Next Article in Journal
Preparation and Characterization of Polyvinylpyrrolidone/Cellulose Nanocrystals Composites
Next Article in Special Issue
Graphene Based Poly(Vinyl Alcohol) Nanocomposites Prepared by In Situ Green Reduction of Graphene Oxide by Ascorbic Acid: Influence of Graphene Content and Glycerol Plasticizer on Properties
Previous Article in Journal
Antimicrobial Effects of Biogenic Nanoparticles
Previous Article in Special Issue
Peculiarities of Synthesis and Properties of Lignin–Silica Nanocomposites Prepared by Sol-Gel Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Protonation, Hydroxylamination, and Hydrazination of g-C3N4 on the Performance of Matrimid®/g-C3N4 Membranes

by
María Soto-Herranz
1,*,
Mercedes Sánchez-Báscones
1,
Antonio Hérnandez-Giménez
2,
José I. Calvo-Díez
2,
Jesús Martín-Gil
3 and
Pablo Martín-Ramos
4,*
1
Department of Agroforestry Sciences, ETSIIAA, University of Valladolid, Avenida de Madrid 44, 34004 Palencia, Spain
2
SMAP, UA-UVa_CSIC, Associated Research Unit to CSIC, Universidad de Valladolid, Paseo Belén 7, 47011 Valladolid, Spain
3
Agriculture and Forestry Engineering Department, ETSIIAA, Universidad de Valladolid, Avenida de Madrid 44, 34004 Palencia, Spain
4
Department of Agricultural and Environmental Sciences, Instituto Universitario de Investigación en Ciencias Ambientales (IUCA), EPS, Universidad de Zaragoza, Carretera de Cuarte s/n, 22071 Huesca, Spain
*
Authors to whom correspondence should be addressed.
Nanomaterials 2018, 8(12), 1010; https://doi.org/10.3390/nano8121010
Submission received: 29 October 2018 / Revised: 24 November 2018 / Accepted: 4 December 2018 / Published: 5 December 2018

Abstract

:
One of the challenges to continue improving polymeric membranes properties involves the development of novel chemically modified fillers, such as nitrogen-rich 2-D nanomaterials. Graphitic carbon nitride (g-C3N4) has attracted significant interest as a new class of these fillers. Protonation is known to afford it desirable functionalities to form unique architectures for various applications. In the work presented herein, doping of Matrimid® with protonated g-C3N4 to yield Matrimid®/g-C3N4 mixed matrix membranes was found to improve gas separation by enhancing the selectivity for CO2/CH4 by up to 36.9% at 0.5 wt % filler doping. With a view to further enhancing the contribution of g-C3N4 to the performance of the composite membrane, oxygen plasma and hydrazine monohydrate treatments were also assayed as alternatives to protonation. Hydroxylamination by oxygen plasma treatment increased the selectivity for CO2/CH4 by up to 52.2% (at 2 wt % doping) and that for O2/N2 by up to 26.3% (at 0.5 wt % doping). Hydrazination led to lower enhancements in CO2/CH4 separation, by up to 11.4%. This study suggests that chemically-modified g-C3N4 may hold promise as an additive for modifying the surface of Matrimid® and other membranes.

Graphical Abstract

1. Introduction

CO2 is significantly present in mixtures where CH4 is the major and valuable component. Its removal from CH4 mixtures is particularly important in processes such as biogas upgrading or natural gas sweetening. Conventional industrial methods used for such CO2 removal include adsorption [1], and water scrubbing and absorption [2] processes. Nonetheless, separation processes mediated by membranes can generally offer advantages over aforementioned techniques in terms of low capital cost, ease of processing, small footprint area, high energy efficiency, and ease of preparation and control [3,4].
Another closely related burgeoning field is that of oxygen production from air separation, since it can be used in many environmental applications, such as in the combustion enhancement of natural gas or in coal gasification [5]. Although cryogenic distillation and pressure swing adsorption are the mainstream technologies for oxygen-enriched air production, the use of membranes has been deemed as specially promising.
Basu et al. [6] and Zhang et al. [7] published two reviews on membrane materials used for CO2/CH4 separation. Among them, and apart from zeolite SAPO-34-based membranes [8,9,10,11,12], polyimides (PI) were claimed to exhibit the best combination of permeability and selectivity, thermal and chemical stability, and water resistance in biogas upgrading. In a similar fashion, glassy polyimides have also proven to exhibit excellent O2/N2 selectivity [13].
The most popular commercially available PI, 3,3’-4,4’-benzophenone tetracarboxylic-dianhydride diaminophenylindane, better known as Matrimid® 5218A, has been shown to combine a remarkable gas separation performance with other desirable features in terms of solubility, processability, thermal stability, chemical resistance, adhesion, film durability, and toughness. The state-of-the art based on this thermoplastic PI for gas separation applications has been recently covered in a thorough review by Castro-Muñoz et al. [14]. In terms of future trends to further enhance Matrimid’s properties, they insisted on the need to continue developing novel chemically modified fillers as a key challenge.
Amongst two-dimensional nanomaterials that are being explored for the development of high-performance separation membranes, discussed in a recent review by Zhu et al. [15], graphitic carbon nitride (g-C3N4) has attracted significant interest as a new class of filler. This mild bandgap metal-free organic semiconductor has remarkable applications as a catalyst for environmental pollution remediation [16,17].
Theoretical studies have preconized its suitability for H2 or He purification from other gases [18,19,20,21], and several experimental works have shown that it can be successfully used in the fabrication of hybrid membranes with enhanced properties. For instance, Hou et al. [22] reported a zeolitic imidazolate framework-8 (ZIF-8) hybrid membrane doped with g-C3N4 nanosheets for H2/CO2 separation, with an excellent selectivity. g-C3N4 nanosheets were also incorporated into the matrix of polymers of intrinsic microporosity (PIM-1) by Tian et al. [23], resulting in an increase in the selectivities for H2/CH4 and H2/N2.
As regards other membrane separation (non-gas) applications, Wang et al. [24] co-doped a polyamide (PA) forward osmosis membrane, based on a porous polyethersulfone substrate, with reduced graphene oxide and graphitic g-C3N4. By doing so, they obtained an osmotic water flux nearly 20% higher than that of the control membrane. Also in a PA nanofiltration membrane, modification with g-C3N4 nanosheets was shown to improve the permeability, antifouling properties, hydrophilicity, negative charge density, roughness and thermal stability [25]. Similar conclusions were reached for reverse osmosis PA membranes doped with acidified g-C3N4 by Gao et al. [26]. Doping g-C3N4 into a sodium alginate matrix or into poly(vinyl alcohol) also improved the performance of hybrid membranes for water/ethanol separation [27,28]; and doping of g-C3N4 co-impregnated with Cu+ and Fe2+ ions into a Pebax® membrane improved gasoline desulfurization [29].
Nonetheless, to the best of the authors’ knowledge, there have been no reports on the utilization of g-C3N4 as a filler for Matrimid® mixed matrix membranes (MMMs). The work presented herein addresses this literature gap by assessing the effects of protonation of g-C3N4 as well as its treatment with hydrazine monohydrate and with oxygen plasma on the performance of MMMs for CO2/CH4 and O2/N2 separation.

