Covalent organic framework

Last updated

Covalent organic frameworks (COFs) are a class of porous polymers that form two- or three-dimensional structures through reactions between organic precursors resulting in strong, covalent bonds to afford porous, stable, and crystalline materials. COFs emerged as a field from the overarching domain of organic materials as researchers optimized both synthetic control and precursor selection. [1] These improvements to coordination chemistry enabled non-porous and amorphous organic materials such as organic polymers to advance into the construction of porous, crystalline materials with rigid structures that granted exceptional material stability in a wide range of solvents and conditions. [1] [2] Through the development of reticular chemistry, precise synthetic control was achieved and resulted in ordered, nano-porous structures with highly preferential structural orientation and properties which could be synergistically enhanced and amplified. [3] With judicious selection of COF secondary building units (SBUs), or precursors, the final structure could be predetermined, and modified with exceptional control enabling fine-tuning of emergent properties. [4] This level of control facilitates the COF material to be designed, synthesized, and utilized in various applications, many times with metrics on scale or surpassing that of the current state-of-the-art approaches.

Contents

History

While at University of Michigan, Omar M. Yaghi (currently at UCBerkeley) and Adrien P Cote published the first paper of COFs in 2005, reporting a series of 2D COFs. [5] They reported the design and successful synthesis of COFs by condensation reactions of phenyl diboronic acid (C6H4[B(OH)2]2) and hexahydroxytriphenylene (C18H6(OH)6). Powder X-ray diffraction studies of the highly crystalline products having empirical formulas (C3H2BO)6·(C9H12)1 (COF-1) and C9H4BO2 (COF-5) revealed 2-dimensional expanded porous graphitic layers that have either staggered conformation (COF-1) or eclipsed conformation (COF-5). Their crystal structures are entirely held by strong bonds between B, C, and O atoms to form rigid porous architectures with pore sizes ranging from 7 to 27 Angstroms. COF-1 and COF-5 exhibit high thermal stability (to temperatures up to 500 to 600 °C), permanent porosity, and high surface areas (711 and 1590 square meters per gram, respectively). [5]

The synthesis of 3D COFs has been hindered by longstanding practical and conceptual challenges until it was first achieved in 2007 by Omar M. Yaghi and colleagues, which received the Newcomb Cleveland Prize. [6] The research team synthesized and designed the first 3D-COF ever; COF-103 and COF-108, helping unleash this new field. Unlike 0D and 1D systems, which are soluble, the insolubility of 2D and 3D structures precludes the use of stepwise synthesis, making their isolation in crystalline form very difficult. This first challenge, however, was overcome by judiciously choosing building blocks and using reversible condensation reactions to crystallize COFs.

Structure

Porous crystalline solids consist of secondary building units (SBUs) which assemble to form a periodic and porous framework. An almost infinite number of frameworks can be formed through various SBU combinations leading to unique material properties for applications in separations, storage, and heterogeneous catalysis. [7]

Types of porous crystalline solids include zeolites, metal-organic frameworks (MOFs), and covalent organic frameworks (COFs). Zeolites are microporous, aluminosilicate minerals commonly used as commercial adsorbents. MOFs are a class of porous polymeric material, consisting of metal ions linked together by organic bridging ligands and are a new development on the interface between molecular coordination chemistry and materials science. [8]

COFs are another class of porous polymeric materials, consisting of porous, crystalline, covalent bonds that usually have rigid structures, exceptional thermal stabilities (to temperatures up to 600 °C), are stable in water and low densities. They exhibit permanent porosity with specific surface areas surpassing those of well-known zeolites and porous silicates. [5]

Secondary building units

Schematic Figure of Reticular Chemistry. Reticular figure.png
Schematic Figure of Reticular Chemistry.

The term ‘secondary building unit’ has been used for some time to describe conceptual fragments which can be compared as bricks used to build a house of zeolites; in the context of this page it refers to the geometry of the units defined by the points of extension. [9]

Reticular synthesis

Reticular synthesis enables facile bottom-up synthesis of the framework materials to introduce precise perturbations in chemical composition, resulting in the highly controlled tunability of framework properties. [4] [10] [11] Through a bottom-up approach, a material is built from atomic or molecular components synthetically as opposed to a top-down approach, which forms a material from the bulk through approaches such as exfoliation, lithography, or other varieties of post-synthetic modification. [3] [12] The bottom-up approach is especially advantageous with respect to materials such as COFs because the synthetic methods are designed to directly result in an extended, highly crosslinked framework that can be tuned with exceptional control at the nanoscale level. [3] [13] [14] Geometrical and dimensional principles govern the framework's resulting topology as the SBUs combine to form predetermined structures. [15] [16] This level of synthetic control has also been termed "molecular engineering", abiding by the concept termed by Arthur R. von Hippel in 1956. [17]

COF topological control through judicious selection of precursors that result in bonding directionality in the final resulting network. Adapted from Jiang and coworkers' Two- and Three-dimensional Covalent Organic Frameworks (COFs). Topological Control via Reticular Synthesis.jpg
COF topological control through judicious selection of precursors that result in bonding directionality in the final resulting network. Adapted from Jiang and coworkers' Two- and Three-dimensional Covalent Organic Frameworks (COFs).

It has been established in the literature that, when integrated into an isoreticular framework, such as a COF, properties from monomeric compounds can be synergistically enhanced and amplified. [3] COF materials possess the unique ability for bottom-up reticular synthesis to afford robust, tunable frameworks that synergistically enhance the properties of the precursors, which, in turn, offers many advantages in terms of improved performance in different applications. As a result, the COF material is highly modular and tuned efficiently by varying the SBUs’ identity, length, and functionality depending on the desired property change on the framework scale.[ citation needed ] Ergo, there exists the ability to introduce diverse functionality directly into the framework scaffold to allow for a variety of functions which would be cumbersome, if not impossible, to achieve through a top-down method, such as lithographic approaches or chemical-based nanofabrication. Through reticular synthesis, it is possible to molecularly engineer modular, framework materials with highly porous scaffolds that exhibit unique electronic, optical, and magnetic properties while simultaneously integrating desired functionality into the COF skeleton.

Reticular synthesis is different from retrosynthesis of organic compounds, because the structural integrity and rigidity of the building blocks in reticular synthesis remain unaltered throughout the construction process—an important aspect that could help to fully realize the benefits of design in crystalline solid-state frameworks. Similarly, reticular synthesis should be distinguished from supramolecular assembly, because in the former, building blocks are linked by strong bonds throughout the crystal. [9]

Synthetic chemistry

Reversible reactions for COF formation featuring boron to form a variety of linkages (boronate, boroxine, and borazine). Boron-based COF Linkages.jpg
Reversible reactions for COF formation featuring boron to form a variety of linkages (boronate, boroxine, and borazine).

