Two-dimensional polymer

Last updated
Structural difference between a linear and a two-dimensional (2D) polymer. In the former, linearly connecting monomers result in a thread-like linear polymer, while in the latter laterally connecting monomers result in a sheet-like 2DP with regularly tessellated repeat units (here of square geometry). The repeat units are marked in red, whereby the number n describes the degree of polymerization. While a linear polymer has two end groups, a 2DP has an infinite number of end groups that are positioned all along the sheet edges (green arrows). 2Dpolymer.png
Structural difference between a linear and a two-dimensional (2D) polymer. In the former, linearly connecting monomers result in a thread-like linear polymer, while in the latter laterally connecting monomers result in a sheet-like 2DP with regularly tessellated repeat units (here of square geometry). The repeat units are marked in red, whereby the number n describes the degree of polymerization. While a linear polymer has two end groups, a 2DP has an infinite number of end groups that are positioned all along the sheet edges (green arrows).

A two-dimensional polymer (2DP) is a sheet-like monomolecular macromolecule consisting of laterally connected repeat units with end groups along all edges. [1] [2] This recent definition of 2DP is based on Hermann Staudinger's polymer concept from the 1920s. [3] [4] [5] [6] 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.

Contents

Moving from one dimension to two offers access to surface morphologies such as increased surface area, porous membranes, and possibly in-plane pi orbital-conjugation for enhanced electronic properties. They are distinct from other families of polymers because 2D polymers can be isolated as multilayer crystals or as individual sheets. [7]

The term 2D polymer has also been used more broadly to include linear polymerizations performed at interfaces, layered non-covalent assemblies, or to irregularly cross-linked polymers confined to surfaces or layered films. [8] 2D polymers can be organized based on these methods of linking (monomer interaction): covalently linked monomers, coordination polymers and supramolecular polymers. 2D polymers containing pores are also known as porous polymers.

Topologically, 2DPs may thus be understood as structures made up from regularly tessellated regular polygons (the repeat units). Figure 1 displays the key features of a linear and a 2DP according to this definition. For usage of the term "2D polymer" in a wider sense, see "History".

Covalently-linked polymers

There are several examples of covalently linked 2DPs which include the individual layers or sheets of graphite (called graphenes), MoS2, (BN)x and layered covalent organic frameworks. As required by the above definition, these sheets have a periodic internal structure.

A well-known example of a 2D polymer is graphene; whose optical, electronic and mechanical properties have been studied in depth. Graphene has a honeycomb lattice of carbon atoms that exhibit semiconducting properties. A potential repeat unit of graphene is a sp2-hybridized carbon atom. Individual sheets can in principle be obtained by exfoliation procedures, though in reality this is a non-trivial enterprise.

Molybdenumdisulfide can exist in two-dimensional, single or layered polymers where each Mo(IV) center occupies a trigonal prismatic coordination sphere.

Boron nitride polymers are stable in its crystalline hexagonal form where it has a two-dimensional layered structure similar to graphene. There are covalent bonds formed between boron and nitrogen atoms, yet the layers are held together by weak van der Waals interactions, in which the boron atoms are eclipsed over the nitrogen.

Figure 2. Surface-mediated 2D polymerization scheme of the tetrafunctional porphyrin monomer. Porphyrn 2.png
Figure 2. Surface-mediated 2D polymerization scheme of the tetrafunctional porphyrin monomer.

Two dimensional covalent organic frameworks (COFs) are one type of microporous coordination polymer that can be fabricated in the 2D plane. The dimensionality and topology of the 2D COFs result from both the shape of the monomers and the relative and dimensional orientations of their reactive groups. These materials contain desirable properties in fields of materials chemistry including thermal stability, tunable porosity, high specific surface area, and the low density of organic material. By careful selection of organic building units, long range π-orbital overlap parallel to the stacking direction of certain organic frameworks can be achieved. [7]

2D polymerization under thermodynamic control (top) versus kinetic control (bottom). Solid black lines represent covalent bond formation 2D polymerization under thermodynamic control (top) versus kinetic control (bottom). Solid black lines represent covalent bond formation..png
2D polymerization under thermodynamic control (top) versus kinetic control (bottom). Solid black lines represent covalent bond formation
Synthetic scheme of covalent organic framework using boronic acid and hexahydroxytriphenylene. COF2.png
Synthetic scheme of covalent organic framework using boronic acid and hexahydroxytriphenylene.

