Scalable and Sustainable Synthesis of Advanced Porous Materials

Jan 13, 2019 - precursors, the materials quality was observed to be on par .... recognized safety or health issues, and face regulatory control in num...
0 downloads 0 Views 2MB Size
Subscriber access provided by University of Winnipeg Library

Perspective

Scalable and Sustainable Synthesis of Advanced Porous Materials Shing Bo Peh, Yuxiang Wang, and Dan Zhao ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b05463 • Publication Date (Web): 13 Jan 2019 Downloaded from http://pubs.acs.org on January 14, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Scalable and Sustainable Synthesis of Advanced Porous Materials Shing Bo Peh,† Yuxiang Wang,† and Dan Zhao*† †

Department of Chemical and Biomolecular Engineering, National University of Singapore, 4

Engineering Drive 4, 117585, Singapore

Keywords: Porous materials; Scale-up production; Green synthesis; Continuous processing

*

D. Zhao. E-mail: [email protected]

1 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 92

ABSTRACT: Advanced porous materials (APMs) constructed from structure-encoded building blocks present an unprecedented degree for molecular design, and have demonstrated class-leading performance in

numerous

applications.

However, the

development of economically viable and sustainable synthetic means of production is a critical prerequisite to translate APMs to real-world applications. In this Perspective, we first consider the diverse synthetic routes to APMs – metal/covalent organic frameworks (MOFs/COFs), porous organic polymers (POPs), and porous molecular solids – and highlight the challenges for scalable and sustainable synthesis. Subsequently, we illustrate concepts to guide the sustainable synthesis of APMs with recent examples in the literature. Lastly, we present our outlook for the translation of molecular design to scalable APM production.

2 ACS Paragon Plus Environment

Page 3 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

INTRODUCTION Porous materials are materials which enclose void spaces into their bulk structures, thus offering extensive surface areas for interaction with guest atoms, ions, or molecules. This provides a natural advantage in applications based on such interactions, including adsorptive separations and catalysis. Significant advances have been made in the past decades to derive materials with highly porous structures. For example, activated carbon may be derived from chemical treatment of naturally-occurring carbonaceous materials with specific surface areas up to 3,000 m2 g-1.1, 2 Naturally obtained or synthetic zeolites with ordered porosities are widely applied in petrochemical industry for size-selective transformations.3, 4 Mesoporous silica with large pores (2-10 nm) may be prepared by surfactant templating for applications including drug delivery and catalysis.5, 6 Despite the well-established nature of these porous materials in commercial applications, precise control over the porous architecture and functionalization to accommodate more demanding host-guest interactions remains a persistent challenge. A recent wave in porous materials development is the utilization of discrete molecular building blocks (MBBs) to prepare porous structures with versatile chemical

compositions.7-10

Beginning

with

coordination

chemistry,

metal-organic

frameworks (MOFs) exhibiting crystalline order or discrete coordination cages (metalorganic cages, or MOCs) may be prepared with defined geometries.11, 12 The adoption of dynamic covalent chemistry leads to their organic analogues, covalent organic frameworks (COFs), porous organic cages (POCs) as well as amorphous but porous structures, referred herein as porous organic polymers (POPs, Figure 1).13-16 To emphasize the common modular design philosophy behind the development of these materials, the term advanced porous materials (APMs) has been adopted. Table 1 summarizes the APMs into four broad categories, on the basis of their compositional character, bonding, long-range order and 3 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 92

solubility. We have lumped MOCs and POCs together in the category of porous molecular materials (PMMs) based on their soluble nature, which allows different treatment from the other APMs.

Figure 1. Modular design of APMs has the potential to reach benchmark performance in a wide-variety of host-guest interaction-dependent applications. (top row, left to right) Examples of APMs include MOFs, COFs, PAFs, POCs, and MOCs. Reproduced with permission.17 Copyright 2018, John Wiley and Sons. (bottom row) Engineered molecular interactions within the pores of APMs may be exploited for different applications. Gas storage image: Reprinted by permission from Nature Chemistry.18 Copyright 2011, Springer Nature. Gas separations image: Reproduced with permission.19 Copyright 2018, American Chemical Society. Sensing image: Reproduced with permission.20 Copyright 2016, American Chemical Society. Catalysis image: Reproduced with permission.21 Copyright 2012, American Chemical Society.

4 ACS Paragon Plus Environment

Page 5 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Table 1: Characteristics of APMs Type of APMs Organic/inorganic character Bonding Long-range order Solubility

MOFs Hybrid

COFs Organic

POPs Organic

Coordination bonding Yes Insoluble

Covalent bonding Yes Insoluble

Covalent bonding No Insoluble

PMMs Hybrid (MOCs)/ Organic (POCs) Coordination/ covalent bonding Variable Soluble

Modularity of design has been exploited to engineer APMs with ideal attributes for the targeted applications. For instance, multivariate MOFs (MTV-MOFs) with statistical linker distributions may lead to highly synergistic effects over the performance.22 Strategies have been developed for site specific installation of distinct pore functionalities.23-25 Furthermore, topology and composition may be rationalized, which is the foundation of reticular chemistry.11 The improved freedom in design has allowed engineered APMs to emerge as class-leading materials in numerous applications. Early development was centred around gas storage, which capitalizes on the ultrahigh porosity reported for MOFs (>7000 m2 g-1),26-28 COFs (>4000 m2 g-1),29, 30 and PAFs (>5000 m2 g-1).31-33 For instance, the MOF HKUST-1 currently holds the record for methane uptake after packing and densification.34 Precise adjustment of pore sizes with narrow increments (0.2 to 1 Å) is achieved on fluorinated MOF adsorbents, which enables molecular exclusion in olefin/paraffin separations.35,

36

Alternatively, selective separation may be realized using membranes containing APM fillers.37 APMs with installed photoactive,38 chiral,39 enzymatic40, 41,40, 41 and metallic centres have shown exceptional catalytic performance.42 Based on the ongoing development, APMs are poised to dominate applications requiring increasing levels of control over the host-guest interactions. In light of the high commercial interest, scale-up processes considering both sustainability and scalability are essential to obtain the large product quantities necessary for

5 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 92

industrial applications. For example, a technoeconomic analysis for MOF-based gas storage systems was conducted recently.43 It was found that a single small-sized vehicular system for natural gas storage will require approximately 50 to 65 kilograms of MOFs on the basis of a material density of 1 g cm-3. For a demand level of 1000 systems per year and 500,000 systems per year, between 50 to 2500 tonnes of MOFs are required.43 Lab-developed protocols are optimized for purity and ease of characterization, so as to clearly elucidate the structure-property relationships. In comparison, the transition to industrial processes requires adjustments for safety, logistical and cost reasons, which necessitate synthetic compromises.44-46 Although several reviews have been published covering the sustainable preparations for MOFs,47, 48 scale-up is rarely considered for other APMs. The shared MBB design philosophy underlying APM synthesis implies that lessons drawn from the synthetic development of MOFs may be generalizable to other categories of APMs. The objective of this Perspective is hence to consider the synthetic differences and similarities of APMs relevant to scale-up development. First, we describe key synthetic challenges of the various materials, emphasizing the considerations relevant to scale-up development. Next, we highlight green chemistry concepts and production strategies which are readily transferable. Finally, we present our outlook for the industrial implementation of APMs.

6 ACS Paragon Plus Environment

Page 7 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

SYNTHETIC CONSIDERATIONS FOR APMs Synthetic aspects of MOFs

Figure 2. (a) Common cluster MBBs employed for the reticular design of MOFs. Reproduced with permission.44 Copyright 2013, AAAS. (b) Rational isoreticular expansion of csq zirconium MOFs. Reproduced with permission.49 Copyright 2018, Elsevier. (c) Trinuclear, octanuclear, dodecanuclear and chain-based clusters are examples of aluminium SBUs arising due to progressive in-situ hydrolysis of linker salts. MOFs are a class of coordination compounds composed of organic ligands and metal ions or clusters. The pioneering work of Robson and Hoskins50 recognized that high 7 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 92

reversibility of coordination bonding enabled an error-checking mechanism towards the spontaneous self-assembly of infinite 3D frameworks. The use of shape-retaining metal-oxide carboxylate clusters as molecular building blocks (MBBs, Figure 2a) enabled the creation of robust frameworks with high porosity.51, 52 The reticular design philosophy (Figure 2b) as well as the topological analysis supporting the molecular design of functional MOFs has been discussed in detail in the literature.44, 53 Biswas and Stock summarized various aspects of MOF synthesis and the application of alternative synthetic technologies.54 Recent reviews have considered the MOF synthetic developments from the perspective of commercialization efforts. For example, Julien et al. highlighted the evaluative criteria and major growth areas with regards to scalable and sustainable MOF synthesis.48 Also in 2017, Rubio-Martinez et al. surveyed novel synthetic routes for MOFs relevant to scale-up processes.55 Benefiting from the increased interest directed towards the scale-up of MOF materials, HKUST-1(Cu),56 ZIF8 (Zn, Co),57 MOF-74 series (M = Zn, Mg, Co, Ni, Mn, Fe, and Cu),40 SIFSIX-3(Zn),58 MIL53-fumarate(Al),59 MIL-100(Fe),60 CAU-10(Al),61 and certain analogues of UiO-66 have attained commercial production through a diversity of manufacturing processes.62 However, these remain a small subset of the thousands of MOFs reported. Inevitably, faithful replication of lab-scale solvothermal synthesis at scale-up volumes is unfeasible and undesirable for safety and economic reasons. Nevertheless, it is instructive to examine the potential of implementing defining MOF design strategies – including topological control, pore expansion and functionalization9, 22, 25, 40 – at commercially relevant throughputs. We identify in this section the divergence of reticular chemistry assumptions with realistic process conditions and emphasize the most promising strategies to design functional MOFs with scale-up compatible production. Synthetic compromises associated with the scale-up production of MOFs

8 ACS Paragon Plus Environment

Page 9 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

The landmark case studies of MOF design demonstrating functionality variation,11, 22 and pore size expansion have invariably relied on linker-centric strategies to modify the materials.40 In contrast, linker diversity in commercial portfolios is dramatically reduced, with restrictions on both the length and functionality. The linkers used to successfully prepare MOFs in scale-up quantities include fumaric acid, pyrazine, trimesic acid, isophthalic acid, and few functionalized variants of terephthalic acid. To the best of our knowledge, only two examples of commercialized MOFs (PCN-250(Fe) and MOF-177(Cu)) incorporate linkers with more than one phenyl ring.55 Notwithstanding the increased costs to produce extended linkers, a further reason is the impaired solubility of larger ligands which is exacerbated by reduction, elimination, or replacement practices imposed at production-scale which restrict the adoption of de-facto polar aprotic solvents.47,

63

Consequently, homogeneous

crystallization protocols which are ubiquitous for materials discovery are impracticable without severely compromising production efficiencies. Thus, alternative approaches for MOF design are required to meet the demand for functional porous materials by the industry. A second departure from reticular chemistry is the violation of cluster shape-persistence under scaled-up synthetic conditions. Pronounced cluster nuclearity variation is inherent in MOFs constructed from high valence cations such as Al3+, Cr3+, and Zr4+, which are expected to exhibit improved stability owing to the stronger metal-oxo bonds as predicted by the HSAB (hard/soft acid/base) theory.64 Nuclearity drifts arise from the progressive hydrolysis of cations in the presence of water. The water involvement is due to its direct inclusion as reaction media, appearance as impurity in organic solvents, or adsorption from the atmosphere as a result of precursor hygroscopicity. During hydrolysis, the metal ions are bridged by hydroxide ions to form polynuclear species, which lowers the overall charge density. The assembly of such high nuclearity clusters into extended frameworks leads to increasingly dense structures with reduced porosity. As a well-studied category of MOFs, 9 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 92

aluminium MOFs (Figure 2c) have been isolated bearing a wide range of MBBs with differing nuclearities and connectivities. Furthermore, MOFs with oxide chain-based SBUs have been observed with a variety of cations, such as MIL-53(Al, Fe, Cr),65 MIL-140(Zr),66 and MIL-177-HT(Ti) (MIL = Materials of Institut Lavoisier).67 This suggests that chainbased MBBs are a common class of structural pseudo-motifs which are targetable by intentional hydrolysis. Although highly porous high-valence MOFs have been discovered, such as MIL-10168 (trinuclear Al/Cr MBB) and MIL-12569 (octanuclear Ti MBB), their reticular functionalization or extension is limited in scope, and extensive optimization should be expected in the isolation of such kinetic phases. Strategies to target kinetic MOF phases, in particular the use of stable multinuclear clusters as precursor reagents,70 have been reported. While successful for Fe3+-based systems, challenges in isolating stable clusters and hydrolytic tendency of more reactive metals71 suggest limited generality in these methods. “Scale-up compatible” design strategies for engineering of functional MOFs The enduring limitations of solubility and hydrolysis imply that alternative design strategies are necessary to realize reticular design at industrially-relevant scales. To this end, variation of the inorganic component appears to be the most feasible means of tuning materials properties. The recent synthetic development regarding two sub-categories of MOFs, namely fluorinated MOFs and rare-earth MOFs, embody this focus. The former category is based on anionic XF62- (X = Si or Ti)35, 58, 72 or XF52- (X = Al, Fe, NbO)36, 73 pillared structures. The rational choice of cation, anion and organic linker combinations allow fine-tuning of pore sizes and chemistry within a very narrow range, allowing benchmark performance in several commercially-important separations. The first-generation fluorinated MOFs incorporating hexafluorosilicate or hexafluorotitanate have attained large-scale production.55 Under the latter category, recent advances using ortho-substituents of carboxylate ligands to favour the in-situ formation of desired clusters have enabled the 10 ACS Paragon Plus Environment

Page 11 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

synthetic preparation of frameworks with predictable topologies.74, 75 The similar chemical properties of the rare-earth cations allow the preparation of isoreticular frameworks without adjusting the reaction conditions, which affords a substantial degree of tunability even when ligand choices are restricted. Another possible direction is to utilize inorganic MBBs with high degrees of condensation. As observed in higher valence MOFs, the thermodynamicallystable chain-based MBBs of MIL-53,76 CAU-10,77 and MIL-14078, 79 offer the greatest scope for reticular chemistry at the expense of their reduced porosity.54, 55 For Al3+ and Ti4+, their low weight, natural abundance, and for the latter, photocatalytic activity, remain valuable attributes justifying the substantial commercial interest for the development of these frameworks. Highly-condensed clusters may also be relevant for divalent metals. For instance, self-assembly of nickel isonicotinates under harsh reaction conditions (>150 °C, >48 h) produces hydrolytically stable materials with good carbon capture performance.80-82 However, the control of polymorphs is a potential issue which demands consideration during the synthetic development.

11 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 92

Synthetic aspects of COFs

Figure 3. (a) Error correction mechanism for self-assembly of COFs distinguishes them from amorphous polymeric materials. Reproduced with permission. Copyright 2018, American Chemical Society.83 (b) Representative coupling reactions and synthetic landmarks of COFs. Adapted with permission.84 Copyright 2018, John Wiley and Sons. COFs are covalently-linked frameworks with porosity and long-range order.45, 85 COF crystallinity distinguishes them from conventional amorphous polymers and facilitates direct comparison with MOFs as tailorable materials at the molecular level. The wholly organic 12 ACS Paragon Plus Environment

Page 13 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

character of COFs presents the advantages of lower densities and possibly high charge carrier mobility over metal-containing frameworks.25, 59, 60 In particular, the organic semiconductor behaviour in some COFs has attracted substantial research interest in photocatalytic applications, where the crystallinity is believed to enable enhanced performance compared to other amorphous organic materials.86-88 In principle, a wide range of coupling reactions with varying extent of reversibility may be used to prepare COFs (Figure 3). By utilizing more robust C-N or C-C related coupling reactions, chemical stabilities surpassing carboxylate MOFs may be attained. For instance, benzoxazole and enamine-linked materials have been shown to tolerate both concentrated acidic and basic solutions, increasing their ability to operate in demanding applications.87, 89,

90

Although the development of COFs has yet to

reach full maturity, their distinct attributes from both MOFs and conventional polymers enable these crystalline organic materials to occupy a unique niche among APMs, which justifies the intense interest in their eventual commercialization. Synthetic challenges for the production of COFs In analogy to the MOF formation process, crystalline ordering arises in COFs due to a reversibility-induced error-correction mechanism, where monomers are oriented in a specific direction within the growing framework to reduce the overall energy.83, 91, 92 The reaction conditions must be controlled to balance the condensation kinetics with equilibration time necessary to obtain an ordered configuration. Experimentally-observed amorphous-tocrystalline transitions in imine COFs and covalent triazine frameworks (CTFs),74, 75 as well as successful building block exchange confirm the existence and feasibility of error-correction.93 The successful preparation of COF single crystals for boronate ester and imine-based systems evidences the expanding toolbox for the precise control of COF growth and nucleation.94, 95 In the first study, the use of nitrile-containing cosolvents to suppress nucleation allowed the preparation of 2D boronate ester COF single crystals from colloidal seed precursors.65 In the 13 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 92

second study, taking inspiration from MOF crystallization, aniline was used as monodentate monomer to serve as a competitive reagent to inhibit nucleation, which enabled the slow growth of 3D imine COF single crystals.95 Although these studies heralded significant progress in the preparation and characterization of COF materials, a large gap still remains before COFs are viable for scale-up production. The first issue relates to the time frame necessary for error correction. The lability for covalent bonds is significantly lower than metal-ligand coordination, which significantly extends the required preparation duration. For example, a recently reported synthesis of imine-linked COF-300 requires 3 to 9 days to achieve microcrystalline product with good crystallinity.96 COF materials synthesized over shorter periods typically exhibit lower degree of long-range order. The second complication specific to 2D COFs is the reliance on interlayer stacking to achieve long-range order in the out-of-plane direction. While in-plane regularity is maintained by strong covalent bonds used for the building block coupling, the strength of the interlayer stacking is governed by weaker secondary interactions such as π-π interactions, electrostatic interactions and hydrogen bonding. As such, the order of 2D COFs in the out-of-plane direction suffers from increased susceptibility to small changes in the chemical environment, which may arise from building block variation during isoreticular synthesis. For instance, Jiang’s group observed the destabilization of stacked layers due to electrostatic repulsion of polarized C=N bonds, which was alleviated by the incorporation of electron-donating methoxy groups.39 The engineering of H-bonding motifs within the COF structure was also found to benefit the crystallinity.97, 98 However, the explicit installation of functional motifs to address stacking problems consumes valuable synthetic resources but contributes only indirectly to the overall material functionality. Hence, it is desirable to search for more convenient solutions. Novel strategies for the synthesis of COFs

14 ACS Paragon Plus Environment

Page 15 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

The ongoing synthetic development has turned to increasing the complexity of the reaction system in order to afford finer control over the condensation reactions. In particular, the in-situ generation of reactive species, and the adoption of catalysts in coupling reactions are emphasized here as promising strategies for COF synthesis which can promote further scale-up development. The first strategy addresses the limitations associated with slow COF error-correction. The insolubility of the product provides the thermodynamic driving force for the covalent assembly of COF monomers. Restricting the availability of monomers helps to control the reaction equilibration, thus facilitating error correction. Dropwise addition of reagents is a common practice at the lab-scale and crucial to the recent success in obtaining 2D single-crystalline COF domains.94 The feasibility of this approach at larger-scales is however questionable. Therefore, reactivity control at the precursor level may extend degrees of freedom in the actual synthesis process. The formal transimination of benzophenone-based monomers was recently demonstrated for the formation of 2D imine COFs.99 Zhao and coworkers introduced primary amines into a β-ketoenamine COF system to transform the rapid condensation route to an imine exchange-based mechanism.100 Furthermore, crystalline covalent triazine frameworks (CTFs) were prepared by generation of aldehyde monomers through the in-situ oxidation of alcohols in dimethyl sulfoxide (DMSO).101 The synthetic conditions applied in this approach (100 °C for 24 h, then 180 °C for 36 h) are significantly milder than de-facto ionothermal synthesis (ZnCl2, 400 °C),102 and may be performed in an open system. Different from the deliberate introduction of protecting groups, benzophenone and alcohol precursors arise in the synthetic process for the monomers, and hence cost savings may be realized through vertical integration of the manufacturing processes. Being essentially a one-pot, multi-step process, the participation of side-products may be an additional concern.