Reticular synthesis was used by Yaghi and coworkers in 2005 to construct the first two COFs reported in the literature: COF-1, using a dehydration reaction of benzenediboronic acid (BDBA), and COF-5, via a condensation reaction between hexahydroxytriphenylene (HHTP) and BDBA. [18] These framework scaffolds were interconnected through the formation of boroxine and boronate linkages, respectively, using solvothermal synthetic methods. [18]

COF linkages

Since Yaghi and coworkers’ seminal work in 2005, COF synthesis has expanded to include a wide range of organic connectivity such as boron-, nitrogen-, other atom-containing linkages. [2] [19] [20] [21] The linkages in the figures shown are not comprehensive as other COF linkages exist in the literature, especially for the formation of 3D COFs.

Skeletal structure of COF-1 consisting of phenyl rings joined by boroxine rings, synthesized by a condensation reaction of phenyldiboronic acid. Boron condensation.png
Skeletal structure of COF-1 consisting of phenyl rings joined by boroxine rings, synthesized by a condensation reaction of phenyldiboronic acid.

Boron condensation

The most popular COF synthesis route is a boron condensation reaction which is a molecular dehydration reaction between boronic acids. In case of COF-1, three boronic acid molecules converge to form a planar six-membered B3O3 (boroxine) ring with the elimination of three water molecules. [5]

Reversible reactions for COF formation featuring nitrogen to form a variety of linkages (imine, hydrazone, azine, squaraine, phenazine, imide, triazine). Nitrogen-based COF Linkages.jpg
Reversible reactions for COF formation featuring nitrogen to form a variety of linkages (imine, hydrazone, azine, squaraine, phenazine, imide, triazine).

Triazine based trimerization

Formation of CTF-1 COF featuring triazine linkages. Triazine trimerization vector.svg
Formation of CTF-1 COF featuring triazine linkages.

Another class of high performance polymer frameworks with regular porosity and high surface area is based on triazine materials which can be achieved by dynamic trimerization reaction of simple, cheap, and abundant aromatic nitriles in ionothermal conditions (molten zinc chloride at high temperature (400 °C)). CTF-1 is a good example of this chemistry. [22]

Imine condensation

A structural representation of the TpOMe-DAQ COF A chemical structure of the TpOMe-DAQ covalent organic framework.png
A structural representation of the TpOMe-DAQ COF
Reversible reactions for COF formation featuring a variety of atoms to form different linkages (a double stage connecting boronate ester and imine linkages, alkene, silicate, nitroso). COF Linkages.jpg
Reversible reactions for COF formation featuring a variety of atoms to form different linkages (a double stage connecting boronate ester and imine linkages, alkene, silicate, nitroso).

The imine condensation reaction which eliminates water (exemplified by reacting aniline with benzaldehyde using an acid catalyst) can be used as a synthetic route to reach a new class of COFs. The 3D COF called COF-300 [23] and the 2D COF named TpOMe-DAQ [24] are good examples of this chemistry. When 1,3,5-triformylphloroglucinol (TFP) is used as one of the SBUs, two complementary tautomerizations occur (an enol to keto and an imine to enamine) which result in a β-ketoenamine moiety [25] as depicted in the DAAQ-TFP [26] framework. Both DAAQ-TFP and TpOMe-DAQ COFs are stable in acidic aqueous conditions and contain the redox active linker 2,6-diaminoanthroquinone which enables these materials to reversibly store and release electrons within a characteristic potential window. [24] [26] Consequently, both of these COFs have been investigated as electrode materials for potential use in supercapacitors. [24] [26]

A structural representation of the DAAQ-TFP COF A chemical structure of the DAAQ-TFP covalent organic framework.png
A structural representation of the DAAQ-TFP COF

Solvothermal synthesis

The solvothermal approach is the most common used in the literature but typically requires long reaction times due to the insolubility of the organic SBUs in nonorganic media and the time necessary to reach thermodynamic COF products. [27]

Templated synthesis

Morphological control on the nanoscale is still limited as COFs lack synthetic control in higher dimensions due to the lack of dynamic chemistry during synthesis. To date, researchers have attempted to establish better control through different synthetic methods such as solvothermal synthesis, interface-assisted synthesis, solid templation as well as seeded growth.[ citation needed ] [28] [29] First one of the precursors is deposited onto the solid support followed by the introduction of the second precursor in vapor form. This results in the deposition of the COF as a thin film on the solid support. [30]

Properties

Porosity

A defining advantage of COFs is the exceptional porosity that results from the substitution of analogous SBUs of varying sizes. Pore sizes range from 7-23 Å and feature a diverse range of shapes and dimensionalities that remain stable during the evacuation of solvent. [14] The rigid scaffold of the COF structure enables the material to be evacuated of solvent and retain its structure, resulting in high surface areas as seen by the Brunauer–Emmett–Teller analysis. [31] This high surface area to volume ratio and incredible stability enables the COF structure to serve as exceptional materials for gas storage and separation.

Crystallinity

There are several COF single crystals synthesized to date. [32] There are a variety of techniques employed to improve crystallinity of COFs. The use of modulators, monofunctional version of precursors, serve to slow the COF formation to allow for more favorable balance between kinetic and thermodynamic control, hereby enabling crystalline growth. This was employed by Yaghi and coworkers for 3D imine-based COFs (COF-300, COF 303, LZU-79, and LZU-111). [32] However, the vast majority of COFs are not able to crystallize into single crystals but instead are insoluble powders. The improvement of crystallinity of these polycrystalline materials can be improved through tuning the reversibility of the linkage formation to allow for corrective particle growth and self-healing of defects that arise during COF formation. [33]

Conductivity

In a fully conjugated 2D COF material such as those synthesized from metallophthalocyanines and highly conjugated organic linkers, charge transport is increased both in-plane, as well as through the stacks, resulting in increased conductivity. 2D COF conductivity.jpg
In a fully conjugated 2D COF material such as those synthesized from metallophthalocyanines and highly conjugated organic linkers, charge transport is increased both in-plane, as well as through the stacks, resulting in increased conductivity.

