Summary

Synthesis and Characterization of Self-Assembled Metal-Organic Framework Monolayers Using Polymer-Coated Particles

Published: June 14, 2024
doi:

Summary

A protocol for the synthesis and characterization of self-assembled metal-organic framework monolayers is provided using polymer-grafted, metal-organic framework (MOF) crystals. The procedure shows that polymer-grafted MOF particles can be self-assembled at an air-water interface resulting in well-formed, free-standing, monolayer structures as evidenced by scanning electron microscopy imaging.

Abstract

Metal-organic frameworks (MOFs) are materials with potential applications in fields such as gas adsorption and separation, catalysis, and biomedicine. Attempts to enhance the utility of MOFs have involved the preparation of various composites, including polymer-grafted MOFs. By directly grafting polymers to the external surface of MOFs, issues of incompatibility between polymers and MOFs can be overcome. Polymer brushes grafted from the surface of MOFs can serve to stabilize the MOF while enabling particle assembly into self-assembled metal-organic framework monolayers (SAMMs) via polymer-polymer interactions.

Control over the chemical composition and molecular weight of the grafted polymer can allow for tuning of the SAMM characteristics. In this work, instructions are provided on how to immobilize a chain transfer agent (CTA) onto the surface of the MOF UiO-66 (UiO = Universitetet i Oslo). The CTA serves as initiation sites for the growth of polymers. Once polymer chains are grown from the MOF surface, the formation of SAMMs is achieved through self-assembly at an air-water interface. The resulting SAMMs are characterized and shown to be freestanding by scanning electron microscopy imaging. The methods presented in this paper are expected to make the preparation of SAMMs more accessible to the research community and thereby expand their potential use as a MOF-polymer composite.

Introduction

Metal-organic frameworks (MOFs) are crystalline, porous materials that offer large surface areas while being readily tunable through modifications of the organic ligands or metal nodes1,2. MOFs are constructed from two components: an organic ligand and metal ions (or metal ion clusters referred to as secondary building units, SBUs). MOFs have been investigated for chemical (e.g., gas) storage, separations, catalysis, sensing, and drug delivery. Generally, MOFs are synthesized in the form of crystalline powders; however, for ease of handling in many applications, formulation into other form factors is desirable if not necessary3,4. For example, mixed matrix membranes (MMMs) of MOFs with polymers have been reported as one particularly useful composite of MOFs and polymers5. However, in some cases, MMMs may have limitations due to the incompatibility/immiscibility between MOF and polymer components5,6. Therefore, strategies have been explored to incorporate polymer grafting directly onto MOF particles to form polymer-grafted MOFs.

Inorganic and metallic nanoparticles exhibit unique behavior in terms of optical, magnetic, catalytic, and mechanical properties7,8. However, they tend to aggregate easily after synthesis, which can hinder their processability. To enhance their processability, polymer chains can be grafted onto the particle surface9. Nanoparticles with high grafting density offer excellent dispersion and stability due to favorable enthalpic interactions between surface polymers and the solvent and entropic repulsion interactions between the particles10. Grafting of polymers onto particle surfaces can be achieved through a variety of strategies11. The most straightforward approach is the 'grafting to' particle strategy, where functional groups, such as thiols or carboxylic acids, are introduced at the ends of polymer chains to directly bind to the nanoparticle. When complementary chemical groups, such as hydroxyls or epoxides, are present on the particle surface, polymer chains can be grafted onto these groups via covalent chemical approaches12,13. The 'grafting from' particle or surface-initiated polymerization method involves anchoring initiators or chain transfer agents (CTAs) to the surface of nanoparticles and then growing polymer chains on the particle surface through surface-initiated polymerization. This method often achieves higher grafting density than the 'grafting to' approach. Furthermore, grafting from enables the synthesis of block copolymers, thereby expanding the diversity of polymer structures that can be immobilized on a particle surface.

Examples of polymer grafting onto MOF particles have begun to emerge, largely focused on installing polymerization sites on the organic ligands of the MOF. In a recent study published by Shojaei and coworkers, vinyl groups were covalently attached to the ligands of Zr(IV)-based MOF UiO-66-NH2 (UiO = Universitetet i Oslo, where the terephthalic acid ligand contains an amino substituent), followed by methyl methacrylate (MMA) polymerization to create polymer-grafted MOFs with a high grafting density (Figure 1A)14. Similarly, Matzger and coworkers functionalized the amine groups on a core-shell MOF-5 (a.k.a., IRMOF-3@MOF-5) particles with 2-bromo-iso-butyl groups. Using polymerization initiated by the 2-bromo-iso-butyl groups, they created poly(methyl methacrylate) (PMMA)-grafted PMMA@IRMOF-3@MOF-515.

