Direct Surface Growth Of UIO-66-NH2 on Polyacrylonitrile Nanofibers

Nov 30, 2017 - MOFabric: Electrospun Nanofiber Mats from PVDF/UiO-66-NH2 for Chemical Protection and Decontamination. ACS Applied Materials & ...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/IECR

Cite This: Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Direct Surface Growth Of UIO-66-NH2 on Polyacrylonitrile Nanofibers for Efficient Toxic Chemical Removal Annie X. Lu,*,†,‡ Ann M. Ploskonka,§ Trenton M. Tovar,‡ Gregory W. Peterson,‡ and Jared B. DeCoste‡ †

Defense Threat Reduction Agency, 2800 Bush River Road, Aberdeen Proving Ground, Maryland 21010, United States Edgewood Chemical Biological Center, 5183 Blackhawk Road, Aberdeen Proving Ground, Maryland 21010, United States § Leidos, Incorporated, P.O. Box 68, Aberdeen Proving Ground, Maryland 21010, United States ‡

S Supporting Information *

ABSTRACT: Direct solvothermal growth of the metal− organic framework (MOF) UiO-66-NH2 on polymer surface was successfully demonstrated. By using acetone as the solvent for synthesis instead of N,N-dimethylformamide, polymers like polyacrylonitrile (PAN) can be used directly in the solvothermal synthesis step to grow MOF on the polymer surface. We use X-ray diffraction and FT-IR to confirm our method produces crystalline UiO-66-NH2 on the surface of electrospun PAN nanofibers. Characterization of this type of composite revealed up to 50 wt % MOF loading according to nitrogen isotherms. Since the MOFs are located on the surface of the polymer fibers, the composites are capable of high loadings of chlorine gas. Compared to electrospun composites made with preformed UiO-66-NH2, the in situ method is a simple alternative that produces composites with higher MOF loading.



INTRODUCTION Metal−organic frameworks (MOFs) are porous crystalline materials composed of metal or metal oxide nodes, known as the secondary building unit (SBU), connected by polydentate organic ligands. Due to their high surface area and tailorability, MOFs have become a leading class of porous materials for applications in gas storage,1 separations,2,3 catalysis,4−6 filtration,7−9 and sensing.10 Until recently, most of the MOF research has focused on the development of new materials and the fundamental use of these materials toward aforementioned applications. An area that has been gaining more interest is the engineering of MOFs into polymers films11−13 (e.g., mixed matrix membranes) and fibers.14−16 The new applications for MOFs in filtration, protective clothing, and wearable sensors all require the attachment of MOF particles to a permeable substrate, namely, polymer fibers. The vast majority of materials in this area focuses on attaching MOF particles on fibers through covalent attachment17,18 or electrospinning polymers containing MOFs in the bulk solution.9,19,20 More recent efforts have focused on in situ techniques, where MOF particles are directly grown on the surface of the polymer21 rather than processing a polymer mixture containing preformed MOF powders into fibers. In situ techniques are attractive “one-pot” approaches in that the polymer substrate and the MOF precursors, such as the linker and the metal salts, can be directly added together in the © XXXX American Chemical Society

solvent. Additionally, since MOFs are located on the surface of the polymer rather than within, the in situ technique can avoid common MOF−polymer compatibility issues, such as poor accessibility of MOFs due to polymer blockage22 or introducing void spaces between the polymer and the MOF, causing poor mechanical properties of the composite material.23 The in situ technique has already been proven successful for several MOF types, for example, HKUST- 1,24 ZIFs,16,25,26 and bio-MOFs.27 One MOF type that is of interest to be grown via the in situ technique is the UiO series. UiO MOFs are of particular interest due to their chemical and mechanical stability stemming from the Zr6O4(OH)4 SBU.28,29 Several MOFs in this series, such as UiO-66, UiO-66-NH2, and UiO-67, have also proven to have exceptional toxic chemical removal capabilities.7,8,30−32 However, direct growth of these MOFs on polymers has been limited, as the MOFs are solvothermally synthesized using the solvent N,N-dimethylformamide (DMF), which would solubilize many polymer materials in the MOF synthesis step, and therefore are not amenable to the in situ technique. Thus, far, Zhao et al. successfully grew UiO MOFs on polyamide-6 nanofibers coated with metal oxides via atomic Received: October 10, 2017 Revised: November 16, 2017 Accepted: November 17, 2017

