Thin Film Nanocomposite Membrane with the Minimum Amount of

5 days ago - The presence and layout of the LS-MIL-101(Cr) monolayer in the TFN membrane was confirmed by STEM-HAADF images, and XPS, EDS, EELS and AF...
1 downloads 12 Views 3MB Size
Subscriber access provided by UNIV OF MISSOURI ST LOUIS

Article

Thin Film Nanocomposite Membrane with the Minimum Amount of MOF by the Langmuir-Schaefer Technique for Nanofiltration Marta Navarro, Javier Benito, Lorena Paseta, Ignacio Gascón, Joaquin Coronas, and Carlos Téllez ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.7b17477 • Publication Date (Web): 15 Dec 2017 Downloaded from http://pubs.acs.org on December 18, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

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

ACS Applied Materials & Interfaces

Thin Film Nanocomposite Membrane with the Minimum Amount of MOF by the LangmuirSchaefer Technique for Nanofiltration Marta Navarro,†* Javier Benito,§ Lorena Paseta,† Ignacio Gascón,§ Joaquín Coronas,† Carlos Téllez, †* † Instituto de Nanociencia de Aragón (INA) and Chemical and Environmental Engineering Department, Universidad de Zaragoza, 50018 Zaragoza, Spain. *Email: [email protected]; [email protected]. § Departamento de Química Física, Facultad de Ciencias, Universidad de Zaragoza, 50009 Zaragoza, Spain.

Keywords: Langmuir-Schaefer (LS) monolayer, Metal-organic framework (MOF), Controlled positioning, Thin film nanocomposite (TFN) membrane, Organic solvent nanofiltration (OSN)

ACS Paragon Plus Environment

1

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

Page 2 of 35

Abstract

An innovative procedure for positioning a monolayer of hydrophilic metal-organic framework (MOF) MIL-101(Cr) nanoparticles (NPs) in thin film nanocomposite (TFN) membranes has been implemented by transferring a Langmuir-Schaefer (LS) film of the MOF in between the polyamide thin layer at the top and the cross-linked asymmetric polyimide (P84®) support at the bottom. The presence and layout of the LS-MIL-101(Cr) monolayer in the TFN membrane was confirmed by STEM-HAADF images, and XPS, EDS, EELS and AFM analysis. This methodology requires the smallest amount of MOF reported to date, 3.8 µg·cm-2, and promotes the formation of a defect-free ultrathin MOF film. Whereas conventional TFN membranes tend to show MOF agglomerates that could contribute to the formation of unselective defects, LSTFN membranes, characterized by a homogeneous and continuous MOF coating, exhibit an optimal membrane performance, without a significant decrease in selectivity. Outstanding methanol permeance, one of the best results reported to date, of 10.1 ±0.5 L·m-2·h-1·bar-1 when filtering sunset yellow and of 9.5 ±2.1 L·m-2·h-1·bar-1 when filtering rose bengal, have been achieved in LS-TFN membranes with a rejection higher than 90% in all cases. Methanol permeates through the polyamide and the LS-MIL-101(Cr) monolayer, greatly enhanced by the MOF pore system, in comparison with thin film composite (TFC) and conventional TFN membranes (7.5 ±0.7 L·m-2·h-1·bar-1 and 7.7 ±1.1 L·m-2·h-1·bar-1 when filtering sunset yellow), respectively, in which polyamide areas free of MOF NPs are present.

ACS Paragon Plus Environment

2

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

ACS Applied Materials & Interfaces

LIST OF ACRONYMS MOF: Metal-organic framework NPs: Nanoparticles MIL: Materials of Institute Lavoisier TFC: Thin film composite TFN: Thin film nanocomposite LB: Langmuir-Blodgett LS: Langmuir-Schaefer OSN: Organic solvent nanofiltration DMF: Dimethylformamide DMSO: Dimethylsulfoxide IPA: Isopropyl alcohol PEG: Polyethylene glycol PXRD: Powder X-ray diffraction TEM: Transmission electron microscopy SEM: Scanning electron microscopy FIB: Cryo-focused ion beam STEM-HAADF: Scanning transmission electron microscopy - high angle annular dark field EDS: Energy-dispersive X-ray spectroscopy EELS: Electron energy loss spectroscopy XPS: X-ray photoelectron spectroscopy AFM: Atomic force microscopy Ra: Average plane roughness Rms: Root mean square

ACS Paragon Plus Environment

3

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

Page 4 of 35

1. INTRODUCTION In the chemical and pharmaceutical industries and in the manufacturing of food, textiles and paper, the nanofiltration membrane system is gaining great importance as a flexible, energyefficient and low-cost separation technology for aqueous and organic feeds.1,

2

In particular,

organic solvent nanofiltration (OSN) focuses on the separation, recovery and disposal of organic solvents. This membrane technology demands materials easily available and capable of retaining multivalent ions and low molecular weight organics in the wide range of 200-1000 g·mol-1.3-6 Acid dyes are present in the waste from several of the above-mentioned industries and have molecular sizes in this range. Their removal from water and organic solvents is a must as they can cause severe damage to the environment and human health due to their toxic and carcinogenic properties.7,

8

In several nanofiltration studies these contaminants have not only

been removed from water9-13 and from more aggressive solvents such as methanol14-17 and toluene,18 but they have also been used for characterizing the molecular weight cut-off of membranes. Polymeric membranes display valuable tuneable properties, namely mechanical stability, structural diversity, relatively low fabrication costs and ease of industrial scale-up.3-6 With regard to organic solvents, solvent resistance should be guaranteed and thus the chemical stability of the polymer is crucial. In order to be applicable to a broad range of solvents, only intrinsically stable and cross-linked polymers should be considered in OSN.19 Thin film composite (TFC) membranes (Table 1) consist of an asymmetric type of polyimide membrane obtained by phase inversion that serves as support. The selective layer is a polyamide thin film grown via interfacial polymerization6, 20 at the top of the previous sub-layer. The cross-

ACS Paragon Plus Environment

4

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

ACS Applied Materials & Interfaces

linked polyamide active layer has been prepared using m-phenylenediamine and trimesoyl chloride as monomers.21, 22 Gorgojo et al.22 described the structure of the polyamide layer as an ensemble of two layers: a loose and negatively charged one on the outside and a cross-linked one in contact with the polyimide support. TFC membranes have been commonly applied in several membrane separation processes. As a matter of fact, in 2001 the Norwegian Statkraft company employed TFC membranes in its large scale osmotic power plant. For their part, Dow® and Koch Membrane Systems companies developed TFC membranes for aqueous separations.23,

24

Furthermore, TFC membranes can be easily employed in OSN since they consist of two polymeric layers chemically stable in a wide range of organic solvents. In fact, their commercialization for nanofiltration of organic solvents has been conducted by several companies, namely Koch Membrane Systems, Evonik, SolSep BV and GE Osmonics.24 Thin film nanocomposite (TFN) membranes (Table 1) were first designed by Jeong et al.25 by embedding molecular sieve nanoparticles inside the polyamide layer, broadly following the mixed matrix membranes (MMMs) approach. Metal-organic frameworks (MOFs), built from organic linkers and metal-containing clusters, were first included in this manner as fillers in TFN membranes by Sorribas et al.,17 showing better affinity with polymer chains than other types of porous nanoparticles. Moreover, Sorribas et al used TFN membranes for OSN application for the first time, and reported higher methanol permeation than with TFC membranes, without sacrificing rejection. A great challenge to be overcome in TFN membrane preparation is the tendency of MOFs to agglomerate and localize heterogeneously inside the polyamide layer.26 To illustrate this phenomenon, Table 1 shows the possible pathways of the permeate flux in the conventional TFN membrane, in which fillers are dispersed heterogeneously.