15 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 92

Secondly, the application of catalysts for the acceleration of reaction rates is an aspect of green chemistry which is highly relevant for COF preparation, in stark contrast with the spontaneous nature of metal-linker coordination. The extension of reticular chemistry towards the utilization of ultra-stable organic connections would hardly be possible without catalytic mediation of the exchange equilibria. The COF domain has increasingly taken inspiration from ongoing research in organic synthesis to extend the scope and viability of framework formation reactions. Moving beyond the acetic acid catalysed protocols adopted for the early imine COFs, Matsumoto et al. applied Lewis-acidic catalysis of transimination reactions, which can dramatically reduce the synthetic duration and temperature.103, 104 The exceptional rate enhancement facilitates COF film formation by interfacial polymerization, which allows their use as membranes.105-107 Nucleophilic catalysts have also been employed in transimination reactions.108, 109 However, their catalytic effect in APM syntheses is still uncertain.100 Unlike conventional dynamic covalent chemistry (DCC) libraries, the reaction of multitopic monomers with nucleophilic catalysts (i.e. monodentate amines) may lead to the formation of insoluble intermediates.110, 111 Ionic liquids (IL), which are organic salts with reduced melting points, have the potential to simultaneously fulfil roles as catalyst and reaction media. Qiu’s group reported IL-assisted syntheses for diamondoid 3D imine COFs with hugely accelerated reaction rates.112 While the use of ILs is generalizable to other materials, including MOFs,113 their widespread use in organic catalysis is a significant advantage for covalently-linked APMs. Baek’s group developed a phosphorous pentoxide (P2O5) catalysed cascade reaction for the preparation of crystalline CTFs from aromatic amides, representing an additional alternative to ionothermal synthesis.114 In the preliminary proof-of-concept studies, sustainability of the catalysts has yet to be considered to a large extent. Nevertheless, their substoichiometric inclusion should not exempt their evaluation

16 ACS Paragon Plus Environment

Page 17 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

under the principles of green chemistry. It is foreseeable that toxicity, efficiency and handling will emerge as relevant factors during synthetic development. Synthetic aspects of POPs POPs are constructed from organic monomers. Although possessing high dimensional (2D and 3D) extended structures, they differ from COFs in their lack of long-range order. POPs may be classified into further sub-groups including polymers of intrinsic microporosity (PIMs), conjugated microporous polymers (CMPs), hyper-crosslinked polymers (HCPs), and porous aromatic frameworks (PAFs).115-118 POPs share numerous common aspects with MOFs and COFs with respect to molecular design, despite their lack of crystallinity. The increase in topicity of the monomer building units, as well as rigidity of the building blocks compared to typical polymers can achieve highly competitive porosities. For instance, PAF-1 reported in 2009 exhibits a remarkable Brunauer–Emmett–Teller (BET) surface area of 5640 m2 g-1, rivalling that of the most porous MOFs.31 The shared organic character of COFs and POPs imply that much of the synthetic considerations summarized in the previous section are applicable to POPs. The relaxation of the condition of crystallinity allows POPs to access a wider synthetic toolbox. In particular, C-C coupling reactions, such as Suzuki reaction,119 Yamamoto reaction,31, 120 Sonogashira-Hagihara (S-H) reaction,121, 122 and oxidative coupling have been applied successfully towards the formation of these APMs.123 The cyclotrimerization reactions are also alternatives to synthesize POPs.102 Microporous POPs have exhibited highly selective uptake of small gas molecules which is promising for practical carbon capture applications, although the maximum uptake capacities remain lower than MOF sorbents.124-127 On the other hand, POPs benefit from exceptional physicochemical stability relative to materials constructed by reversible linkages, which has allowed flexibility in postsynthetic functionalization even for ultra-porous materials. Harsh treatments, such as lithiation and sulfonation, have been successfully carried out on PAF-1 and PPN-4 leading to 17 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 92

impressive gas uptake performance.128-130 The extended conjugation of CMPs implies potential applications in sensing, capacitance and photocatalysis.131-134 While POPs have yet to attain benchmark performance, their tunability, unique attributes and stability indicate considerable promise in the future. Challenges to the synthesis and scale-up production of POPs Numerous synthetic challenges exist preventing the further scale-up of POPs. In line with the development trajectory of other APMs, the existing studies have focused on evaluating design strategies and optimizing for performance. Utilizing porosity as a metric, it is apparent that the condensation parameters have major influence over the product quality. Cooper’s group explicitly evaluated the influence of solvent choice in the S-H coupling of different CMPs, observing up to 4-fold enhancement in surface area when the condensation was performed in an optimal solvent.135 The coupling approach is observed to have overriding influence, even when choosing similar monomers. Recently, Ben et al. compiled the specific surface areas of PAFs prepared by different coupling approaches; the PAFs exhibited significant porosity variance with Yamamoto coupling being the most effective method.16 The noble-metal catalysed approaches have the advantage by involving accessible monomers, being widely adopted in the pharmaceutical industry and in the synthesis of fine chemicals. For use as adsorbents, the unit cost of material is required to be low, which makes the economic feasibility of noble-metal catalysis questionable. As an alternative to the Yamamoto, Suzuki and S-H couplings, a few recent examples have been reported using freeradical induced polymerization,136 Diels-Alder cycloaddition,137 and azo coupling138 in the absence of metal catalysis. The diverse synthetic toolbox of POPs may inspire the development of new routes to achieve similar topological structures, with CTFs serving as a good case study of this development. Moving from the original ionothermal synthesis,102 amorphous CTFs were prepared using superacid catalysis at mild conditions.139 Subsequently, 18 ACS Paragon Plus Environment

Page 19 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

cascade Schiff-base and Michael addition condensation approach enabled a mild basecatalysed gram-scale synthesis of amorphous CTF materials.140 Mild synthesis conditions compared to the high temperature ionothermal synthesis are essential to avoid carbonization, which unveils the fluorescence and photocatalysis performance of these materials. 141 POP synthesis has made a substantial progress in transitioning to industrially-accessible synthetic conditions. Nevertheless, the control of monomer and catalyst costs, as well as the optimization of reaction efficiencies represent the next step towards scale-up production. Synthetic aspects of porous molecular materials

Figure 4. (a) Assembly strategies for MOCs. Subcomponent self-assembly image: Reproduced with permission.142 Copyright 2008, John Wiley and Sons. Middle cage image: Reproduced with permission.143 Copyright 2014, Royal Society of Chemistry. Catenated cage image: Reproduced with permission.144 Copyright 2018, John Wiley and Sons. (b) Coupling

19 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 92

reactions employed for the synthesis of POCs. Reprinted by permission from Nature Reviews Materials.46 Copyright 2016, Springer Nature. Porous molecular materials (PMMs) may be formed by solution self-assembly using coordination linkages (e.g., metal-organic cages, MOCs) or covalent bonds (e.g., porous organic cages, POCs). A diverse synthetic toolbox has been developed to prepare coordination cages with ever-increasing complexity (Figure 4a). The early development of coordination MOCs is based on directional bonding approach similar to MOFs, which forms highly symmetric products with classical Archimedean and Platonic geometries.145 The elaborate synthesis needed to prepare ligands with the desired directionality is a drawback to this approach. The need to derive considerable molecular complexity from simple and accessible precursors has benefited from two distinct strategies. Firstly, taking inspiration from MOF chemistry, the use of multinuclear clusters as edges or vertices result in elaborate polyhedral structures, even with relatively simple linker geometries.146 Alternatively, the reversibility of coordination enables use of orthogonal interactions or reactions to govern the assembly of multicomponent mixtures to form elaborate thermodynamic products. As an example of the latter, Nitschke and coworkers developed a subcomponent assembly approach to generate in-situ iminopyridyl ligands from simple aldehyde and amine precursors, yielding a diverse array of structures encompassing tetrahedra, cubes and less symmetric architectures.147 The template-controlled interpenetration of coordination cages provides another route to structural diversity.144, 148, 149 Mirroring the development in COFs, POCs utilize reversible organic couplings such as imine condensation, alkyne metathesis and boronate ester condensation (Figure 4b) to form porous molecular solids.46 The solubility attribute of POCs is advantageous for their processing into films or membranes. CC3, as a particularly robust example of POC, has been evaluated in separation and catalysis.14,

150-152

A further interesting development involved 20

ACS Paragon Plus Environment

Page 21 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

POCs with exceptionally high solubilities, which are able to form freely-flowing liquids exhibiting permanent porosity.153, 154 The ongoing work in POCs remains concerned with implementation

of

molecular-level

design

and

elucidating

the

structure-function

relationship.155-158 The extreme sensitivity of POC synthesis to factors such as geometry, sterics and building block reactivity require address before the materials may meet the requirements of scale-up synthesis. The soluble nature of PMMs simplifies their processing in comparison to extended framework APMs. Doonan’s and Cooper’s groups have reported continuous flow protocols of organic cage molecules using alkyne homocoupling and Schiff-base reactions performed on commercially-available flow platforms.111, 159 The reaction homogeneity facilitates multigram synthesis. However, successful self-assembly of cages alone is insufficient to derive functional materials. The coordination or covalent bonding in PMMs ensures shape-defined pore cavities, whereas weaker secondary interactions control the intermolecular packing. The variable packing of the cages has considerable influence over the bulk material properties, but it is difficult to ensure shape persistence of PMMs against solvent evacuation or exposure to acidic or basic environments.155,

160

Post-assembly modification strategies such as

crosslinking may aid the retention of porosity and even macroscopic structure. For example, Coskun’s group synthesized organic cages based on 1,3,5-tris(p-hydroxyphenyl)benzene and induced further polymerization to convert these cages into extended nanoporous polymers by thermal treatment.161 Nam et al. prepared stable metal-organic polyhedra (MOPs) based on zirconium and crosslinked the products in crystalline packed state using aliphatic acyl chloride linkers.162 The crosslinked CLMOP-1c exhibited decreased solubility in methanol although the material became amorphous after activation. The problems related to desolvation of porous molecular solids may be circumvented by in-situ crystallization in the functional devices or composites. Organic polymer/CC3 composite membranes were 21 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 92

prepared by combining PIM-1 and POC solutions before solvent evaporation.163 Mixedmatrix membranes incorporating MOP-18 have been prepared using a similar method.164 We foresee that post-synthetic processing will constitute an integral part in the manufacturing process for PMM-based materials, for which the present research attention remains lacking.

22 ACS Paragon Plus Environment

Page 23 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

CONCEPTS IN SUSTAINABLE SYNTHESIS OF APMS Selection of building blocks Selection of building blocks is a key consideration in APM synthesis, given their influence over the product structure and functionality. Atom economy is a key principle of green chemistry introduced by Trost, which considers the reaction efficiency in terms of the extent whereby reactants are assimilated into the final product.165 This principle advocates that the usage of auxiliary materials is minimized where possible. Being fundamentally a coupling or condensation of molecular building blocks, a first source of wastage involves the leaving group of the precursors, for example halides in coupling reactions or the balancing anions in metal salts. This wastage is inherent even at quantitative yields. Moreover, additives for reaction control may contribute a significant portion of wastage. For example, monocarboxylate modulators for zirconium MOFs are employed at super-stoichiometric ratios (20 to >100 eq.) for one-pot synthesis.166 However, optimization of atom economy should be in light of other criteria such as process robustness and economic feasibility. Alternative metal precursors for the synthesis of APMs Traditional studies on coordination chemistry and organic-inorganic hybrid materials rely on the introduction of metal precursor as salts, which are typically highly soluble. The counter-anions are potential ligands and may compete with the organic linkers for coordination to the metal atoms. The reactivity of the salt may hence be proxied by the ability of the counter-anion to penetrate the coordination sphere. For example, in the case of Cr3+, the coordination tendency (from weakest to strongest) follows the order ClO4- > NO3- > Cl- > SO42- > HCOO-.167 Weak coordination activity of counter-anions is preferred to minimize disruptions to the assembly process. However, perchlorates, nitrates, or chlorides introduce process concerns at scale-up levels due to safety issues with oxidation or corrosion.59 23 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 92

Sulphate salts may avoid the corrosion problems and safety hazards while retaining moderate solubility for solution-based processes. Aluminium sulphate was successfully used as the precursor for the aqueous tonne-scale production of MIL-53-fumarate by BASF.59 However, a similar attempt to synthesize UiO-66-NH2 and UiO-66-(COOH)2 using zirconium sulphate yielded unexpected 8-fold connected bcu frameworks due to increased coordination tendency of the anion.47 The use of linker in large excess can avoid the reduced connectivity and anion occlusion, but the reactant efficiency was sacrificed in this approach.168 The direct use of oxide or hydroxides, which can be harvested from natural sources, can further reduce the cost of the metal precursor. Despite their low solubility, the transformation of oxides and hydroxides to MOFs was observed to be possible in a few rare cases. Benefiting from the exceptional lability of copper cations, HKUST-1 may be derived from variety of coppercontaining metal sources, including hydroxide,169 hydroxy double salts,170 and carbonate minerals.171 In these reports with lab-prepared or high purity commercial precursors, the materials quality was observed to be on par with products of conventional syntheses. However, for the last study, when HKUST-1 was prepared from naturally occurring minerals, the reduction of surface area (from ~1600 m2 g-1 to 340 ± 5 m2 g-1 and 740 ± 5 m2 g-1) was observed and attributed to the presence of insoluble contaminants. Helms’ investigation of the preparation of M2(dobpdc) MOFs (dobpdc = 4,4’-dioxido-3,3’-biphenyldicarboxylate) from different divalent metal oxide precursors revealed a dissolution-controlled crystallization mechanism governing the transformation.172 This suggests that additional process adjustments such as size control of the precursors are necessary to overcome the lower solubility and reactivity of the replacements. Expectedly, successful examples involving the less reactive metal cations are less common. Hydrothermal synthesis of MIL53(Al) has been attempted using aluminium-containing metal oxide and hydroxide precursors.173 In a further development, Walton’s group used aluminium carbide for this

24 ACS Paragon Plus Environment

Page 25 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

synthesis which led to a needle-like morphology.174 The slow kinetics of the abovementioned solution-based conversion – requiring extended duration (>24 h) and high temperatures (>200 °C) – must be overcome in order to be industrially feasible. The harsh conditions may also promote rapid hydrolytic polymerization of the metal oxoclusters, which is not conducive towards the formation of non-chain-based frameworks. Recently, 2-dimensional transition metal dichalogenides (TiS2) was employed as the metal source for the synthesis of titanium MOFs MIL-125-NH2 and MIL-167, furthering widening the pool of possible metal sources.175 The disulphide precursor is more air-stable compared to other titanium sources. However, the release of hydrogen sulphide during the solvothermal transformation remains a sustainability concern.

Figure 5. (a) Schematic representation of metal-oxide based vapour processing method for the fabrication of ZIF-8 thin films. (b) Comparison of thin film PXRD patterns with simulated pattern. (c) AFM depth profile of ZIF-8 film. (d) HAADF TEM elemental analysis of incompletely converted ZIF-8 film. Reprinted by permission from Nature Materials.176 Copyright 2015, Springer Nature. 25 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 92

The use of less reactive precursors may not always entail synthetic compromises. Leveraging well-established techniques for oxide deposition, oxide-to-MOF transformation is useful to prepare patterned MOF films and coatings (Figure 5). Inspired by existing solventfree protocols,177, 178 Ameloot and coworkers prepared ZIF-8 films using liquid-phase179 and vapour-phase transformations.176 Recently, Tsapatsis and coworkers prepared continuous ZIF-8 film membranes for propylene/propane separations using vapour-phase conversion.180 The use of nonporous ZnO precursor avoids the formation of defects which is frequently encountered using solution-phase direct growth methods, thereby enhancing the scale-up reliability.180 Developing sustainable organic building blocks for the synthesis of APMs The feasibility of isoreticular synthesis in APMs enables building block replacement to be a viable strategy in increasing process sustainability. Substitutions may be implemented provided the assumptions of isoreticular synthesis remain valid. In particular, a degree of rigidity in the building block is crucial to maintain directionality of the bonding and avoid undesired structural outcomes during synthesis. Phenyl-based building blocks, which are derivatised from the petrochemical industry, are commonly used because the linker geometry is enforced by aromaticity. For example, the prototypical building block for carboxylate MOFs – terephthalic acid – is manufactured from the oxidation of p-xylene with a price of about US$1,000/mt. Biomass conversion is another route to derive building blocks with increased sustainability. In 2004, the U.S. Department of Energy (D.O.E) identified 12 platform chemicals as targets for biotransformation from organic sugars.181 In particular, 1,4diacids (succinic, fumaric, and malic acid) and 2,5-furandicarboxylic acid (FDCA)182 possess established pathways owing to strong demand for their derivative products. As dicarboxylates, fumaric acid and FDCA have already been explored for MOF synthesis.59, 66, 183, 184

26 ACS Paragon Plus Environment

Page 27 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

A few reports of linker substitution are available in the literature. For example, iron fumarate (MIL-88A) may be used as nanocarriers for drug encapsulation, although its performance is surpassed by more porous structures.185 Aluminium fumarate (MIL-53) is one of the earliest MOFs to reach commercial production and was employed in early prototypes for vehicular methane storage.59 The aluminium FDCA MOF (MIL-160(Al)) is isostructural to an isophthalate based analogue (CAU-10) and exhibited potential for heat transformation applications due to its excellent water sorption performance.183 As an incidental advantage, bio-based linkers tend to show appreciable solubility in water and alcoholic solvents compared to similar-sized rigid phenyl-based ligands, which benefits their scale-up by solution-based processing. Thus, straightforward substitution is feasible when the linkers possess sufficient shape similarity. The design opportunities for bio-based building blocks are restricted by routes for functionalization and extension. While aromatic phenyls are readily functionalized by electrophilic substitution, the 1,4-diacids and FDCA do not have many available sites for further modification. As for extension, the elongated building blocks may result in awkward angles or too much flexibility for the reliable formation of frameworks. The stability of bio-based ligands is another potential concern. Recently, the in-silico design of MOF-74 analogue based on 2,3-dihydroxyfumarate indicated promising carbon capture performance, but the experimental attempts to obtain this isoreticular structure were not successful due to linker oxidation, even under mild conditions.186 The applicability of building block replacement strategies to other APMs is uncertain. For COFs, the flexibility and unconventional bond angle may not be well tolerated, although not impossible.187 Whereas for PMMs, the flexibility of ligands may be beneficial to achieve energetically favourable conformations. Besides the simple linkers derived from C6 sugars, more complex biologically-derived precursors may be used for APM synthesis. Serendipitously, γ-CD was found to coordinate to various alkali metals (K+, Rb+, and Cs+) and form infinitely extended

27 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 92

structures.188 Reticular chemistry, such as linker functionalization and mixed-metal strategies, have been used to improve the performance of the moderately porous CD-MOF materials (BET surface area >1000 m2 g-1) in CO2 capture applications.189-191 The γ-CD spatial orientation within the network may be exploited for various industrially demanded separations, including BTEX (benzene, toluene, ethylbenzene and xylene) isolation and chiral resolution.192,

193

By a more rational design, crystalline γ-CD COFs were synthesized by

reaction with trimethyl borate to form an anionic 3D structure. However, the compounds have poor stability in water.194 The CD examples suggest that APMs may be obtained from highly complex biologically-derived precursors, fully leveraging on the intricate synthetic machinery found in nature. However, currently established methods remain inadequate to understand the dynamic host-guest behaviour of these molecules, which complicates design efforts in this direction. Solvent management The solvent plays an integral role in APM synthesis, often influencing the structural and functional outcomes of the product. The polar aprotic solvents, including dimethylformamide (DMF), dimethylacetamide (DMAc), 1,4-dioxane, acetonitrile and dichloromethane, have been routinely applied in laboratory preparations of APMs. The ability to solubilize diverse range of precursors is emphasized in exploratory syntheses, so as to ensure reaction reversibility and controlled product formation (particularly for crystalline APMs). However, the examples listed above have recognized safety or health issues, and face regulatory control in numerous jurisdictions. As a result, there is a pressing need to develop processes using identified greener alternatives.195 Besides the safety aspect, there is also an economic incentive to manage solvent use effectively. DeSantis et al. performed a technoeconomic analysis on large scale production of four MOF sorbents, identifying solvent costs as the key cost driver.43 Such a trend is likely to be common to APM manufacturing. 28 ACS Paragon Plus Environment

Page 29 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Encouragingly, significant advances have been made in improving the solvent efficiency, which can be classified into two categories: (1) substitution by benign solvents, (2) development of solvent-free methods. Substitution of solvents Substitution approach constitutes the first level of process adjustment which avoids radical impact on the reaction mechanism. The general strategy is to adopt reaction media with chemical characteristics so as to maintain the solvation power and environment whilst reducing the level of health, environmental or flammability hazards. Water is regarded as the “greenest” option, being free of health or flammability issues. Alcoholic and ester solvents are also of little detriment to human health. However, the solubility of aromatic linkers in these solvents is severely compromised. The progress of scale-up development in MOFs has enabled explicit consideration of solvent replacements, while studies remain rare for the other APMs. Due to the dependence of solubility on temperature, the operation of processes near critical temperature may permit reactions, even in water. This approach is employed to a great success in the discovery of the MIL-series MOFs. Of these, MIL-100(Fe) and MIL101(Cr) stand out as highly porous materials with BET surface areas of >2000 m2 g-1.60, 196, 197 In addition, a wide variety of UiO-66-type MOFs have been derived from aqueous reflux conditions, even when using linkers with sparing solubility.198, 199 The inclusion of substantial amounts of monocarboxylates in the reaction system ensures the reproducible formation of hexanuclear zirconium clusters in aqueous solution.200 Interestingly, even though the reaction mixture is heterogeneous, strategies such as mixed-linker approach remain valid.201 Aqueous reaction protocols are possible even for COFs. Banerjee’s group prepared a series of COFs based on 1,3,5-triformylphloroglucinol using a water/acetic acid mixture, albeit with moderate surface areas.202 Tautomerism to form keto-enamines was observed to impart the requisite water stability to overcome the hydrolysis back-reaction. 29 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 92

It should be noted that the synthesis of numerous commercially relevant MOFs incorporating divalent cations, such as ZIF-8, HKUST-1 and SIFSIX series, is feasible in alcoholic solvents. For MOFs constructed using higher valence metal cations, the use of organic solvent substitutes is less reported. Among the more successful examples is the synthesis of Fe-soc-MOF (MIL-127) in 2-propanol.203 Solvothermal protocols for UiO-66 have also been developed with acetone, where the reduced surface tension of the replacement solvent permits direct activation of the framework.204 The dearth of examples may be due to undesired interactions with reaction components and the potential replacements. For instance, Katz’s group reported the issue of aldol condensation during the attempted synthesis of UiO66 in Cyrene (dihydrolevoglucosenone), which led to reduced surface areas.205 In general, the compromised solubility imposes a soft constraint over the size of the building blocks, and indirectly, the attainable porosity of the materials. It is foreseeable that further synthetic breakthroughs are necessary before widespread solvent substitutions become viable. Solvent-free chemistry Considering the exorbitant costs, complete elimination of solvent is an attractive prospect for APM manufacturing. The development of solvent-free approaches requires a means for the activation and contacting of the reagents. We highlight in this section three major approaches to solvent-free chemistry, namely (1) neat chemistry (involving the melt phase or vapour phase contacting of reagents), (2) vapour assisted conversion, and (3) mechanochemistry. Neat chemistry is typically applied for organic transformations in which at least one reagent exists as fluid under the reaction conditions. As ionic bonds in metal salts are difficult to break at low temperatures, the applicability of neat chemistry to APM synthesis is limited by the phase transition of the reactant linkers. For imidazolate and azolate based frameworks, synthesis in melt-phase has been reported at relatively low temperatures.147, 148 On the other hand, melt-phase synthesis is not viable for carboxylate30 ACS Paragon Plus Environment

Page 31 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

based ligands because of their high melting points (>300 °C) resulting from strong intermolecular hydrogen-bonding. Such temperatures are well in excess of that required for solution-phase synthesis and may even lead to decomposition of the other reaction components. The second approach is the dry-gel conversion (DGC) or vapour assisted conversion (VAC) method. By carefully tuning the composition of precursor and vapour solutions, it is possible to prepare well-ordered films of COFs and MOFs.206-209 Vapour-assisted methods are also employed to obtain bulk materials, including ZIFs,210,

211

UiO-66,212,

213

MIL-

100(Fe)214 and MIL-101.215 In addition, an example of DGC-derived imine-linked COFs was reported with somewhat reduced materials quality.216 Typically, the solid reactants are supported above the solvent reservoir in a closed system. The physical separation of solid and liquid reagents suggests the possibility of solvent recycle. However, this remains subject to solvent stability. Intriguingly, DGC/VAC is notable for permitting solvent substitutions that appear impracticable under typical solution-phase processing protocols. For example, COF-1 and COF-5 have been prepared in humid environments,206 whereas UiO-66 has been crystallized in the presence of alcohol vapours.213 A relatively wide temperature range has been explored for the existing DGC/VAC studies, ranging from 45 °C to 240 °C. DGC/VAC approach was originally developed for zeolites, and it has been shown that the dramatic reduction in solvent volume may allow attainment of higher temperatures without generating excessive autogenous pressures, which is preferable for industrial equipment.217 Exploration of temperature effects in further optimization studies may address the typically low spacetime efficiencies implied in the small scale experiments conducted for APM synthesis. The third approach – mechanochemistry – has emerged as a highly efficient alternative synthetic route for various chemical transformations.218 Mechanochemical synthesis may provide numerous advantages for scale-up. Firstly, space-time-efficiency is dramatically 31 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 92

increased due to the reduction in reactor volume required by the solvent. Secondly, the product is obtained directly in solids form which simplifies the separation. The use of extrusion as a continuous analogue to mechanical grinding has been shown to be highly effective, which will be discussed in a later section. The James group demonstrated the formation of porous frameworks by neat grinding of copper acetate and isonicotinic acid.219 By introducing solvent or ionic salt additives into the process, termed liquid-assisted grinding (LAG) or ion- and liquid-assisted grinding (ILAG), Friscic and coworkers achieved significant acceleration in the MOF synthesis, which was attributed to a templating effect. 220 A parameter η reflecting the additive volume to reactant mass ratio was defined to demarcate the boundary between LAG and slurry reactions, which were controlled by reactant solubility.221 To systematically understand the mechanism behind the self-assembly, a suite of techniques has been developed by Friscic and collaborators for the in-situ monitoring of mechanochemical

reactions.

