Hierarchically Porous Covalent Organic Framework Nanostructures for

Figure 1. a) Schematic representation of COF–foam synthesis using an in situ gas-foaming technique. Different amines ... the interconnected pore cha...
0 downloads 0 Views 6MB Size
Subscriber access provided by University of Otago Library

Article

Inducing Disorder in Order: Hierarchically Porous Covalent Organic Framework Nanostructures for Rapid Removal of Persistent Organic Pollutants Suvendu Karak, Kaushik Dey, Arun Torris A. T., Arjun Halder, Saibal Bera, Fayis KP, and Rahul Banerjee J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.9b02706 • Publication Date (Web): 21 Apr 2019 Downloaded from http://pubs.acs.org on April 21, 2019

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 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 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.

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 18 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

Journal of the American Chemical Society

Inducing Disorder in Order: Hierarchically Porous Covalent Organic Framework Nanostructures for Rapid Removal of Persistent Organic Pollutants Suvendu Karak,1,2 Kaushik Dey,3 Arun Torris,4 Arjun Halder,1,2 Saibal Bera1,2 Fayis Kanheerampockil4 and Rahul Banerjee3,* 1Academy

of Scientific and Innovative Research, New Delhi, India. /Materials Chemistry Division, CSIR-National Chemical Laboratory, Dr. Homi Bhabha Road, Pune-411008, India. 3Department of Chemical Sciences, Indian Institute of Science Education and Research, Kolkata, Mohanpur 741246, India. 4Polymer Science and Engineering Division, CSIR-National Chemical Laboratory, Dr. Homi Bhabha Road, Pune-411008, India. 2Physical

*Email: [email protected]; Tel: +91-20-2590-2535. ABSTRACT: The key factor responsible for fast diffusion and mass transfer through a porous material is the availability of a widely open pore-interior having complete accessibility from their surface. But, due to their highly stacked nature, ordered two dimensional (2D) materials fail to find real-world applicability, as it is difficult to take advantage of their complete structure, especially the inner cores. In this regard, three dimensional (3D) nano-structures constructed from layered two-dimensional crystallites could prove advantageous. However, the real challenge is to cultivate a porous nanostructure with ordered pores where the pores are surrounded by crystalline walls. Herein, a simple yet versatile in situ gas phase foaming technique has been employed to address these cardinal issues. The use of baking soda leads to the continuous effervescence of CO2 during the crystallization of foam, which creates ripples and fluctuations on the surface of the 2D crystallites. The induction of ordered micropores within the disordered 3D architecture synergistically renders fast diffusion of various guests through the interconnected pore network. The high-density defects in the hierarchically porous structure help in ultrafast adsorption (99% of total pore volume)

Fig. 4. The mechanism of foam synthesis. The plausible mechanism has been represented depending upon the SEM and

TEM images. The stepwise synthesis reveals that the foams are formed by controlled gas foaming through the 2D nano-sheets. The numerical values that have been mentioned herein represent the scale bars for the sequential steps: fibers (5 μm), sheets (2 μm), sheets having defects (1 μm), COF-foam (5 μm).