ACS Paragon Plus Environment

5

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

Page 6 of 35

Membranes and their schema (not to scale) TFC

LS-P84

LS-TFN

Conventional TFN

Table 1. Membrane types synthesized in this work. Numbers indicate the components of each membrane: (1) indicates the cross-linked asymmetric polyimide (P84®) support, (2) show the LS-MIL-101(Cr) monolayer, and (3) and (4) indicate the polyamide selective layer and the polyamide layer with MIL-10(Cr) NPs inside, respectively. Arrows show the possible flux pathways.

As previously reported,17,

26

MOF nanoparticles (NPs) facilitate and enhance methanol

permeance due to their pore system. In conventional TFN membranes, MOF NPs are heterogeneously embedded inside the polyamide layer and the permeate flows either through the combination of the polyamide and MOF pores or only through the polyamide layer, since polyamide areas free of MOF nanoparticles are more likely to be present. On the other hand, the presence of MOF agglomerates can be a drawback, contributing to the deterioration of filler dispersion and access to MOF pores and provoking a levelling-off the solvent flux.17 Additionally, it can facilitate the formation of unselective defects between the MOF NPs and the thin polyamide layer, reducing the degree of cross-linking of the polyamide11 and consequently decreasing selectivity.13 Moreover, Morris27 suggested that the control of the orientation of a

ACS Paragon Plus Environment

6

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

ACS Applied Materials & Interfaces

porous filler and therefore its pore system ensures high quality films, while Bètard and Fischer28 compiled many strategies to produce thin MOF layers suitable for countless applications. For their part, Ha et al.29 emphasized the importance of the organization of uniformly oriented zeolite monolayers in order to be widely applied in industry, for molecular separations and size-selective chemical sensors.30 Accordingly, recent experiments have been done to avoid MOF NPs agglomeration and to control their growth process, especially their size, in order to design thin MOF membranes. For instance, by means of the interfacial synthesis method, continuous ZIF-8 layers12, 31 of 250 - 300 nm in thickness were obtained after only one synthesis cycle at room temperature. Moreover, a multilayer structure in nanocomposite membranes was produced via solvothermal synthesis by the in situ growth of ZIF-8, using a layer-by-layer procedure.11 Another determining factor in the performance of OSN membranes is the surface roughness and morphology, owing to the fact that this directly relates to the enlargement of the effective surface area that may well result in flux increase32 and eventually in fouling.33,

34

The fouling phenomenon constitutes an important

limiting parameter in the performance of nanofiltration membranes and its reduction affects operation costs and energy and chemical consumption. Designing a hydrophilic top selective layer is a determining factor to reduce fouling in the OSN membrane by preventing chemical interactions with foulants.25,

35

Since improving hydrophilicity in the membrane surface is

desirable, dimethylformamide (DMF) post-treatments22 and the incorporation of negatively charged-hydrophilic fillers such as MOF MIL-101(Cr) NPs14,

16, 17

have been implemented to

enhance OSN membrane performance. Other strategies, like the bottom-up Langmuir-Blodgett (LB) methodology (vertical deposition), have recently been applied to improve the performance of polysulfone and

ACS Paragon Plus Environment

7

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

Page 8 of 35

polyacrylonitrile ultrafiltration membranes with thin monomolecular coatings of poly-4vinylpyridine (PVPyr) or latex particles.36 Previous studies to obtain smooth, uniform, ordered and high crystallinity polymer thin films have been conducted to enhance charge transport properties in electronic devices, by using LB or Langmuir-Schaefer (LS) (horizontal deposition) methodologies.37-39 Nowadays, LB technology focuses on the industrial-scale production of defect-free films on a square-meter substrates.40 Commercial troughs and two alternative film deposition routes are being developed for that purpose: automatic Roll-to-Roll LB processing for large device areas, and the LS methodology, since it allows to fabricate resistant, uniform and high-quality monolayers. In regard with this, we report the fabrication of a hydrophilic MOF monolayer on cross-linked asymmetric polyimide (P84®) supports before the interfacial polymerization stage, by the LS methodology. Although we have previously demonstrated that LB-MIL-101(Cr) films deposited onto different substrates (glass, quartz and QCM crystals) were highly compact and homogeneous,41 in this case we found that vertical deposition was not suitable when the substrate has a porous and hydrophilic nature, which is the case of P84®, and horizontal deposition (LS methodology) was needed. As a result, a LS-MIL-101(Cr) film has been incorporated for the first time in TFN membranes for nanofiltration and in particular for OSN application, giving rise to the LS-TFN membrane (Table 1). Through the LS approach, we have been able to position the MOF in a controlled manner forming a monolayer free of agglomerates. In these membranes, a thin polyamide layer is well formed on top of the MOF monolayer and the methanol continuously permeates through the combination of the polyamide layer and the MOF pores. Thus, higher permeance is expected in LS-TFN than in conventional TFN membranes.

ACS Paragon Plus Environment

8

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

ACS Applied Materials & Interfaces

Thus, organic solvent permeance has been efficiently enhanced by the hydrophilic porous structure of the MOF MIL-101(Cr)42 (MIL; Materials of Institute Lavoisier), built up from trimers of Cr octahedral, whose large pores and cavities are 1.2 - 2.9 nm and 2.9 - 3.4 nm in size, respectively.43 On the other hand, negligible loadings of MOF nanoparticles have been used to prepare the LS-MIL-101(Cr) film. In particular, less than 82% of MOF (MIL-101(Cr)) was required to fabricate LS-TFN membranes in comparison with the TFN membranes of Van Goethem et al.,13 who claimed to have considerably lessened (80 times reduction) the amount of MOF filler (ZIF-8) in comparison to other conventional MOF-TFN membranes.14, 16, 17, 44 As a consequence, this may result in the reduction of the costs associated with membrane preparation. A comparative study of the OSN performance of the three types of membranes (TFC, LS-TFN and conventional TFN) has been conducted, with an exceptional increase in permeance for LSTFN membranes, in which the nanosized MIL-101(Cr) fillers were organized in a monolayer. 4. MATERIALS AND METHODS 4.1. MIL-101(Cr) synthesis The synthesis of MIL-101(Cr) nanoparticles of 60 ± 20 nm in size was done following a procedure previously reported.17 Briefly, 0.7 g of chromium chloride hexahydrate (CrCl6·H2O >98%, Sigma Aldrich) was dissolved in 26 mL of deionized water. Then 0.45 g of terephthalic acid (C6H4-1,4-(CO2H)2 – 98%, Sigma Aldrich) was added. The resulting solution was poured into a PTFE-TFM pressure vessel. The reaction was performed in a microwave oven (Multiwave 3000, Anton Paar) at 453 K for 30 min. The green product was recovered by centrifugation at 10 000 rpm and washed several times with deionized water. To remove unreacted terephthalic acid, the sample was purified with dimethylformamide (DMF – 99.5%, Scharlab) at 393 K overnight.