A

DOI: 10.1021/acs.iecr.7b04202 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

Figure 1. Schematic of in situ growth of UiO-66-NH2 on plain polyacrylonitrile fibers and polyacrylonitrile (PAN) fibers seeded with the aminoterephthalic acid (ATA linker). Electrospun swatches are sandwiched between two cellulose filter papers before being placed in a parr bomb with the UiO-66-NH2 precursors: equimolar concentrations of ATA linker and zirconium chloride (metal), in acetone. Increasing concentrations of precursors are used for synthesis, 1, 2, and 3 mmol, to observe effects of concentration of precursors on UiO-66-NH2 synthesized on the fiber: (a) control PAN swatch; (b) PAN with ATA linker. (c) Control PAN postsynthesis (d) PAN with ATA linker postsynthesis. Scale bar 3 μm.

Table 1. BET Surface Area and Chlorine Uptake of Electrospun PAN, PAN with ATA Linker, Each Sample Undergoing 1, 2, and 3 mmol Synthesis Conditionsa control

PAN PAN-ATA preformed UiO-66-NH2 in PAN UiO-66-NH2 (control) a

PS 1 mmol

PS 2 mmol

PS 3 mmol

BET (m2/g)

chlorine (mol/kg sample)

BET (m2/g)

chlorine (mol/kg sample)

BET (m2/g)

chlorine (mol/kg sample)

BET (m2/g)

chlorine (mol/kg sample)

4 7 342

0.30 0.38 2.21

62 250

0.30 2.43

211 270

1.99 2.95

392 500

2.83 3.29

1015

6.6

BET surface area and chlorine uptake of PAN with preformed UiO-66-NH2 (40 wt %) and powder UiO-66-NH2 are listed for comparison.

layer deposition (ALD);15 the metal oxide forms a protective coating around the nanofiber, thus avoiding dissolution during solvothermal synthesis while also providing nucleation points for MOF growth. Recently, Lee et al. successfully assembled preformed UiO-66-NH2 onto the surface of untreated polypropylene fibers.33 Here, we present for the first time the direct solvothermal growth of UiO-66-NH2 on unmodified PAN fibers, utilizing acetone as the solvent for synthesis, a technique developed by Ploskonka et al.34 Since acetone is not a solvent for PAN, electrospun PAN fibers can be directly placed in the synthesis solution with the MOF precursors without being solubilized. We use a seeding strategy to optimize the in situ growth of UiO-66-NH2 on electrospun PAN nanofibers. The linker, amino-terephthalic acid (ATA), is first dissolved in the polymer solution before electrospinning. Once the polymer solution is electrospun, the resulting nanofibers contain the ATA linkers embedded within the polymer, which gives it a yellow appearance (Figure 1b). The embedded linker is to serve asan anchoring point or nucleation point for MOF growth. For comparison, when pristine PAN is electrospun, the result is a white nanofiber swatch (Figure 1a). After electrospinning, all samples are submerged in acetone with equimolar concentrations of zirconium chloride (ZrCl4) and ATA. Postsynthesis, the swatches are washed and sonicated 3 times for 30 min each to remove any residual MOFs or precursors trapped between the interstitial spaces of the nanofibers. Our study illustrates how some variables, such as the concentration of the MOF precursors, can be tuned to optimize the MOF loading on the nanofibers.