Many covalent organic frameworks derive their topology from the directionality of the covalent linkages, thus small changes in organic linkers can dramatically affect their mechanical and electronic properties. [7] Even small changes in their structure can induce dramatic changes in stacking behavior of molecular semiconductors.

Porphyrins are an additional class of conjugated, heterocyclic macrocycles. Control of monomer assembly through covalent assembly has also been demonstrated using covalent interactions with porphyrins. Upon thermal activation of porphyrin building blocks, covalent bonds form to create a conductive polymer, a versatile route for bottom-up construction of electronic circuits has been demonstrated. [9]

COF synthesis

Boronate ester equilibria used to prepare various 2D COFs EQUILBRIA.png
Boronate ester equilibria used to prepare various 2D COFs

It is possible to synthesize COFs using both dynamic covalent and non-covalent chemistry. The kinetic approach involves a stepwise process of polymerizing pre-assembled 2D-monomer while thermodynamic control exploits reversible covalent chemistry to allow simultaneous monomer assembly and polymerization. Under thermodynamic control, bond formation and crystallization also occur simultaneously. Covalent organic frameworks formed by dynamic covalent bond formation involves chemical reactions carried out reversibly under conditions of equilibrium control. [7] Because the formation of COFs in dynamic covalent formation occurs under thermodynamic control, product distributions depend only on the relative stabilities of the final products. Covalent assembly to form 2D COFs has been previously done using boronate esters from catechol acetonides in the presence of a lewis acid (BF3*OEt2). [10]

2D polymerization under kinetic control relies on non-covalent interactions and monomer assembly prior to bond formation. The monomers can be held together in a pre-organized position by non-covalent interactions, such as hydrogen bonding or van der Waals. [11]

Coordination polymers

Synthetic scheme for metal organic framework (MOF) using hexahydroxytriphenylene (HHTP) and Cu(II) metal. MOF2.png
Synthetic scheme for metal organic framework (MOF) using hexahydroxytriphenylene (HHTP) and Cu(II) metal.

Metal organic frameworks

Self-assembly can also be observed in the presence of organic ligands and various metals centers through coordinative bonds or supramolecular interactions. Molecular self- assembly involves the association by many weak, reversible interactions to obtain a final structure that represents a thermodynamic minimum. [12] A class of coordination polymers, known also as metal-organic frameworks (MOFs), are metal-ligand compounds that extend "infinitely" into one, two or three dimensions. [13]

MOF synthesis

Availability of modular metal centers and organic building blocks generate wide diversity in synthetic versatility. Their applications range from industrial use [14] to chemiresistive sensors. [15] The ordered structure of the frame is largely determined by the coordination geometry of the metal and directionality of functional groups upon the organic linker. Consequently, MOFs contain highly defined pore dimensions when compared with conventional amorphous nanoporous materials and polymers. [16] Reticular Synthesis of MOFs is a term that has been recently coined to describe the bottom-up method of assembling cautiously designed rigid molecular building blocks into prearranged structures held together by strong chemical bonds. [17] The synthesis of two-dimensional MOFs begins with the knowledge of a target "blueprint" or a network, followed by identification of the required building blocks for its assembly. [18]

By interchanging metal centers and organic ligands, one can fine-tune electronic and magnetic properties observed in MOFs. There have been recent efforts synthesize conductive MOFs using triyphenylene linkers. [19] Additionally, MOFs have been utilized as reversible chemiresistive sensors. [13] [15]

Supramolecular polymers

Supramolecular aggregates of (CA*M) cyanuric acid (CA) and melamine (M). C**AM.png
Supramolecular aggregates of (CA*M) cyanuric acid (CA) and melamine (M).

Supramolecular assembly requires non-covalent interactions directing the formation of 2D polymers by relying on electrostatic interactions such as hydrogen bonding and van der Waals forces. To design artificial assemblies capable of high selectivity requires correct manipulation of energetic and stereochemical features of non-covalent forces. [11] Some benefits of non-covalent interactions is their reversible nature and response to external factors such as temperature and concentration. [20] The mechanism of non-covalent polymerization in supramolecular chemistry is highly dependent on the interactions during the self-assembly process. The degree of polymerization depends highly on temperature and concentration. The mechanisms may be divided into three categories: isodesmic, ring-chain, and cooperative. [20]

Self-assembly of a PTCDI-melamine supramolecular network. Dotted lines represent the stabilizing hydrogen bonds between the molecules. Supra AM 2****.png
Self-assembly of a PTCDI–melamine supramolecular network. Dotted lines represent the stabilizing hydrogen bonds between the molecules.