ACS Paragon Plus Environment

9

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

Page 10 of 35

Then DMF was exchanged with methanol (HPLC grade, Scharlab) by refluxing for 12 h. Finally, the MOF was washed with methanol and stored in the solvent to avoid NP agglomeration.

4.2. OSN support preparation In the preparation of the dope solution, dimethylsulfoxide (DMSO – synthesis grade, Scharlab) was used as the solvent because its solvent properties are similar to those of DMF, as reported by Solomon et al.20 Polyimide Lenzing P84® (HP polymer GmbH) was dissolved in DMSO to form a 24 wt% polymer dope solution. The solution was allowed to stand until the complete removal of air bubbles and then cast onto a polypropylene non-woven 40 cm x 40 cm sheet support (Freudenberg) at room temperature, using an adjustable casting knife set at a thickness of 250 µm. To design an asymmetric membrane, the resultant polymer sheet was immersed in a distilled water coagulation bath, set at 23 °C, where the phase inversion took place. After 10 min it was transferred to a fresh water bath and left for about 1 h. To remove the remaining water and DMSO, it was immersed into two successive 1 h solvent exchange baths of isopropyl alcohol (IPA - 99.5%, Scharlab). Then the cross-linking process took place to obtain a membrane resistant to organic solvents, and the cast asymmetric polyimide P84® supports were treated with a solution of 120 g·L-1 of hexanediamine (98%, Sigma Aldrich) in IPA for 16 h. Afterwards, the supports were washed with 4 cycles of 1 h of IPA bath and then conditioned in a polyethylene glycol (PEG - synthesis grade, Scharlab) bath, with a 3:2 volume ratio PEG/IPA solution, to avoid pore collapse. Finally, the supports were dried with tissue paper and properly stored. 4.3. Langmuir (L) and Langmuir-Schaefer (LS) fabrication

ACS Paragon Plus Environment

10

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

ACS Applied Materials & Interfaces

L films were assembled in a Teflon-made Langmuir trough KSV-NIMA, model KN200, whose dimensions are 580 mm × 145 mm, equipped with a double-barrier compression system. This trough was kept inside a closed cabinet placed in a clean room at 293 K (±1 K). In all the experiments, ultra-pure Milli-Q water (ρ = 18.2 MΩ·cm) was used as a subphase. For the general procedure for preparing Langmuir-Blodgett (LB) films on cross-linked asymmetric polyimide (P84®) supports, these supports were initially immersed into the water subphase, but the transfer ratio was very low due to the poor adhesion of MOF NPs onto the wet surface of the porous polymer. This difficulty has been overcome using horizontal deposition (LS technique) by means of a vacuum pump-based horizontal dipping clamp. The suction cup where the substrates were attached has a diameter of 18 mm. In this method, the horizontal substrate descends in a controlled manner and interacts directly with the film floating on the aqueous subphase. Before starting, MIL-101(Cr) NPs were suspended (the dispersion concentration was 0.2 mg·mL-1) in a solvent mixture of chloroform (>99.9%, Panreac) and methanol (>99.9%, SigmaAldrich) in a 4:1 volume ratio. To reduce NPs agglomeration in the film, a 1 wt% of behenic acid (CH3(CH2)20COOH – 99%, Sigma Aldrich) was added, and its total amount was calculated using the following equation: % Behenic Acid =

Behenic Acid mass · 100 MOF mass + Behenic Acid mass

To ensure a complete mixing of all the components, the suspensions were stirred for 24 h and sonicated during 5 min before using. To obtain homogeneous L films, 2 mL of MIL101(Cr):Behenic Acid suspension was spread onto the water surface. The solvent was allowed to evaporate for 15 min, and the compact film was formed by compressing the trough barriers at a constant speed of 6 cm2·min−1 until a surface pressure of 12 mN·m-1 was achieved. Only one LS

ACS Paragon Plus Environment

11

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

Page 12 of 35

film was transferred by direct contact of the cross-linked asymmetric polyimide (P84®) substrate (60.8 cm2) at a dipping speed of 1 cm·min-1, forming the hereinafter called LS-P84 (Table 1).

4.4. Synthesis of the thin polyamide layer A thin polyamide layer was fabricated on a cross-linked asymmetric polyimide P84® support, giving rise to the so-called TFC membranes, or on top of a LS-P84 support resulting in LS-TFN membranes. Moreover, for the purposes of comparison, MIL-101(Cr) NPs were included in the polyamide matrix following the conventional methodology,17 producing TFN membranes. In all cases, the polyamide layer was hand-cast by interfacial polymerization using mphenylenediamine (99%, Sigma Aldrich) as a monomer with amine groups in the aqueous phase and trimesoyl chloride (98%, Sigma Aldrich) as a monomer with acyl chloride groups in the organic phase. For the fabrication of the three types of membranes (TFC, LS-TFN and conventional TFN), the cross-linked asymmetric polyimide P84® OSN support was placed in a glass filtration holder. 30 mL of m-phenylenediamine / distilled water solution (2% (w/v)) was added and soaked into the support for 2 min. The excess solution was disposed of and the membrane was wiped with tissue paper. 30 mL of the organic solution made of 0.1% (w/v) of trimesoyl chloride in hexane (extra pure, Scharlab) was poured into the membrane and left reacting for 1 min, when 10 mL of hexane was added to stop the reaction. After removing the excess, an extra 10 mL of hexane was added to remove unreacted trimesoyl chloride. Finally, the hexane was washed out with 10 mL of distilled water. For the preparation of TFN membranes, 0.2% (w/v) of MIL-101(Cr) NPs were dispersed in the organic phase by sonication and stirring,

ACS Paragon Plus Environment

12

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

ACS Applied Materials & Interfaces

following the procedure reported in Sorribas et al.17 Then, the same process as for synthesizing the TFC membranes was followed. Two post-treatments were applied sequentially to improve the performance of the TFC and LS-TFN membranes.14,