Integration of SBUs into a covalent framework results in the synergistic emergence of conductivities much greater than the monomeric values. The nature of the SBUs can improve conductivity. Through the use of highly conjugated linkers throughout the COF scaffold, the material can be engineered to be fully conjugated, enabling high charge carrier density as well as through- and in-plane charge transport. For instance, Mirica and coworkers synthesized a COF material (NiPc-Pyr COF) from nickel phthalocyanine (NiPc) and pyrene organic linkers that had a conductivity of 2.51 x 10−3 S/m, which was several orders of magnitude larger than the undoped molecular NiPc, 10−11 S/m. [34] A similar COF structure made by Jiang and coworkers, CoPc-Pyr COF, exhibited a conductivity of 3.69 x 10−3 S/m. [35] In both previously mentioned COFs, the 2D lattice allows for full π-conjugation in the x and y directions as well as π-conduction along the z axis due to the fully conjugated, aromatic scaffold and π-π stacking, respectively. [34] [35] Emergent electrical conductivity in COF structures is especially important for applications such as catalysis and energy storage where quick and efficient charge transport is required for optimal performance.

Characterization

There exists a wide range of characterization methods for COF materials. There are several COF single crystals synthesized to date. For these highly crystalline materials, X-ray diffraction (XRD) is a powerful tool capable of determining COF crystal structure. [36] The majority of COF materials suffer from decreased crystallinity so powder X-ray diffraction (PXRD) is used. In conjunction with simulated powder packing models, PXRD can determine COF crystal structure.[ citation needed ]

In order to verify and analyze COF linkage formation, various techniques can be employed such as infrared (IR) spectroscopy, and nuclear magnetic resonance (NMR) spectroscopy. [36] Precursor and COF IR spectra enables comparison between vibrational peaks to ascertain that certain key bonds present in the COF linkages appear and that peaks of precursor functional groups disappear. In addition, solid-state NMR enables probing of linkage formation as well and is well suited for large, insoluble materials like COFs. Gas adsorption-desorption studies quantify the porosity of the material via calculation of the Brunauer–Emmett–Teller (BET) surface area and pore diameter from gas adsorption isotherms. [36] Electron imagine techniques such as scanning electron microscope (SEM), and transmission electron microscopy (TEM) can resolve surface structure and morphology, and microstructural information, respectively. [36] Scanning tunneling microscope (STM) and atomic force microscopy (AFM) have also been used to characterize COF microstructural information as well. [36] Additionally, methods like X-ray photoelectron spectroscopy (XPS), inductively coupled plasma mass spectrometry (ICP-MS), and combustion analysis can be used to identify elemental composition and ratios. [36]

Applications

Gas storage and separation

Due to the exceptional porosity of COFs, they have been used extensively in the storage and separation of gases such as hydrogen, methane, etc.

Hydrogen storage

Omar M. Yaghi and William A. Goddard III reported COFs as exceptional hydrogen storage materials. They predicted the highest excess H2 uptakes at 77 K are 10.0 wt % at 80 bar for COF-105, and 10.0 wt % at 100 bar for COF-108, which have higher surface area and free volume, by grand canonical Monte Carlo (GCMC) simulations as a function of temperature and pressure. This is the highest value reported for associative H2 storage of any material. Thus 3D COFs are most promising new candidates in the quest for practical H2 storage materials. [37] In 2012, the lab of William A. Goddard III reported the uptake for COF102, COF103, and COF202 at 298 K and they also proposed new strategies to obtain higher interaction with H2. Such strategy consists of metalating the COF with alkali metals such as Li. [38] These complexes composed of Li, Na and K with benzene ligands (such as 1,3,5-benzenetribenzoate, the ligand used in MOF-177) have been synthesized by Krieck et al. [39] and Goddard showed that the THF is important to their stability. If the metalation with alkali meals is performed in the COFs, Goddard et al. calculated that some COFs can reach 2010 DOE gravimetric target in delivery units at 298 K of 4.5 wt %: COF102-Li (5.16 wt %), COF103-Li (4.75 wt %), COF102-Na (4.75 wt %) and COF103-Na (4.72 wt %). COFs also perform better in delivery units than MOFs because the best volumetric performance is for COF102-Na (24.9), COF102-Li (23.8), COF103-Na (22.8), and COF103-Li (21.7), all using delivery g H2/L units for 1–100 bar. These are the highest gravimetric molecular hydrogen uptakes for a porous material under these thermodynamic conditions.

Methane storage

Omar M. Yaghi and William A. Goddard III also reported COFs as exceptional methane storage materials. The best COF in terms of total volume of CH4 per unit volume COF adsorbent is COF-1, which can store 195 v/v at 298 K and 30 bar, exceeding the U.S. Department of Energy target for CH4 storage of 180 v/v at 298 K and 35 bar. The best COFs on a delivery amount basis (volume adsorbed from 5 to 100 bar) are COF-102 and COF-103 with values of 230 and 234 v(STP: 298 K, 1.01 bar)/v, respectively, making these promising materials for practical methane storage. More recently, new COFs with better delivery amount have been designed in the lab of William A. Goddard III, and they have been shown to be stable and overcome the DOE target in delivery basis. COF-103-Eth-trans and COF-102-Ant, are found to exceed the DOE target of 180 v(STP)/v at 35 bar for methane storage. They reported that using thin vinyl bridging groups aids performance by minimizing the interaction methane-COF at low pressure.

Gas separation

In addition to storage, COF materials are exceptional at gas separation. For instance, COFs like imine-linked COF LZU1 and azine-linked COF ACOF-1 were used as a bilayer membrane for the selective separation of the following mixtures: H2/CO2, H2/N2, and H2/CH4. [40] The COFs outperformed molecular sieves due to the inherent thermal and operational stability of the structures. [40] It has also been shown that COFs inherently act as adsorbents, adhering to the gaseous molecules to enable storage and separation. [41]

Optical properties

A highly ordered π-conjugation TP-COF, consisting of pyrene and triphenylene functionalities alternately linked in a mesoporous hexagonal skeleton, is highly luminescent, harvests a wide wavelength range of photons, and allows energy transfer and migration. Furthermore, TP-COF is electrically conductive and capable of repetitive on–off current switching at room temperature. [42]

Porosity/surface-area effects

Most studies to date have focused on the development of synthetic methodologies with the aim of maximizing pore size and surface area for gas storage. That means the functions of COFs have not yet been well explored, but COFs can be used as catalysts, [43] or for gas separation, etc. [5]

Carbon capture

In 2015 the use of highly porous, catalyst-decorated COFs for converting carbon dioxide into carbon monoxide was reported. [44] MOF under solvent-free conditions can also be used for catalytic activity in the cycloaddition of CO2 and epoxides into cyclic organic carbonates with enhanced catalyst recyclability. [45]

