Open access peer-reviewed chapter

Coordination Polymer Frameworks for Next Generation Optoelectronic Devices

Written By

Hemali Rathnayake and Sheeba Dawood

Submitted: 10 June 2020 Reviewed: 05 October 2020 Published: 28 October 2020

DOI: 10.5772/intechopen.94335

From the Edited Volume

Optoelectronics

Edited by Mike Haidar Shahine

Chapter metrics overview

736 Chapter Downloads

View Full Metrics

Abstract

Metal–organic frameworks (MOFs), which belong to a sub-class of coordination polymers, have been significantly studied in the fields of gas storage and separation over the last two decades. There are 80,000 synthetically known MOFs in the current database with known crystal structures and some physical properties. However, recently, numerous functional MOFs have been exploited to use in the optoelectronic field owing to some unique properties of MOFs with enhanced luminescence, electrical, and chemical stability. This book chapter provides a comprehensive summary of MOFs chemistry, isoreticular synthesis, and properties of isoreticular MOFs, synthesis advancements to tailor optical and electrical properties. The chapter mainly discusses the research advancement made towards investigating optoelectronic properties of IRMOFs. We also discuss the future prospective of MOFs for electronic devices with a proposed roadmap suggested by us. We believe that the MOFs-device roadmap should be one meaningful way to reach MOFs milestones for optoelectronic devices, particularly providing the potential roadmap to MOF-based field-effect transistors, photovoltaics, thermoelectric devices, and solid-state electrolytes and lithium ion battery components. It may enable MOFs to be performed in their best, as well as allowing the necessary integration with other materials to fabricate fully functional devices in the next few decades.

Keywords

  • MOF
  • coordination polymers
  • optoelectronics
  • isoreticular MOF
  • semiconducting MOFs

1. Introduction

Over last few decades, crystalline microporous materials, from zeolites, to coordination polymers and its subclass, metal organic frameworks (MOFs) have gained enormous attention in the scientific community due to their structural versatility and tailorable properties like nanoscale porosity, high surface area, and functional density [1, 2]. Metal organic frameworks have evolved in last few years as a revolutionary material that are self-assembled nanostructure [3, 4] built from metal ions and organic ligands. The first MOF, MOF-5 or IRMOF-1 (Zn4O(BDC)3) reported by Omar M. Yaghi was used in gas adsorption applications accounting to its high surface area of 2900 m2/g [5, 6]. To date, 80,000 MOFs [7] have been reported owing to its diverse structure, compositions, tunable porosity, specific surface area, [8] ease of functionalization, unsaturated metal sites [9] and biocompatibility [10] . As a result, MOFs were used in a wide range of applications such as gas storage and separation, drug delivery and storage, chemical separation, sensing, catalysis, and bio-imaging [3, 7, 11, 12, 13]. In terms of structural orientation, the coordination bonding between a metal ion and organic ligand results in the formation of extended networks of one, two, and three-dimensional framework with potential voids [6, 14]. The coordination bonding facilitated through a suitable molecular approach, involving reticular synthesis, provides the flexibility to alter the pore size and transform its structure, targeting specific applications. Thus, utilizing the advantage of various combinations of metal-ligands and interaction of metal-ligands, MOFs are ideal candidates in the field of material science, offering an attractive property of structural tunability, providing a pathway to introduce and tailor intrinsic characteristics, such as optical, electrical, and magnetic properties.

There has been a growing interest exploring MOF as emerging semiconducting materials to meet the current demand in the electronic devices [15]. In particular, the electronic characteristics such as electrical, optical, and magnetic properties of MOFs have become an interesting topic of research attributing to their applications in microelectronic and optical devices. The implementation of MOFs in the electronic industry was first reported by Allendorf and co-workers [16]. MOF-5 with Zn4O metal nodes and orthogonally interconnected six units of terephthalate is the most-studied MOF as a semiconductor. In 2007, Garcia and co-workers reported on the semiconducting behavior of MOF-5 synthesized at room temperature, with a bandgap of 3.4 eV [17]. Since then, intense research has been carried out to develop MOFs with semiconducting properties, opening new research domains for the scientific community in nanoscience.

The presence of narrow band gap structure either direct or indirect and charge mobility contribute to the semiconducting behavior of MOFs. To design MOFs with semiconducting behavior, significant amount of research is ongoing to identify the general structural requirements for enhancing the orbital overlapping between the building components. The main advantage of MOFs is the ability to tune the crystalline structure and functionality through phenomenal conceptual approaches such as rational designing and synthetic flexibility. In reticular chemistry, which is also known as rational designing, the coordination bonding between metal node and organic ligand provides an understanding of atomic positions precisely contributing to determine the fundamental structure–property relationships. Thus, the crystalline structure of MOFs consists of self-assembled ordered nanostructure with defined organized spatial space that is constructed via coordination chemistry between the building components.

Moreover, the sub-angstrom knowledge of atomic positions helps to eliminate any disorder in the structure that contributes to poor mobility in the structure. Considering synthetic flexibility, the electronic properties of MOFs could be tailored, resulting in potential applications such as a photovoltaic device tuned for solar cells, electroluminescent devices, field effect transistors, spintronic devices, and sensors. These developments have led many researchers to explore electrical, magnetic, and optical properties of MOFs [15, 18, 19]. However, the electrical properties of MOFs and integration of them in micro-electronic devices is still at an early stage and remain under research when compared to other types of existing conducting materials [4, 15] due to their insulating character. Although MOFs possess the properties of both organic and inorganic counterparts, they behave as electrical insulators or poor electrical conductors due to the poor overlapping between the π-orbitals of organic ligands and d-orbitals of the metal ion [20]. Yet, MOFs serving as an interface between (inorganic) hard and (organic) soft materials provide an opportunity for adapting various structure–property relationships that is related to wide range of parameters such as choice of metal ion, organic linker, and molecular designing approach. In general, the structure–property relationship in MOFs is a consequence of cooperative mechanism, i.e. the interaction between the metal and ligand, which could be readily identified by taking advantage of the knowledge of their detailed atomic structure, enabling fine tuning of their functionalities [7, 11]. According to the literature, Bastian Hoppe and his co-workers reported Cu-2, 3, 6, 7, 10, 11-hexahydroxytriphenylene (Cu3hhtp2-MOF), a copper-based graphene-like framework with inherent electrical conductivity about 0.045 S cm−1 [21]. MOFs with electrical conductivity higher than 0.1 S cm−1 was achieved by Talin and co-workers [22]. Thus, the designing of MOFs with conducting or semiconducting properties is necessary to enhance the sensitivity of electrical or demonstrate a sensing concept; but rarely have MOFs been an integral part of an actual device [23].

The purpose of this chapter is to provide comprehensive discussion on optoelectronic MOFs developed up to date and identify focus points to bring MOFs with optoelectronic properties for the realization of integrating MOFs into actual devices for electronic device applications. We first provide a MOFs chemistry and isoreticular synthesis advancements to make isoreticular MOFs (IRMOFs) with tailored optical and electronic properties. Then we summarize the current state of MOF research relevant to optoelectronics, particularly discussing the synthesis, electronic structure, and photophysical properties of three selected IRMOFs (IRMOF-1, 8, and 10). Finally, we propose a MOFs-device roadmap, focusing on MOF-based field-effect transistors, photovoltaics, thermoelectric devices, and solid-state electrolytes and lithium ion battery components.

Advertisement

2. Chemistry of MOFs

2.1 Dimensional classification and evolution of MOFs

Coordination polymers are organic–inorganic hybrid materials where organic moieties are bonded to metal ion or metal clusters via coordination bonds. The energy of such bonding is usually between 50 and 200 KJ mol−1. Apart from strong coordination bonding, weaker interaction such as hydrogen bonds, van der Waal forces and π-π interactions also influence the formation of coordination polymers. Depending on the geometry, coordination polymers are classified into three subclasses: (1) One-dimensional (1-D) coordination polymers, (2) Two-dimensional (2-D) coordination polymers, and (3) Three-dimensional coordination polymers (Figure 1).

Figure 1.

Dimensional structures of coordination polymers.

The coordination polymer assembled from organic ligand and metal ion into three dimensional hierarchical crystalline structures is often regarded as metal organic framework. Since then, the term coordination polymer and metal organic framework have been used interchangeably. The term MOFs was first introduced by Omar Yaghi in 1995 [4, 9]. The framework of MOFs is either porous or non-porous. However, the porosity of MOFs was reported to be reversible due to various environmental factors (temperature, pressure, light intensity) contributing to the weak intermolecular interactions between building components. Thus, efforts have been made to modulate the strong structural rigidity that could incorporate permanent porosity. Based on this, in 1998 Kitagawa classified MOFs into three categories; 1st, 2nd, and 3rd generation coordinated network. Among three generations of coordinated networks, 3rd generation coordinated networks were defined to have permanent porosity with structural flexibility [10]. This led to numerous applications and implementation of coordinated networks in the gas storage community. The intermolecular interaction between organic ligand and metal ions, choice of building units, crystallization, environment, and guest molecules determine the crystal structural rigidity and dimensionality of MOF’s coordination network. This major advance in the field of coordination polymer depicted that coordinated networks of MOFs could be modified and developed in a highly periodic manner, with a defined understanding of the crystalline structure, porosity, and chemical functionality. Thus, the ability to design and control the arrangement of metal ions with extended organic spaces in three-dimensional fashion led to the origin of the term reticular chemistry which was first introduced by Yaghi and coworkers [4].