16,

17

Immediately after the interfacial polymerization reaction,

membranes were immersed in a DMF bath for 10 min and then stored in water at 4 °C until tested for OSN. After 30 min of OSN (see below), these membranes were submitted to DMF filtration for 10 min at a pressure of 20 bar, using the same membrane module as for the OSN applications. 4.5. Characterization of MOFs and membranes Powder X-ray diffraction (PXRD) patterns of MIL-101(Cr) NPs were recorded in a D-Max 2500 Rigaku diffractometer with a Cu Kα (λ = 1.542 Å) rotating anode, at 40 kV and 80 mA. A LS film was directly deposited onto a carbon mesh grid to observe the size, morphology and distribution of the MOF crystals that form the LS film by transmission electron microscopy (TEM). Images of the LS-MIL-101(Cr) monolayer were taken using a FEI Tecnai T20, operated at 200 kV. Surface and cross-section morphology images of the LS film and the LS-TFN membrane were taken using a scanning electron microscope (SEM) FEI Inspect F20, with 10 and 20 kV of acceleration voltage. Samples were coated with a layer of Pt (10 nm). The SEM images were digitally treated to calculate the real coverage of the cross-linked asymmetric polyimide (P84®) support with the LS-MIL-101(Cr) monolayer. To verify the presence of the MOF monolayer in between the two polymers, a lamella (~ 80 nm thick) was obtained from the upper 2 µm of the LS-TFN membrane using cryo-focused ion beam (FIB) equipment (Dual Beam 3 Nova 200) and Ga atoms for the etching. Scanningtransmission electron microscopy imaging with a high angle annular dark field detector (STEM-

ACS Paragon Plus Environment

13

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

Page 14 of 35

HAADF) was applied to illustrate the polyamide/MOF/polyimide system interface and the thickness of the polyamide nanocomposite layer was also measured by averaging four measurements from different areas of the membrane. A cryo holder station in a FEI Tecnai F30 microscope operated at 300 kV was used for this purpose. Furthermore, energy-dispersive X-ray spectroscopy (EDS) and electron energy loss spectroscopy (EELS) were used to confirm the presence of MOFs (Cr). Chemical evidence of the LS-MIL-101(Cr) monolayer in between the polymer films (the top 300 nm at most) was assessed using X-ray photoelectron spectroscopy (XPS, AXIS ultra DLD (Kratos Analytical) analysis system), using monochromated Al Kα (1486.6 eV) excitation radiation at 12 kV and 10 mA. Ar+ bombardment was performed before the measurement and then the concentration of Cr, C and N were profiled, etching the membrane from the top (beam energy of 3 keV, 5 mA) and taking data every hour. Peaks were fitted using the software CasaXPS. The membrane surface roughness of a cross-linked asymmetric polyimide P84® support and TFC, conventional TFN and LS-TFN membranes was measured using AFM (atomic force microscopy), by means of a Veeco MultiMode 8 scanning probe microscope, in tapping mode under ambient conditions. AFM experiments were carried out at a resonant frequency of 300 kHz with a force constant of 40 mN, using a silicon cantilever provided by Bruker. Images were recorded with a scan rate of 1 Hz, and an amplitude set-point lower than 1 V. After the measurement, the average plane roughness (Ra), the root mean square (Rms) and the relative surface area were obtained. 4.6. OSN experiments Nanofiltration experiments were carried out in a dead-end membrane module (Sterlitech HP4750), at 20 bar and at room temperature, under constant stirring. The membranes were cut

ACS Paragon Plus Environment

14

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

ACS Applied Materials & Interfaces

into disks with an effective area of 12 cm2 and located at the bottom of the module. Membrane performance was evaluated by feeding a solution (20 mg·L-1) of a dye in methanol. In the OSN process, the permeate flux, in units of L·m-2·h-1·bar-1, and dye rejection efficiency (%) can be expressed by the equations already reported in Echaide-Gorriz et al.14 In a typical OSN analysis, the feeding solution was constantly pushed through the membrane for 30 min and then the membrane was submitted to DMF filtration as a post-treatment process for 10 min using the same 20 bar of pressure. Afterwards, the OSN process was restarted for 30 min more and then 3 mL of permeate flux and another 3 mL of feeding solution were collected for evaluating the rejection. Methanol from both solutions was allowed to evaporate and then 3 mL of distilled water was added as a solvent. A UV spectrophotometer (Jasco V-670 spectrophotometer) was used to measure the concentrations of the permeate and feeding. This apparatus detects the light absorbed at the maximum wavelength, which is 546 nm for rose bengal (Sigma-Aldrich, 95% dye content) and 480 nm for sunset yellow (Sigma-Aldrich, 90% dye content), both dyes being used as solutes in the feeding solution. Absorbance and concentration were related by a calibration curve. 3. RESULTS AND DISCUSSION 3. 1. Characterization of LS-TFN membranes The membranes fabricated in this work are listed and represented in Table 1. The MOF MIL101(Cr) was used as the filler for the TFN membranes and was synthesized by microwave irradiation, following a previous publication.41 Pure nano-MIL-101(Cr) crystals (see XRD pattern in Figure S1) were produced with 60 ± 20 nm in size (Figure 1a). LS-MIL-101(Cr) film transferred onto a carbon mesh grid was used to check the layout of the crystals by means of

ACS Paragon Plus Environment

15

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

Page 16 of 35

TEM, confirming the formation of a monolayer. It should be noted that some crystal faces are in contact, as Figure 1a shows. However, they are beam sensitive and apart from suffering loss of crystallinity, the crystals tend to segregate during TEM observation. The longer the time they were under the beam, the more the original LS layout was broken apart (Figure 1a). Overall, a negligible amount of MIL-101(Cr) NPs was required to achieve their successful transfer onto the support. The SEM images in Figures 1b and S2 illustrate how the LS film of MIL-101(Cr) nearly covers the whole surface of the cross-linked asymmetric polyimide (P84®) support and mimics its superficial morphology. It is important to note that the area of the support in Figure S2a is 0.47 mm2, and it is completely covered with the LS-MOF monolayer. Figures 1b and S2b have been used to calculate the coated surface area of the cross-linked asymmetric polyimide (P84®) with the LS-MIL-101(Cr) film, and it has proved to be 70.5 ± 1%.

Figure 1. a) TEM image of a LS-MIL-101(Cr) film transferred onto a carbon mesh TEM grid; b) SEM image of the surface of a LS - MIL-101(Cr) film over a cross-linked asymmetric polyimide (P84®) support (LS-P84); c) SEM image of the surface of a LS-TFN membrane.

ACS Paragon Plus Environment

16

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

ACS Applied Materials & Interfaces

Although LS films on polyimide are slightly less dense than the LB films previously obtained in other substrates,41,

45

they are compact enough for our purposes and the continuity and

integrity of the LS-MIL-101(Cr) monolayer have been demonstrated using Figures 1b, S2b and 2a. Regarding to the Benito et al.41 surface pressure-area isotherm (Figure S3), the theoretical density of MIL-101(Cr) NPs at the transference surface pressure of 12 nN·m-1 is 3.8 µg·cm-2. This isotherm is consistent with the formation of a compact film, owing to a considerable fast increase of the surface pressure up to 40 mN·m-1. 3.8 µg·cm-2 is the maximum amount of material that is added to the support by the LS methodology. In fact, the improvement in the reduction of the MOF used in comparison with conventional membranes14, 16, 17, 44 is remarkable. When a MOF-TFN membrane is prepared by the conventional procedure, a 0.2% (w/v) 30 mL solution of dispersed MOF in hexane is used for a support with identical unit area, but only small amounts of MOF NPs are effectively incorporated as fillers in the TFN membrane. Van Goethem et al.13 followed the evaporation-controlled filler positioning (EFP) procedure to design TFN membranes, requiring amounts of MOF 80 times smaller than those required by conventional routes and without any MOF loss. Taking into account that Van Goethem’s optimal MOF loading for achieving the best membrane performance was 0.005% (w/v) in 12.5 mL of hexane for a 30 cm2 support and no MOF loss occurred, we have calculated that in this case 20.8 µg·cm-2 of ZIF-8 NPs were incorporated in the TFN membranes. This indicates that the LS-TFN membranes require about 82% less MOF per support unit area compared to the EFP-TFN membranes.

