3.1. Characterization of gCN nanosheets
To corroborate the formation of 2D gCN nanosheets, the functional groups, crystalline structure, and morphology of the synthesized nanosheets were conducted using FTIR, XRD, SEM, and TEM analysis, respectively.
The chemical nature of different functional groups present on the surface of the gCN sample was explored through FTIR analysis (Fig. 2a). The spectrum of the as-synthesized gCN illustrates several characteristic infrared absorption bands that are in coordination with those of the reported spectra by many other publications [56, 57, 59–61, 67, 70–78]. The sharp absorption peak, known as a specific characteristic band of gCN, was recorded at 812 cm− 1, which represents the ring-sextant out-of-plane bending vibration of heptazine rings. In addition, the absorption band at 890 cm− 1 is related to the deformation mode of N\(-\)H in amino groups [56, 57, 59–61, 72, 73, 76]. The absorption bands in the 1200–1700 cm− 1 region correspond to the typical stretching patterns of aromatic heterocyclic C\(-\)N and C\(=\)N in gCN structure [56, 57, 59, 67, 70–72, 76–78]. With a more detailed review on the peaks of the mentioned region, it can be found that the peaks at 1237 and 1316 cm−1 are allocated to the stretching vibration of interconnected trigonal units of C\(-\)N(\(-\)C)\(-\)C (full condensation) or bridging C\(-\)NH\(-\)C (partial condensation) [60, 70, 75]. Moreover, the peaks at 1407, 1462, 1575, and 1641 cm− 1 are ascribed to the typical stretching vibration modes of heptazine-derived repeating units [70, 76]. On the other hand, as narrated in some references [57, 74], the aromatic C\(-\)N stretching bond of tri-s-triazine is identified by the peaks at the wavenumbers of 1237, 1316, 1407, and 1462 cm− 1. The existence of the C\(=\)N heterocycle stretching vibration bond is also proved by the peaks appearing in the vicinity of 1540, 1575, and 1641 cm− 1 wavenumbers. Finally, the broad absorption bands in the range of 3000–3350 cm−1 are derived from the stretching vibration modes of the terminal N\(-\)H band, which is at the deficiency sites of the aromatic ring encompassing the residual \(-\)NH or \(-\)NH2 groups [70, 73, 74, 76, 78]. It is worth mentioning that such primary and secondary amines arise from the incomplete condensation of melamine precursor [52, 56, 72]. Furthermore, a group of researchers [57, 60, 67, 72] has pointed out the presence of the O\(-\)H stretching vibration modes in this region due to the physically adsorbed water molecules.
Depending on the synthesis method and the precursor type, diverse allotropes of carbon nitride with different properties can be obtained. Due to its unique physicochemical properties, gCN is known as the most energetically stable allotrope in ambient conditions [38, 48–50]. Generally, the configuration of gCN has been constituted of carbon and nitrogen atoms, which are covalently linked by sp2 hybrid orbitals [48, 50]. The periodically connected tri-s-triazine units have also formed the whole structural framework of gCN. Furthermore, the layered material has been contained of a lattice structure analogous to graphite, which are coupled with weak van der Waals interactions. Within the layers, nitrogen atoms alternate instead of carbon atoms in a hexagonal arrangement [39, 49, 50]. The structural formula of gCN displays many intrinsic triangular nanopores and defective pore structures (see the inset in Fig. 2a).
In order to inquire about the crystalline structure of the as-synthesized gCN nanosheets, XRD analysis was performed and the resultant pattern is represented in Fig. 2b. The diffractogram displays two typical diffraction peaks at around 13.08º and 27.68º of 2θ, indicating the successful synthesis of the gCN by thermal polymerization from melamine precursor. In detail, the weak reflection peak located at 2θ = 13.08º is attributed to the (100) in-planar consecutive motif of tri-s-triazine polymeric building units. Meanwhile, the strong peak with high intensity at 2θ = 27.68º corresponding to the (002) crystallographic planes of graphitic materials, which originates from the presence of the interlayer periodic stacking of the conjugated aromatic graphite-like framework in gCN nanosheets. These values are in accordance with JCPDS No. 87-1526 [56, 57, 72–74, 77, 79, 80].