One example of isodesmic associations in supramolecular aggregates is seen in Figure 7, (CA*M) cyanuric acid (CA) and melamine (M) interactions and assembly through hydrogen bonding. [12] Hydrogen bonding has been used to guide assembly of molecules into two-dimensional networks, that can then serve as new surface templates and offer an array of pores of sufficient capacity to accommodate large guest molecules. [21] An example of utilizing surface structures through non-covalent assembly uses adsorbed monolayers to create binding sites for target molecules through hydrogen bonding interactions. Hydrogen bonding is used to guide the assembly of two different molecules into a 2D honeycomb porous network under ultra high vacuum seen in figure 8. [21] 2D polymers based on DNA have been reported [22]

Characterization

2DPs as two dimensional sheet macromolecules have a crystal lattice, that is they consist of monomer units that repeat in two dimensions. Therefore, a clear diffraction pattern from their crystal lattice should be observed as a proof of crystallinity. The internal periodicity is supported by electron microscopy imaging, electron diffraction and Raman-spectroscopic analysis.

2DPs should in principle also be obtainable by, e.g., an interfacial approach whereby proving the internal structure, however, is more challenging and has not yet been achieved. [23] [24] [25]

In 2014 a 2DP was reported synthesised from a trifunctional photoreactive anthracene derived monomer, preorganised in a lamellar crystal and photopolymerised in a [4+4]cycloaddition. [26] Another reported 2DP also involved an anthracene-derived monomer [27]

Applications

2DPs are expected to be superb membrane materials because of their defined pore sizes. Furthermore, they can serve as ultrasensitive pressure sensors, as precisely defined catalyst supports, for surface coatings and patterning, as ultrathin support for cryo-TEM, and many other applications.

Since 2D polymers provide an availability of large surface area and uniformity in sheets, they also found useful applications in areas such as selective gas adsorption and separation. [7] Metal organic frameworks have become popular recently due to the variability of structures and topology which provide tunable pore structures and electronic properties. There are also ongoing methods for creation of nanocrystals of MOFs and their incorporation into nanodevices. [28] Additionally, metal-organic surfaces have been synthesized with cobalt dithionlene catalysts for efficient hydrogen production through reduction of water as an important strategy for fields of renewable energy. [29]

The fabrication of 2D organic frameworks, have also synthesized two-dimensional, porous covalent organic frameworks to be used as storage media for hydrogen, methane and carbon dioxide in clean energy applications. [13]

History

First attempts to synthesize 2DPs date back to the 1930s when Gee reported interfacial polymerizations at the air/water interface in which a monolayer of an unsaturated fatty acid derivative was laterally polymerized to give a 2D cross-linked material. [30] [31] [32] Since then a number of important attempts were reported in terms of cross-linking polymerization of monomers confined to layered templates or various interfaces. [1] [33] These approaches provide easy accesses to sheet-like polymers. However, the sheets' internal network structures are intrinsically irregular and the term "repeat unit" is not applicable (See for example: [34] [35] [36] ). In organic chemistry, creation of 2D periodic network structures has been a dream for decades. [37] Another noteworthy approach is "on-surface polymerization" [38] [39] whereby 2DPs with lateral dimensions not exceeding some tens of nanometers were reported. [40] [41] [42] Laminar crystals are readily available, each layer of which can ideally be regarded as latent 2DP. There have been a number of attempts to isolate the individual layers by exfoliation techniques (see for example: [43] [44] [45] ).

Related Research Articles

Supramolecular chemistry refers to the branch of chemistry concerning chemical systems composed of a discrete number of molecules. The strength of the forces responsible for spatial organization of the system range from weak intermolecular forces, electrostatic charge, or hydrogen bonding to strong covalent bonding, provided that the electronic coupling strength remains small relative to the energy parameters of the component. While traditional chemistry concentrates on the covalent bond, supramolecular chemistry examines the weaker and reversible non-covalent interactions between molecules. These forces include hydrogen bonding, metal coordination, hydrophobic forces, van der Waals forces, pi–pi interactions and electrostatic effects.

<span class="mw-page-title-main">Polycatenane</span> Mechanically interlocked molecular architecture

A polycatenane is a chemical substance that, like polymers, is chemically constituted by a large number of units. These units are made up of concatenated rings into a chain-like structure.