ACS Paragon Plus Environment

17

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

Page 18 of 35

After the LS-MOF transference, a polyamide thin film layer is formed on top of the LS-P84 support and the irregular morphology of the polymer layer, characterized by ridges and valleys, veils the existence of the LS-MIL-101(Cr) monolayer underneath (Figure 1c). It is also important to note that the total amount of MOF incorporated in the membrane resisted attached to the support during the interfacial polymerization process to prepare the polyamide film. This is due to the fact that no NPs were lost towards the water and organic solutions. An achievement of this kind could be implemented in other composite membranes where compact and homogeneous films that totally cover the support surface are mandatory. In addition, this will help save expenses associated to MOF synthesis since the small excess of NPs remains in the LS bath and could be easily reused. This would be more difficult if the NPs were combined with the interfacial polymerization reactants, as typically happens when preparing conventional TFN membranes. SEM or EDX techniques on the surface of the membrane were not powerful enough to image or detect the presence of MOF crystals in the membrane. Therefore, in order to reveal the good formation of the MOF-monolayer and its resistant location even after the deposit of the polyamide layer, a composition profile was analyzed by XPS. The idea was to profile the sample from the top, where the polyamide is, and then to descend until reaching the LS-MIL-101(Cr) monolayer. LS-P84 sample was used as a reference for the atomic percentages of Cr present in the LS-MIL-101(Cr) film and the C/N ratio of the polyimide. Then, starting from the top of the membrane, Ar+ bombardment was performed in the LS-TFN membrane and an average composition was calculated from three measurements taken at the same depth level along the depth profile. The atomic percentages of Cr, C and N are shown for each sample at different etching cycles that correspond to different profile depths. Ideally a decline in the Cr signal would

ACS Paragon Plus Environment

18

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

ACS Applied Materials & Interfaces

take place while reaching the cross-linked asymmetric polyimide (P84®) support, and the C/N ratio is expected to be higher for polyimide than for polyamide. The XPS data shown in Table 2 confirm this previously formulated hypothesis. Table 2. Atomic compositions of C, N and Cr of a LS-P84 and a LS-TFN systems analyzed via XPS, after a given number of bombardment cycles with Ar+. System

N° etching cycles

%C

%N

% Cr

95.28

4.50

0.21

0

86.90

13.02

0.08

20

89.10

10.68

0.21

50

95.17

4.70

0.13

LS-P84 LS-TFN

Before starting the etching process, the contribution of Cr to the atomic percentage of the three elements could be detected but was the smallest, indicating that the thickness of the polyamide layer is particularly low. After 20 Ar+ etching cycles, a slight increase in the C/N ratio was observed. In addition, the Cr percentage rose to a value close to that of the LS-P84 system (0.21%). The nearly complete removal of the polyamide layer was achieved after 50 cycles, as the C/N ratio showed a sharper increase. Additionally, the Cr percentage started decreasing (0.13%), meaning that the XPS data was then taken closed to the bottom of the LS-MOF film. Figures 2a and 2b show the STEM images of a LS-TFN lamella obtained by FIB, illustrating the continuity of the LS thin monolayer in between the “sandwich” of polymers: the cross-linked asymmetric polyimide (P84®) support at the bottom and the polyamide thin layer at the top. The preparation of the lamella required the deposition of two Pt coatings that have also been imaged, represented by letters d and e (electron and ion-beam-deposited-layer, respectively). In addition, it is important to note that prior to FIB Pt deposition, the LS-TFN membrane was coated with Pt

ACS Paragon Plus Environment

19

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

Page 20 of 35

to make the sample conductive (Figure 2b), in the same way as for other microscopy techniques. This Pt coating stands out (whiter in the image) from the other Pt depositions and the crosslinked asymmetric polyimide (P84®) support. As a consequence of the FIB Pt depositions, the edges of the MIL-101(Cr) NPs were impregnated with Pt, highlighting their localization and distribution and the fact that the LS-MOF monolayer was almost free of MOF agglomerates that could hinder the membrane separation performance. In fact, the Pt contribution in the EDS line spectrum of 2 (inset in Figure 2c), which corresponds to the edge of a MIL-101(Cr) nanoparticle, contrasts with the absence of Cr whose contribution has only been detected in the inside of a MIL-101(Cr) NP (EELS spectrum 3, Figure 3c). Although Pt is still detected in the inside of a MIL-101(Cr) NP by EDS, its peak is less intense, while the Cr contribution is clearly represented (Figure S4a). The EELS line scan of 1, which corresponds to the cross-linked asymmetric polyimide (P84®) support (Figure 2c), and the EDS spectrum (Figure S4b) evinces neither the presence of Pt nor that of Cr but only confirms the presence of C, as expected.

ACS Paragon Plus Environment

20

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

ACS Applied Materials & Interfaces

Figure 2. a) and b) STEM images of the LS-TFN lamella that illustrate the LS film of MIL101(Cr) NPs in between the polymer system (polyimide at the bottom, polyamide at the top): a) the elements that constitute the FIB lamella imaged by STEM are indicated as a, b, c, d and e; b) magnified area from a) (red square). 1, 2 and 3 indicate the areas where EELS and EDS spectra (represented in c)) have been recorded. Moreover, the thickness of the LS-MIL-101(Cr) monolayer was measured using Figures 2a, 2b and S4, being 60-110 nm that is in good agreement with our filler particle size. The LS-MIL101(Cr) monolayer in combination with the thin polyamide layer was also measured, being 130275 nm, depending on the irregular morphology of the polyamide layer. The polyamide nanocomposite layer thickness is in good agreement with the findings reported by Lind et al.46

ACS Paragon Plus Environment

21

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

Page 22 of 35

(100-300 nm) and Jeong et al.25 (50-200 nm) for their conventional zeolite-TFN membranes, and by Van Goethem et al.13 (100-200 nm) for their EFP-TFN membranes. AFM images of LS-P84, TFC, LS-TFN and conventional TFN membranes are given in Figures 3 and 4, and the corresponding Ra and Rms values are presented in Table 3. In particular, the Rms parameter is more sensitive to big peaks and valleys. The surface topography of the bare cross-linked asymmetric polyimide (P84®) support (Figures 3a and 3b) inspected by AFM was used as a reference to characterize the changes in roughness and morphology in comparison with the rest of the composite membranes, especially focusing on the samples containing the MOF. All the membranes were analyzed before any DMF post-treatment; therefore, they will show rougher surfaces than after having been tested in the OSN experiments. Compared to the bare cross-linked asymmetric polyimide (P84®) support (Ra = ±2.1) (Figure 3a, b), the characteristic ridge and valley polyamide surface in TFC membranes increases the membrane roughness (Ra = ±24.8) (Figure 4a), and so does the LS-MIL-101(Cr) film procedure in LS-P84 membranes (Ra = ±50.6) (Figure 3c, d). Interestingly, when adding the thin polyamide layer to form the LS-TFN membrane, the Ra value decreases considerably down to ±47.0 and its surface becomes smoother (Figure 4b). On the other hand, the increase in roughness when comparing the TFC with the two TFN membranes is enough to provoke an enhancement in the membrane permeance, as it will be showed in due course, but it is not too marked so as to harm rejection and produce a defective membrane.35

ACS Paragon Plus Environment

22

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

ACS Applied Materials & Interfaces

Figure 3. a) 3D and b) 2D AFM images of the bare cross-linked asymmetric polyimide (P84®) support; c) 3D and d) 2D AFM images of the LS film of MIL-101(Cr) NPs on the cross-linked asymmetric polyimide (P84®) support (LS-P84).