The morphological properties of the synthesized nanosheets together with EDS analysis are presented in Fig. 2. The images of SEM and TEM of gCN are demonstrated in Fig. 2c and d, respectively. As can be observed from the SEM images, a sheet-like lamellar structure with irregular wrinkles and ripples has been the result of the nanomaterial synthesis. The TEM images, however, can provide detailed descriptions concerning the morphology of the nanosheets. They exhibit a transparent graphene-like 2D morphology composed of poly-crystallized texture with a size of about several hundred nanometers. Figure 2e shows the EDS spectrum of the gCN nanosheets and confirms the presence of carbon and nitrogen elements without any impurities.
"The location of Fig. 2"
3.2. Membrane characterization
Figure 3a shows the digital photo of different dope solutions prepared in this study for membrane fabrication after standing for 24 h at 30 ºC. Compared with the PVDF clear solution (vial number 1), adding gCN nanofillers caused pale-yellow coloring opaque solutions (vials number 2–5), indicating that an even mixing as well as sustainable solutions were being achieved. Moreover, from Fig. 3a, it can be evidently observed that the color of the colloidal solutions is gradually getting darker as the amount of gCN is increased.
ATR-FTIR and XRD were used to characterize the chemical composition and crystalline phase structure of the as-prepared membranes, respectively. The typical spectra of the pristine PVDF membrane and membranes with various contents of gCN nanosheets are depicted in Fig. 3(b and c).
The FTIR spectrum of PVDF makes it possible to discern between the different crystalline forms. Besides, the same film of PVDF can contain one or more crystalline phase structures depending on the preparation conditions. In the FTIR spectrum of pure PVDF (Fig. 3b), the peak at 878 cm− 1 is assigned to the amorphous phase. From literatures on the spectral features of the crystalline phases in PVDF, the peaks at 762, 795, and 976 cm− 1 represent typical bands of α phase, while the absorption bands at 840 and 1277 cm− 1 are imputed to the all-trans conformations of β phase vibration [23, 70, 76, 81]. Likewise, the characteristic peaks at 1403, (1072 and 1179), and 878 cm− 1 are related to C\(-\)H stretching vibrations of \(-\)CH2 groups, C\(-\)F stretching vibrations of \(-\)CF2 groups, and C\(-\)C stretching vibration bond, respectively [23, 69]. Additionally, two vibrational bands and reflection peaks at the wavenumbers of 1235 and 1277 cm− 1 can be ascribed to the stretching frequencies of C\(-\)F and C\(-\)F2 [18]. The weak absorption bonds at 2981 and 3019 cm− 1 could also correspond to the sp3 C\(-\)H symmetric and asymmetric stretch of methylene groups of PVDF [69, 70].
As can be seen from the curves, all the gCN incorporated membranes exhibit similar IR spectra to the pure PVDF membrane without any new peaks appearing related to gCN. There could be several reasons for this issue, including: (1) The amount of gCN powder embedded in the polymer matrix is not enough to be detected. In other words, adding a trace amount of nanosheets does not change the surface chemical structure of the PVDF membrane [18, 71]; (2) The sheets are covered by or buried within the PVDF polymer matrix; and (3) The nanosheets are mixed with PVDF physically or by weak chemical bonding [32, 82].
It is worth noting that the interference and interfacial interaction between polymer and nanoparticles results in a slight diminution of bond intensity in IR spectra for fabricated mixed matrix membranes. This decline of transmittance intensity becomes bolder with the enhancement of nanofiller content as a consequence of the increased interfacial influence [82]. These results indicate a good compatibility between gCN nanosheets and PVDF molecules [18].