In addition to functionalizing the ligand of the MOF for grafting from polymerization, new methods that create sites for polymer grafting via coordination to the metal centers (a.k.a., SBUs) of the MOF have also been explored. For example, a ligand that can bind to the MOF metal centers, such as catechol (Figure 1B), can be used to coordinate to exposed metal sites on the MOF surface. Using a catechol-functionalized chain-transfer agent (cat-CTA, Figure 1B) the MOF surface can be functionalized and made suitable for a grafting from polymerization.

Recently, the aforementioned strategy for synthesizing MOFs-polymer composites has also been used for the creation of free-standing MOF monolayers16,17,18. MOFs such as UiO-66 and MIL-88B-NH2 (MIL = Materials of Institute Lavoisier) were surface-functionalized with pMMA using a ligand-CTA strategy (Figure 1B). The polymer-grafted MOF particles were self-assembled at an air-water interface to form self-supporting, self-assembled metal-organic framework monolayers (SAMMs) with a thickness of ~250 nm. The polymer content in these composites was ~20 wt%, indicating that SAMMs contained ~80 wt% MOF loading. Followup studies showed that different vinyl polymers could be grafted onto UiO-66 to produce SAMMs with different characteristics19. Analytical techniques such as thermogravimetric analysis (TGA), dynamic light scattering (DLS), and gel permeation chromatography (GPC) were used to calculate polymer brush height and grafting density of the surface-grafted MOF-polymer composites.

Herein, the preparation of SAMMs from UiO-66-pMA (pMA = poly(methyl acrylate)) is presented. For the polymerization of methyl acrylate (MA), 2-(dodecylthiocarbonothioylthio)-2-methylpropionic acid (DDMAT, Figure 1B) is used as the CTA19. The functionalization of the UiO-66 particles with cat-DDMAT is essential for the grafting of pMA. Cat-DDMAT can be synthesized through a two-step acylation procedure from a commercially available CTA and dopamine hydrochloride19. It is also crucial to use UiO-66 particles of uniform size for the successful formation of SAMMs19; therefore, the UiO-66 used in this study was prepared using the continuous addition method20. The polymerization method employed for forming the polymer-grafted MOF particles is photoinduced reversible addition-fragmentation chain transfer (RAFT) conducted under blue LED light (using an in-house-built photoreactor, Figure 2) with a tris(2-phenylpyridine)iridium (Ir(ppy)3) photocatalyst. RAFT polymerization gives exceptionally narrow polymer dispersity that can be finely controlled. Free CTA is included during the polymerization reaction because the ratio of transfer agent to monomer allows for control over the molecular weight during polymerization. The amount of cat-DDMAT transfer agent on the surface of the MOF particles is small; therefore, excess free CTA is added and the amount of monomer to be used is calculated based off of the amount of free CTA present21. After polymerization, the free polymer produced from the free CTA is removed by washing, leaving only the polymer-grafted UiO-66-pMA. Subsequently, this composite is dispersed in toluene at a high concentration and used to form SAMMs at an air-water interface.

Protocol

1. Surface modification of UiO-66 with cat-DDMAT Exchange the solvent of UiO-66 from methanol with water. Prepare UiO-66 in methanol at a concentration of 20 mg/mL. NOTE: According to Wang et al.20, homogeneous UiO-66 is washed with DMF and methanol after synthesis and then stored in a dispersed state in methanol. Transfer the 10 mL of UiO-66 suspension to a 15 mL conical centrifuge tube using a pipette. Perform centrifugation …

Representative Results

When the polymer-grafted MOFs are gently dropped onto water from a concentrated toluene dispersion (as illustrated in Figure 4A), a monolayer is formed in a few seconds with an iridescent appearance. Furthermore, using a mold made from copper wire to lift this monolayer and subsequently drying the obtained water allows for the formation of free-standing SAMMs (Figure 4B). After transferring the monolayer to a glass microscope cover slip and drying it, SEM imagin…

Discussion

There are several critical steps where specific attention to detail is required to successfully synthesize polymer-grafted MOFs that will produce SAMMs. First, the monomers utilized in RAFT polymerization are supplemented with inhibitors or stabilizers during storage to prevent undesired polymerization (e.g., hydroquinone or monomethyl ether of hydroquinone, MEHQ). To remove these additives, purification through distillation is required before use22. In protocol step 2.4, it is essential to dilute…

Divulgaciones

The authors have nothing to disclose.