Upon comparing Ra and Rms values of LS-TFN and conventional TFN membranes (Table 3), the LS-TFN membrane has the smoothest and homogeneous surface. AFM images of a conventional TFN membrane (Figure 4c) show a random distribution of MIL-101(Cr) NPs in the polyamide layer in comparison with LS-TFN membranes, due to the absence of a well-formed MOF monolayer that it has been clearly formed in the LS-TFN membrane.

ACS Paragon Plus Environment

23

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

Page 24 of 35

Figure 4. AFM images of the surface of a) TFC membrane; b) LS-TFN membrane and c) conventional TFN membrane. Table 3. Mean surface roughness (Ra) and root mean square (Rms) values of the cross-linked asymmetric polyimide (P84®) support, LS-P84, TFC, LS-TFN and conventional TFN membranes. Each value was calculated from 3 images taken from 4 µm2 of different substrates. cross-linked asymmetric polyimide (P84®) support

LS-P84

TFC

LS-TFN

Conventional TFN

Ra (nm)

2.1 ± 0.0

50.6 ± 1.2

24.8 ± 1.7

47.0 ± 0.2

52.7 ± 0.6

Rms (nm)

2.7 ± 0.1

62.8 ± 0.1

30.7 ± 1.2

57.5 ± 1.7

64.7 ± 0.2

3.2. OSN results Figure 5 shows the performance of TFC and LS-TFN membranes in terms of the methanol/dye system (Figure 5a) and rejection (Figure 5b) when using sunset yellow (in orange) or rose bengal (in pink) as solutes. In addition, two conventional TFN membranes were tested for OSN. Membrane separation was reproducible with relatively small errors (obtained by averaging the performance of two different membranes) in the permeance and rejection values. Regardless of the solute used, it should be noted that permeance was markedly higher for the LS-TFN than for

ACS Paragon Plus Environment

24

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

ACS Applied Materials & Interfaces

the TFC membranes. Nevertheless, there are several differences related with the type of solute and the type of membrane (a membrane with fillers organized in a monolayer (LS-TFN), with inhomogeneously distributed fillers (conventional TFN) or without fillers (TFC)). Focusing exclusively on sunset yellow, the methanol permeance for the TFC, LS-TFN and conventional TFN membranes is surprisingly high (from 7.5 ±0.7 to 10.1 ±0.5 and 7.7 ±1.1 L·m-2·h-1·bar-1, respectively) with steady and high sunset yellow rejections greater than 90% for all membranes (from 94.0 ±6.4 to 91.1 ±0.9 and 97.8 ±0.8, respectively). The overall OSN results reported here are among the highest for methanol permeance to date, only improved by Peyravi et al.15 using functionalized TiO2 nanoparticles and obtaining a methanol permeance of 25.2 L·m-2·h-1·bar-1 and a rejection of 93% for crystal violet (409 Da, in the range of sunset yellow). Our results show a marked improvement even compared with the TFC and TFN membranes in which MOF NPs are used as fillers, like the MIL-101(Cr)-conventional TFN membranes designed by Sorribas et al.,17 measured in a cross-flow unit for the filtration of a methanol/styrene oligomers suspension, and the Echaide-Gorriz et al.14,

16

membranes, whose OSN performance was

measured in a dead-end module, filtrating methanol with sunset yellow as the solute. Conversely, in these two cases, DMF was used to prepare the dope solution for the polyimide membrane casting and as the activating solvent instead of using DMSO. The methanol permeance for TFC and TFN membranes was 1.5 ±0.1 and 3.9 ±0.3 L·m-2·h-1·bar-1, respectively, for Sorribas et al.,17 and 3.3 ±0.9 to 3.9 ±1.0 L·m-2·h-1·bar-1, respectively, for Echaide-Gorriz et al.14,

16

Styrene

oligomers and sunset yellow rejection were higher than 90% in all cases.

ACS Paragon Plus Environment

25

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

Page 26 of 35

Figure 5. OSN performance at 20 bar and 20 °C: a) permeance of methanol for TFC and LSTFN membranes and b) rejection of solutes for LS-TFN and TFC membranes. Orange polygons correspond to sunset yellow (SY, 450 Da) and pink ones correspond to rose bengal (RB, 1017 Da).

Regarding the rose bengal OSN experiments, the rejection values were very satisfactory, being > 96% for all membranes, and the methanol permeance increased significantly in the LSTFN membranes compared to the TFC membranes, from 6.0 ±0.7 to 9.5 ±2.1 L·m-2·h-1·bar-1, respectively. Finally, it is noteworthy that regardless of the solute used, the LS-TFN membranes showed an exceptional membrane performance, with a minor loss in the already high rejection of the two solutes used: smaller sunset yellow (450 Da) and larger rose bengal (1017 Da). The higher roughness in LS-TFN and conventional TFN membranes compared to TFC membranes could facilitate the methanol permeance due to an increase in the effective membrane surface area, without altering the solute rejection. Moreover, the large pores and cavities of the MOF MIL-101(Cr) NPs used as fillers, their hydrophilic nature, and the LS-MIL-101(Cr) layout

ACS Paragon Plus Environment

26

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

ACS Applied Materials & Interfaces

forming a monolayer with no MOF agglomerates, contribute to their remarkable methanol permeance. Additionally, the absence of obstacles (other MOF NPs) that can hinder the methanol permeance through the MOF pore system as they are organized in a thin monolayer, and the controlled localization of the MOF NPs over practically the total surface of the support, make the LS-TFN membranes more effective and suitable for OSN performance than conventional TFN membranes.