Figure 3c illustrates XRD patterns of the membranes embedded with 0.03 and 0.06 wt.% of gCN nanosheets compared to the pristine PVDF membrane. As is well known, PVDF is a semi-crystalline polymer, which has two main crystalline phases, namely nonpolar α and polar β forms [70, 83]. For pristine PVDF membrane, two characteristic diffraction peaks at angles (2θ) of 18.04º and 26.28º represent the crystal plane reflections of (020) and (021), respectively, which belong to the characteristic α-phase of PVDF. Moreover, there is a relatively small peak at 2θ = 20.4º which indicates the characteristic β-phase of PVDF [70, 76, 81]. After the addition of gCN nanosheets, the former packing of the backbone is maintained. In other words, it can be seen that the XRD patterns of the mixed matrix membranes have not changed significantly when compared to that of the pristine membrane. It should also be acknowledged that the peaks imputed to the reflections of α and β phases of the PVDF still appeared in nearly the same situations, but their intensities are lower when compared to those of the pristine PVDF membrane and decrease with increasing the concentration of nanofillers as represented in Fig. 3c [70]. In addition, the characteristic peaks of gCN are not visible in the resultant mixed matrix membranes, which is attributed to the meager amount of 2D nanofillers [67]. These observations were further confirmed by the IR spectra.
"The location of Fig. 3"
To investigate the impact of gCN on the structure of the resultant mixed matrix membranes, visual characterization under the SEM analysis was performed (Figs. 4 and 5). The surface morphology of the unmodified and the gCN modified PVDF membranes at two different magnifications are depicted by top-view SEM micrographs in Fig. 4(a-e1). All membranes exhibit a uniform and defect-free surface. The gCN nanosheets are not visible on the surface of the mixed matrix membranes owing to the low concentration as well as good dispersion of the nanosheets. Comparing the surface SEM images of the membranes indicates that adding gCN into the polymeric solution up to 0.03 wt.% leads to an increment in the number and the size of surface pores. However, it is worth noting that no tangible change was observed in the number of surface holes following the addition of 0.03 and 0.045 wt.% of gCN in the PVDF matrix. This phenomenon can be clarified by the fact as follows:
In general, thermodynamics and kinetics are two determining factors, which control the format of membranes prepared with phase inversion. The presence of polymeric nanosheets into the dope solution enhances its thermodynamic instability, which favor the progress of phase separation and increasing the rate of demixing [84].
Moreover, the number of surface pores diminishes obviously since the percentage of nanofillers raises to 0.06 wt.%. This is maybe because the relatively higher viscosity of the dope solution comprising 0.06 wt.% of gCN, which hinders the exchange rate of the solvent and non-solvent during the phase inversion process, resulting in lower membrane pore size and porosity. Generally, the addition of inorganic or carbon-based nanomaterials will increase the solution viscosity and thus hinder the solution demixing process [85].
"The location of Fig. 4"
As previously mentioned, the membranes in this study were formed via the phase inversion of the PVDF induced by the immersion precipitation (IP) method in the presence of water as the non-solvent. Phase inversion induced by IP can be considered as the key technique in the production of industrial asymmetric membranes owing to the simplicity of the process. Most of the available PVDF membranes are also fabricated via the same method due to their easy dissolution in prevalent organic solvents. Since PVDF is well-known as a semi-crystalline polymer, its behavior during the phase separation process is more complicated than amorphous polymers and is governed by two mechanisms (i.e., liquid–liquid demixing and crystallization). The coagulation bath is one of the main factors in determining the sequence of the mechanisms as mentioned above, which dictates the ultimate morphology of the PVDF membrane formed through the IP method. The presence of water (acts as the harsh non-solvent) as the coagulation medium during IP often accelerates the liquid–liquid demixing process and consequently, an asymmetric structure like ordinary amorphous membranes consisting of dense skin layer along with finger-like voids could be created [86].