Synchrotron

X-ray

powder

diffraction

during

the

mechanosynthesis of Zn-MeIm and Zn-EtIm (HMeIM = 2-methyl imidazole and HEtIM = 2-ethyl imidazole) revealed the existence of numerous topologically distinct polymorphs which progressively transformed into increasingly denser phases (Figure 6).222-224 Subsequently, control in the Zn-Im system favouring the highly porous RHO phase was achieved by adopting a non-ionic macrocyclic template in a milling-activated synthesis.225 Similar in-situ studies of industrially relevant MOF systems, including MOF-74226 and UiO66/NU-901227, have facilitated the development of optimized synthetic procedures. Beyond the obvious utility of X-ray based procedures in phase identification, Raman spectra228 and thermal profiles229 have been employed as auxiliary characterization tools to gain further mechanistic insight.

32 ACS Paragon Plus Environment

Page 33 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 6. In-situ monitoring of mechanochemical synthesis. (a) Milling jars of various materials for mechanosynthesis. (b) Adapted ball mill for X-ray monitoring. (c) Simulated XRD patterns and phase transition scheme for Zn HEtIM. (d-e) Time-resolved diffractograms and characteristic reflection monitoring in ILAG reactions. Images a and b: Reprinted with permission from Nature Protocols.230 Copyright 2013, Springer Nature. Images c to e: Reprinted with permission from Nature Chemistry.222 Copyright 2012, Springer Nature. Recognizing the compatibility of mechanosynthesis with Schiff-base chemistry, Banerjee’s team developed a solvent-free protocol for ketoenamine COFs.231 While moderate success was obtained using LAG,232 grinding of solid reactants in the presence of ptoluenesulfonic acid (PTSA) as a Bronsted acidic template followed by heating in a programmable oven produces the lamellar COFs with impressive BET surface areas (up to 3000 m2 g-1).233 The template synthesis may be used to fabricate freestanding COF membranes by a scalable knife-casting procedure with stable performance for 33 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 92

nanofiltration.234 To propose a guideline to direct the choice of acidic template, Karak et al. performed a comprehensive study which correlates the hydrogen-bonding distance within various acid-diamine salts (i.e. dav(Namine--H···Oacid)) with the crystallinity and porosity of the product materials.235 Clearly, varying the length and functional substituents of the monomers demands the use of different acidic templates to optimize the reversibility of the solvent-free reaction, which is a fine example of the synthetic complexities involved in the development of more sustainable manufacturing procedures. The scope of mechanochemistry is readily extended to porous organic polymers. Ball-milling synthesis of PIM-1 and PIM-4 is capable of achieving products with increased average molecular weight compared to solution-phase synthesis.236 The derived PIM materials may be redissolved for further processing, which could facilitate the elimination of harmful solvents. Dai and coworkers recently reported a mechanochemical route to prepare nanoporous polycarbazoles.237 The elimination of harmful chlorinated solvents serves as a further sustainability improvement to the relatively mild oxidative coupling synthesis.238-240 Furthermore, by tuning the quantity of FeCl3 catalyst and milling duration, it is possible to obtain soluble products.241 Consequently, mixed matrix membranes with high filler loadings (up to 60 wt. %) and good CO2 capture performance may be prepared from the mechanochemically-derived materials. The carbazole based monomers have also been used in solvent-free Friedel-Crafts alkylation with cyanuric chloride to form CTFs.242 Factorial methods of optimization have also been used to obtain high quality thiophene CMPs at quantitative conversion.243 As an alternative synthetic approach, mechanochemistry allows the circumvention of reactivity bottlenecks associated with the solution-phase. However, the phase complexity necessitates substantial optimization and mechanistic understanding. Strategies for process intensification

34 ACS Paragon Plus Environment

Page 35 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Optimization of process efficiency is necessary to ensure the industrial viability of the process. MOFs, which contributed a few early examples of APMs to reach commercialization, have originally relied on strategic modification of lab solvothermal protocols to rapidly attain scale-up capacities using batch reactors. Subsequently, due to the inspiration from pharmaceutical manufacturing and production of other value-added products, alternative synthetic routes – such as tubular flow reactors – have been developed. Rubio-Martinez et al. provided a recent review of the routes to scale up MOF production.55 We categorize the process intensification strategies into 3 categories: (1) alternative energy sources, (2) continuous synthesis, and (3) telescoping of manufacturing processes. The space-time-yield (STY) parameter, which correlates the amount of product to the volume of the reaction vessel and time required for synthesis, has been defined to facilitate benchmarking of synthetic performance. For a lab-based batch synthesis of MOFs, the STY is approximately 5 kg m-3 d1

, which may be increased to 150 kg m-3 d-1 by improved optimization. Orders of magnitude

improvement in efficiency (i.e. extrusion, STY ~ 104 –105 kg m-3 d-1,244 continuous microwave processing, STY ~ 103 – 106 kg m-3 d-1)245 may be realized by adopting intensified production methods. Although process intensification emphasizes mainly on the production efficiency, the materials quality and actual production rate remain important factors to be considered. Alternative energy sources Conventional heating as an energy source is ubiquitous and typically economical, but the thermal lag is inherent with increasing vessel sizes. Importantly, the effect is nonlinear and impairs fine control of the reaction. A homogeneous local environment devoid of sharp concentration or thermal gradients is desirable for crystallization. The electrochemical and microwave routes are distinct strategies for the introduction of energy into the process. For electrosynthesis, two possible routes – anodic dissolution and cathodic deposition – have 35 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 92

been developed for differing purposes. Anodic dissolution was developed as an anion-free strategy to prepare MOFs.246 In this process, the reactant metal is used as the electrode and released into the solution as ions by applying a suitable potential. The method is highly general and has been demonstrated for many archetypical MOFs and also high valence UiO66.247-249 As a scale-up approach, the electrochemical process realizes energy savings relative to thermally heated processes, and may be converted to continuous flow operation for uninterrupted processing. However, the solution conductivity and reaction at the counter electrode serve as complications to the process.250 Furthermore, reaction duration remains similar to batch phase solution synthesis, implying that the STY is inferior to other popular processing methods. On the other hand, the cathodic deposition route may still be important to APM production due to its utility in film formation, which is useful for device fabrication.251 For MOFs, Dinca’s group has designed the electrosynthesis protocol to exploit the pH drift from the cathodic reduction reaction as the self-assembly driving force.252 Gu et al. showed that the conductive characteristic of conjugated microporous polymers is conducive for their electrochemical deposition into thin-films with controllable thickness.131, 133, 253

The carbazole or thiophene monomers possess low oxidation potential for the coupling

reaction. Scherf group introduced boron trifluoride which further optimized the porosity.254, 255

While electrosynthesis may be used for organic transformations, the use of more resistive

organic solvents generally necessitates electrolyte addition. Microwave energy is another alternative energy source that is credited for remarkable rate accelerations for batch synthetic procedures. The delivery of energy is effected through the interaction of the reaction mixture with an alternating electromagnetic field, which requires the chemical species to possess a polar moment.256 The microwave-developed protocols are able to tolerate relatively similar reaction conditions whilst producing highly porous products, such as MOF-5,257 MIL-101(Cr),196, 258 and COF-5.259 The implication is 36 ACS Paragon Plus Environment

Page 37 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

that orthogonal synthetic improvements, such as building block protection, will be retained even if the energy source is changed. This was verified during the case study of BND-TFP COF prepared by transimination.99 If the penetration depth is sufficient, microwave radiation can produce volumetric heating (as opposed to conductive heating starting from the walls of the reaction vessel). Typically, materials encounter reduction in crystal sizes and narrower size distributions under microwave synthesis. The nanosized materials are difficult to separate, but may be beneficial when their usage requires high dispersion, such as in cellular applications.185, 260, 261 Similar to mechanochemistry, the rapid reactions can be exploited for the isolation of kinetic phases78,

262

or the preparation of materials with core-shell

structures.263 Metal-incorporating APMs possess compatible dielectric constants to exploit the use of microwave energy, and examples of pronounced rate accelerations have been reported. For example, Laybourn et al. used a purpose-built single mode microwave to reduce MIL-53(Al) synthetic duration to the order of seconds.264 The obtained yield data suggest that the product yield is correlated with the average absorbed power, keeping the total delivered energy constant. For pure organic materials, additives may have to be introduced to facilitate the energy absorption. Continuous synthesis The conversion of batch operations into continuous ones may improve the scalability of a process. Since reactants and products may be continuously fed or removed from the reactor, substantial time savings may be realized by avoiding the need to constantly charge and remove materials. A further improvement in the process efficiency results from high surface area to volume ratio (SA/V), which facilitates the mass and heat transfer. The synthetic duration of the batch process is mapped to the residence time parameter, τ, under continuous flow, which is a function of the vessel size and flow rate (i.e.

). It is hence

37 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 38 of 92

feasible to control the effective synthesis duration within a fine range, providing access to novel process windows which are rarely considered.265,

266

For APMs, extensive

modifications are required to make the transition from batch to continuous processing. The key factors to be considered are (1) the reagent or product insolubility, (2) the required process conditions (viz. temperature and pressure) for the transformation, and (3) the sensitivity of the formation process to mixing characteristics. The risks of flow disruption by insoluble particles are the highest for pumping equipment (at the feed inlet) and pressure regulation valves installed at the reactor exit. The ongoing development to address solidsrelated issues has led to various engineering solutions implemented at the lab-scale. For example, a ballast-inspired design has been proposed to prevent the direct contact of the pressure-regulating valve with solid materials.267, 268 However, the size of the ballast must occupy the major portion of the system volume to reduce pressure fluctuations, which raises questions over the scalability of this solution. More importantly, such designs are not yet commercially available, which hinders the scale-up development. Processes involving feed slurries are confined to operating zones below the boiling temperature, which is a major barrier in the scale-up of heterogeneous hydrothermal reactions.269, 270 To circumvent this, the organic building blocks are introduced as salts to enhance the aqueous solubility,63 or are dissolved in polar aprotic solvents, as is the case for numerous examples.271-274 Next, reactor configuration (Figure 7) and process conditions have strong influence over the mixing characteristics. The mixing is described by the dimensionless Reynolds number (

). The bulk fluid density

and viscosity

mixture, whereas for tubular reactors, the pipe diameter

are inherent for the reaction and flow velocity

may be

controlled by the system designer. A few studies of continuous flow synthesis of APMs have focused on microfluidic systems (Dt ~ 1.6 mm) operating in the laminar flow region. For this category, “droplet flow” approach (Figure 7c) – wherein immiscible solutions are separated 38 ACS Paragon Plus Environment

Page 39 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

into a continuous and dispersed phase after mixing – is widely adopted. Interfacial reactions at the droplet wall provide possibilities for shaping and guest encapsulation.273, 275 From the operations perspective, crystallization away from the reactor walls reduces the threat of fouling.271, 276 Not only can the defect concentration and nanosheet thickness be controlled, sequential synthesis can readily afford core-shell composite structures.271,

277, 278

The

extension to other APMs is also possible. The facile crystallization of two COFs – LZU-1 and RTCOF-1 – at room temperature aids their swift preparation (τ = 11 s) using microfluidic devices,279, 280 while Kim’s group presented a modified droplet protocol for the preparation of ketoenamine COFs.274 In the latter case, the solvent has to be adjusted from 1,4-dioxane to dimethylacetamide/water mixture to avoid mutual solubility with the carrier fluid. The efficiency and precision of microfluidic reactors is limited nevertheless by the low reactor volumes attainable. Mesoscale reactors (Dt ~ 3 to 10 mm) preserve the mixing characteristics while enabling production quantities approaching pilot plant levels. The counter-current reactor (Figure 7a) and plug flow reactors have been used to synthesize many archetypical MOFs.269,

272, 281, 282

The reported STYs range from 103 to 104 kg m-3 d-1, which is

approximately an order of magnitude improvement over batch processes. Another significant advantage of continuous processing is its orthogonal nature to other technologies, such as the practice of solvent-free chemistry or the use of alternative energy sources. A continuous analogue to mechanosynthesis was realized by James’ group using extrusion equipment, which enabled the synthesis of numerous MOFs at kg h-1 levels.244 The long history of extrusion and widespread application in food, metal and polymer processing (often at tonne scale) ensured the ready availability of industrial scale equipment, which is based on the screw conveyance of solid materials. As a solvent-free process, benchmark space-time-yields were reported for aluminium fumarate, HKUST-1, and ZIF-8 in the order of 104 to 105 kg m-3 d-1. The integration of microwave power with continuous flow has led to 39 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 40 of 92

the highest space-time efficiencies for solution-phase synthesis reported in the literature.245, 283

Figure 7. Process schematics of various flow reactor geometries: (a) countercurrent-flow (hydro)solvothermal synthesis (CFHS/CFSS); (b) continuous-stirred tank reactors (CSTR) and (c) micro- or mesofluidic systems. (d) (left) Schematic of coupled CF/spray-drying process. (right) Microscope images of dense MOF product. Images a-c: Reproduced with permission.284 Copyright 2016, Royal Society of Chemistry. Image d: Reproduced with permission.270 Copyright 2018, Royal Society of Chemistry. 40 ACS Paragon Plus Environment

Page 41 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Banerjee’s group showed a smaller-scale extrusion of TpPa-1 under PTSA templating followed by a second thermal activation step to obtain the COF (Figure 8).233 The possibility to control the temperature during the extrusion suggests that with more sophisticated equipment, it may be possible to obtain the product in a single stage process.

Figure 8. Extension of COF mechanosynthetic protocol to continuous processing and shaping. (a) Schematic of COF synthesis by extrusion. (b) Schematic of COF bead fabrication process. (c) Schematic of COF molding process. (d-f) Digital photographs of membrane, hollow tube and cylindrical morphologies of shaped COF structures. Reproduced with permission.233 Copyright 2018, American Chemical Society. A further option of continuous processing is the materials transfer to downstream processes. By eliminating work-up between each step, it may be possible to reduce 41 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 42 of 92

processing time and resources. This concept is known as “telescoping”. Telescoping may allow the integration of post-synthetic APM treatment with the main synthesis stage, directly affording dried and shaped materials ready for application. A case-study has been shown for MOFs which involves direct integration of spray drying with the solution-phase synthetic process.285-287 The co-atomization of precursor containing solutions and evaporation in a heated gas stream led to hollow superstructures made up of numerous MOF nanocrystals. Spray-drying may be used to encapsulate a diverse range of materials, such as hygroscopic salts for adsorptive heat transformation.288 The method is readily coupled to continuous flow processing, which overcomes the residence time limitation in full-fledged “aerosol synthesis” (Figure 7d).270,

289

A further difference is the formation of dense beads without hollow

interiors. A considerable residence time reduction is realized upon the coupling, which may be due to the rapid evaporative concentration of the precursors during the drying of the atomized microdroplets. The spray drying configuration was further shown to be compatible with Schiff-base chemistry and illustrated for the post-synthetic functionalization of MOFs.290 At the research level, effectiveness and generality of post-synthetic processes are prioritized whereas the operational efficiency is frequently neglected. In particular, solvent exchange and activation duration in conventional protocols have been observed by Matzger to be well in excess of the time actually required, suggesting much room for optimization.291 Film fabrication Rational design approaches towards tailoring the electrical transport and lightharvesting properties of APMs have expanded their applicability in several technologically important areas, including photovoltaic systems and electronic devices.292, 293 The deposition of crystalline thin films is attractive for attaining optimal performance in these applications.294 While the nanoscale thicknesses of the films reduce the emphasis on synthesizing bulk quantities of material, opportunities for process intensification have been 42 ACS Paragon Plus Environment

Page 43 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

identified to expand the industrial relevance of APM-based devices. Thin film deposition techniques and the challenges associated with device fabrication have been reviewed recently for MOF-based systems.294, 295 Notably, the liquid-phase epitaxial (LPE) growth of pillaredlayered MOFs based on paddlewheel SBUs has enabled precise control over film thickness and even the crystal orientation, which is challenging for strategies involving pre-synthesized particles.296 The scalability of film fabrication has been enhanced by process automation and high-throughput equipment, such as in the recent fabrication of highly-efficient porphyrinMOF based photovoltaic devices. Solution-phase processing methods provide further scope for post-synthetic transformations.297 Goswami et al. performed a controlled substitution of pillaring ligands to achieve a dimensionality reduction in a pillared-paddlewheel structure, extending the range of exciton propagation.298 2D COFs with extended conjugation and ordered electron donor/acceptor configurations are also shown to be ideal candidates for device fabrication.88,

299, 300

These examples serve to show the applicability of process

intensification techniques even at device fabrication level, which is promising for the advent of APMs in sophisticated technological applications.

43 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 44 of 92

OUTLOOK AND SUMMARY APMs have demonstrated their potential as disruptive technologies for several applications involving host-guest chemistry. Given the increasing number of APMs nearing pilot or commercial production, process sustainability is becoming a critical issue that warrants consideration even at early-stage development. We highlight the following challenges and directions that may deserve research attention in the near future. (1) The existing literature on APMs has focused on design principles, with less attention paid to understanding the synthetic mechanisms. Significant progress has been achieved for the in-situ observation of MOF synthetic mechanisms, which are recently extended to other APMs like COFs or POCs.301 For example, time-resolved X-ray diffraction,222, 302 mass spectroscopy303, 304 and NMR techniques may be employed to understand processrelevant conditions ranging from self-assembly and linker exchange to post-synthetic modification and activation.291, 305, 306 A balance between minimizing the measurement interference and maintaining the fidelity to optimal process conditions is required. (2) The utilization of catalysts in APM synthesis remains a nascent field which we feel requires deeper consideration. Several inspiring examples have shown that the adoption of superior catalysts may provide solutions to poor materials quality, extended synthetic durations and harsh reaction conditions.104,

112, 233

These advantages may outweigh

additional complexities involved due to introducing new components into the reaction mixture. Notably, the imine-233 and carbazole- based linkages242 have benefited significantly from catalyst integration into the process development. Considering the integral role of the catalyst in the process operation, the lifecycle costs and other safety or sustainability concerns should be carefully considered in the process evaluation of catalyst-assisted syntheses. 44 ACS Paragon Plus Environment

Page 45 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(3) The third major development for scalable and sustainable synthesis is solvent-free chemistry, which has proved to be applicable to organic and even hybrid materials. Even with the rapidly expanding portfolio of mechanosynthesis-derived APMs, the dissimilarity with solution-phase protocols demands systematic approaches towards understanding the processes, as has been amply demonstrated in MOFs.222, 224 (4) Finally, the synthetic development of APMs will benefit greatly by leveraging on established manufacturing equipment. The transition to continuous processes such as extrusion or spray drying can lead to orders-of-magnitude improvement in production efficiency even over optimized lab-scale processes.244 Further increase in production scales could justify dedicated process modelling studies on their operation. In summary, the practice of reticular design of APMs at commercially relevant scales is an ongoing challenge. The consideration of sustainability issues at the materials development stage is important to facilitate the transition to industrial production. A strong mechanistic understanding of the synthesis, the adoption of novel synthetic strategies such as solvent-free approaches and catalysis, as well as the utilization of robust manufacturing technologies, can promote more widespread implementation of APMs in the near future.

45 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 46 of 92

AUTHOR INFORMATION Corresponding Author * E-mail: [email protected] Notes The authors declare no competing financial interest. ACKNOWLEDGMENT This work is supported by National University of Singapore (CENGas R-261-508-001-646), Ministry of Education - Singapore (MOE AcRF Tier 1 R-279-000-472-112, R-279-000-540114), and Agency for Science, Technology and Research (PSF 1521200078, IRG A1783c0015, IAF-PP A1789a0024).