throughout its volume. The identification of pores in submicron size was not possible due to the resolution limit of the imaging facility. As evident from the pore-size distribution profile, crystalline COF powders contain a maximum distribution of pores in the lower size range (3.1 to 82.7 microns), whereas the corresponding crystalline COF-foam contains more pores with volume fraction in the higher size range (2.3 to 390 microns) (Figure 2m and Figure S15, SI). These pores with higher diameter contribute significantly to the overall pore volume fraction of the foam. To check the mechanical strength of the TpPa-2-foam, nano-mechanical mapping measurement was performed using atomic force microscopy (AFM). The calculation reveals the Young’s modulus of 4.21 GPa at a scanning range of 2 μm × 2 μm (Figure S16, SI). The N2 adsorption analyses of the TpPa-2, TpPa-NO2, TpAzo, TpBD-Me2–foams provide the key information related to their porosity and the nature of the pore channels (Figure 3, Section S-8, SI). The N2 adsorption analysis was performed at 77 K temperature with the activated foam samples. TpPa-2, TpPaNO2, TpAzo, TpBD-Me2–foams show surface area of 579, 254, 1054 and 797 m2g-1 respectively (Figure 3a). The hysteresis present in the isotherm could be assigned to the defects present i.e. the random pores within the foam matrix. The high gas adsorption near to 1 bar pressure could be attributed to the capillary condensation, which further confirms the presence of mesopores. Such hysteresis loop is absent in the N2 adsorption isotherms of the powdered crystalline COF samples.5 The pore size distributions have been calculated by non-linear density functional theory (NLDFT), and they clearly confirm the presence of  2 nm sized micropores within the foam (Figure 3b and Figure S17, SI). These micropores are due to the presence of the microporous 2D crystallites within the foam matrix. The low intensity ( 0.05 unit) of the pore size distribution is due to the lower order among the pore channels. The Barrett-JoynerHalenda (BJH) analysis suggests that instead of regular pore size, the macropores in the COF-foam range from 10 to 150 nm, which proves the irregularities within the network (Figure 3c and Figure S18, SI). We have employed the SEM and TEM imaging techniques to understand the mechanism behind the induction of the controlled disorder in the ordered microporous COFs (Figure 4). First of all, the mixing of amine and PTSA leads to the formation of the PTSA-amine salt, which adopts a fiber-like morphology. These fibers (~2-6 μm of length ~50-100 nm of width) are formed via the Namine–HOacid [D, d, ; Davg = 2.76 Å, davg = 1.88 Å;  = 176°; PTSA-Pa-2 salt]12d hydrogenbonding between the amine functionality (–NH3+) of the diamine molecules and the sulphonate groups (–SO3–) from the PTSA.12d,15 The addition of aldehyde Tp initiates the interconnection among the fibers, which leads to the formation of 2D sheets (with ~5 μm of width). The addition of water further improves the stacking of the sheets. When bicarbonate is added to the mixture, the excess PTSA, after the imination reaction, readily reacts with the bicarbonate resulting in continuous CO2 effervescence. The generated CO2 passes

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

through the 2D COF-sheets and creates pores, thereby inducing disorder among the ordered COF crystallites. The emergence of random pores is reasonable, as sequential addition of bicarbonate was performed for a couple of minutes. The continuous effervescence helps in raising the COF-paste into the foam. Freeze-drying helps in holding the shape of the 3D nano-structure and creates additional pores

Page 6 of 18

due to the fast removal of water from the micropores. The dried foam was then kept inside the oven at 90 °C for 12 hrs. The unreacted bicarbonate releases excess CO2 during heating, which further generates extrinsic porosity. The persistent gas releasing results in open cell foams in the polymer like intermediate until the crystallization starts at 90 °C. The Bronsted acid

Figure 5. a) Chem Draw representation of the pollutants that have been removed efficiently. b) The activity of foam at various pH ranges. c) First adsorption kinetics of different pollutants from water using TpPa-2-foam. d) The pollutant removal efficiency of TpPa-2-foam. e) Model representation of dye molecules inside the foam micropores. f) The confocal image hints at the nature of adsorption and the morphological evolution in the foam as well. g) Recyclability of TpPa-2-foam shows their insignificant changes in performance after 5 cycles.

controls the dynamic covalent connections between the building units, thereby reconstructing the framework and maintaining the crystallinity of the system.16 PTSA could further help in retaining the crystallinity during the foaming in the presence of the NaHCO3 base (if any bond-breakage takes place). This is crucial for the foam making, as inducing defects by any base after the stacking (completion of crystallization)

could lead to the delamination of the sheets or degradation of the framework (Figure S19, SI). Again, the bicarbonate addition is limited to a certain extent, as excess addition could lead to the total disruption of the framework or complete exfoliation of the sheets. However, the extent of porosity could be controlled through the limited effervescence to achieve increased mass transfer.

ACS Paragon Plus Environment

Page 7 of 18 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

Journal of the American Chemical Society

Unlike oxidative degradation, photocatalysis or other methods of separating pollutants, adsorption-based separation is advantageous owing to the cost-effectiveness, ease of operation and its property of fast removal. In parallel, the incorporation of defects within the crystalline foam material has provided the richness of solution-state host–guest chemistry in the solid state. The microstructure merits of the TpPa-2, TpPa-NO2, TpAzo, TpBD-Me2–foams has endowed

them very low density. This low density, along with their super strong adsorption ability is illustrated by the rapid removal of various dyes and inorganic-organic pollutants from the water (Section S-10, SI). Among all the foams, TpPa-2 shows the highest efficiency, and it removes pollutants within 10 seconds of the addition (Fig. 5c-5d). Pollutants such as Methylene Blue (MB, ~ 99%), Rhodamine B (RhB, ~ 96%), Acid Fuchsin (AF, ~ 99%), Rose Bengal