Figure 5(a-e1) illustrates the cross-sectional morphology of the pristine and the gCN embedded membranes. It can easily be recognized by these SEM images that all membranes have an asymmetric structure composed of a thin compact top layer and a porous sublayer divided into two distinct regions (i.e., finger-like and sponge-like). These observed morphologies are consistent with the contents of water immersion induced structures mentioned above. As shown in Fig. 5, the incorporating of the gCN nanosheets changes the cross-section microstructure of the prepared PVDF/gCN membranes. Beneath the skin is a region composed of parallel columnar macro voids, which expand to the middle part of the membrane. Adding the relatively hydrophobic gCN nanosheets in the dope solution, turns the texture of this region into spherical-like nodules connected to each other with pores that are resided between these adjacent nodule aggregates. Moreover, the size of these spherical nodules appears to be growing over the entire membrane cross-section by increasing the content of the polymeric nanosheets. The lower half of the gCN embedded membranes cross-section also demonstrates a cellular morphology that became more porous and mixed with spherical nodules compared with pristine membrane. However, for the pure PVDF membrane, the sponge-like morphology instead of the cellular one was observed both at the bottom and among the finger-like macropores. As can be apparently seen, interconnected inner pores of the pure PVDF membrane joined to form larger voids in the 0.015 wt. % gCN embedded membrane. Moreover, by adding 0.015 wt.% of gCN nanosheets, the finger-like voids changed to pear-shaped macro voids with larger sizes and looser than those of the pure membrane. In addition, by incorporating the nanofillers, a slight decline in the skin layer thickness was detected. The finger-like structures seem to become slightly narrow, more regular, and with a rather uniform size after adding 0.03 wt.% gCN into the PVDF membranes. In the case of the 0.045 wt.% gCN embedded membrane, the finger-like structure is similar to the 0.03 wt.% counterpart but, to some extent, narrower. However, a further increase in the loading of gCN to 0.06 wt.%, increased the viscosity of the solution and, as already explained, resulted in a lag in the phase inversion kinetics (i.e., the rate of solvent-outflow and nonsolvent-inflow) during the NIPS process and the cellular structure was more pronounced than the finger-like one.
Besides, the upper layer of the finger-like zone in the resultant gCN embedded membranes was composed of small cavities. Also, the number of these cavities in the membrane incorporating 0.03 wt.% of gCN seems to be more than the other three mixed matrix membranes. It is expected that such regular and uniform finger-like voids combined with more pores underneath the top layer led to an increase in porosity and would promote the permeability of the MD membrane.
On the other hand, from higher magnification SEM images in Fig. 5(c1, d1, and e1), it can be seen that the gCN nanosheets are well embedded in the spherical nodules of the resultant mixed matrix membranes, as marked with yellow arrows in the corresponding images. Furthermore, it can be concluded that when using higher loadings of gCN (i.e., 0.045 wt.% and particularly 0.06 wt.%) in the polymer matrix, the aggregation of nanomaterials takes place.
"The location of Fig. 5"
Generally, membrane porosity is defined as the void space volume fraction of the membrane. The higher the porosity, the higher the permeate flux and consequently the lower the heat loss by conduction [10]. Anyway, there must be a counterbalance between the high porosity and the mechanical stability of the membranes because the porosity is restricted by the required mechanical strength. El-Bourawi et al. [6] have reported that the porosity for MD membranes typically stands within the range of 30–85% and indeed, the values obtained in this study lie within this range. The porosity amounts for the as-fabricated membranes are given in Table 1, which demonstrate an increase by adding the nanosheets into the polymer solution. The highest and the lowest porosity values of 84.27 \(\pm\) 1.84% and 53.01 \(\pm\) 3.45% were obtained for the PVDF membranes with 0.03 wt.% gCN loading and without nanofiller, respectively. It is likely that the higher nanosheets concentration increases the thermodynamic instability, accelerates the rate of precipitation in the water coagulation bath, and results in a membrane with higher porosity, as previously discussed in the section of SEM images analysis of the membranes. It is noteworthy that the membrane with 0.045 wt.% gCN has shown a porosity close but a little less (i.e., 80.88 \(\pm\) 1.2%) to that of the membrane embedded with 0.03 wt.% gCN. However, the increment in gCN content to 0.06 wt.% produces a far more viscous dope solution, which prevents the solution demixing process due to the enhanced kinetic hindrance, again as explained earlier in the SEM section, resulting in lower membrane pore size and porosity (Table 1). These results are in keeping with the calculated flux values that are rendered in the coming sections.