2. Materials and Methods

2.1. Materials

Matrimid® 5218 polymer (CAS no. 104983-64-4) was obtained from Huntsman Chemical, Advanced Materials Americas Inc. (Los Angeles, CA, USA). Melamine cyanurate (CAS no. 37640-57-6) was supplied by Ferro-plast S.r.l. (Vimodrone, Italy) with a purity superior to 99%. N-N-Dimethylacetamide (DMAc; CAS no. CAS: 127-19-5; ≥99%) and hydrazine monohydrate (CAS no. 7803-57-8; 98%) were purchased from Sigma-Aldrich (Munich, Germany). Reagents were used as received, without any further purification.

2.2. Methods

2.2.1. Synthesis of Graphitic Carbon Nitride (g-C3N4)

Graphitic carbon nitride was prepared by a modification of the thermal oxidation method [30], using melamine cyanurate as a starting material for the polymerization/condensation reaction [31]. The reagent was placed in a Vycor® glass vial and heated at 650 °C for 4 h under aeration, with a cooling/heating ramp of 10 °C·min−1. The resulting yellow powder was transferred to an agate mortar and ground. Subsequently, the samples were exposed to ultrasonic treatment for approximately 2 h using a Branson (FisherScientific, Hampton, NH, USA) Sonifier 450. Samples were dried in an oven at 105 °C for 24 h to remove moisture.

2.2.2. Synthesis of Protonated g-C3N4

Due to the presence of abundant –C–N– bonds in the g-C3N4 framework, it can be easily protonated by HCl, resulting in a modification of the surface charge, from a negatively to positively charged (Figure 1a) [32]. The process consisted in adding 2 g of pristine g-C3N4 sample to a 12 mL diluted 3% solution of HCl and treating it with ultrasound for 2 h. Samples were not subsequently washed. They were placed in an oven at 105 °C for approximately 24 h and were then ground.

2.2.3. Modifications of g-C3N4 with Oxygen Plasma and Hydrazine Monohydrate

g-C3N4 is a laminar material with non-oxidized aromatic regions and aliphatic regions containing phenolic, carboxyl, and oxygen epoxide groups that make it hydrophilic in aqueous media. This behavior can be modified through treatment with oxygen plasma or with hydrazine monohydrate.
In the former approach, interfacial forces are improved via plasma treatment, an ecologically benign process that does not produce liquid chemical residues. When the plasma is used with oxygen gas, it dissociates in order to generate oxygen-containing radicals [33], which incorporate –OH, as hydroxylamine groups, into the surface of g-C3N4, improving its dispersibility [34] (Figure 1b). The oxygen plasma treatment was conducted using a Harrick Plasma (Ithaca, NY, USA) PDC-002 apparatus. 150 mg of g-C3N4 were placed on a Petri dish on a quartz sample holder within the plasma cavity. The operating power remained at 10.2 W, the chamber pressure was 300 Torr with an O2 flow of 35 cm3·min−1, and the treatment time was 90 min.
In the later approach, diazanyl group modified g-C3N4-NHNH2 (Figure 1c) was obtained according to the procedure described by Chen et al. [35]: 1 g of as-synthesized g-C3N4 was mixed with 20 mL water and 4 mL hydrazine hydrate, followed by stirring at 80 °C for 40 min.

2.2.4. Preparation of the Matrimid/g-C3N4 Mixed Matrix Membranes

The Matrimid/g-C3N4 MMMs were prepared from a Matrimid commercial polymer solution doped with 0.5 wt % or 2 wt % g-C3N4 loading in 5 mL of DMAc solvent at adequate concentration (6% w/v) to achieve a reasonable viscosity. Working at such g-C3N4 loadings, previous work showed that measurable effects were obtained as compared with pristine Matrimid membranes, while much higher loadings led to an inhomogeneous distribution of g-C3N4 across the membrane film, thus compromising the mechanical stability of the resulting membranes. After stirring until complete homogenization, all solutions were sonicated for 5 min at an amplitude of 60% in a sonicator with a work/rest period of 20 s and 10 s. Next, membranes were prepared by the casting method following the protocol described by Recio et al. [36], using a glass plate at 25 °C. The resulting films were dried at 60 °C for 24 h to complete the removal of the solvent and were then placed in a vacuum oven, where they remained at 120 °C for 4 hours before raising the temperature to 180 °C for another 4–6 h. Finally, the membranes were separated from the glass plates and their thickness was determined using a Fischer (Sindelfingen, Germany) Dualscope MP0R. All thicknesses were in the 40–60 μm range. Three samples of each formulation were prepared, and at least two pieces of each membrane were tested. Their results were averaged, with a sample to sample variability in the usual range for laboratory prepared membranes (5–10%).

2.2.5. Characterization Methods

Vibrational information of the dopant and the Matrimid/g-C3N4 MMMs was retrieved by using a Thermo Scientific (Waltham, MA, USA) Nicolet iS50 Fourier-Transform Infrared (FTIR) spectrometer. 8–10 mm Ø samples were cut from the films and their transmittance was measured at room temperature in a scanning range between 410 and 4000 cm−1, with a 1 cm−1 spectral resolution and 64 scans.
The morphology of the as-prepared samples was examined using field emission scanning electron microscopy (FESEM), with a FEI (Hillsboro, OR, USA) QUANTA 200F device.
Thermogravimetric analysis (TG) was used to determine the thermal and/or oxidative stability of the materials. TG experiments were performed with a Mettler Toledo (Columbus, OH, USA) DMA/SDTA 861 device, heating the samples from 50 °C to 850 °C at a heating rate of 10 °C·min−1, under a N2 flow of 20 cm3·min−1.