Sensing

Due to defining molecule-framework interactions, COFs can be used as chemical sensors in a wide range of environments and applications. Properties of the COF change when their functionalities interact with various analytes enabling the materials to serve as devices in various conditions: as chemiresistive sensors, [34] as well as electrochemical sensors for small molecules. [46]

Catalysis

Due to the ability to introduce diverse functionality into COFs’ structure, catalytic sites can be fine-tuned in conjunction with other advantageous properties like conductivity and stability to afford efficient and selective catalysts. COFs have been used as heterogeneous catalysts in organic, [47] electrochemical, [35] [48] as well as photochemical reactions. [27]

Electrocatalysis

COFs have been studied as non-metallic electrocatalysts for energy-related catalysis, including carbon dioxide electro-reduction and water splitting reaction. [49] However, such researches are still in the very early stage. Most of the efforts have been focusing on solving the key issues, such as conductivity, [50] stability in electrochemical processes. [51]

Energy storage

A few COFs possess the stability and conductivity necessary to perform well in energy storage applications like lithium-ion batteries, [52] [53] and various different metal-ion batteries and cathodes. [54] [55]

Water filtration

A prototype 2 nanometer thick COF layer on a graphene substrate was used to filter dye from industrial wastewater. Once full, the COF can be cleaned and reused. [56]

Pharmaceutical drug delivery

A 3D COF was created, characterised by an interconnected mesoporous scaffold that showed effective drug loading and release in a simulated body fluid environment, making it useful as a nanocarrier for pharmaceutical drugs. [57]

See also

Related Research Articles

Dynamic covalent chemistry (DCvC) is a synthetic strategy employed by chemists to make complex molecular and supramolecular assemblies from discrete molecular building blocks. DCvC has allowed access to complex assemblies such as covalent organic frameworks, molecular knots, polymers, and novel macrocycles. Not to be confused with dynamic combinatorial chemistry, DCvC concerns only covalent bonding interactions. As such, it only encompasses a subset of supramolecular chemistries.

<span class="mw-page-title-main">Metal–organic framework</span> Class of chemical substance

Metal–organic frameworks (MOFs) are a class of porous polymers consisting of metal clusters coordinated to organic ligands to form one-, two-, or three-dimensional structures. The organic ligands included are sometimes referred to as "struts" or "linkers", one example being 1,4-benzenedicarboxylic acid (BDC).

<span class="mw-page-title-main">Zeolitic imidazolate framework</span>

Zeolitic imidazolate frameworks (ZIFs) are a class of metal-organic frameworks (MOFs) that are topologically isomorphic with zeolites. ZIF glasses can be synthesized by the melt-quench method, and the first melt-quenched ZIF glass was firstly made and reported by Bennett et al. back in 2015. ZIFs are composed of tetrahedrally-coordinated transition metal ions connected by imidazolate linkers. Since the metal-imidazole-metal angle is similar to the 145° Si-O-Si angle in zeolites, ZIFs have zeolite-like topologies. As of 2010, 105 ZIF topologies have been reported in the literature. Due to their robust porosity, resistance to thermal changes, and chemical stability, ZIFs are being investigated for applications such as carbon dioxide capture.

<span class="mw-page-title-main">Omar M. Yaghi</span> Chemist

Omar M. Yaghi is the James and Neeltje Tretter Chair Professor of Chemistry at the University of California, Berkeley, an affiliate scientist at Lawrence Berkeley National Laboratory, the Founding Director of the Berkeley Global Science Institute, and an elected member of the US National Academy of Sciences as well as the German National Academy of Sciences Leopoldina.

<span class="mw-page-title-main">Solvothermal synthesis</span>

Solvothermal synthesis is a method of producing chemical compounds, in which a solvent containing reagents is put under high pressure and temperature in an autoclave. Many substances dissolve better in the same solvent in such conditions than at standard conditions, enabling reactions that would not otherwise occur and leading to new compounds or polymorphs. Solvothermal synthesis is very similar to the hydrothermal route; both are typically conducted in a stainless steel autoclave. The only difference being that the precursor solution is usually non-aqueous.

<span class="mw-page-title-main">Periodic graph (crystallography)</span>

In crystallography, a periodic graph or crystal net is a three-dimensional periodic graph, i.e., a three-dimensional Euclidean graph whose vertices or nodes are points in three-dimensional Euclidean space, and whose edges are line segments connecting pairs of vertices, periodic in three linearly independent axial directions. There is usually an implicit assumption that the set of vertices are uniformly discrete, i.e., that there is a fixed minimum distance between any two vertices. The vertices may represent positions of atoms or complexes or clusters of atoms such as single-metal ions, molecular building blocks, or secondary building units, while each edge represents a chemical bond or a polymeric ligand.

<span class="mw-page-title-main">Two-dimensional polymer</span>

A two-dimensional polymer (2DP) is a sheet-like monomolecular macromolecule consisting of laterally connected repeat units with end groups along all edges. This recent definition of 2DP is based on Hermann Staudinger's polymer concept from the 1920s. According to this, covalent long chain molecules ("Makromoleküle") do exist and are composed of a sequence of linearly connected repeat units and end groups at both termini.

<span class="mw-page-title-main">Hong-Cai (Joe) Zhou</span> Chinese–American chemist and academic (born c. 1964)

Hong-Cai (Joe) Zhou is a Chinese–American chemist and academic. He is the Davidson Professor of Science and Robert A. Welch Chair in Chemistry at Texas A&M University. He is the associate editor of the journal Inorganic Chemistry.

<span class="mw-page-title-main">Mircea Dincă</span> Romanian-American inorganic chemist

Mircea Dincă is a Romanian-American inorganic chemist. He is a Professor of Chemistry and W. M. Keck Professor of Energy at the Massachusetts Institute of Technology (MIT). At MIT, Dincă leads a research group that focuses on the synthesis of functional metal-organic frameworks (MOFs), which possess conductive, catalytic, and other material-favorable properties.

Rahul Banerjee is a Bengali Indian organic chemist and a professor at the department of chemical sciences of the Indian Institute of Science Education and Research, Kolkata. Banerjee, a fellow of the Royal Society of Chemistry, is known for his studies in the field of Metal–organic framework designing. The Council of Scientific and Industrial Research, the apex agency of the Government of India for scientific research, awarded him the Shanti Swarup Bhatnagar Prize for Science and Technology, one of the highest Indian science awards, for his contributions to chemical sciences in 2018. Currently he is one of the associate editor of international peer-review journal Journal of the American Chemical Society.

Natalia B. Shustova is a Peter and Bonnie McCausland Professor of Chemistry at the University of South Carolina. She focuses on developing materials for sustainable energy conversion, metal-organic frameworks (MOFs), covalent organic frameworks (COFs), and graphitic supramolecular structures.