With the variability of organic and inorganic components and their interaction, the freedom of modulating the structure of MOFs into highly ordered hierarchical structures with tunable pore volume and adjustable surface area has become feasible that made MOFs stand out compared to the other porous materials. Taking advantage of one of these hallmarks of MOFs i.e. designing of topologically diverse structures with desirable properties has been explored extensively attracting wide range of applications in gas storage, separation, catalysis, sensing and drug delivery [5]. Since 1990s, this area of chemistry has experienced tremendous growth in the field of material science and modern chemistry [4]. The flexibility with geometry, size, and functionality led to the “design” of a large number of MOFs. The organic units are generally ditopic or polytopic organic carboxylates, linked to metal-containing units, such as transition metals (e.g., Cu, Zn, Fe, Co, and Ni), alkaline earth elements (e.g. Sr., Ba), p-block elements (e.g. In, Ga), and actinides (e.g. U, Th) [6]. A major advance in the chemistry of MOFs came in 1999 with the invention of two structures i.e. MOF-5 (IRMOF-1) and HKUST-1 [11] reported by Omar et al. and Chui et al., respectively. Subsequently, in the coming years around 2002, the flexible and non-flexible structures of MIL-88/53 [12] was reported by Ferey et al.

2.2 Reticular chemistry and isoreticular MOFs

The demand for the synthesis of new materials to perform highly specific and cooperative functions has been increasing rapidly in parallel with advanced technology [13, 14]. Recently, the field of metal organic framework has evolved significantly due to its practical and conceptual approach to design and develop the target material. Intrinsically, the reticular chemistry is described as the process of assembly of molecular building blocks held together by strong bonding that pattern into periodic arrays of the ordered net like structures [13, 14, 15, 16]. Some of the advantages of this approach are: (1) Molecular approach, which provides the ability to design and control the structure of frameworks [17]; (2) Bonding in which the strong bonding between the building blocks could impart superior functionalities like thermal and chemical stability into the framework; and (3) Engineered crystallinity, which is based on the type of the interactions (intermolecular or intramolecular) design and synthesis with controlled and desired properties.

After the introduction of the parent MOF, MOF-5, taking advantage of reticular chemistry that includes reticulating metal ions and organic carboxylate, the group of Omar M. Yaghi synthesized a new class of materials called IRMOFs. Thus, the theory of isoreticular chemistry was established in the year 2002 with the development of IRMOFs. These class of materials were developed to improve the surface area and pore volume by incorporation of different topological linkers. In IRMOF, IR stands for isoreticular, which means it is a series of MOFs with the same topology, but different pore size [14, 20, 22, 23]. A series of different IRMOFs share similar pcu topology of IRMOF-n (n = 1–16). As shown in the Figure 2, the pore volume and porosity vary with the variation in the organic linker. Applying the concept of isoreticular chemistry, various kinds of MOFs were developed.

Figure 2.

Crystal structures of IRMOFs-n series [n = 1–16]. The non-interpenetrated structures from (n = 1,2,3,4,5,6,7,8,10,12,14,16). The yellow sphere represents the pore volume. Zn atoms are in green, O in red, C in gray, Br atoms in Orange, and amino groups in blue [17].

2.3 Synthetic advancements of MOFs

The conceptual approach of designing and assembling a metal–organic framework follows reticular synthesis and is based upon identification of how building blocks come together to form a net, or reticulate. The synthesis of MOFs is often regarded as “design” which implies to construct, built, execute, or create according to the target material. The synthesis approach for a new MOF should follow several factors asides from the geometric principles that are considered during its design. Among such factors, by far the most important is the maintenance of the integrity of the building blocks. A great deal of research effort has been demonstrated on the synthesis of a novel organic link and synthesis conditions that are mild enough to maintain the functionality and conformation of organic ligand, yet reactive enough to establish the metal–organic bonds. In situ generation of a desired secondary subunit (SBU) is required carful design of synthetic conditions and must be compatible with the mobilization and preservation of the linking units [24]. Typically, this is achieved by precipitation of the product from a solution of the precursors where solubility is a necessary attribute of the building blocks but is quite often circumvented by using solvothermal techniques [24].

Traditional goal of MOF synthesis is to obtain high quality single crystal for deducing the structure and understand the crystal packing, geometry, and pore volume with respect to the organic ligand’s length. Thus, prior to begin elucidating the concept of reticular synthesis, most early studies were exploratory and early stage synthesis has mainly involved simple, highly soluble precursors, and labile metal ions of the late transition series. The polymerization that leads to make 3D-network of MOFs needs an assembly process where an insoluble entity is quickly formed that precludes recrystallization. Fortunately, it differs in the degree of reversibility of the bond formation event, allowing detachment of incoherently matched monomers followed by reattachment with continued defect free crystal growth. The framework assembly occurs as a single synthetic step, where all of the desired attributes of the target material constructs by the building blocks. This often requires a combinatorial approach, which involves subtle changes in concentration, solvent polarity, pH, or temperature. Any subtle changes in these parameters leads to poorer quality crystals, reduced yields, or the formation of entirely new phases [24].

Augmenting simple crystal growth processes used to grow simple inorganic salts, early efforts of producing highly crystalline MOFs involved the slow introduction of the building blocks to reduce the rate of crystallite nucleation. Methods included slow evaporation of a solution of the precursors, layering of solutions, or slow diffusion of one component solution into another through a membrane or an immobilizing gel [24]. During the nucleation stage, the ligand deprotonation prior to the coordination onto metal ion is catalyzed introducing a volatile amine gradually via vapor diffusion. Just as for many of the polar solvents used, suitable choice of base is necessary to avoid competitive coordination with the organic links for the available metal sites. While in some cases, blocking of metal coordination sites is necessary for the formation of a particular SBU. However, this approach has generally been regarded as leading to low-dimensional structures that are less likely to define an open framework.

With the need for more robust frameworks, having larger pore volumes and higher surface area, introducing bulker, longer length organic linkers are necessary, but greater difficulties in crystal growth were encountered. Thus, later, solvothermal techniques were found to be a convenient solution to overcome this challenge and have largely benefit over often time-consuming methods involving slow coupling of the coordinating species. The typical solvothermal method combines the precursors as dilute solutions in polar solvents such as water, alcohols, acetone or acetonitrile and heated in sealed vessels such as Teflon-lined stainless-steel bombs or glass tubes, generating autogenous pressure. The crystal growth process is enhanced by using mixed solvent systems where the solution polarity and the kinetics of solvent-ligand exchange can tune to achieve rapid crystal growth. It has found that, exposing the growing framework to a variety of space-filling solvent molecules may also be an effective way to stabilize its defect-free construction as they efficiently pack within the defined channels [24]. For deprotonation of the linking molecule alkyl formamides and pyrrolidinones have been particularly useful, as they are also excellent solubilizing agents.

In recent years, modifying the solvothermal method, there are several rapid synthesis methods were proposed by researchers to develop MOF crystals within a short duration of time. Some of the external parameters implemented to develop MOFs include the use of Microwave energy (Microwave synthesis), [25] Ultrasonic waves (Sonochemical synthesis), Mechanical energy (Mechano-chemical synthesis) and electrical energy (Electrochemical synthesis). The synthetic strategies developed up to date to make different type of MOFs are summarized in the Table 1 along with reaction conditions [26]. Additionally, a surfactant driven-templating method, [22] a CO2-expanded liquid route, [27] a post-synthetic method, [28] and an ionic liquid-based method [29] are developed to create hierarchical mesoporous microstructures and thin films of MOFs [25, 27, 28, 29].

Synthesis methodReaction timeTemperature (K)
Slow evaporation7 days to 7 months298
Sonochemical method30–180 mins272–313
Solvothermal method48–96 hours353–453
Mechano-chemical method30 min to 2 hours298
Electrochemical method10–30 mins273–303
Microwave Synthesis4 mins to 4 hours303–373

Table 1.

Synthesis methods developed up to date to make MOFs.

Advertisement

3. Zn4O(L)3-based isoreticular MOFs with cubic topology for optoelectronics

3.1 Road map to electrically conductive MOFs

In the area of MOFs, the main desire is to design MOFs with optoelectronic properties and to optimize the charge transport mechanism suitable for developing electronic devices. Although numerous applications of MOFs with different types of synthesis methods are being investigated, a versatile and scalable synthesis approach for the preparation of MOFs with semiconducting properties for optoelectronic devices are still in the early stage and a little research work so far done towards tailoring MOFs structure–property relationship to use as active materials in optoelectronic devices, such as solar cells, field-effect transistors, and photoluminescence devices.

To introduce MOFs as semiconducting materials, tuning of band gap such as lowering the bandgap or increasing the charge mobility is required. This tunability is again dependent upon the type of interaction i.e. Intermolecular interaction: metal ion and the organic ligand or Intramolecular interaction - π stacking [18]. The two key factors responsible for poor electrical conductivity in MOFs are: (1) the insulating character of organic ligand and (2) due to poor overlapping between the π-orbitals of organic ligand and d-orbitals of metal ions [16]. The common strategies for constructing MOFs with electrical conductivity involves three possible charge pathways.

Pathway 1: A long range of charge transport in this pathway is facilitated through bonds. This mechanism is promoted by interaction between ligand π and metal d orbital [16]. This mechanism is based on the tunneling of electron between the donor and acceptor portions of the framework. Typically, the electrical conductivity in the range 10−7 to 10−10 S cm−1 is considered as insulator. This is caused due to poor overlapping between the metal ion and organic linker as the electronegative nature of oxygen atom in the carboxylate group of the linker is so high that it requires high voltage for tunneling of the electrons [30]. Various MOFs that exhibit conductivity through this mechanism have been reported, of which [[Cu2(6-Hmna) (6-mn)]·NH4]n, a copper-sulfur based MOF constructed from 1,6-Hmna = 6-mercaptonicotinic acid, 6-mn = 6-mercaptonicotinate shows highest electrical conductivity of 10.96 S/cm (Table 2).

MaterialsMechanismConductivity (Scm−1)Charge carrierMobility (cm2V−1 S−1)Ref.
Metals
Cu
Au
Fe
Tunneling6.5 x105
4.1x105
1.0 x 105
e46[31]
Organic polymers
Polyacetylene
Polythiophenes
Rubrene
Charge transfer (electron hopping)10–9
1975
h1–10
4
[31]
TTF-TCNQ
Ni3 (HITP)2
Zn2 (TTFTB)
Through-space700
40
4.0x10–6
h or e48.6
0.2
[16, 31, 32]
Cu3(BTC)2- TCNQ
NU-901-C60
Guest molecule0.07
1x10–3
h[33, 34]
Fe2(DSBDC)
{[Cu2(6-Hmna) (6-mn)] ·NH4]n
Cu [Ni(pdt)2]
Through-bond1x10–6
10.96
1x10–4
[18, 35, 36, 37]

Table 2.