4. CONCLUSIONS MIL-101(Cr) NPs have been assembled into monolayers using the Langmuir-Schaefer (LS) technique over cross-linked asymmetric polyimide (P84®) supports. Innovative LS-TFN membranes have been designed for OSN application following a procedure in which there is no filler loss and the amount of MOF used is the smallest reported to date. Furthermore, LS film fabrication promotes a controlled positioning of the MOF monolayer ensuring a continuous and homogeneous coating of the support surface, with the LS-MIL-101(Cr) NPs in between the polymer system (polyamide/LS-MOF/polyimide). Via XPS analysis, a profile-composition of the MOF was carried out and STEM imaging of a cross–section lamella obtained by FIB was performed to confirm the presence and layout of the LS-MOF film. Defect-free LS-TFN membranes with a homogeneous distribution of the MOF fillers forming a monolayer have been achieved with an excellent nanofiltration performance of methanol/dye solutions due to the large MIL-101(Cr) pore system, its hydrophilic nature and the LS layout free of MOF agglomerates, therefore facilitating an effective methanol transport and a high dye rejection.

ACS Paragon Plus Environment

27

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

Page 28 of 35

ASSOCIATED CONTENT Supporting Information. Characterization of MIL-101(Cr) NPs (XRD), additional SEM images of the surface of a LS-MIL-101(Cr) film over a cross-linked asymmetric polyimide (P84®) support (LS-P84), surface pressure vs Area per mg of MOF isotherm and characterization of the LS-TFN membrane (STEM images and EDS analysis). ACKNOWLEDGMENT Financial support from the Spanish Ministry of Economy and Competitiveness (MINECO) and FEDER (MAT2016-77290-R), the European Social Fund (ESF) and the Aragón Government (DGA, T05) is gratefully acknowledged. L.P. also thanks the MINECO for a Ph.D. grant. The Laboratorio de Microscopías Avanzadas at the Instituto de Nanociencia de Aragón (LMA-INA, Universidad de Zaragoza) is thanked for providing the microscopy equipment and expertise. Moreover, the authors would like to acknowledge the use of the Servicio General de Apoyo a la Investigación-SAI (Universidad de Zaragoza). REFERENCES (1) Adler, S.; Beaver, E.; Bryan, P.; Robinson, S.; Watson, J. Vision 2020: 2000 Separations Roadmap, Center for Waste Reduction Technologies of the AIChE and Dept. of Energy of the United States of America 2000. (2)

Marchetti, P.; Solomon, M. F. J.; Szekely, G.; Livingston, A. G. Molecular Separation

with Organic Solvent Nanofiltration: A Critical Review. Chem. Rev. 2014, 114, 10735-10806.

ACS Paragon Plus Environment

28

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

ACS Applied Materials & Interfaces

(3) Han, S.; Wong, H. T.; Livingston, A. G. Application of Organic Solvent Nanofiltration to Separation of Ionic Liquids and Products from Ionic Liquid Mediated Reactions. Chem. Eng. Res. Des. 2005, 83, 309-316. (4) Wong, H. T. C.; Pink, J.; Ferreira, F. C.; Livingston, A. G. Recovery and Reuse of Ionic Liquids and Palladium Catalyst for Suzuki Reactions Using Organic Solvent Nanofiltration. Green Chem. 2006, 8, 373-379. (5) Wood, J.; Gifford, J.; Arba, J.; Shaw, M. Production of Ultrapure Water by Continuous Electrodeionization. Desalination. 2010, 250, 973-976. (6) Vandezande, P.; Gevers, L. E. M.; Vankelecom, I. F. J.; Solvent Resistant Nanofiltration: Separating on a Molecular Level. Chem. Soc. Rev. 2008, 37, 365-405. (7) Ge, P.; Wang, Q.; Wan, C.; Chung, T.-S. Polyelectrolyte-Promoted Forward Osmosis– Membrane Distillation (FO–MD) Hybrid Process for Dye Wastewater Treatment. Environ. Sci. Technol. 2012, 46, 6236-6243. (8) Minear, R. A.; Amy, G. L. Water disinfection and natural organic matter, American Chemical Society, 1996. (9) Xu, Y. C.; Wang, Z. X.; Cheng, X. Q.; Xiao, Y. C.; Shao, L. Positively Charged Nanofiltration Membranes Via Economically Mussel-Substance-Simulated co-Deposition for Textile Wastewater Treatment. Chem. Eng. J. 2016, 303, 555-564. (10) An, A. K.; Guo, J. X.; Jeong, S.; Lee, E. J.; Tabatabai, S. A. A.; Leiknes, T. High Flux and Antifouling Properties of Negatively Charged Membrane for Dyeing Wastewater Treatment by Membrane Distillation. Water Res. 2016, 103, 362-371.

ACS Paragon Plus Environment

29

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

Page 30 of 35

(11) Wang, L. Y.; Fang, M. Q.; Liu, J. He, J.; Li, J. D.; Lei, J. D.; Layer-by-Layer Fabrication of High-Performance Polyamide/ZIF-8 Nanocomposite Membrane for Nanofiltration Applications. ACS Appl. Mater. Interfaces. 2015, 7, 24082-24093. (12) Li, Y.; Wee, L. H.; Martens, J. A.; Vankelecom, I. F. Interfacial Synthesis of ZIF-8 Membranes with Improved Nanofiltration Performance. J. Membr. Sci. 2017, 523, 561-566. (13) Van Goethem, C.; Verbeke, R.; Hermans, S.; Bernstein, R.; Vankelecom, I. Controlled Positioning of MOFs in Interfacially Polymerized Thin-Film Nanocomposites. J. Mater. Chem. A. 2016, 4, 16368-16376. (14) Echaide-Gorriz, C.; Sorribas, S.; Tellez, C; . Coronas, J. MOF Nanoparticles of MIL-68 (Al), MIL-101 (Cr) and ZIF-11 for Thin Film Nanocomposite Organic Solvent Nanofiltration Membranes. RSC Adv. 2016, 6, 90417-90426. (15) Peyravi, M.; Jahanshahi, M.; Rahimpour, A.; Javadi, A.; Hajavi, S. Novel Thin Film Nanocomposite Membranes Incorporated with Functionalized TiO2 Nanoparticles for Organic Solvent Nanofiltration. Chem. Eng. J. 2014, 241, 155-166. (16) Echaide-Gorriz, C.; Navarro, M.; Tellez, C.; Coronas, J. Simultaneous Use of MOFs MIL-101 (Cr) and ZIF-11 in Thin Film Nanocomposite Membranes for Organic Solvent Nanofiltration. Dalton Trans. 2017, 46, 6244-6252. (17) Sorribas, S.; Gorgojo, P.; Tellez, C.; Coronas, J.; Livingston, A. G. High Flux Thin Film Nanocomposite Membranes Based on Metal–Organic Frameworks for Organic Solvent Nanofiltration. J. Am. Chem. Soc. 2013, 135, 15201-15208.