46 ACS Paragon Plus Environment

Page 47 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

REFERENCESs 1.

Hu, Z.; Srinivasan, M. P., Preparation of high-surface-area activated carbons from

coconut shell. Microporous Mesoporous Mater. 1999, 27 (1), 11-18, 10.1016/S13871811(98)00183-8. 2.

Hu, Z.; Srinivasan, M. P., Mesoporous high-surface-area activated carbon.

Microporous Mesoporous Mater. 2001, 43 (3), 267-275, 10.1016/S1387-1811(00)00355-3. 3.

Vermeiren, W.; Gilson, J. P., Impact of Zeolites on the Petroleum and Petrochemical

Industry. Top. Catal. 2009, 52 (9), 1131-1161, 10.1007/s11244-009-9271-8. 4.

Vogt, E. T. C.; Weckhuysen, B. M., Fluid catalytic cracking: recent developments on

the grand old lady of zeolite catalysis. Chem. Soc. Rev. 2015, 44 (20), 7342-7370, 10.1039/C5CS00376H. 5.

Tang, F.;

Biocompatibility

Li, L.; Chen, D., Mesoporous Silica Nanoparticles: Synthesis, and

Drug

Delivery.

Adv.

Mater.

2012,

24

(12),

1504-1534,

10.1002/adma.201104763. 6.

Sayari, A., Catalysis by Crystalline Mesoporous Molecular Sieves. Chem. Mater.

1996, 8 (8), 1840-1852, 10.1021/cm950585+. 7.

Blake, A. J.; Champness, N. R.; Hubberstey, P.; Li, W.-S.; Withersby, M. A.;

Schröder, M., Inorganic crystal engineering using self-assembly of tailored building-blocks. Coord. Chem. Rev. 1999, 183 (1), 117-138, 10.1016/S0010-8545(98)00173-8. 8.

Férey, G., Building Units Design and Scale Chemistry. J. Solid State Chem. 2000, 152

(1), 37-48, 10.1006/jssc.2000.8667. 9.

Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; Keeffe, M.; Yaghi, O. M.,

Systematic Design of Pore Size and Functionality in Isoreticular MOFs and Their Application in Methane Storage. Science 2002, 295 (5554), 469. 47 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

10.

Page 48 of 92

Kitagawa, S.; Kitaura, R.; Noro, S.-i., Functional Porous Coordination Polymers.

Angew. Chem. Int. Ed. 2004, 43 (18), 2334-2375, 10.1002/anie.200300610. 11.

Yaghi, O. M.; O'Keeffe, M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J.,

Reticular synthesis and the design of new materials. Nature 2003, 423, 705, 10.1038/nature01650. 12.

Mendoza-Cortés, J. L.;

Tranchemontagne, D. J.;

O’Keeffe, M.; Yaghi, O. M.,

Secondary building units, nets and bonding in the chemistry of metal–organic frameworks. Chem. Soc. Rev. 2009, 38 (5), 1257-1283, 10.1039/B817735J. 13.

Côté, A. P.; Benin, A. I.; Ockwig, N. W.; Keeffe, M.; Matzger, A. J.; Yaghi, O. M.,

Porous, Crystalline, Covalent Organic Frameworks. Science 2005, 310 (5751), 1166. 14.

Tozawa, T.; Jones, J. T. A.; Swamy, S. I.; Jiang, S.; Adams, D. J.; Shakespeare, S.;

Clowes, R.; Bradshaw, D.; Hasell, T.; Chong, S. Y.; Tang, C.; Thompson, S.; Parker, J.; Trewin, A.; Bacsa, J.; Slawin, A. M. Z.; Steiner, A.; Cooper, A. I., Porous organic cages. Nat. Mater. 2009, 8, 973, 10.1038/nmat2545. 15.

Wu, D.; Xu, F.; Sun, B.; Fu, R.; He, H.; Matyjaszewski, K., Design and Preparation

of Porous Polymers. Chem. Rev. 2012, 112 (7), 3959-4015, 10.1021/cr200440z. 16. and

Das, S.; Heasman, P.; Ben, T.; Qiu, S., Porous Organic Materials: Strategic Design Structure–Function

Correlation.

Chem.

Rev.

2017,

117

(3),

1515-1563,

10.1021/acs.chemrev.6b00439. 17.

Cheng, Y.; Ying, Y.; Japip, S.; Jiang, S.-D.; Chung, T.-S.; Zhang, S.; Zhao, D.,

Advanced Porous Materials in Mixed Matrix Membranes. Adv. Mater. 2018, 0 (0), 1802401, 10.1002/adma.201802401. 18.

Wilmer, C. E.; Leaf, M.; Lee, C. Y.; Farha, O. K.; Hauser, B. G.; Hupp, J. T.;

Snurr, R. Q., Large-scale screening of hypothetical metal–organic frameworks. Nat Chem. 2011, 4, 83, 10.1038/nchem.1192.

48 ACS Paragon Plus Environment

Page 49 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

19.

Gonzalez, M. I.; Kapelewski, M. T.; Bloch, E. D.; Milner, P. J.; Reed, D. A.;

Hudson, M. R.; Mason, J. A.; Barin, G.; Brown, C. M.; Long, J. R., Separation of Xylene Isomers through Multiple Metal Site Interactions in Metal–Organic Frameworks. J. Am. Chem. Soc. 2018, 140 (9), 3412-3422, 10.1021/jacs.7b13825. 20.

Dong, J.; Tummanapelli, A. K.; Li, X.; Ying, S.; Hirao, H.; Zhao, D., Fluorescent

Porous Organic Frameworks Containing Molecular Rotors for Size-Selective Recognition. Chem. Mater. 2016, 28 (21), 7889-7897, 10.1021/acs.chemmater.6b03376. 21.

Zhu, C.;

Yuan, G.;

Chen, X.;

Yang, Z.; Cui, Y., Chiral Nanoporous Metal–

Metallosalen Frameworks for Hydrolytic Kinetic Resolution of Epoxides. J. Am. Chem. Soc. 2012, 134 (19), 8058-8061, 10.1021/ja302340b. 22.

Deng, H.; Doonan, C. J.; Furukawa, H.; Ferreira, R. B.; Towne, J.; Knobler, C. B.;

Wang, B.; Yaghi, O. M., Multiple Functional Groups of Varying Ratios in Metal-Organic Frameworks. Science 2010, 327 (5967), 846. 23.

Liu, L.; Konstas, K.; Hill, M. R.; Telfer, S. G., Programmed Pore Architectures in

Modular Quaternary Metal–Organic Frameworks. J. Am. Chem. Soc. 2013, 135 (47), 1773117734, 10.1021/ja4100244. 24.

Pang, J.; Yuan, S.; Qin, J.; Wu, M.; Lollar, C. T.; Li, J.; Huang, N.; Li, B.; Zhang,

P.; Zhou, H.-C., Enhancing Pore-Environment Complexity Using a Trapezoidal Linker: Toward Stepwise Assembly of Multivariate Quinary Metal–Organic Frameworks. J. Am. Chem. Soc. 2018, 10.1021/jacs.8b07411. 25.

Zhang, Y.-B.; Furukawa, H.; Ko, N.; Nie, W.; Park, H. J.; Okajima, S.; Cordova,

K. E.; Deng, H.; Kim, J.; Yaghi, O. M., Introduction of Functionality, Selection of Topology, and Enhancement of Gas Adsorption in Multivariate Metal–Organic Framework-177. J. Am. Chem. Soc. 2015, 137 (7), 2641-2650, 10.1021/ja512311a.

49 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

26.

Page 50 of 92

Farha, O. K.; Eryazici, I.; Jeong, N. C.; Hauser, B. G.; Wilmer, C. E.; Sarjeant, A.

A.; Snurr, R. Q.; Nguyen, S. T.; Yazaydın, A. Ö.; Hupp, J. T., Metal–Organic Framework Materials with Ultrahigh Surface Areas: Is the Sky the Limit? J. Am. Chem. Soc. 2012, 134 (36), 15016-15021, 10.1021/ja3055639. 27.

Wang, T. C.; Bury, W.; Gómez-Gualdrón, D. A.; Vermeulen, N. A.; Mondloch, J.

E.; Deria, P.; Zhang, K.; Moghadam, P. Z.; Sarjeant, A. A.; Snurr, R. Q.; Stoddart, J. F.; Hupp, J. T.; Farha, O. K., Ultrahigh Surface Area Zirconium MOFs and Insights into the Applicability of the BET Theory. J. Am. Chem. Soc. 2015, 137 (10), 3585-3591, 10.1021/ja512973b. 28.

Hönicke, I.; Senkovska, I.; Bon, V.; Baburin, I.; Boenisch, N.; Raschke, S.; Evans,

J. D.; Kaskel, S., Balancing Mechanical Stability and Ultrahigh Porosity in Crystalline Framework Materials. Angew. Chem. Int. Ed. 2018, 10.1002/anie.201808240. 29.

El-Kaderi, H. M.; Hunt, J. R.; Mendoza-Cortés, J. L.; Côté, A. P.; Taylor, R. E.;

O'Keeffe, M.; Yaghi, O. M., Designed Synthesis of 3D Covalent Organic Frameworks. Science 2007, 316 (5822), 268-272, 10.1126/science.1139915. 30.

Baldwin, L. A.; Crowe, J. W.; Pyles, D. A.; McGrier, P. L., Metalation of a

Mesoporous Three-Dimensional Covalent Organic Framework. J. Am. Chem. Soc. 2016, 138 (46), 15134-15137, 10.1021/jacs.6b10316. 31.

Ben, T.; Ren, H.; Ma, S.; Cao, D.; Lan, J.; Jing, X.; Wang, W.; Xu, J.; Deng, F.;

Simmons, J. M.; Qiu, S.; Zhu, G., Targeted Synthesis of a Porous Aromatic Framework with High Stability and Exceptionally High Surface Area. Angew. Chem. Int. Ed. 2009, 48 (50), 9457-9460, 10.1002/anie.200904637. 32.

Li, M.; Ren, H.; Sun, F.; Tian, Y.; Zhu, Y.; Li, J.; Mu, X.; Xu, J.; Deng, F.; Zhu,

G., Construction of Porous Aromatic Frameworks with Exceptional Porosity via Building Unit Engineering. Adv. Mater. 2018, 0 (0), 1804169, 10.1002/adma.201804169.

50 ACS Paragon Plus Environment

Page 51 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

33.

Yuan, D.; Lu, W.; Zhao, D.; Zhou, H.-C., Highly Stable Porous Polymer Networks

with Exceptionally High Gas-Uptake Capacities. Adv. Mater. 2011, 23 (32), 3723-3725, 10.1002/adma.201101759. 34.

Tian, T.; Zeng, Z.; Vulpe, D.; Casco, M. E.; Divitini, G.; Midgley, P. A.;

Silvestre-Albero, J.; Tan, J.-C.; Moghadam, P. Z.; Fairen-Jimenez, D., A sol–gel monolithic metal–organic framework with enhanced methane uptake. Nat. Mater. 2017, 17, 174, 10.1038/nmat5050. 35.

Cui, X.; Chen, K.; Xing, H.; Yang, Q.; Krishna, R.; Bao, Z.; Wu, H.; Zhou, W.;

Dong, X.; Han, Y.; Li, B.; Ren, Q.; Zaworotko, M. J.; Chen, B., Pore chemistry and size control in hybrid porous materials for acetylene capture from ethylene. Science 2016, 353 (6295), 141-144, 10.1126/science.aaf2458. 36.

Cadiau, A.; Adil, K.; Bhatt, P. M.; Belmabkhout, Y.; Eddaoudi, M., A metal-

organic framework–based splitter for separating propylene from propane. Science 2016, 353 (6295), 137-140, 10.1126/science.aaf6323. 37.

Rodenas, T.; Luz, I.; Prieto, G.; Seoane, B.; Miro, H.; Corma, A.; Kapteijn, F.;

Llabrés i Xamena, F. X.; Gascon, J., Metal–organic framework nanosheets in polymer composite materials for gas separation. Nat. Mater. 2014, 14, 48, 10.1038/nmat4113. 38.

Lin, S.; Diercks, C. S.; Zhang, Y.-B.; Kornienko, N.; Nichols, E. M.; Zhao, Y.;

Paris, A. R.; Kim, D.; Yang, P.; Yaghi, O. M.; Chang, C. J., Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water. Science 2015, 349 (6253), 1208-1213, 10.1126/science.aac8343. 39.

Xu, H.; Gao, J.; Jiang, D., Stable, crystalline, porous, covalent organic frameworks as

a platform for chiral organocatalysts. Nat Chem. 2015, 7, 905, 10.1038/nchem.2352. 40.

Deng, H.; Grunder, S.; Cordova, K. E.; Valente, C.; Furukawa, H.; Hmadeh, M.;

Gándara, F.; Whalley, A. C.; Liu, Z.; Asahina, S.; Kazumori, H.; O’Keeffe, M.; Terasaki,

51 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

O.;

Page 52 of 92

Stoddart, J. F.; Yaghi, O. M., Large-Pore Apertures in a Series of Metal-Organic

Frameworks. Science 2012, 336 (6084), 1018. 41.

Chen, W.-H.;

Vázquez-González, M.;

Zoabi, A.;

Abu-Reziq, R.; Willner, I.,

Biocatalytic cascades driven by enzymes encapsulated in metal–organic framework nanoparticles. Nat. Catal. 2018, 1 (9), 689-695, 10.1038/s41929-018-0117-2. 42.

Fortea-Pérez, F. R.; Mon, M.; Ferrando-Soria, J.; Boronat, M.; Leyva-Pérez, A.;

Corma, A.; Herrera, J. M.; Osadchii, D.; Gascon, J.; Armentano, D.; Pardo, E., The MOFdriven synthesis of supported palladium clusters with catalytic activity for carbene-mediated chemistry. Nat. Mater. 2017, 16, 760, 10.1038/nmat4910. 43.

DeSantis, D.; Mason, J. A.; James, B. D.; Houchins, C.; Long, J. R.; Veenstra, M.,

Techno-economic Analysis of Metal–Organic Frameworks for Hydrogen and Natural Gas Storage. Energy Fuels 2017, 31 (2), 2024-2032, 10.1021/acs.energyfuels.6b02510. 44.

Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M., The Chemistry and

Applications of Metal-Organic Frameworks. Science 2013, 341 (6149). 45.

Diercks, C. S.; Yaghi, O. M., The atom, the molecule, and the covalent organic

framework. Science 2017, 355 (6328). 46.

Hasell, T.; Cooper, A. I., Porous organic cages: soluble, modular and molecular pores.

Nat. Rev. Mat. 2016, 1, 16053, 10.1038/natrevmats.2016.53. 47.

Frohlich, D.;

Pantatosaki, E.;

Kolokathis, P. D.;

Markey, K.;

Reinsch, H.;

Baumgartner, M.; van der Veen, M. A.; De Vos, D. E.; Stock, N.; Papadopoulos, G. K.; Henninger, S. K.; Janiak, C., Water adsorption behaviour of CAU-10-H: a thorough investigation of its structure-property relationships. J. Mater. Chem. A 2016, 4 (30), 1185911869, 10.1039/C6TA01757F. 48.

Julien, P. A.; Mottillo, C.; Friščić, T., Metal–organic frameworks meet scalable and

sustainable synthesis. Green Chem. 2017, 19 (12), 2729-2747, 10.1039/C7GC01078H.

52 ACS Paragon Plus Environment

Page 53 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

49.

Li, P.; Chen, Q.; Wang, T. C.; Vermeulen, N. A.; Mehdi, B. L.; Dohnalkova, A.;

Browning, N. D.; Shen, D.; Anderson, R.; Gómez-Gualdrón, D. A.; Cetin, F. M.; Jagiello, J.; Asiri, A. M.; Stoddart, J. F.; Farha, O. K., Hierarchically Engineered Mesoporous MetalOrganic Frameworks toward Cell-free Immobilized Enzyme Systems. Chem 2018, 4 (5), 1022-1034, 10.1016/j.chempr.2018.03.001. 50.

Hoskins, B. F.; Robson, R., Design and construction of a new class of scaffolding-like

materials comprising infinite polymeric frameworks of 3D-linked molecular rods. A reappraisal of the zinc cyanide and cadmium cyanide structures and the synthesis and structure of the diamond-related frameworks [N(CH3)4][CuIZnII(CN)4] and CuI[4,4',4'',4'''tetracyanotetraphenylmethane]BF4.xC6H5NO2. J. Am. Chem. Soc. 1990, 112 (4), 1546-1554, 10.1021/ja00160a038. 51.

Li, H.; Eddaoudi, M.; O'Keeffe, M.; Yaghi, O. M., Design and synthesis of an

exceptionally stable and highly porous metal-organic framework. Nature 1999, 402, 276, 10.1038/46248. 52.

Eddaoudi, M.; Moler, D. B.; Li, H.; Chen, B.; Reineke, T. M.; O'Keeffe, M.; Yaghi,

O. M., Modular Chemistry:  Secondary Building Units as a Basis for the Design of Highly Porous and Robust Metal−Organic Carboxylate Frameworks. Acc. Chem. Res 2001, 34 (4), 319-330, 10.1021/ar000034b. 53.

Li, M.; Li, D.; O’Keeffe, M.; Yaghi, O. M., Topological Analysis of Metal–Organic

Frameworks with Polytopic Linkers and/or Multiple Building Units and the Minimal Transitivity Principle. Chem. Rev. 2014, 114 (2), 1343-1370, 10.1021/cr400392k. 54.

Stock, N.; Biswas, S., Synthesis of Metal-Organic Frameworks (MOFs): Routes to

Various MOF Topologies, Morphologies, and Composites. Chem. Rev. 2012, 112 (2), 933969, 10.1021/cr200304e.

53 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

55.

Page 54 of 92

Rubio-Martinez, M.; Avci-Camur, C.; Thornton, A. W.; Imaz, I.; Maspoch, D.; Hill,

M. R., New synthetic routes towards MOF production at scale. Chem. Soc. Rev. 2017, 46 (11), 3453-3480, 10.1039/C7CS00109F. 56.

Chui, S. S.-Y.; Lo, S. M.-F.; Charmant, J. P. H.; Orpen, A. G.; Williams, I. D., A

Chemically Functionalizable Nanoporous Material [Cu3(TMA)2(H2O)3]n. Science 1999, 283 (5405), 1148-1150, 10.1126/science.283.5405.1148. 57.

Park, K. S.; Ni, Z.; Côté, A. P.; Choi, J. Y.; Huang, R.; Uribe-Romo, F. J.; Chae,

H. K.; O’Keeffe, M.; Yaghi, O. M., Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. U.S.A. 2006, 103 (27), 10186. 58.

Nugent, P.; Belmabkhout, Y.; Burd, S. D.; Cairns, A. J.; Luebke, R.; Forrest, K.;

Pham, T.; Ma, S.; Space, B.; Wojtas, L.; Eddaoudi, M.; Zaworotko, M. J., Porous materials with optimal adsorption thermodynamics and kinetics for CO2 separation. Nature 2013, 495, 80, 10.1038/nature11893. 59.

Gaab, M.; Trukhan, N.; Maurer, S.; Gummaraju, R.; Müller, U., The progression of

Al-based metal-organic frameworks – From academic research to industrial production and applications.

Microporous

Mesoporous

Mater.

2012,

157,

131-136,

https://doi.org/10.1016/j.micromeso.2011.08.016. 60.

Seo, Y.-K.; Yoon, J. W.; Lee, J. S.; Lee, U. H.; Hwang, Y. K.; Jun, C.-H.;

Horcajada, P.; Serre, C.; Chang, J.-S., Large scale fluorine-free synthesis of hierarchically porous iron(III) trimesate MIL-100(Fe) with a zeolite MTN topology. Microporous Mesoporous Mater. 2012, 157, 137-145, https://doi.org/10.1016/j.micromeso.2012.02.027. 61.

Lenzen, D.; Bendix, P.; Reinsch, H.; Fröhlich, D.; Kummer, H.; Möllers, M.;

Hügenell, P. P. C.; Gläser, R.; Henninger, S.; Stock, N., Scalable Green Synthesis and FullScale Test of the Metal–Organic Framework CAU-10-H for Use in Adsorption-Driven Chillers. Adv. Mater. 2017, 30 (6), 1705869, 10.1002/adma.201705869.

54 ACS Paragon Plus Environment

Page 55 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

62.

Khabzina, Y.; Dhainaut, J.; Ahlhelm, M.; Richter, H.-J.; Reinsch, H.; Stock, N.;

Farrusseng, D., Synthesis and Shaping Scale-up Study of Functionalized UiO-66 MOF for Ammonia Air Purification Filters. Ind. Eng. Chem. Res. 2018, 57 (24), 8200-8208, 10.1021/acs.iecr.8b00808. 63.

Sánchez-Sánchez, M.; Getachew, N.; Díaz, K.; Díaz-García, M.; Chebude, Y.; Díaz,

I., Synthesis of metal–organic frameworks in water at room temperature: salts as linker sources. Green Chem. 2015, 17 (3), 1500-1509, 10.1039/C4GC01861C. 64.

Yuan, S.; Feng, L.; Wang, K.; Pang, J.; Bosch, M.; Lollar, C.; Sun, Y.; Qin, J.;

Yang, X.; Zhang, P.; Wang, Q.; Zou, L.; Zhang, Y.; Zhang, L.; Fang, Y.; Li, J.; Zhou, H. C., Stable Metal–Organic Frameworks: Design, Synthesis, and Applications. Adv. Mater. 2018, 10.1002/adma.201704303. 65.

Serre, C.; Millange, F.; Thouvenot, C.; Noguès, M.; Marsolier, G.; Louër, D.;

Férey, G., Very Large Breathing Effect in the First Nanoporous Chromium(III)-Based Solids: MIL-53 or CrIII(OH)·{O2C−C6H4−CO2}·{HO2C−C6H4−CO2H}x·H2Oy. J. Am. Chem. Soc. 2002, 124 (45), 13519-13526, 10.1021/ja0276974. 66.