Figure 6. a) Comparison of uptake capacity of methylene blue between the COF-foam and pristine COF from the absorbance spectra. b) The time-dependent removal efficiency of COF-foam and pristine COF. c) Comparison of the histogram of pore-size distribution from Xray computed tomography between the COF-foam and COF. d) Confocal imaging shows the adsorption of dyes in the pores and the presence of various defects within the foam. e), f) and g) The 2.5D confocal image shows the distribution of dye inside the foam. h) Surface confocal imaging (Surface-rendered projection from 3d confocal micrograph) of COF shows the adsorption of dyes on the surface only. i, j, and k) Adsorption of dye only on the surface for the COF can be seen by the nature of adsorption. l) and m) Color-coded visualization of flow velocity inside the COF-foam matrix. n) Color-coded visualization of pressure variation inside a foam block and o) 3D image of COF and visualization of its flow velocity. All the experiments, including the tomographic experiments have been performed with respect to TpPa-2–foam.

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

(RB, ~ 99%), Iodine (I2, ~ 99%), Potassium Permanganate (KMnO4, ~ 96%), Bisphenol A (BPA, ~ 79%) and Decabromodiphenyl Ether (decaBDE, ~ 70%) have been removed efficiently from water using TpPa-2-foam (Figure 5d, Figure S20, SI). TpPa-NO2-foam also removes almost 99% of dyes (MB ~ 99%, RhB ~ 98%, AF ~ 99%, RB~ 99%) from the water (Figure S21, SI). Other foams such as TpBD-Me2 and TpAzo-foam exhibit efficient removal of all the dyes (7299%) from water (Figures S22 and S23, SI). The high percentage of removal efficiency has also been validated in different pH (1-12) to make the foam more vulnerable to the environmental conditions. The experiment reveals that the TpPa-2-foam can retain its activity (removal efficiency of MB up to 99%) in the entire pH range of 1-12 (Figure 5b and Figure S24, SI).17 The TpPa-2-foam has also been utilized for 5 cycles without affecting its performance (Figure 5g and Figure S25, SI). After performing one cycle, the foam was treated with acetone to collect the adsorbent. The foam was then reused for the next cycle. The stability of the foam in water also eliminates the chance of any secondary pollution from the adsorbent (Figures S26 and S27, SI). Due to the continuous pore channels, pollutants diffused through the macropores and remained adsorbed inside the micropores and meso-pores (Figure 5f, Section S-11). In order to explore their removal efficiency, we have measured their kinetics of adsorption. It is important to note that TpPa-2 foam almost reaches 90% of its equilibrium efficiency within 10 seconds (Figure 5c, insight). The pseudo-second order rate constant (kobs) of MB, RhB, AF, RB, BPA and I2 adsorption are 11.8, 7.4, 1.4, 30.3, 182.3 and 4.3 g mg-1 min-1 (Figures S28 and S29, SI). Among all, the crucial plastic pollutant, BPA, removal rate constant is the highest among all reported porous pollutant removers such as β-cyclodextrin-based polymers (~10 times higher) and other carbon-based porous materials (almost 100 times higher) (Table S3, SI).18 The high rate constants of iodine and BPA could be attributed to the small size of the pollutants that fit the ordered micro-pores. Interconnected pores also offer the complete access of the ordered intrinsic nano-pores of the COF-foam unlike the random networks within the porous amorphous polymers. The maximum uptake capacity of MB at equilibrium is found to be 108 mg g-1, which is also very high (Figures S30 and S31, SI). The high surface area of the foam enhances the efficiency of the dye adsorption. The fast uptake has been further probed by determining the flow-through experiment where MB solution of 50 μM was passed through a packed column of TpPa-2foam. The flow rate was also improved with air pressure without altering the removal efficiency of ~ 99% (Figure S32, SI). In order to support the fast removal efficiency of the foams, the flow velocity of water through the interconnected pores has been estimated by numerical simulations using the FlowDict software package (Section S-12, SI). The flow of water was simulated along the Z-axis of the segmented 3D image model of COF-foam with a pressure drop of 20 Pa (Figure 6n). The average flow velocity in the TpPa-2-foam in the X-direction is 0.0015 ms-1 and in the Y- and Z-direction is 0.0022 and 0.0028 ms-1 respectively. This contributes significantly towards the high pollutant adsorption into the foam matrix (Figures 6l-n). Furthermore, due to the interconnected hierarchical porous nature of the foam, the range of velocities was observed within a block in all the directions (Figure 6l). TpPa-NO2, TpBD-Me2 and TpAzo-