When compared to swatches made from electrospun PAN containing preformed UiO-66-NH2 (Figure S1), the swatches with embedded linkers offer better chlorine uptake at all synthesis conditions. Overall, our method offers a simple, scalable way to create MOF on fiber composites from unmodified polymer materials.



RESULTS AND DISCUSSION SEM images of the PAN and PAN-ATA electrospun swatches prior to in situ synthesis show smooth fiber surfaces, as shown in Figure 1a and 1b. Features of roughness can be seen in the SEM of electrospun PAN with preformed UiO-66-NH2 (Figure S1), as these samples have MOF crystals intercalated within the nanofibers. Previous work done in our lab showed that loadings of up to 40 wt % preformed MOF in polymer solution can be consistently spun into nanofibers, which is the weight loading used in this work to produce the preformed UiO-66-NH2 in PAN. For in situ synthesis conditions, 3 different concentrations of equimolar ZrCl4 and ATA linker were used for growing UiO-66-NH2 on PAN and PAN-ATA swatches. For brevity, samples postsynthesis (PS) will be indicated by the concentration of MOF precursors. For example, PAN PS 1 mmol is electrospun pristine PAN treated postsynthetically with 1 mmol of equimolar concentrations of ZrCl4 and ATA linker. Specific synthesis conditions are provided in the Supporting Information (SI). SEM images of PAN and PANATA swatches postsynthesis at 1 mmol can be found in Figure 1c and 1d, and SEM images of the same swatches postsynthesis at 2 and 3 mmol can be found in Figure S2. The coverage of the nanofibers by UiO-66-NH2 is not apparent at 1 mmol synthesis B

DOI: 10.1021/acs.iecr.7b04202 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

Figure 2. X-ray diffraction patterns of electrospun samples different growth conditions: (a) electrospun PAN PS 1, 2, and 3 mmol synthesis conditions, electrospun PAN with preformed UiO-66-NH2 (40 wt %) PS 1, 2, and 3 mmol synthesis conditions; (b) electrospun PAN with ATA linker PS 1, 2, and 3 mmol synthesis conditions, powder UiO-66-NH2.

conditions on PAN. However, it is still visible to see some particles attached to the nanofibers at this condition. As the concentration of the precursors is increased, more MOF particles are visibly attached to the pristine PAN nanofibers (Figure S2a and S2b). This observation is supported by BET measurements using nitrogen isotherms, shown in Table 1, where the surface area of the swatches increases from 62 to 392 m2/g at 1−3 mmol synthesis conditions. While we were not expecting MOF growth on pristine PAN nanofibers, we postulate that since the polymer swells at the temperature used for synthesis, it is likely that MOF precursors anchored to the softened polymer surface and initiated growth. SEM images of PAN-ATA postsynthesis at different concentrations show expected surface morphology, where there is MOF growth continuously covering the surface of the nanofibers even at 1 mmol concentrations. Increasing synthesis concentrations at 2 and 3 mmol resulted in more homogeneous coverage and larger MOF clusters, seen in Figure S2c and S2d. BET data supports the visual observation of increasing MOF growth, where surface area increases from 250 to 500 m2/g at 1 and 3 mmol concentrations, respectively. It is important to note that during synthesis the linker in the PAN-ATA samples leaches out into the synthesis solution, which can be visually observed when PAN-ATA fibers are incubated with acetone solvent alone (results not shown). The additional ATA from the polymer could lead to changing the ratio of concentration of the linker to metal salt in the solution; this could explain the morphology change of the MOF grown on PAN-ATA at 3 mmol concentrations, where a saturated solution of precursors causes larger MOF clusters to form. The adsorption isotherms for all electrospun swatches can be found in Figure S3. PXRD patterns of electrospun PAN and PAN-ATA before and after in situ synthesis are show in Figure 2. PXRD patterns