Guillerm, V.; Ragon, F.; Dan-Hardi, M.; Devic, T.; Vishnuvarthan, M.; Campo, B.;

Vimont, A.; Clet, G.; Yang, Q.; Maurin, G.; Férey, G.; Vittadini, A.; Gross, S.; Serre, C., A Series of Isoreticular, Highly Stable, Porous Zirconium Oxide Based Metal–Organic Frameworks. Angew. Chem. Int. Ed. 2012, 51 (37), 9267-9271, 10.1002/anie.201204806. 67.

Wang, S.; Kitao, T.; Guillou, N.; Wahiduzzaman, M.; Martineau-Corcos, C.;

Nouar, F.; Tissot, A.; Binet, L.; Ramsahye, N.; Devautour-Vinot, S.; Kitagawa, S.; Seki, S.; Tsutsui, Y.; Briois, V.; Steunou, N.; Maurin, G.; Uemura, T.; Serre, C., A phase transformable ultrastable titanium-carboxylate framework for photoconduction. Nat. Commun. 2018, 9 (1), 1660, 10.1038/s41467-018-04034-w.

55 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

68.

Page 56 of 92

Férey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S.;

Margiolaki, I., A Chromium Terephthalate-Based Solid with Unusually Large Pore Volumes and Surface Area. Science 2005, 309 (5743), 2040. 69.

Dan-Hardi, M.; Serre, C.; Frot, T.; Rozes, L.; Maurin, G.; Sanchez, C.; Férey, G.,

A New Photoactive Crystalline Highly Porous Titanium(IV) Dicarboxylate. J. Am. Chem. Soc. 2009, 131 (31), 10857-10859, 10.1021/ja903726m. 70.

Feng, D.; Wang, K.; Wei, Z.; Chen, Y.-P.; Simon, C. M.; Arvapally, R. K.; Martin,

R. L.; Bosch, M.; Liu, T.-F.; Fordham, S.; Yuan, D.; Omary, M. A.; Haranczyk, M.; Smit, B.; Zhou, H.-C., Kinetically tuned dimensional augmentation as a versatile synthetic route towards

robust

metal–organic

frameworks.

Nat.

Commun.

2014,

5,

5723,

10.1038/ncomms6723. 71.

Yuan, S.; Liu, T.-F.; Feng, D.; Tian, J.; Wang, K.; Qin, J.; Zhang, Q.; Chen, Y.-P.;

Bosch, M.;

Zou, L.;

Teat, S. J.;

Dalgarno, S. J.; Zhou, H.-C., A single crystalline

porphyrinic titanium metal–organic framework. Chem. Sci. 2015, 6 (7), 3926-3930, 10.1039/C5SC00916B. 72.

Subramanian, S.; Zaworotko, M. J., Porous Solids by Design: [Zn(4,4 ′ -

bpy)2(SiF6)]n·xDMF, a Single Framework Octahedral Coordination Polymer with Large Square Channels. Angew. Chem. Int. Ed. 1995, 34 (19), 2127-2129, 10.1002/anie.199521271. 73.

Cadiau, A.; Belmabkhout, Y.; Adil, K.; Bhatt, P. M.; Pillai, R. S.; Shkurenko, A.;

Martineau-Corcos, C.; Maurin, G.; Eddaoudi, M., Hydrolytically stable fluorinated metalorganic frameworks for energy-efficient dehydration. Science 2017, 356 (6339), 731-735, 10.1126/science.aam8310. 74.

Xue, D.-X.; Belmabkhout, Y.; Shekhah, O.; Jiang, H.; Adil, K.; Cairns, A. J.;

Eddaoudi, M., Tunable Rare Earth fcu-MOF Platform: Access to Adsorption Kinetics Driven

56 ACS Paragon Plus Environment

Page 57 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Gas/Vapor Separations via Pore Size Contraction. J. Am. Chem. Soc. 2015, 137 (15), 50345040, 10.1021/ja5131403. 75.

Luo, T.-Y.; Liu, C.; Eliseeva, S. V.; Muldoon, P. F.; Petoud, S.; Rosi, N. L., Rare

Earth pcu Metal–Organic Framework Platform Based on RE4(μ3-OH)4(COO)62+ Clusters: Rational Design, Directed Synthesis, and Deliberate Tuning of Excitation Wavelengths. J. Am. Chem. Soc. 2017, 139 (27), 9333-9340, 10.1021/jacs.7b04532. 76.

Biswas, S.; Ahnfeldt, T.; Stock, N., New Functionalized Flexible Al-MIL-53-X (X =

-Cl, -Br, -CH3, -NO2, -(OH)2) Solids: Syntheses, Characterization, Sorption, and Breathing Behavior. Inorg. Chem. 2011, 50 (19), 9518-9526, 10.1021/ic201219g. 77.

Reinsch, H.; van der Veen, M. A.; Gil, B.; Marszalek, B.; Verbiest, T.; de Vos, D.;

Stock, N., Structures, Sorption Characteristics, and Nonlinear Optical Properties of a New Series of Highly Stable Aluminum MOFs. Chem. Mater. 2013, 25 (1), 17-26, 10.1021/cm3025445. 78.

Liang, W.; D'Alessandro, D. M., Microwave-assisted solvothermal synthesis of

zirconium oxide based metal–organic frameworks. Chem. Commun. 2013, 49 (35), 37063708, 10.1039/C3CC40368H. 79.

Liang, W.; Babarao, R.; D’Alessandro, D. M., Microwave-Assisted Solvothermal

Synthesis and Optical Properties of Tagged MIL-140A Metal–Organic Frameworks. Inorg. Chem. 2013, 52 (22), 12878-12880, 10.1021/ic4024234. 80.

Cortijo, M.; Herrero, S.; Jiménez-Aparicio, R.; Perles, J.; Priego, J. L.; Torralvo,

M. J.; Torroba, J., Hybrid Polyfunctional Systems Based on Nickel(II) Isonicotinate. Eur. J. Inorg. Chem. 2013, 2013 (14), 2580-2590, 10.1002/ejic.201201461. 81.

Nandi, S.; De Luna, P.; Daff, T. D.; Rother, J.; Liu, M.; Buchanan, W.; Hawari, A.

I.;

Woo, T. K.; Vaidhyanathan, R., A single-ligand ultra-microporous MOF for

57 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 58 of 92

precombustion CO2 capture and hydrogen purification. Sci Adv. 2015, 1 (11), 10.1126/sciadv.1500421. 82.

Nandi, S.; Collins, S.; Chakraborty, D.; Banerjee, D.; Thallapally, P. K.; Woo, T.

K.; Vaidhyanathan, R., Ultralow Parasitic Energy for Postcombustion CO2 Capture Realized in a Nickel Isonicotinate Metal–Organic Framework with Excellent Moisture Stability. J. Am. Chem. Soc. 2017, 139 (5), 1734-1737, 10.1021/jacs.6b10455. 83.

Nguyen, V.; Grünwald, M., Microscopic Origins of Poor Crystallinity in the

Synthesis of Covalent Organic Framework COF-5. J. Am. Chem. Soc. 2018, 140 (9), 33063311, 10.1021/jacs.7b12529. 84.

Lohse, M. S.; Bein, T., Covalent Organic Frameworks: Structures, Synthesis, and

Applications. Adv. Funct. Mater. 2018, 28 (33), 1705553, 10.1002/adfm.201705553. 85.

Ding, S.-Y.; Wang, W., Covalent organic frameworks (COFs): from design to

applications. Chem. Soc. Rev. 2013, 42 (2), 548-568, 10.1039/C2CS35072F. 86.

Vyas, V. S.; Haase, F.; Stegbauer, L.; Savasci, G.; Podjaski, F.; Ochsenfeld, C.;

Lotsch, B. V., A tunable azine covalent organic framework platform for visible light-induced hydrogen generation. Nat. Commun. 2015, 6, 8508, 10.1038/ncomms9508. 87.

Wei, P.-F.; Qi, M.-Z.; Wang, Z.-P.; Ding, S.-Y.; Yu, W.; Liu, Q.; Wang, L.-K.;

Wang, H.-Z.; An, W.-K.; Wang, W., Benzoxazole-Linked Ultrastable Covalent Organic Frameworks for Photocatalysis. J. Am. Chem. Soc. 2018, 140 (13), 4623-4631, 10.1021/jacs.8b00571. 88.

Wang, X.; Chen, L.; Chong, S. Y.; Little, M. A.; Wu, Y.; Zhu, W.-H.; Clowes, R.;

Yan, Y.; Zwijnenburg, M. A.; Sprick, R. S.; Cooper, A. I., Sulfone-containing covalent organic frameworks for photocatalytic hydrogen evolution from water. Nat Chem. 2018, 10.1038/s41557-018-0141-5.

58 ACS Paragon Plus Environment

Page 59 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

89.

Halder, A.; Karak, S.; Addicoat, M.; Bera, S.; Chakraborty, A.; Kunjattu, S. H.;

Pachfule, P.;

Heine, T.; Banerjee, R., Ultrastable Imine-Based Covalent Organic

Frameworks for Sulfuric Acid Recovery: An Effect of Interlayer Hydrogen Bonding. Angew. Chem. Int. Ed. 2018, 57 (20), 5797-5802, 10.1002/anie.201802220. 90.

Yang, Y.; Faheem, M.; Wang, L.; Meng, Q.; Sha, H.; Yang, N.; Yuan, Y.; Zhu, G.,

Surface Pore Engineering of Covalent Organic Frameworks for Ammonia Capture through Synergistic Multivariate and Open Metal Site Approaches. ACS Cent. Sci. 2018, 4 (6), 748754, 10.1021/acscentsci.8b00232. 91.

Smith, B. J.; Dichtel, W. R., Mechanistic Studies of Two-Dimensional Covalent

Organic Frameworks Rapidly Polymerized from Initially Homogenous Conditions. J. Am. Chem. Soc. 2014, 136 (24), 8783-8789, 10.1021/ja5037868. 92.

Smith, B. J.;

Overholts, A. C.;

Hwang, N.; Dichtel, W. R., Insight into the

crystallization of amorphous imine-linked polymer networks to 2D covalent organic frameworks. Chem. Commun. 2016, 52 (18), 3690-3693, 10.1039/C5CC10221A. 93.

Qian, C.; Qi, Q.-Y.; Jiang, G.-F.; Cui, F.-Z.; Tian, Y.; Zhao, X., Toward Covalent

Organic Frameworks Bearing Three Different Kinds of Pores: The Strategy for Construction and COF-to-COF Transformation via Heterogeneous Linker Exchange. J. Am. Chem. Soc. 2017, 139 (19), 6736-6743, 10.1021/jacs.7b02303. 94.

Evans, A. M.; Parent, L. R.; Flanders, N. C.; Bisbey, R. P.; Vitaku, E.; Kirschner,

M. S.; Schaller, R. D.; Chen, L. X.; Gianneschi, N. C.; Dichtel, W. R., Seeded growth of single-crystal two-dimensional covalent organic frameworks. Science 2018. 95.

Ma, T.; Kapustin, E. A.; Yin, S. X.; Liang, L.; Zhou, Z.; Niu, J.; Li, L.-H.; Wang,

Y.; Su, J.; Li, J.; Wang, X.; Wang, W. D.; Wang, W.; Sun, J.; Yaghi, O. M., Singlecrystal x-ray diffraction structures of covalent organic frameworks. Science 2018, 361 (6397), 48.

59 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

96.

Page 60 of 92

Ma, T.; Li, J.; Niu, J.; Zhang, L.; Etman, A. S.; Lin, C.; Shi, D.; Chen, P.; Li, L.-

H.; Du, X.; Sun, J.; Wang, W., Observation of Interpenetration Isomerism in Covalent Organic Frameworks. J. Am. Chem. Soc. 2018, 140 (22), 6763-6766, 10.1021/jacs.8b03169. 97.

Kandambeth, S.; Shinde, D. B.; Panda, M. K.; Lukose, B.; Heine, T.; Banerjee, R.,

Enhancement of Chemical Stability and Crystallinity in Porphyrin-Containing Covalent Organic Frameworks by Intramolecular Hydrogen Bonds. Angew. Chem. Int. Ed. 2013, 52 (49), 13052-13056, 10.1002/anie.201306775. 98.

Chen, X.; Addicoat, M.; Jin, E.; Zhai, L.; Xu, H.; Huang, N.; Guo, Z.; Liu, L.;

Irle, S.; Jiang, D., Locking Covalent Organic Frameworks with Hydrogen Bonds: General and Remarkable Effects on Crystalline Structure, Physical Properties, and Photochemical Activity. J. Am. Chem. Soc. 2015, 137 (9), 3241-3247, 10.1021/ja509602c. 99.

Vitaku, E.; Dichtel, W. R., Synthesis of 2D Imine-Linked Covalent Organic

Frameworks through Formal Transimination Reactions. J. Am. Chem. Soc. 2017, 139 (37), 12911-12914, 10.1021/jacs.7b06913. 100.

Zhao, W.; Qiao, J.; Ning, T.-L.; Liu, X.-K., Scalable ambient pressure synthesis of

covalent organic frameworks and their colorimetric nanocomposites through dynamic imine exchange reactions. Chin. J. Polym. Sci. 2018, 36 (1), 1-7, 10.1007/s10118-018-2010-5. 101.

Liu, M.; Huang, Q.; Wang, S.; Li, Z.; Li, B.; Jin, S.; Tan, B., Crystalline Covalent

Triazine Frameworks by In Situ Oxidation of Alcohols to Aldehyde Monomers. Angew. Chem. Int. Ed. 2018, 57 (37), 11968-11972, 10.1002/anie.201806664. 102.

Kuhn, P.; Antonietti, M.; Thomas, A., Porous, Covalent Triazine-Based Frameworks

Prepared by Ionothermal Synthesis. Angew. Chem. Int. Ed. 2008, 47 (18), 3450-3453, 10.1002/anie.200705710.

60 ACS Paragon Plus Environment

Page 61 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

103.

Giuseppone, N.; Schmitt, J.-L.; Schwartz, E.; Lehn, J.-M., Scandium(III) Catalysis

of Transimination Reactions. Independent and Constitutionally Coupled Reversible Processes. J. Am. Chem. Soc. 2005, 127 (15), 5528-5539, 10.1021/ja042469q. 104.

Matsumoto, M.; Dasari, R. R.; Ji, W.; Feriante, C. H.; Parker, T. C.; Marder, S. R.;

Dichtel, W. R., Rapid, Low Temperature Formation of Imine-Linked Covalent Organic Frameworks Catalyzed by Metal Triflates. J. Am. Chem. Soc. 2017, 139 (14), 4999-5002, 10.1021/jacs.7b01240. 105. F.;

Matsumoto, M.; Valentino, L.; Stiehl, G. M.; Balch, H. B.; Corcos, A. R.; Wang, Ralph, D. C.;

Mariñas, B. J.; Dichtel, W. R., Lewis-Acid-Catalyzed Interfacial

Polymerization of Covalent Organic Framework Films. Chem 2018, 4 (2), 308-317, 10.1016/j.chempr.2017.12.011. 106.

Valentino, L.; Matsumoto, M.; Dichtel, W. R.; Mariñas, B. J., Development and

Performance Characterization of a Polyimine Covalent Organic Framework Thin-Film Composite Nanofiltration Membrane. Environ. Sci. Technol. 2017, 51 (24), 14352-14359, 10.1021/acs.est.7b04056. 107.

Wang, R.; Shi, X.; Xiao, A.; Zhou, W.; Wang, Y., Interfacial Polymerization of

Covalent Organic Frameworks (COFs) on Polymeric Substrates for Molecular Separations. J. Membrane Sci. 2018, 10.1016/j.memsci.2018.08.044. 108.

Dirksen, A.;

Hackeng, T. M.; Dawson, P. E., Nucleophilic Catalysis of Oxime

Ligation. Angew. Chem. Int. Ed. 2006, 45 (45), 7581-7584, 10.1002/anie.200602877. 109.

Dirksen, A.; Dirksen, S.; Hackeng, T. M.; Dawson, P. E., Nucleophilic Catalysis of

Hydrazone Formation and Transimination:  Implications for Dynamic Covalent Chemistry. J. Am. Chem. Soc. 2006, 128 (49), 15602-15603, 10.1021/ja067189k. 110.

Miller, B. L., Catalysing dynamic libraries.

Nat Chem. 2010, 2, 433,

10.1038/nchem.659.

61 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

111.

Page 62 of 92

Briggs, M. E.; Slater, A. G.; Lunt, N.; Jiang, S.; Little, M. A.; Greenaway, R. L.;

Hasell, T.; Battilocchio, C.; Ley, S. V.; Cooper, A. I., Dynamic flow synthesis of porous organic cages. Chem. Commun. 2015, 51 (98), 17390-17393, 10.1039/C5CC07447A. 112.

Guan, X.; Ma, Y.; Li, H.; Yusran, Y.; Xue, M.; Fang, Q.; Yan, Y.; Valtchev, V.;

Qiu, S., Fast, Ambient Temperature and Pressure Ionothermal Synthesis of ThreeDimensional Covalent Organic Frameworks. J. Am. Chem. Soc. 2018, 140 (13), 4494-4498, 10.1021/jacs.8b01320. 113.

Sang, X.; Zhang, J.; Xiang, J.; Cui, J.; Zheng, L.; Zhang, J.; Wu, Z.; Li, Z.; Mo,

G.; Xu, Y.; Song, J.; Liu, C.; Tan, X.; Luo, T.; Zhang, B.; Han, B., Ionic liquid accelerates the crystallization of Zr-based metal–organic frameworks. Nat. Commun. 2017, 8 (1), 175, 10.1038/s41467-017-00226-y. 114.

Yu, S.-Y.; Mahmood, J.; Noh, H.-J.; Seo, J.-M.; Jung, S.-M.; Shin, S.-H.; Im, Y.-

K.; Jeon, I.-Y.; Baek, J.-B., Direct Synthesis of a Covalent Triazine-Based Framework from Aromatic Amides. Angew. Chem. 2018, 130 (28), 8574-8578, 10.1002/ange.201801128. 115.

McKeown, N. B.; Budd, P. M., Polymers of intrinsic microporosity (PIMs): organic

materials for membrane separations, heterogeneous catalysis and hydrogen storage. Chem. Soc. Rev. 2006, 35 (8), 675-683, 10.1039/B600349D. 116.

Xu, Y.; Jin, S.; Xu, H.; Nagai, A.; Jiang, D., Conjugated microporous polymers:

design, synthesis and application. Chem. Soc. Rev. 2013, 42 (20), 8012-8031, 10.1039/C3CS60160A. 117.

Tan, L.; Tan, B., Hypercrosslinked porous polymer materials: design, synthesis, and

applications. Chem. Soc. Rev. 2017, 46 (11), 3322-3356, 10.1039/C6CS00851H. 118.

Ben, T.; Qiu, S., Porous aromatic frameworks: Synthesis, structure and functions.

CrystEngComm 2013, 15 (1), 17-26, 10.1039/C2CE25409C.

62 ACS Paragon Plus Environment

Page 63 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

119.

Chen, L.; Honsho, Y.; Seki, S.; Jiang, D., Light-Harvesting Conjugated Microporous

Polymers: Rapid and Highly Efficient Flow of Light Energy with a Porous Polyphenylene Framework as Antenna. J. Am. Chem. Soc. 2010, 132 (19), 6742-6748, 10.1021/ja100327h. 120. via

Schmidt, J.; Werner, M.; Thomas, A., Conjugated Microporous Polymer Networks Yamamoto

Polymerization.

Macromolecules

2009,

42

(13),

4426-4429,

10.1021/ma9005473. 121. Jiang, J.-X.; Su, F.; Trewin, A.; Wood, C. D.; Campbell, N. L.; Niu, H.; Dickinson, C.;

Ganin, A. Y.;

Rosseinsky, M. J.;

Khimyak, Y. Z.; Cooper, A. I., Conjugated

Microporous Poly(aryleneethynylene) Networks. Angew. Chem. Int. Ed. 2007, 46 (45), 85748578, 10.1002/anie.200701595. 122.

Jiang, J.-X.; Su, F.; Trewin, A.; Wood, C. D.; Niu, H.; Jones, J. T. A.; Khimyak,

Y. Z.; Cooper, A. I., Synthetic Control of the Pore Dimension and Surface Area in Conjugated Microporous Polymer and Copolymer Networks. J. Am. Chem. Soc. 2008, 130 (24), 7710-7720, 10.1021/ja8010176. 123.

Schmidt, J.; Weber, J.; Epping, J. D.; Antonietti, M.; Thomas, A., Microporous

Conjugated Poly(thienylene arylene) Networks. Adv. Mater. 2008, 21 (6), 702-705, 10.1002/adma.200802692. 124.

Katsoulidis, A. P.; Kanatzidis, M. G., Phloroglucinol Based Microporous Polymeric

Organic Frameworks with −OH Functional Groups and High CO2 Capture Capacity. Chem. Mater. 2011, 23 (7), 1818-1824, 10.1021/cm103206x. 125.

Rabbani, M. G.; El-Kaderi, H. M., Synthesis and Characterization of Porous

Benzimidazole-Linked Polymers and Their Performance in Small Gas Storage and Selective Uptake. Chem. Mater. 2012, 24 (8), 1511-1517, 10.1021/cm300407h. 126.

Arab, P.; Rabbani, M. G.; Sekizkardes, A. K.; İslamoğlu, T.; El-Kaderi, H. M.,

Copper(I)-Catalyzed Synthesis of Nanoporous Azo-Linked Polymers: Impact of Textural

63 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 64 of 92

Properties on Gas Storage and Selective Carbon Dioxide Capture. Chem. Mater. 2014, 26 (3), 1385-1392, 10.1021/cm403161e. 127.

Zhao, Y.; Yao, K. X.; Teng, B.; Zhang, T.; Han, Y., A perfluorinated covalent

triazine-based framework for highly selective and water–tolerant CO2 capture. Energy Environ. Sci. 2013, 6 (12), 3684-3692, 10.1039/C3EE42548G. 128.