Jeffrey R. Long is a professor of chemistry at University of California, Berkeley known for his work in metal−organic frameworks and molecular magnetism. He was elected to the American Academy of Arts and Sciences in 2019 and is the 2019 F. Albert Cotton Award recipient. His research interests include: synthesis of inorganic clusters and porous materials, investigating the electronic and magnetic properties of inorganic materials; metal-organic frameworks, and gas storage/capture.

<span class="mw-page-title-main">HKUST-1</span>

HKUST-1, which is also called MOF-199, is a material in the class of metal-organic frameworks (MOFs). Metal-organic frameworks are crystalline materials, in which metals are linked by ligands to form repeating coordination motives extending in three dimensions. The HKUST-1 framework is built up of dimeric metal units, which are connected by benzene-1,3,5-tricarboxylate linker molecules. The paddlewheel unit is the commonly used structural motif to describe the coordination environment of the metal centers and also called secondary building unit (SBU) of the HKUST-1 structure. The paddlewheel is built up of four benzene-1,3,5-tricarboxylate linkers molecules, which bridge two metal centers. One water molecules is coordinated to each of the two metal centers at the axial position of the paddlewheel unit in the hydrated state, which is usually found if the material is handled in air. After an activation process, these water molecules can be removed and the coordination site at the metal atoms is left unoccupied. This unoccupied coordination site is called coordinatively unsaturated site (CUS) and can be accessed by other molecules.

<span class="mw-page-title-main">Wendy Lee Queen</span> American chemist and material scientist

Wendy Lee Queen is an American chemist and material scientist. Her research interest focus on development design and production of hybrid organic/inorganic materials at the intersection of chemistry, chemical engineering and material sciences. As of 2020 she is a tenure-track assistant professor at the École polytechnique fédérale de Lausanne (EPFL) in Switzerland, where she directs the Laboratory for Functional Inorganic Materials.

2,6-Diformylpyridine is an organic compound with the formula C5H3N(CHO)2, and typically appears as a solid powder at room temperature. The molecule features formyl groups adjacent to the nitrogen of pyridine. The compound is prepared by oxidation of 2,6-dimethylpyridine.

<span class="mw-page-title-main">Conductive metal−organic frameworks</span>

Conductive metal−organic frameworks are a class of metal–organic frameworks with intrinsic ability of electronic conduction. Metal ions and organic linker self-assemble to form a framework which can be 1D/2D/3D in connectivity. The first conductive MOF, Cu[Cu(2,3-pyrazinedithiol)2] was described in 2009 and exhibited electrical conductivity of 6 × 10−4 S cm−1 at 300 K.

Song Lin is a Chinese-American organic electrochemist who is an associate professor at Cornell University. His research involves the development of new synthetic organic methodologies that utilize electrochemistry to forge new chemical bonds. He is an Associate Editor of the journal Organic Letters, and serves on the Early Career Advisory Board of Chemistry - A European Journal. He was named by Chemical & Engineering News as one of their Trailblazers of 2022, a feature highlighting LGBTQ+ chemists in academia.

Carboxylate–based metal–organic frameworks are metal–organic frameworks that are based on organic molecules comprising carboxylate functional groups.

<span class="mw-page-title-main">Hydrogen-bonded organic framework</span>

Hydrogen-bonded organic frameworks (HOFs) are a class of porous polymers formed by hydrogen bonds among molecular monomer units to afford porosity and structural flexibility. There are diverse hydrogen bonding pair choices that could be used in HOFs construction, including identical or nonidentical hydrogen bonding donors and acceptors. For organic groups acting as hydrogen bonding units, species like carboxylic acid, amide, 2,4-diaminotriazine, and imidazole, etc., are commonly used for the formation of hydrogen bonding interaction. Compared with other organic frameworks, like COF and MOF, the binding force of HOFs is relatively weaker, and the activation of HOFs is more difficult than other frameworks, while the reversibility of hydrogen bonds guarantees a high crystallinity of the materials. Though the stability and pore size expansion of HOFs has potential problems, HOFs still show strong potential for applications in different areas.

Porous polymers are a class of porous media materials in which monomers form 2D and 3D polymers containing angstrom- to nanometer-scale pores formed by the arrangement of the monomers. They may be either crystalline or amorphous. Subclasses include covalent organic frameworks (COFs), hydrogen-bonded organic frameworks (HOFs), metal-organic frameworks (MOFs), and porous organic polymers (POPs). The subfield of chemistry specializing in porous polymers is called reticular chemistry.