confirm that all PAN-ATA samples display short- and longrange order consistent with the crystal structure of UiO-66NH2 synthesized in acetone (Figure 2b). The PXRD patterns of UiO-66-NH2 are less pronounced for PAN samples, especially at 1 mmol synthesis conditions. This is consistent with SEM and nitrogen isotherm data, where the MOF grown at this condition is either of poor crystalline structure or the has a quantity that is low compared to the mass of the polymer. The presence of UiO-66-NH2 on each sample is further confirmed by ATR-FTIR (Figure S4), where all samples postsynthesis contain the same peaks found in the fingerprint region of UiO-66-NH2 synthesized in acetone. It has been reported that UiO-66-NH2 has the ability to remove chlorine gas from air via electrophilic aromatic substitution.7 Here, we use chlorine gas as a probe molecule not only to assess the reactivity and accessibility of the surfacegrown MOFs but also to explore potential applications of the MOFs synthesized via the in situ technique for applications in chemical protection. UiO-66-NH2 synthesized from acetone is capable of 6.6 mol of chlorine uptake per kilogram of MOF, whereas PAN nanofibers with preformed UiO-66-NH2 have one-third the capacity of powder UiO-66-NH2 (Table 1). However, with the in situ technique, chlorine loading of PANATA samples postsynthesis is 2.43, 2.95, and 3.29 mol of chlorine per kilogram of sample (MOF and fiber) at 1, 2, and 3 mmol synthesis conditions. This is an improvement when compared to chlorine loading of preformed UiO-66-NH2 in PAN at 2.21 mol per kilogram of sample. Chlorine loading for PAN samples is negligible at 1 mmol synthesis conditions, which is consistent with other characterization methods that showed minimal MOF formation on the nanofibers at this concentration. At 2 and 3 mmol synthesis conditions, chlorine loading increases to 1.99 and 2.83 mol per kilogram of sample, C

DOI: 10.1021/acs.iecr.7b04202 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

Figure 3. (a) Percent of MOF loading on PAN and PAN-ATA swatches postsynthesis at 1, 2, and 3 mmol conditions. Percentage is calculated by taking the ratio of the BET surface area value of the swatch to the BET surface area of powder UiO- 66-NH2. (b) Percent of activity of PAN and PAN-ATA swatches postsynthesis at 1, 2, and 3 mmol conditions. Percentage is calculated by taking the ratio of chlorine loading of the swatch to chlorine loading of powder UiO-66-NH2. For reference, the percent MOF loading and percent activity of electrospun PAN with preformed UiO-66NH2 is depicted by the gray dot connected to the horizontal dotted line.

used to attach UiO-66-NH2 to PAN-based polymers for applications in chemical protection and detoxification.

respectively. Compared to chlorine loading of PAN-ATA samples, this data suggests that having the embedded ATA linker in the polymer produces MOFs that have higher capacity toward chlorine. Electrospun PAN without UiO-66-NH2 has a negligible chlorine loading at 0.3 mol/kg, some of which is due to artificial loading due to instrumentation. Electrospun PAN with seeded ATA linker has a slightly higher chlorine loading at 0.38 mol/kg, which can be explained by the incorporated amines in the fiber. To better illustrate the performance of the MOFs grown on the surface via the in situ technique, a percent loading of the MOF on fiber after each synthesis procedure is calculated by taking the ratio of the BET surface area postsynthesis to the surface area of the powder UiO-66-NH2 synthesized in acetone. Similarly, a percent activity is calculated by taking the ratio of chlorine loading of the samples postsynthesis to the powder UiO-66-NH2. This data is displayed graphically in Figure 3. This calculation is done in lieu of TGA due to the ATA linker polymerizing at high temperatures,34 therefore making MOF loading calculations via TGA difficult for PAN with embedded ATA linker. The MOF loading and activity of preformed UiO66-NH2 in PAN is used as a baseline, shown by the dashed lines. In Figure 3a, it is evident that the calculated percent MOF loading for PAN and PAN-ATA based on the surface area is below the calculated loading for preformed UiO-66-NH2 in PAN at 1 and 2 mmol synthesis conditions but exceeds the baseline at 3 mmol concentrations, with up to 50% surface area of the UiO-66-NH2 powders for PAN-ATA. However, at all of the synthesis concentrations, the calculated activity of PAN-ATA exceeds that of the baseline, suggesting that the quality of UiO-66-NH2 grown on PAN-ATA swatches is more accessible and therefore resulting in higher reactive capacity for chlorine. On the basis of the chlorine loading measurements, the in situ technique produces MOF fiber composites that have up to 50% MOF loading and activity when compared to powder UiO-66-NH2.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.iecr.7b04202. Materials and Methods, instruments used for characterization, image of electrospun PAN with preformed PAN, SEM images of PAN and PAN-ATA at different synthesis conditions, nitrogen adsorption isotherms, ATR-FTIR data (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Annie X. Lu: 0000-0001-7179-2015 Gregory W. Peterson: 0000-0003-3467-5295 Jared B. DeCoste: 0000-0003-0345-2697 Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Funding