Konstas, K.; Taylor, J. W.; Thornton, A. W.; Doherty, C. M.; Lim, W. X.; Bastow,

T. J.; Kennedy, D. F.; Wood, C. D.; Cox, B. J.; Hill, J. M.; Hill, A. J.; Hill, M. R., Lithiated Porous Aromatic Frameworks with Exceptional Gas Storage Capacity. Angew. Chem. Int. Ed. 2012, 51 (27), 6639-6642, 10.1002/anie.201201381. 129.

Lu, W.; Yuan, D.; Sculley, J.; Zhao, D.; Krishna, R.; Zhou, H.-C., Sulfonate-

Grafted Porous Polymer Networks for Preferential CO2 Adsorption at Low Pressure. J. Am. Chem. Soc. 2011, 133 (45), 18126-18129, 10.1021/ja2087773. 130.

Lu, W.; Verdegaal, W. M.; Yu, J.; Balbuena, P. B.; Jeong, H.-K.; Zhou, H.-C.,

Building multiple adsorption sites in porous polymer networks for carbon capture applications. Energy Environ. Sci. 2013, 6 (12), 3559-3564, 10.1039/C3EE42226G. 131.

Gu, C.; Huang, N.; Gao, J.; Xu, F.; Xu, Y.; Jiang, D., Controlled Synthesis of

Conjugated Microporous Polymer Films: Versatile Platforms for Highly Sensitive and LabelFree Chemo- and Biosensing. Angew. Chem. Int. Ed. 2014, 53 (19), 4850-4855, 10.1002/anie.201402141. 132.

Kou, Y.; Xu, Y.; Guo, Z.; Jiang, D., Supercapacitive Energy Storage and Electric

Power Supply Using an Aza-Fused π-Conjugated Microporous Framework. Angew. Chem. Int. Ed. 2011, 50 (37), 8753-8757, 10.1002/anie.201103493. 133.

Gu, C.; Huang, N.; Wu, Y.; Xu, H.; Jiang, D., Design of Highly Photofunctional

Porous Polymer Films with Controlled Thickness and Prominent Microporosity. Angew. Chem. Int. Ed. 2015, 54 (39), 11540-11544, 10.1002/anie.201504786.

64 ACS Paragon Plus Environment

Page 65 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

134.

Gu, C.; Huang, N.; Chen, Y.; Qin, L.; Xu, H.; Zhang, S.; Li, F.; Ma, Y.; Jiang, D.,

π-Conjugated Microporous Polymer Films: Designed Synthesis, Conducting Properties, and Photoenergy Conversions. Angew. Chem. Int. Ed. 2015, 54 (46), 13594-13598, 10.1002/anie.201506570. 135.

Dawson, R.; Laybourn, A.; Khimyak, Y. Z.; Adams, D. J.; Cooper, A. I., High

Surface Area Conjugated Microporous Polymers: The Importance of Reaction Solvent Choice. Macromolecules 2010, 43 (20), 8524-8530, 10.1021/ma101541h. 136.

Sun, Q.; Aguila, B.; Perman, J.; Ivanov, A. S.; Bryantsev, V. S.; Earl, L. D.;

Abney, C. W.; Wojtas, L.; Ma, S., Bio-inspired nano-traps for uranium extraction from seawater and recovery from nuclear waste. Nat. Commun. 2018, 9 (1), 1644, 10.1038/s41467018-04032-y. 137.

Byun, Y.; Coskun, A., Epoxy-Functionalized Porous Organic Polymers via the Diels–

Alder Cycloaddition Reaction for Atmospheric Water Capture. Angew. Chem. Int. Ed. 2018, 57 (12), 3173-3177, 10.1002/anie.201800380. 138. Patel, H. A.; Hyun Je, S.; Park, J.; Chen, D. P.; Jung, Y.; Yavuz, C. T.; Coskun, A., Unprecedented high-temperature CO2 selectivity in N2-phobic nanoporous covalent organic polymers. Nat. Commun. 2013, 4, 1357, 10.1038/ncomms2359. 139.

Ren, S.; Bojdys, M. J.; Dawson, R.; Laybourn, A.; Khimyak, Y. Z.; Adams, D. J.;

Cooper, A. I., Porous, Fluorescent, Covalent Triazine-Based Frameworks Via RoomTemperature and Microwave-Assisted Synthesis. Adv. Mater. 2012, 24 (17), 2357-2361, 10.1002/adma.201200751. 140.

Wang, K.; Yang, L. M.; Wang, X.; Guo, L.; Cheng, G.; Zhang, C.; Jin, S.; Tan,

B.; Cooper, A., Covalent Triazine Frameworks via a Low‐Temperature Polycondensation Approach. Angew. Chem. Int. Ed. 2017, 56 (45), 14149-14153, 10.1002/anie.201708548.

65 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

141.

Page 66 of 92

Xie, J.; Shevlin, S. A.; Ruan, Q.; Moniz, S. J. A.; Liu, Y.; Liu, X.; Li, Y.; Lau, C.

C.; Guo, Z. X.; Tang, J., Efficient visible light-driven water oxidation and proton reduction by an ordered covalent triazine-based framework. Energy Environ. Sci. 2018, 11 (6), 16171624, 10.1039/C7EE02981K. 142.

Mal, P.; Schultz, D.; Beyeh, K.; Rissanen, K.; Nitschke, J. R., An Unlockable–

Relockable Iron Cage by Subcomponent Self-Assembly. Angew. Chem. Int. Ed. 2008, 47 (43), 8297-8301, 10.1002/anie.200803066. 143.

Guillerm, V.; Kim, D.; Eubank, J. F.; Luebke, R.; Liu, X.; Adil, K.; Lah, M. S.;

Eddaoudi, M., A supermolecular building approach for the design and construction of metal– organic frameworks. Chem. Soc. Rev. 2014, 43 (16), 6141-6172, 10.1039/C4CS00135D. 144.

Bloch, W. M.; Holstein, J. J.; Dittrich, B.; Hiller, W.; Clever, G. H., Hierarchical

Assembly of an Interlocked M8L16 Container. Angew. Chem. Int. Ed. 2018, 57 (19), 55345538, 10.1002/anie.201800490. 145.

Seidel, S. R.; Stang, P. J., High-Symmetry Coordination Cages via Self-Assembly.

Acc. Chem. Res 2002, 35 (11), 972-983, 10.1021/ar010142d. 146.

Tranchemontagne, D. J.; Ni, Z.; O'Keeffe, M.; Yaghi, O. M., Reticular Chemistry of

Metal–Organic

Polyhedra.

Angew.

Chem.

Int.

Ed.

2008,

47

(28),

5136-5147,

10.1002/anie.200705008. 147.

Ronson, T. K.; Zarra, S.; Black, S. P.; Nitschke, J. R., Metal–organic container

molecules through subcomponent self-assembly. Chem. Commun. 2013, 49 (25), 2476-2490, 10.1039/C2CC36363A. 148.

Zhu, R.; Lübben, J.; Dittrich, B.; Clever, G. H., Stepwise Halide-Triggered Double

and Triple Catenation of Self-Assembled Coordination Cages. Angew. Chem. Int. Ed. 2014, 54 (9), 2796-2800, 10.1002/anie.201408068.

66 ACS Paragon Plus Environment

Page 67 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

149.

Yang, L.; Jing, X.; An, B.; He, C.; Yang, Y.; Duan, C., Binding of anions in triply

interlocked coordination catenanes and dynamic allostery for dehalogenation reactions. Chem. Sci. 2018, 9 (4), 1050-1057, 10.1039/C7SC04070A. 150.

Song, Q.; Jiang, S.; Hasell, T.; Liu, M.; Sun, S.; Cheetham, A. K.; Sivaniah, E.;

Cooper, A. I., Porous Organic Cage Thin Films and Molecular-Sieving Membranes. Adv. Mater. 2016, 28 (13), 2629-2637, 10.1002/adma.201505688. 151.

Uemura, T.; Nakanishi, R.; Mochizuki, S.; Kitagawa, S.; Mizuno, M., Radical

Polymerization of Vinyl Monomers in Porous Organic Cages. Angew. Chem. 2016, 128 (22), 6553-6557, 10.1002/ange.201601587. 152.

Yang, X.; Sun, J.-K.; Kitta, M.; Pang, H.; Xu, Q., Encapsulating highly catalytically

active metal nanoclusters inside porous organic cages. Nat. Catal. 2018, 1 (3), 214-220, 10.1038/s41929-018-0030-8. 153.

Giri, N.; Davidson, C. E.; Melaugh, G.; Del Pópolo, M. G.; Jones, J. T. A.; Hasell,

T.; Cooper, A. I.; Horton, P. N.; Hursthouse, M. B.; James, S. L., Alkylated organic cages: from

porous

crystals

to

neat

liquids.

Chem.

Sci.

2012,

3

(6),

2153-2157,

10.1039/C2SC01007K. 154. Giri, N.; Del Pópolo, M. G.; Melaugh, G.; Greenaway, R. L.; Rätzke, K.; Koschine, T.; Pison, L.; Gomes, M. F. C.; Cooper, A. I.; James, S. L., Liquids with permanent porosity. Nature 2015, 527, 216, 10.1038/nature16072. 155.

Jones, J. T. A.; Holden, D.; Mitra, T.; Hasell, T.; Adams, D. J.; Jelfs, K. E.;

Trewin, A.; Willock, D. J.; Day, G. M.; Bacsa, J.; Steiner, A.; Cooper, A. I., On–Off Porosity Switching in a Molecular Organic Solid. Angew. Chem. Int. Ed. 2010, 50 (3), 749753, 10.1002/anie.201006030. 156.

Jones, J. T. A.; Hasell, T.; Wu, X.; Bacsa, J.; Jelfs, K. E.; Schmidtmann, M.;

Chong, S. Y.; Adams, D. J.; Trewin, A.; Schiffman, F.; Cora, F.; Slater, B.; Steiner, A.;

67 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 68 of 92

Day, G. M.; Cooper, A. I., Modular and predictable assembly of porous organic molecular crystals. Nature 2011, 474, 367, 10.1038/nature10125. 157. Greenaway, R. L.; Santolini, V.; Bennison, M. J.; Alston, B. M.; Pugh, C. J.; Little, M. A.; Miklitz, M.; Eden-Rump, E. G. B.; Clowes, R.; Shakil, A.; Cuthbertson, H. J.; Armstrong, H.; Briggs, M. E.; Jelfs, K. E.; Cooper, A. I., High-throughput discovery of organic cages and catenanes using computational screening fused with robotic synthesis. Nat. Commun. 2018, 9 (1), 2849, 10.1038/s41467-018-05271-9. 158.

Pulido, A.; Chen, L.; Kaczorowski, T.; Holden, D.; Little, M. A.; Chong, S. Y.;

Slater, B. J.; McMahon, D. P.; Bonillo, B.; Stackhouse, C. J.; Stephenson, A.; Kane, C. M.; Clowes, R.; Hasell, T.; Cooper, A. I.; Day, G. M., Functional materials discovery using energy–structure–function maps. Nature 2017, 543, 657, 10.1038/nature21419. 159.

Kitchin, M.; Konstas, K.; Sumby, C. J.; Czyz, M. L.; Valente, P.; Hill, M. R.;

Polyzos, A.; Doonan, C. J., Continuous flow synthesis of a carbon-based molecular cage macrocycle via a three-fold homocoupling reaction. Chem. Commun. 2015, 51 (75), 1423114234, 10.1039/C5CC05181A. 160.

Moneypenny, T. P.; Walter, N. P.; Cai, Z.; Miao, Y.-R.; Gray, D. L.; Hinman, J. J.;

Lee, S.; Zhang, Y.; Moore, J. S., Impact of Shape Persistence on the Porosity of Molecular Cages. J. Am. Chem. Soc. 2017, 139 (8), 3259-3264, 10.1021/jacs.7b00189. 161.

Buyukcakir, O.; Seo, Y.; Coskun, A., Thinking Outside the Cage: Controlling the

Extrinsic Porosity and Gas Uptake Properties of Shape-Persistent Molecular Cages in Nanoporous

Polymers.

Chem.

Mater.

2015,

27

(11),

4149-4155,

10.1021/acs.chemmater.5b01346. 162.

Nam, D.; Huh, J.; Lee, J.; Kwak, J. H.; Jeong, H. Y.; Choi, K.; Choe, W., Cross-

linking Zr-based metal–organic polyhedra via postsynthetic polymerization. Chem. Sci. 2017, 8 (11), 7765-7771, 10.1039/C7SC03847J.

68 ACS Paragon Plus Environment

Page 69 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

163. I.;

Bushell, A. F.; Budd, P. M.; Attfield, M. P.; Jones, J. T. A.; Hasell, T.; Cooper, A. Bernardo, P.;

Bazzarelli, F.;

Clarizia, G.; Jansen, J. C., Nanoporous Organic

Polymer/Cage Composite Membranes. Angew. Chem. Int. Ed. 2012, 52 (4), 1253-1256, 10.1002/anie.201206339. 164.

Perez, E. V.;

Balkus, K. J.;

Ferraris, J. P.; Musselman, I. H., Metal-organic

polyhedra 18 mixed-matrix membranes for gas separation. J. Membrane Sci. 2014, 463, 8293, 10.1016/j.memsci.2014.03.045. 165.

Trost, B. M., The atom economy--a search for synthetic efficiency. Science 1991, 254

(5037), 1471. 166.

Schaate, A.; Roy, P.; Godt, A.; Lippke, J.; Waltz, F.; Wiebcke, M.; Behrens, P.,

Modulated Synthesis of Zr-Based Metal–Organic Frameworks: From Nano to Single Crystals. Chem. Eur. J. 2011, 17 (24), 6643-6651, 10.1002/chem.201003211. 167.

Mandich, N., Chemistry & Theory of Chromium Deposition: Part 1--Chemistry. Plat.

Surf. Finish 1997, 84 (5), 108-115. 168. Reinsch, H.; Fröhlich, D.; Waitschat, S.; Chavan, S.; Lillerud, K.-P.; Henninger, S.; Stock, N., Optimisation of synthesis conditions for UiO-66-CO2H towards scale-up and its vapour sorption properties. React. Chem. Eng. 2018, 3 (3), 365-370, 10.1039/C7RE00214A. 169. Majano, G.; Pérez-Ramírez, J., Scalable Room-Temperature Conversion of Copper(II) Hydroxide into HKUST-1 (Cu3(btc)2). Adv. Mater. 2012, 25 (7), 1052-1057, 10.1002/adma.201203664. 170.

Zhao, J.; Nunn, W. T.; Lemaire, P. C.; Lin, Y.; Dickey, M. D.; Oldham, C. J.;

Walls, H. J.; Peterson, G. W.; Losego, M. D.; Parsons, G. N., Facile Conversion of Hydroxy Double Salts to Metal–Organic Frameworks Using Metal Oxide Particles and Atomic Layer Deposition Thin-Film Templates. J. Am. Chem. Soc. 2015, 137 (43), 13756-13759, 10.1021/jacs.5b08752.

69 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

171.

Page 70 of 92

Riccò, R.; Linder-Patton, O.; Sumida, K.; Styles, M. J.; Liang, K.; Amenitsch, H.;

Doonan, C. J.; Falcaro, P., Conversion of Copper Carbonate into a Metal–Organic Framework. Chem. Mater. 2018, 30 (16), 5630-5638, 10.1021/acs.chemmater.8b01891. 172.

Maserati, L.;

Meckler, S. M.;

Li, C.; Helms, B. A., Minute-MOFs: Ultrafast

Synthesis of M2(dobpdc) Metal–Organic Frameworks from Divalent Metal Oxide Colloidal Nanocrystals. Chem. Mater. 2016, 28 (5), 1581-1588, 10.1021/acs.chemmater.6b00494. 173.

Li, Z.; Wu, Y.-n.; Li, J.; Zhang, Y.; Zou, X.; Li, F., The Metal–Organic Framework

MIL-53(Al) Constructed from Multiple Metal Sources: Alumina, Aluminum Hydroxide, and Boehmite. Chem. Eur. J. 2015, 21 (18), 6913-6920, 10.1002/chem.201406531. 174.

Moran, C. M.; Joshi, J. N.; Marti, R. M.; Hayes, S. E.; Walton, K. S., Structured

Growth of Metal–Organic Framework MIL-53(Al) from Solid Aluminum Carbide Precursor. J. Am. Chem. Soc. 2018, 140 (29), 9148-9153, 10.1021/jacs.8b04369. 175.

Sun, Y.; Hu, S.; Song, C.; Miao, S.; Jiang, Z.; Jiang, X.; Zhao, J.; Guo, X.; Liu,

Y., Two-dimensional transition metal dichalcogenides as metal sources of metal–organic frameworks. Chem. Commun. 2018, 54 (29), 3664-3667, 10.1039/C8CC01258J. 176.

Stassen, I.; Styles, M.; Grenci, G.; Gorp, Hans V.; Vanderlinden, W.; Feyter,

Steven D.;

Falcaro, P.;

Vos, D. D.;

Vereecken, P.; Ameloot, R., Chemical vapour

deposition of zeolitic imidazolate framework thin films. Nat. Mater. 2015, 15, 304, 10.1038/nmat4509. 177.

Lin, J.-B.; Lin, R.-B.; Cheng, X.-N.; Zhang, J.-P.; Chen, X.-M., Solvent/additive-

free synthesis of porous/zeolitic metal azolate frameworks from metal oxide/hydroxide. Chem. Commun. 2011, 47 (32), 9185-9187, 10.1039/C1CC12763B. 178.

Lanchas, M.; Vallejo-Sánchez, D.; Beobide, G.; Castillo, O.; Aguayo, A. T.;

Luque, A.; Román, P., A direct reaction approach for the synthesis of zeolitic imidazolate

70 ACS Paragon Plus Environment

Page 71 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

frameworks: template and temperature mediated control on network topology and crystal size. Chem. Commun. 2012, 48 (79), 9930-9932, 10.1039/C2CC34787C. 179.

Stassen, I.; Campagnol, N.; Fransaer, J.; Vereecken, P.; De Vos, D.; Ameloot, R.,

Solvent-free synthesis of supported ZIF-8 films and patterns through transformation of deposited

zinc

oxide

precursors.

CrystEngComm

2013,

15

(45),

9308-9311,

10.1039/C3CE41025K. 180.

Ma, X.; Kumar, P.; Mittal, N.; Khlyustova, A.; Daoutidis, P.; Mkhoyan, K. A.;

Tsapatsis, M., Zeolitic imidazolate framework membranes made by ligand-induced permselectivation. Science 2018, 361 (6406), 1008. 181.

Corma, A.; Iborra, S.; Velty, A., Chemical Routes for the Transformation of Biomass

into Chemicals. Chem. Rev. 2007, 107 (6), 2411-2502, 10.1021/cr050989d. 182.

Motagamwala, A. H.; Won, W.; Sener, C.; Alonso, D. M.; Maravelias, C. T.;

Dumesic, J. A., Toward biomass-derived renewable plastics: Production of 2,5furandicarboxylic acid from fructose. Sci Adv. 2018, 4 (1), 10.1126/sciadv.aap9722. 183.

Cadiau, A.; Lee, J. S.; Damasceno Borges, D.; Fabry, P.; Devic, T.; Wharmby, M.

T.; Martineau, C.; Foucher, D.; Taulelle, F.; Jun, C.-H.; Hwang, Y. K.; Stock, N.; De Lange, M. F.; Kapteijn, F.; Gascon, J.; Maurin, G.; Chang, J.-S.; Serre, C., Design of Hydrophilic Metal Organic Framework Water Adsorbents for Heat Reallocation. Adv. Mater. 2015, 27 (32), 4775-4780, 10.1002/adma.201502418. 184.

Dreischarf, A. C.; Lammert, M.; Stock, N.; Reinsch, H., Green Synthesis of Zr-

CAU-28: Structure and Properties of the First Zr-MOF Based on 2,5-Furandicarboxylic Acid. Inorg. Chem. 2017, 56 (4), 2270-2277, 10.1021/acs.inorgchem.6b02969. 185.

Horcajada, P.; Chalati, T.; Serre, C.; Gillet, B.; Sebrie, C.; Baati, T.; Eubank, J. F.;

Heurtaux, D.; Clayette, P.; Kreuz, C.; Chang, J.-S.; Hwang, Y. K.; Marsaud, V.; Bories, P.-N.; Cynober, L.; Gil, S.; Férey, G.; Couvreur, P.; Gref, R., Porous metal–organic-

71 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 72 of 92

framework nanoscale carriers as a potential platform for drug delivery and imaging. Nat. Mater. 2009, 9, 172, 10.1038/nmat2608. 186.

Witman, M.; Ling, S.; Gladysiak, A.; Stylianou, K. C.; Smit, B.; Slater, B.;

Haranczyk, M., Rational Design of a Low-Cost, High-Performance Metal–Organic Framework for Hydrogen Storage and Carbon Capture. J. Phys. Chem. C 2017, 121 (2), 1171-1181, 10.1021/acs.jpcc.6b10363. 187.

Guo, X.; Tian, Y.; Zhang, M.; Li, Y.; Wen, R.; Li, X.; Li, X.; Xue, Y.; Ma, L.;

Xia, C.; Li, S., Mechanistic Insight into Hydrogen-Bond-Controlled Crystallinity and Adsorption Property of Covalent Organic Frameworks from Flexible Building Blocks. Chem. Mater. 2018, 30 (7), 2299-2308, 10.1021/acs.chemmater.7b05121. 188. Smaldone, R. A.; Forgan, R. S.; Furukawa, H.; Gassensmith, J. J.; Slawin, A. M. Z.; Yaghi, O. M.; Stoddart, J. F., Metal–Organic Frameworks from Edible Natural Products. Angew. Chem. Int. Ed. 2010, 49 (46), 8630-8634, 10.1002/anie.201002343. 189.

Forgan, R. S.; Smaldone, R. A.; Gassensmith, J. J.; Furukawa, H.; Cordes, D. B.;

Li, Q.; Wilmer, C. E.; Botros, Y. Y.; Snurr, R. Q.; Slawin, A. M. Z.; Stoddart, J. F., Nanoporous Carbohydrate Metal–Organic Frameworks. J. Am. Chem. Soc. 2012, 134 (1), 406-417, 10.1021/ja208224f. 190.