Figure 6 offers the two- and three-dimensional surface AFM images of the pristine and gCN embedded PVDF membranes at a scan size of 10 µm \(\times\) 10 µm. Although all fabricated membranes presented ridge and valley topology, it is visually observed that their surface topography has been affected by the addition of nanosheets in the polymer scaffold. The surface roughness parameters of the membranes obtained from AFM images were measured in a scan area of 10 µm \(\times\) 10 µm utilizing DME SPM software and the results are summarized in Table 1. It can be seen in Table 1 that the mixed matrix membranes possess a rougher appearance relative to the pristine PVDF membrane originating from the presence of the nanofillers. Here, the question may arise that the gCN nanosheets are not visible in the surface SEM images. However, it should be said in response that those same images prove the morphological changes of the surface and/or underneath the skin of the membranes due to the addition of the nanosheets. As we know, the rough surface is usually desirable for the MD process, whereas it enables higher free surface energy, and thus, more resistance against wetting [87].
According to Table 1, when the nanosheets are integrated into the membrane matrix, the Sa value is increased from 20.02 nm for the pure membrane to 26.51 nm for the membrane comprising 0.015 wt.% gCN. By adding the gCN content to 0.03 wt.%, the Sa value is considerably increased to 46.6 nm. By more adding gCN to 0.045 wt.% into the polymer network, the mean roughness is slightly decreased to 43.88 nm. And finally, the value of Sa reaches to 32.23 nm at the content of 0.06 wt.% gCN. These observations can be justified by the morphological alterations during the phase separation from two aspects of the number of surface pores and the size of subsurface spherical nodules, which were discussed earlier in the analysis of membranes surface and cross-sectional SEM images. Some authors have also pointed out that the entrapping of large quantities of the gCN nanomaterials (i.e., in the case of 0.06 wt.% gCN) into the PVDF matrix may lead to filling up the membrane surface pores, defects, or vacancies and accordingly, a smoother surface is obtained [50, 88, 89]. Anyhow, a large number of surface pores as well as smaller subsurface spherical nodules lead to the increased surface roughness by virtue of the increase in bumps and depressions and vice versa. It can be finally concluded from the data of Table 1 that the resultant surface roughness parameters might have been derived from a tradeoff between the number of surface pores and the subsurface spherical nodules.
"The location of Fig. 6"
Additionally, the size of the pores on the membrane surface is considered as an effective factor of surface roughness according to the reviewed literatures [89, 90]. Regarding the dependence of the roughness parameter on the Z value, when the surface of the membrane includes relatively larger pores, the vertical distance to be moved up and down by the piezoelectric scanner tip increases in comparison to shallow depressions, thus, the roughness parameter also increases [91, 92].
In order to further clarify this issue, the pore size distribution of the membranes was calculated from AFM images by using the Nanosurf Report (please see Fig. 7) and the mean surface pore sizes are presented in Table 1. The results reveal that the mean surface pore size of the PVDF membranes enhances after loading gCN nanosheets up to 0.03 wt.% and then follow a downward trend (Table 1). In other words, the pore size distribution shifts to the larger pore diameters by increasing the gCN concentration up to 0.03 wt.% in the dope solution and then it moves towards smaller pore diameters (Fig. 7). These evidences are in good agreement with the results of the surface roughness analysis.