Significant progress in the last few years made towards developing electrically conductive MOFs and their conductive properties compared with conventional metals.

Pathway 2: In this pathway, the charge transport is facilitated through space via π stacked aromatic ligands which was proposed as an alternative to through bond strategy. This mechanism typically promotes electron hopping mechanism by employing electroactive molecules [16, 30]. TTF-TCNQ i.e. tetrathiafulvalene- tetracyano quinomethane is one of the MOFs that demonstrate metallic conductivity (shown in the Table 2) through-space (π-π stacking) mechanism [38]. Recently, TTF-based ligand consisting of benzoate spacers is used to develop Zn based MOF reported by Dincă et al. These MOFs shows columnar stacks of TTF (3.8 Å) with the charge mobility of a magnitude that resembles some best conductive organic polymers [35, 36].

Pathway 3: The other alternative strategy to increase the conductivity of MOFs is via incorporating an appropriate guest molecule within the MOF. These molecules can activate long range delocalization either through bonds or through space or that can inject mobile charge carriers by oxidizing or reducing the organic ligand and metal ions [16, 30] NU-901, a MOF consisting of Zr63-O)43-OH)4 (H2O)4 (OH)4 nodes and tetratopic 1,3,6,8-tetrakis (p- benzoate) pyrene (TBAPy4-) linkers. These materials were chosen for the encapsulation of C60. After installation of C60, the NU-901-C60 shows electrical conductivity higher than that of NU-901 (shown in the Table 2). As per reports, the donor-acceptor interactions between TBAPy4-/C60 contribute to the electrical conductivity of the framework [32, 38].

3.2 Synthesis and optoelectronic properties of IRMOF-1

IRMOF-1, which is commonly known as MOF-5, invented by Yaghi and co-workers [39] in 1996, has become one of mostly studied MOF with promising application in high capacity hydrogen storage [40, 41]. MOF-5, consists of Zn4O units connected by linear 1,4-benzenedicarboxylate units to form a cubic network, having the primitive cubic unit cell. Syntheses demonstrated for MOF-5 in which the starting materials are mixed in solution at ambient temperature. Subsequent addition of triethylamine promotes the deprotonation of the organic linker to precipitate MOF-5. Depending on the addition of base either slowly by diffusion as described in the original synthesis method [39] or rapidly as an aliquot [42] the product can be either single crystal mixtures, which must be mechanically separated, or microcrystalline powders. The ambient temperature synthesis method described above following the fast addition of base, is easy to scale up. However, metal precursor, zinc nitrate poses potential safety concerns, especially for large-scale production. Furthermore, reports of such synthetic conditions have been largely limited to MOF-5 and IRMOF-8 [42, 43, 44].

Later, a rapid, simple, room-temperature high yielding synthesis method was introduced by Yaghi and co-workers that can apply to make a wide range of new MOFs, including IRMOF-0, which uses acetylenedicarboxylate as the linker [45]. This synthesis method follows a room temperature synthesis, wherein separate N,N-dimethylformamide (DMF) solutions of terephthalic acid (BDC) with triethylamine and zinc acetate dihydrate are prepared, then the zinc salt solution is added to the organic solution with rapid stirring at ambient temperature. Upon immediately of the formation of a white powder followed by 2.5 hours of reaction time, pure MOF-5 is collected and confirmed by powder XRD. The same synthesis without a base (triethylamine) has also yielded pure MOF-5, confirming that addition of a base is unnecessary when zinc acetate is used as a source of Zn (II) in the MOF-5 synthesis [45].

This synthesis method has later modified by Rathnayake et.al to make IRMOFs (IRMOF-1, 8, and 10) by cutting down the reaction time from 2.5 hours to 7–9 minutes [23]. As depicted in Figure 3, our group is able to make a wide range of hierarchical microstructures of highly crystalline MOFs, including IRMOF-1. Microstructures of IRMOF-1 prepared from the modified solvothermal method (Figure 3), are visualized using scanning electron microscope and are depicted in Figure 4(a). Crystal structure of IRMOF-1, retrieved by matching its simulated XRD with experimental powder XRD is depicted in Figure 4(b), and follows cubic lattice cell, which belongs to Fm3m cubic space groups. The electron density potential distribution modeled from VESTA (Figure 4(c)) evidences that the electron potential is localized on Zn4O clusters and there is no electron delocalization with the organic linkers, confirming no orbital overlap for energy transfer through metal–ligand charge transfer processes.

Figure 3.

Reaction scheme for synthesis of Isoreticular MOFs using modified solvothermal method followed by solvent driven self-assembly.

Figure 4.

(a) A SEM image of IRMOF-1 microstructures, (b) crystal structure of IRMOF-1 retrieved from crystallographic open database, and (c) electron density potential distribution of IRMOF-1 modeled from VESTA.

As a first member of isoreticular series, IRMOF-1 has explored for luminescence due to ZnO quantum dots behavior, which has been believed, contributing to luminescence. The ZnO QD absorption and emission spectra from electronic transitions have been investigated, suggesting that that the luminescent behavior of IRMOF-1 arises from a O2 − Zn+ → O − Zn+ charge-transfer transition within each tetrahedral Zn4O metal cluster, which has been described as a ZnO-like QD [46]. The photoluminescence emissions of IRMOF-1 with intensity peak maximum at 525 nm, was ascribed to energy harvesting and LMCT from 1,4-benzenedicarboxylate (BDC) linked to the Zn4O cluster. The nature of the luminescence transitions in IRMOF-1 nanoparticles has been investigated by Tachikawa et al. where the transition responsible for the green emission of IRMOF-1 is similar to that of ZnO [47]. Therefore, the emission observed in IRMOF-1 has been speculated to originate from the ZnO QD not from the ligand. However, Further investigations demonstrated that ZnO impurities in the material gave rise to the emission assigned to the quantum dot like luminescence and that pure MOF-5 displays a luminescence behavior that is more closely relevant to that of the ligand. [9] However, the exact nature of the luminescence of MOF-5 is still under dispute with ligand− ligand charge transfer, [10] ligand-centered, [9] and ligand–metal charge transfer [11] mechanisms as primary suggestions.

In an on-going study, our group has been investigating optoelectronic behavior of IRMOF-1. As depicted in Figure 5, UV–vis absorption spectrum shows absorption vibrionic features similar to the linker with two absorptions peaks at 208 nm and 240 nm along with a shoulder peak at 285 nm. The emission spectrum collected by exciting at 240 nm exhibits linker-based emission with three well-resolved vibrionic transitions at 328 nm, 364 nm, and 377 nm. We observed a small high energy shoulder peak at 464 nm, which corresponds to an excitonic transition of Zn4O nodes. However, we have no observed a longer wavelength emission peak at 525 nm, which has claimed in prior studies to energy harvesting and LMCT from 1,4-benzenedicarboxylate (BDC) linked to the Zn4O cluster. Therefore, our findings support that IRMOF-1’s luminescence comes from linker emission rather than the charge transfer processes. This further excludes the emission originating from the ZnO quantum dots like clusters of Zn4O. The optical band gap calculated from the UV–visible spectrum on-set is found to be 3.97 eV. There are no experimental band gaps reported for IRMOF-1 up to date.

Figure 5.

Photophysical properties of IRMOF-1 – (a) UV–visible spectrum and (b) photoluminescence spectrum in solution (ethanol).

3.3 Synthesis and optoelectronic properties of IRMOF-8

Significant research efforts have demonstrated successful synthesis of a variety of isoreticular MOFs (IRMOFs) with the formula of Zn4O(L)3 (where L is a rigid linear dicarboxylates) using traditional solvothermal method, which uses zinc nitrate as metal precursor and the respective organic ligands in an amide-based solvent system. These IRMOFs follow the same cubic topology as the prototypical MOF-5, a framework with octahedral Zn4O(CO2)6 clusters, which are linked along orthogonal axes by phenylene rings [3, 26, 48, 49]. This family of IRMOFs-n (n = 1–16) gained significant attention in gas storage community due to its high pore volume and surface area. Among the IRMOFs series, IRMOF-1 and 8 have been extensively studied for gas adsorption and photoluminescence properties [395051] but have not explored their optoelectronic properties until recently.

IRMOF-8 is constructed from the linkage of basic zinc acetate clusters and naphthalene-2,6-dicarboxylic acid units (NDC). Originally reported IRMOF-8 with non-interpenetrated cubic crystal lattice has only been extensively studied for gas sorption and storage applications [50, 51]. Later, a number of interpenetrated phases of Zn4O(ndc)3-based systems have been discovered [52, 53, 54]. Although the synthesis of interpenetrated IRMOF-8 (INT-IRMOF-8) are similar to that of IRMOF-8, the possibility that typical solvothermally synthesized IRMOF-8 contains at least a significant amount of an interpenetrated phase. There are modified synthesis methods have been introduced to make fully non-interpenetrated IRMOF-8 [55] and INT-IRMOF-8 [23, 55]. The crystal structures of non-interpenetrated IRMOF-8 and INT-IRMOF-8 along with their space filling structures, acquired from the Crystallographic Open Database (COD) and generated using VESTA are depicted in Figure 6.

Figure 6.

Crystal structures of: (a) non-interpenetrated IRMOF-8 and (b) its space filling view, (c) INT-IRMOF-8 and (d) its space filling view.