Hartlieb, K. J.; Peters, A. W.; Wang, T. C.; Deria, P.; Farha, O. K.; Hupp, J. T.;

Stoddart, J. F., Functionalised cyclodextrin-based metal–organic frameworks. Chem. Commun. 2017, 53 (54), 7561-7564, 10.1039/C7CC03345A. 191.

Patel, H. A.;

Anamimoghadam, O.;

Islamoglu, T.;

Liu, Z.;

Malliakas, C. D.;

Nalluri, S. K. M.;

Samanta, A.;

Farha, O. K.; Stoddart, J. F., Noninvasive

Substitution of K+ Sites in Cyclodextrin Metal–Organic Frameworks by Li+ Ions. J. Am. Chem. Soc. 2017, 139 (32), 11020-11023, 10.1021/jacs.7b06287.

72 ACS Paragon Plus Environment

Page 73 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

192.

Holcroft, J. M.; Hartlieb, K. J.; Moghadam, P. Z.; Bell, J. G.; Barin, G.; Ferris, D.

P.; Bloch, E. D.; Algaradah, M. M.; Nassar, M. S.; Botros, Y. Y.; Thomas, K. M.; Long, J. R.; Snurr, R. Q.; Stoddart, J. F., Carbohydrate-Mediated Purification of Petrochemicals. J. Am. Chem. Soc. 2015, 137 (17), 5706-5719, 10.1021/ja511878b. 193.

Hartlieb, K. J.; Holcroft, J. M.; Moghadam, P. Z.; Vermeulen, N. A.; Algaradah, M.

M.; Nassar, M. S.; Botros, Y. Y.; Snurr, R. Q.; Stoddart, J. F., CD-MOF: A Versatile Separation Medium. J. Am. Chem. Soc. 2016, 138 (7), 2292-2301, 10.1021/jacs.5b12860. 194.

Zhang, Y.; Duan, J.; Ma, D.; Li, P.; Li, S.; Li, H.; Zhou, J.; Ma, X.; Feng, X.;

Wang, B., Three-Dimensional Anionic Cyclodextrin-Based Covalent Organic Frameworks. Angew. Chem. Int. Ed. 2017, 56 (51), 16313-16317, 10.1002/anie.201710633. 195.

Henderson, R. K.; Jiménez-González, C.; Constable, D. J. C.; Alston, S. R.; Inglis,

G. G. A.; Fisher, G.; Sherwood, J.; Binks, S. P.; Curzons, A. D., Expanding GSK's solvent selection guide – embedding sustainability into solvent selection starting at medicinal chemistry. Green Chem. 2011, 13 (4), 854-862, 10.1039/C0GC00918K. 196.

Bromberg, L.; Diao, Y.; Wu, H.; Speakman, S. A.; Hatton, T. A., Chromium(III)

Terephthalate Metal Organic Framework (MIL-101): HF-Free Synthesis, Structure, Polyoxometalate Composites, and Catalytic Properties. Chem. Mater. 2012, 24 (9), 16641675, 10.1021/cm2034382. 197.

Zhao, T.; Jeremias, F.; Boldog, I.; Nguyen, B.; Henninger, S. K.; Janiak, C., High-

yield, fluoride-free and large-scale synthesis of MIL-101(Cr). Dalton Trans. 2015, 44 (38), 16791-16801, 10.1039/C5DT02625C. 198.

Hu, Z.; Peng, Y.; Kang, Z.; Qian, Y.; Zhao, D., A Modulated Hydrothermal (MHT)

Approach for the Facile Synthesis of UiO-66-Type MOFs. Inorg. Chem. 2015, 54 (10), 48624868, 10.1021/acs.inorgchem.5b00435.

73 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

199.

Page 74 of 92

Hu, Z.; Nalaparaju, A.; Peng, Y.; Jiang, J.; Zhao, D., Modulated Hydrothermal

Synthesis of UiO-66(Hf)-Type Metal–Organic Frameworks for Optimal Carbon Dioxide Separation. Inorg. Chem. 2016, 55 (3), 1134-1141, 10.1021/acs.inorgchem.5b02312. 200.

Hennig, C.; Weiss, S.; Kraus, W.; Kretzschmar, J.; Scheinost, A. C., Solution

Species and Crystal Structure of Zr(IV) Acetate. Inorg. Chem. 2017, 56 (5), 2473-2480, 10.1021/acs.inorgchem.6b01624. 201.

Hu, Z.; Gami, A.; Wang, Y.; Zhao, D., A Triphasic Modulated Hydrothermal

Approach for the Synthesis of Multivariate Metal–Organic Frameworks with Hydrophobic Moieties for Highly Efficient Moisture-Resistant CO2 Capture. Adv. Sustainable Syst. 2017, 1 (11), 1700092, 10.1002/adsu.201700092. 202.

Thote, J.; Barike Aiyappa, H.; Rahul Kumar, R.; Kandambeth, S.; Biswal, B. P.;

Balaji Shinde, D.; Chaki Roy, N.; Banerjee, R., Constructing covalent organic frameworks in water

via

dynamic

covalent

bonding.

IUCrJ

2016,

3

(6),

402-407,

10.1107/S2052252516013762. 203.

Chevreau, H.; Permyakova, A.; Nouar, F.; Fabry, P.; Livage, C.; Ragon, F.;

Garcia-Marquez, A.; Devic, T.; Steunou, N.; Serre, C.; Horcajada, P., Synthesis of the biocompatible and highly stable MIL-127(Fe): from large scale synthesis to particle size control. CrystEngComm 2016, 18 (22), 4094-4101, 10.1039/C5CE01864A. 204.

Ploskonka, A. M.;

Marzen, S. E.; DeCoste, J. B., Facile Synthesis and Direct

Activation of Zirconium Based Metal–Organic Frameworks from Acetone. Ind. Eng. Chem. Res. 2017, 56 (6), 1478-1484, 10.1021/acs.iecr.6b04361. 205.

Zhang, J.;

White, G. B.;

Ryan, M. D.;

Hunt, A. J.; Katz, M. J.,

Dihydrolevoglucosenone (Cyrene) As a Green Alternative to N,N-Dimethylformamide (DMF) in

MOF

Synthesis.

ACS

Sustainable

Chem.

Eng.

2016,

4

(12),

7186-7192,

10.1021/acssuschemeng.6b02115.

74 ACS Paragon Plus Environment

Page 75 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

206.

Dienstmaier, J. F.; Gigler, A. M.; Goetz, A. J.; Knochel, P.; Bein, T.; Lyapin, A.;

Reichlmaier, S.; Heckl, W. M.; Lackinger, M., Synthesis of Well-Ordered COF Monolayers: Surface Growth of Nanocrystalline Precursors versus Direct On-Surface Polycondensation. ACS Nano 2011, 5 (12), 9737-9745, 10.1021/nn2032616. 207.

Dienstmaier, J. F.; Medina, D. D.; Dogru, M.; Knochel, P.; Bein, T.; Heckl, W. M.;

Lackinger, M., Isoreticular Two-Dimensional Covalent Organic Frameworks Synthesized by On-Surface Condensation of Diboronic Acids. ACS Nano 2012, 6 (8), 7234-7242, 10.1021/nn302363d. 208.

Medina, D. D.;

Rotter, J. M.;

Hu, Y.;

Dogru, M.;

Werner, V.;

Auras, F.;

Markiewicz, J. T.; Knochel, P.; Bein, T., Room Temperature Synthesis of Covalent–Organic Framework Films through Vapor-Assisted Conversion. J. Am. Chem. Soc. 2015, 137 (3), 1016-1019, 10.1021/ja510895m. 209.

Virmani, E.; Rotter, J. M.; Mähringer, A.; von Zons, T.; Godt, A.; Bein, T.;

Wuttke, S.; Medina, D. D., On-Surface Synthesis of Highly Oriented Thin Metal–Organic Framework Films through Vapor-Assisted Conversion. J. Am. Chem. Soc. 2018, 140 (14), 4812-4819, 10.1021/jacs.7b08174. 210.

Shi, Q.; Chen, Z.; Song, Z.; Li, J.; Dong, J., Synthesis of ZIF-8 and ZIF-67 by

Steam-Assisted Conversion and an Investigation of Their Tribological Behaviors. Angew. Chem. Int. Ed. 2010, 50 (3), 672-675, 10.1002/anie.201004937. 211.

Mottillo, C.; Lu, Y.; Pham, M.-H.; Cliffe, M. J.; Do, T.-O.; Friščić, T., Mineral

neogenesis as an inspiration for mild, solvent-free synthesis of bulk microporous metal– organic frameworks from metal (Zn, Co) oxides. Green Chem. 2013, 15 (8), 2121-2131, 10.1039/C3GC40520F.

75 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

212.

Page 76 of 92

Užarević, K.; Wang, T. C.; Moon, S.-Y.; Fidelli, A. M.; Hupp, J. T.; Farha, O. K.;

Friščić, T., Mechanochemical and solvent-free assembly of zirconium-based metal–organic frameworks. Chem. Commun. 2016, 52 (10), 2133-2136, 10.1039/C5CC08972G. 213.

Lu, N.; Zhou, F.; Jia, H.; Wang, H.; Fan, B.; Li, R., Dry-Gel Conversion Synthesis

of Zr-Based Metal–Organic Frameworks. Ind. Eng. Chem. Res. 2017, 56 (48), 14155-14163, 10.1021/acs.iecr.7b04010. 214.

Ahmed, I.; Jeon, J.; Khan, N. A.; Jhung, S. H., Synthesis of a Metal–Organic

Framework, Iron-Benezenetricarboxylate, from Dry Gels in the Absence of Acid and Salt. Cryst Growth Des. 2012, 12 (12), 5878-5881, 10.1021/cg3014317. 215.

Kim, J.; Lee, Y.-R.; Ahn, W.-S., Dry-gel conversion synthesis of Cr-MIL-101 aided

by grinding: high surface area and high yield synthesis with minimum purification. Chem. Commun. 2013, 49 (69), 7647-7649, 10.1039/C3CC44559C. 216.

Jiang, Y.; Huang, W.; Wang, J.; Wu, Q.; Wang, H.; Pan, L.; Liu, X., Green,

scalable and morphology controlled synthesis of nanofibrous covalent organic frameworks and their nanohybrids through a vapor-assisted solid-state approach. J. Mater. Chem. A 2014, 2 (22), 8201-8204, 10.1039/C4TA00555D. 217.

Bian, C.; Zhang, C.; Pan, S.; Chen, F.; Zhang, W.; Meng, X.; Maurer, S.; Dai, D.;

Parvulescu, A.-N.; Müller, U.; Xiao, F.-S., Generalized high-temperature synthesis of zeolite catalysts with unpredictably high space-time yields (STYs). J. Mater. Chem. A 2017, 5 (6), 2613-2618, 10.1039/C6TA09866E. 218.

Do, J.-L.; Friščić, T., Mechanochemistry: A Force of Synthesis. ACS Cent. Sci. 2017,

3 (1), 13-19, 10.1021/acscentsci.6b00277. 219.

Pichon, A.; Lazuen-Garay, A.; James, S. L., Solvent-free synthesis of a microporous

metal–organic framework. CrystEngComm 2006, 8 (3), 211-214, 10.1039/B513750K.

76 ACS Paragon Plus Environment

Page 77 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

220.

Friščić, T.; Reid, D. G.; Halasz, I.; Stein, R. S.; Dinnebier, R. E.; Duer, M. J., Ion-

and Liquid-Assisted Grinding: Improved Mechanochemical Synthesis of Metal–Organic Frameworks Reveals Salt Inclusion and Anion Templating. Angew. Chem. Int. Ed. 2010, 49 (4), 712-715, 10.1002/anie.200906583. 221.

Friščić, T.;

Childs, S. L.;

Rizvi, S. A. A.; Jones, W., The role of solvent in

mechanochemical and sonochemical cocrystal formation: a solubility-based approach for predicting

cocrystallisation

outcome.

CrystEngComm

2009,

11

(3),

418-426,

10.1039/B815174A. 222. Friščić, T.; Halasz, I.; Beldon, P. J.; Belenguer, A. M.; Adams, F.; Kimber, S. A. J.; Honkimäki, V.; Dinnebier, R. E., Real-time and in situ monitoring of mechanochemical milling reactions. Nat Chem. 2012, 5, 66, 10.1038/nchem.1505. 223.

Katsenis, A. D.; Puškarić, A.; Štrukil, V.; Mottillo, C.; Julien, P. A.; Užarević, K.;

Pham, M.-H.; Do, T.-O.; Kimber, S. A. J.; Lazić, P.; Magdysyuk, O.; Dinnebier, R. E.; Halasz, I.; Friščić, T., In situ X-ray diffraction monitoring of a mechanochemical reaction reveals a unique topology metal-organic framework. Nat. Commun. 2015, 6, 6662, 10.1038/ncomms7662. 224.

Akimbekov, Z.; Katsenis, A. D.; Nagabhushana, G. P.; Ayoub, G.; Arhangelskis,

M.; Morris, A. J.; Friščić, T.; Navrotsky, A., Experimental and Theoretical Evaluation of the Stability of True MOF Polymorphs Explains Their Mechanochemical Interconversions. J. Am. Chem. Soc. 2017, 139 (23), 7952-7957, 10.1021/jacs.7b03144. 225.

Brekalo, I.; Kane, C. M.; Ley, A. N.; Ramirez, J. R.; Friščić, T.; Holman, K. T.,

Use of a “Shoe-Last” Solid-State Template in the Mechanochemical Synthesis of HighPorosity RHO-Zinc Imidazolate. J. Am. Chem. Soc. 2018, 140 (32), 10104-10108, 10.1021/jacs.8b05471.

77 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

226.

Page 78 of 92

Julien, P. A.; Užarević, K.; Katsenis, A. D.; Kimber, S. A. J.; Wang, T.; Farha, O.

K.; Zhang, Y.; Casaban, J.; Germann, L. S.; Etter, M.; Dinnebier, R. E.; James, S. L.; Halasz, I.; Friščić, T., In Situ Monitoring and Mechanism of the Mechanochemical Formation of a Microporous MOF-74 Framework. J. Am. Chem. Soc. 2016, 138 (9), 2929-2932, 10.1021/jacs.5b13038. 227. Fidelli, A. M.; Karadeniz, B.; Howarth, A. J.; Huskić, I.; Germann, L. S.; Halasz, I.; Etter, M.; Moon, S.-Y.; Dinnebier, R. E.; Stilinović, V.; Farha, O. K.; Friščić, T.; Užarević, K., Green and rapid mechanosynthesis of high-porosity NU- and UiO-type metal–organic frameworks. Chem. Commun. 2018, 54 (51), 6999-7002, 10.1039/C8CC03189D. 228.

Gracin, D.; Štrukil, V.; Friščić, T.; Halasz, I.; Užarević, K., Laboratory Real-Time

and In Situ Monitoring of Mechanochemical Milling Reactions by Raman Spectroscopy. Angew. Chem. Int. Ed. 2014, 53 (24), 6193-6197, 10.1002/anie.201402334. 229.

Užarević, K.; Ferdelji, N.; Mrla, T.; Julien, P. A.; Halasz, B.; Friščić, T.; Halasz, I.,

Enthalpy vs. friction: heat flow modelling of unexpected temperature profiles in mechanochemistry of metal–organic frameworks. Chem. Sci. 2018, 9 (9), 2525-2532, 10.1039/C7SC05312F. 230.

Halasz, I.;

Kimber, S. A. J.;

Beldon, P. J.;

Belenguer, A. M.;

Adams, F.;

Honkimäki, V.; Nightingale, R. C.; Dinnebier, R. E.; Friščić, T., In situ and real-time monitoring of mechanochemical milling reactions using synchrotron X-ray diffraction. Nat. Protoc. 2013, 8, 1718, 10.1038/nprot.2013.100. 231.

Biswal, B. P.; Chandra, S.; Kandambeth, S.; Lukose, B.; Heine, T.; Banerjee, R.,

Mechanochemical Synthesis of Chemically Stable Isoreticular Covalent Organic Frameworks. J. Am. Chem. Soc. 2013, 135 (14), 5328-5331, 10.1021/ja4017842. 232.

Das, G.;

Balaji Shinde, D.;

Kandambeth, S.;

Biswal, B. P.; Banerjee, R.,

Mechanosynthesis of imine, β-ketoenamine, and hydrogen-bonded imine-linked covalent

78 ACS Paragon Plus Environment

Page 79 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

organic frameworks using liquid-assisted grinding. Chem. Commun. 2014, 50 (84), 1261512618, 10.1039/C4CC03389B. 233.

Karak, S.; Kandambeth, S.; Biswal, B. P.; Sasmal, H. S.; Kumar, S.; Pachfule, P.;

Banerjee, R., Constructing Ultraporous Covalent Organic Frameworks in Seconds via an Organic

Terracotta

Process.

J.

Am.

Chem.

Soc.

2017,

139

(5),

1856-1862,

10.1021/jacs.6b08815. 234.

Kandambeth, S.; Biswal, B. P.; Chaudhari, H. D.; Rout, K. C.; H, S. K.; Mitra, S.;

Karak, S.; Das, A.; Mukherjee, R.; Kharul, U. K.; Banerjee, R., Selective Molecular Sieving in Self‐Standing Porous Covalent‐Organic‐Framework Membranes. Adv. Mater. 2017, 29 (2), 10.1002/adma.201603945. 235.

Karak, S.;

Kumar, S.;

Pachfule, P.; Banerjee, R., Porosity Prediction through

Hydrogen Bonding in Covalent Organic Frameworks. J. Am. Chem. Soc. 2018, 140 (15), 5138-5145, 10.1021/jacs.7b13558. 236.

Zhang, P.;

Jiang, X.;

Wan, S.; Dai, S., Advancing polymers of intrinsic

microporosity by mechanochemistry. J. Mater. Chem. A 2015, 3 (13), 6739-6741, 10.1039/C4TA07196D. 237.

Zhu, X.; Tian, C.; Jin, T.; Browning, K. L.; Sacci, R. L.; Veith, G. M.; Dai, S.,

Solid-State Synthesis of Conjugated Nanoporous Polycarbazoles. ACS Macro Lett. 2017, 6 (10), 1056-1059, 10.1021/acsmacrolett.7b00480. 238.

Chen, Q.; Luo, M.; Hammershøj, P.; Zhou, D.; Han, Y.; Laursen, B. W.; Yan, C.-

G.; Han, B.-H., Microporous Polycarbazole with High Specific Surface Area for Gas Storage and Separation. J. Am. Chem. Soc. 2012, 134 (14), 6084-6087, 10.1021/ja300438w. 239.

Jin, T.; Xiong, Y.; Zhu, X.; Tian, Z.; Tao, D.-J.; Hu, J.; Jiang, D.-e.; Wang, H.;

Liu, H.; Dai, S., Rational design and synthesis of a porous, task-specific polycarbazole for efficient CO2 capture. Chem. Commun. 2016, 52 (24), 4454-4457, 10.1039/C6CC00573J.

79 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

240.

Page 80 of 92

Jiang, F.; Jin, T.; Zhu, X.; Tian, Z.; Do-Thanh, C.-L.; Hu, J.; Jiang, D.-e.; Wang,

H.; Liu, H.; Dai, S., Substitution Effect Guided Synthesis of Task-Specific Nanoporous Polycarbazoles with Enhanced Carbon Capture. Macromolecules 2016, 49 (15), 5325-5330, 10.1021/acs.macromol.6b01342. 241.

Zhu, X.; Hua, Y.; Tian, C.; Abney, C. W.; Zhang, P.; Jin, T.; Liu, G.; Browning,

K. L.; Sacci, R. L.; Veith, G. M.; Zhou, H.-C.; Jin, W.; Dai, S., Accelerating Membranebased CO2 Separation by Soluble Nanoporous Polymer Networks Produced by Mechanochemical Oxidative Coupling. Angew. Chem. Int. Ed. 2017, 57 (11), 2816-2821, 10.1002/anie.201710420. 242.

Troschke, E.; Grätz, S.; Lübken, T.; Borchardt, L., Mechanochemical Friedel–Crafts

Alkylation—A Sustainable Pathway Towards Porous Organic Polymers. Angew. Chem. 2017, 129 (24), 6963-6967, 10.1002/ange.201702303. 243.

Grätz, S.; Oltermann, M.; Troschke, E.; Paasch, S.; Krause, S.; Brunner, E.;

Borchardt, L., Solvent-free synthesis of a porous thiophene polymer by mechanochemical oxidative polymerization. J. Mater. Chem. A 2018, 10.1039/C8TA03684E. 244.

Crawford, D.; Casaban, J.; Haydon, R.; Giri, N.; McNally, T.; James, S. L.,

Synthesis by extrusion: continuous, large-scale preparation of MOFs using little or no solvent. Chem. Sci. 2015, 6 (3), 1645-1649, 10.1039/C4SC03217A. 245.

Laybourn, A.; López-Fernández, A. M.; Thomas-Hillman, I.; Katrib, J.; Lewis, W.;

Dodds, C.; Harvey, A. P.; Kingman, S. W., Combining continuous flow oscillatory baffled reactors and microwave heating: Process intensification and accelerated synthesis of metalorganic frameworks. Chem. Eng. J. 2019, 356, 170-177, 10.1016/j.cej.2018.09.011. 246.

Czaja, A. U.;

Trukhan, N.; Müller, U., Industrial applications of metal–organic

frameworks. Chem. Soc. Rev. 2009, 38 (5), 1284-1293, 10.1039/B804680H.

80 ACS Paragon Plus Environment

Page 81 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

247.

Martinez Joaristi, A.; Juan-Alcañiz, J.; Serra-Crespo, P.; Kapteijn, F.; Gascon, J.,

Electrochemical Synthesis of Some Archetypical Zn2+, Cu2+, and Al3+ Metal Organic Frameworks. Cryst Growth Des. 2012, 12 (7), 3489-3498, 10.1021/cg300552w. 248.