"The location of Fig. 7"
The obtained contact angles are also in harmony with the resulting surface roughness parameters, so that the maximum contact angle has shown the maximum surface roughness [88]. The water contact angle (θ) on the surface of the membrane can quantify the wettability of the material and reflect the strength of the interaction on the interface [93]. As can be seen in Table 1, the water contact angle of the mixed matrix membranes is enhanced by increasing the content of the gCN nanosheets in the polymeric solution and then decreased by continuing this trend. Among the mixed matrix membranes, the highest contact angle belongs to the membrane incorporated with 0.03 wt.% of gCN (θ = 95.2 \(\pm\) 1º). The pristine PVDF membrane has also rendered the lowest contact angle (θ = 80.5 \(\pm\) 1º).
LEPw was measured to inquire about the effect of the gCN nanosheets on membrane resistance against wetting and the obtained values can be observed in Table 1. The PVDF membrane embedded with 0.03 wt.% gCN has demonstrated the highest LEPw equals 8 \(\pm\) 0.5 bar, making it a better candidate for MD. LEPw is strongly influenced by surface hydrophobicity and mean surface pore size [94]. A membrane with higher hydrophobicity and smaller mean surface pore size will possess a higher LEP value according to the Cantor-Laplace equation [6]. Although the PVDF membrane modified with 0.03 wt.% gCN has a larger mean surface pore size than all other membranes, the higher hydrophobicity of the former compensates for its larger pore size and leads to a more water-resistant membrane. From the measured values of LEPw in Table 1, it can be proven that all membranes behave in an identical manner.
Table 1
Characteristics of different fabricated membranes.
Membrane | Thickness (µm) | Mean surface pore size (nm) | Porosity (%) | Contact angle, θ (º) | LEPw (bar) | Roughness parameters (nm) |
Sa | Sq | Sz |
Pure PVDF | 26.9 \(\pm\) 0.84 | 289.63 | 53.01 \(\pm\) 3.45 | 80.5 \(\pm\) 1 | 6 \(\pm\) 0.5 | 20.02 | 25.19 | 198.01 |
gCN-PVDF 0.015 | 25 \(\pm\) 1.5 | 327.03 | 67.85 \(\pm\) 2.27 | 84 \(\pm\) 2 | 6.5 \(\pm\) 0.5 | 26.51 | 33.36 | 258.50 |
gCN-PVDF 0.03 | 29.3 \(\pm\) 1.6 | 486.28 | 84.27 \(\pm\) 1.84 | 95.2 \(\pm\) 1 | 8 \(\pm\) 0.5 | 46.60 | 62.13 | 472.35 |
gCN-PVDF 0.045 | 26.2 \(\pm\) 0.92 | 403.15 | 80.88 \(\pm\) 1.2 | 91.3 \(\pm\) 2 | 7.5 | 43.88 | 55.29 | 351.55 |
gCN-PVDF 0.06 | 30.6 \(\pm\) 1.1 | 356.5 | 71.73 \(\pm\) 2.07 | 87.6 \(\pm\) 2 | 7 | 32.23 | 40.25 | 333.73 |
"The location of Table 1"
MD is typically operated without hydraulic pressure and it is taken for granted that unlike pressure-driven membrane separation technologies, it has less membrane mechanical robustness requirements. Nevertheless, having sufficient mechanical resistance of MD membranes are still needed by virtue of withstanding the stress from compression in modules under the high-temperature long-term operation [30]. The effect of the gCN nanosheets on the mechanical behavior of the mixed matrix membranes against the pristine PVDF membrane was evaluated and the related data of stress-strain measurements are summarized in Table 2. As can be seen in Table 2, the ever-increasing tensile strength by increasing gCN content in the PVDF matrix occurs in all the mixed matrix membranes compared with the pure PVDF membrane. So that, tensile strength increases from 6.05 \(\pm\) 0.21 MPa for the pure PVDF membrane to its maximum value of 7.34 \(\pm\) 0.26 MPa for the membrane containing 0.06 wt.% gCN. In other words, the maximal increase by 21.32% is obtained for 0.06 wt.% gCN embedded membrane relative to the pure one. Thus, the increase in membranes' tensile strength is mainly interpreted by the presence of gCN nanosheets with high intrinsic mechanical stability within the polymer matrix [75, 95]. Also, the improvement of membranes' tensile strength may be due to the creation of morphological changes through introducing gCN nanosheets in the structure of the membranes. In line with the previous sentence, it should be said that the good reinforcement property of gCN is the fundamental reason to employ it as a filler material, as investigated by Shi and his coworkers [96]. They have developed sodium alginate (SA) nanocomposite films using 0.5-6 wt.% gCN and observed that with the addition of 6 wt.% gCN, the tensile strength of SA nanocomposite films was dramatically enhanced by 103%. At the same time, the Young’s modulus increased to an extremely high value of 3540 MPa. Regarding the latter, the cross-sectional SEM images had proved that when a high concentration of the nanosheets (with special emphasis on 0.06 wt.% gCN) was incorporated into the membrane scaffold, the sponge-like morphology became more apparent. Since membranes with a sponge-like structure possess superior mechanical attributes according to the literature reviews [97, 98], it is expected that the mechanical resistance of membranes with higher levels of gCN loading will increase.