Page 8 of 18

foam also exhibit high average flow rate of 0.0012, 0.0014 and 0.0011 ms-1 respectively (Figure S35, SI). Both TpPa-2–foam and TpPa-2 COF powder were tested for the removal of Methylene Blue from the water. It has been observed that COF-foam can remove more than 99% of pollutant (MB) within 10 seconds as compared to the COF powder that can only efficiently remove 50% after 3 hrs (Figures 6a and b). This can be attributed to their less interconnected pore channels and highly stacked 2D structures, which make the inner surfaces less accessible for guest inclusion. This can be easily shown from the 2.5D confocal images (Figures 6h-k). The time-dependent confocal study reveals that in any time interval, COF adsorbs the dyes only on the crystallite surface. The periphery of dye molecules around the COF powder proves the inability of the dyes in penetrating the surface of the COF. However, in the case of the foam, the macropores on the wall provide easy infiltration for the pollutants (distribution of dyes), and interconnected pore channels allow continuous mass flow, unlike that of the COF powder (Figures 6d-g). The random distribution of dye molecule within the foam-matrix proves the aspect mentioned above. The confocal images, represented herein, are collected from different layers in different time intervals in the equilibrium condition. The flow velocity simulation study from tomography also proves the superiority of foams over the COFs. The average flow velocity in TpPa-2-foam is 0.0028 ms-1 as compared to 0.0004 ms-1 in COF powder (Figures 6m and 6o). Again, the pseudo-second order rate constant (kobs) for the adsorption of MB for TpPa-2 COF is 0.0026 g mg-1 min-1 whereas the TpPa-2-foam exhibits the value of 11.8 g mg-1 min-1 (Figure S33, SI). Thus, a wide range of pore size distribution (mentioned earlier), high flow velocity and ease of accessibility to the internal pores make the foam a suitable candidate for ultrafast pollutant removal. CONCLUSION In pursuit of a new strategy for boosting the adsorption property of two-dimensional COFs, we have endeavored to synthesize hierarchically porous covalent organic framework foams via a simple yet widely applicable gas foaming technique. The use of baking soda helps in the rising of foam, and this further creates different pores within the framework matrix. Employing this technique, we have succeeded in forming a 3D architecture decorated with a 2D covalent network. The intentionally created surface defects and interconnected pore structures help in fast adsorption followed by diffusion of pollutants through the pore channel. The foam has been used to remove various pollutants e.g. dyes, iodine, potassium permanganate and Bisphenol A and persistent organic pollutant, Decabromodiphenyl Ether. The foam has been used as a scavenger for the removal of micropollutants from the water even at very low concentrations. The fast uptake property, recyclability and ability to adsorb different pollutants indicate that these newly developed COF-foams deserve further exploration.

ASSOCIATED CONTENT Supporting Information. Synthesis, crystallography and characterization details are provided in Supporting Information file. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION

ACS Paragon Plus Environment

Page 9 of 18 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

Journal of the American Chemical Society

Corresponding Author

4.

* [email protected]

Author Contributions Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT S. K acknowledges University Grants Commission, Government of India for Senior Research Fellowship. K.D acknowledges CSIR for fellowship. R. B. acknowledges Swarna Jayanti Fellowship grant [DST/SJF/CSA-02/2016-2017] for funding for funding. Authors would also like to thank Dr. Mehdi Azimian (Math2Market GmbH) for the short-term license to PoroDict software module and Dr. Matthew Addicoat for the theoretical calculation.

REFERENCES 1.

2.

3.