Campagnol, N.; Van Assche, T.; Boudewijns, T.; Denayer, J.; Binnemans, K.; De

Vos, D.; Fransaer, J., High pressure, high temperature electrochemical synthesis of metal– organic frameworks: films of MIL-100 (Fe) and HKUST-1 in different morphologies. J. Mater. Chem. A 2013, 1 (19), 5827-5830, 10.1039/C3TA10419B. 249.

Stassen, I.; Styles, M.; Van Assche, T.; Campagnol, N.; Fransaer, J.; Denayer, J.;

Tan, J.-C.; Falcaro, P.; De Vos, D.; Ameloot, R., Electrochemical Film Deposition of the Zirconium Metal–Organic Framework UiO-66 and Application in a Miniaturized Sorbent Trap. Chem. Mater. 2015, 27 (5), 1801-1807, 10.1021/cm504806p. 250.

Campagnol, N.; Van Assche, T. R. C.; Li, M.; Stappers, L.; Dincă, M.; Denayer, J.

F. M.; Binnemans, K.; De Vos, D. E.; Fransaer, J., On the electrochemical deposition of metal–organic

frameworks.

J.

Mater.

Chem.

A

2016,

4

(10),

3914-3925,

10.1039/C5TA10782B. 251.

Li, W.-J.; Tu, M.; Cao, R.; Fischer, R. A., Metal–organic framework thin films:

electrochemical fabrication techniques and corresponding applications & perspectives. J. Mater. Chem. A 2016, 4 (32), 12356-12369, 10.1039/C6TA02118B. 252.

Li, M.; Dincă, M., Reductive Electrosynthesis of Crystalline Metal–Organic

Frameworks. J. Am. Chem. Soc. 2011, 133 (33), 12926-12929, 10.1021/ja2041546. 253.

Gu, C.; Chen, Y.; Zhang, Z.; Xue, S.; Sun, S.; Zhang, K.; Zhong, C.; Zhang, H.;

Pan, Y.; Lv, Y.; Yang, Y.; Li, F.; Zhang, S.; Huang, F.; Ma, Y., Electrochemical Route to Fabricate Film-Like Conjugated Microporous Polymers and Application for Organic Electronics. Adv. Mater. 2013, 25 (25), 3443-3448, 10.1002/adma.201300839.

81 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

254.

Page 82 of 92

Palma-Cando, A.; Brunklaus, G.; Scherf, U., Thiophene-Based Microporous Polymer

Networks via Chemical or Electrochemical Oxidative Coupling. Macromolecules 2015, 48 (19), 6816-6824, 10.1021/acs.macromol.5b01821. 255.

Palma-Cando, A.; Preis, E.; Scherf, U., Silicon- or Carbon-Cored Multifunctional

Carbazolyl Monomers for the Electrochemical Generation of Microporous Polymer Films. Macromolecules 2016, 49 (21), 8041-8047, 10.1021/acs.macromol.6b02025. 256.

Oliver Kappe, C., Microwave dielectric heating in synthetic organic chemistry. Chem.

Soc. Rev. 2008, 37 (6), 1127-1139, 10.1039/B803001B. 257.

Ni, Z.; Masel, R. I., Rapid Production of Metal−Organic Frameworks via Microwave-

Assisted Solvothermal Synthesis. J. Am. Chem. Soc. 2006, 128 (38), 12394-12395, 10.1021/ja0635231. 258.

Jhung, S. H.; Lee, J. H.; Yoon, J. W.; Serre, C.; Férey, G.; Chang, J. S., Microwave

Synthesis of Chromium Terephthalate MIL-101 and Its Benzene Sorption Ability. Adv. Mater. 2006, 19 (1), 121-124, 10.1002/adma.200601604. 259.

Campbell, N. L.; Clowes, R.; Ritchie, L. K.; Cooper, A. I., Rapid Microwave

Synthesis and Purification of Porous Covalent Organic Frameworks. Chem. Mater. 2009, 21 (2), 204-206, 10.1021/cm802981m. 260. for

Majewski, M. B.; Noh, H.; Islamoglu, T.; Farha, O. K., NanoMOFs: little crystallites substantial

applications.

J.

Mater.

Chem.

A

2018,

6

(17),

7338-7350,

10.1039/C8TA02132E. 261.

Khan, N. A.; Hasan, Z.; Jhung, S. H., Beyond pristine metal-organic frameworks:

Preparation and application of nanostructured, nanosized, and analogous MOFs. Coord. Chem. Rev. 2018, 376, 20-45, 10.1016/j.ccr.2018.07.016. 262.

Khan, N. A.; Lee, J. S.; Jeon, J.; Jun, C.-H.; Jhung, S. H., Phase-selective synthesis

and phase-conversion of porous aluminum-benzenetricarboxylates with microwave

82 ACS Paragon Plus Environment

Page 83 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

irradiation.

Microporous

Mesoporous

Mater.

2012,

152,

235-239,

10.1016/j.micromeso.2011.11.025. 263.

Depauw, H.; Nevjestić, I.; De Winne, J.; Wang, G.; Haustraete, K.; Leus, K.;

Verberckmoes, A.; Detavernier, C.; Callens, F.; De Canck, E.; Vrielinck, H.; Van Der Voort, P., Microwave induced “egg yolk” structure in Cr/V-MIL-53. Chem. Commun. 2017, 53 (60), 8478-8481, 10.1039/C7CC04651K. 264.

Laybourn, A.; Katrib, J.; Ferrari-John, R. S.; Morris, C. G.; Yang, S.; Udoudo, O.;

Easun, T. L.; Dodds, C.; Champness, N. R.; Kingman, S. W.; Schröder, M., Metal–organic frameworks in seconds via selective microwave heating. J. Mater. Chem. A 2017, 5 (16), 7333-7338, 10.1039/C7TA01493G. 265.

Yoshida, J.-i.;

Takahashi, Y.; Nagaki, A., Flash chemistry: flow chemistry that

cannot be done in batch. Chem. Commun. 2013, 49 (85), 9896-9904, 10.1039/C3CC44709J. 266.

Nagaki, A.; Ichinari, D.; Yoshida, J.-i., Three-Component Coupling Based on Flash

Chemistry. Carbolithiation of Benzyne with Functionalized Aryllithiums Followed by Reactions with Electrophiles. J. Am. Chem. Soc. 2014, 136 (35), 12245-12248, 10.1021/ja5071762. 267.

Sauks, J. M.; Mallik, D.; Lawryshyn, Y.; Bender, T.; Organ, M., A Continuous-

Flow Microwave Reactor for Conducting High-Temperature and High-Pressure Chemical Reactions. Org. Process Res. Dev. 2014, 18 (11), 1310-1314, 10.1021/op400026g. 268.

Deadman, B. J.;

Browne, D. L.;

Baxendale, I. R.; Ley, S. V., Back Pressure

Regulation of Slurry-Forming Reactions in Continuous Flow. Chem. Eng. Technol. 2014, 38 (2), 259-264, 10.1002/ceat.201400445. 269.

Reinsch, H.; Waitschat, S.; Chavan, S. M.; Lillerud, K. P.; Stock, N., A Facile

“Green” Route for Scalable Batch Production and Continuous Synthesis of Zirconium MOFs. Eur. J. Inorg. Chem. 2016, 2016 (27), 4490-4498, 10.1002/ejic.201600295.

83 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 84 of 92

270. Avci-Camur, C.; Troyano, J.; Pérez-Carvajal, J.; Legrand, A.; Farrusseng, D.; Imaz, I.; Maspoch, D., Aqueous production of spherical Zr-MOF beads via continuous-flow spraydrying. Green Chem. 2018, 20 (4), 873-878, 10.1039/C7GC03132G. 271.

Faustini, M.; Kim, J.; Jeong, G.-Y.; Kim, J. Y.; Moon, H. R.; Ahn, W.-S.; Kim,

D.-P., Microfluidic Approach toward Continuous and Ultrafast Synthesis of Metal–Organic Framework Crystals and Hetero Structures in Confined Microdroplets. J. Am. Chem. Soc. 2013, 135 (39), 14619-14626, 10.1021/ja4039642. 272.

Rubio-Martinez, M.; Batten, M. P.; Polyzos, A.; Carey, K.-C.; Mardel, J. I.; Lim,

K.-S.; Hill, M. R., Versatile, High Quality and Scalable Continuous Flow Production of Metal-Organic Frameworks. Sci. Rep. 2014, 4, 5443, 10.1038/srep05443. 273.

Jeong, G.-Y.; Ricco, R.; Liang, K.; Ludwig, J.; Kim, J.-O.; Falcaro, P.; Kim, D.-P.,

Bioactive MIL-88A Framework Hollow Spheres via Interfacial Reaction In-Droplet Microfluidics for Enzyme and Nanoparticle Encapsulation. Chem. Mater. 2015, 27 (23), 7903-7909, 10.1021/acs.chemmater.5b02847. 274.

Singh, V.; Jang, S.; Vishwakarma, N. K.; Kim, D.-P., Intensified synthesis and post-

synthetic modification of covalent organic frameworks using a continuous flow of microdroplets technique. NPG Asia Mater. 2018, 10, e456, 10.1038/am.2017.209. 275.

Ameloot, R.; Vermoortele, F.; Vanhove, W.; Roeffaers, M. B. J.; Sels, B. F.; De

Vos, D. E., Interfacial synthesis of hollow metal–organic framework capsules demonstrating selective permeability. Nat Chem. 2011, 3, 382, 10.1038/nchem.1026. 276.

Paseta, L.;

Seoane, B.;

Julve, D.;

Sebastián, V.;

Téllez, C.; Coronas, J.,

Accelerating the Controlled Synthesis of Metal–Organic Frameworks by a Microfluidic Approach: A Nanoliter Continuous Reactor. ACS Appl. Mater. Interfaces 2013, 5 (19), 94059410, 10.1021/am4029872.

84 ACS Paragon Plus Environment

Page 85 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

277.

Wang, Y.; Li, L.; Dai, P.; Yan, L.; Cao, L.; Gu, X.; Zhao, X., Missing-node

directed synthesis of hierarchical pores on a zirconium metal–organic framework with tunable porosity and enhanced surface acidity via a microdroplet flow reaction. J. Mater. Chem. A 2017, 5 (42), 22372-22379, 10.1039/C7TA06060B. 278.

Wang, Y.; Li, L.; Yan, L.; Gu, X.; Dai, P.; Liu, D.; Bell, J. G.; Zhao, G.; Zhao,

X.; Thomas, K. M., Bottom-Up Fabrication of Ultrathin 2D Zr Metal–Organic Framework Nanosheets through a Facile Continuous Microdroplet Flow Reaction. Chem. Mater. 2018, 30 (9), 3048-3059, 10.1021/acs.chemmater.8b00765. 279.

Peng, Y.; Wong, W. K.; Hu, Z.; Cheng, Y.; Yuan, D.; Khan, S. A.; Zhao, D.,

Room Temperature Batch and Continuous Flow Synthesis of Water-Stable Covalent Organic Frameworks

(COFs).

Chem.

Mater.

2016,

28

(14),

5095-5101,

10.1021/acs.chemmater.6b01954. 280.

Rodríguez-San-Miguel, D.; Abrishamkar, A.; Navarro, J. A. R.; Rodriguez-Trujillo,

R.; Amabilino, D. B.; Mas-Ballesté, R.; Zamora, F.; Puigmartí-Luis, J., Crystalline fibres of a covalent organic framework through bottom-up microfluidic synthesis. Chem. Commun. 2016, 52 (59), 9212-9215, 10.1039/C6CC04013F. 281.

Gimeno-Fabra, M.; Munn, A. S.; Stevens, L. A.; Drage, T. C.; Grant, D. M.;

Kashtiban, R. J.; Sloan, J.; Lester, E.; Walton, R. I., Instant MOFs: continuous synthesis of metal–organic frameworks by rapid solvent mixing. Chem. Commun. 2012, 48 (86), 1064210644, 10.1039/C2CC34493A. 282.

Bayliss, P. A.; Ibarra, I. A.; Pérez, E.; Yang, S.; Tang, C. C.; Poliakoff, M.;

Schröder, M., Synthesis of metal–organic frameworks by continuous flow. Green Chem. 2014, 16 (8), 3796-3802, 10.1039/C4GC00313F.

85 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

283.

Page 86 of 92

Taddei, M.; Steitz, D. A.; van Bokhoven, J. A.; Ranocchiari, M., Continuous-Flow

Microwave Synthesis of Metal–Organic Frameworks: A Highly Efficient Method for LargeScale Production. Chem. Eur. J. 2016, 22 (10), 3245-3249, 10.1002/chem.201505139. 284.

Dunne, P. W.; Lester, E.; Walton, R. I., Towards scalable and controlled synthesis of

metal–organic framework materials using continuous flow reactors. React. Chem. Eng. 2016, 1 (4), 352-360, 10.1039/C6RE00107F. 285.

Carné-Sánchez, A.;

Imaz, I.;

Cano-Sarabia, M.; Maspoch, D., A spray-drying

strategy for synthesis of nanoscale metal–organic frameworks and their assembly into hollow superstructures. Nat Chem. 2013, 5, 203, 10.1038/nchem.1569. 286.

Guillerm, V.; Garzón-Tovar, L.; Yazdi, A.; Imaz, I.; Juanhuix, J.; Maspoch, D.,

Continuous One-Step Synthesis of Porous M-XF6-Based Metal-Organic and HydrogenBonded Frameworks. Chem. Eur. J. 2017, 23 (28), 6829-6835, 10.1002/chem.201605507. 287.

Garcia Marquez, A.; Horcajada, P.; Grosso, D.; Ferey, G.; Serre, C.; Sanchez, C.;

Boissiere, C., Green scalable aerosol synthesis of porous metal–organic frameworks. Chem. Commun. 2013, 49 (37), 3848-3850, 10.1039/C3CC39191D. 288.

Garzón-Tovar, L.; Pérez-Carvajal, J.; Imaz, I.; Maspoch, D., Composite Salt in

Porous Metal-Organic Frameworks for Adsorption Heat Transformation. Adv. Funct. Mater. 2017, 27 (21), 1606424, 10.1002/adfm.201606424. 289.

Garzón-Tovar, L.; Cano-Sarabia, M.; Carné-Sánchez, A.; Carbonell, C.; Imaz, I.;

Maspoch, D., A spray-drying continuous-flow method for simultaneous synthesis and shaping of microspherical high nuclearity MOF beads. React. Chem. Eng. 2016, 1 (5), 533539, 10.1039/C6RE00065G. 290.

Garzón-Tovar, L.; Rodríguez-Hermida, S.; Imaz, I.; Maspoch, D., Spray Drying for

Making Covalent Chemistry: Postsynthetic Modification of Metal–Organic Frameworks. J. Am. Chem. Soc. 2017, 139 (2), 897-903, 10.1021/jacs.6b11240.

86 ACS Paragon Plus Environment

Page 87 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

291.

Ma, J.; Kalenak, A. P.; Wong-Foy, A. G.; Matzger, A. J., Rapid Guest Exchange and

Ultra-Low Surface Tension Solvents Optimize Metal–Organic Framework Activation. Angew. Chem. Int. Ed. 2017, 56 (46), 14618-14621, 10.1002/anie.201709187. 292.

Sun, L.; Campbell, M. G.; Dincă, M., Electrically Conductive Porous Metal–Organic

Frameworks. Angew. Chem. Int. Ed. 2016, 55 (11), 3566-3579, 10.1002/anie.201506219. 293.

Medina, D. D.; Sick, T.; Bein, T., Photoactive and Conducting Covalent Organic

Frameworks. Adv. Energy Mater. 2017, 7 (16), 1700387, 10.1002/aenm.201700387. 294.

Liu, J.; Wöll, C., Surface-supported metal–organic framework thin films: fabrication

methods, applications, and challenges. Chem. Soc. Rev. 2017, 46 (19), 5730-5770, 10.1039/C7CS00315C. 295.

Stassen, I.; Burtch, N.; Talin, A.; Falcaro, P.; Allendorf, M.; Ameloot, R., An

updated roadmap for the integration of metal–organic frameworks with electronic devices and chemical sensors. Chem. Soc. Rev. 2017, 46 (11), 3185-3241, 10.1039/C7CS00122C. 296.

Liu, J.; Lukose, B.; Shekhah, O.; Arslan, H. K.; Weidler, P.; Gliemann, H.; Bräse,

S.; Grosjean, S.; Godt, A.; Feng, X.; Müllen, K.; Magdau, I.-B.; Heine, T.; Wöll, C., A novel series of isoreticular metal organic frameworks: realizing metastable structures by liquid phase epitaxy. Sci. Rep. 2012, 2, 921, 10.1038/srep00921. 297.

Liu, J.; Zhou, W.; Liu, J.; Howard, I.; Kilibarda, G.; Schlabach, S.; Coupry, D.;

Addicoat, M.; Yoneda, S.; Tsutsui, Y.; Sakurai, T.; Seki, S.; Wang, Z.; Lindemann, P.; Redel, E.; Heine, T.; Wöll, C., Photoinduced Charge-Carrier Generation in Epitaxial MOF Thin Films: High Efficiency as a Result of an Indirect Electronic Band Gap? Angew. Chem. Int. Ed. 2015, 54 (25), 7441-7445, 10.1002/anie.201501862. 298. Goswami, S.; Ma, L.; Martinson, A. B. F.; Wasielewski, M. R.; Farha, O. K.; Hupp, J. T., Toward Metal–Organic Framework-Based Solar Cells: Enhancing Directional Exciton

87 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 88 of 92

Transport by Collapsing Three-Dimensional Film Structures. ACS Appl. Mater. Interfaces 2016, 8 (45), 30863-30870, 10.1021/acsami.6b08552. 299.

Dogru, M.; Handloser, M.; Auras, F.; Kunz, T.; Medina, D.; Hartschuh, A.;

Knochel, P.; Bein, T., A Photoconductive Thienothiophene-Based Covalent Organic Framework Showing Charge Transfer Towards Included Fullerene. Angew. Chem. Int. Ed. 2013, 52 (10), 2920-2924, 10.1002/anie.201208514. 300.

Calik, M.; Auras, F.; Salonen, L. M.; Bader, K.; Grill, I.; Handloser, M.; Medina,

D. D.; Dogru, M.; Löbermann, F.; Trauner, D.; Hartschuh, A.; Bein, T., Extraction of Photogenerated Electrons and Holes from a Covalent Organic Framework Integrated Heterojunction. J. Am. Chem. Soc. 2014, 136 (51), 17802-17807, 10.1021/ja509551m. 301.

Van Vleet, M. J.; Weng, T.; Li, X.; Schmidt, J. R., In Situ, Time-Resolved, and

Mechanistic Studies of Metal–Organic Framework Nucleation and Growth. Chem. Rev. 2018, 118 (7), 3681-3721, 10.1021/acs.chemrev.7b00582. 302.

Taddei, M.; Casati, N.; Steitz, D. A.; Dümbgen, K. C.; van Bokhoven, J. A.;

Ranocchiari, M., In situ high-resolution powder X-ray diffraction study of UiO-66 under synthesis conditions in a continuous-flow microwave reactor. CrystEngComm 2017, 19 (23), 3206-3214, 10.1039/C7CE00867H. 303.

Zhu, G.; Liu, Y.; Flores, L.; Lee, Z. R.; Jones, C. W.; Dixon, D. A.; Sholl, D. S.;

Lively, R. P., Formation Mechanisms and Defect Engineering of Imine-Based Porous Organic Cages. Chem. Mater. 2018, 30 (1), 262-272, 10.1021/acs.chemmater.7b04323. 304.

Liu, G.; Di Yuan, Y.; Wang, J.; Cheng, Y.; Peh, S. B.; Wang, Y.; Qian, Y.; Dong,

J.; Yuan, D.; Zhao, D., Process-Tracing Study on the Postassembly Modification of Highly Stable Zirconium Metal–Organic Cages. J. Am. Chem. Soc. 2018, 140 (20), 6231-6234, 10.1021/jacs.8b03517.

88 ACS Paragon Plus Environment

Page 89 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

305.

Boissonnault, J. A.; Wong-Foy, A. G.; Matzger, A. J., Core–Shell Structures Arise

Naturally During Ligand Exchange in Metal–Organic Frameworks. J. Am. Chem. Soc. 2017, 139 (42), 14841-14844, 10.1021/jacs.7b08349. 306.

Taddei, M.; Wakeham, R. J.; Koutsianos, A.; Andreoli, E.; Barron, A. R., Post-

Synthetic Ligand Exchange in Zirconium-Based Metal–Organic Frameworks: Beware of The Defects! Angew. Chem. Int. Ed. 2018, 57 (36), 11706-11710, 10.1002/anie.201806910.

89 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 90 of 92

Insert Table of Contents artwork and synopsis here

Strategies for the sustainable scale-up of advanced porous materials with highly tunable compositions and functionalities are summarized and discussed.

90 ACS Paragon Plus Environment

Page 91 of 92 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Author biography

Shing Bo Peh obtained his B.S. degree in Chemical Engineering from National University of Singapore. He is currently a Ph.D. candidate of Chemical Engineering at National University of Singapore under the supervision of Dr. Dan Zhao, working on the scaled-up synthesis of porous materials.

Yuxiang Wang obtained his Bachelor of Engineering in Polymer Materials and Engineering from Zhejiang University in 2015. He is currently a Ph.D. candidate of Chemical Engineering at National University of Singapore under the supervision of Dr. Dan Zhao, focusing on the design and synthesis of microporous metal-organic frameworks for gas separation.

Dan Zhao obtained his Ph.D. degree in Inorganic Chemistry under the supervision of Prof. Hong-Cai Joe Zhou at Texas A&M University in 2010. After finishing his postdoctoral 91 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 92 of 92

training at Argonne National Laboratory, he joined the Department of Chemical & Biomolecular Engineering at National University of Singapore in July 2012 as an Assistant Professor, and was promoted to Associate Professor with tenure in July 2018. His research interests include advanced porous materials and membranes with the applications in clean energy and environmental sustainability.

92 ACS Paragon Plus Environment