2.2.6. Gas Separation Performance Measurement

The pure gas permeation tests were performed on an isochoric (constant volume, variable pressure) permeation system to measure the permeability properties of the membranes. Five pure gas species (He, N2, O2, CH4, and CO2) were applied as test gases, and the permeability and perm-selectivity of Matrimid/g-C3N4 MMMs were evaluated. Experiments were carried out at 35 °C using a constant feed pressure of 3 bar. To assure constant flow of each gas tested, desired pressure was maintained during 1 h and then flow measurements were taken by using a gas flow meter.
The permeability of a given membrane was calculated from Equation (1)
P = Q × L Δ P × A
where P is the permeability in barrer (1 barrer = (10−10 cm3 (STP) × cm)/(cm2 × s × cmHg)), Q is the volumetric flow rate [cm3 (STP)/s], L is the membrane thickness [cm], ΔP is the pressure difference between two sides of the membrane [cmHg], and A is the effective membrane area [cm2]. The perm-selectivity αA/B is defined as the ratio of the permeability coefficients of two gases A and B (PA and PB), according to Equation (2)
α A / B = P A P B

3. Results and Discussion

3.1. Vibrational Characterization

3.1.1. ATR-FTIR Spectra of the Modified g-C3N4 Fillers

The spectra of the protonated g-C3N4 samples revealed similar characteristic features to those found in the as-prepared g-C3N4, thus confirming that the structural integrity of g-C3N4 remained intact after protonation (Figure 2). Strong absorption bands could be observed in the 1200-1650 cm−1 range, arising from the skeletal stretching of C–N heterocycles and comprising both trigonal (N–(C)3) (full condensation) and bridging C–NH–C units (partial condensation) [37]. This would point at a successful development of an extended C–N–C network. The broad band ranging from 3000 to 3700 cm−1 was assigned to N–H and O–H stretching, due to the free amino groups and adsorbed hydroxyl species, respectively, whereas the sharp band at ca. 805 cm−1 was originated from the breathing vibration of tri-s-triazine units [38].
After oxygen plasma treatment, all the characteristic bands were retained (Figure 2), featuring bands at 800 cm−1 (assigned to the tri-s-triazine ring); and at 1200, 1309, 1395, 1455, 1533, and 1627 cm−1 (related to C–NH–C and N–(C) stretching modes). The peaks located at 1010 cm−1, 1079 cm−1, and above 3000 cm−1 corresponded to signals from N–OH groups. These results indicated that the oxygen plasma treatment led to the incorporation of N–OH on the surface of the C3N4.
The treatment of g-C3N4 with hydrazine monohydrate (Figure 2) appeared to decrease the signal intensity of all peaks with respect to the original untreated g-C3N4 sample or to that treated with oxygen plasma.

3.1.2. ATR-FTIR Spectra of Matrimid/g-C3N4 MMMs

A first comparison of the spectra of pure Matrimid and of the Matrimid MMMs doped with as-prepared or protonated g-C3N4 (Figure 3) evidenced shifts in the bands at 819 cm−1 (shifted to 824 cm−1), 849 cm−1 (to 858 cm−1), 919 cm−1 (to 924 cm−1), 1475 cm−1 (to 1489 cm−1), and 1606 cm−1 (to 1618 cm−1). These shifts were sufficient to suggest an interaction between Matrimid and g-C3N4. As regards other strong bands in the doped samples, they were found at 708 cm−1 (v(CNC) out-of-plane bending); 1090 cm−1 (v(CNC) transversal stretching); 1364 cm−1 (v(CNC) axial stretching); 1510 cm−1 (v(C=C) aromatic stretching); 1663 cm−1 (v(C=O) benzophenone carbonyl stretching); 1715 cm−1 and 1784 cm−1 (v(C=O) symmetric and asymmetric stretching, respectively); 2861 and 2924 cm−1 (aliphatic C–H stretching); and 2958 cm−1 (aromatic C–H stretching).
Differences in the nature of the dopant, i.e., protonated vs. non-protonated g-C3N4, mainly affected the band at 1672 cm−1 (amide C=O stretching), which, after protonation, was shifted to 1663 cm−1. This finding is important because it suggests that the carbonyl groups in Matrimid can act as electron acceptors from –NH+ units from exfoliated protonated g-C3N4, resulting in a material assembled through electrostatic interactions (this matter will be discussed in detail in Section 3.4).
In relation to the ATR-FTIR spectra of MMMs doped with oxygen plasma and hydrazine treated g-C3N4, also depicted in Figure 3, it is worth noting that no bands associated with the N–OH or –NHNH2 groups of g-C3N4 (which would result from those treatments) were observed. This may be due to low dopant doses, and would be in agreement with the findings of Chen et al. [35], who reported no remarkable differences between pristine g-C3N4 and g-C3N4-NHNH2. On the other hand, the band at 1606 cm−1, present in the pristine matrix, was found to disappear upon doping in both cases.

3.2. SEM Analysis

SEM was used to study the MMMs cross-section morphology. The micrographs of the neat Matrimid membrane (not shown) featured a smooth surface without obvious imperfections, in good agreement with those reported, for instance, by Ebadi Amooghin et al. [39].
Representative micrographs of Matrimid/g-C3N4 MMMs with different g-C3N4 loadings and sonication times are shown in Figure 4a–c. At low loadings (0.5 wt %, Figure 4a,b) and sonication times (30 min), the g-C3N4 dopant was typically well dispersed. However, at higher loadings (e.g., 10 wt %, Figure 4c) and moderate sonication times (4 h), several large clusters of g-C3N4 could be found. These clusters ranged in size from 20 to 50 nm. This would be in agreement with the findings of Tian et al. [23], who evaluated the improvement of gas separation performance of mixed matrix membranes composed of PIM-1 doped with different g-C3N4 loads as compared to pure PIM-1 membranes. They found that, using CHCl3 as a solvent and after sonication for 12 h, the g-C3N4 dopant was well dispersed throughout the matrix at low loading levels (0.5–1 wt %), and also observed the presence of partial agglomeration for higher filler contents (>1 wt %).
It is also worth noting that, in contrast with the neat Matrimid membrane, a scalloped morphology was observed for the MMMs doped with protonated g-C3N4 (Figure 4a,b). As noted by Venna et al. [40] (and references therein), it may be attributed to the formation of elongated polymer segments with increased plastic deformation of the polymer and can be regarded as an indication of a good interaction between the polymer and the filler, further supported by the absence of a sieve-in-a-cage morphology (i.e., when there are interfacial voids larger than the penetrating molecules, so a by-pass around the particles is produced, enhancing the permeability and reducing the apparent selectivity [41]) at each loading.
SEM analysis was also conducted upon Matrimid doping with 0.5 wt % of g-C3N4 treated with oxygen plasma (Figure 4d,e) or with hydrazine monohydrate (Figure 4f). In the hydroxylaminated g-C3N4-doped samples, one could observe a crater-like pattern in which the eye of each crater was formed by nanosized fillers, similar to that obtained for protonated g-C3N4, albeit more evident. The filler also appeared to be uniformly distributed along the polymer matrix. According to Ebadi Amooghin et al. [39], such crater-like morphology of MMMs arises from an adequate compatibility existing between two phases, possibly due to the interfacial stress concentrations because of the polymer matrix/filler consecutive debonding. Thus, this deformation of the polymer matrix was again indicative of a strong interaction of polymer chains and filler. On the other hand, the MMMs doped with hydrazinolyzed g-C3N4 did not show a significant modification of the smooth aspect of pristine Matrimid, thus suggesting a weaker interaction.