(a) Jie, K.; Liu, M.; Zhou, Y.; Little, M. A.; Pulido, A.; Chong, S. Y.; Stephenson, A.; Hughes, A. R.; Sakakibara, F.; Ogoshi, T.; Blanc, F.; Day, G. M.; Huang, F.; Cooper, A. I. Near-Ideal Xylene Selectivity in Adaptive Molecular Pillar[n]arene Crystals. J. Am. Chem. Soc. 2018, 140, 6921−6930. (a) Yaghi, O. M.; O'Keeffe, M; Ockwig, N.; Chae, H. K.; Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new materials. Nature 2003, 423, 705−714. (b) Kitagawa, S.; Kitaura, R.; Noro, S. Functional Porous Coordination Polymers. Angew. Chem. Int. Ed. 2004, 43, 2334−2375. (c) Lehn, J.-M. Supramolecular Chemistry (VCH, 1995). (d) Inokuma, Y.; Arai, T.; Fujita, M. Networked molecular cages as crystalline sponges for fullerenes and other guests. Nature Chemistry 2010, 2, 780−783. (e) Sanchez, C.; Arribart, H.; Guille, M. M. G. Biomimetism and bioinspiration as tools for the design of innovative materials and systems. Nat. Mater. 2005, 4, 277−288. (f) Bae, W.-G.; Kim, H. N.; Kim, D.; Park, S.-H.; Jeong, H. E.; Suh, K.-Y. 25th Anniversary Article: Scalable Multiscale Patterned Structures Inspired by Nature: the Role of Hierarchy. Adv. Mater. 2014, 26, 675−700. (a) Côté, A. P.; Benin, A. I.; Ockwig, N. W.; O’Keeffe, M.; Matzger, A. J.; Yaghi, O. M. Porous, Crystalline, Covalent Organic Frameworks. Science 2005, 310, 1166−1170. (b) Kandambeth, S.; Mallick, A.; Lukose, B.; Mane, M. V.; Heine, T.; Banerjee, R. Construction of Crystalline 2D Covalent Organic Frameworks with Remarkable Chemical (Acid/Base) Stability via a Combined Reversible and Irreversible Route. J. Am. Chem. Soc. 2012, 134, 19524−19527. (c) Vyas, V. S.; Vishwakarma, M.; Moudrakovski, I.; Haase, F.; Savasci, G.; Ochsenfeld, C.; Spatz, J. P.; Lotsch, B. V. Exploiting Noncovalent Interactions in an Imine‐Based Covalent Organic Framework for Quercetin Delivery. Adv. Mater. 2016, 28, 8749−8754. (d) Kuhn, P.; Antonietti, M.; Thomas, A. Porous, Covalent Triazine‐Based Frameworks Prepared by Ionothermal Synthesis. Angew. Chem. Int. Ed. 2008, 47, 3450−3453 (e) Liu, X-H.; Guan, C-Z.; Ding, S-Y.; Wang, W.; Yan, H-J.; Wang, D.; Wan, L-J. On-Surface Synthesis of Single-Layered TwoDimensional Covalent Organic Frameworks via Solid–Vapor Interface Reactions. J. Am. Chem. Soc. 2013, 135, 10470−10474. (f) Ding, S. Y.; Gao, J.; Wang, Q.; Zhang, Y.; Song, W. G.; Su, C. Y.; Wang, W. Construction of Covalent Organic Framework for Catalysis: Pd/COF-LZU1 in Suzuki– Miyaura Coupling Reaction. J. Am. Chem. Soc. 2011, 133, 19816−19822. (g) Ma, H.; Ren, H.; Meng, S.; Yan, Z.; Zhao, H.; Sun, F.; Zhu, G. A 3D microporous covalent organic framework with exceedingly high C3H8/CH4 and C2 hydrocarbon/CH4 selectivity. Chem. Commun. 2013, 49, 9773−9775. (h) Dogru, M.; Sonnauer, A.; Gavryushin, A.; Knochel, P.; Bein, T. A. Covalent Organic Framework with 4 nm open pores. Chem. Commun. 2011, 47, 1707-1709.