References

  1. 1 2 Ding, San-Yuan; Wang, Wei (2013). "Covalent organic frameworks (COFs): from design to applications". Chem. Soc. Rev. 42 (2): 548–568. doi:10.1039/C2CS35072F. ISSN   0306-0012. PMID   23060270.
  2. 1 2 Huang, Ning; Wang, Ping; Jiang, Donglin (2016-09-20). "Covalent organic frameworks: a materials platform for structural and functional designs". Nature Reviews Materials. 1 (10): 16068. Bibcode:2016NatRM...116068H. doi:10.1038/natrevmats.2016.68. ISSN   2058-8437. S2CID   138892338.
  3. 1 2 3 4 Aykanat, Aylin; Meng, Zheng; Benedetto, Georganna; Mirica, Katherine A. (2020-07-14). "Molecular Engineering of Multifunctional Metallophthalocyanine-Containing Framework Materials". Chemistry of Materials. 32 (13): 5372–5409. doi:10.1021/acs.chemmater.9b05289. ISSN   0897-4756. S2CID   225664378.
  4. 1 2 Feng, Xiao; Ding, Xuesong; Jiang, Donglin (2012). "Covalent organic frameworks". Chemical Society Reviews. 41 (18): 6010–22. doi:10.1039/c2cs35157a. ISSN   0306-0012. PMID   22821129.
  5. 1 2 3 4 5 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, pp 1166-1170. doi : 10.1126/science.1120411
  6. El-Kaderi, H. M.; Hunt, J. R.; Mendoza-Cortes, J. L.; Cote, A. P.; Taylor, R. E.; O'Keeffe, M.; Yaghi, O. M. (2007). "Designed Synthesis of 3D Covalent Organic Frameworks". Science. 316 (5822): 268–272. Bibcode:2007Sci...316..268E. doi:10.1126/science.1139915. PMID   17431178. S2CID   19555677.
  7. Kitagawa, S.; Kitaura, R.; Noro, S.; Functional Porous Coordination Polymers. Angew. Chem. Int. Ed.2004, 43, pp 2334-2375. doi : 10.1002/anie.200300610
  8. James, S. L.; Metal-organic frameworks. Chem. Soc. Rev.2003, 32, pp 276-288. doi : 10.1039/B200393G
  9. 1 2 Yaghi, O. M.; O'Keeffe, M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J.; Reticular synthesis and the design of new materials. Nature. 2003, 423, pp 705-714. doi : 10.1038/nature01650
  10. Yaghi, Omar M. (2016-12-07). "Reticular Chemistry—Construction, Properties, and Precision Reactions of Frameworks". Journal of the American Chemical Society. 138 (48): 15507–15509. doi: 10.1021/jacs.6b11821 . ISSN   0002-7863. PMID   27934016.
  11. Yaghi, Omar M.; O'Keeffe, Michael; Ockwig, Nathan W.; Chae, Hee K.; Eddaoudi, Mohamed; Kim, Jaheon (2003-06-12). "Reticular synthesis and the design of new materials". Nature. 423 (6941): 705–714. doi:10.1038/nature01650. hdl: 2027.42/62718 . ISSN   0028-0836. PMID   12802325. S2CID   4300639.
  12. Yu, Hai-Dong; Regulacio, Michelle D.; Ye, Enyi; Han, Ming-Yong (2013). "Chemical routes to top-down nanofabrication". Chemical Society Reviews. 42 (14): 6006–18. doi:10.1039/c3cs60113g. ISSN   0306-0012. PMID   23653019.
  13. Teo, Boon K.; Sun, X. H. (2006-12-05). "From Top-Down to Bottom-Up to Hybrid Nanotechnologies: Road to Nanodevices". Journal of Cluster Science. 17 (4): 529–540. doi:10.1007/s10876-006-0086-5. ISSN   1040-7278. S2CID   98710293.
  14. 1 2 Feng, Liang; Wang, Kun-Yu; Lv, Xiu-Liang; Yan, Tian-Hao; Li, Jian-Rong; Zhou, Hong-Cai (2020-02-12). "Modular Total Synthesis in Reticular Chemistry". Journal of the American Chemical Society. 142 (6): 3069–3076. doi:10.1021/jacs.9b12408. ISSN   0002-7863. PMID   31971790. S2CID   210882977.
  15. 1 2 Chen, Q.; Dalapati, S.; Jiang, D. (2017), "Two- and Three-dimensional Covalent Organic Frameworks (COFs)", Comprehensive Supramolecular Chemistry II, Elsevier, pp. 271–290, doi:10.1016/b978-0-12-409547-2.12608-3, ISBN   978-0-12-803199-5 , retrieved 2021-03-01
  16. Zhang, Yue-Biao; Li, Qiaowei; Deng, Hexiang (2021-11-28). "Reticular chemistry at the atomic, molecular, and framework scales". Nano Research. 14 (2): 335–337. Bibcode:2021NaRes..14..335Z. doi: 10.1007/s12274-020-3226-6 . ISSN   1998-0124.
  17. von Hippel, A. (1956-02-24). "Molecular Engineering". Science. 123 (3191): 315–317. Bibcode:1956Sci...123..315V. doi:10.1126/science.123.3191.315. ISSN   0036-8075. PMID   17774519.
  18. 1 2 Cote, A. P. (2005-11-18). "Porous, Crystalline, Covalent Organic Frameworks". Science. 310 (5751): 1166–1170. Bibcode:2005Sci...310.1166C. doi:10.1126/science.1120411. ISSN   0036-8075. PMID   16293756. S2CID   35798005.
  19. Yusran, Yusran; Li, Hui; Guan, Xinyu; Fang, Qianrong; Qiu, Shilun (June 2020). "Covalent Organic Frameworks for Catalysis". EnergyChem. 2 (3): 100035. doi:10.1016/j.enchem.2020.100035. S2CID   219459194.
  20. Jackson, Karl T.; Rabbani, Mohammad G.; Reich, Thomas E.; El-Kaderi, Hani M. (2011). "Synthesis of highly porous borazine-linked polymers and their application to H2, CO2, and CH4 storage". Polymer Chemistry. 2 (12): 2775. doi:10.1039/c1py00374g. ISSN   1759-9954.
  21. Dogru, Mirjam; Bein, Thomas (2014). "On the road towards electroactive covalent organic frameworks". Chem. Commun. 50 (42): 5531–5546. doi: 10.1039/C3CC46767H . ISSN   1359-7345. PMID   24667827.
  22. Kuhn, P.; Antonietti, M.; Thomas, A.; Porous, Covalent Triazine-Based Frameworks Prepared by Ionothermal Synthesis. Angew. Chem. Int. Ed.2008. 47, pp 3450-3453. PMID   18330878
  23. Uribe-Romo, F. J.; Hunt, J. R.; Furukawa, H.; Klck, C.; O'Keeffe, M.; Yaghi, O. M.; A Crystalline Imine-Linked 3-D Porous Covalent Organic Framework. J. Am. Chem. Soc. 2009, 131, pp 4570-4571. doi : 10.1021/ja8096256