This project was supported by the Joint Science and Technology Office for Chemical Biological Defense (JSTO− CBD) under project number CB3934/BA13PHM210. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors thank Mr. Matthew Browe for performing the microbreakthrough experiments, Dr. Erica Valdes for assistance with SEM imaging, and Professor Gregory Parsons for helpful discussions. This research was performed while Dr. Trenton Tovar held a National Research Council (NRC) Post-Doctoral Fellowship at the Edgewood Chemical Biological Center (ECBC).



CONCLUSIONS In conclusion, we present a scalable method to grow UiO-66NH2 directly on the surface of PAN nanofibers. Through various characterization methods, we show that we can grow up to 50 wt %, highly crystalline, high-performance MOFs. All of the surface-grown MOFs are considered accessible as our synthesis method does not involve mixing the MOFs with the polymer solution. We envision that this method that can be



REFERENCES

(1) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keeffe, M.; Yaghi, O. M. Systematic design of pore size and functionality in

D

DOI: 10.1021/acs.iecr.7b04202 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research isoreticular MOFs and their application in methane storage. Science 2002, 295, 469−472. (2) Li, J. R.; Kuppler, R. J.; Zhou, H. C. Selective gas adsorption and separation in metal- organic frameworks. Chem. Soc. Rev. 2009, 38, 1477−1504. (3) Li, J. R.; Sculley, J.; Zhou, H. C. Metal-organic frameworks for separations. Chem. Rev. 2012, 112, 869−932. (4) Ma, L. Q.; Abney, C.; Lin, W. B. Enantioselective catalysis with homochiral metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1248−1256. (5) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Metal-organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38, 1450−1459. (6) Mondloch, J. E.; Katz, M. J.; Isley, W. C., III; Ghosh, P.; Liao, P. L.; Bury, W.; Wagner, G.; Hall, M. G.; DeCoste, J. B.; Peterson, G. W.; Snurr, R. Q.; Cramer, C. J.; Hupp, J. T.; Farha, O. K. Destruction of chemical warfare agents using metal-organic frameworks. Nat. Mater. 2015, 14, 512−516. (7) DeCoste, J. B.; Browe, M. A.; Wagner, G. W.; Rossin, J. A.; Peterson, G. W. Removal of chlorine gas by an amine functionalized metal-organic framework via electrophilic aromatic substitution. Chem. Commun. 