3.3. Studies on Thermal Stability of the Matrimid/g-C3N4 Membranes

The thermal stability of the Matrimid/g-C3N4 membranes was evaluated in comparison to pristine Matrimid (Figure 5). The thermograms showed that all the studied membranes were stable up to 470 °C and that only a small weight loss was observed between 200 °C and 250 °C (see Table 1), which could be attributed to residual solvent evaporation (the boiling point of DMAc is 202–204 °C). The inflection point for this first stage occurred at 351 °C for pristine Matrimid, and at lower temperatures for the MMMs membranes: 295 °C for Matrimid loaded with protonated g-C3N4, 247 °C for Matrimid doped with oxygen plasm-treated g-C3N4, and 287 °C for Matrimid with hydrazine-treated g-C3N4. Above 450 °C, a decomposition in two stages took place, with inflection points at around 517 °C and 591 °C (for pristine Matrimid they appeared at 518 °C and 580 °C, respectively). This second weight loss could be ascribed to the evolution of CO, CO2, and CH4 from the cleavage of the benzene ring pattern of the Matrimid matrix. Degradation of the Matrimid/g-C3N4 composites ended at 850 °C after a brief carbonization process.
Taking in account the delayed TG inflection points for MMMs vs. neat Matrimid at around 591 °C (588 °C for Matrimid/hydrazinated g-C3N4; 589 °C for Matrimid/hydroxylaminated g-C3N4 and 593 °C for Matrimid/protonated g-C3N4 vs. 580 °C for pristine Matrimid, as shown in Figure 5), it may be inferred that thermally stable structures were created in the membrane matrix after treatment with either g-C3N4(N–OH) or g-C3N4(N–H+).

3.4. CO2/CH4 and O2/N2 Gas Separation Performance of Matrimid/g-C3N4 MMMs

Apropos of the pristine Matrimid membrane, the CO2/CH4 selectivity value was within the range reported in other recent works (34.2 [42], 34.9 [43], 35.7 [44], 36.3 [45]), and so was the O2/N2 selectivity (6.5–6.8 [42,46,47]).
Doping of the Matrimid MMMs with protonated g-C3N particulates at 0.5 wt % increased the CO2/CH4 selectivity by 36.9% (from 36.3 to 49.6) as compared to the neat Matrimid membrane, although the O2/N2 selectivity decreased by 5.7% (from 7.0 to 6.6). Doping at 2 wt % led to a lower increase in the CO2/CH4 selectivity (by 16.5%, from 36.3 to 42.2), but the O2/N2 selectivity was preserved (Table 2).
The behavior for the MMMs doped with oxygen plasma-treated g-C3N4 treatment as a function of the filler doping values was the opposite: doping at 2 wt % led to a larger increase in the CO2/CH4 selectivity (by 52.2%, from 36.3 to 55.2) than doping at 0.5 wt % (4.4% increase). On the other hand, doping at 0.5 wt % led to a significant enhancement of O2/N2 selectivity (26.3% increase, from 7.0 to 8.9), while the increase for 2 wt % doping was only 2.7%.
In relation to the hydrazine treatment, it enhanced CO2/CH4 selectivities (by 7.5% and by 11.4% for 0.5 wt % and 2 wt % doping, respectively), while O2/N2 separation values remained similar to those of the Matrimid matrix (0.7% and −3.7% variation).
The results in the two latter treatments would be in agreement with observations by other authors, such as Tian et al. [23], who synthesized MMMs with a PIM matrix doped with g-C3N4, and an increase in the permeability values was observed as the filler concentration was increased to reach >1 wt %.
The enhancement of gas separation in the Matrimid membranes after doping with modified g-C3N4 may be referred either to an increase of hydrogen bonding in the Matrimid/g-C3N4 composite structure or to cross-linking modifications in Matrimid mediated by g-C3N4.
The first hypothesis suggests that the good results obtained for Matrimid MMMs doped with protonated g-C3N4 would be referred to, on the one hand, their exfoliated nature and the associated surface gain for interaction, and, on the other hand, to the presence of –NH+ groups capable of bonding, through hydrogen bonding, with the carbonyl groups of the Matrimid matrix. The benzene rings of both g-C3N4 and Matrimid would lend themselves to the formation of “π-staking” (Figure 6a).
In the case of g-C3N4 treated with oxygen plasma, designed to provide hydroxylamine N–OH groups that would bond with the carbonyl groups of Matrimid, it is unclear whether interactions between filler and matrix would be mediated by subsequent hydrogen bonding between such groups. There is also the possibility of formation of charge transfer between the –OH (behaving as electron acceptor group) and the aromatic nucleus in Matrimid (behaving as an electron donor group), as depicted in Figure 6b.
In the case of Matrimid/hydrazinated g-C3N4 MMM, its behavior can be associated to a potential cross-linking effect of the g-C3N4(NH–NH2) filler. Addition of H2N–R–NH2 fillers has been reported to lead to cross-linking modification of polyimides [49], increasing chain packing and inhibiting the intra-segmental and inter-segmental mobilities of the matrix, resulting in higher gas selectivity [50]. For instance, p-phenylenediamine (PPD) was reported to be a good cross-linking agent for Matrimid, showing a remarkable enhancement of selectivity for O2/N2 as compared to the unmodified membrane (from 6.2 to 10.0 due to the affinity of nitrogen-containing molecules towards oxygen) [13]. Thus, the hydrazinated g-C3N4 discussed herein could be regarded as an alternative to PPD, albeit the associated enhancement of gas selectivity for O2/N2 would be lower.