ACS Paragon Plus Environment

30

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

ACS Applied Materials & Interfaces

(18) Darvishmanesh, S.; Degreve, J.; Van der Bruggen, B. Performance of Solvent-Pretreated Polyimide Nanofiltration Membranes for Separation of Dissolved Dyes from Toluene. Ind. Eng. Chem. Res. 2010, 49, 9330-9338. (19) Koros, W. J.; Mahajan, R. Pushing the Limits on Possibilities for Large Scale Gas Separation: Which Strategies?. J. Membr. Sci. 2000, 175, 181-196. (20) Soroko, I.; Lopes, M. P.; Livingston, A. The Effect of Membrane Formation Parameters on Performance of Polyimide Membranes for Organic Solvent Nanofiltration (OSN): Part A. Effect of Polymer/Solvent/Non-Solvent System Choice. J. Membr. Sci. 2011, 381, 152-162. (21) Jimenez-Solomon, M. F.; Bhole, Y.; Livingston, A. G. High Flux Hydrophobic Membranes for Organic Solvent Nanofiltration (OSN)—Interfacial Polymerization, Surface Modification and Solvent Activation. J. Membr. Sci. 2013, 434, 193-203. (22) Gorgojo, P.; Jimenez-Solomon, M.; Livingston, A. G. Polyamide Thin Film Composite Membranes on Cross-Linked Polyimide Supports: Improvement of RO Performance via Activating Solvent. Desalination. 2014, 344, 181-188. (23) Buonomenna, M. G. Membrane processes for a sustainable industrial growth. RSC Adv. 2013, 3, 5694-5740. (24) Buonomenna, M. G.; Bae, J. Organic Solvent Nanofiltration in Pharmaceutical Industry. Sep. Purif. Rev. 2015, 44, 157-182. (25) Jeong, B.-H.; Hoek, E. M. V.; Yan, Y.; Subramani, A.; Huang, X.; Hurwitz, G.; Ghosh, A. K.; Jawor, A. Interfacial polymerization of Thin Film Composites: a New Concept for Reverse Osmosis Membranes. J. Membr. Sci. 2007, 294, 1-7.

ACS Paragon Plus Environment

31

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

Page 32 of 35

(26) Lau, W. J.; Gray, S.; Matsuura, T.; Emadzadeh, D.; Chen, J. P.; Ismail, A. F. A Review on Polyamide Thin Film Nanocomposite (TFN) Membranes: History, Applications, Challenges and Approaches. Water Res. 2015, 80, 306-324. (27) Morris, R. E. How does your MOF grow? ChemPhysChem, 2009, 10, 327-329. (28) Bétard, A.; Fischer, R. A. Metal–Organic Framework Thin Films: From Fundamentals to Applications. Chem. Rev. 2012, 112, 1055-1083. (29) Ha, K.; Lee, Y. J.; Jung, D. Y.; Lee, J. H.; Yoon, K. B. Micropatterning of Oriented Zeolite Monolayers on Glass by Covalent Linkage. Adv. Mater. 2000, 12, 1614-1617. (30) Urbiztondo, M.; Pellejero, I.; Rodriguez, A.; Pina, M. P.; Santamaria, J. Zeolite-coated interdigital capacitors for humidity sensing. Sens. Actuators, B. 2011, 157, 450-459. (31) Li, Y.; Wee, L. H.; Volodin, A.; Martens, J. A.; Vankelecom, I. F. Polymer Supported ZIF-8 Membranes Prepared via an Interfacial Synthesis Method. Chem. Comm. 2015, 51, 918920. (32) Hirose, M.; Ito, H;. Kamiyama, Y. Effect of Skin Layer Surface Structures on the Flux Behaviour of RO Membranes. J. Membr. Sci. 1996, 121, 209-215. (33) Vrijenhoek, E. M.; Hong, S.; Elimelech, M. J. Membr. Sci. 2001, 188, 115-128. (34) Tijing, L. D.; Woo, Y. C.; Choi, J.-S.; Lee, S.; Kim, S.-H.; Shon, H. K. J. Membr. Sci. 2015, 475, 215-244. (35) Jhaveri, J. H.; Murthy, Z. Desalination. 2016, 379, 137-154.

ACS Paragon Plus Environment

32

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

ACS Applied Materials & Interfaces

(36) Melnikova, G.; Zhavnerko, G.; Chizhik, S.; Bildyukevich, A. Structure and Mechanical Properties of Ultrafiltration Membranes Modified with Langmuir–Blodgett Films. Pet. Chem. 2016, 56, 406-412. (37) Pandey, R. K.; Upadhyay, C.; Prakash, R. Pressure dependent surface morphology and Raman studies of semicrystalline poly(indole-5-carboxylic acid) by the Langmuir-Blodgett technique. RSC Adv. 2013, 3, 15712-15718. (38) Pandey, R. K.; Singh, A. K.; Upadhyay, C.; Prakash, R. Molecular self ordering and charge transport in layer by layer deposited poly (3,3‴-dialkylquarterthiophene) films formed by Langmuir-Schaefer technique. J. Appl. Phys. 2014, 116, 094311-094318. (39) Pandey, R. K.; Yadav, S. K.; Upadhyay, C.; Prakash, R.; Mishra, H. Surface plasmon coupled metal enhanced spectral and charge transport properties of poly(3,3 '''dialkylquarterthiophene) Langmuir Schaefer films. Nanoscale. 2015, 7, 6083-6092. (40) Zhavnerko, G.; Marletta, G. Developing Langmuir-Blodgett strategies towards practical devices. Mater. Sci. Eng., B. 2010, 169, 43-48. (41) Benito, J.; Fenero, M;. Sorribas, S.; Zornoza, B.; Msayib, K. J.; McKeown, N. B.; Tellez, C.; Coronas, J.; Gascon, I. Fabrication of Ultrathin Films Containing the Metal Organic Framework Fe-MIL-88B-NH 2 by the Langmuir–Blodgett Technique. Colloids Surf., A. 2015, 470, 161-170. (42) Férey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S.; Margiolaki, I. A Chromium Terephthalate-Based Solid with Unusually Large Pore Volumes and Surface Area. Science. 2005, 309, 2040-2042.

ACS Paragon Plus Environment

33

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

Page 34 of 35

(43) Khan, N. A.; Kang, I. J.; Seok H. Y.; Jhung, S. H. Facile Synthesis of Nano-Sized MetalOrganic Frameworks, Chromium-Benzenedicarboxylate, MIL-101. Chem. Eng. J. 2011, 166, 1152-1157. (44) Duan, J.; Pan, Y.; Pacheco, F.; Litwiller, E.; Lai, Z.; Pinnau, I. High-Performance Polyamide Thin-Film-Nanocomposite Reverse Osmosis Membranes Containing Hydrophobic Zeolitic Imidazolate Framework-8. J. Membr. Sci. 2015, 476, 303-310. (45) Benito, J.; Sorribas, S.; Lucas, I.; Coronas, J.; Gascon, I. Langmuir–Blodgett Films of the Metal–Organic Framework MIL-101 (Cr): Preparation, Characterization, and CO2 Adsorption Study Using a QCM-Based Setup. ACS Appl. Mater. Interfaces. 2016, 8, 16486-16492. (46) Lind, M. L.; Ghosh, A. K.; Jawor, A.; Huang, X. F.; Hou, W.; Yang, Y.; Hoek, E. M. V. Influence of Zeolite Crystal Size on Zeolite-Polyamide Thin Film Nanocomposite Membranes. Langmuir. 2009, 25, 10139-10145.

ACS Paragon Plus Environment

34

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

ACS Applied Materials & Interfaces

Table of Contents Graphic

ACS Paragon Plus Environment

35