2015, 51, 12474−12477. (8) Peterson, G. W.; Mahle, J. J.; DeCoste, J. B.; Gordon, W. O.; Rossin, J. A. Extraordinary NO2 Removal by the Metal-Organic Framework UiO-66-NH2. Angew. Chem., Int. Ed. 2016, 55, 6235− 6238. (9) Zhang, Y.; Yuan, S.; Feng, X.; Li, H.; Zhou, J.; Wang, B. Preparation of Nanofibrous Metal-Organic Framework Filters for Efficient Air Pollution Control. J. Am. Chem. Soc. 2016, 138, 5785−8. (10) Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent functional metal-organic frameworks. Chem. Rev. 2012, 112, 1126−62. (11) Moreton, J. C.; Denny, M. S.; Cohen, S. M. High MOF loading in mixed-matrix membranes utilizing styrene/butadiene copolymers. Chem. Commun. 2016, 52, 14376−14379. (12) DeCoste, J. B.; Denny, M. S., Jr.; Peterson, G. W.; Mahle, J. J.; Cohen, S. M. Enhanced aging properties of HKUST-1 in hydrophobic mixed-matrix membranes for ammonia adsorption. Chem. Sci. 2016, 7, 2711−2716. (13) Khdhayyer, M. R.; Esposito, E.; Fuoco, A.; Monteleone, M.; Giorno, L.; Jansen, J. C.; Attfield, M. P.; Budd, P. M. Mixed matrix membranes based on UiO-66 MOFs in the polymer of intrinsic microporosity PIM-1. Sep. Purif. Technol. 2017, 173, 304−313. (14) Zhang, Y. Y.; Yuan, S.; Feng, X.; Li, H. W.; Zhou, J. W.; Wang, B. Preparation of Nanofibrous Metal-Organic Framework Filters for Efficient Air Pollution Control. J. Am. Chem. Soc. 2016, 138, 5785− 5788. (15) Zhao, J. J.; Lee, D. T.; Yaga, R. W.; Hall, M. G.; Barton, H. F.; Woodward, I. R.; Oldham, J.; Walls, H. J.; Peterson, G. W.; Parsons, G. N. Ultra-Fast Degradation of Chemical Warfare Agents Using MOFNanofiber Kebabs. Angew. Chem., Int. Ed. 2016, 55, 13224−13228. (16) Gao, M.; Zeng, L. W.; Nie, J.; Ma, G. P. Polymer-metal-organic framework core-shell framework nanofibers via electrospinning and their gas adsorption activities. RSC Adv. 2016, 6, 7078−7085. (17) Bunge, M. A.; Ruckart, K. N.; Leavesley, S.; Peterson, G. W.; Nguyen, N.; West, K. N.; Glover, T. G. Modification of Fibers with Nanostructures Using Reactive Dye Chemistry. Ind. Eng. Chem. Res. 2015, 54, 3821−3827. (18) Yu, M.; Li, W. X.; Wang, Z. Q.; Zhang, B. W.; Ma, H. J.; Li, L. F.; Li, J. Y. Covalent immobilization of metal-organic frameworks onto the surface of nylon-a new approach to the functionalization and coloration of textiles. Sci. Rep. 2016, 6.10.1038/srep22796 (19) Ostermann, R.; Cravillon, J.; Weidmann, C.; Wiebcke, M.; Smarsly, B. M. Metal-organic framework nanofibers via electrospinning. Chem. Commun. 2011, 47, 442−4. (20) Wu, Y. N.; Li, F. T.; Liu, H. M.; Zhu, W.; Teng, M. M.; Jiang, Y.; Li, W. N.; Xu, D.; He, H.; Hannam, P.; Li, G. T. Electrospun fibrous mats as skeletons to produce free-standing MOF membranes. J. Mater. Chem. 2012, 22, 16971−16978.