Furthermore, the elongation at break increases with the addition of the nanofiller content into the polymeric matrix up to 0.03 wt.%. When the nanofiller content exceeded 0.03 wt.%, the elongation at break started reversing, probably due to the agglomeration of gCN nanosheets [95, 99]. The mixed matrix membrane with 0.03 wt.% gCN possesses the maximum level of elongation at break equal to 134.54 \(\pm\) 4.2%. The elongation at break of the mixed matrix membranes is still higher than that of the pure PVDF membranes even at high gCN content, confirming that the incorporating of gCN nanosheets into the PVDF matrix can effectively enhance the mechanical properties of mixed matrix membranes [96].
Table 2
Mechanical strength of different fabricated membranes.
Membrane | Tensile strength (MPa) | Elongation at break (%) |
Pure PVDF | 6.05 \(\pm\) 0.21 | 98.62 \(\pm\) 2.5 |
gCN-PVDF 0.015 | 6.21 \(\pm\) 0.23 | 116.44 \(\pm\) 6.5 |
gCN-PVDF 0.03 | 6.88 \(\pm\) 0.18 | 134.54 \(\pm\) 4.2 |
gCN-PVDF 0.045 | 7.16 \(\pm\) 0.11 | 129.08 \(\pm\) 3.68 |
gCN-PVDF 0.06 | 7.34 \(\pm\) 0.26 | 122.78 \(\pm\) 5.01 |
"The location of Table 2"
3.3. Membrane desalination performance using DCMD
The desalination performance of the as-prepared membranes was evaluated through the DCMD configuration under the temperature of 68°C and with a feed concentration of 35000 ppm. The obtained values for water flux and salt rejection of the pristine PVDF and mixed matrix membranes over a duration of 5 h are displayed in Fig. 8. As shown, the pure PVDF membrane exhibits the lowest flux value of 16.25 \(\pm\) 0.8 kg/m2h, while much higher fluxes are obtained for the mixed matrix membranes. By adding a small amount of nanofiller to 0.015 wt.%, the flux enhances to 20.64 \(\pm\) 0.95 kg/m2h. The membrane incorporated with 0.03 wt.% gCN renders the highest flux of 27.63 \(\pm\) 0.75 kg/m2h; however, at a loading of 0.045 and 0.06 wt.%, this drops to 25.03 \(\pm\) 0.85 and 21.94 \(\pm\) 0.7 kg/m2h, respectively. These findings conform truly to the alterations in pore size and porosity of the mentioned membranes (see Table 1). The treatment efficiency in the field of water and wastewater via gCN based hybrid membranes could be influenced by their physicochemical properties such as surface roughness, surface contact angle, pore size, and porosity, etc. [50]. While these morphological differences play a crucial role in the membranes' performance, other factors may also be influential. In order to figure out the flux improvement for membranes containing graphene-based nanofillers, various suggestions have been offered in the literatures [16, 17]. Based on the analogy with graphene-like nanomaterials structure, it can be assumed that gCN nanosheets are able to provide effectual sorption sites for water vapor but repel liquid water molecules. Moreover, the existence of the remaining terminal polar functional groups in the gCN structure (i.e., amino groups) can desirably alter the interaction between the membrane and water vapor molecules, thus enhancing the overall rate of vapor permeation [16]. Certainly, the combination of all these has been effective in the flux promotion of the membranes, but in some cases, one or more factors may prevail over the other(s). For example, in the cases of the 0.045 and 0.06 wt.% gCN embedded membranes, the morphological factors have overcome the gCN concentration.