Acknowledgements

M.K. was supported by a grant from the National Science Foundation, Division of Chemistry under Award No. CHE-2153240. Additional support for materials and supplies was provided by the Department of Energy, Office of Basic Energy Sciences, Division of Materials Science and Engineering under award no. DE-FG02-08ER46519. SEM imaging was performed in part at the San Diego Nano-Technology Infrastructure (SDNI) of U.C. San Diego, a member of the National Nanotechnology Coordinated Infrastructure, which is supported by the National Science Foundation (ECCS-1542148).

Materials

2-(dodecylthiocarbonothioylthio)-2-methylpropionic acid (DDMAT) Sigma-Aldrich 723010 98%
10 mL Single Neck RBF Chemglass CG-1506-82 14/20 Outer Joint
Acetone Fisher Chemical A18-20 ACS Grade
Allegra X-30R Centrifuge BECKMAN COULTER B06320 1.6 L max capacity, 18,000 RPM, 29,756 x g
Analog Vortex Mixer VWR 10153-838 300 – 3,200 rpm
cat-DDMAT Prepared according to literature procedure (ref. 17).
Centrifuge Tube, 50 mL / 15 mL CORNING 430291 / 430766 Conical Bottom with plug seal cap, polypropylene
Chloroform Fisher Chemical AC423550040 99.8%
Conventional needles Becton Dickinson 382903051670 21 G x 1 1/2
Copper wire Malin Co. No. 30 B & S GAUGE
Dimethyl Sulfoxide (DMSO) Fisher Bioreagents BP231-1 >=99.7%
Disposable Pasteur Pipets Fisher Scientific 13-678-20C Borosilicate Glass
Ethanol KOPTEC V1001 200 proof ethanol
Glass Scintillation Vial, 20 mL KIMBIL 74508-20
Graduated Cylinder, 10 mL KIMBIL 20024-10
Hypodermic Needles Air-Tite N224 22 G x 4''
Methanol Fisher Chemical A412-20 99.8%
Methyl Acrylate Aldrich Chemistry M27301 99%, contains =< 100 ppm monomethyl ether hydroquinone as inhibitor
Micropipette P10 (1 – 10 µL) GILSON F144055M PIPETMAN, Metal Ejector
Micropipette P1000 (100 – 1,000 µL) GILSON F144059M PIPETMAN, Metal Ejector
Micropipette P20 (2 – 20 µL) GILSON F144056M PIPETMAN, Metal Ejector
Microscope cover glass Fisher Scientific 12542A 18 mm x 18 mm
NN-Dimerhylformamide (DMF) Fisher Chemical D119-4 99.8%
Petri Dish, Stackable Lid Fisher Scientific FB0875713A 60 mm x 15 mm
Septum Stopper Chemglass CG302401 14/20 – 14/35
Stir Bar Chemglass CG-2005T-01 Magnetic, PTFE, Turbo, Rare Earth, Elliptical, 10 x 6mm
SuperNuova+ Stirring Hot Plate Thermo Scientific SP88857190 50 – 1,500 rpm, 30 – 450 °C
Toluene Fisher Chemical T324-4 99.5%
Tris[2-phenylpyridinato-C2,N]iridium(III) (Ir(ppy)3) Sigma-Aldrich 688096 97%
UiO-66 (120 nm edge length) Prepared according to literature procedure (ref. 18).
Ultrasonic Cleaner CPX3800H EMERSON / BRANSON CPX-952-318R 40 kHz, 5.7 L
Waterproof Flexible LED Strip Light ALITOVE ALT-5B300WPBK 16.4 ft 5050 Blue LED