4. Conclusions

Modified g-C3N4 was assessed as a filler for the doping of Matrimid matrices with a view to enhancing their gas separation properties. The resulting Matrimid/g-C3N4 MMMs were characterized through ATR-FTIR vibrational analysis, SEM microscopy, thermal analysis techniques, and gas perm-selectivity assays. Due to the strong interfacial interactions among the modified g-C3N4 and the Matrimid matrix, the hybrid nanocomposite membranes featured high swelling resistance and mechanical stability. In the assays presented herein, the doping of Matrimid with modified g-C3N4 led to an enhancement of CO2/CH4 selectivity—as compared to that of the pure Matrimid membrane—in all cases. In particular, MMMs doped with 0.5 wt % protonated g-C3N4 and with 2 wt % hydroxylaminated g-C3N4 showed the best CO2/CH4 separation performances (with an increase by 36.9% and 52.2% vs. neat Matrimid, respectively). On the other hand, doping with 0.5 wt % oxygen plasma-treated g-C3N4 improved the O2/N2 selectivity by 26.3%, as compared to pristine Matrimid. Thus, the chemical modification of g-C3N4 may hold promise as an efficient pathway toward the doping of Matrimid and other membranes.

Author Contributions

Conceptualization, A.H.-G., J.M.-G., and P.M.-R.; Data curation, M.S.-H.; Formal analysis, M.S.-H., J.M.-G., and P.M.-R.; Funding acquisition, M.S.-B., A.H.-G., and J.I.C.-D.; Investigation, M.S.-H., J.M.-G., and P.M.-R.; Methodology, A.H.-G. and J.M.-G.; Resources, A.H.-G., J.I.C.-D., and J.M.-G.; Supervision, M.S.-B., A.H.-G., and J.M.-G.; Validation, J.M.-G. and P.M.-R.; Visualization, M.S.-H. and P.M.-R.; Writing—original draft, M.S.-H., J.M.-G., and P.M.-R.; Writing—review & editing, M.S.-H. and P.M.-R.

Funding

This research was funded by the European Union, project LIFE+ AMMONIA TRAPPING (LIFE15-ENV/ES/000284); and by Spanish MINECO, projects MAT2016-76413-C2-R1 and MAT2016-76413-C2-R2. The APC was funded by LIFE+ AMMONIA TRAPPING (LIFE15-ENV/ES/000284) project.