Furthermore, the rejection of NaCl is > 99.9% for all tested membranes, while it is approximately 100% for the membrane comprising 0.03 wt.% gCN nanosheets. To have a better comparison between the permeate quality of the membranes, the permeate conductivity values after 5 h continuous desalting operation were arranged in tabular form, as can be observed in Table 3. The addition of gCN results in a lower permeate conductivity relative to the pure PVDF membrane. The comparison of the data in Tables 1 and 3 shows that the resultant permeate conductivities follows the contact angle values of the corresponding membranes. Accordingly, the lowest permeate conductivity, even lower than that for the coolant (i.e., 1.71 \(\pm\) 0.3 µS/cm), is achieved for 0.03 wt.% gCN embedded membrane with the highest contact angle value. The presence of the relatively hydrophobic gCN nanosheets in the structure of the mixed matrix membranes may be considered as the main cause for the reduction of the permeate conductivity in addition to the measured contact angle values.
In association with the surface wettability of gCN, many scientific theories have been proposed. All of them agree on the same concept and that is the relative hydrophobicity of this 2D nanosheet. Some references have pointed out that gCN possesses a hydrophobic surface [64, 73, 100, 101], and some others have used the phrase “the relative hydrophobic surface” for gCN [102–104]. Liu et al. [93] have studied the mechanisms of water desalination across the nanoporous multilayer gCN membranes with different pore diameters using non-equilibrium molecular dynamics simulations. They have reported that the surface of gCN is somewhat hydrophobic with a contact angle of about 70º. Cao and coworkers [103] have investigated the mechanisms of humic acid impact on the photocatalytic antialgal activities of gCN and copper oxide (CuO) nanoparticles doped gCN to harmful algae. They have introduced gCN as a hydrophobic nanomaterial with a contact angle equal to 77.4 \(\pm\) 0.9º. In addition, the performance of nanoparticulate gCN as an amphiphile has been evaluated in research by Xu et al. [105]. Amphiphiles refer to substances that own both hydrophilic and hydrophobic moieties. According to the authors, the amphiphilic feature of gCN is rooted in its intrinsic structural characteristics, i.e., the hydrophobic conjugated basal framework and hydrophilic edge groups and that’s why it is expected to be capable of dispersing hydrophobic solids in the aqueous phase.
Finally, to evaluate the long-time performance stability, the pristine and the optimum-performing membrane containing 0.03 wt.% gCN nanosheets were tested for 24 h. Results are demonstrated in Fig. 9 and offer a good stability of the 0.03 wt.% gCN embedded membrane under the test conditions, without diminution in salt rejection.
Table 3
Permeate conductivity of all tested membranes after five hours continuous operation. Feed conductivity was (50 \(\pm\) 3) \(\times\) 103 µS/cm.
Membrane | Permeate conductivity (µS/cm) |
Pure PVDF | 39.6 \(\pm\) 4 |
gCN-PVDF 0.015 | 26.5 \(\pm\) 3 |
gCN-PVDF 0.03 | 1.71 \(\pm\) 0.3 |
gCN-PVDF 0.045 | 8.12 \(\pm\) 2 |
gCN-PVDF 0.06 | 22.83 \(\pm\) 3 |
"The location of Fig. 8"
"The location of Fig. 9"
"The location of Table 3"