Recently, our group has introduced a modified solvothermal synthesis method, which involves a solvent polarity driven self-assembly process to make hierarchical microstructures of INT- IRMOF-8, exhibiting promising optoelectronic properties for the first time [23]. Instead using zinc nitrate as the metal precursor, the synthesis we developed utilizes zinc(II)acetate as the metal precursor. Hierarchical microstructures of INT-IRMOF-8 nanocrystals can be prepared in high yield in the presence of minimum volume of dimethyl formamide by mixing zinc(II) precursor with naphthalene-2,6-dicarboxylic acid at room temperature followed by subjecting to solvothermal annealing at 260°C for 7 minutes [23]. Microstructures visualized under TEM (Figure 7(b)) reveal that they are hierarchical layers of self-assembled nanocrystals with randomly arranged voids among the nanocrystals. The wide-angel X-ray scattering (WAXS) pattern along with the selective area electron diffraction (SAED) pattern have shown that the microstructures are made from self-assembled nanocrystals of INT-IRMOF-8, which exhibits lamella packing pattern (Figure 7(c) and (d)), benefiting for optoelectronic behavior.

Figure 7.

(a) Interpenetrated view of INT-IRMOF-8’s crystal structure; (b) a TEM image of a microstructure; (c) the SAED pattern of a microstructure taken from the TEM under dark field diffraction mode along with (d) a TEM image of the respective microstructure. [Figure 7 is re-created from the original data].

The photophysical properties of INT-IRMOF-8 exhibit mainly linker based optical properties. The presence of high intensity absorption peak set from 220 nm to 360 nm, which corresponds to vibronic π-π* absorption transitions of naphthalene core, evidencing the linker-based absorption, resulting from the lack of favorable spatial and energetic overlap of the metal and the ligand orbitals [21, 49]. Typically, MOFs’ photoluminescence behavior arises as a result of different types of charge transfer processes, which include metal-to-ligand charge transfer (MLCT), ligand-to-metal charge transfer (LMCT), ligand–ligand charge transfer (LLCT), ligand-centered luminescence, and metal-to-metal charge transfer (MMCT) processes [56]. However, this metal-centered luminescence depends on the metal type, ligand type, and their spatial orientations. The emission spectrum of INT-IRMOF-8 microstructures exhibits three emission bands, in which vibrionic transitions corresponds to only linker-based emission with no indication of additional emissions due to any charge transfer processes. The optical band gap reported for INT-IRMOF-8 is 2.82 eV [23] and the theoretical band gap reported in the past for non-interpenetrated IRMOF-8 was ranged from 2.83 eV to 3.27 eV [57]. There are no experimental band gaps reported for IRMOF-8 up to date.

The charge transfer ability of IRMOF-8 for the first time is evaluated by our group. The average electrical conductivity of INT-IRMOF-8 microstructures was ranged from 3.98 x 10−2 to 2.16 x 10−2 S cm−1, which is higher than the electrical conductivities reported for most MOFs (<10−10 S/cm). The interpenetrated structure, high crystallinity, and narrow band gap contribute to the to the electrical conductivity of hierarchical structures of INT-IRMOF-8 nanocrystals.

3.4 Synthesis and optoelectronic properties of IRMOF-10

Among the series of IRMOFs (IRMOF 1–16) introduced by Yaghi and coworkers [58, 59, 60], several IRMOFs have shown effective selective preconcentration properties, including IRMOF-10 [61, 62, 63, 64].

Compared to IRMOF-1, physicochemical, optical, and electronic properties of IRMOF-10 with its 4,4′-biphenyldicarboxylate linkers has received much less attention. IRMOF-10 was first synthesized by Yaghi and coworkers [50, 58, 59, 60], Owing to its higher surface area and larger pore sizes, use of IRMOF-10 for gas absorption and separation and hydrogen storage have been widely investigated, but scarce attention has been paid to other properties of IRMOF-10, such as structural stability, optical and electrical properties, electronic structure, and chemical bonding. The first publication about biphenyl MOFs already anticipated the major challenges related to Zn-biphenyl MOFs: the growth of single-crystals and interpenetration. A structure from single-crystal XRD of non-functionalized IRMOF-10 is not yet available. A single-crystal X-ray structure analysis of a non-interpenetrated IRMOF-10 derivative was not reported until the breakthrough of the group of Telfer, which showed how interpenetration can be effectively suppressed by using thermolabile protecting groups in the synthesis of amino-MOFs [65]. Following the modified solvothermal synthesis method introduced by Rathnayake et al., microstructures of non-interpenetrated IRMOF-10 was successfully synthesized, and crystal structure was confirmed by matching the powder XRD traces with the simulated XRD pattern. The microstructures morphology is depicted in Figure 8(a) and crystal structure retrieved from the Crystallographic Open Database is depicted in Figure 8(b). IRMOF-10’s single crystal structure reveals three-dimensional coordination framework, formed by periodic arrangement of Zn(II) atoms, which is tetrahedrally coordinated by four oxygen atoms from four biphenyl linker units, following the unit formula of Zn4O(L)3 with cubic topology as IRMOF-1.

Figure 8.

(a) A SEM image of IRMOF-10 microstructures, (b) crystal structure of IRMOF-10 retrieved from crystallographic open database, and (c) electron density potential distribution of IRMOF-10 modeled from VESTA.

The electron potential density localization surrounding the metal oxide nodes and organic linker units in IRMOF-10’s unit cell reveals the electron density distribution with respect to the biphenyl conjugation length. As shown in Figure 8(c), the electron potential is delocalized within metal oxide nodes and bi-phenyl units, and partial distribution of charges has increased around bi-phenyl units compared to naphthalene units of IRMOF-8. Thus, the findings suggest that linker length has more pronounced effect on the individual material’s electronic band structure and density of state, providing clear visualization on the localization of electronic potential within the crystal lattice. The delocalization of electron density potential through biphenyl linkers evidences its potential to be used as optoelectronic materials. Thus, exploring its electronic structure, band gap, optical, and electrical properties is a major interest to the materials science community. However, despite computational investigations on theoretical prediction of optoelectronic properties, [66] there are no experimental investigations on IRMOF-10’s optoelectronic behavior has been conducted up to date.

The equilibrium solid-state structure, electronic structure, formation enthalpy, chemical bonding, and optical properties of IRMOF-10 have investigated with density functional calculations. Electronic density of states and band structures study have shown that the band gap for the IRMOF-10 is ranged from 2.9 eV to 3.0 eV, resulting in a nonmetallic character [66]. Until now, there are no experimental studies available to verify theoretical predictions on IRMOF-10’s electronic structure. The calculated optical properties of IRMOF-10 provide useful information for future experimental exploration. The optical properties (dielectric function, refractive index, absorption coefficient, optical conductivity s(v), reflectivity, and electron energy-loss spectrum of the IRMOF-10 have computed in the past, [66] but there is no experimental investigation conducted up to date.

Recently, our group has studied optoelectronic behavior of non-interpenetrated IRMOF-10. As shown in Figure 9, we explored the photophysical properties of non-interpenetrated IRMOF-10 and calculated its optical band gap. IRMOF-10 exhibits linker-based absorption with absorption maximum at 282 nm along with a shoulder peak at 222 nm. IRMOF-10 shown blue luminescence with broader emission ranged from 310 nm to 450 nm along with the emission maximum at 353 nm. The optical band gap calculated from the UV–visible spectrum on-set is 3.80 eV, which is narrower than the optical band gap of IRMOF-1 and larger than the theoretical band gap predicted from computational analysis.

Figure 9.

Photophysical properties of IRMOF-10 – (a) UV–visible spectrum and (b) photoluminescence spectrum in solution (ethanol).

In summary, the conjugation length of the organic linker in IRMOFs contributes to the semiconducting properties rather than the periodic pattern or the distances between the Zn4O clusters. The conjugation length of organic linkers of IRMOF-1, 8, and 10 described here differs from one aromatic unit (benzene) to one and half aromatic unit (naphthalene) to two aromatic units (biphenyl). The resonance effect arises due to the conjugation speaks directly to the photophysical behavior and optical band structure characteristics, reflecting a clear trend in narrowing the band gap with gradual increase in the conjugation length of the ligand. The dramatic change in the optical band gap upon changing the organic linker in the MOF structure has also been reported in the past [66]. Thus, these studies evidences that the semiconducting properties of MOFs strongly depends on the resonance effects from the organic linker [67].

Advertisement

4. Future prospective

With a growing demand for continuous miniaturization and functional scaling, the complementary metal-oxide semiconductor (CMOS) platform continues to drive advances in integrated circuits (IC), nanoelectronics and information processing technologies. While it is now possible to produce an amazing array of nanoscale materials and morphologies, the assembly and integration of these nanostructures into ordered arrays, with other materials, remain key challenges. Moore’s Second Law projects a need for new, high throughput fabrication approaches that can produce useful and defect free nanostructures for future silicon-based CMOS related technologies. Recent advances in nanomaterial synthesis enable new families of emerging research materials (ERMs) that show potential for extending and augmenting existing CMOS technology, with respect to wafer level manufacturability, uniformity, reliability, performance and cost, and they warrant additional research focus and verification. The integration of More-than-Moore, application specific, materials and structures on a CMOS platform leverages the best of both technologies, though this added complexity also challenges the extensibility of conventional fabrication and patterning methods. Consequently, there remains a need for simple fabrication methods that can create two- and three-dimensional ordered functional nanostructures, which can adapt to a wide variety of materials, patterning, and application needs.

Highly crystalline microstructures of MOFs have been paving the path, addressing the current challenges in fabrication needs that create two- and three-dimensional ordered structures and which are adaptable to a wide variety of materials specific applications. These nanoscale building blocks, and their assemblies combine the flexibility, conductivity, transparency, and ease of processability of soft matter (organic) with electrical, thermal, and mechanical properties of hard matter (inorganic). They offer a new window for fine-tuning structural nodes with known geometries and coordination environments. With respect to the fabrication of ordered nanoscale structures, MOFs have several advantages. First, since they are themselves a highly ordered self-assembled nanostructure, as a result of their crystallinity, their pore dimensions are completely defined, making knowledge of atomic positions possible. Second, the nanoporosity of their structure results from geometric factors associated with the bonding between their inorganic and organic components, enabling rational template design [68]. Third, unlike the conventional template materials, MOFs possess a high degree of synthetic flexibility with potentially widely tunable electrical, optical, and mechanical properties. Surely, the development of simple, versatile low-cost methodologies for the design, production, and nanoscale manipulation of innovative functional organic–inorganic hybrid building blocks will provide a powerful new capability for designing, integrating, and patterning new nanoscale materials with tunable properties onto a CMOS platform. Recent milestones of MOFs in photovoltaic, optical and chemical sensing, and field effect transistors highlight the potential of these materials for future electronic devices, heterogeneous platforms, non-traditional patterning opportunities [16, 69, 70, 71].