  24. 1 2 3 Halder, Arjun; Ghosh, Meena; Khayum M, Abdul; Bera, Saibal; Addicoat, Matthew; Sasmal, Himadri Sekhar; Karak, Suvendu; Kurungot, Sreekumar; Banerjee, Rahul (2018-09-05). "Interlayer Hydrogen-Bonded Covalent Organic Frameworks as High-Performance Supercapacitors". Journal of the American Chemical Society. 140 (35): 10941–10945. doi:10.1021/jacs.8b06460. ISSN   0002-7863. PMID   30132332. S2CID   207193051.
  25. Kandambeth, Sharath; Mallick, Arijit; Lukose, Binit; Mane, Manoj V.; Heine, Thomas; Banerjee, Rahul (2012-12-05). "Construction of Crystalline 2D Covalent Organic Frameworks with Remarkable Chemical (Acid/Base) Stability via a Combined Reversible and Irreversible Route". Journal of the American Chemical Society. 134 (48): 19524–19527. doi:10.1021/ja308278w. ISSN   0002-7863. PMID   23153356.
  26. 1 2 3 DeBlase, Catherine R.; Silberstein, Katharine E.; Truong, Thanh-Tam; Abruña, Héctor D.; Dichtel, William R. (2013-11-13). "β-Ketoenamine-Linked Covalent Organic Frameworks Capable of Pseudocapacitive Energy Storage". Journal of the American Chemical Society. 135 (45): 16821–16824. doi:10.1021/ja409421d. ISSN   0002-7863. PMID   24147596.
  27. 1 2 Sharma, Rakesh Kumar; Yadav, Priya; Yadav, Manavi; Gupta, Radhika; Rana, Pooja; Srivastava, Anju; Zbořil, Radek; Varma, Rajender S.; Antonietti, Markus; Gawande, Manoj B. (2020). "Recent development of covalent organic frameworks (COFs): synthesis and catalytic (organic-electro-photo) applications". Materials Horizons. 7 (2): 411–454. doi:10.1039/C9MH00856J. ISSN   2051-6347. S2CID   204292382.
  28. Allendorf, Mark D.; Dong, Renhao; Feng, Xinliang; Kaskel, Stefan; Matoga, Dariusz; Stavila, Vitalie (2020-08-26). "Electronic Devices Using Open Framework Materials". Chemical Reviews. 120 (16): 8581–8640. doi:10.1021/acs.chemrev.0c00033. ISSN   0009-2665. PMID   32692163. S2CID   220670221.
  29. Evans, Austin M.; Parent, Lucas R.; Flanders, Nathan C.; Bisbey, Ryan P.; Vitaku, Edon; Kirschner, Matthew S.; Schaller, Richard D.; Chen, Lin X.; Gianneschi, Nathan C.; Dichtel, William R. (2018-07-06). "Seeded growth of single-crystal two-dimensional covalent organic frameworks". Science. 361 (6397): 52–57. Bibcode:2018Sci...361...52E. doi: 10.1126/science.aar7883 . ISSN   0036-8075. PMID   29930093.
  30. Chen, T.; Wang, D. (2018), "Synthesis of 2D Covalent Organic Frameworks at the Solid–Vapor Interface", Encyclopedia of Interfacial Chemistry, Elsevier, pp. 446–452, doi:10.1016/b978-0-12-409547-2.13071-9, ISBN   978-0-12-809894-3 , retrieved 2021-03-01
  31. Ben, Teng; Ren, Hao; Ma, Shengqian; Cao, Dapeng; Lan, Jianhui; Jing, Xiaofei; Wang, Wenchuan; Xu, Jun; Deng, Feng; Simmons, Jason M.; Qiu, Shilun (2009-12-07). "Targeted Synthesis of a Porous Aromatic Framework with High Stability and Exceptionally High Surface Area". Angewandte Chemie International Edition. 48 (50): 9457–9460. doi:10.1002/anie.200904637. PMID   19921728.
  32. 1 2 Ma, Tianqiong; Kapustin, Eugene A.; Yin, Shawn X.; Liang, Lin; Zhou, Zhengyang; Niu, Jing; Li, Li-Hua; Wang, Yingying; Su, Jie; Li, Jian; Wang, Xiaoge (2018-07-06). "Single-crystal x-ray diffraction structures of covalent organic frameworks". Science. 361 (6397): 48–52. Bibcode:2018Sci...361...48M. doi: 10.1126/science.aat7679 . ISSN   0036-8075. PMID   29976818.
  33. Haase, Frederik; Lotsch, Bettina V. (2020). "Solving the COF trilemma: towards crystalline, stable and functional covalent organic frameworks". Chemical Society Reviews. 49 (23): 8469–8500. doi: 10.1039/D0CS01027H . ISSN   0306-0012. PMID   33155009.
  34. 1 2 3 Meng, Zheng; Stolz, Robert M.; Mirica, Katherine A. (2019-07-31). "Two-Dimensional Chemiresistive Covalent Organic Framework with High Intrinsic Conductivity". Journal of the American Chemical Society. 141 (30): 11929–11937. doi:10.1021/jacs.9b03441. ISSN   0002-7863. PMID   31241936. S2CID   195694903.
  35. 1 2 3 Huang, Ning; Lee, Ka Hung; Yue, Yan; Xu, Xiaoyi; Irle, Stefan; Jiang, Qiuhong; Jiang, Donglin (2020-09-14). "A Stable and Conductive Metallophthalocyanine Framework for Electrocatalytic Carbon Dioxide Reduction in Water". Angewandte Chemie International Edition. 59 (38): 16587–16593. doi:10.1002/anie.202005274. ISSN   1433-7851. PMID   32436331. S2CID   218765357.
  36. 1 2 3 4 5 6 Guo, Hao; Zhang, Longwen; Xue, Rui; Ma, Baolong; Yang, Wu (2019-03-26). "Eyes of covalent organic frameworks: cooperation between analytical chemistry and COFs". Reviews in Analytical Chemistry. 38 (1). doi: 10.1515/revac-2017-0023 . ISSN   2191-0189.
  37. Han, S.; Hurukawa, H.; Yaghi, O. M.; Goddard, W. A.; Covalent Organic Frameworks as Exceptional Hydrogen Storage Materials. J. Am. Chem. Soc.2008, 130, pp 11580–11581. doi : 10.1021/ja803247y
  38. Mendoza-Cortés, José L.; Han, Sang Soo; Goddard, William A. (2012-02-16). "High H 2 Uptake in Li-, Na-, K-Metalated Covalent Organic Frameworks and Metal Organic Frameworks at 298 K". The Journal of Physical Chemistry A. 116 (6): 1621–1631. Bibcode:2012JPCA..116.1621M. doi:10.1021/jp206981d. ISSN   1089-5639. PMID   22188543.
  39. Krieck, S.; Gorls, H.; Westerhausen, M., Alkali Metal-Stabilized 1,3,5-Triphenylbenzene Monoanions: Synthesis and Characterization of the Lithium, Sodium, and Potassium Complexes. Organometallics. 2010, 29, pp 6790–6800. doi : 10.1021/om1009632
  40. 1 2 Fan, Hongwei; Mundstock, Alexander; Feldhoff, Armin; Knebel, Alexander; Gu, Jiahui; Meng, Hong; Caro, Jürgen (2018-08-15). "Covalent Organic Framework–Covalent Organic Framework Bilayer Membranes for Highly Selective Gas Separation". Journal of the American Chemical Society. 140 (32): 10094–10098. doi:10.1021/jacs.8b05136. ISSN   0002-7863. PMID   30021065. S2CID   51696424.