<span class="mw-page-title-main">Supramolecular assembly</span> Complex of molecules non-covalently bound together

In chemistry, a supramolecular assembly is a complex of molecules held together by noncovalent bonds. While a supramolecular assembly can be simply composed of two molecules, or a defined number of stoichiometrically interacting molecules within a quaternary complex, it is more often used to denote larger complexes composed of indefinite numbers of molecules that form sphere-, rod-, or sheet-like species. Colloids, liquid crystals, biomolecular condensates, micelles, liposomes and biological membranes are examples of supramolecular assemblies, and their realm of study is known as supramolecular chemistry. The dimensions of supramolecular assemblies can range from nanometers to micrometers. Thus they allow access to nanoscale objects using a bottom-up approach in far fewer steps than a single molecule of similar dimensions.

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">Coordination polymer</span> Polymer consisting of repeating units of a coordination complex

A coordination polymer is an inorganic or organometallic polymer structure containing metal cation centers linked by ligands. More formally a coordination polymer is a coordination compound with repeating coordination entities extending in 1, 2, or 3 dimensions.

<span class="mw-page-title-main">Crystal engineering</span> Designing solid structures with tailored properties

Crystal engineering studies the design and synthesis of solid-state structures with desired properties through deliberate control of intermolecular interactions. It is an interdisciplinary academic field, bridging solid-state and supramolecular chemistry.

In polymer chemistry and materials science, the term "polymer" refers to large molecules whose structure is composed of multiple repeating units. Supramolecular polymers are a new category of polymers that can potentially be used for material applications beyond the limits of conventional polymers. By definition, supramolecular polymers are polymeric arrays of monomeric units that are connected by reversible and highly directional secondary interactions–that is, non-covalent bonds. These non-covalent interactions include van der Waals interactions, hydrogen bonding, Coulomb or ionic interactions, π-π stacking, metal coordination, halogen bonding, chalcogen bonding, and host–guest interaction. The direction and strength of the interactions are precisely tuned so that the array of molecules behaves as a polymer in dilute and concentrated solution, as well as in the bulk.

<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">Molecular self-assembly</span> Movement of molecules into a defined arrangement without outside influence

In chemistry and materials science, molecular self-assembly is the process by which molecules adopt a defined arrangement without guidance or management from an outside source. There are two types of self-assembly: intermolecular and intramolecular. Commonly, the term molecular self-assembly refers to the former, while the latter is more commonly called folding.

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">Conjugated microporous polymer</span>

Conjugated microporous polymers (CMPs) are a sub-class of porous materials that are related to structures such as zeolites, metal-organic frameworks, and covalent organic frameworks, but are amorphous in nature, rather than crystalline. CMPs are also a sub-class of conjugated polymers and possess many of the same properties such as conductivity, mechanical rigidity, and insolubility. CMPs are created through the linking of building blocks in a π-conjugated fashion and possess 3-D networks. Conjugation extends through the system of CMPs and lends conductive properties to CMPs. Building blocks of CMPs are attractive in that the blocks possess broad diversity in the π units that can be used and allow for tuning and optimization of the skeleton and subsequently the properties of CMPs. Most building blocks have rigid components such as alkynes that cause the microporosity. CMPs have applications in gas storage, heterogeneous catalysis, light emitting, light harvesting, and electric energy storage.

Coordination cages are three-dimensional ordered structures in solution that act as hosts in host–guest chemistry. They are self-assembled in solution from organometallic precursors, and often rely solely on noncovalent interactions rather than covalent bonds. Coordinate bonds are useful in such supramolecular self-assembly because of their versatile geometries. However, there is controversy over calling coordinate bonds noncovalent, as they are typically strong bonds and have covalent character. The combination of a coordination cage and a guest is a type of inclusion compound. Coordination complexes can be used as "nano-laboratories" for synthesis, and to isolate interesting intermediates. The inclusion complexes of a guest inside a coordination cage show intriguing chemistry as well; often, the properties of the cage will change depending on the guest. Coordination complexes are molecular moieties, so they are distinct from clathrates and metal-organic frameworks.