Acknowledgments

The authors would like to gratefully acknowledge Manuel Avella-Romero for the SEM characterization at the Microscopy Facilities of the University of Valladolid; Cristina Álvarez and Ángel Emilio Lozano for the thermal characterization carried out at ICTP-CSIC; and Sara Rodriguez for the gas separation measurements. The guidance and help provided by Carla Aguilar-Lugo is also acknowledged. P.M.-R. acknowledges the financial support of Santander Universidades through “Becas Iberoamérica. Santander Investigación—Santander Universidades, España” program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Drioli, E.; Romano, M. Progress and new perspectives on integrated membrane operations for sustainable industrial growth. Ind. Eng. Chem. Res. 2001, 40, 1277–1300. [Google Scholar] [CrossRef]
  2. Tuinier, M.J.; van Sint Annaland, M.; Kramer, G.J.; Kuipers, J.A.M. Cryogenic CO2 capture using dynamically operated packed beds. Chem. Eng. Sci. 2010, 65, 114–119. [Google Scholar] [CrossRef]
  3. Brunetti, A.; Sun, Y.; Caravella, A.; Drioli, E.; Barbieri, G. Process intensification for greenhouse gas separation from biogas: More efficient process schemes based on membrane-integrated systems. Int. J. Greenh. Gas Control 2015, 35, 18–29. [Google Scholar] [CrossRef]
  4. Dorosti, F.; Omidkhah, M.; Abedini, R. Fabrication and characterization of Matrimid/MIL-53 mixed matrix membrane for CO2/CH4 separation. Chem. Eng. Res. Des. 2014, 92, 2439–2448. [Google Scholar] [CrossRef]
  5. Burdyny, T.; Struchtrup, H. Hybrid membrane/cryogenic separation of oxygen from air for use in the oxy-fuel process. Energy 2010, 35, 1884–1897. [Google Scholar] [CrossRef]
  6. Basu, S.; Khan, A.L.; Cano-Odena, A.; Liu, C.; Vankelecom, I.F.J. Membrane-based technologies for biogas separations. Chem. Soc. Rev. 2010, 39, 750–768. [Google Scholar] [CrossRef]
  7. Zhang, Y.; Sunarso, J.; Liu, S.; Wang, R. Current status and development of membranes for CO2/CH4 separation: A review. Int. J. Greenh. Gas Control 2013, 12, 84–107. [Google Scholar] [CrossRef]
  8. Carreon, M.A.; Li, S.; Falconer, J.L.; Noble, R.D. Alumina-supported SAPO-34 membranes for CO2/CH4 separation. J. Am. Chem. Soc. 2008, 130, 5412–5413. [Google Scholar] [CrossRef]
  9. Carreon, M.A.; Li, S.; Falconer, J.L.; Noble, R.D. SAPO-34 seeds and membranes prepared using multiple structure directing agents. Adv. Mater. 2008, 20, 729–732. [Google Scholar] [CrossRef]
  10. Li, S.; Carreon, M.A.; Zhang, Y.; Funke, H.H.; Noble, R.D.; Falconer, J.L. Scale-up of SAPO-34 membranes for CO2/CH4 separation. J. Membr. Sci. 2010, 352, 7–13. [Google Scholar] [CrossRef]
  11. Li, S. SAPO-34 membranes for CO2/CH4 separation. J. Membr. Sci. 2004, 241, 121–135. [Google Scholar] [CrossRef]
  12. Falconer, J.L.; Carreon, M.A.; Li, S.; Noble, R.D. Synthesis of Zeolites and Zeolite Membranes Using Multiple Structure Directing Agents. U.S. Patent 8,302,782 B2, 6 November 2012. [Google Scholar]
  13. Aziz, F.; Ismail, A.F. Preparation and characterization of cross-linked Matrimid® membranes using para-phenylenediamine for O2/N2 separation. Sep. Purif. Technol. 2010, 73, 421–428. [Google Scholar] [CrossRef]
  14. Castro-Muñoz, R.; Martin-Gil, V.; Ahmad, M.Z.; Fíla, V. Matrimid® 5218 in preparation of membranes for gas separation: Current state-of-the-art. Chem. Eng. Commun. 2017, 205, 161–196. [Google Scholar] [CrossRef]
  15. Zhu, J.; Hou, J.; Uliana, A.; Zhang, Y.; Tian, M.; Van der Bruggen, B. The rapid emergence of two-dimensional nanomaterials for high-performance separation membranes. J. Mater. Chem. A 2018, 6, 3773–3792. [Google Scholar] [CrossRef]
  16. Mamba, G.; Mishra, A.K. Graphitic carbon nitride (g-C3N4) nanocomposites: A new and exciting generation of visible light driven photocatalysts for environmental pollution remediation. Appl. Catal. B 2016, 198, 347–377. [Google Scholar] [CrossRef]
  17. Ding, F.; Yang, D.; Tong, Z.; Nan, Y.; Wang, Y.; Zou, X.; Jiang, Z. Graphitic carbon nitride-based nanocomposites as visible-light driven photocatalysts for environmental purification. Environ. Sci. Nano 2017, 4, 1455–1469. [Google Scholar] [CrossRef]
  18. De Silva, S.W.; Du, A.; Senadeera, W.; Gu, Y. Strained graphitic carbon nitride for hydrogen purification. J. Membr. Sci. 2017, 528, 201–205. [Google Scholar] [CrossRef]
  19. Ji, Y.; Dong, H.; Lin, H.; Zhang, L.; Hou, T.; Li, Y. Heptazine-based graphitic carbon nitride as an effective hydrogen purification membrane. RSC Adv. 2016, 6, 52377–52383. [Google Scholar] [CrossRef]
  20. Li, F.; Qu, Y.; Zhao, M. Efficient helium separation of graphitic carbon nitride membrane. Carbon 2015, 95, 51–57. [Google Scholar] [CrossRef]
  21. Qu, Y.; Li, F.; Zhao, M. Theoretical design of highly efficient CO2/N2 separation membranes based on electric quadrupole distinction. J. Phys. Chem. C 2017, 121, 17925–17931. [Google Scholar] [CrossRef]
  22. Hou, J.; Wei, Y.; Zhou, S.; Wang, Y.; Wang, H. Highly efficient H2/CO2 separation via an ultrathin metal-organic framework membrane. Chem. Eng. Sci. 2018, 182, 180–188. [Google Scholar] [CrossRef]
  23. Tian, Z.; Wang, S.; Wang, Y.; Ma, X.; Cao, K.; Peng, D.; Wu, X.; Wu, H.; Jiang, Z. Enhanced gas separation performance of mixed matrix membranes from graphitic carbon nitride nanosheets and polymers of intrinsic microporosity. J. Membr. Sci. 2016, 514, 15–24. [Google Scholar] [CrossRef]
  24. Wang, Y.; Ou, R.; Wang, H.; Xu, T. Graphene oxide modified graphitic carbon nitride as a modifier for thin film composite forward osmosis membrane. J. Membr. Sci. 2015, 475, 281–289. [Google Scholar] [CrossRef]
  25. Chen, J.; Li, Z.; Wang, C.; Wu, H.; Liu, G. Synthesis and characterization of g-C3N4 nanosheet modified polyamide nanofiltration membranes with good permeation and antifouling properties. RSC Adv. 2016, 6, 112148–112157. [Google Scholar] [CrossRef]
  26. Gao, X.; Li, Y.; Yang, X.; Shang, Y.; Wang, Y.; Gao, B.; Wang, Z. Highly permeable and antifouling reverse osmosis membranes with acidified graphitic carbon nitride nanosheets as nanofillers. J. Mater. Chem. A 2017, 5, 19875–19883. [Google Scholar] [CrossRef]
  27. Cao, K.; Jiang, Z.; Zhang, X.; Zhang, Y.; Zhao, J.; Xing, R.; Yang, S.; Gao, C.; Pan, F. Highly water-selective hybrid membrane by incorporating g-C3N4 nanosheets into polymer matrix. J. Membr. Sci. 2015, 490, 72–83. [Google Scholar] [CrossRef]
  28. Wang, J.; Li, M.; Zhou, S.; Xue, A.; Zhang, Y.; Zhao, Y.; Zhong, J.; Zhang, Q. Graphitic carbon nitride nanosheets embedded in poly(vinyl alcohol) nanocomposite membranes for ethanol dehydration via pervaporation. Sep. Purif. Technol. 2017, 188, 24–37. [Google Scholar] [CrossRef]
  29. Ding, H.; Pan, F.; Mulalic, E.; Gomaa, H.; Li, W.; Yang, H.; Wu, H.; Jiang, Z.; Wang, B.; Cao, X.; et al. Enhanced desulfurization performance and stability of Pebax membrane by incorporating Cu+ and Fe2+ ions co-impregnated carbon nitride. J. Membr. Sci. 2017, 526, 94–105. [Google Scholar] [CrossRef]
  30. Niu, P.; Zhang, L.; Liu, G.; Cheng, H.-M. Graphene-Like Carbon Nitride Nanosheets for Improved Photocatalytic Activities. Adv. Funct. Mater. 2012, 22, 4763–4770. [Google Scholar] [CrossRef]
  31. Dante, R.C.; Martín-Ramos, P.; Sánchez-Arévalo, F.M.; Huerta, L.; Bizarro, M.; Navas-Gracia, L.M.; Martín-Gil, J. Synthesis of crumpled nanosheets of polymeric carbon nitride from melamine cyanurate. J. Solid State Chem. 2013, 201, 153–163. [Google Scholar] [CrossRef]
  32. Ong, W.-J.; Tan, L.-L.; Chai, S.-P.; Yong, S.-T.; Mohamed, A.R. Surface charge modification via protonation of graphitic carbon nitride (g-C3N4) for electrostatic self-assembly construction of 2D/2D reduced graphene oxide (rGO)/g-C3N4 nanostructures toward enhanced photocatalytic reduction of carbon dioxide to methane. Nano Energy 2015, 13, 757–770. [Google Scholar] [CrossRef]
  33. Park, O.-K.; Young Kim, W.; Min Kim, S.; You, N.-H.; Jeong, Y.; Su Lee, H.; Ku, B.-C. Effect of oxygen plasma treatment on the mechanical properties of carbon nanotube fibers. Mater. Lett. 2015, 156, 17–20. [Google Scholar] [CrossRef]
  34. Bu, X.; Li, J.; Yang, S.; Sun, J.; Deng, Y.; Yang, Y.; Wang, G.; Peng, Z.; He, P.; Wang, X.; et al. Surface modification of C3N4 through oxygen-plasma treatment: A simple way toward excellent hydrophilicity. ACS Appl. Mater. Interfaces 2016, 8, 31419–31425. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, Y.; Lin, B.; Wang, H.; Yang, Y.; Zhu, H.; Yu, W.; Basset, J.-M. Surface modification of g-C3N4 by hydrazine: Simple way for noble-metal free hydrogen evolution catalysts. Chem. Eng. J. 2016, 286, 339–346. [Google Scholar] [CrossRef]