Interest in using these materials in fields such as gas storage [72], separations [73], [sensing [21], and catalysis [74] is rapidly accelerating. The advantages of MOFs for above applications are promising due to the wide range of possibilities of the rational design inherited in these materials. Thus, superior properties and new understanding with respect to the interaction of small molecules with nanoporous materials are being achieved. Although most MOFs are found to be dielectrics, a few semiconducting frameworks are known [23, 37, 75, 76]. The theoretical predictions conduced up to date on variety of MOFs suggest that there are possible MOFs with semiconducting properties [77, 78, 79] . MOFs that are magnetic [80], ferroelectric [81, 82], proton-conducting [83, 84, 85, 86], and luminescent [87, 88] are also known. Additionally, their porosity creates the potential to introduce non-native functionality to a given structure by infusing the accessible volume with a second molecule or material. Moreover, because the chemical environment within the pore can be modified, it is possible to tailor the interface between the MOF and a templated material to stabilize specific materials or nanostructures. Consequently, MOFs and the coordination polymers of crystalline nanoporous frameworks possess many of the properties of an ideal template.

Despite the endless possibilities for how MOFs could be used for device applications, when using MOFs for semiconductor microelectronic devices such as sensors, field-effect transistors, light harvesting and absorbing, light-emitting diodes, thermoelectric devices, energy storages and lithium ion batteries, and scintillators, it is necessary to understand how these materials function within the device and how they will interface with other functional and structural elements. Therefore, this section focuses on providing a future prospective for advances that must be made for their realization in electronic devices. A possible MOF-device roadmap, which identifies MOFs applications in electronic devices along with machine learning for new MOFs developments and MOFs database screening for novel properties is depicted in Figure 10. Our intention of providing this roadmap is to stimulate future endeavors of MOFs roadmap for electronic industry by translating current MOFs basic research agenda into applied research in the future. The roadmap that we identified here is created by combining the prospective previously provided by Allendorf et al., focusing five major fields pertinent to device fabrication [89]. These previously proposed areas include: (1) Fundamental Properties, (2) Thin film growth and processing, (3) MOFs hybrid and multilevel structures, (4) Device integration, and (5) Manufacturing issues. Our prospective for the proposed potential MOF-device roadmap particularly concentrated on member-specific applications in electronic industry, where functional density of MOFs can be utilized in subcategories of a wide variety of electronic devices. As the MOF-based optoelectronic field is fairly new and fall within the basic research stage, our roadmap is structured based on MOFs relative progress made so far and build upon the future road map comparing with the progress made in the field of organic electronics.

Figure 10.

A possible MOF-device roadmap for electronic industry proposed by Rathnayake.

Exploring electronic properties, such as electronic structure, band gap, conductivity, electron, and hole mobilities, and dielectric constants of MOFs need to be one of the priority areas in the next decades and must be understood. Additionally, understanding lattice defects and their relationship to electronic properties must be explored combining theoretical and experimental approaches as they likely will limit the ultimate performance of a device. The field-effect transistor (FETs), which is the basic device building block for modern electronics, dictates the materials properties relevant to electronic applications. The FET performance is determined by the carrier mobility, source and drain contact resistance, and the capacitance of the gate electrode. Si is the preeminent materials for FET fabrication because of its bandgap of 1.1 eV, high carrier mobility, availability of multiple n- and p-type dopants, environmental stability, stable oxide, and high terrestrial abundance. However, Si based device fabrication requires enormous capital investment and Si is not compatible with a variety of low cost, flexible, transparent, and low melting temperature substrates. For these reasons, alternative materials including polymers, organic molecules, and more recently nanotubes and nanowires have been gaining a lot of attention for various emerging applications. The long-range crystalline order of MOFs implies that charge transport through delocalized conduction and valence bands typical of crystalline inorganic semiconductors is possible. Emergence of delocalized bands in MOFs will require that the π orbitals in the linker groups overlap effectively with the metal d orbitals. Such overlap is absent in the majority of synthetically known MOFs where carboxylate oxygen atoms are coordinated to the metal center through σ bonds. Therefore, most MOFs are electrical insulators. This barrier needs to be overcome in next decades, perhaps by synthesizing novel MOFs using higher order conjugated linkers and increasing the functional density of the MOFs. Modifying the linker structure could lead to better charge transfer between linker and the metal cations of the framework. One possible route has suggested replacing the carboxylate terminating linkers with isocyanide groups [89]. It has been shown that, Prussian Blue, a mixed valence crystalline compound with Fe(II) and Fe(III) ions coordinated with isocyanide ligands, is electrically conducting [90]. Another approach, suggested by Allendorf et al., is to introduce conducting phases into the MOF channels [89]. Some other approaches have been taken place to enhance electronic transport properties of MOFs by introducing other conductive nanomaterials, inorganic oxides, polymers, and carbon nanotubes into MOFs framework [91, 92, 93].

Atomic level fundamental understanding that cannot be obtained readily from experimental methods, is necessary to address MOFs electronic band structure, density of state, band gap, and electron and hole mobilities. There has been increasing accuracy of predictive results using molecular dynamics (MD) force fields (FF) and DFT approximations for various MOFs’ property studies [94, 95, 96, 97, 98, 99, 100]. Density Functional Theory (DFT) methods using periodic boundary conditions have been popular for predicting the electronic structure of MOFs [57, 67, 101, 102]. However, DFT-based computational calculations are underestimate excited state energies by about a factor of two. Adapting high-accuracy methods, such as Quantum Monte Carlo (QMC), DFT + U, and GW are not feasible for systems with large numbers of electrons. For example, practical QMC calculations currently could not apply to the systems that exceed 1000 electrons. One formula unit of IRMOF-1 has 760 electrons and 106 atoms. Owing to these limitations in current computational approaches, MOFs are much more challenging than traditional electronic materials with much smaller unit cells. The computational methods for predicting properties of MOFs are at an early stage of development, in particularly for predicting electronic properties of MOFs [57, 101]. Developing simple and rapid analytical approaches are not only a necessary tool to experimental investigations but also can be used as an accelerate investigation and predictive tool by themselves to screen semiconducting MOFs from the database of synthetically known MOFs. Such analytical approaches combined with computational studies will eventually enable the design of machine learning approaches for large-scale screening of existing and hypothetical MOF structures for specific applications [103, 104, 105].

MOFs are also showing promise in their use as electrolytes due to their low electronic conductivity, tunable polarity, and high porosity [106]. There are many ways that MOFs have been employed to elevate the downfalls of current electrolytes. For example, they have been using as hosts for liquid electrolyte solutions or ionic liquids [107, 108]. However, the drying of the electrolyte solution within the MOFs presents an issue since the ion transport is mostly achieved by the solvent molecules within the electrolyte rather than by the MOF itself. Furthermore, MOFs is used as a filler to reduce the crystallinity of SPEs [107, 108]. However, up to date MOFs have not been explored to be used as a solid electrolyte excepts in a composite form [109]. One way to achieve this is designing lithium-based metal organic frameworks (Li MOF) where excess lithium is transferred through the defects in the MOF structure. However, research regarding Li MOFs as solid electrolytes is currently lacking. The majority of MOF/electrolyte studies are only focused on employing MOFs as a host of ionically conductive materials rather than utilizing MOFs as solid-state electrolytes. Therefore, we identified this area of research in the proposed roadmap to stimulate investigating the potential of Li-based MOFs as solid electrolytes. There are different types of Li-MOFs already developed [110, 111, 112], but many of them are designed for applications other than battery electrolytes. We believe that Li- MOF structures can be tuned for lithium transport. Overall, Li MOFs show potential for the use as solid ionic conductors and much research should be performed to explore their possibility for solid state electrolytes and battery components.

Exploring thermoelectric properties of MOFs emerges five years ago along with exploring the electronic properties of MOFs by systematic structural modifications and introducing guest molecules onto MOFs. The first thermoelectric property measurements on MOFs has introduced by Erikson in 2015 [113]. then, up to date, there have been less than ten publications in thermoelectric MOFs, thus this field of research is relatively new. Highly nanoporous MOFs are promising since porosity can reduce the lattice thermal conductivity. The effect the conjugation length of the organic linker that tailors the pore dimension for lattice thermal conductivity must be investigated. The thermoelectric figure of merit that measures the efficiency of a thermoelectric device can be improved by decreasing the lattice thermal conductivity. It is believed that changing the conjugation length or the complexity of the organic linker changes phonon scattering, thereby changing the lattice thermal conductivity [77, 114]. The ligand modifications can be successfully achieved by isoreticular synthesis approaches. Also, increasing the porosity of MOFs increases phonon scattering that also reduces thermal conductivity [114]. Therefore, in order to utilize MOFs as active materials in thermoelectric devices, understanding the contribution of phonon vibrations to lattice thermal conductivity is essential and must be investigated. Directing future research on thermoelectric MOFs towards experimentally investigating thermoelectric properties of MOF based thin films to find ways of decreasing thermal conductivity by structural modifications to the organic ligand is beneficial.

In order to use MOFs as photoactive layer for energy harvesting and conversion, MOFs should possess decent light harvesting capability in the region from visible light to near-infrared (NIR). As the material’s light-harvesting window is primarily determined by its band gap, synthesizing a MOF with a semiconducting band gap that can absorb light in the solar spectrum should be a requirement for it to serve as the photoactive material. Given that the electronic configuration of MOFs is contributed by both the constituent metal ion and the organic linker, the resultant bandgap and semiconducting properties of MOFs can thus be tailored by their structural design and engineering. Since most MOFs possess large band gap due to lack of overlap between metal ion and the organic linker and low degree of conjugation, they cannot effectively absorb light in the solar spectrum. The ligand center of MOFs plays a dominant role in its resulting light harvesting behavior [77114]. Tailoring the structure and its composition, MOFs charge transfer processes can be improved to enable the photocurrent of MOFs and fulfilling the photoactive functions.