<span class="mw-page-title-main">Kim Kimoon</span> South Korean chemist

Kim Kimoon is a South Korean chemist and professor in the Department of Chemistry at Pohang University of Science and Technology (POSTECH). He is the first and current director of the Center for Self-assembly and Complexity at the Institute for Basic Science. Kim has authored or coauthored 300 papers which have been cited more than 30,000 times and he holds a number of patents. His work has been published in Nature, Nature Chemistry, Angewandte Chemie, and JACS, among others. He has been a Clarivate Analytics Highly Cited Researcher in the field of chemistry in 2014, 2015, 2016.

In materials science, the term single-layer materials or 2D materials refers to crystalline solids consisting of a single layer of atoms. These materials are promising for some applications but remain the focus of research. Single-layer materials derived from single elements generally carry the -ene suffix in their names, e.g. graphene. Single-layer materials that are compounds of two or more elements have -ane or -ide suffixes. 2D materials can generally be categorized as either 2D allotropes of various elements or as compounds.

Jürgen P. Rabe is a German physicist and nanoscientist.

<span class="mw-page-title-main">Dmitrii Perepichka</span>

Dmitrii "Dima" F. Perepichka is the Chair of Chemistry Department and Sir William C. MacDonald Chair Professor in Chemistry at McGill University. His research interest are primarily in the area of organic electronics. He has contributed in the understanding of structural electronics effects of organic conjugated materials at molecular, supramolecular, and macromolecular levels via the study of small molecules, supramolecular (co-)assemblies, polymers, covalent organic frameworks, and on-surface assemblies/polymers.

<span class="mw-page-title-main">Macromolecular cages</span>

Macromolecular cages have three dimensional chambers surrounded by a molecular framework. Macromolecular cage architectures come in various sizes ranging from 1-50 nm and have varying topologies as well as functions. They can be synthesized through covalent bonding or self-assembly through non-covalent interactions. Most macromolecular cages that are formed through self-assembly are sensitive to pH, temperature, and solvent polarity.

Light harvesting materials harvest solar energy that can then be converted into chemical energy through photochemical processes. Synthetic light harvesting materials are inspired by photosynthetic biological systems such as light harvesting complexes and pigments that are present in plants and some photosynthetic bacteria. The dynamic and efficient antenna complexes that are present in photosynthetic organisms has inspired the design of synthetic light harvesting materials that mimic light harvesting machinery in biological systems. Examples of synthetic light harvesting materials are dendrimers, porphyrin arrays and assemblies, organic gels, biosynthetic and synthetic peptides, organic-inorganic hybrid materials, and semiconductor materials. Synthetic and biosynthetic light harvesting materials have applications in photovoltaics, photocatalysis, and photopolymerization.

<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.

<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.