Witman, M.; Ling, S.; Jawahery, S.; Boyd, P. G.; Haranczyk, M.; Slater, B.; Smit, B. The Influence of Intrinsic Framework Flexibility on Adsorption in Nanoporous Materials. . J. Am. Chem. Soc., 2017, 139, 5547–5557. 5. (a) Cao, X.; Yin, Z.; Zhang, H. Three-dimensional graphene materials: preparation, structures and application in supercapacitors. Energy Environ. Sci. 2014, 7, 1850−1865. (b) Li, C.; Shi, G. Three-dimensional graphene architectures. Nanoscale 2012, 4, 5549−5563. (c) Lu, A.; Hao, G.; Sun, Q. Design of three-dimensional porous carbon materials: from static to dynamic skeletons. Angew. Chem. Int. Ed. 2013, 52, 7930−7932. 6. (a) Baldwin, L. A.; Crowe, J. W.; Shannon, M.D.; Jaroniec, C.P.; McGrier, P.L. 2D Covalent Organic Frameworks with Alternating Triangular and Hexagonal Pores. Chem. Mater. 2015, 27, 6169−6172. (b) Qian, C.; Qi, Q.Y.; Jiang, G.F.; Cui, F.Z.; Tian, Y.; Zhao, X. Toward Covalent Organic Frameworks Bearing Three Different Kinds of Pores: The Strategy for Construction and COF-to-COF Transformation via Heterogeneous Linker Exchange. J. Am. Chem. Soc. 2017, 139, 6736−6743. (c) Dalapati, S.; Jin, E.; Addicoat, M.; Heine, T.; Jiang, D. Highly emissive covalent organic frameworks. J. Am. Chem. Soc. 2016, 138, 5797−5800. (d) Du, Y.; Yang, H.; Whiteley, J. M.; Wan, S.; Jin, Y.; Lee, S.-H.; Zhang, W. Ionic covalent organic frameworks with spiroborate linkage. Angew. Chem.Int. Ed. 2016, 55, 1737-1741. (e) Chen, Y.; Huang, X.; Zhang, S.; Li, S.; Cao, S.; Pei, X.; Zhou, J.; Feng, X.; Wang, B. Shaping of Metal–Organic Frameworks: From Fluid to Shaped Bodies and Robust Foams. J. Am. Chem. Soc., 2016, 138, 10810−10813. 7. (a) Xie. A. F.; Granick. S. Phospholipid membranes as substrates for polymer adsorption. Nat. Mater. 2002, 1, 129−133. (b) Zhao, H.; Song, X.; Zeng. H. 3D white graphene foam scavengers: vesicant-assisted foaming boosts the gram-level yield and forms hierarchical pores for superstrong pollutant removal applications. NPG Asia Materials 2015, 7, 168. (c) Hartmann, S.; Brandhuber, D.; Hüsing, N. Glycol-Modified Silanes: Novel Possibilities for the Synthesis of Hierarchically Organized (Hybrid) Porous Materials. Acc. Chem. Res. 2007, 40, 885−894. (d) Davis. M. E. Ordered porous materials for emerging applications. Nature 2002, 417, 813−821. (e) Yang, P.; Deng, T.; Zhao, D.; Feng, P.; Pine, D.; Chmelka, B. F.; Whitesides, G. M.; Stucky, G. D. Hierarchically ordered oxides. Science 1998, 282, 2244−2246. 8. (a) Batten, S. R.; Robson, R. Interpenetrating Nets: Ordered, Periodic Entanglement. Angew. Chem. 1998, 37, 1460−1494. (b) Ockwig, N. W.; Delgado-Friedrichs, O.; OKeeffe, M.; Yaghi, O. M. Reticular Chemistry:  Occurrence and Taxonomy of Nets and Grammar for the Design of Frameworks. Acc. Chem. Res. 2005, 38, 176−182. (c) Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastre, J. Metal–organic frameworks— prospective industrial applications. J. Mater. Chem. 2006, 16, 626−636. 9. (a) Seo, M.; Kim, S.; Oh, J.; Kim, S-J.; Hillmyer M. A. Hierarchically Porous Polymers from Hyper-cross-linked Block Polymer Precursors. J. Am. Chem. Soc. 2015, 137, 600−603. (b) Schwab, M. G.; Senkovska, I.; Rose, M.; Klein, N.; Koch, M.; Pahnke, J.; Jonschker, G.; Schmitz, B.; Hirscher, M.; Kaskel, S. High surface area poly HIPEs with hierarchical pore system. Soft Matter 2009, 5, 1055−1059. (c) Pulko, I.; Wall, J.; Krajnc, P.; Cameron, N. R. Ultra‐High Surface Area Functional Porous Polymers by Emulsion Templating and Hypercross linking: Efficient Nucleophilic Catalyst Supports. Chem. Eur. J. 2010, 16, 2350−2354. 10. (a) Kuang, D.; Brezesinski, T.; Smarsly, B. Hierarchical Porous Silica Materials with a Trimodal Pore System Using Surfactant Templates. J. Am. Chem. Soc. 2004, 126, 10534−10535. (b) Mandlmeier, B.; Szeifert, J. M.; Fattakhova-Rohlfing, D.; Amenitsch, H.; Bein, T. Formation of Interpenetrating Hierarchical Titania Structures by Confined Synthesis in Inverse

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

11.

12.

13.

14.

15.