To effectively reduce the band gap of MOFs and enrich their semiconducting properties for photovoltaic applications, three strategies can be implemented and have been identified [115]. These strategies are: (1) selecting electron rich metal nodes and conjugated-based organic molecules, (2) enhancing the conjugation of the organic linker, and (3) functionalizing the organic linker with electron-donating groups, such as hydroxyl, nitro, and amino groups. Additionally, facilitating electron delocalization through guest-mediated p-donor/acceptor stacks can also effectively diminish the band gaps of the materials [115]. Besides narrowing the band gap, electronic structure that contributes the semiconducting properties of MOFs also play a vital role as sufficient dissociation of the photoexcitons generated in the MOFs is required to produce a reasonable photocurrent. In this regard, MOFs exhibit critical barriers to use as the photoactive materials directly and impedes its progress in photovoltaic applications to date. However, up to date, besides acting as the photoactive materials, the MOFs has been contributing to the photovoltaic community by serving as functional additives or interlayers to improve the performance and stability of the derived solar cell devices. In order to utilize MOFs for photoactive layer in photovoltaics, it is necessary to design electrically conductive MOFs. The research efforts developing more functional conducting MOFs are required in the coming decade.

Advertisement

5. Conclusions

Owing to synthetic processability using reticular chemistry, MOFs offer unusual properties paving the path for many opportunities and their use in optoelectronic devices. Their use in devices so far is limited to sensors and gas storage. However, MOFs field is moving towards exploring their optical and electrical properties to use in electronic devices. There are many MOFs with tunable bandgap, both ultralow-k and high-k dielectric constants, varied magnetic properties, luminescence, and a few with semiconducting behavior, suggesting MOFs as emerging material with unique properties exceeding any other class of materials. Combining the solvothermal synthesis method with self-assembly processes, we can achieve highly ordered nanoporous structures with precise dimensionality that creates the potential for electronics and self-assembly with atomic-scale resolution and precision. In order to become MOFs for electronic devices, many challenges must be solved, and electronic structures of MOFs should be revealed. The MOFs-device roadmap should be one meaningful way to reach MOFs milestones for optoelectronic devices and will enable MOFs to be performed in their best, as well as allowing the necessary integration with other materials to fabricate fully functional devices.

Advertisement

Acknowledgments

Authors acknowledge the Joint School of Nanoscience and Nanoengineering, a member of the South-eastern Nanotechnology Infrastructure Corridor (SENIC) and National Nanotechnology Coordinated Infrastructure (NNCI), supported by the NSF (Grant ECCS-1542174).

Advertisement

Conflict of interest

There is no conflict of interest to declare.