  36. Recio, R.; Lozano, Á.E.; Prádanos, P.; Marcos, Á.; Tejerina, F.; Hernández, A. Effect of fractional free volume and Tg on gas separation through membranes made with different glassy polymers. J. Appl. Polym. Sci. 2008, 107, 1039–1046. [Google Scholar] [CrossRef]
  37. Dong, F.; Zhao, Z.; Xiong, T.; Ni, Z.; Zhang, W.; Sun, Y.; Ho, W.-K. In situ construction of g-C3N4/g-C3N4 metal-free heterojunction for enhanced visible-light photocatalysis. ACS Appl. Mater. Interfaces 2013, 5, 11392–11401. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, Y.; Mori, T.; Niu, L.; Ye, J. Non-covalent doping of graphitic carbon nitride polymer with graphene: Controlled electronic structure and enhanced optoelectronic conversion. Energy Environ. Sci. 2011, 4, 4517. [Google Scholar] [CrossRef]
  39. Ebadi Amooghin, A.; Omidkhah, M.; Kargari, A. The effects of aminosilane grafting on NaY zeolite–Matrimid®5218 mixed matrix membranes for CO2/CH4 separation. J. Membr. Sci. 2015, 490, 364–379. [Google Scholar] [CrossRef]
  40. Venna, S.R.; Lartey, M.; Li, T.; Spore, A.; Kumar, S.; Nulwala, H.B.; Luebke, D.R.; Rosi, N.L.; Albenze, E. Fabrication of MMMs with improved gas separation properties using externally-functionalized MOF particles. J. Mater. Chem. A 2015, 3, 5014–5022. [Google Scholar] [CrossRef]
  41. Zornoza, B.; Casado, C.; Navajas, A. Advances in hydrogen separation and purification with membrane technology. In Renewable Hydrogen Technologies; Gandía, L.M., Arzamendi, G., Diéguez, P.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 245–268. [Google Scholar]
  42. Shao, L.; Liu, L.; Cheng, S.-X.; Huang, Y.-D.; Ma, J. Comparison of diamino cross-linking in different polyimide solutions and membranes by precipitation observation and gas transport. J. Membr. Sci. 2008, 312, 174–185. [Google Scholar] [CrossRef]
  43. Loloei, M.; Omidkhah, M.; Moghadassi, A.; Amooghin, A.E. Preparation and characterization of Matrimid® 5218 based binary and ternary mixed matrix membranes for CO2 separation. Int. J. Greenh. Gas Control 2015, 39, 225–235. [Google Scholar] [CrossRef]
  44. Rahmani, M.; Kazemi, A.; Talebnia, F.; Abbasszadeh Gamali, P. Fabrication and characterization of brominated Matrimid® 5218 membranes for CO2/CH4 separation: Application of response surface methodology (RSM). e-Polymers 2016, 16, 481–492. [Google Scholar] [CrossRef]
  45. Ebadi Amooghin, A.; Omidkhah, M.; Sanaeepur, H.; Kargari, A. Preparation and characterization of Ag+ ion-exchanged zeolite-Matrimid®5218 mixed matrix membrane for CO2/CH4 separation. J. Energy Chem. 2016, 25, 450–462. [Google Scholar] [CrossRef]
  46. Sterescu, D.M.; Stamatialis, D.F.; Wessling, M. Boltorn-modified polyimide gas separation membranes. J. Membr. Sci. 2008, 310, 512–521. [Google Scholar] [CrossRef]
  47. Clausi, D.T.; Koros, W.J. Formation of defect-free polyimide hollow fiber membranes for gas separations. J. Membr. Sci. 2000, 167, 79–89. [Google Scholar] [CrossRef]
  48. Tejero, R.; Lozano, Á.E.; Álvarez, C.; de Abajo, J. Synthesis, characterization, and evaluation of novel polyhydantoins as gas separation membranes. J. Polym. Sci. Part A Polym. Chem. 2013, 51, 4052–4060. [Google Scholar] [CrossRef]
  49. Yong, W.F.; Li, F.Y.; Chung, T.-S.; Tong, Y.W. Highly permeable chemically modified PIM-1/Matrimid membranes for green hydrogen purification. J. Mater. Chem. A 2013, 1, 13914–13925. [Google Scholar] [CrossRef]
  50. Xiao, Y.; Chung, T.; Guan, H.; Guiver, M. Synthesis, cross-linking and carbonization of co-polyimides containing internal acetylene units for gas separation. J. Membr. Sci. 2007, 302, 254–264. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Scheme of the synthesis of: (a) protonated g-C3N4; (b) oxygen plasm treatment used to introduce hydroxylamine structure (N-OH) on the surface of g-C3N4 (hydroxylaminated g-C3N4); and (c) hydrazinated g-C3N4.
Figure 1. Scheme of the synthesis of: (a) protonated g-C3N4; (b) oxygen plasm treatment used to introduce hydroxylamine structure (N-OH) on the surface of g-C3N4 (hydroxylaminated g-C3N4); and (c) hydrazinated g-C3N4.
Nanomaterials 08 01010 g001
Figure 2. ATR-FTIR spectra of: (a) pristine g-C3N4, (b) protonated g-C3N4, (c) hydroxylaminated g-C3N4, and (d) hydrazinated g-C3N4.
Figure 2. ATR-FTIR spectra of: (a) pristine g-C3N4, (b) protonated g-C3N4, (c) hydroxylaminated g-C3N4, and (d) hydrazinated g-C3N4.
Nanomaterials 08 01010 g002
Figure 3. ATR-FTIR spectra of (a) a pristine Matrimid matrix and of Matrimid MMMs doped at 2 wt % with different fillers: (b) as-prepared g-C3N4, (c) protonated g-C3N4, (d) oxygen plasma-treated g-C3N4, and (e) hydrazinated g-C3N4.
Figure 3. ATR-FTIR spectra of (a) a pristine Matrimid matrix and of Matrimid MMMs doped at 2 wt % with different fillers: (b) as-prepared g-C3N4, (c) protonated g-C3N4, (d) oxygen plasma-treated g-C3N4, and (e) hydrazinated g-C3N4.
Nanomaterials 08 01010 g003
Figure 4. Cross-sectional SEM micrographs for MMMs doped with g-C3N4 subjected to different treatments and prepared with different sonication times: Matrimid membrane doped with 0.5 wt % protonated g-C3N4 at 3000× (a) and 6000× (b) magnification; Matrimid membrane doped with 10 wt % protonated g-C3N4 loading after 4 hours of sonication (c) at 3000×; Matrimid loaded with 0.5 wt % g-C3N4 treated with oxygen plasma at 3000× (d) and 10000× (e); Matrimid loaded with 0.5 wt % hydrazine monohydrate-treated g-C3N4 at 3000× (f). Micrographs of MMMs at 2 wt % loading were analogous to the ones at 0.5 wt %.
Figure 4. Cross-sectional SEM micrographs for MMMs doped with g-C3N4 subjected to different treatments and prepared with different sonication times: Matrimid membrane doped with 0.5 wt % protonated g-C3N4 at 3000× (a) and 6000× (b) magnification; Matrimid membrane doped with 10 wt % protonated g-C3N4 loading after 4 hours of sonication (c) at 3000×; Matrimid loaded with 0.5 wt % g-C3N4 treated with oxygen plasma at 3000× (d) and 10000× (e); Matrimid loaded with 0.5 wt % hydrazine monohydrate-treated g-C3N4 at 3000× (f). Micrographs of MMMs at 2 wt % loading were analogous to the ones at 0.5 wt %.
Nanomaterials 08 01010 g004
Figure 5. TG (solid line, y-axis on the left side of the graph) and DTG (dashed line, y-axis on the right side of the graph) curves for neat Matrimid and Matrimid-based membranes doped with 0.5 wt % of protonated g-C3N4, hydroxylaminated g-C3N4, and hydrazinated g-C3N4. The inset shows the TG/DTG curves for the untreated g-C3N4 filler.
Figure 5. TG (solid line, y-axis on the left side of the graph) and DTG (dashed line, y-axis on the right side of the graph) curves for neat Matrimid and Matrimid-based membranes doped with 0.5 wt % of protonated g-C3N4, hydroxylaminated g-C3N4, and hydrazinated g-C3N4. The inset shows the TG/DTG curves for the untreated g-C3N4 filler.
Nanomaterials 08 01010 g005
Figure 6. (a) Linking modes between protonated g-C3N4 and Matrimid; (b) a possible linking mode between oxygen plasma-treated g-C3N4 and Matrimid.
Figure 6. (a) Linking modes between protonated g-C3N4 and Matrimid; (b) a possible linking mode between oxygen plasma-treated g-C3N4 and Matrimid.
Nanomaterials 08 01010 g006
Table 1. TG curves features for neat Matrimid, the g-C3N4 filler and the three MMMs under study
Table 1. TG curves features for neat Matrimid, the g-C3N4 filler and the three MMMs under study
Pristine Matrimidg-C3N4Protonated g-C3N4Oxygen Plasma-Treated g-C3N4Hydrazinated g-C3N4
Weight Loss (%)Midpoint (°C)Weight Loss (%)Midpoint (°C)Weight Loss (%)Midpoint (°C)Weight Loss (%)Midpoint (°C)Weight Loss (%)Midpoint (°C)
2.828797.67004.229.42.702842.2280
23.451623.351824.851825.9517
14.960915.961117.061117.1611
Table 2. Permeabilities/selectivities of the Matrimid/g-C3N4 MMMs for He, N2, O2, CH4, CO2, O2/N2, and CO2/CH4 as a function of different filler treatments at different doping levels (0.5 wt % and 2 wt %)
Table 2. Permeabilities/selectivities of the Matrimid/g-C3N4 MMMs for He, N2, O2, CH4, CO2, O2/N2, and CO2/CH4 as a function of different filler treatments at different doping levels (0.5 wt % and 2 wt %)
MMMPermeability (Barrer)SelectivityRef.
HeN2O2CH4CO2O2/N2CO2/CH4
Matrimid (control)22.000.271.900.248.707.036.3[48]
Matrimid/protonated g-C3N4 0.5 wt %23.530.261.700.157.696.649.6This study
Matrimid/protonated g-C3N4 2 wt %20.170.201.440.156.457.142.2This study
Matrimid/hydroxylaminated g-C3N4 0.5 wt %24.120.191.710.217.868.937.9This study
Matrimid/hydroxylaminated g-C3N4 2 wt %23.630.231.690.137.157.255.2This study
Matrimid/hydrazinated g-C3N4 0.5 wt %20.600.201.440.176.567.139.0This study
Matrimid/hydrazinated g-C3N4 2 wt %18.650.201.360.156.196.840.4This study