Referencias

  1. Eddaoudi, M., et al. Systematic design of pore size and functionality in isoreticular mofs and their application in methane storage. Science. 295, 469-472 (2002).
  2. Yaghi, O. M., et al. Reticular synthesis and the design of new materials. Nature. 423, 705-714 (2003).
  3. Kitao, T., Zhang, Y., Kitagawa, S., Wang, B., Uemura, T. Hybridization of mofs and polymers. Chem Soc Rev. 46 (11), 3108-3133 (2017).
  4. Kalaj, M., et al. Mof-polymer hybrid materials: From simple composites to tailored architectures. Chem Rev. 120 (16), 8267-8302 (2020).
  5. Lin, R., Villacorta Hernandez, B., Ge, L., Zhu, Z. Metal organic framework based mixed matrix membranes: An overview on filler/polymer interfaces. J Mater Chem A. 6 (2), 293-312 (2018).
  6. Semino, R., Moreton, J. C., Ramsahye, N. A., Cohen, S. M., Maurin, G. Understanding the origins of metal-organic framework/polymer compatibility. Chem Sci. 9 (2), 315-324 (2018).
  7. Daniel, M. -. C., Astruc, D. Gold nanoparticles: Assembly, supramolecular chemistry, quantum-size-related properties, and applications toward biology, catalysis, and nanotechnology. Chem Rev. 104, 293-346 (2004).
  8. Zhou, J., Yang, Y., Zhang, C. Y. Toward biocompatible semiconductor quantum dots: From biosynthesis and bioconjugation to biomedical application. Chem Rev. 115 (21), 11669-11717 (2015).
  9. Chancellor, A. J., Seymour, B. T., Zhao, B. Characterizing polymer-grafted nanoparticles: From basic defining parameters to behavior in solvents and self-assembled structures. Anal Chem. 91 (10), 6391-6402 (2019).
  10. Wright, R. A., Wang, K., Qu, J., Zhao, B. Oil-soluble polymer brush grafted nanoparticles as effective lubricant additives for friction and wear reduction. Angew Chem Int Ed. 55 (30), 8656-8660 (2016).
  11. Pastore, V. J., Cook, T. R. Coordination-driven self-assembly in polymer-inorganic hybrid materials. Chem Mater. 32 (9), 3680-3700 (2020).
  12. Chiu, J. J., Kim, B. J., Kramer, E. J., Pine, D. J. Control of nanoparticle location in block copolymers. J Am Chem Soc. 127, 5036-5037 (2005).
  13. Zubarev, E. R., Xu, J., Sayyad, A., Gibson, J. D. Amphiphilic gold nanoparticles with v-shaped arms. J Am Chem Soc. 128 (15), 4958-4959 (2006).
  14. Molavi, H., Shojaei, A., Mousavi, S. A. Improving mixed-matrix membrane performance via pmma grafting from functionalized nh2-uio-66. J Mater Chem. A. 6 (6), 2775-2791 (2018).
  15. Mcdonald, K. A., Feldblyum, J. I., Koh, K., Wong-Foy, A. G., Matzger, A. J. Polymer@mof@mof: "Grafting from" atom transfer radical polymerization for the synthesis of hybrid porous solids. Chem Commun. 51 (60), 11994-11996 (2015).
  16. Barcus, K., Cohen, S. M. Free-standing metal-organic framework (mof) monolayers by self-assembly of polymer-grafted nanoparticles. Chem Sci. 11 (32), 8433-8437 (2020).
  17. Xiao, J., et al. Photoswitchable nanoporous metal-organic framework monolayer film for light-gated ion nanochannel. ACS Appl Nano Mater. 6 (4), 2813-2821 (2023).
  18. Xiao, J., et al. Self-assembled nanoporous metal-organic framework monolayer film for osmotic energy harvesting. Adv Funct Mater. 34 (2), 2307996 (2024).
  19. Barcus, K., Lin, P. A., Zhou, Y., Arya, G., Cohen, S. M. Influence of polymer characteristics on the self-assembly of polymer-grafted metal-organic framework particles. ACS Nano. 16 (11), 18168-18177 (2022).
  20. Wang, X. G., Cheng, Q., Yu, Y., Zhang, X. Z. Controlled nucleation and controlled growth for size predicable synthesis of nanoscale metal-organic frameworks (mofs): A general and scalable approach. Angew Chem Int Ed. 57 (26), 7836-7840 (2018).
  21. Moad, C. L., Mood, G. Fundamentals of reversible addition-fragmentation chain transfer (raft). Chem Teach Int. 3 (2), 3-17 (2021).
  22. Van Keulen, H., Mulder, T. H. M., Goedhart, M. J., Verdonk, A. H. Teaching and learning distillation in chemistry laboratory courses. J Res Sci Teach. 32 (7), 715-734 (2006).
  23. Pérez, L. D., Giraldo, L. F., Brostow, W., López, B. L. Poly(methyl acrylate) plus mesoporous silica nanohybrids: Mechanical and thermophysical properties. e-poly. 7 (1), 29 (2007).
This article has been published
Video Coming Soon
Keep me updated:

.

Citar este artículo
Kang, M., Cohen, S. M. Synthesis and Characterization of Self-Assembled Metal-Organic Framework Monolayers Using Polymer-Coated Particles. J. Vis. Exp. (208), e66497, doi:10.3791/66497 (2024).

View Video