References

  1. 1. Zafar F, Sharmin E (2016) Metal-Organic Frameworks. BoD–Books on Demand
  2. 2. Ogawa T, Iyoki K, Fukushima T, Kajikawa Y (2017) Landscape of research areas for zeolites and metal-organic frameworks using computational classification based on citation networks. Materials 10:1428
  3. 3. Furukawa H, Cordova KE, O’Keeffe M, Yaghi OM (2013) The chemistry and applications of metal-organic frameworks. Science 341:
  4. 4. Eslava S, Zhang L, Esconjauregui S, Yang J, Vanstreels K, Baklanov MR, Saiz E (2013) Metal-organic framework ZIF-8 films as low-κ dielectrics in microelectronics. Chemistry of Materials 25:27-33
  5. 5. Yap MH, Fow KL, Chen GZ (2017) Synthesis and applications of MOF-derived porous nanostructures. Green Energy & Environment 2:218-245
  6. 6. Logar NZ, Kaucic V (2006) Nanoporous materials: from catalysis and hydrogen storage to wastewater treatment. Acta chimica slovenica 53:117
  7. 7. Espallargas GM, Coronado E (2018) Magnetic functionalities in MOFs: from the framework to the pore. Chemical Society Reviews 47:533-557
  8. 8. Vikrant K, Kumar V, Kim K-H, Kukkar D (2017) Metal–organic frameworks (MOFs): potential and challenges for capture and abatement of ammonia. Journal of Materials Chemistry A 5:22877-22896
  9. 9. Ma M, Lu L, Li H, Xiong Y, Dong F (2019) Functional metal organic framework/sio2 nanocomposites: From versatile synthesis to advanced applications. Polymers 11:1823
  10. 10. Liu Y, Zhao Y, Chen X (2019) Bioengineering of metal-organic frameworks for nanomedicine. Theranostics 9:3122
  11. 11. Rowsell JL, Yaghi OM (2004) Metal–organic frameworks: a new class of porous materials. Microporous and mesoporous materials 73:3-14
  12. 12. Kotzabasaki M, Froudakis GE (2018) Review of computer simulations on anti-cancer drug delivery in MOFs. Inorganic Chemistry Frontiers 5:1255-1272
  13. 13. Rowsell JL, Spencer EC, Eckert J, Howard JA, Yaghi OM (2005) Gas adsorption sites in a large-pore metal-organic framework. Science 309:1350-1354
  14. 14. Yang J, Trickett CA, Alahmadi SB, Alshammari AS, Yaghi OM (2017) Calcium L-lactate frameworks as naturally degradable carriers for pesticides. Journal of the American Chemical Society 139:8118-8121
  15. 15. Silva CG, Corma A, García H (2010) Metal–organic frameworks as semiconductors. Journal of Materials Chemistry 20:3141-3156
  16. 16. Stassen I, Burtch N, Talin A, Falcaro P, Allendorf M, Ameloot R (2017) An updated roadmap for the integration of metal–organic frameworks with electronic devices and chemical sensors. Chemical Society Reviews 46:3185-3241
  17. 17. Usman M, Mendiratta S, Lu K-L (2017) Semiconductor metal–organic frameworks: Future low-bandgap materials. Advanced Materials 29:1605071
  18. 18. Stavila V, Talin AA, Allendorf MD (2014) MOF-based electronic and opto-electronic devices. Chemical Society Reviews 43:5994-6010
  19. 19. Guo Z, Panda DK, Gordillo MA, Khatun A, Wu H, Zhou W, Saha S (2017) Lowering Band Gap of an Electroactive Metal–Organic Framework via Complementary Guest Intercalation. ACS applied materials & interfaces 9:32413-32417
  20. 20. Deng X, Hu J-Y, Luo J, Liao W-M, He J (2020) Conductive Metal–Organic Frameworks: Mechanisms, Design Strategies and Recent Advances. Topics in Current Chemistry 378:1-50
  21. 21. Allendorf MD, Bauer CA, Bhakta RK, Houk RJT (2009) Luminescent metal–organic frameworks. Chemical Society Reviews 38:1330-1352
  22. 22. Yang C, Dong R, Wang M, Petkov PS, Zhang Z, Wang M, Han P, Ballabio M, Bräuninger SA, Liao Z (2019) A semiconducting layered metal-organic framework magnet. Nature communications 10:1-9
  23. 23. Dawood S, Yarbrough R, Davis K, Rathnayake H (2019) Self-assembly and optoelectronic properties of isoreticular MOF nanocrystals. Synthetic Metals 252:107-112
  24. 24. Rowsell JL, Yaghi OM (2004) Metal–organic frameworks: a new class of porous materials. Microporous and mesoporous materials 73:3-14
  25. 25. Jhung SH, Lee J-H, Yoon JW, Serre C, Férey G, Chang J-S (2007) Microwave synthesis of chromium terephthalate MIL-101 and its benzene sorption ability. Advanced Materials 19:121-124
  26. 26. Castaldelli E, Imalka Jayawardena KDG, Cox DC, Clarkson GJ, Walton RI, Le-Quang L, Chauvin J, Silva SRP, Demets GJ-F (2017) Electrical semiconduction modulated by light in a cobalt and naphthalene diimide metal-organic framework. Nat Commun 8:2139
  27. 27. Qiu L-G, Xu T, Li Z-Q , Wang W, Wu Y, Jiang X, Tian X-Y, Zhang L-D (2008) Hierarchically micro-and mesoporous metal–organic frameworks with tunable porosity. Angewandte Chemie International Edition 47:9487-9491
  28. 28. Peng L, Zhang J, Xue Z, Han B, Sang X, Liu C, Yang G (2014) Highly mesoporous metal–organic framework assembled in a switchable solvent. Nat Commun 5:4465
  29. 29. Zhao Y, Zhang J, Han B, Song J, Li J, Wang Q (2011) Metal–organic framework nanospheres with well-ordered mesopores synthesized in an ionic liquid/CO2/surfactant system. Angewandte Chemie International Edition 50:636-639
  30. 30. Kung C-W, Han P-C, Chuang C-H, Wu KC-W (2019) Electronically conductive metal–organic framework-based materials. APL Materials 7:110902
  31. 31. Goswami S, Ray D, Otake K, Kung C-W, Garibay SJ, Islamoglu T, Atilgan A, Cui Y, Cramer CJ, Farha OK (2018) A porous, electrically conductive hexa-zirconium (IV) metal–organic framework. Chemical science 9:4477-4482
  32. 32. Sun L, Hendon CH, Minier MA, Walsh A, Dincă M (2015) Million-fold electrical conductivity enhancement in Fe2 (DEBDC) versus Mn2 (DEBDC)(E= S, O). Journal of the American Chemical Society 137:6164-6167
  33. 33. Talin AA, Centrone A, Ford AC, Foster ME, Stavila V, Haney P, Kinney RA, Szalai V, El Gabaly F, Yoon HP (2014) Tunable electrical conductivity in metal-organic framework thin-film devices. Science 343:66-69
  34. 34. Banerjee S, Tyagi AK (2011) Functional materials: preparation, processing and applications. Elsevier
  35. 35. Wu G, Huang J, Zang Y, He J, Xu G (2017) Porous field-effect transistors based on a semiconductive metal–organic framework. Journal of the American Chemical Society 139:1360-1363
  36. 36. Narayan TC, Miyakai T, Seki S, Dincă M (2012) High charge mobility in a tetrathiafulvalene-based microporous metal–organic framework. Journal of the American Chemical Society 134:12932-12935
  37. 37. Kobayashi Y, Jacobs B, Allendorf MD, Long JR (2010) Conductivity, Doping, and Redox Chemistry of a Microporous Dithiolene-Based Metal−Organic Framework. Chem Mater 22:4120-4122
  38. 38. Pathak A, Shen J-W, Usman M, Wei L-F, Mendiratta S, Chang Y-S, Sainbileg B, Ngue C-M, Chen R-S, Hayashi M (2019) Integration of a (–Cu–S–) n plane in a metal–organic framework affords high electrical conductivity. Nature communications 10:1-7
  39. 39. Li H, Eddaoudi M, O’Keeffe M, Yaghi OM (1999) Design and synthesis of an exceptionally stable and highly porous metal-organic framework. nature 402:276-279
  40. 40. Mueller T, Ceder G (2005) A density functional theory study of hydrogen adsorption in MOF-5. The Journal of Physical Chemistry B 109:17974-17983
  41. 41. Bordiga S, Vitillo JG, Ricchiardi G, Regli L, Cocina D, Zecchina A, Arstad B, Bjørgen M, Hafizovic J, Lillerud KP (2005) Interaction of hydrogen with MOF-5. The Journal of Physical Chemistry B 109:18237-18242
  42. 42. Li Y, Yang RT (2006) Significantly enhanced hydrogen storage in metal− organic frameworks via spillover. Journal of the American Chemical Society 128:726-727
  43. 43. Hafizovic J, Bjørgen M, Olsbye U, Dietzel PD, Bordiga S, Prestipino C, Lamberti C, Lillerud KP (2007) The inconsistency in adsorption properties and powder XRD data of MOF-5 is rationalized by framework interpenetration and the presence of organic and inorganic species in the nanocavities. Journal of the American Chemical Society 129:3612-3620
  44. 44. Huang L, Wang H, Chen J, Wang Z, Sun J, Zhao D, Yan Y (2003) Synthesis, morphology control, and properties of porous metal–organic coordination polymers. Microporous and mesoporous materials 58:105-114
  45. 45. Tranchemontagne DJ, Hunt JR, Yaghi OM (2008) Room temperature synthesis of metal-organic frameworks: MOF-5, MOF-74, MOF-177, MOF-199, and IRMOF-0. Tetrahedron 64:8553-8557
  46. 46. Bordiga S, Lamberti C, Ricchiardi G, Regli L, Bonino F, Damin A, Lillerud K-P, Bjorgen M, Zecchina A (2004) Electronic and vibrational properties of a MOF-5 metal–organic framework: ZnO quantum dot behaviour. Chemical communications 2300-2301
  47. 47. Tachikawa T, Choi JR, Fujitsuka M, Majima T (2008) Photoinduced Charge-Transfer Processes on MOF-5 Nanoparticles: Elucidating Differences between Metal-Organic Frameworks and Semiconductor Metal Oxides. J Phys Chem C 112:14090-14101
  48. 48. Stock N, Biswas S (2012) Synthesis of Metal-Organic Frameworks (MOFs): Routes to Various MOF Topologies, Morphologies, and Composites. Chem Rev 112:933-969
  49. 49. Sun L, Campbell MG, Dincă M (2016) Electrically Conductive Porous Metal-Organic Frameworks. Angew Chem Int Ed 55:3566-3579
  50. 50. Eddaoudi M, Kim J, Rosi N, Vodak D, Wachter J, O’Keeffe M, Yaghi OM (2002) Systematic Design of Pore Size and Functionality in Isoreticular MOFs and Their Application in Methane Storage. Science 295:469-472
  51. 51. Rowsell JLC, Millward AR, Park KS, Yaghi OM (2004) Hydrogen Sorption in Functionalized Metal−Organic Frameworks. J Am Chem Soc 126:5666-5667
  52. 52. Yao Q , Su J, Cheung O, Liu Q , Hedin N, Zou X (2012) Interpenetrated metal–organic frameworks and their uptake of CO2 at relatively low pressures. J Mater Chem 22:10345
  53. 53. Das MC, Xu H, Wang Z, Srinivas G, Zhou W, Yue Y-F, Nesterov VN, Qian G, Chen B (2011) A Zn4O-containing doubly interpenetrated porous metal–organic framework for photocatalytic decomposition of methyl orange. Chem Commun 47:11715
  54. 54. Perry IV JJ, Feng PL, Meek ST, Leong K, Doty FP, Allendorf MD (2012) Connecting structure with function in metal–organic frameworks to design novel photo- and radioluminescent materials. J Mater Chem 22:10235
  55. 55. Feldblyum JI, Wong-Foy AG, Matzger AJ (2012) Non-interpenetrated IRMOF-8: synthesis, activation, and gas sorption. Chem Commun 48:9828
  56. 56. Dong W, Sun Y-Q , Yu B, et al (2004) Synthesis, crystal structures and luminescent properties of two supramolecular assemblies containing [Au(CN)2]− building block. New J Chem 28:1347-1351
  57. 57. Pham HQ , Mai T, Pham-Tran N-N, Kawazoe Y, Mizuseki H, Nguyen-Manh D (2014) Engineering of Band Gap in Metal–Organic Frameworks by Functionalizing Organic Linker: A Systematic Density Functional Theory Investigation. J Phys Chem C 118:4567-4577
  58. 58. Li H, Eddaoudi M, Groy TL, Yaghi OM (1998) Establishing Microporosity in Open Metal−Organic Frameworks: Gas Sorption Isotherms for Zn(BDC) (BDC = 1,4-Benzenedicarboxylate). J Am Chem Soc 120:8571-8572
  59. 59. Eddaoudi M, Moler DB, Li H, Chen B, Reineke TM, O’Keeffe M, Yaghi OM (2001) Modular Chemistry: Secondary Building Units as a Basis for the Design of Highly Porous and Robust Metal−Organic Carboxylate Frameworks. Acc Chem Res 34:319-330