Opal. J. Am. Chem. Soc. 2011, 33, 17274−17282. (c) Dacquin, J.-P.; Dhainaut, J.; Duprez, D.; Royer, S.; Lee, A. F.; Wilson, K. An Efficient Route to Highly Organized, Tunable Macroporous−Mesoporous Alumina. J. Am. Chem. Soc. 2009, 131, 12896−12897. (d) Shi, Y.; Zhang, F.; Hu, Y.-S.; Sun, X.; Zhang, Y.; Lee, H. I.; Chen, L.; Stucky, G. D. Low-Temperature Pseudomorphic Transformation of Ordered Hierarchical Macromesoporous SiO2/C Nanocomposite to SiC via Magnesiothermic Reduction. J. Am. Chem. Soc. 2010, 132, 5552−5553. (a) Worsley, M. A.; Pauzauskie, P. J.; Olson, T. Y.; Biener, J.; Satcher, J. H.; Baumann, T. F. Synthesis of Graphene Aerogel with High Electrical Conductivity. J. Am. Chem. Soc. 2010, 132, 14067−14069. (b) Xu, Y.; Wu, Q. O.; Sun, Y. Q.; Bai, H.; Shi, G. Q. Three-Dimensional Self-Assembly of Graphene Oxide and DNA into Multifunctional Hydrogels. ACS Nano 2010, 4, 7358−7362. (c) Tang, Z. H.; Shen, S. L.; Zhuang, J.; Wang, X. Noble‐Metal‐Promoted Three‐Dimensional Macroassembly of Single‐Layered Graphene Oxide. Angew. Chem. Int. Ed. 2010, 49, 4603−4607. (d) Lee, S. H.; Kim, H. W.; Hwang, J. O.; Lee, W. J.; Kwon, J.; Bielawski, C. W.; Ruoff, R. S.; Kim, S. O. Three-dimensional self-assembly of graphene oxide platelets into mechanically flexible macroporous carbon films. Angew. Chem. Int. Ed. 2010, 49, 10084−10088. (e) Chen, Z. P.; Ren, W. C.; Gao, L. B.; Liu, B. L.; Pei, S. F.; Cheng, H. M. Threedimensional flexible and conductive interconnected graphene networks grown by chemical vapour deposition. Nat. Mater. 2011, 10, 424−428. (f) Kang, L.; Carlos, C.; Mark J. S.; Raffaele, R.; Jiwei, C.; Joseph J. R.; Daniel, M.; Frank, C.; Paolo, F. Biomimetic Replication of Microscopic Metal–Organic Framework Patterns Using Printed Protein Patterns. Adv. Mater. 2015, 27, 7293–7298. (a) Kandambeth, S.; Biswal, B. P.; Chaudhari, H. D.; Rout, K. C.; Kunjattu, H. S.; Mitra, S.; Karak, S.; Das, A.; Mukherjee, R.; Kharul, U. K.; Banerjee, R. Selective Molecular Sieving in Self‐Standing Porous Covalent‐Organic‐Framework Membranes. Adv. Mater. 2017, 29, 1603945. (b) Karak, S.; Kandambeth, S.; Biswal, B. P.; Sasmal, H. S.; Kumar, S.; Pachfule, P.; Banerjee, R. Constructing Ultraporous Covalent Organic Frameworks in Seconds via an Organic Terracotta Process. J. Am. Chem. Soc. 2017, 139, 1856−1862. (c) Mitra, S.; Sasmal, H. S.; Kundu,T.; Kandambeth, S.; Illath, K.; Díaz, D. D.; Banerjee, R. Targeted Drug Delivery in Covalent Organic Nanosheets (CONs) via Sequential Postsynthetic Modification. J. Am. Chem. Soc. 2017, 139, 4513−4520. (d) Karak, S.; Kumar, S.; Pachfule, P.; Banerjee, R. Porosity Prediction through Hydrogen Bonding in Covalent Organic Frameworks. J. Am. Chem. Soc. 2018, 140, 5138−5145. The Stockholm Convention on Persistent Organic Pollutants aimed to restrict or eliminate few persistent organic pollutants (POPs) from the environment that are resistant to degradation by any means. Decabromo biphenyl ether (decaBDE) has been included in the elimination list in the fourth conference in 2009 that has severe impact on animal health. (a) Geim, A. K.; Novoselov, K. S. The rise of grapheme. Nat. Mater. 2007, 6, 183−191. (b) Rao, C. N. R.; Sood, A. K.; Subrahmanyam, K. S.; Govindaraj, A. Graphene: the new twodimensional nanomaterial. Angew. Chem. Int. Ed. 2009, 48, 7752−7777. (c) Geim, A. K. Graphene: Status and Prospects. Science 2009, 324, 1530−1534. (d) Raccichini, R.; Varzi, A.; Passerini, S.; Scrosati, B. The role of graphene for electrochemical energy storage. Nat. Mater. 2015, 14, 271−279. (a) Junggeburth, S. C.; Diehl, L.; Werner, S.; Duppel, V.; Sigle, W.; Lotsch, B. V. Ultrathin 2D coordination polymer nanosheets by surfactant-mediated synthesis. J. Am. Chem. Soc. 2013, 135, 6157−6164. (b) Xu, C.; Zeng, Y.; Rui, X.; Xiao, N.; Zhu, J.; Zhang, W.; Chen, J.; Liu, W.; Tan, H.; Hng, H. H.; Yan, Q. Controlled soft-template synthesis of ultrathin C@ FeS nanosheets with high-Li-storage performance. ACS Nano 2012, 6, 4713−4721. (c) Zaworotko, M. J. Superstructural diversity in two dimensions: crystal engineering of laminated solids Chem.