  41. Fenton, Julie L.; Burke, David W.; Qian, Dingwen; Olvera de la Cruz, Monica; Dichtel, William R. (2021-01-27). "Polycrystalline Covalent Organic Framework Films Act as Adsorbents, Not Membranes". Journal of the American Chemical Society. 143 (3): 1466–1473. doi:10.1021/jacs.0c11159. ISSN   0002-7863. PMID   33438399. S2CID   231596406.
  42. Shun, W.; Jia, G.; Jangbae, K.; Hyotcherl, I.; Donglin, J.; A Belt-Shaped, Blue Luminescent, and Semiconducting Covalent Organic Framework. Angew. Chem. Int. Ed.2008, 47, pp 8826-8830. doi : 10.1002/anie.200890235
  43. Marco, B.; Cortizo-Lacalle, D.; Perez-Miqueo, C.; Valenti, G.; Boni, A.; Plas, J.; Strutynski, K.; De Feyter, S.; Paolucci, F.; Montes, M.; Khlobystov, K.; Melle-Franco, M.; Mateo-Alonso, A. (2017). "Twisted Aromatic Frameworks: Readily Exfoliable and Solution-Processable Two-Dimensional Conjugated Microporous Polymers". Angew. Chem. Int. Ed. 56 (24): 6946–6951. doi:10.1002/anie.201700271. PMC   5485174 . PMID   28318084.
  44. Martin, Richard (September 24, 2015). "New Technology to Capture, Convert Carbon Dioxide | MIT Technology Review" . Retrieved 2015-09-27.
  45. Hussain, MD. Waseem; Bhardwaj, Vipin; GIRI, ARKAPRABHA; Chande, Ajit; Patra, Abhijit (2019-11-27). "Functional Ionic Porous Frameworks Based on Triaminoguanidinium for CO2 Conversion and Combating Microbes". doi:10.26434/chemrxiv.10332431 . Retrieved 2022-06-22.{{cite journal}}: Cite journal requires |journal= (help)
  46. Liang, Huihui; Xu, Mengli; Zhu, Yongmei; Wang, Linyu; Xie, Yi; Song, Yonghai; Wang, Li (2020-01-24). "H2O2 Ratiometric Electrochemical Sensors Based on Nanospheres Derived from Ferrocence-Modified Covalent Organic Frameworks". ACS Applied Nano Materials. 3 (1): 555–562. doi: 10.1021/acsanm.9b02117 . ISSN   2574-0970. S2CID   214062588.
  47. Hu, Hui; Yan, Qianqian; Ge, Rile; Gao, Yanan (July 2018). "Covalent organic frameworks as heterogeneous catalysts". Chinese Journal of Catalysis. 39 (7): 1167–1179. doi:10.1016/S1872-2067(18)63057-8. S2CID   102933312.
  48. Guo, Jia; Jiang, Donglin (2020-06-24). "Covalent Organic Frameworks for Heterogeneous Catalysis: Principle, Current Status, and Challenges". ACS Central Science. 6 (6): 869–879. doi:10.1021/acscentsci.0c00463. ISSN   2374-7943. PMC   7318070 . PMID   32607434.
  49. Zheng, Weiran; Tsang, Chui-Shan; Lee, Lawrence Yoon Suk; Wong, Kwok-Yin (June 2019). "Two-dimensional metal-organic framework and covalent-organic framework: synthesis and their energy-related applications". Materials Today Chemistry. 12: 34–60. doi:10.1016/j.mtchem.2018.12.002. hdl: 10397/101525 . S2CID   139305086.
  50. Yang, Hui; Zhang, Shengliang; Han, Liheng; Zhang, Zhou; Xue, Zheng; Gao, Juan; Li, Yongjun; Huang, Changshui; Yi, Yuanping; Liu, Huibiao; Li, Yuliang (16 February 2016). "High Conductive Two-Dimensional Covalent Organic Framework for Lithium Storage with Large Capacity". ACS Applied Materials & Interfaces. 8 (8): 5366–5375. doi:10.1021/acsami.5b12370. PMID   26840757.
  51. Diercks, Christian S.; Lin, Song; Kornienko, Nikolay; Kapustin, Eugene A.; Nichols, Eva M.; Zhu, Chenhui; Zhao, Yingbo; Chang, Christopher J.; Yaghi, Omar M. (16 January 2018). "Reticular Electronic Tuning of Porphyrin Active Sites in Covalent Organic Frameworks for Electrocatalytic Carbon Dioxide Reduction" (PDF). Journal of the American Chemical Society. 140 (3): 1116–1122. doi:10.1021/jacs.7b11940. PMID   29284263. S2CID   207188096.
  52. Li, Xing; Wang, Hui; Chen, Zhongxin; Xu, Hai‐Sen; Yu, Wei; Liu, Cuibo; Wang, Xiaowei; Zhang, Kun; Xie, Keyu; Loh, Kian Ping (2019-10-14). "Covalent‐Organic‐Framework‐Based Li–CO 2 Batteries". Advanced Materials. 31 (48): 1905879. Bibcode:2019AdM....3105879L. doi:10.1002/adma.201905879. ISSN   0935-9648. PMID   31609043. S2CID   204545588.
  53. Luo, Zhiqiang; Liu, Luojia; Ning, Jiaxin; Lei, Kaixiang; Lu, Yong; Li, Fujun; Chen, Jun (2018-07-20). "A Microporous Covalent-Organic Framework with Abundant Accessible Carbonyl Groups for Lithium-Ion Batteries". Angewandte Chemie International Edition. 57 (30): 9443–9446. doi:10.1002/anie.201805540. PMID   29863784. S2CID   205407552.
  54. Miner, Elise M.; Dincă, Mircea (2019-07-15). "Metal- and covalent-organic frameworks as solid-state electrolytes for metal-ion batteries". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 377 (2149): 20180225. Bibcode:2019RSPTA.37780225M. doi:10.1098/rsta.2018.0225. ISSN   1364-503X. PMC   6562342 . PMID   31130094.
  55. Vitaku, Edon; Gannett, Cara N.; Carpenter, Keith L.; Shen, Luxi; Abruña, Héctor D.; Dichtel, William R. (2020-01-08). "Phenazine-Based Covalent Organic Framework Cathode Materials with High Energy and Power Densities". Journal of the American Chemical Society. 142 (1): 16–20. doi:10.1021/jacs.9b08147. ISSN   0002-7863. PMID   31820958. S2CID   209317683.
  56. Irving, Michael (2022-08-05). "Nano-sponges on graphene make efficient filters of industrial wastewater". New Atlas. Retrieved 2022-08-06.
  57. Zhao, Yu; Das, Saikat; Sekine, Taishu; Mabuchi, Haruna; Irie, Tsukase; Sakai, Jin; Wen, Dan; Zhu, Weidong; Ben, Teng; Negishi, Yuichi (2023-01-23). "Record Ultralarge-Pores, Low Density Three-Dimensional Covalent Organic Framework for Controlled Drug Delivery". Angewandte Chemie. 62 (13): e202300172. doi: 10.1002/anie.202300172 . PMID   36688253.