  60. 60. Ockwig NW, Delgado-Friedrichs O, O’Keeffe M, Yaghi OM (2005) Reticular Chemistry: Occurrence and Taxonomy of Nets and Grammar for the Design of Frameworks. Acc Chem Res 38:176-182
  61. 61. Xiong R, Odbadrakh K, Michalkova A, Luna JP, Petrova T, Keffer DJ, Nicholson DM, Fuentes-Cabrera MA, Lewis JP, Leszczynski J (2010) Evaluation of functionalized isoreticular metal organic frameworks (IRMOFs) as smart nanoporous preconcentrators of RDX. Sensors and Actuators B: Chemical 148:459-468
  62. 62. Odbadrakh K, Lewis JP, Nicholson DM, Petrova T, Michalkova A, Leszczynski J (2010) Interactions of Cyclotrimethylene Trinitramine (RDX) with Metal−Organic Framework IRMOF-1. J Phys Chem C 114:3732-3736
  63. 63. Petrova T, Michalkova A, Leszczynski J (2010) Adsorption of RDX and TATP on IRMOF-1: an ab initio study. Struct Chem 21:391-404
  64. 64. Xiong R, Keffer DJ, Fuentes-Cabrera M, Nicholson DM, Michalkova A, Petrova T, Leszczynski J, Odbadrakh K, Doss BL, Lewis JP (2010) Effect of Charge Distribution on RDX Adsorption in IRMOF-10. Langmuir 26:5942-5950
  65. 65. Deshpande RK, Minnaar JL, Telfer SG (2010) Thermolabile Groups in Metal-Organic Frameworks: Suppression of Network Interpenetration, Post-Synthetic Cavity Expansion, and Protection of Reactive Functional Groups. Angew Chem Int Ed 49:4598-4602
  66. 66. Yang L-M, Ravindran P, Vajeeston P, Tilset M (2012) Ab initio investigations on the crystal structure, formation enthalpy, electronic structure, chemical bonding, and optical properties of experimentally synthesized isoreticular metal–organic framework-10 and its analogues: M-IRMOF-10 (M = Zn, Cd, Be, Mg, Ca, Sr and Ba). RSC Adv 2:1618-1631
  67. 67. Gascon J, Hernández-Alonso MD, Almeida AR, van Klink GPM, Kapteijn F, Mul G (2008) Isoreticular MOFs as Efficient Photocatalysts with Tunable Band Gap: An Operando FTIR Study of the Photoinduced Oxidation of Propylene. ChemSusChem 1:981-983
  68. 68. Houk RJT, Jacobs BW, Gabaly FE, Chang NN, Talin AA, Graham DD, House SD, Robertson IM, Allendorf MD (2009) Silver Cluster Formation, Dynamics, and Chemistry in Metal−Organic Frameworks. Nano Lett 9:3413-3418
  69. 69. Cao A, Zhu W, Shang J, Klootwijk JH, Sudhölter EJR, Huskens J, de Smet LCPM (2017) Metal–Organic Polyhedra-Coated Si Nanowires for the Sensitive Detection of Trace Explosives. Nano Lett 17:1-7
  70. 70. Aubrey ML, Ameloot R, Wiers BM, Long JR (2014) Metal–organic frameworks as solid magnesium electrolytes. Energy Environ Sci 7:667-671
  71. 71. Sheberla D, Sun L, Blood-Forsythe MA, Er S, Wade CR, Brozek CK, Aspuru-Guzik A, Dincă M (2014) High Electrical Conductivity in Ni3(2,3,6,7,10,11-hexaiminotriphenylene)2, a Semiconducting Metal–Organic Graphene Analogue. J Am Chem Soc 136:8859-8862
  72. 72. Schröder M (ed) (2010) Functional metal-organic frameworks: gas storage, separation, and catalysis. Springer, Berlin ; New York
  73. 73. Li J-R, Kuppler RJ, Zhou H-C (2009) Selective gas adsorption and separation in metal–organic frameworks. Chem Soc Rev 38:1477-1504
  74. 74. Lee J, Farha OK, Roberts J, Scheidt KA, Nguyen ST, Hupp JT (2009) Metal–organic framework materials as catalysts. Chem Soc Rev 38:1450-1459
  75. 75. Alvaro M, Carbonell E, Ferrer B, Llabrés i Xamena FX, Garcia H (2007) Semiconductor Behavior of a Metal-Organic Framework (MOF). Chem Eur J 13:5106-5112
  76. 76. Takaishi S, Hosoda M, Kajiwara T, Miyasaka H, Yamashita M, Nakanishi Y, Kitagawa Y, Yamaguchi K, Kobayashi A, Kitagawa H (2009) Electroconductive Porous Coordination Polymer Cu[Cu(pdt)2] Composed of Donor and Acceptor Building Units. Inorg Chem 48:9048-9050
  77. 77. Civalleri B, Napoli F, Noël Y, Roetti C, Dovesi R (2006) Ab-initio prediction of materials properties with CRYSTAL: MOF-5 as a case study. CrystEngComm 8:364-371
  78. 78. Fuentes-Cabrera M, Nicholson DM, Sumpter BG, Widom M (2005) Electronic structure and properties of isoreticular metal-organic frameworks: The case of M-IRMOF1 (M=Zn, Cd, Be, Mg, and Ca). The Journal of Chemical Physics 123:124713
  79. 79. Kuc A, Enyashin A, Seifert G (2007) Metal−Organic Frameworks: Structural, Energetic, Electronic, and Mechanical Properties. J Phys Chem B 111:8179-8186
  80. 80. Kurmoo M (2009) Magnetic metal–organic frameworks. Chem Soc Rev 38:1353-1379
  81. 81. Guo Z, Cao R, Wang X, Li H, Yuan W, Wang G, Wu H, Li J (2009) A Multifunctional 3D Ferroelectric and NLO-Active Porous Metal−Organic Framework. J Am Chem Soc 131:6894-6895
  82. 82. Xie Y-M, Liu J-H, Wu X-Y, Zhao Z-G, Zhang Q-S, Wang F, Chen S-C, Lu C-Z (2008) New Ferroelectric and Nonlinear Optical Porous Coordination Polymer Constructed from a Rare (CuBr)∞ Castellated Chain. Crystal Growth & Design 8:3914-3916
  83. 83. Bureekaew S, Horike S, Higuchi M, Mizuno M, Kawamura T, Tanaka D, Yanai N, Kitagawa S (2009) One-dimensional imidazole aggregate in aluminium porous coordination polymers with high proton conductivity. Nature Mater 8:831-836
  84. 84. Hurd JA, Vaidhyanathan R, Thangadurai V, Ratcliffe CI, Moudrakovski IL, Shimizu GKH (2009) Anhydrous proton conduction at 150 °C in a crystalline metal–organic framework. Nature Chem 1:705-710
  85. 85. Sadakiyo M, Yamada T, Kitagawa H (2009) Rational Designs for Highly Proton-Conductive Metal−Organic Frameworks. J Am Chem Soc 131:9906-9907
  86. 86. Taylor JM, Mah RK, Moudrakovski IL, Ratcliffe CI, Vaidhyanathan R, Shimizu GKH (2010) Facile Proton Conduction via Ordered Water Molecules in a Phosphonate Metal−Organic Framework. J Am Chem Soc 132:14055-14057
  87. 87. Fletcher AJ, Thomas KM, Rosseinsky MJ (2005) Flexibility in metal-organic framework materials: Impact on sorption properties. Journal of Solid State Chemistry 178:2491-2510
  88. 88. Uemura K, Matsuda R, Kitagawa S (2005) Flexible microporous coordination polymers. Journal of Solid State Chemistry 178:2420-2429
  89. 89. Allendorf MD, Schwartzberg A, Stavila V, Talin AA (2011) A Roadmap to Implementing Metal-Organic Frameworks in Electronic Devices: Challenges and Critical Directions. Chem Eur J 17:11372-11388
  90. 90. Behera JN, D’Alessandro DM, Soheilnia N, Long JR (2009) Synthesis and Characterization of Ruthenium and Iron−Ruthenium Prussian Blue Analogues. Chem Mater 21:1922-1926
  91. 91. Sosa J, Bennett T, Nelms K, Liu B, Tovar R, Liu Y (2018) Metal–Organic Framework Hybrid Materials and Their Applications. Crystals 8:325
  92. 92. Zhang Z, Nguyen HTH, Miller SA, Cohen SM (2015) polyMOFs: A Class of Interconvertible Polymer-Metal-Organic-Framework Hybrid Materials. Angew Chem Int Ed 54:6152-6157
  93. 93. Parnham ER, Morris RE (2007) Ionothermal Synthesis of Zeolites, Metal–Organic Frameworks, and Inorganic–Organic Hybrids. Acc Chem Res 40:1005-1013
  94. 94. Tafipolsky M, Amirjalayer S, Schmid R (2007) Ab initio parametrized MM3 force field for the metal-organic framework MOF-5. J Comput Chem 28:1169-1176
  95. 95. Bureekaew S, Amirjalayer S, Tafipolsky M, Spickermann C, Roy TK, Schmid R (2013) MOF-FF - A flexible first-principles derived force field for metal-organic frameworks: MOF-FF - A force field for metal-organic frameworks. Phys Status Solidi B 250:1128-1141
  96. 96. Han SS, Choi S-H, van Duin ACT (2010) Molecular dynamics simulations of stability of metal–organic frameworks against H2O using the ReaxFF reactive force field. Chem Commun 46:5713-5715
  97. 97. Yang Q , Liu D, Zhong C, Li J-R (2013) Development of Computational Methodologies for Metal–Organic Frameworks and Their Application in Gas Separations. Chem Rev 113:8261-8323
  98. 98. Odoh SO, Cramer CJ, Truhlar DG, Gagliardi L (2015) Quantum-Chemical Characterization of the Properties and Reactivities of Metal–Organic Frameworks. Chem Rev 115:6051-6111
  99. 99. Grajciar L, Wiersum AD, Llewellyn PL, Chang J-S, Nachtigall P (2011) Understanding CO2 Adsorption in CuBTC MOF: Comparing Combined DFT–ab Initio Calculations with Microcalorimetry Experiments. J Phys Chem C 115:17925-17933
  100. 100. Chavan S, Vitillo JG, Gianolio D, et al (2012) H2storage in isostructural UiO-67 and UiO-66 MOFs. Phys Chem Chem Phys 14:1614-1626
  101. 101. Kumar A, Banerjee K, Foster AS, Liljeroth P (2018) Two-Dimensional Band Structure in Honeycomb Metal–Organic Frameworks. Nano Lett 18:5596-5602
  102. 102. Li J, Musho T, Bright J, Wu N (2019) Functionalization of a Metal-Organic Framework Semiconductor for Tuned Band Structure and Catalytic Activity. J Electrochem Soc 166:H3029–H3034
  103. 103. Weitzner SE, Dabo I (2017) Quantum–continuum simulation of underpotential deposition at electrified metal–solution interfaces. npj Comput Mater 3:1
  104. 104. Wilmer CE, Leaf M, Lee CY, Farha OK, Hauser BG, Hupp JT, Snurr RQ (2012) Large-scale screening of hypothetical metal–organic frameworks. Nature Chem 4:83-89
  105. 105. Witman M, Ling S, Anderson S, Tong L, Stylianou KC, Slater B, Smit B, Haranczyk M (2016) In silico design and screening of hypothetical MOF-74 analogs and their experimental synthesis. Chem Sci 7:6263-6272
  106. 106. Zhao R, Liang Z, Zou R, Xu Q (2018) Metal-Organic Frameworks for Batteries. Joule 2:2235-2259
  107. 107. Fu X, Yu D, Zhou J, Li S, Gao X, Han Y, Qi P, Feng X, Wang B (2016) Inorganic and organic hybrid solid electrolytes for lithium-ion batteries. CrystEngComm 18:4236-4258
  108. 108. Wiers BM, Foo M-L, Balsara NP, Long JR (2011) A Solid Lithium Electrolyte via Addition of Lithium Isopropoxide to a Metal–Organic Framework with Open Metal Sites. J Am Chem Soc 133:14522-14525
  109. 109. Yuan C, Li J, Han P, Lai Y, Zhang Z, Liu J (2013) Enhanced electrochemical performance of poly(ethylene oxide) based composite polymer electrolyte by incorporation of nano-sized metal-organic framework. Journal of Power Sources 240:653-658
  110. 110. Zhao X, Wu T, Zheng S-T, Wang L, Bu X, Feng P (2011) A zeolitic porous lithium–organic framework constructed from cubane clusters. Chem Commun 47:5536-5538
  111. 111. Banerjee D, Borkowski LA, Kim SJ, Parise JB (2009) Synthesis and Structural Characterization of Lithium-Based Metal−Organic Frameworks. Crystal Growth & Design 9:4922-4926
  112. 112. Ogihara N, Ohba N, Kishida Y (2017) On/off switchable electronic conduction in intercalated metal-organic frameworks. Sci Adv 3:e1603103
  113. 113. Erickson KJ, Léonard F, Stavila V, Foster ME, Spataru CD, Jones RE, Foley BM, Hopkins PE, Allendorf MD, Talin AA (2015) Thin Film Thermoelectric Metal-Organic Framework with High Seebeck Coefficient and Low Thermal Conductivity. Adv Mater 27:3453-3459
  114. 114. Yang L-M, Vajeeston P, Ravindran P, Fjellvåg H, Tilset M (2010) Theoretical Investigations on the Chemical Bonding, Electronic Structure, And Optical Properties of the Metal−Organic Framework MOF-5. Inorg Chem 49:10283-10290
  115. 115. Chueh C-C, Chen C-I, Su Y-A, Konnerth H, Gu Y-J, Kung C-W, Wu KC-W (2019) Harnessing MOF materials in photovoltaic devices: recent advances, challenges, and perspectives. J Mater Chem A 7:17079-17095

Written By

Hemali Rathnayake and Sheeba Dawood

Submitted: 10 June 2020 Reviewed: 05 October 2020 Published: 28 October 2020