(21) da Silva Pinto, M.; Sierra-Avila, C. A.; Hinestroza, J. P. In situ synthesis of a Cu-BTC metal- organic framework (MOF 199) onto cellulosic fibrous substrates: cotton. Cellulose 2012, 19, 1771−1779. (22) Zhang, Y. Y.; Feng, X.; Yuan, S.; Zhou, J. W.; Wang, B. Challenges and recent advances in MOF-polymer composite membranes for gas separation. Inorg. Chem. Front. 2016, 3, 896−909. (23) Semino, R.; Ramsahye, N. A.; Ghoufi, A.; Maurin, G. Microscopic Model of the Metal Organic Framework/Polymer Interface: A First Step toward Understanding the Compatibility in Mixed Matrix Membranes. ACS Appl. Mater. Interfaces 2016, 8, 809− 819. (24) Wahiduzzaman; Khan, M. R.; Harp, S.; Neumann, J.; Sultana, Q. N. Processing and Performance of MOF (Metal Organic Framework)Loaded PAN Nanofibrous Membrane for CO2 Adsorption. J. Mater. Eng. Perform. 2016, 25, 1276−1283. (25) Bechelany, M.; Drobek, M.; Vallicari, C.; Abou Chaaya, A.; Julbe, A.; Miele, P. Highly crystalline MOF-based materials grown on electrospun nanofibers. Nanoscale 2015, 7, 5794−5802. (26) Brown, A. J.; Johnson, J. R.; Lydon, M. E.; Koros, W. J.; Jones, C. W.; Nair, S. Continuous Polycrystalline Zeolitic Imidazolate Framework-90 Membranes on Polymeric Hollow Fibers. Angew. Chem., Int. Ed. 2012, 51, 10615−10618. (27) Huo, S. H.; Yu, J.; Fu, Y. Y.; Zhou, P. X. In situ hydrothermal growth of a dual-ligand metal-organic framework film on a stainless steel fiber for solid-phase microextraction of polycyclic aromatic hydrocarbons in environmental water samples. RSC Adv. 2016, 6, 14042−14048. (28) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K. P. A new zirconium inorganic building brick forming metal organic frameworks with exceptional stability. J. Am. Chem. Soc. 2008, 130, 13850−13851. (29) Chavan, S.; Vitillo, J. G.; Gianolio, D.; Zavorotynska, O.; Civalleri, B.; Jakobsen, S.; Nilsen, M. H.; Valenzano, L.; Lamberti, C.; Lillerud, K. P.; Bordiga, S. H-2 storage in isostructural UiO-67 and UiO-66 MOFs. Phys. Chem. Chem. Phys. 2012, 14, 1614−1626. (30) Katz, M. J.; Moon, S. Y.; Mondloch, J. E.; Beyzavi, M. H.; Stephenson, C. J.; Hupp, J. T.; Farha, O. K. Exploiting parameter space in MOFs: a 20-fold enhancement of phosphate-ester hydrolysis with UiO-66-NH2. Chem. Sci. 2015, 6, 2286−2291. (31) Moon, S. Y.; Wagner, G. W.; Mondloch, J. E.; Peterson, G. W.; DeCoste, J. B.; Hupp, J. T.; Farha, O. K. Effective, Facile, and Selective Hydrolysis of the Chemical Warfare Agent VX Using Zr-6-Based Metal-Organic Frameworks. Inorg. Chem. 2015, 54, 10829−10833. (32) Peterson, G. W.; DeCoste, J. B.; Fatollahi-Fard, F.; Britt, D. K. Engineering UiO-66-NH2 for Toxic Gas Removal. Ind. Eng. Chem. Res. 2014, 53, 701−707. (33) Lee, D. T.; Zhao, J. J.; Peterson, G. W.; Parsons, G. N. Catalytic ″MOF-Cloth″ Formed via Directed Supramolecular Assembly of UiO66-NH2 Crystals on Atomic Layer Deposition-Coated Textiles for Rapid Degradation of Chemical Warfare Agent Simulants. Chem. Mater. 2017, 29, 4894−4903. (34) Ploskonka, A. M.; Marzen, S. E.; DeCoste, J. B. Facile Synthesis and Direct Activation of Zirconium Based Metal-Organic Frameworks from Acetone. Ind. Eng. Chem. Res. 2017, 56, 1478−1484.

E

DOI: 10.1021/acs.iecr.7b04202 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX