Hydrogen-bonded organic framework

Last updated

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. [1] [2] [3] [4] 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. [3] 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. [5] [6]

Contents

An important consequence[ editorializing ] of the natural porous architecture of hydrogen-bonded organic frameworks is to realize the adsorption of guest molecules. This character accelerates the emergence of various applications of different HOFs structures, including gas removal/storage/separation, molecule recognition, proton conduction, and biomedical applications, etc. [1] [7] [8]

(a) 1,3,5-Tris(4-carboxyphenyl)benzene as a HOF monomer unit. (b) The corresponding HOF structure with porous packing. Hydrogen-bonded organic frameworks.jpg
(a) 1,3,5-Tris(4-carboxyphenyl)benzene as a HOF monomer unit. (b) The corresponding HOF structure with porous packing.

History

Reports of extended 2D hydrogen-bonding-based porous frameworks can be traced back to the 1960s. In 1969, Duchamp and Marsh reported a 2D interpenetrated nonporous crystal structure with a honeycomb network constructed by benzene-1,3,5-tricarboxylic acid (trimesic acid or TMA). [9] Then Ermer reported an adamantane-1,3,5,7-tetracarboxylic acid (ADTA) based hydrogen-bonded network with interpenetrated diamond topology. [10] Meanwhile diverse works of guest-induced hydrogen-bonded frameworks were reported successively, which gradually developed the concept of hydrogen-bonded organic frameworks. [11] [12] [13] [14] Another milestone in the evolution of hydrogen-bonded organic frameworks was set by Chen. In 2011, Chen reported a porous organic framework with hydrogen bonding as binding force and demonstrated its porosity by gas adsorption for the first time. [4] Since then, numerous HOF structures have been designed and constructed, meanwhile various applications related to porous frameworks have been attempted and applied to HOFs, whose effectiveness has been proved. [2] [3]  

Hydrogen bonding pairs in HOFs

Hydrogen bonds formed among various monomers guarantee the construction of hydrogen-bonded organic frameworks with different assembly architectures. [15] [16] [17] The constitution of the hydrogen pairs is based on the structural and functional design of the HOFs, therefore different hydrogen bonding pairs should be selected following systematic requirements. The hydrogen bonding pairs generally include 2,4-diaminotriazine, carboxylic acid, amide, imide, imidazole, imidazolone and resorcinol, etc. [2] [3] [18] [19] [20] [21] Assorting with appropriate backbones, in every crystallization condition, the hydrogen-bonding pairs will exhibit specific assembly states, which means the morphologies with favored energy for this crystallization condition could be assembled by the monomers. In order to realize 2D or 3D HOFs, monomers with more than one hydrogen bonding pair are generally considered: the rigidity and directionality are also in favor of HOF construction.    

Organic group examples as potential hydrogen bonding units. Fig. 2. Organic group examples as potential hydrogen bonding units.jpg
Organic group examples as potential hydrogen bonding units.

Backbones of HOF monomer

Rigidity and directionality of the constructional units offer HOFs various pore structures, topologies, and further applications. [2] Therefore, a proper choice of monomer backbones plays an important role in the construction of HOFs. These backbones not only can combine with different hydrogen bonding pairs mentioned above to realize stable HOF structural design and expand pore size, but also give opportunities to offer more topologies of HOFs. Also, by using backbones with similar geometry and same connection pattern to generate the monomers and HOFs, the isoreticular expansion of the frameworks becomes a reliable method to expand the pore size effectively. [22] [23] As mentioned, for the sake of constructing porous and stable HOFs, multiple aspects should be considered simultaneously, such as the rigidity of the backbones, the orientation and binding strength of the hydrogen pairs, and other intermolecular interactions for orderly stacking. Therefore, the design of HOF monomers should focus on their H-bonds orientations and structural rigidity, and consequent framework stability and porosity.  

A series of isoreticular expansion carboxylic acid based monomers (a-c) and HOFs (d-f). Fig. 3. A series of isoreticular expansion carboxylic acid based monomers (a-c) and HOFs (d-f)..jpg
A series of isoreticular expansion carboxylic acid based monomers (a-c) and HOFs (d-f).

Synthetic methods

In principle, HOFs could be crystallized from solvents. [4] However, the factors of solvent types, precursor concentration, crystallization time and temperature, etc., can have significant influence on HOFs crystallization process. Generally, the crystal products can correspond to kinetics through high concentration and short crystallization time, while slowing down the crystallization rate might yield thermodynamic crystals. One common method to produce HOF crystal is to slowly evaporate the solvent of the solution, which benefits the stacking of the monomers. [4] Another widely used method is to diffuse low boiling point poor solvents into monomer solution with higher boiling point good solvents, in order to induce the assembly of the monomers. [4] Depending on different crystallization systems, other methods have also been applied to HOF construction. [4]           

Characterization methods

There are various methods to characterize HOF materials and their monomers. Nuclear magnetic resonance (NMR) spectroscopy and high-resolution mass spectrometry (HR-MS) are generally used for characterizing the synthesis of monomers. [1] [2] Single crystal X-ray diffraction (SCXRD) is the powerful tool for determining the structure of the HOF crystal packing. Powder X-ray diffraction (PXRD) is also a supported technique to demonstrate the pure phase formation of HOFs. [1] [2] The gas adsorption and desorption study through Brunauer-Emmett-Teller (BET) method could reasonably demonstrate some key parameters of HOFs, like pore size, specific gas adsorption amount and surface area from the adsorption isotherms. Depending on application directions and study fields, diverse techniques have been applied to the characterization of HOFs. [1] [2]  

Applications

The porous structures and unique properties guarantee HOFs good application performance in practical fields. The applications include but are not limited to gas adsorption, hydrocarbon separation, proton conductivity, and molecular recognition, etc.

HOF applications HOF Fig. 4.jpg
HOF applications

Gas adsorption

As a kind of networks with tailorable pore size, HOFs could serve as storage containers for gas molecules with certain sizes and interactions. [24] [25] The relatively constrained pore size in HOFs could help to store, capture, or separate different small gas molecules, including H2, N2, CO2, CH4, C2H2, C2H4, C2H6 and so on. [2] Mastalerz and Oppel reported a special 3D HOF with triptycene trisbenzimidazolone (TTBI) as constitutional monomers. Because of the molecular rigidity and stereo construction, 1D channels were formed through the frameworks and the surface area was largely enhanced, to the extent of 2796 m2/g as shown by BET. [26] The HOF also presented good adsorption ability of H2 and CO2, as 243 and 80.7 cm3/g at 1 bar at 77 and 273 K, separately.[ citation needed ]

CO2 adsorption

As a typical greenhouse gas that could cause serious problems in many aspects, the capture of carbon dioxide is always under big concern. Meanwhile, carbon dioxide has also been widely used as a gas resource or emitted as waste gas in manufacturing and industry, therefore the storage and separation of CO2 have always been emphasized as an important application. Chen and co-workers reported a structural transformation HOF with high CO2 adsorption capacity in 2015. [27] The N–H···N hydrogen bond is formed between the units to realize the assembly of the HOF architecture with binodal topology. The CO2 uptake capacity of the HOF could reach 117.1 cm3/g at 273 K.[ citation needed ]

Hydrocarbon separation

The hydrogen-bonded organic framework used for C2H2/C2H4 separation was reported by Chen and coworkers. In the structure of this HOF, each 4,4',4'',4'''-tetra(4,6-diamino-s-triazin-2-yl)tetraphenylmethane unit connected with eight other units by N–H···N hydrogen bonds. [3] Due to certain structural flexibility, the framework was able to uptake C2H2 up to 63.2 cm3/g while the adsorption amount of C2H4 was 8.3 cm3/g at 273 K, showing effective C2H2/C2H4 separation.[ citation needed ]

Molecules recognition

The non-covalent interactions existing in the hydrogen-bonded organic frameworks, e.g., hydrogen bonding, π-π interaction and Van der Waals force, are considered as important intermolecular interactions for molecules recognition. Meanwhile, the multiple binding sites and adaptable structures also make HOFs good molecules recognition platform. By exploiting these features, so far different kinds of recognition have been realized, including gas molecules recognition, fullerene recognition, aniline recognition, pyridine recognition, etc. [28] [27] [29] [30]  

Optical materials

Some luminescence molecules with large π conjugation structures are also used for HOFs construction. Therefore, various luminescent HOFs are designed and assembled in order to realize the non-covalent controlled luminescence adjustment which could introduce more functions to the HOF materials. [31] For example, by using tetraphenylethylene (TPE) as backbones, a series of HOFs combined with solvents presenting different color emission have been reported. [32]    

Proton conduction

The hydrogen-bonded organic frameworks constructed with proton carriers have been widely used for proton conduction. The hydrogen bonds can also serve as proton sources in the frameworks to transfer protons. As an example, porphyrin-based structures and guanidinium sulfonate salt monomers have been studied and included in HOF design and construction for proton conduction since the certain conductivity they have. [33] [34]

Biological applications

As kinds of metal-free porous materials, hydrogen-bonded organic frameworks are also ideal platform for drug delivery and disease treatment. [35] Meanwhile, with proper monomer selection and reasonable arrangement, Cao reported a robust HOF which could effectively encapsulate a cancer drug Doxorubicin and yield singlet oxygen by embedded photoactive pyrene moiety in order to realize dual functions of drug release and photodynamic therapy for cancer remedy. [36]

Related Research Articles

<span class="mw-page-title-main">Clathrate compound</span> Chemical substance consisting of a lattice that traps or contains molecules

A clathrate is a chemical substance consisting of a lattice that traps or contains molecules. The word clathrate is derived from the Latin clathratus, meaning 'with bars, latticed'. Most clathrate compounds are polymeric and completely envelop the guest molecule, but in modern usage clathrates also include host–guest complexes and inclusion compounds. According to IUPAC, clathrates are inclusion compounds "in which the guest molecule is in a cage formed by the host molecule or by a lattice of host molecules." The term refers to many molecular hosts, including calixarenes and cyclodextrins and even some inorganic polymers such as zeolites.

<span class="mw-page-title-main">Fluoroantimonic acid</span> Chemical compound

Fluoroantimonic acid is a mixture of hydrogen fluoride and antimony penta­fluoride, containing various cations and anions. This mixture is a superacid that, in terms of corrosiveness, is trillions of times stronger than 100% sulfuric acid in terms of its protonating ability measured by Hammett function. It even protonates some hydro­carbons to afford pentacoordinate carbo­cations. Like its precursor hydrogen fluoride, it attacks glass, but can be stored in containers lined with PTFE (Teflon) or PFA.

<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">Transition metal dinitrogen complex</span> Coordination compounds with N2

Transition metal dinitrogen complexes are coordination compounds that contain transition metals as ion centers the dinitrogen molecules (N2) as ligands.

In chemistry, a halogen bond occurs when there is evidence of a net attractive interaction between an electrophilic region associated with a halogen atom in a molecular entity and a nucleophilic region in another, or the same, molecular entity. Like a hydrogen bond, the result is not a formal chemical bond, but rather a strong electrostatic attraction. Mathematically, the interaction can be decomposed in two terms: one describing an electrostatic, orbital-mixing charge-transfer and another describing electron-cloud dispersion. Halogen bonds find application in supramolecular chemistry; drug design and biochemistry; crystal engineering and liquid crystals; and organic catalysis.

<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.

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. 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. 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. 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. 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.

<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">Carbon nanothread</span>

A carbon nanothread is a sp3-bonded, one-dimensional carbon crystalline nanomaterial. The tetrahedral sp3-bonding of its carbon is similar to that of diamond. Nanothreads are only a few atoms across, more than 20,000 times thinner than a human hair. They consist of a stiff, strong carbon core surrounded by hydrogen atoms. Carbon nanotubes, although also one-dimensional nanomaterials, in contrast have sp2-carbon bonding as is found in graphite. The smallest carbon nanothread has a diameter of only 0.2 nanometers, much smaller than the diameter of a single-wall carbon nanotube.

Diborane(2), also known as diborene, is an inorganic compound with the formula B2H2. The number 2 in diborane(2) indicates the number of hydrogen atoms bonded to the boron complex. There are other forms of diborane with different numbers of hydrogen atoms, including diborane(4) and diborane(6).

In chemistry, a chalcogen bond (ChB) is an attractive interaction in the family of σ-hole interactions, along with halogen bonds. Electrostatic, charge-transfer (CT) and dispersion terms have been identified as contributing to this type of interaction. In terms of CT contribution, this family of attractive interactions has been modeled as an electron donor ) interacting with the σ* orbital of a C-X bond of the bond donor. In terms of electrostatic interactions, the molecular electrostatic potential (MEP) maps is often invoked to visualize the electron density of the donor and an electrophilic region on the acceptor, where the potential is depleted, referred to as a σ-hole. ChBs, much like hydrogen and halogen bonds, have been invoked in various non-covalent interactions, such as protein folding, crystal engineering, self-assembly, catalysis, transport, sensing, templation, and drug design.

<span class="mw-page-title-main">Field effect (chemistry)</span>

A field effect is the polarization of a molecule through space. The effect is a result of an electric field produced by charge localization in a molecule. This field, which is substituent and conformation dependent, can influence structure and reactivity by manipulating the location of electron density in bonds and/or the overall molecule. The polarization of a molecule through its bonds is a separate phenomenon known as induction. Field effects are relatively weak, and diminish rapidly with distance, but have still been found to alter molecular properties such as acidity.

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.

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">MIL-53</span> Chemical compound

MIL-53 belongs to the class of metal-organic framework (MOF) materials. The first synthesis and the name was established by the group of Gérard Férey in 2002. The MIL-53 structure consists of inorganic [M-OH] chains, which are connected to four neighboring inorganic chains by therephthalate-based linker molecules. Each metal center is octahedrally coordinated by six oxygen atoms. Four of these oxygen atoms originate from four different carboxylate groups and the remaining two oxygen atoms belong to two different μ-OH moieties, which bridge neighboring metal centers. The resulting framework structure contains one-dimensional diamond-shaped pores. Many research group have investigated the flexibility of the MIL-53 structure. This flexible behavior, during which the pore cross-section changes reversibly, was termed 'breathing.effect' and describes the ability of the MIL-53 framework to respond to external stimuli.

<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">Hoffmann Institute of Advanced Materials</span>

The Hoffmann Institute of Advanced Materials (HIAM) is a science research institute affiliated to Shenzhen Polytechnic in Shenzhen, China. As the eighth institute at Shenzhen named after a Nobel laureate, it was founded in February 2018 under the tutelage of the theoretical chemist Roald Hoffmann. The institute was officially opened with a formal ceremony in May 2019. 

<span class="mw-page-title-main">Topochemical polymerization</span>

Topochemical polymerization is a polymerization method performed by monomers aligned in the crystal state. In this process, the monomers are crystallised and polymerised under external stimuli such as heat, light, or pressure. Compared to traditional polymerisation, the movement of monomers was confined by the crystal lattice in topochemical polymerisation, giving rise to polymers with high crystallinity, tacticity, and purity. Topochemical polymerisation can also be used to synthesise unique polymers such as polydiacetylene that are otherwise hard to prepare.

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

Porous polymers are a class of porous media materials in which monomers form 2D and 3D polymers containing pores. 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 3 4 5 Li, Penghao; Ryder, Matthew R.; Stoddart, J. Fraser (2020). "Hydrogen-Bonded Organic Frameworks: A Rising Class of Porous Molecular Materials". Accounts of Materials Research. 1: 77–87. doi: 10.1021/accountsmr.0c00019 . S2CID   225131091.
  2. 1 2 3 4 5 6 7 8 Lin, Rui-Biao; Chen, Banglin (2022). "Hydrogen-bonded organic frameworks: Chemistry and functions". Chem. 8 (8): 2114–2135. doi: 10.1016/j.chempr.2022.06.015 . S2CID   250942238.
  3. 1 2 3 4 5 Lin, Rui-Biao; He, Yabing; Li, Peng; Wang, Hailong; Zhou, Wei; Chen, Banglin (2019). "Multifunctional porous hydrogen-bonded organic framework materials". Chemical Society Reviews. 48 (5): 1362–1389. doi:10.1039/C8CS00155C. ISSN   0306-0012. PMID   30676603. S2CID   59225974.
  4. 1 2 3 4 5 6 He, Yabing; Xiang, Shengchang; Chen, Banglin (2011-09-21). "A Microporous Hydrogen-Bonded Organic Framework for Highly Selective C 2 H 2 /C 2 H 4 Separation at Ambient Temperature". Journal of the American Chemical Society. 133 (37): 14570–14573. doi:10.1021/ja2066016. ISSN   0002-7863. PMID   21863860.
  5. Luo, Jie; Wang, Jia-Wei; Zhang, Ji-Hong; Lai, Shan; Zhong, Di-Chang (2018). "Hydrogen-bonded organic frameworks: design, structures and potential applications". CrystEngComm. 20 (39): 5884–5898. doi:10.1039/C8CE00655E. ISSN   1466-8033.
  6. Hisaki, Ichiro; Xin, Chen; Takahashi, Kiyonori; Nakamura, Takayoshi (2019-08-12). "Designing Hydrogen-Bonded Organic Frameworks (HOFs) with Permanent Porosity". Angewandte Chemie International Edition. 58 (33): 11160–11170. doi: 10.1002/anie.201902147 . ISSN   1433-7851. PMID   30891889. S2CID   84184025.
  7. Wang, Bin; Lin, Rui-Biao; Zhang, Zhangjing; Xiang, Shengchang; Chen, Banglin (2020-08-26). "Hydrogen-Bonded Organic Frameworks as a Tunable Platform for Functional Materials". Journal of the American Chemical Society. 142 (34): 14399–14416. doi:10.1021/jacs.0c06473. ISSN   0002-7863. PMID   32786796. S2CID   225425763.
  8. Lü, Jian; Cao, Rong (2016-08-08). "Porous Organic Molecular Frameworks with Extrinsic Porosity: A Platform for Carbon Storage and Separation". Angewandte Chemie International Edition. 55 (33): 9474–9480. doi:10.1002/anie.201602116. PMID   27410190.
  9. Duchamp, D. J.; Marsh, R. E. (1969-01-15). "The crystal structure of trimesic acid (benzene-1,3,5-tricarboxylic acid)". Acta Crystallographica Section B Structural Crystallography and Crystal Chemistry. 25 (1): 5–19. Bibcode:1969AcCrB..25....5D. doi:10.1107/S0567740869001713.
  10. Ermer, Otto. (1988). "Five-fold diamond structure of adamantane-1,3,5,7-tetracarboxylic acid". Journal of the American Chemical Society. 110 (12): 3747–3754. doi:10.1021/ja00220a005. ISSN   0002-7863.
  11. Herbstein, F. H.; Kapon, M.; Reisner, G. M. (1987). "Catenated and non-catenated inclusion complexes of trimesic acid". Journal of Inclusion Phenomena. 5 (2): 211–214. doi:10.1007/BF00655650. ISSN   0167-7861. S2CID   93043649.
  12. Ibragimov, B. T.; Talipov, S. A.; Aripov, T. F. (1994). "Inclusion complexes of the natural product gossypol. Recognition by gossypol of halogeno methanes. Structure of the dichloromethane complex of gossypol and single crystal conservation after decomposition". Journal of Inclusion Phenomena and Molecular Recognition in Chemistry. 17 (4): 317–324. doi:10.1007/BF00707127. ISSN   0923-0750. S2CID   96181739.
  13. Simard, Michel; Su, Dan; Wuest, James D. (1991). "Use of hydrogen bonds to control molecular aggregation. Self-assembly of three-dimensional networks with large chambers". Journal of the American Chemical Society. 113 (12): 4696–4698. doi:10.1021/ja00012a057. ISSN   0002-7863.
  14. Ermer, Otto; Lindenberg, Lorenz (1991-06-19). "Double-Diamond Inclusion Compounds of 2,6-Dimethylydeneadamantane-1,3,5,7-tetracarboxylic Acid". Helvetica Chimica Acta (in German). 74 (4): 825–877. doi:10.1002/hlca.19910740417. ISSN   0018-019X.
  15. Chen, Teng-Hao; Popov, Ilya; Kaveevivitchai, Watchareeya; Chuang, Yu-Chun; Chen, Yu-Sheng; Daugulis, Olafs; Jacobson, Allan J.; Miljanić, Ognjen Š. (2014-10-13). "Thermally robust and porous noncovalent organic framework with high affinity for fluorocarbons and CFCs". Nature Communications. 5 (1): 5131. Bibcode:2014NatCo...5.5131C. doi: 10.1038/ncomms6131 . ISSN   2041-1723. PMID   25307413. S2CID   1078.
  16. Zhou, Yue; Liu, Bing; Sun, Xiaodong; Li, Jiantang; Li, Guanghua; Huo, Qisheng; Liu, Yunling (2017-12-06). "Self-assembly of Homochiral Porous Supramolecular Organic Frameworks with Significant CO 2 Capture and CO 2 /N 2 Selectivity". Crystal Growth & Design. 17 (12): 6653–6659. doi:10.1021/acs.cgd.7b01282. ISSN   1528-7483.
  17. Wahl, Helene; Haynes, Delia A.; le Roex, Tanya (2017-08-02). "Guest Exchange in a Robust Hydrogen-Bonded Organic Framework: Single-Crystal to Single-Crystal Exchange and Kinetic Studies". Crystal Growth & Design. 17 (8): 4377–4383. doi:10.1021/acs.cgd.7b00684. ISSN   1528-7483.
  18. Maly, Kenneth E.; Gagnon, Eric; Maris, Thierry; Wuest, James D. (2007-04-01). "Engineering Hydrogen-Bonded Molecular Crystals Built from Derivatives of Hexaphenylbenzene and Related Compounds". Journal of the American Chemical Society. 129 (14): 4306–4322. doi:10.1021/ja067571x. ISSN   0002-7863. PMID   17358060.
  19. Luo, Xu-Zhong; Jia, Xin-Jian; Deng, Ji-Hua; Zhong, Jin-Lian; Liu, Hui-Jin; Wang, Ke-Jun; Zhong, Di-Chang (2013-08-14). "A Microporous Hydrogen-Bonded Organic Framework: Exceptional Stability and Highly Selective Adsorption of Gas and Liquid". Journal of the American Chemical Society. 135 (32): 11684–11687. doi:10.1021/ja403002m. ISSN   0002-7863. PMID   23885835.
  20. Fournier, Jean-Hugues; Maris, Thierry; Wuest, James D.; Guo, Wenzhuo; Galoppini, Elena (2003-01-01). "Molecular Tectonics. Use of the Hydrogen Bonding of Boronic Acids To Direct Supramolecular Construction". Journal of the American Chemical Society. 125 (4): 1002–1006. doi:10.1021/ja0276772. ISSN   0002-7863. PMID   12537499.
  21. Maly, Kenneth E.; Buck, William; Dawe, Louise N. (2017). "Open network structures from 2D hydrogen bonded networks: diaminotriazyl tetraoxapentacenes". CrystEngComm. 19 (43): 6401–6405. doi:10.1039/C7CE01247K. ISSN   1466-8033.
  22. Hisaki, Ichiro; Nakagawa, Shoichi; Ikenaka, Nobuaki; Imamura, Yutaka; Katouda, Michio; Tashiro, Motomichi; Tsuchida, Hiromu; Ogoshi, Tomoki; Sato, Hiroyasu; Tohnai, Norimitsu; Miyata, Mikiji (2016-05-25). "A Series of Layered Assemblies of Hydrogen-Bonded, Hexagonal Networks of C 3 -Symmetric π-Conjugated Molecules: A Potential Motif of Porous Organic Materials". Journal of the American Chemical Society. 138 (20): 6617–6628. doi:10.1021/jacs.6b02968. ISSN   0002-7863. PMID   27133443.
  23. Hisaki, Ichiro; Suzuki, Yuto; Gomez, Eduardo; Cohen, Boiko; Tohnai, Norimitsu; Douhal, Abderrazzak (2018-09-24). "Docking Strategy To Construct Thermostable, Single-Crystalline, Hydrogen-Bonded Organic Framework with High Surface Area". Angewandte Chemie International Edition. 57 (39): 12650–12655. doi: 10.1002/anie.201805472 . PMID   29885200. S2CID   205407547.
  24. Yang, Wenbin; Greenaway, Alex; Lin, Xiang; Matsuda, Ryotaro; Blake, Alexander J.; Wilson, Claire; Lewis, William; Hubberstey, Peter; Kitagawa, Susumu; Champness, Neil R.; Schröder, Martin (2010-10-20). "Exceptional Thermal Stability in a Supramolecular Organic Framework: Porosity and Gas Storage". Journal of the American Chemical Society. 132 (41): 14457–14469. doi:10.1021/ja1042935. ISSN   0002-7863. PMID   20866087.
  25. Chen, Teng-Hao; Kaveevivitchai, Watchareeya; Jacobson, Allan J.; Miljanić, Ognjen Š. (2015). "Adsorption of fluorinated anesthetics within the pores of a molecular crystal". Chemical Communications. 51 (74): 14096–14098. doi:10.1039/C5CC04885K. ISSN   1359-7345. PMID   26252729.
  26. Mastalerz, Michael; Oppel, Iris M. (2012-05-21). "Rational Construction of an Extrinsic Porous Molecular Crystal with an Extraordinary High Specific Surface Area". Angewandte Chemie International Edition. 51 (21): 5252–5255. doi: 10.1002/anie.201201174 . PMID   22473702.
  27. 1 2 Wang, Hailong; Li, Bin; Wu, Hui; Hu, Tong-Liang; Yao, Zizhu; Zhou, Wei; Xiang, Shengchang; Chen, Banglin (2015-08-12). "A Flexible Microporous Hydrogen-Bonded Organic Framework for Gas Sorption and Separation". Journal of the American Chemical Society. 137 (31): 9963–9970. doi:10.1021/jacs.5b05644. ISSN   0002-7863. PMID   26214340.
  28. Natarajan, Ramalingam; Bridgland, Lydia; Sirikulkajorn, Anchalee; Lee, Ji-Hun; Haddow, Mairi F.; Magro, Germinal; Ali, Bakhat; Narayanan, Sampriya; Strickland, Peter; Charmant, Jonathan P. H.; Orpen, A. Guy; McKeown, Neil B.; Bezzu, C. Grazia; Davis, Anthony P. (2013-11-13). "Tunable Porous Organic Crystals: Structural Scope and Adsorption Properties of Nanoporous Steroidal Ureas". Journal of the American Chemical Society. 135 (45): 16912–16925. doi:10.1021/ja405701u. ISSN   0002-7863. PMC   3880060 . PMID   24147834.
  29. Wang, Hailong; Wu, Hui; Kan, Jinglan; Chang, Ganggang; Yao, Zizhu; Li, Bin; Zhou, Wei; Xiang, Shengchang; Cong-Gui Zhao, John; Chen, Banglin (2017). "A microporous hydrogen-bonded organic framework with amine sites for selective recognition of small molecules". Journal of Materials Chemistry A. 5 (18): 8292–8296. doi:10.1039/C7TA01364G. ISSN   2050-7488.
  30. Yan, Wenqing; Yu, Xiaopeng; Yan, Tao; Wu, Doufeng; Ning, Erlong; Qi, Yi; Han, Ying-Feng; Li, Qiaowei (2017). "A triptycene-based porous hydrogen-bonded organic framework for guest incorporation with tailored fitting". Chemical Communications. 53 (26): 3677–3680. doi:10.1039/C7CC00557A. ISSN   1359-7345. PMID   28265598.
  31. Bian, Lifang; Shi, Huifang; Wang, Xuan; Ling, Kun; Ma, Huili; Li, Mengping; Cheng, Zhichao; Ma, Chaoqun; Cai, Suzhi; Wu, Qi; Gan, Nan; Xu, Xiangfei; An, Zhongfu; Huang, Wei (2018-08-29). "Simultaneously Enhancing Efficiency and Lifetime of Ultralong Organic Phosphorescence Materials by Molecular Self-Assembly". Journal of the American Chemical Society. 140 (34): 10734–10739. doi:10.1021/jacs.8b03867. ISSN   0002-7863. PMID   30078313. S2CID   51921536.
  32. Huang, Qiuyi; Li, Wenlang; Mao, Zhu; Qu, Lunjun; Li, Yang; Zhang, Hao; Yu, Tao; Yang, Zhiyong; Zhao, Juan; Zhang, Yi; Aldred, Matthew P.; Chi, Zhenguo (2019-07-12). "An exceptionally flexible hydrogen-bonded organic framework with large-scale void regulation and adaptive guest accommodation abilities". Nature Communications. 10 (1): 3074. Bibcode:2019NatCo..10.3074H. doi:10.1038/s41467-019-10575-5. ISSN   2041-1723. PMC   6625987 . PMID   31300644.
  33. Yang, Wei; Yang, Fan; Hu, Tong-Liang; King, Stephen Charles; Wang, Hailong; Wu, Hui; Zhou, Wei; Li, Jian-Rong; Arman, Hadi D.; Chen, Banglin (2016-10-05). "Microporous Diaminotriazine-Decorated Porphyrin-Based Hydrogen-Bonded Organic Framework: Permanent Porosity and Proton Conduction". Crystal Growth & Design. 16 (10): 5831–5835. doi:10.1021/acs.cgd.6b00924. ISSN   1528-7483.
  34. Karmakar, Avishek; Illathvalappil, Rajith; Anothumakkool, Bihag; Sen, Arunabha; Samanta, Partha; Desai, Aamod V.; Kurungot, Sreekumar; Ghosh, Sujit K. (2016-08-26). "Hydrogen-Bonded Organic Frameworks (HOFs): A New Class of Porous Crystalline Proton-Conducting Materials". Angewandte Chemie International Edition. 55 (36): 10667–10671. doi:10.1002/anie.201604534. PMID   27464784.
  35. Bracco, S.; Asnaghi, D.; Negroni, M.; Sozzani, P.; Comotti, A. (2018). "Porous dipeptide crystals as volatile-drug vessels". Chemical Communications. 54 (2): 148–151. doi:10.1039/C7CC06534E. hdl: 10281/183765 . ISSN   1359-7345. PMID   29210379. S2CID   206070122.
  36. Yin, Qi; Zhao, Peng; Sa, Rong-Jian; Chen, Guang-Cun; Lü, Jian; Liu, Tian-Fu; Cao, Rong (2018-06-25). "An Ultra-Robust and Crystalline Redeemable Hydrogen-Bonded Organic Framework for Synergistic Chemo-Photodynamic Therapy". Angewandte Chemie. 130 (26): 7817–7822. Bibcode:2018AngCh.130.7817Y. doi:10.1002/ange.201800354. S2CID   242082705.