Page 10 of 18

Commun. 2001, 0, 1. (d) Russell, V. A.; Evans, C. C.; Li, W.; Ward, M. D. Nanoporous molecular sandwiches: pillared twodimensional hydrogen-bonded networks with adjustable porosity. Science 1997, 276, 575−579. 16. (a) Mal, P.; Schultz, D.; Beyeh, K.; Rissanen, K.; Nitschke, J. R. An unlockable–relockable iron cage by subcomponent selfassembly. Angew. Chem. Int. Ed. 2008, 47, 8297−8301. (b) Tanoue, R.; Higuchi, R.; Enoki, N.; Miyasato, Y.; Uemura, S.;Kimizuka, N.; Stieg, A. Z.; Gimzewski, J. K.; Kunitake, M. Thermodynamically controlled self-assembly of covalent nanoarchitectures in aqueous solution. ACS Nano 2011, 5, 3923−3929. (c) Belowich, M. E.; Stoddart, J. F. Dynamic imine chemistry. Chem. Soc. Rev. 2012, 41, 2003−2024. 17. Órfão, J. J. M. Silva, A. I. M.; Pereira, J. C. V.; Barata, S. A.; onseca, I. M. F; Faria, P. C. C.; Pereira, M. F. R. Adsorption of a reactive dye on chemically modifid activated carbons — inflence of pH. J. Colloid Interf. Sci. 2006, 296, 480−489. 18. (a) Putra, E. K.; Pranowo, R.; Sunarso, J.; Indraswati, N.; Ismadji, S. Performance of activated carbon and bentonite for adsorption of amoxicillin from wastewater: mechanisms, isotherms and kinetics. Water Res. 2009, 43, 2419−2430. (b) Wan Ngah, W. S.; Teong, L. C.; Hanafih, M. A. K. M. Adsorption of dyes and heavy metal ions by chitosan composites: a review. Carbohydr. Polym. 2011, 83, 1446−1456. (c) Alsbaiee, A.; Smith, B. J.; Xiao1, L.; Ling,Y.; Helbling, D. E.; Dichtel, W. R. Rapid removal of organic micropollutants from water by a porous β-cyclodextrin polymer. Nature 2016, 529, 190−194. (d) Shetty, D.; Jahovic, I.; Raya, J.; Asfari, Z.; Olsen, J-C.; Trabolsi, A. Porous Polycalix[4]arenes for Fast and Efficient Removal of Organic Micropollutants from Water. ACS Appl. Mater. Interfaces 2018, 10, 2976−2981.

ACS Paragon Plus Environment

Page 11 of 18 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

Journal of the American Chemical Society

Table of contents

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Figure 1 1189x764mm (55 x 55 DPI)

ACS Paragon Plus Environment

Page 12 of 18

Page 13 of 18 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

Journal of the American Chemical Society

Figure 2 414x450mm (93 x 93 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Figure 3 487x812mm (52 x 52 DPI)

ACS Paragon Plus Environment

Page 14 of 18

Page 15 of 18 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

Journal of the American Chemical Society

Figure 4 381x848mm (49 x 49 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Figure 5 958x990mm (42 x 42 DPI)

ACS Paragon Plus Environment

Page 16 of 18

Page 17 of 18 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

Journal of the American Chemical Society

Figure 6 421x443mm (94 x 94 DPI)

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Figure_TOC

ACS Paragon Plus Environment

Page 18 of 18