29. M. C. Duke, B. Zhu, C. M. Doherty, M. R. Hill, A. J. Hill, and M. A. Carreon, “Structural effects on SAPO-34 and ZIF-8 materials exposed to seawater solutions, and their potential as desalination membranes”, Desalination, 377, 128 (2016).
    30. L. C. Lin, J. Choi, and J. C. Grossman, “Two-dimensional covalent triazine framework as an ultrathin- film nanoporous membrane for desalination”, Chem. Commun., 51, 14921 (2015).