References

  1. 1 2 Sakamoto, J.; van Heijst, J.; Lukin, O.; Schlüter, A. D. (2009). "Two-dimensional polymers: just a dream of synthetic chemists?". Angew. Chem. Int. Ed. 48 (6): 1030–69. doi:10.1002/anie.200801863. PMID   19130514.
  2. Colson, John W.; Dichtel, William R. (2013). "Rationally synthesized two-dimensional polymers". Nature Chemistry. 5 (6): 453–465. Bibcode:2013NatCh...5..453C. doi:10.1038/nchem.1628. PMID   23695626.
  3. Staudinger, H. (1920). "Über Polymerisation". Ber. Dtsch. Chem. Ges. 53 (6): 1073. doi:10.1002/cber.19200530627.
  4. Staudinger, H.; Fritschi, J. (1922). "Uber Isopren und Kautschuk. 5. Mitteilung. Uber die Hydrierung des Kautschuks und uber seine Konstitution". Helv. Chim. Acta. 5 (5): 785. doi:10.1002/hlca.19220050517.
  5. Mark, H. F. (1980). "Aus den fruhen Tagen der Makromolekularen Chemie". Naturwissenschaften. 67 (10): 477. Bibcode:1980NW.....67..477M. doi:10.1007/bf01047626. S2CID   27327793.
  6. Ringsdorf, H. (2004). "Hermann Staudinger and the Future of Polymer Research Jubilees—Beloved Occasions for Cultural Piety". Angew. Chem. Int. Ed. 43 (9): 1064–76. doi:10.1002/anie.200330071. PMID   14983439.
  7. 1 2 3 4 5 Colson, John W.; Dichtel, William R. (2013-06-01). "Rationally synthesized two-dimensional polymers". Nature Chemistry. 5 (6): 453–465. Bibcode:2013NatCh...5..453C. doi:10.1038/nchem.1628. ISSN   1755-4330. PMID   23695626.
  8. Sakamoto, Junji; van Heijst, Jeroen; Lukin, Oleg; Schlüter, A. Dieter (2009-01-26). "Two-Dimensional Polymers: Just a Dream of Synthetic Chemists?". Angewandte Chemie International Edition. 48 (6): 1030–1069. doi:10.1002/anie.200801863. ISSN   1521-3773. PMID   19130514.
  9. Grill, Leonhard; Dyer, Matthew; Lafferentz, Leif; Persson, Mats; Peters, Maike V.; Hecht, Stefan (2007-11-01). "Nano-architectures by covalent assembly of molecular building blocks". Nature Nanotechnology. 2 (11): 687–691. Bibcode:2007NatNa...2..687G. doi:10.1038/nnano.2007.346. ISSN   1748-3387. PMID   18654406.
  10. Côté, Adrien P.; Benin, Annabelle I.; Ockwig, Nathan W.; O'Keeffe, Michael; Matzger, Adam J.; Yaghi, Omar M. (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.
  11. 1 2 Lehn, Supramolecular Chemistry (1995). Supramolecular Chemistry. VCH.
  12. 1 2 Whitesides, George M.; Grzybowski, Bartosz (2002-03-29). "Self-Assembly at All Scales". Science. 295 (5564): 2418–2421. Bibcode:2002Sci...295.2418W. doi:10.1126/science.1070821. ISSN   0036-8075. PMID   11923529. S2CID   40684317.
  13. 1 2 3 Janiak, Christoph (2003). "Engineering coordination polymers towards applications". Dalton Transactions (14): 2781–2804. doi:10.1039/B305705B.
  14. Furukawa, Hiroyasu; Yaghi, Omar M. (2009-07-01). "Storage of Hydrogen, Methane, and Carbon Dioxide in Highly Porous Covalent Organic Frameworks for Clean Energy Applications". Journal of the American Chemical Society. 131 (25): 8875–8883. doi:10.1021/ja9015765. ISSN   0002-7863. PMID   19496589.
  15. 1 2 Campbell, Michael G.; Sheberla, Dennis; Liu, Sophie F.; Swager, Timothy M.; Dincă, Mircea (2015-03-27). "Cu3(hexaiminotriphenylene)2: An Electrically Conductive 2D Metal–Organic Framework for Chemiresistive Sensing". Angewandte Chemie International Edition. 54 (14): 4349–4352. doi:10.1002/anie.201411854. ISSN   1521-3773. PMID   25678397.
  16. Stavila, V.; Talin, A. A.; Allendorf, M. D. (2014-07-22). "MOF-based electronic and opto-electronic devices". Chemical Society Reviews. 43 (16): 5994–6010. doi:10.1039/C4CS00096J. PMID   24802763.
  17. 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" (PDF). Nature. 423 (6941): 705–714. doi:10.1038/nature01650. hdl: 2027.42/62718 . ISSN   0028-0836. PMID   12802325. S2CID   4300639.
  18. Liu, Jiewei; Chen, Lianfen; Cui, Hao; Zhang, Jianyong; Zhang, Li; Su, Cheng-Yong (2014-07-22). "Applications of metal–organic frameworks in heterogeneous supramolecular catalysis". Chemical Society Reviews. 43 (16): 6011–6061. doi:10.1039/C4CS00094C. PMID   24871268.
  19. Hmadeh, Mohamad; Lu, Zheng; Liu, Zheng; Gándara, Felipe; Furukawa, Hiroyasu; Wan, Shun; Augustyn, Veronica; Chang, Rui; Liao, Lei (2012-09-25). "New Porous Crystals of Extended Metal-Catecholates". Chemistry of Materials. 24 (18): 3511–3513. doi:10.1021/cm301194a. ISSN   0897-4756.
  20. 1 2 Harada, Akira (2012). Supramolecular Polymer Chemistry. Wiley BCH.
  21. 1 2 Theobald, James A.; Oxtoby, Neil S.; Phillips, Michael A.; Champness, Neil R.; Beton, Peter H. (2003-08-28). "Controlling molecular deposition and layer structure with supramolecular surface assemblies". Nature. 424 (6952): 1029–1031. Bibcode:2003Natur.424.1029T. doi:10.1038/nature01915. ISSN   0028-0836. PMID   12944962. S2CID   4373112.
  22. Yu, Hao; Alexander, Duncan T. L.; Aschauer, Ulrich; Häner, Robert (2017). "Synthesis of Responsive Two-Dimensional Polymers via Self-Assembled DNA Networks". Angewandte Chemie International Edition. 56 (18): 5040–5044. doi:10.1002/anie.201701342. PMID   28370933.
  23. Payamyar, P.; Kaja, K.; Ruiz-Vargas, C.; Stemmer, A.; Murray, D. J; Johnson, C. J; King, B. T.; Schiffmann, F.; VandeVondele, J.; Renn, A.; Götzinger, S.; Ceroni, P.; Schütz, A.; Lee, L.-T.; Zheng, Z.; Sakamoto, J.; Schlüter, A. D. (2014). "Synthesis of a Covalent Monolayer Sheet by Photochemical Anthracene Dimerization at the Air/Water Interface and its Mechanical Characterization by AFM Indentation". Adv. Mater. 26 (13): 2052–2058. doi:10.1002/adma.201304705. PMID   24347495.
  24. For a report aiming at synthesis of 2D copolymers see: Payamyar, P.; Servalli, M.; Hungerland, T.; Schütz, A. P.; Zheng, Z.; Borgschulte, A.; Schlüter, A. D. (2015). "Approaching Two-Dimensional Copolymers: Photoirradiation of Anthracene- and Diaza-Anthracene-Bearing Monomers in Langmuir Monolayers". Macromol. Rapid Commun. 36 (2): 151–158. doi:10.1002/marc.201400569. PMID   25475710.
  25. For another possible case on metal organic frameworks see: Bauer, T.; Zheng, Z.; Renn, A.; Enning, R.; Stemmer, A.; Sakamoto, J.; Schlüter, A. D. (2011). "Synthesis of Free-Standing, Monolayered Organometallic Sheets at the Air/Water Interface". Angew. Chem. Int. Ed. 50 (34): 7879–84. doi:10.1002/anie.201100669. PMID   21726022.
  26. Murray, Daniel J.; Wulftange, William J.; Catalano, Vincent J.; King, Benjamin T. (2014). "A nanoporous two-dimensional polymer by single-crystal-to-single-crystal photopolymerization Patrick Kissel". Nature Chemistry. 6 (9): 774–778. Bibcode:2014NatCh...6..774K. doi:10.1038/nchem.2008. PMID   25143211.
  27. Kory, Max J.; Wörle, Michael; Weber, Thomas; Payamyar, Payam; van de Poll, Stan W.; Dshemuchadse, Julia; Trapp, Nils; Schlüter, A. Dieter (2014). "Gram-scale synthesis of two-dimensional polymer crystals and their structure analysis by X-ray diffraction". Nature Chemistry. 6 (9): 779–784. Bibcode:2014NatCh...6..779K. doi:10.1038/nchem.2007. PMID   25143212.
  28. Makiura, Rie; Konovalov, Oleg (2013-08-26). "Interfacial growth of large-area single-layer metal-organic framework nanosheets". Scientific Reports. 3: 2506. Bibcode:2013NatSR...3E2506M. doi:10.1038/srep02506. PMC   3752618 . PMID   23974345.
  29. Clough, Andrew J.; Yoo, Joseph W.; Mecklenburg, Matthew H.; Marinescu, Smaranda C. (2015-01-14). "Two-Dimensional Metal–Organic Surfaces for Efficient Hydrogen Evolution from Water". Journal of the American Chemical Society. 137 (1): 118–121. doi:10.1021/ja5116937. ISSN   0002-7863. PMID   25525864.
  30. Gee, G.; Rideal, E. K. (1935). "Reactions in Monolayers of Drying Oils. I. The Oxidation of the Maleic Anhydride Compound of Formula-Elaeostearin". Proc. R. Soc. Lond. A. 153 (878): 116. Bibcode:1935RSPSA.153..116G. doi: 10.1098/rspa.1935.0224 .
  31. Gee, G. (1936). "Polymerisation in monolayers". Trans. Faraday Soc. 32: 187. doi:10.1039/tf9363200187.
  32. Gee, G. (1937). "156. Catalysed polymerisation in monolayers of drying oils". J. Chem. Soc. 1937: 772. doi:10.1039/jr9370000772.
  33. J. Sakamoto and A. D. Schlüter in Synthesis of Polymers, New Structures and Methods (eds. A. D. Schlüter, C. J. Hawker, J. Sakamoto), Wiley-VCH, Weinheim, Germany, 2012, Volume 2, pg 841.
  34. Asakusa, S.; Okada, H.; Kunitake, T. (1991). "Template synthesis of two-dimensional network of crosslinked acrylate polymer in a cast multibilayer film". J. Am. Chem. Soc. 113 (5): 1749. doi:10.1021/ja00005a044.
  35. Stupp, S. I.; Son, S.; Lin, H. C.; Li, L. S. (1993). "Synthesis of Two-Dimensional Polymers". Science. 259 (5091): 59–63. Bibcode:1993Sci...259...59S. doi:10.1126/science.259.5091.59. PMID   17757473. S2CID   34503747.
  36. Stupp, S. I.; Son, S.; Li, L. S.; Lin, H. C.; Keser, M. (1995). "Bulk Synthesis of Two-Dimensional Polymers: The Molecular Recognition Approach". J. Am. Chem. Soc. 117 (19): 5212. doi:10.1021/ja00124a005.
  37. Diederich, F. (1994). "Carbon scaffolding: building acetylenic all-carbon and carbon-rich compounds". Nature. 369 (6477): 199. Bibcode:1994Natur.369..199D. doi:10.1038/369199a0. S2CID   4330198.
  38. Perepichka, D. F.; Rosei, F. (2009). "CHEMISTRY: Extending Polymer Conjugation into the Second Dimension". Science. 323 (5911): 216–7. doi:10.1126/science.1165429. PMID   19131618. S2CID   206516046.
  39. Grill, L.; Dyer, M.; Lafferentz, L.; Persson, M.; Peters, M. V.; Hecht, S. (2007). "Nano-architectures by covalent assembly of molecular building blocks". Nat. Nanotechnol. 2 (11): 687–91. Bibcode:2007NatNa...2..687G. doi:10.1038/nnano.2007.346. PMID   18654406.
  40. Bieri, M.; Treier, M.; Cai, J.; Ait-Mansour, K.; Ruffieux, P.; Groning, O.; Groning, P.; Kastler, M.; Rieger, R.; Feng, X.; Müllen, K.; Fasel, R. (2009). "Porous graphenes: two-dimensional polymer synthesis with atomic precision". Chem. Commun. 45 (45): 6919–21. doi:10.1039/b915190g. PMID   19904347.
  41. Bieri, M.; Blankenburg, S.; Kivala, M.; Pignedoli, C. A.; Ruffieux, P.; Müllen, K.; Fasel, R. (2011). "Surface-supported 2D heterotriangulene polymers". Chem. Commun. 47 (37): 10239–41. doi:10.1039/c1cc12490k. PMID   21850288.
  42. Abel, M.; Clair, S.; Ourdjini, O.; Mossoyan, M.; Porte, L. (2011). "Single Layer of Polymeric Fe-Phthalocyanine: An Organometallic Sheet on Metal and Thin Insulating Film". J. Am. Chem. Soc. 133 (5): 1203–5. doi:10.1021/ja108628r. PMID   21192107.
  43. Hernandez, Y.; Nicolosi, V.; Lotya, M.; Blighe, F. M.; Sun, Z.; De, S.; McGovern, I. T.; Holland, B.; Byrne, M.; Gun'ko, Y.; Boland, J.; Niraj, P.; Duesberg, G.; Krishnamurti, S.; Goodhue, R.; Hutchison, J.; Scardaci, V.; Ferrari, A. C.; Coleman, J. N. (2008). "High-yield production of graphene by liquid-phase exfoliation of graphite". Nat. Nanotechnol. 3 (9): 563–8. arXiv: 0805.2850 . Bibcode:2008NatNa...3..563H. doi:10.1038/nnano.2008.215. PMID   18772919. S2CID   205443620.
  44. Ma, R.; Sasaki, T. (2010). "Nanosheets of Oxides and Hydroxides: Ultimate 2D Charge-Bearing Functional Crystallites". Adv. Mater. 22 (45): 5082–104. doi:10.1002/adma.201001722. PMID   20925100.
  45. Berlanga, I.; Ruiz-Gonzalez, M. L.; Gonzalez-Calbet, J. M.; Fierro, J. L. G.; Mas-Balleste, R.; Zamora, F. (2011). "Delamination of Layered Covalent Organic Frameworks". Small. 7 (9): 1207–11. doi:10.1002/smll.201002264. PMID   21491587.