Share and Cite

MDPI and ACS Style

Soto-Herranz, M.; Sánchez-Báscones, M.; Hérnandez-Giménez, A.; Calvo-Díez, J.I.; Martín-Gil, J.; Martín-Ramos, P. Effects of Protonation, Hydroxylamination, and Hydrazination of g-C3N4 on the Performance of Matrimid®/g-C3N4 Membranes. Nanomaterials 2018, 8, 1010. https://doi.org/10.3390/nano8121010

AMA Style

Soto-Herranz M, Sánchez-Báscones M, Hérnandez-Giménez A, Calvo-Díez JI, Martín-Gil J, Martín-Ramos P. Effects of Protonation, Hydroxylamination, and Hydrazination of g-C3N4 on the Performance of Matrimid®/g-C3N4 Membranes. Nanomaterials. 2018; 8(12):1010. https://doi.org/10.3390/nano8121010

Chicago/Turabian Style

Soto-Herranz, María, Mercedes Sánchez-Báscones, Antonio Hérnandez-Giménez, José I. Calvo-Díez, Jesús Martín-Gil, and Pablo Martín-Ramos. 2018. "Effects of Protonation, Hydroxylamination, and Hydrazination of g-C3N4 on the Performance of Matrimid®/g-C3N4 Membranes" Nanomaterials 8, no. 12: 1010. https://doi.org/10.3390/nano8121010

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop