Metal-Organic Frameworks (MOFs) Derived Effective Solid Catalysts

Sep 28, 2018 - Over the past decade, the development in the valorization of biomass technologies keeps increasing because the biomass utilization for ...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

Perspective

Metal-Organic Frameworks (MOFs) Derived Effective Solid Catalysts for the Valorization of Lignocellulosic Biomass Yu-Te Liao, Babasaheb Mansub Matsagar, and Kevin C.-W. Wu ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b03683 • Publication Date (Web): 28 Sep 2018 Downloaded from http://pubs.acs.org on September 29, 2018

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 48 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

Metal-Organic Frameworks (MOFs) Derived Effective Solid Catalysts for the Valorization Lignocellulosic Biomass Yu-Te Liao,1‡ Babasaheb M. Matsagar1‡ and Kevin C.-W. Wu1-3* 1

Department of Chemical Engineering, National Taiwan University, No. 1, Sec. 4, Roosevelt Rd., Taipei 10617, Taiwan

2

Center of Atomic Initiative for New Materials (AI-MAT), National Taiwan University, No. 1, Sec. 4, Roosevelt Rd., Taipei 10617, Taiwan 3

International Graduate Program of Molecular Science and Technology, National Taiwan University (NTU-MST), No. 1, Sec. 4, Roosevelt Rd., Taipei 10617, Taiwan Corresponding Author Email: [email protected]

ACS Paragon Plus Environment

1

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 48

Abstract

Over the past decade, the development in the valorization of biomass technologies keeps increasing because the biomass utilization for the manufacturing fine chemical and fuels has diverse advantages over fossil feedstock. The review focuses on the utilization of metal-organic framework-derived (MOF-derived) materials as effective solid catalysts for the valorization of biomass into platform chemicals. MOFs compose of abundant organic ligands and metal cluster, and additional functional groups could be modified on ligands (or metal clusters), serving as active sites. On the other hand, MOFs could also be converted into porous carbons or metal oxide composites by calcination at nitrogen or air, respectively, for catalytic reactions. These MOF-derived catalysts feature the advantages like high specific surface area, porosity and active sites from mother MOFs. More importantly, stronger interactions between guests (i.e. metal or alloy NPs) and hosts (i.e. MOF-derived carbon or metal oxides) make these catalysts more efficient than conventional catalysts where guests are deposited on hosts by impregnation. We summarize the studies of lignocellulosic biomass conversion including (1) dehydration of sugars such as glucose, fructose, and xylose into furans, (2) hydrogenation of furans into fine chemicals and (3) sugars into sugar alcohols using MOF-derived catalysts. The challenges and prospective of MOF-derived materials applied in biomass conversion are also described.

KEYWORDS: metal organic frameworks • biomass • heterogeneous catalyst • porous material

ACS Paragon Plus Environment

2

Page 3 of 48 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 Lignocellulosic biomass The fossil feedstocks which are coal, natural gas, and crude oil are used to satisfy our energy requirements and to produce the variety of chemicals that enrich our lives. The difficulties with the large-scale utilization of fossil feedstock are the uneven geographical distribution of reserves, declining availability, and global warming. To avoid these problems with fossil feedstocks currently, lignocellulosic biomass is extensively used all over the world for chemical synthesis and energy applications. It has been suggested that biomass is the only reliable alternative source for replacing fossil feedstock for the synthesis chemicals and fuels by promoting environmentally benign pathways.1-2 Because biomass is renewable, abundant, and readily available, it can make carbon-neutral process upon utilization, equally distributed and has a lower impact on the environment. Hence, it becomes a natural choice for researchers to use biomass as an alternate resource for fossil feedstocks. The lignocellulosic biomass is the most abundant form of biomass present with an annual production of ca. 170 billion metric tons.2 As shown in Fig. 1, it is comprised of lignin (15-30%) cellulose (30-45%) and hemicellulose (20-35%). The lignin is an aromatic polymer made up of methoxylated phenylpropane units,3 while the cellulose and hemicellulose are polysaccharides. Cellulose is homo-polysaccharide made up of β-D-glucose units. The hemicellulose can be homo- or hetero-polysaccharide which are made up of pentose sugar, hexose sugars and sugar acids linked together by β-(1→4), β-(1→3) and β-(1→6) glycosidic linkages. Hemicellulose is classified as xylan, mannan, xyloglucan, galactan, and arabinan depending on the cell-wall polysaccharide types.

ACS Paragon Plus Environment

3

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 48

Figure 1. Composition of lignocellulosic biomass. As hemicellulose is made up of various sugars such as glucose, xylose, mannose, galactose, arabinose and rhamnose respectively. The recent studies demonstrate that hemicellulose present in this raw biomass could be efficiently converted into C5 sugar monomers.4-5 The xylan type of hemicellulose is most common which is basically made up of xylose sugar. As it is shown in Scheme 1 the hydrolysis of xylan produces xylose and arabinose as major C5 sugars which could be further converted into wide variety of value-added chemicals, including xylitol and furfural. For instance, furfural has many applications in several industries such as Lube oil refining,6 synthesis of phenolic resins and pharmaceuticals respectively,7-8 and the derivatives, furfuryl alcohol (FOL) and methyl furan (MF) could be produced by hydrogenation of furfural (FAL) which further finds potential application as solvent and fuel.6 Extensive literature is available for the conversion of isolated hemicellulose into C5 sugars and furfural using solid acid, mineral acid and acidic ionic liquid catalysts.2, 9-10

ACS Paragon Plus Environment

4

Page 5 of 48 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

Scheme 1. Conversion of hemicellulose into diverse chemicals. Cellulose, on the other hand, is a homopolymer composed of glucose with a higher degree of polymerization (DP) = 100-2000 compared to hemicellulose (DP = 100-200).11 Cellulose is difficult to be hydrolyzed compared to hemicellulose and other saccharides because of the presence of inter- and intra-molecular hydrogen bonding between various anhydro-glucan units.12 This extensive hydrogen bonding of cellulose makes it partly crystalline and robust material towards chemical reactivity. As presented in Scheme 2, various potentially important chemicals can be generated from cellulose. Cellulose is decomposed into glucose by hydrolysis.13 Further dehydration of glucose yields 5-hydroxymethylfurfural (HMF) which is a platform chemical used for the synthesis of verity of industrially significant chemicals such hydrogenation into 2,5-dimethylfuran (DMF) via bis(hydroxymethyl)furfural (BHMF) intermediate,14-15 and oxidation into 2,5-furandicarboxylic acid (FDCA) via furan-2,5-

ACS Paragon Plus Environment

5

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 48

diformylfuran (DFF), 5-hydroxymethyl-2- furancarboxylic acid (HMFCA) and 5-formyl-2furancarboxylic acid (FFCA) intermediates.16 Furthermore, various significant chemicals such as γ-valerolactone (γ-GVL), levulinic acid (LA), 1,6-hexanediols can be manufactured from HMF, too.17-18

HO

+H 2

HO

OH

1,6-hexanediol

OH

HO O

O

O OH

HO

+H 2

OH

O DMF

n

Cellulose

OH O

OH OH Glucose

+H

HO O

BHMF

O OH

2O

OH

OH

O

+H 2 OH OH HO

OH

HO

+H 2

+H 2

OH

+H 2

OH O BHMTHF

O DMTHF

-H2O

Re

HO

Condensation O

O

OH

Eth

HMF

Fructose +O 2 O

FA

LA

n tio dr a y h

O

n Alkane

erifi cati on

+O 2

O

RO O

O

HO

O

O

DFF

HMFCA

+O 2

CO 2H

AMF

+O 2 O

HO 2C

O

CO 2H

+O 2

FDCA

O

CO2H

FFCA

Scheme 2. Conversion of cellulose into various promising chemicals. Many catalysts have been reported for the biomass conversion to value-added chemicals such as mineral acids (HCl, H2SO4),19-20 organic acid (p-toluenesulfonic acid, maleic acid),6, 21 solid acid catalysts (SAPO, HUSY, sulfonated carbon),22-24 Zeolites (ZSM-5, HMOR),24-25 ion exchange resins (Amberlyst-15, Nafion),26 heteropoly acids (heteropoly tungstate, MOF-based heteropoly acid),27-28 metal catalyst (Ru/C, Pt/C, Pd/C),29 ionic liquids ([BMIM][Cl], [C3SO3HMIM][HSO4]) respectively.6, 10, 30

ACS Paragon Plus Environment

6

Page 7 of 48 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

Metal-organic frameworks (MOFs) and their derived nanomaterial Metal organic frameworks (MOFs), also known as porous coordination polymers (PCPs), are a series of porous structure composed of metal ions/clusters (joint) and organic ligands (linker).31 The metal ions/clusters coordinate with multipodal organic ligands, forming secondary building units (SBUs). SBUs are arranged into a porous framework with particular morphology by sharing podals with each other. Most of transition metals, post-transition metals and lanthanides have been used as joints and coordinated with different linkers such as bi, tricarboxylate contained benzene ring and N-contained cycloalkane composites.32 The diverse combination of joints and linkers makes MOFs with different texture properties and morphology. Until now, more than 20,000 kinds of MOFs have been either synthesized or design by simulation (e.g. ZIF-8,33 HKUST-1,34 MIL-53,35 UiO-66,36 and PCN-22237).38 MOFs have the features like large Brunauer-Emmett-Teller (BET) specific surface area, large pore volume, abundant Lewis/Brønsted acid/base sites, which make MOFs as a better material for catalysis, separation, storage, and sensing applications by either directly using MOFs or combination of MOFs with polymer,39 metal oxide,40 graphene,41 porous carbon,42 cell43 and virus.44 The diversity of MOFs materials makes them potentially useful catalysts with various functionalities. Besides directly using the active sites in the framework of MOFs, functionalization of MOFs with other ligands or the thermal treatment of MOFs to make porous carbon or metal oxides are two types of popular MOF-derived catalysts (Fig. 2). The ligands on MOFs are either modified with functional group or replaced with a functional group-containing ligand. The functional group on ligands can change the microenvironment around the SBUs which has apparent effect on the catalytic reaction. The MOFs are inherent porous structure

ACS Paragon Plus Environment

7

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 48

where the organic ligands and SBUs are placed along the edge of pores which could be regarded as a self-template.45 MOFs could be converted into porous carbons (PCs), metal on PCs and porous metal oxides (PMOs) by annulling under inert or oxygen-rich environment. The MOFderived materials feature high specific surface area, porosity, and abundant active sites from mother MOFs. Moreover, by carefully selecting ligands and metal ions, the texture properties of MOF-derived materials could be customized for different applications. Till now, MOF-derived materials have been applied for catalysis,46 separation,47 energy storage48-49 and so on owing to the advantages mentioned above. To the best of our knowledge, so far MOF-derived materials have not been used as catalysts for hemicellulose conversion, while recent studies show the applications of MOF-derived materials for the xylose dehydration, FAL hydrogenation into cyclopentanone (CPO) and FOL.50-52 In the case of lignocellulosic biomass conversion, MOFderived materials have been reported as an efficient catalyst for hydrolysis of cellulose to glucose, isomerization of glucose, dehydration of glucose and fructose, hydrogenation of HMF, oxidation of HMF, conversion of fructose into DFF, etherification of HMF, etc.53-57

ACS Paragon Plus Environment

8

Page 9 of 48 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 2. Synthesis of MOF-derived materials via functionalization and thermal treatment. The conversion of lignocellulose biomass into value-added chemicals and biofuels are widely studied recently, and we attempt to provide a broad overview of lignocellulose biomass conversion using MOF-derived materials as catalysts. An overview of recent fabrication of MOF-derived materials including functionalized MOF, porous carbons, metal composites on porous carbons and metal oxides are provided, followed by case study of reaction including dehydration of sugars, hydrogenation of HMF and FAL, oxidation of HMF, etc. This review focuses on the efficient and economically viable processes and the modifications in the MOFderived materials for the optimization of efficient biomass conversion. Various properties of MOF-derived materials are explained in detail and advantages of these materials for biomass conversion are reviewed. Although there are few reports on biomass conversion into value-added chemicals using MOF-derived materials because of the unexplored applications of MOF-derived

ACS Paragon Plus Environment

9

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 48

materials for biomass conversion,58 this review would help readers find novel applications of MOF-derived materials and usability for biomass conversion as a catalyst for developing new economical, selective, environmentally friendly, and sustainable technology. Finally, the challenges and future trends and prospects associated with the development of MOF-derived materials are examined. This review will be a beneficial source for guiding industrial and academic researchers working in the biomass and MOF research areas. Also, it will help readers to understand the proficiency of MOF-derived materials over conventional heterogeneous catalysts (e.g. zeolites) and homogeneous catalysts (e.g. mineral acids) when using MOF-derived materials in the field of biomass conversion. MOF-derived catalysts In this section, many MOF-derived catalysts will be introduced. These catalysts could be briefly divided into two groups: a) MOF with a functional group, b) metal nanoparticles in MOFderived carbon or in MOF-derived metal oxides. The synthesis of catalyst will be introduced and the advantages of MOF-derived catalyst on application will be discussed in this section. MOF-derived catalysts via modification of functional groups Terephthalic acid is one of popular organic ligands used in the synthesis of MOFs. A functional group containing terephthalic acid could be used to synthesize the isoreticular MOF (IRMOF) with same SBUs due to their similar topology while an additional function is given on MOF framework.59 For example, zinc oxide clusters coordinated with terephthalic acid is IRMOF-1 (or MOF-5), while zinc oxide clusters coordinated with aminoterephthalic acid is IRMOF-3. The additional amino group could serve as Lewis base in the catalytic reaction. Cmarik et al.60 synthesized a series of isoreticular UiO-66, including –H, –NH2, –NO2, –1,4Naphthyl and –2,5-(OMe)2, and showed the effect of ligand on adsorption properties. The result

ACS Paragon Plus Environment

10

Page 11 of 48 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

showed that the functional groups change the adsorption heat and the water affinity, which leads to different ability toward gas separation. Zhou and co-workers synthesized PCN-125 with various functionalized ligands (Fig. 3a).61 Different from Cmarik’s work, Zhou’s group modified terphenyltetracarboxylates (TPTCs, ligands of PCN-125) by breaking a TPTCs into functional group-containing isophthalate (R-isoph, R represents methyl, amino, nitro, sulfo, etc). The Risoph not only provide additional function for PCN-125 but also enlarge the pore size of PCN125. The pore size was enlarged from 1 nm to 1.1 as 50% of H-isoph was added into PCN-125growth solution. As comparison, two functional groups are modified on a TPTCs (2R-TPTC), too. The R-isoph-contianing PCN-125 showed higher CO2 uptake with lower heat of adsorption than 2R-TPTC-containing PCN-125, demonstrating that functional groups change the adsorption while the enlarged pore size is helpful for adsorption, too. Hupp and co-workers synthesized a squaramide moiety-containing biphenyl-4,4’dicarboxylate (squar-bpdc) and fabricated squar-UiO-67 as a hydrogen-bond-donating organocatalyst (Fig. 3b).62 The NMR spectrum showed the ratio of squar-bpdc to bpdc was one on the UiO-67. The free squaramide-based catalyst is easily deactivated due to combination of detrimental intermolecular hydrogen bonding of the squaramide moieties, while squar-UiO-67 showed good catalytic activity even after several cycles. They demonstrated that the modification of functional moiety on ligand of MOF could prevent the deactivation of functional moiety since the ligands were separated by SBUs and well align in a 3D framework. However, part of functional group-containing ligands could not be used for synthesis of ISMOFs due to thermodynamically unfavorable characteristics. An alternative way to fabricate these kind of thermodynamically unfavorable ligand is using solvent-assisted linker exchange (SALE) where the ligands in as-synthesized MOF are

ACS Paragon Plus Environment

11

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 48

replaced by functional group-containing ligands in a favorable condition.63 For example, Kim et al.64 successfully fabricated a dual-functional groups containing UiO-66 (i.e. –Br and –NH2) by SALE. Takaishi et al.65 accessed new versions of robust porphyrinic MOF (RPM) by changing the pillaring linkers with SALE. RPM composes of two types of porphyrins whose metal centers need not match. It looks like a chance to synthesize a series of dual-metallic RPM for multi-stage chemical catalysis; however, the variety of metal ions using conventional procedure is limited due to the thermodynamically unfavorable characteristic. A ZnZn-RPM was synthesized with conventional procedure. A metal-free or metal-chelated ligand were used to replace the second Zn-ligand with SALE. For example, ZnAl-RPM fabricated by SALE showed catalytic activity on ring-opening reaction which could be achieved by ZnM-RPM synthesized with conventional procedure (Fig. 3c).

ACS Paragon Plus Environment

12

Page 13 of 48 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 3. MOF-derived materials by modification with a functional moiety. (a) Coassembly of ligand and ligand-fragment into a PCN-125. Reprinted with permission from Ref.61 Copyright 2012 American Chemical Society. (b) Synthesis of ISMOF with a functional group on ligand as a catalyst. Reprinted with permission from Ref.62 Copyright 2015 American Chemical Society.

ACS Paragon Plus Environment

13

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 48

(c) Modification with a functional metal by SALE as a catalyst. Reprinted with permission from Ref.65 Copyright 2013 Royal Society of Chemistry. The functional group could also be grafted on ligands or SBUs of MOF to promote the catalytic activity. Canivet et al.66 synthesized carboxylimidazolate-containing zeolitic imidazolate frameworks (ZIFs) and grafted dodecylamine on ligands. A hydrophobic shell around ZIFs was formed, which apparently prevents ZIFs from water adsorption. A better catalytic activity of Knoevenagel was achieved by grafted a hydrophobic group on the ligand of SIM-1 (isostructural ZIF-8, C10H10N4O2Zn). Goesten et al.67 also directly functionalized sulfate groups on ligands of MIL-53. Based on the assumption that the ligand was mono-sulfated, 50% of ligands in a MIL-53 became sulfated terephthalate. This procedure was processed on MIL101, showing that 20% of ligands were successfully sulfated, too. The sulfate groups on ligand showed good thermal stability and catalytic activity on esterification of acetic acid with butanol. MIL-101-H2SO4 showed a comparable turnover frequency on esterification to sulfuric acid while it also demonstrated the recyclability which could not be achieved by homogeneous mineral acid catalyst such as sulfuric acid. Besides serving as active sites, the functional group-containing ligands could also serve as metal-binding sites for metalation. An isolated single atoms-coordinated MOF was fabricated owing to the isolated ligands homogeneously aligned in MOF.68 Cohen’s group performed a series of metalation on MOF by modifying catechol groups on ligands of MOF and stabilizing organometallic ions.69-70 A robust UiO-66 was selected as the template and a 2,3dihydroxyterephthalic acid (or a 2,3-dimercaptoterephthalic acid) was used to replace the part of terephthalic acid in UiO-66 by SALE. The NMR spectrum showed 75% of terephthalic acid was replaced by 2,3-dihydroxyterephthalic acid while 40% of that was replace by 2,3-

ACS Paragon Plus Environment

14

Page 15 of 48 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

dimercaptoterephthalic acid. Both noble metal and transition metal were chelated with catechol (or thiocatechol), showing an isolated metallic atom on every functionalized ligand by X-ray absorption spectrum and electron spin resonance spectrum. The resulting catalysts demonstrated activities on regioselective functionalization of sp2 C–H bond and secondary alcohol oxidation on various reactants. Besides the outstanding activity, the catalyst also showed the stability on recycle experiment which might be difficult to achieve by homogeneous catalysts. MOF-derived catalysts via thermal treatment The MOFs are assembled by coordination between SBUs and ligands while the decoordination between SBUs and ligands processes in an acidic environment which makes MOFs unstable during the catalytic reaction. Several kinds of MOFs are proposed to be stable in acidic environments, such as MIL-101, UiO-66, PCN-222, etc.; however, it remains several thousand types of MOF which is not stable in acidic environment and could not be applied for chemical catalysis.71 The thermal treatment is used to convert MOFs into diverse types of porous materials, including porous carbons (PCs) and porous metal oxides (PMOs) which are more stable than their mother MOFs. In addition, if the second (or third) metal ions were added upon the synthesis, after thermal treatment it would result in metal nanoparticles-embedded PC or PMOs. The synthesis of PCs has been attracted much attention owing to the advantages of large specific surface area, conductivity, thermal stability and chemical stability.72-73 PCs are typically synthesized from the direct pyrolysis of carbon-rich composites (e.g. biomass and polymer) followed by chemical or physical activation.74-76 The organic ligands in MOFs are separated and aligned regularly by SBUs. Besides, they can be regarded as the resource of carbon.

ACS Paragon Plus Environment

15

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 48

Consequently, direct pyrolysis of MOFs followed by removal of SBUs provides an efficient way for synthesis of PCs.47, 77-80 Xu’s group prepared MOF-derived PCs by pyrolysis of polymerized-furfuryl alcoholstuffed MOF-5, and demonstrated a good performance on H2 storage.81 MOF-5 is composed of zinc oxide as SBUs and terephthalic acid as ligands. After pyrolysis under Ar flow at 1000 °C, the as-synthesized PCs showed amorphous carbon without existence of Zn element by X-ray diffractometer. The SBUs of MOF-5 (Zn4O) were reduced into Zn metal at 800 °C followed by evaporation of Zn metal. The MOF-5-derived PCs exhibits a high BET specific surface area of 2872 m2/g and a large pore volume of 2.06 cm3/g with hierarchical porous structure which were comparable to theoretical surface area of double-sided separated graphitic sheet (2965 m2/g).82 Xu’s group further used ZIF-8 as template and stuffed with polymerized-furfuryl alcohol followed by pyrolysis under Ar flow at 1000 °C.83 The furfuryl alcohol-stuffed ZIF-8-derived PCs showed high percentage of H2-storage (2.77%) than most of carbon based materials including activated carbons and MOF-5 derived carbons owing to the nature property of the ligands used for synthesizing ZIFs.83 The higher percentage of C atom in organic ligand (2MIM) enhances the H2 storage of PCs. ZIF-8 is one of the eye-catching branches in MOFs for fabricating PCs because its organic ligand contains only C and N. One 2-methylimidazole ligand contains 34 wt% of N which is higher than that of 2,5-diaminoterephthalic acid (14 wt%). Direct carbonization of ZIF8 into nitrogen-containing PCs (denoted as NPCs) has been reported widely as a good material for electrocatalysis and chemical catalysis.46-47, 84 The nitrogen atoms on ZIF-8 were converted into pyridinic-N, pyrrolic-N or graphitic-N (Fig. 4a).85 The pyridinic-N on NPCs serves as electrochemically active sites for enhancing the capacitive behaviors while graphitic-N serves as

ACS Paragon Plus Environment

16

Page 17 of 48 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

active sites for aerobic oxidation by formation of oxygen radicals on N-O2 adducts.86 In our previous work, we demonstrated that ZIF-8-derived NPCs showed a superior catalytic activity on oxidation of HMF to FDCA using graphitic-N as the active site.46 The functionalized MOFs could also be used as a template during thermal treatment. For instance, Fu et al.87 synthesized amine contained UiO-66 and functionalized with glyphosine, a P-containing molecule, to form a P-containing UiO-66. After pyrolysis at 1000 °C, P-containing UiO-66 nanoparticles (NPs) became defect rich carbon which is favorable for oxygen reduction reaction (ORR). The resulting P-containing PCs showed comparable limiting current density and half way potential to that of platinum on carbon, a state-of-the-art of electrode for ORR. The transition metals (Ni, Co, Fe) are used for graphitization of amorphous carbon.88 Thanks to the homogeneous distribution of SBUs and ligands, the MOF-derived PC could be in situ graphitized into graphite using SBUs. Yamauchi’s group synthesized ZIF-8/ZIF-67 hybrid NPs and further converted into graphitized PCs.79 ZIF-8 derived PCs show high content of N with appreciable surface area while the carbon is amorphous. On the contrary, ZIF-67 derived PCs show high degree of graphitization while low content of N and low surface area was observed. The PCs derived from hybrid NPs showed comparable N content and surface area to ZIF-8-derived PCs while the degree of graphitization is much higher than ZIF-8-derived PCs. The resulting catalyst exhibited a distinguishable specific capacitance owing to the in-situ graphitization.79

ACS Paragon Plus Environment

17

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 48

Figure 4. Synthesis of MOF-derived materials by thermal treatment. (a) Direct pyrolysis of ZIF8 into nitrogen containing porous carbons (NPCs) as catalyst. (b) Direct pyrolysis of ZIF-8/ZIF67 hybrid nanoparticles into Co single atoms on nitrogen-doped porous carbon (Co SAs/N-C). (c) High resolution TEM image of Co SAs/N-C. The red circles represent Co single atoms. (d) The corresponding EXAFS fitting curves with schematic model of Co SAs/N-C, Co (purple), N (blue), O (red), and C (gray). Reprinted with permission from Ref.89 Copyright 2016 John Wiley and Sons (e) Direct calcination of MOF embedded KIT-6 into metal-oxides@KIT-6 as catalyst.

ACS Paragon Plus Environment

18

Page 19 of 48 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

(f) High resolution TEM image of Co3O4@KIT-6 derived from ZIF-67@KIT-6. Reprinted with permission from Ref.90 Copyright 2018 Royal Society of Chemistry. Metal composite NPs, including metal oxide and metal are widely used in catalysis and gas sensing. MOF-derived PCs have tunable porosity and specific surface area which is a benefit for deposition of metal composites. The combination of metal NPs with MOF-derived PCs is demonstrated to have better performance than either PCs or metal composite NPs itself. The metal NPs-containing MOF-derived PCs could be synthesized by either introducing a selected metal source into a MOF or a MOF-derived PC or directly converting the SBUs into metal NPs by thermal treatment. Fischer’s group91 has summarized a series of methodology for introducing metal NPs into a MOF, including impregnation,92 solid grinding,93 chemical vapor infiltration.94 Xu’s group demonstrated an improved incipient wetness impregnation to introduce single metal source and bi-metal sources into a MOF using the co-solvent method.95 The metal ions are dissolved in water which is in same volume with a total pore volume of MIL-101. Both MIL-101 and metal ion-containing water were suspended in hexane, becoming a MIL-101/water in hexane system. They demonstrated that using co-solvent method, the metal nanoparticles evenly distributed in a MOF with a narrow size distribution than using impregnation or chemical vapor infiltration. Furthermore, an alloy metal NPs could be synthesized with the co-solvent method. In our previous work, we introduced Au NPs into a ZIF-8 by de novo approach where metal ions were infiltrated into a MOF during assembly of MOF.96 Our results also demonstrated an even distribution of Au NPs with narrow size distribution, too. We further converted the Au NPs embedded ZIF-8 into Au NPs embedded PCs by pyrolysis under N2 environment. As shown in the previous section that the nitrogen atoms on ZIF-8 had many useful functions after pyrolysis

ACS Paragon Plus Environment

19

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 48

of ZIF-8 into PCs, the nitrogen atoms on PCs changed the micro-environment of Au NPs and promoted the performance of Au NPs on the reduction of 4-nitrophenol into 4-aminophenol. Chen et al.97 directly carbonized ZIF-8 into NPCs. The abundant nitrogen atoms were used to stabilize palladium precursors, followed by reduction of Pd precursors into Pd clusters using H2 gas. The abundant nitrogen atoms on NPCs server two function in the catalyst. It works as anchor point for Pd ions while it also generates hydrogen bond with water molecules by lone pair of electrons, which make the surface of carbon hydrophilic. The Pd-NPCs were applied for cascade conversion of vanillin into 2-methoxy-4-methylphenol, demonstrating high conversion and selectivity. The results were attributed to the hydrophilic surface of Pd-NPCs since the intermediate, vanillin alcohol is hydrophilic molecule while the reactant, vanillin and the product, 2-methoxy-4-methylphenol are hydrophobic molecules. Li et al.98 similarly applied MOF-253 which is composed of 2,2’-bipyridine-5,5’-dicarboxylic acid and aluminum as the template, and decorated Pd NPs on MOF-253-derived NPCs for Knoevenagel condensationhydrogenation. The inherent N atoms worked as anchor points for Pd NPs, at the same time, provided Lewis basic sites for a tandem reaction with a distinguishable result. Wang et al.99 directly used SBUs as metal source and directly annulled MOF into MPCs. A Ni-BTC (BTC: benzene tricarboxylic acid) was added with zinc ions during synthesis. As proofed by Xu that zinc-SBUs would evaporate during pyrolysis, generating cavities in the carbon structure.81 The zinc-SBUs act as pore-forming agents, making MOF-derived carbon become porous by evaporating at 900 °C (pyrolysis). The porous Ni/C showed an improved electrocatalytic performance in hydrogen evolution because the porous structure apparently reduced the charge transfer resistance in carbon substrate, displaying a semi double-layer capacitance for porous Ni/C.

Li’s group further reduced the scale of the SBU-derived metal

ACS Paragon Plus Environment

20

Page 21 of 48 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

composites from NPs into isolated single atoms (Fig. 4b).89 A hybrid ZIF-8/ZIF-67 NPs were synthesized to dilute the density of Co atoms in a MOF particle. The hybrid ZIF-8/ZIF-67 NPs were further annulled into Co single atoms on nitrogen-doped porous carbon (Co SAs/N-C) under Ar environment at 900 °C (Fig. 4c). Co ions were reduced into Co atoms by carbon and stabilized by nitrogen while Zn ions evaporated during pyrolysis. The EDS mapping of Co atoms in different Z-axis reflected different intensity, indicating the distribution of Co was homogeneous in 3-D space rather than on the surface. The simulated model showed that one Co atom coordinated with two N atoms and two C atoms, showing both higher value of limiting current density and halfway potential in ORR owing to the strong interaction between Co-2N (Fig. 4d). Besides PCs and MPCs, MOF could be converted into metal oxide, serving as active sites for catalysis and/or as a host for immobilization of catalyst. Porous metal oxide structures (PMOs) are fabricated from SBUs. Thanks to the wide combination between SBUs and ligands, a desired PMO could be easily synthesized by selecting a desired metal precursor as SBU. For instance, Co3O4 NPs,100 CuO NPs101 and TiO2 NPs102 were synthesized by calcination of ZIF-67 NPs, MOF-199 (HKUST-1) and MIL-125(Ti) particles under air atmosphere. Furthermore, not only monometallic PMOs but also bimetallic PMOs could be fabricated by using MOFs as the template. We delivered secondary metal precursor into MOF by de novo approach.103 Cu(NO3)2 was introduced into MIL-125(Ti) upon the synthesis. The Cu-containing MIL-125(Ti) was directly calcined into CuO embedded mesoporous titania tablets (CuO@MTs). The element distribution showed homogeneous distribution of Cu atoms in a MT. The results of H2temperature programed reduction (H2-TPR) and X-ray absorption near edge structure (XANES) showed that some parts of CuO NPs were inserted into TiO2 layers, which promoted the

ACS Paragon Plus Environment

21

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 48

performance of CuO@MTs on photocatalytic hydrogen evolution by effective charge transfer between CuO and TiO2. However, the rapid and severe reaction between MOF and oxygen, like carbon combustion and metal oxidation, the MOF-derived PMOs were not sufficiently well controlled to generate nanomaterials. To prevent the severe reaction, Li’s group synthesized ZIF-67 in mesoporous silica, KIT-6, followed by calcination under air at 250 °C to remove organic ligands (Fig. 4e).90 Co3O4 NPs were confined within the channel of KIT-6 with an average size of ca. 2 nm (Fig. 4f). On the contrary, Co3O4 NPs aggregating on the outer surface of KIT-6 had an average size of ca. 13 nm, indicating the function of porous substrate on controlling the particle size of metal oxide. The MOF-derived metal oxide composites showed catalytic activity on oxidation of HMF into FDCA under milder conditions. The results could be attributed to the separated SBUs by ligands and confinement of metal oxide in the silica matrix. In this section, we introduce three types of MOF-derived materials (i.e. PCs, MPCs and PMOs) that can be fabricated by thermal treatment. The heteroatoms on ligands provide additional function for catalyst. The overgrowth of metal seeds is prevented owing to separated and well aligned SBUs in a MOF particle. A catalyst with catalytic activity to selected reaction could be fabricated by choosing a favorable MOF containing SBUs whose derivative has activity to that reaction. The flexible combination of ligands and SBUs makes MOF-derived materials efficient alternative catalyst in biomass conversion. Biomass conversion over MOF-derived catalysts Hydrolysis of saccharides into sugar monomers The UiO-66 and MIL-101(Cr) based MOF-derived materials have been significantly used in the biomass conversion so far. Both UiO-66 and MIL-101(Cr) are proved to be robust MOFs

ACS Paragon Plus Environment

22

Page 23 of 48 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

which are stable against acid and high temperature. The MIL-101(Cr) features large pore size (1.2 nm and 1.6 nm), open metal sites, and the feasibility of functionalization for catalytic biomass conversion. For example, the Brønsted acidity can be provided to MIL-101 by functionalizing the organic ligand (terephthalic acid) with –SO3H groups. Hydrolysis of cellulose was studied using MIL-101(Cr)-SO3H by Kitagawa’s group,53 where the results showed a very lower yield of monosaccharides (5%) in water. The inferior yield of monosaccharide (5%) is observed because of lower Brønsted acidity and heterogeneous nature of the catalyst which causes diffusion limitation for bulky cellulose. The Brønsted acidic MOF (MIL-101(Cr)-SO3H) used for cellulose hydrolysis was synthesized directly from 2sulfoterephthalate monosodium salt and chromium oxide and has lower acidity (1.8 mmol/g). The difficulties in synthesizing highly Brønsted acidic MOF is its stability in presence of higher acid concentration because most of the MOFs are not stable under strongly acidic conditions. The recycle test demonstrated the stability of MIL-101(Cr)-SO3H even after testing recycling experiment for 13 times which indicates that this MOF is extremely stable compared with a popular solid acid catalyst like Amberlyst-15. The MIL-101(Cr)-SO3H was also used as the catalyst for glucose isomerization for the first time in 2014 by the Kitagawa group.54 Both MIL-101(Cr) and MIL-100(Cr) are chromiumbased MOF while chromium oxide coordinated with trimesic acid in MIL-100(Cr) and chromium oxide coordinated with terephthalic acid in MIL-101(Cr). MIL-101(Cr) exhibited a better performance than MIL-100(Cr) because the aperture size of MIL-101(Cr) (1.2 nm and 1.6 nm) was larger than the kinetic diameter of glucose (0.8 nm), making easy diffusion of glucose into the pore of MIL-101(Cr). The electron withdrawing functional group such as –NO2 and – SO3H presenting on the ligand of MIL-101 can enhance the activity of MOF for glucose

ACS Paragon Plus Environment

23

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 48

isomerization. Because the electron withdrawing groups increase Lewis acidity of SBUs, the MIL-101(Cr)-SO3H not only produce higher fructose yield but also showed 100% glucose and fructose recovery. Conversely, electron donating groups (–NH2 and –(CH3)2) increases electron density on the SBUs resulting in the decrease in Lewis acidity. Hence, electron donating groups on the ligands decreases the activity of MIL-101(Cr) towards glucose isomerization.54 Similar to their previous work, MIL-101(Cr)-SO3H showed a good stability in recycling test for the glucose isomerization. The results demonstrate that MOF-derived materials can be a potential catalyst for hydrolysis reactions. Conversion of sugars into furan derivatives The dehydration of biomass-derived C5 and C6 sugars (fructose and glucose) is an important reaction for the synthesis of platform chemicals such as FAL and HMF which could be further converted into various value-added furan derivatives by hydrogenation and oxidation reactions. Recently, the synthesis of FAL and HMF has attracted much attention, and an extensive research work for the production of furans from C5 and C6 sugars have been reported.6, 10 Several catalysts have been reported for the dehydration reactions of fructose and glucose into HMF apart from MOF-derived materials such as acidic ionic liquids (ILs), metal complexes and oxides, heteropoly acid-based materials, mineral acid and organic acids, metal halides, ion-exchange resins, zeolites, functionalized carbonaceous materials, functionalized mesoporous materials, etc.6, 104-108 The employment of MOF-derived materials for fructose and glucose dehydration reactions has been extensively studied because in many cases MOF-derived catalysts are easy to be separated and recycled from the reaction mixture in addition to their high catalytic ability. Moreover, MOF-derived materials can provide Lewis as well as Brønsted acidity which is useful for glucose dehydration reaction, as shown in Scheme 3. The Lewis

ACS Paragon Plus Environment

24

Page 25 of 48 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

acidity comes from open metal sites present at the SBUs, and Brønsted acidity comes from acid functionalized ligands. The presence of Lewis acidity and Brønsted acidity helps the conversion of glucose into HMF in a more selective pathway and hence inhibit side reactions. In addition, the presence of regular porous structure helps the diffusion of sugar molecules. OH

HO HO

O OH Glucose

OH BA

LA

More selective pathway

OH

O HO

HO

O O

HO HO

OH

Less selective pathway

OH HO Fructose

Levoglucosan

BA

BA

HO

O

Humin formation

O HMF BA

O OH

HO

O

O LA

FA

Scheme 3. Pathways for glucose cyclodehydration into HMF. A MIL-101(Cr) based MOF-derived material was first employed in a THF/H2O solvent system for glucose dehydration to HMF.109 The MIL-101(Cr)-SO3H exhibited 29% yield of HMF from glucose.109 The MOF catalysts (MIL-101(Cr)-SO3H and MIL-101(Cr)-NO2) are selective for HMF synthesis compared to conventional catalysts such as sulfuric acid and Amberlyst-15.109 On the other hand, the specific surface area, X-ray diffraction peaks of catalyst, and the yield of HMF after four cycles of reaction remained unchanged, demonstrating a good potential of MIL-101(Cr)-SO3H for glucose dehydration. The MIL-101(Cr)-SO3H was also

ACS Paragon Plus Environment

25

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 48

reported for continuous reaction (fixed bed reactor) for glucose dehydration in γ-GVL/H2O (9:1 v/v) solvent and showed improvement in yield of HMF (45%) with a short reaction time of 3.5 h.110 Fukuoka and co-workers demonstrated the dehydration of glucose into HMF using phosphate modified NU-1000 (material prepared with equimolar phosphoric acid concentration PO4/NU(eq) and half-equimolar phosphoric acid concentration PO4/NU(half)).111 The glucose dehydration using unmodified NU-1000 showed only 2.3% yield of HMF and 19% yield of fructose in H2O at 140 ºC in 5 h, whereas the PO4/NU(half) showed 15% yield of HMF under similar reaction condition. The modified phosphate molecules work as a negative molecule to poison the Brønsted acid sites on zirconia (SBUs) which generate undesired glucose transformation (e.g. humins) during reaction. However, PO4/NU(eq) showed the results similar to bare NU-1000, indicating the Brønsted acid sites still have some positive effect on the reaction. Hence PO4/NU(half) catalyst exhibits higher HMF yield compared to unmodified NU-1000 and PO4/NU(eq). Furthermore, the chromium-based porous coordination polymer and its sulfonic functionalized derivatives was successfully employed by Du et al. for glucose dehydration which showed over 99% conversion of glucose and 80.7% yield of HMF in water/THF/NaCl solvent system in 4 h at 180 ºC.112 This higher yield of HMF was possible because of presence of Brønsted and Lewis acidity in the catalyst. Furthermore, the NaCl will help increase the immiscibility between water and THF phase, thus lower side reactions. A five-times recycle test showed only 3% loss of yield (from 80% to 77%) of glucose on the catalyst in the glucose dehydration. Till now this is a promising catalytic system reported using the MOF-derived material for the synthesis of HMF using glucose. The fructose dehydration into HMF using a MOF-derived catalyst was reported recently, too. Hatton et al.39 reported the synthesis of a MOF-polymer composite (MOF-PMAi-Br,

ACS Paragon Plus Environment

26

Page 27 of 48 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

produced using MIL-101(Cr) surrounded in a crosslinked poly(N-bromo-maleimide)) for fructose dehydration with a high HMF yield of 87% in DMSO solvent. The higher yield (87%) with MOF-PMAi-Br compared to PMAi-Br (50%) could be attributed to 3D structure of MOF which unfold the polymer chain and expose bromine to fructose. In Table 1, the dehydration reactions of sugars into furans are presented using MOF or MOF-derived materials as the catalysts. As shown in Table 1, the dehydration of fructose into HMF was performed in DMSO using various MOF-SO3H as solid acidic catalysts at 120 ºC for 1 h.113 The catalytic efficiency of MOF-SO3H increases in the subsequent order as MIL-101(Cr)-SO3H (90%) > UiO-66-SO3H (85%) > MIL-53(Al)-SO3H (79%) in fructose-to-HMF dehydration. The catalyst with higher sulfonic acid groups among MOF-SO3H catalysts showed better performance; however, the MIL101(Cr)-SO3H catalyst was not very active for the dehydration of glucose into HMF because of the inefficiency of MIL-101(Cr)-SO3H catalyst for the glucose isomerization into fructose. For xylose dehydration into FAL, very limited reports are available using MOF-derived materials as the catalyst. The MOF-based mixed matrix membrane reactor (Zn2(bim)4-PMPS) was used by Weishen and coworkers for the xylose dehydration and obtained 41% FAL yield in 9 h at 140 ºC.50

The use of a heterogeneous MOF and MOF-derived material such as MIL-101(Cr), UiO66, PCP(Cr)-SO3H, etc. for the dehydration of sugars offers higher surface area, metal sites, and bi-functional acid sites such as Brønsted and Lewis. The MOF and MOF-derived materials are also easily recyclable and offers similar activity in recycling experiment in most of the cases. Although several MOF-derived materials are used for the dehydration reactions of C6 sugars (glucose and fructose) to synthesize HMF, there is no successful MOF-derived material which can efficiently convert glucose sugar into HMF in an environmentally benign solvent (H2O) under milder reaction condition. To make HMF synthesis sustainable and more economically

ACS Paragon Plus Environment

27

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 48

viable, selection of H2O as reaction solvent and glucose substrate over fructose is necessary due to its lower cost and wider supply compared to fructose.

Table 1. Dehydration of C5 and C6 sugars (glucose and fructose) using MOF and MOF-derived materials as catalyst.

Entry

Substrate

Catalyst

Solvent

Temp (ºC)

Time (h)

Yield (%)

Ref.

1

Glucose

MIL-101Cr-SO3H

THF/H2O (39:1 v/v)

130

24

29 HMF

109

2

Glucose

PO4/NU-1000 (half)

H2O/2-propanol (1:9 v/v)

140

7

64 HMF

111

3

Glucose

MOF-PMAi-Br

DMSO

100

6

16 HMF

39

4

Glucose

PCP(Cr)-SO3H

H2O/THF/NaCl

180

4

80.7 HMF

112

5

Glucose

MIL-101(Cr)-SO3H

GVL/H2O (9:1 v/v)

150

2

45 HMF

110

6

Glucose

MIL-101(Cr)-SO3H

[BMIM][Cl]

120

2

8 HMF

113

7

Glucose

NU-1000

H2O

140

5

2.3 HMF

111

8

Fructose

MIL-101(Cr)-SO3H

DMSO

120

1

90 HMF

113

9

Fructose

UIO-66-SO3H

DMSO

120

1

85 HMF

113

10

Fructose

MIL-53(Al)-SO3H

DMSO

120

1

79 HMF

113

11

Fructose

NUS-6(Hf)

DMSO

100

1

98 HMF

114

12

Fructose

PTA (3)/MIL-101

[EMIM][Cl]

80

1

63 HMF

115

13

Fructose

PMAi-Br/MIL-101

DMSO

100

1

78 HMF

39

14

Fructose

UiO-66(Hf)

DMSO

100

1

8 HMF

114

Upgrading furans into fine chemicals Furan-based chemicals like HMF and FAL are regarded as platform materials which could be further converted into value-added products via hydrogenation, oxidation, ring-opening,

ACS Paragon Plus Environment

28

Page 29 of 48 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

hydroxylation, or combination of above reactions.16, 56, 116 Here we disclose the advantages of MOF-derived materials as the catalysts for the conversion of furan-based materials into fine chemicals. Rafael et al. employed the MIL-45B as a sacrificial template for the synthesis of Fe-Co non-noble catalyst. This catalyst was used for the oxidation of HMF to 2,5-diformylfuran (DFF) which exhibited very high yield of DFF (99%) under mild conditions.56 The unique hollow structure of the MOF-derived Fe-Co catalyst favors the adsorption of HMF and quick desorption of the formed DFF due to weak interaction with MOF-derived Fe-Co which results in higher yield of DFF. Furthermore, the catalyst is stable under reaction condition and showed no drop in DFF

yield during recycling test. The same research group also reported the MOF-derived Fe3O4 catalyst for the conversion of fructose into DFF in a one-pot fashion, as shown in Scheme 4. This catalyst was stable under reaction condition and showed complete conversion of fructose with DFF selectivity >99% in 5 h, at 100 ºC.57 The Fe/C-S synthesized from MIL-88B offers strong adsorption of HMF and weak adsorption of DFF because the two formyl groups of DFF molecule improves the electrostatic repulsive interaction with catalyst surface which helps in the HMF conversion and decreasing side reactions which helps in the HMF conversion and decreasing

side reactions. Moreover, the Fe/C-S catalyst has magnetic properties which make the easy separation of the catalyst from a reaction mixture. Furthermore, the atomic absorption spectroscopy (AAS) analysis result showed no metal leaching in the reaction solution.

OH

O HO

HO

3-H 2O OH

HO

O

O

O

O

O

O

HMF

DFF

HO Fructose

Scheme 4. Selective oxidation of fructose into DFF.

ACS Paragon Plus Environment

29

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 48

The cyclopentanone (CPO) is an important chemical used in the manufacturing of insecticides, perfumes, medicines, rubber chemicals and solvent in the electronics industry. Generally, CPO is produced by oxidation of cyclopentene, catalytic vapor-phase cyclization of adipic esters or 1,6-hexanediols,51, 117 while recently the synthesis of CPO was also carried out from FAL which is a renewable chemical obtained from biomass and hence has received much interest.4, 9 MOF-derived CuNi@C catalyst was reported for an efficient conversion of FAL into CPO.51 The pyrolysis of Cu-based MOFs impregnated with Ni(NO3)2 is used for the synthesis of MOF-derived CuNi@C catalyst. The porous carbon used could prevent the accumulation of metal particles and acts as a support for the metal. The CuNi0.5@C catalyst provides the highest catalytic performance with 97% CPO yield and over 99% FAL conversion. It was found that Cu@C catalyst useful for FAL rearrangement and produces CPO with 52% yield and showed 79% conversion of FAL.51 On the other hand, the addition of Ni in HKUST-1 showed dramatic improvement in the performance of the catalytic activity (99% FAL conversion with 97% CPO yield) owing to the synergistic effect of metal NPs. However, when Ni-Cu bimetallic catalysts supported on SBA-15 was used for FAL hydrogenation, 39% CPO selectivity with 46% FAL conversion was observed within 4 h at 160 ºC (4 MPa H2).118 These results suggested that MOFderived CuNi@C could perform better compared to the bimetallic Ni-Cu@SBA-15 catalyst. Furthermore, the selective hydrogenation of FAL into CPO was reported using Ru nanoparticle supported on acidic MOF (Ru/MIL-101) catalyst.119 The Ru/MIL-101 exhibited complete conversion of FAL within 2.5 h at 160 ºC (4 MPa H2) with very high CPO selectivity (96%). The activity of Ru/MIL-101 was compared with Ru supported on active carbon (Ru/AC) under similar reaction condition the inferior results are seen (72% furfural conversion with 40% CPO selectivity) owing to the inherent acidic microenvironment provided by Lewis acid sites on MIL-

ACS Paragon Plus Environment

30

Page 31 of 48 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

101. Catalyst characterization (AAS analysis) study revealed that no metal leaching observed in the reaction solution. Moreover, the crystal structure of the catalyst remains unchanged after the reaction. In hydrogenation reactions of FAL, it was found that MOF-derived CuNi@C can help to increase the concentration of reactants around the active metal site and also contribute to the hydrogenation.120 A bimetallic CuCo/C was fabricated with similar procedure by Xiao’s group and the catalyst was applied for hydrogenation of FAL into FOL.52 The conversion was achieved to 99% with 97.3% selectivity of FAL. Compared with bimetallic MPC, MOF-derived Cu/C showed only 46.6% of conversion with 81.3% selectivity of FAL. The CuCo/C demonstrated the stability of catalyst in the recycle test by showing an unapparent difference of selectivity between each cycle. These two results suggested that MOF are potential template for synthesis of MPCs since metal precursors are tightly confined in MOF, which could prevent the overgrowth of metal NPs during pyrolysis, and secondary metal source could evenly distribute around SBUs, which would lead to an even distribution of bimetallic NPs. A further hydrogenation of FAL would make the product become THFA. Another case of hydrogenation of FAL to THFA was done by Wang’s group who prepared Ni-based MPCs for hydrogenation of FAL by directly pyrolysis of Ni-MOFs under nitrogen.121 The TEM images showed an even distribution of Ni NPs on carbon substrate owing to the formation of Ni-carbines which plays a crucial role on preventing the aggregation of Ni NPs. The MOF-derived Ni/C showed total conversion of FAL into tetrahydrofurfuryl alcohol (THFA) as MOFs were pyrolyzed at 500 °C. The amount of nickel active site and nickel carbine were changed with the pyrolysis temperature. Ni active sites dominate the reaction while nickel carbines stabilize nickel NPs. The Ni NPs were stable at 400 ºC during the reaction while the amount of active sites was not enough for complete conversion

ACS Paragon Plus Environment

31

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 48

of FAL into THFA. At 600 °C, most of Ni-carbines were converted into nickel active sites for hydrogenation while the Ni NPs became unstable due to inexistence of Ni-carbine. It suggests that the coordination between SBUs and ligands could stabilize the metal NPs by forming a metal-carbon bond. Ning et al. synthesized a 2D Co-based MOF which was pyrolyzed under argon environment from 150 °C to 800 °C.122 The oxidation condensation of FAL to 3-(furan-2-yl-)-2methylacrylaldehyde showed bi-modal distribution of conversion and selectivity over MOFderived material. The first peak value of conversion and selectivity showed up as MOFs were pyrolyzed at 300 °C. According to thermogravimetric analysis, the SBUs are confined by organic ligands while the moisture covering SBUs were removed as pyrolysis temperature was lower than 400 °C. The unsaturated active sites on SBUs were exposed to reactants (FAL) for oxidation condensation reaction. As the temperature was increased to 800 °C, the framework started to decompose and SBUs transformed into α-Co NPs immobilized on carbon matrix. The α-Co NPs showed better conversion and selectivity on the oxidation condensation of FAL than the unsaturated active sites on SBUs. Although the catalysts pyrolyzed at different temperature showed different activity for oxidation condensation reaction, both catalysts demonstrated the good recyclability in the reuse test. The selective hydrogenation of HMF into 2,5-dihydroxymethyl-tetrahydrofuran (DHMTHF) was reported using Pd immobilized on amine-functionalized MOF [Pd/MIL101(Al)-NH2] catalyst.55 This catalyst offers preferential adsorption to intermediate 2,5dihydroxymethylfuran (DHMF) formed during the hydrogenation compared to HMF adsorption. Because of enhanced hydrophilic nature of DHMF, it adsorbs extensively over Pd/MIL-101(Al)NH2 catalyst which results in the further hydrogenation of DHMF into DHMTHF. The optimized

ACS Paragon Plus Environment

32

Page 33 of 48 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

reaction condition showed 96% DHMTHF yield with complete HMF conversion using Pd/MIL101(Al)-NH2 catalyst.55 The other catalysts such as Pd/MIL-53(Al)-NH2, Pd/MIL-53(Al), and Pd/MIL-101(Cr) exhibited lower activity than MIL-101(Al)-NH2 catalyst. The higher activity in case of MIL-101(Al)-NH2 was seen because of well-dispersed Pd NPs over MIL-101(Al)-NH2 and –NH2 functional group cooperates for selective hydrogenation of HMF into DHMTHF. The presence of –NH2 group in the frameworks helps for the formation of highly well-dispersed Pd NPs and also preferential adsorption of DHMF intermediate through hydrogen bonding formed during the hydrogenation reaction. Valorization of lignin As per our knowledge, no reports have been published yet for the valorization of real lignin using MOF-derived catalysts. However, few reports are available on the conversion of lignin model compounds. The model compounds used so far have included benzyl phenyl ether (BPE), phenyl ethyl phenyl ether (PPE), diphenyl ether (DPE), vanillyl alcohol (VAL), etc., respectively.123-124 Two types of isorecticular MOF-74(Mg) (IRMOF-74 (I, II)) were used efficiently to cleave β-O-4 and 4-O-5 linkages of phenyl ethers. To further promote the selectivity of the catalyst, titanium and nickel were doped into MOF framework. Compared to other commercial catalysts, the doping of Titanium (Ti) and Nickel (Ni) in these catalysts improved the conversion and selectivity for hydrogenolysis versus ring hydrogenation.124 Carbon supported cobalt (CSC) nanocomposite catalyst derived from ZIF-67 is reported for the conversion of VAL into vanillin. The CSC catalyst is stable under reaction condition and offers excellent selectivity for vanillin (>95%).123 Furthermore, hydro-deoxygenation (HDO) of vanillin into 2-methoxy-4-methyl phenol was reported by Zhang and co-workers using Pd@MIL-101(Cr)-SO3H and Pd@UiO-66-NH2 catalysts. The reference Pd@MIL-101(Cr)

ACS Paragon Plus Environment

33

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 48

catalyst showed lower conversion of vanillin (67%) and selectivity of 2-methoxy-4methylphenol (58%) compared to Pd@MIL-101(Cr)-SO3H and [email protected] The higher activity Pd@UiO-66-NH2 is observed because of higher adsorption of vanillin and vanillin alcohol due to the interaction of hydrophilic substrate molecule and catalyst. This reveals that MOF-derived materials can also act as a promising catalyst for hydrogenolysis of ligninderived chemicals. In conclusion, MOF-derived materials showed catalytic activity for many important biomass conversions including hydrolysis of saccharide, dehydration of sugars (xylose, glucose and fructose) into furan, and upgradation of furans into fine chemical. The good performance of MOF-derived materials on biomass conversion could be attributed to: a) the synergism between active sites and functionalized ligand and b) the separated active sites oriented from MOF. Conclusion and future prospect The MOF-derived materials inherit the porous and tunable structure from mother MOF while the surface functionalization makes MOF materials more active, and the annulling process significantly improved stability and electrical conductivity of material. As highlighted in this review, the diversity of MOFs has been demonstrated by a highly versatile combination between SBUs and organic ligands. The diversity of mother MOFs gives the MOF-derived materials an excellent performance on many biomass reactions such as hydrolysis, dehydration, hydrogenation reactions of carbohydrates and hydrogenation of oxidation reactions of furans (FAL and HMF), respectively. As a heterogeneous catalyst, MOF-derived materials demonstrate the recyclability and stability after recovery of sample from reaction media which is highly beneficial in industrial production. For the efficient catalytic valorization of biomass concentrated substrate solution selection will help for making economically promising and

ACS Paragon Plus Environment

34

Page 35 of 48 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

sustainable processes. Moreover, non-edible biomass should be preferred as a substrate over edible biomass which will not conflict with food demand. It is always desirable that to practice the conversion of lignocellulosic biomass into sugars and furans in one-pot fashion without using any pretreatment for reducing the cost of the experiment. Apart from these advantages of MOFs and MOF-derived materials for biomass conversion reactions, there are several future prospects described as follows: (1) There have been more than 20,000 kinds of MOFs which are not explored for MOF-derived materials yet. The potential and possibility of these MOFs in catalysis should be explored urgently. (2) The effect of physical properties of mother MOF on the catalysis over MOF-derived materials is seldom discussed. For example, a hollow MOF-derived material could be produced by synthesis of a layer of MOF on a soft template which could accelerate the catalytic activity by enhancing the diffusion of molecules. (3) It has been reported for the synthesis of alloy NPs or single atom in the MOF, showing good performance in chemical catalysis, photocatalysis and oxygen reduction reaction. While best of our knowledge, none of the paper reported for biomass conversion using alloy NPs or single atom-containing MOF-derived materials. The higher activity of alloy NPs and single atom are expected to accelerate the complex biomass conversion. (4) The serious weight loss after annulling process makes the cost of MOF-derived increase which is not favorable for industrial application. The development of large-scale synthesis process with high yields to lower the cost of MOFs is also urgent. (5) Functionalized MOF-derived materials might not provide enough efficiency on more complicated cascade reaction such as conversion of glucose to FDCA including isomerization, dehydration, and three step oxidations. A way is to combine MOFs and substrates such as carbon fiber cloth or silica NPs can help for improving catalytic efficiency.

ACS Paragon Plus Environment

35

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 48

Overall, the well-developed MOF-derived materials make the biomass conversion more efficient and valuable. Although there are still unknown and the possibility of MOF-derived materials waiting for exploration, we expect that future researchers could focus on these prospects to prepare high-performance MOF-derived materials for not only biomass conversion but also electrochemistry, photo-catalysis and gas sensing applications.

AUTHOR INFORMATION Corresponding Author *Kevin C.-W. Wu [email protected] Author Contributions ‡These authors contributed equally.

ACKNOWLEDGMENT We would like to thank the Ministry of Science and Technology (MOST) of Taiwan (104-2628E-002-008-MY3; 105-2218-E-155-007; 105-2221-E-002-003-MY3; 105-2221-E-002-227-MY3; 105-2622-E-155-003-CC2) and the Aim for Top University Project at National Taiwan University (105R7706) for the funding support.

References

ACS Paragon Plus Environment

36

Page 37 of 48 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

(1) Huber, G. W.; Iborra, S.; Corma, A., Synthesis of transportation fuels from biomass:  Chemistry, catalysts, and engineering. Chem. Rev. 2006, 106 (9), 4044-4098. https://doi.org/10.1021/cr068360d (2) Zhang, Z.; Song, J.; Han, B., Catalytic transformation of lignocellulose into chemicals and fuel products in ionic liquids. Chem. Rev. 2017, 117 (10), 6834-6880. https://doi.org/10.1021/acs.chemrev.6b00457 (3) Deepa, A. K.; Dhepe, P. L., Lignin depolymerization into aromatic monomers over solid acid catalysts. ACS Catal. 2015, 5 (1), 365-379. https://doi.org/10.1021/cs501371q (4) Matsagar, B. M.; Hossain, S. A.; Islam, T.; Alamri, H. R.; Alothman, Z. A.; Yamauchi, Y.; Dhepe, P. L.; Wu, K. C. W., Direct production of furfural in one-pot fashion from raw biomass using Brønsted acidic ionic liquids. Sci. Rep. 2017, 7 (1), 13508. https://doi.org/10.1038/s41598-017-13946-4 (5) Matsagar, B. M.; Hossain, S. A.; Islam, T.; Yamauchi, Y.; Wu, K. C. W., A novel method for the pentosan analysis present in jute biomass and its conversion into sugar monomers using acidic ionic liquid. J. Vis. Exp. 2018, (136), e57613. https://doi.org/10.3791/57613 (6) Matsagar, B. M.; Munshi, M. K.; Kelkar, A. A.; Dhepe, P. L., Conversion of concentrated sugar solutions into 5-hydroxymethyl furfural and furfural using Brønsted acidic ionic liquids. Catal. Sci. Technol. 2015, 5 (12), 5086-5090. https://doi.org/10.1039/C5CY00858A (7) Mika, L. T.; Cséfalvay, E.; Németh, Á., Catalytic conversion of carbohydrates to initial platform chemicals: Chemistry and sustainability. Chem. Rev. 2018, 118 (2), 505-613. https://doi.org/10.1021/acs.chemrev.7b00395 (8) van Putten, R.-J.; van der Waal, J. C.; de Jong, E.; Rasrendra, C. B.; Heeres, H. J.; de Vries, J. G., Hydroxymethylfurfural, a versatile platform chemical made from renewable resources. Chem. Rev. 2013, 113 (3), 1499-1597. https://doi.org/10.1021/cr300182k (9) Matsagar, B. M.; Dhepe, P. L., Brönsted acidic ionic liquid-catalyzed conversion of hemicellulose into sugars. Catal. Sci. Technol. 2015, 5 (1), 531-539. https://doi.org/10.1039/c4cy01047g (10) Matsagar, B. M.; Dhepe, P. L., Effects of cations, anions and H+ concentration of acidic ionic liquids on the valorization of polysaccharides into furfural. New J. Chem. 2017, 41 (14), 6137-6144. https://doi.org/10.1039/c7nj00342k (11) Dhepe, P. L.; Sahu, R., A solid-acid-based process for the conversion of hemicellulose. Green Chem. 2010, 12 (12), 2153-2156. https://doi.org/10.1039/c004128a (12) Huang, Y. B.; Fu, Y., Hydrolysis of cellulose to glucose by solid acid catalysts. Green Chem. 2013, 15 (5), 1095-1111. https://doi.org/10.1039/C3GC40136G (13) Shrotri, A.; Kobayashi, H.; Fukuoka, A., Cellulose depolymerization over heterogeneous catalysts. Acc. Chem. Res. 2018, 51 (3), 761-768. https://doi.org/10.1021/acs.accounts.7b00614 (14) Priecel, P.; Endot, N. A.; Carà, P. D.; Lopez-Sanchez, J. A., Fast catalytic hydrogenation of 2,5-hydroxymethylfurfural to 2,5-dimethylfuran with ruthenium on carbon nanotubes. Ind. Eng. Chem. Res. 2018, 57 (6), 1991-2002. https://doi.org/10.1021/acs.iecr.7b04715 (15) Iriondo, A.; Mendiguren, A.; Güemez, M. B.; Requies, J.; Cambra, J. F., 2,5-DMF production through hydrogenation of real and synthetic 5-HMF over transition metal catalysts supported on carriers with different nature. Catal. Today 2017, 279 286-295. https://doi.org/10.1016/j.cattod.2016.02.019 (16) McKenna, S. M.; Mines, P.; Law, P.; Kovacs-Schreiner, K.; Birmingham, W. R.; Turner, N. J.; Leimkühler, S.; Carnell, A. J., The continuous oxidation of HMF to FDCA and the

ACS Paragon Plus Environment

37

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 48

immobilisation and stabilisation of periplasmic aldehyde oxidase (PaoABC). Green Chem. 2017, 19 (19), 4660-4665. https://doi.org/10.1039/C7GC01696D (17) Xiao, B.; Zheng, M.; Li, X.; Pang, J.; Sun, R.; Wang, H.; Pang, X.; Wang, A.; Wang, X.; Zhang, T., Synthesis of 1,6-hexanediol from HMF over double-layered catalysts of Pd/SiO2 + Ir– ReOx/SiO2 in a fixed-bed reactor. Green Chem. 2016, 18 (7), 2175-2184. https://doi.org/10.1039/C5GC02228B (18) Kuwahara, Y.; Kango, H.; Yamashita, H., Catalytic transfer hydrogenation of biomassderived levulinic acid and its esters to gamma-valerolactone over sulfonic acid-functionalized UiO-66. ACS Sustainable Chem. Eng. 2017, 5 (1), 1141-1152. https://doi.org/10.1021/acssuschemeng.6b02464 (19) Bhaumik, P.; Dhepe, P. L., Chapter 1 Conversion of Biomass into Sugars. In Biomass Sugars for Non-Fuel Applications, R. Soc. Chem.: 2016, 1-53. (20) Jiang, L.; Wu, N.; Zheng, A.; Zhao, Z.; He, F.; Li, H., The integration of dilute acid hydrolysis of xylan and fast pyrolysis of glucan to obtain fermentable sugars. Biotechnol. Biofuel 2016, 9 (1), 196-205. https://doi.org/10.1186/s13068-016-0612-0 (21) Kim, E. S.; Liu, S.; Abu-Omar, M. M.; Mosier, N. S., Selective conversion of biomass hemicellulose to furfural using maleic acid with microwave heating. Energy Fuels 2012, 26 (2), 1298-1304. https://doi.org/10.1021/ef2014106 (22) Bhaumik, P.; Dhepe, P. L., Efficient, stable, and reusable silicoaluminophosphate for the one-pot production of furfural from hemicellulose. ACS Catal. 2013, 3 (10), 2299-2303. https://doi.org/10.1021/cs400495j (23) Hara, M., Biomass conversion by a solid acid catalyst. Energy Environ. Sci. 2010, 3 (5), 601-607. https://doi.org/10.1039/B922917E (24) Gürbüz, E. I.; Gallo, J. M. R.; Alonso, D. M.; Wettstein, S. G.; Lim, W. Y.; Dumesic, J. A., Conversion of hemicellulose into furfural using solid acid catalysts in γ-valerolactone. Angew. Chem. Int. Edit. 2013, 52 (4), 1270-1274. https://doi.org/10.1002/anie.201207334 (25) Zhang, H.; Cheng, Y.-T.; Vispute, T. P.; Xiao, R.; Huber, G. W., Catalytic conversion of biomass-derived feedstocks into olefins and aromatics with ZSM-5: The hydrogen to carbon effective ratio. Energy Environ. Sci. 2011, 4 (6), 2297-2307. https://doi.org/10.1039/C1EE01230D (26) Rinaldi, R.; Palkovits, R.; Schüth, F., Depolymerization of cellulose using solid catalysts in ionic liquids. Angew. Chem. Int. Edit. 2008, 47 (42), 8047-8050. https://doi.org/10.1002/anie.200802879 (27) Srinivasa Rao, B.; Krishna Kumari, P.; Dhana Lakshmi, D.; Lingaiah, N., One pot selective transformation of biomass derived chemicals towards alkyl levulinates over titanium exchanged heteropoly tungstate catalysts. Catal. Today 2018, 309 269-275. https://doi.org/10.1016/j.cattod.2017.05.040 (28) Wang, Z.; Chen, Q., Conversion of 5-hydroxymethylfurfural into 5-ethoxymethylfurfural and ethyl levulinate catalyzed by MOF-based heteropolyacid materials. Green Chem. 2016, 18 (21), 5884-5889. https://doi.org/10.1039/C6GC01206J (29) Besson, M.; Gallezot, P.; Pinel, C., Conversion of biomass into chemicals over metal catalysts. Chem. Rev. 2014, 114 (3), 1827-1870. https://doi.org/10.1021/cr4002269 (30) Swatloski, R. P.; Spear, S. K.; Holbrey, J. D.; Rogers, R. D., Dissolution of cellose with ionic liquids. J. Am. Chem. Soc. 2002, 124 (18), 4974-4975. https://doi.org/10.1021/ja025790m

ACS Paragon Plus Environment

38

Page 39 of 48 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

(31) 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 (6941), 705-714. https://doi.org/10.1038/nature01650 (32) Tranchemontagne, D. J.; Mendoza-Cortes, J. L.; 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. https://doi.org/10.1039/B817735J (33) Park, K. S.; Ni, Z.; Cote, A. P.; Choi, J. Y.; Huang, R. D.; 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), 10186-10191. https://doi.org/10.1073/pnas.0602439103 (34) 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. https://doi.org/10.1126/science.283.5405.1148 (35) Millange, F.; Serre, C.; Ferey, G., Synthesis, structure determination and properties of MIL-53as and MIL-53ht: the first CrIII hybrid inorganic-organic microporous solids: CrIII(OH){O2C-C6H4-CO2}{HO2C-C6H4-CO2H}x. Chem. Commun. 2002, (8), 822-823. https://doi.org/10.1039/B201381A (36) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K. P., A new zirconium inorganic building brick forming metal organic frameworks with exceptional stability. J. Am. Chem. Soc. 2008, 130 (42), 13850-13851. https://doi.org/10.1021/ja8057953 (37) Feng, D. W.; Gu, Z. Y.; Li, J. R.; Jiang, H. L.; Wei, Z. W.; Zhou, H. C., Zirconiummetalloporphyrin PCN-222: Mesoporous metal-organic frameworks with utrahigh stability as bomimetic calysts. Angew. Chem. Int. Edit. 2012, 51 (41), 10307-10310. https://doi.org/10.1002/anie.201204475 (38) Furukawa, H.; Cordova, K. E.; O'Keeffe, M.; Yaghi, O. M., The chemistry and applications of metal-organic frameworks. Science 2013, 341 (6149), 974-975. https://doi.org/10.1126/science.1230444 (39) Bromberg, L.; Su, X.; Hatton, T. A., Functional networks of organic and coordination polymers: catalysis of fructose conversion. Chem. Mater. 2014, 26 (21), 6257-6264. https://doi.org/10.1021/cm503098p (40) Zhan, W. W.; Kuang, Q.; Zhou, J. Z.; Kong, X. J.; Xie, Z. X.; Zheng, L. S., Semiconductor@metal-organic framework core-shell heterostructures: A case of ZnO@ZIF-8 nanorods with selective photoelectrochemical response. J. Am. Chem. Soc. 2013, 135 (5), 19261933. https://doi.org/10.1021/ja311085e (41) Xia, W.; Qu, C.; Liang, Z. B.; Zhao, B. T.; Dai, S. G.; Qiu, B.; Jiao, Y.; Zhang, Q. B.; Huang, X. Y.; Guo, W. H.; Dang, D.; Zou, R. Q.; Xia, D. G.; Xu, Q.; Liu, M. L., Highperformance energy storage and conversion materials derived from a single metal organic framework/graphene aerogel composite. Nano Lett. 2017, 17 (5), 2788-2795. https://doi.org/10.1021/acs.nanolett.6b05004 (42) Campagnol, N.; Romero-Vara, R.; Deleu, W.; Stappers, L.; Binnemans, K.; De Vos, D. E.; Fransaer, J., A hybrid supercapacitor based on porous carbon and the metal-organic framework MIL-100(Fe). Chemelectrochem 2014, 1 (7), 1182-1188. https://doi.org/10.1002/celc.201402022 (43) Liang, K.; Richardson, J. J.; Doonan, C. J.; Mulet, X.; Ju, Y.; Cui, J. W.; Caruso, F.; Falcaro, P., An enzyme-coated metal-organic framework shell for synthetically adaptive cell

ACS Paragon Plus Environment

39

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 48

survival. Angew. Chem. Int. Edit. 2017, 56 (29), 8510-8515. https://doi.org/10.1002/anie.201704120 (44) Li, S. B.; Dharmarwardana, M.; Welch, R. P.; Ren, Y. X.; Thompson, C. M.; Smaldone, R. A.; Gassensmith, J. J., Template-directed synthesis of porous and protective core-shell bionanoparticles. Angew. Chem. Int. Edit. 2016, 55 (36), 10691-10696. https://doi.org/10.1002/anie.201604879 (45) Lim, S.; Suh, K.; Kim, Y.; Yoon, M.; Park, H.; Dybtsev, D. N.; Kim, K., Porous carbon materials with a controllable surface area synthesized from metal-organic frameworks. Chem. Commun. 2012, 48 (60), 7447-7449. https://doi.org/10.1039/c2cc33439a (46) Nguyen, C. V.; Liao, Y. T.; Kang, T. C.; Chen, J. E.; Yoshikawa, T.; Nakasaka, Y.; Masuda, T.; Wu, K. C. W., A metal-free, high nitrogen-doped nanoporous graphitic carbon catalyst for an effective aerobic HMF-to-FDCA conversion. Green Chem. 2016, 18 (22), 59575961. https://doi.org/10.1039/C6GC02118B (47) Liu, N. L.; Dutta, S.; Salunkhe, R. R.; Ahamad, T.; Alshehri, S. M.; Yamauchi, Y.; Hou, C. H.; Wu, K. C. W., ZIF-8 derived, nitrogen-doped porous electrodes of carbon polyhedron particles for high-performance electrosorption of salt ions. Sci. Rep. 2016, 6 28847. https://doi.org/10.1038/srep28847 (48) Xu, J.; Liu, S. C.; Liu, Y., Co3O4/ZnO nanoheterostructure derived from core-shell ZIF8@ZIF-67 for supercapacitors. RSC Adv. 2016, 6 (57), 52137-52142. https://doi.org/10.1039/C6RA07773K (49) Wang, X.; Wu, X. L.; Guo, Y. G.; Zhong, Y. T.; Cao, X. Q.; Ma, Y.; Yao, J. N., Synthesis and lithium storage properties of Co3O4 nanosheet-assembled multishelled hollow spheres. Adv. Funct. Mater. 2010, 20 (10), 1680-1686. https://doi.org/10.1002/adfm.200902295 (50) Jin, H.; Liu, X.; Ban, Y.; Peng, Y.; Jiao, W.; Wang, P.; Guo, A.; Li, Y.; Yang, W., Conversion of xylose into furfural in a MOF-based mixed matrix membrane reactor. Chem. Eng. J. 2016, 305 12-18. http://dx.doi.org/10.1016/j.cej.2015.10.115 (51) Wang, Y.; Sang, S.; Zhu, W.; Gao, L.; Xiao, G., CuNi@C catalysts with high activity derived from metal–organic frameworks precursor for conversion of furfural to cyclopentanone. Chem. Eng. J. 2016, 299 104-111. https://doi.org/10.1016/j.cej.2016.04.068 (52) Wang, Y.; Miao, Y. N.; Li, S.; Gao, L. J.; Xiao, G. M., Metal-organic frameworks derived bimetallic Cu-Co catalyst for efficient and selective hydrogenation of biomass-derived furfural to furfuryl alcohol. Mol. Catal. 2017, 436 128-137. https://doi.org/10.1016/j.mcat.2017.04.018 (53) Akiyama, G.; Matsuda, R.; Sato, H.; Takata, M.; Kitagawa, S., Cellulose hydrolysis by a new porous coordination polymer decorated with sulfonic acid functional groups. Adv. Mater. 2011, 23 (29), 3294-3297. https://doi.org/10.1002/adma.201101356 (54) Akiyama, G.; Matsuda, R.; Sato, H.; Kitagawa, S., Catalytic glucose isomerization by porous coordination polymers with open metal sites. Chem-Asian J. 2014, 9 (10), 2772-2777. https://doi.org/10.1002/asia.201402119 (55) Chen, J.; Liu, R.; Guo, Y.; Chen, L.; Gao, H., Selective hydrogenation of biomass-based 5-hydroxymethylfurfural over catalyst of palladium immobilized on amine-functionalized metal– organic frameworks. ACS Catal. 2015, 5 (2), 722-733. https://doi.org/10.1021/cs5012926 (56) Fang, R.; Luque, R.; Li, Y., Selective aerobic oxidation of biomass-derived HMF to 2,5diformylfuran using a MOF-derived magnetic hollow Fe-Co nanocatalyst. Green Chem. 2016, 18 (10), 3152-3157. https://doi.org/10.1039/C5GC03051J

ACS Paragon Plus Environment

40

Page 41 of 48 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

(57) Fang, R.; Luque, R.; Li, Y., Efficient one-pot fructose to DFF conversion using sulfonated magnetically separable MOF-derived Fe3O4 (111) catalysts. Green Chem. 2017, 19 (3), 647-655. https://doi.org/10.1039/C6GC02018F (58) Herbst, A.; Janiak, C., MOF catalysts in biomass upgrading towards value-added fine chemicals. CrystEngComm 2017, 19 (29), 4092-4117. https://doi.org/10.1039/C6CE01782G (59) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O'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), 469-472. https://doi.org/10.1126/science.1067208 (60) Cmarik, G. E.; Kim, M.; Cohen, S. M.; Walton, K. S., Tuning the adsorption properties of UiO-66 via ligand functionalization. Langmuir 2012, 28 (44), 15606-15613. https://doi.org/10.1021/la3035352 (61) Park, J.; Wang, Z. Y. U.; Sun, L. B.; Chen, Y. P.; Zhou, H. C., Introduction of functionalized mesopores to metal-organic frameworks via metal-ligand-fragment coassembly. J. Am. Chem. Soc. 2012, 134 (49), 20110-20116. https://doi.org/10.1021/ja3085884 (62) McGuirk, C. M.; Katz, M. J.; Stern, C. L.; Sarjeant, A. A.; Hupp, J. T.; Farha, O. K.; Mirkin, C. A., Turning on catalysis: Incorporation of a hydrogen-bond-donating squaramide moiety into a Zr metal-organic framework. J. Am. Chem. Soc. 2015, 137 (2), 919-925. https://doi.org/10.1021/ja511403t (63) Karagiaridi, O.; Bury, W.; Mondloch, J. E.; Hupp, J. T.; Farha, O. K., Solvent-assisted linker exchange: An alternative to the de novo synthesis of unattainable metal-organic frameworks. Angew Chem. Int. Edit. 2014, 53 (18), 4530-4540. https://doi.org/10.1002/anie.201306923 (64) Kim, M.; Cahill, J. F.; Su, Y. X.; Prather, K. A.; Cohen, S. M., Postsynthetic ligand exchange as a route to functionalization of 'inert' metal-organic frameworks. Chem. Sci. 2012, 3 (1), 126-130. https://doi.org/10.1039/C1SC00394A (65) Takaishi, S.; DeMarco, E. J.; Pellin, M. J.; Farha, O. K.; Hupp, J. T., Solvent-assisted linker exchange (SALE) and post-assembly metallation in porphyrinic metal-organic framework materials. Chem. Sci. 2013, 4 (4), 1509-1513. https://doi.org/10.1039/C2SC21516K (66) Canivet, J.; Aguado, S.; Daniel, C.; Farrusseng, D., Engineering the environment of a catalytic metal-organic framework by postsynthetic hydrophobization. Chemcatchem 2011, 3 (4), 675-678. https://doi.org/10.1002/cctc.201000386 (67) Goesten, M. G.; Juan-Alcaniz, J.; Ramos-Fernandez, E. V.; Gupta, K. B. S. S.; Stavitski, E.; van Bekkum, H.; Gascon, J.; Kapteijn, F., Sulfation of metal-organic frameworks: Opportunities for acid catalysis and proton conductivity. J. Catal. 2011, 281 (1), 177-187. https://doi.org/10.1016/j.jcat.2011.04.015 (68) Kuwahara, Y.; Yoshimura, Y.; Yamashita, H., Liquid-phase oxidation of alkylaromatics to aromatic ketones with molecular oxygen over a Mn-based metal-organic framework. Dalton Trans. 2017, 46 (26), 8415-8421. https://doi.org/10.1039/C7DT01351E (69) Fei, H. H.; Cohen, S. M., Metalation of a thiocatechol-functionalized Zr(IV)-based metalorganic framework for selective C-H functionalization. J. Am. Chem. Soc. 2015, 137 (6), 21912194. https://doi.org/10.1021/ja5126885 (70) Fei, H. H.; Shin, J. W.; Meng, Y. S.; Adelhardt, M.; Sutter, J.; Meyer, K.; Cohen, S. M., Reusable oxidation catalysis using metal-monocatecholato species in a robust metal-organic framework. J. Am. Chem. Soc. 2014, 136 (13), 4965-4973. https://doi.org/10.1021/ja411627z

ACS Paragon Plus Environment

41

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 48

(71) Aijaz, A.; Xu, Q., Catalysis with metal nanoparticles immobilized within the pores of metal-organic frameworks. J. Phys. Chem. Lett. 2014, 5 (8), 1400-1411. https://doi.org/10.1021/jz5004044 (72) Jiang, H.; Lee, P. S.; Li, C. Z., 3D carbon based nanostructures for advanced supercapacitors. Energ. Environ. Sci. 2013, 6 (1), 41-53. https://doi.org/10.1039/C2EE23284G (73) Zhai, Y. P.; Dou, Y. Q.; Zhao, D. Y.; Fulvio, P. F.; Mayes, R. T.; Dai, S., Carbon materials for chemical capacitive energy storage. Adv. Mater. 2011, 23 (42), 4828-4850. https://doi.org/10.1002/adma.201100984 (74) Wang, R. T.; Wang, P. Y.; Yan, X. B.; Lang, J. W.; Peng, C.; Xue, Q. J., Promising porous carbon derived from celtuce leaves with outstanding supercapacitance and CO2 capture performance. ACS Appl. Mater. Inter. 2012, 4 (11), 5800-5806. https://doi.org/10.1021/am302077c (75) Hong, K. L.; Qie, L.; Zeng, R.; Yi, Z. Q.; Zhang, W.; Wang, D.; Yin, W.; Wu, C.; Fan, Q. J.; Zhang, W. X.; Huang, Y. H., Biomass derived hard carbon used as a high performance anode material for sodium ion batteries. J. Mater. Chem. A 2014, 2 (32), 12733-12738. https://doi.org/10.1039/C4TA02068E (76) Hao, G. P.; Li, W. C.; Qian, D.; Lu, A. H., Rapid synthesis of nitrogen-doped porous carbon monolith for CO2 capture. Adv. Mater. 2010, 22 (7), 853-854. https://doi.org/10.1002/adma.200903765 (77) Tang, J.; Salunkhe, R. R.; Zhang, H. B.; Malgras, V.; Ahamad, T.; Alshehri, S. M.; Kobayashi, N.; Tominaka, S.; Ide, Y.; Kim, J. H.; Yamauchi, Y., Bimetallic metal-organic frameworks for controlled catalytic graphitization of nanoporous carbons. Sci. Rep. 2016, 6 30295. https://doi.org/10.1038/srep30295 (78) Yang, S. J.; Kim, T.; Im, J. H.; Kim, Y. S.; Lee, K.; Jung, H.; Park, C. R., MOF-Derived Hierarchically Porous Carbon with Exceptional Porosity and Hydrogen Storage Capacity. Chem. Mater. 2012, 24 (3), 464-470. https://doi.org/10.1021/cm202554j (79) Tang, J.; Salunkhe, R. R.; Liu, J.; Torad, N. L.; Imura, M.; Furukawa, S.; Yamauchi, Y., Thermal conversion of core-shell metal-organic frameworks: A new method for selectively functionalized nanoporous hybrid carbon. J. Am. Chem. Soc. 2015, 137 (4), 1572-1580. https://doi.org/10.1021/ja511539a (80) Liu, S. H.; Wang, Z. Y.; Zhou, S.; Yu, F. J.; Yu, M. Z.; Chiang, C. Y.; Zhou, W. Z.; Zhao, J. J.; Qiu, J. S., Metal-organic-framework-derived hybrid carbon nanocages as a bifunctional electrocatalyst for oxygen reduction and evolution. Adv. Mater. 2017, 29 (31), 1700874. https://doi.org/10.1002/adma.201700874 (81) Liu, B.; Shioyama, H.; Akita, T.; Xu, Q., Metal-organic framework as a template for porous carbon synthesis. J. Am. Chem. Soc. 2008, 130 (16), 5390-5391. https://doi.org/10.1021/ja7106146 (82) Chae, H. K.; Siberio-Perez, D. Y.; Kim, J.; Go, Y.; Eddaoudi, M.; Matzger, A. J.; O'Keeffe, M.; Yaghi, O. M., A route to high surface area, porosity and inclusion of large molecules in crystals. Nature 2004, 427 (6974), 523-527. https://doi.org/10.1038/nature02311 (83) Jiang, H. L.; Liu, B.; Lan, Y. Q.; Kuratani, K.; Akita, T.; Shioyama, H.; Zong, F. Q.; Xu, Q., From metal-organic framework to nanoporous carbon: Toward a very high surface area and hydrogen uptake. J. Am. Chem. Soc. 2011, 133 (31), 11854-11857. https://doi.org/10.1021/ja203184k (84) Chaikittisilp, W.; Hu, M.; Wang, H. J.; Huang, H. S.; Fujita, T.; Wu, K. C. W.; Chen, L. C.; Yamauchi, Y.; Ariga, K., Nanoporous carbons through direct carbonization of a zeolitic

ACS Paragon Plus Environment

42

Page 43 of 48 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

imidazolate framework for supercapacitor electrodes. Chem. Commun. 2012, 48 (58), 72597261. https://doi.org/10.1039/C2CC33433J (85) Zheng, F. C.; Yang, Y.; Chen, Q. W., High lithium anodic performance of highly nitrogen-doped porous carbon prepared from a metal-organic framework. Nat. Commun. 2014, 5 5261-5270. https://doi.org/10.1038/ncomms6261 (86) Wu, J. J.; Ma, L. L.; Yadav, R. M.; Yang, Y. C.; Zhang, X.; Vajtai, R.; Lou, J.; Ajayan, P. M., Nitrogen-doped graphene with pyridinic dominance as a highly active and stable electrocatalyst for oxygen reduction. ACS Appl. Mater. Inter. 2015, 7 (27), 14763-14769. https://doi.org/10.1021/acsami.5b02902 (87) Fu, Y.; Huang, Y.; Xiang, Z. H.; Liu, G. Q.; Cao, D. P., Phosphorous-nitrogen-codoped carbon materials derived from metal-organic frameworks as efficient electrocatalysts for oxygen reduction reactions. Eur. J. Inorg. Chem. 2016, (13-14), 2100-2105. https://doi.org/10.1002/ejic.201500822 (88) Schaefer, Z. L.; Gross, M. L.; Hickner, M. A.; Schaak, R. E., Uniform hollow carbon shells: Nanostructured graphitic supports for improved oxygen-reduction catalysis. Angew. Chem. Int. Edit. 2010, 49 (39), 7045-7048. https://doi.org/10.1002/anie.201003213 (89) Yin, P. Q.; Yao, T.; Wu, Y.; Zheng, L. R.; Lin, Y.; Liu, W.; Ju, H. X.; Zhu, J. F.; Hong, X.; Deng, Z. X.; Zhou, G.; Wei, S. Q.; Li, Y. D., Single cobalt atoms with precise Ncoordination as superior oxygen reduction reaction catalysts. Angew. Chem. Int. Edit. 2016, 55 (36), 10800-10805. https://doi.org/10.1002/anie.201604802 (90) Fang, R. Q.; Tian, P. L.; Yang, X. F.; Luque, R.; Li, Y. W., Encapsulation of ultrafine metal-oxide nanoparticles within mesopores for biomass-derived catalytic applications. Chem. Sci. 2018, 9 (7), 1854-1859. https://doi.org/10.1039/C7SC04724J (91) Rosler, C.; Fischer, R. A., Metal-organic frameworks as hosts for nanoparticles. CryStengComm 2015, 17 (2), 199-217. https://doi.org/10.1039/C4CE01251H (92) Liu, H. L.; Liu, Y. L.; Li, Y. W.; Tang, Z. Y.; Jiang, H. F., Metal-organic framework supported gold nanoparticles as a highly active heterogeneous catalyst for aerobic oxidation of alcohols. J. Phys. Chem. C 2010, 114 (31), 13362-13369. https://doi.org/10.1021/jp105666f (93) Jiang, H. L.; Liu, B.; Akita, T.; Haruta, M.; Sakurai, H.; Xu, Q., Au@ZIF-8: CO oxidation over gold nanoparticles deposited to metal-organic framework. J. Am. Chem. Soc. 2009, 131 (32), 11302-11303. https://doi.org/10.1021/ja9047653 (94) Hermes, S.; Schroter, M. K.; Schmid, R.; Khodeir, L.; Muhler, M.; Tissler, A.; Fischer, R. W.; Fischer, R. A., Metal@MOF: Loading of highly porous coordination polymers host lattices by metal organic chemical vapor deposition. Angew. Chem. Int. Edit. 2005, 44 (38), 6237-6241. https://doi.org/10.1002/anie.200462515 (95) Zhu, Q. L.; Li, J.; Xu, Q., Immobilizing metal nanoparticles to metal-organic frameworks with size and location control for optimizing catalytic performance. J. Am. Chem. Soc. 2013, 135 (28), 10210-10213. https://doi.org/10.1021/ja403330m (96) Liao, Y. T.; Chen, J. E.; Isida, Y.; Yonezawa, T.; Chang, W. C.; Alshehri, S. M.; Yamauchi, Y.; Wu, K. C. W., De novo synthesis of gold-nanoparticle-embedded, nitrogen-doped nanoporous carbon nanoparticles (Au@NC) with enhanced reduction ability. Chemcatchem 2016, 8 (3), 502-509. https://doi.org/10.1002/cctc.201501020 (97) Chen, Y. Z.; Cai, G. R.; Wang, Y. M.; Xu, Q.; Yu, S. H.; Jiang, H. L., Palladium nanoparticles stabilized with N-doped porous carbons derived from metal-organic frameworks for selective catalysis in biofuel upgrade: the role of catalyst wettability. Green Chem. 2016, 18 (5), 1212-1217. https://doi.org/10.1039/C5GC02530C

ACS Paragon Plus Environment

43

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 48

(98) Li, X. L.; Zhang, B. Y.; Fang, Y. H.; Sun, W. J.; Qi, Z. Y.; Pei, Y. C.; Qi, S. Y.; Yuan, P. Y.; Luan, X. C.; Goh, T. W.; Huang, W. Y., Metal-organic-framework-derived carbons: Applications as solid-base catalyst and support for Pd nanoparticles in tandem catalysis. ChemEur. J. 2017, 23 (18), 4266-4270. https://doi.org/10.1002/chem.201605852 (99) Wang, L.; Ren, L. T.; Wang, X. R.; Feng, X.; Zhou, J. W.; Wang, B., Multivariate MOFTemplated pomegranate-like Ni/C as efficient bifunctional electrocatalyst for hydrogen evolution and urea oxidation. ACS Appl. Mater. Inter. 2018, 10 (5), 4750-4756. https://doi.org/10.1021/acsami.7b18650 (100) Wu, R. B.; Qian, X. K.; Rui, X. H.; Liu, H.; Yadian, B. L.; Zhou, K.; Wei, J.; Yan, Q. Y.; Feng, X. Q.; Long, Y.; Wang, L. Y.; Huang, Y. Z., Zeolitic imidazolate framework 67-derived high symmetric porous Co3O4 hollow dodecahedra with highly enhanced lithium storage capability. Small 2014, 10 (10), 1932-1938. https://doi.org/10.1002/smll.201303520 (101) Banerjee, A.; Singh, U.; Aravindan, V.; Srinivasan, M.; Ogale, S., Synthesis of CuO nanostructures from Cu-based metal organic framework (MOF-199) for application as anode for Li-ion batteries. Nano Energy 2013, 2 (6), 1158-1163. https://doi.org/10.1016/j.nanoen.2013.04.008 (102) Xiu, Z.; Alfaruqi, M. H.; Gim, J.; Song, J.; Kim, S.; Thi, T. V.; Duong, P. T.; Baboo, J. P.; Mathew, V.; Kim, J., Hierarchical porous anatase TiO2 derived from a titanium metal-organic framework as a superior anode material for lithium ion batteries. Chem. Commun. 2015, 51 (61), 12274-12277. https://doi.org/10.1039/C5CC03381K (103) Liao, Y. T.; Huang, Y. Y.; Chen, H. M.; Komaguchi, K.; Hou, C. H.; Henzie, J.; Yamauchi, Y.; Ide, Y.; Wu, K. C. W., Mesoporous TiO2 embedded with a uniform distribution of CuO exhibit enhanced charge separation and photocatalytic efficiency. ACS Appl. Mater. Inter. 2017, 9 (49), 42425-42429. https://doi.org/10.1021/acsami.7b13912 (104) Matsagar, B. M.; Nguyen, C. V.; Hossain, M. S.; Islam, M. T.; Yamauchi, Y.; Dhepe, P. L.; Wu, K. C. W., Glucose isomerization catalyzed by bone char and selective production of 5hydroxymethylfurfural in aqueous media. 2018, 2 (10), 2148-2153. https://doi.org/10.1039/C8SE00339D (105) Peng, W. H.; Lee, Y. Y.; Wu, C.; Wu, K. C. W., Acid-base bi-functionalized, large-pored mesoporous silica nanoparticles for cooperative catalysis of one-pot cellulose-to-HMF conversion. J. Mater. Chem. 2012, 22 (43), 23181-23185. https://doi.org/10.1039/C2JM35391A (106) Hou, Q.; Zhen, M.; Liu, L.; Chen, Y.; Huang, F.; Zhang, S.; Li, W.; Ju, M., Tin phosphate as a heterogeneous catalyst for efficient dehydration of glucose into 5hydroxymethylfurfural in ionic liquid. Appl. Catal. B 2018, 224 183-193. https://doi.org/10.1016/j.apcatb.2017.09.049 (107) Lv, G.; Deng, L.; Lu, B.; Li, J.; Hou, X.; Yang, Y., Efficient dehydration of fructose into 5-hydroxymethylfurfural in aqueous medium over silica-included heteropolyacids. J. Clean. Prod. 2017, 142 (4), 2244-2251. https://doi.org/10.1016/j.jclepro.2016.11.053 (108) Choudhary, V.; Mushrif, S. H.; Ho, C.; Anderko, A.; Nikolakis, V.; Marinkovic, N. S.; Frenkel, A. I.; Sandler, S. I.; Vlachos, D. G., Insights into the interplay of lewis and brønsted acid catalysts in glucose and fructose conversion to 5-(hydroxymethyl)furfural and levulinic acid in aqueous media. J. Am. Chem. Soc. 2013, 135 (10), 3997-4006. https://doi.org/10.1021/ja3122763 (109) Herbst, A.; Janiak, C., Selective glucose conversion to 5-hydroxymethylfurfural (5-HMF) instead of levulinic acid with MIL-101Cr MOF-derivatives. New J. Chem. 2016, 40 (9), 79587967. https://doi.org/10.1039/C6NJ01399F

ACS Paragon Plus Environment

44

Page 45 of 48 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

(110) Su, Y.; Chang, G.; Zhang, Z.; Xing, H.; Su, B.; Yang, Q.; Ren, Q.; Yang, Y.; Bao, Z., Catalytic dehydration of glucose to 5-hydroxymethylfurfural with a bifunctional metal-organic framework. Am. Inst. Chem. Eng. J. 2016, 62 (12), 4403-4417. https://doi.org/10.1002/aic.15356 (111) Yabushita, M.; Li, P.; Islamoglu, T.; Kobayashi, H.; Fukuoka, A.; Farha, O. K.; Katz, A., Selective metal-organic framework catalysis of glucose to 5-hydroxymethylfurfural using phosphate-modified NU-1000. Ind. Eng. Chem. Res. 2017, 56 (25), 7141-7148. https://doi.org/10.1021/acs.iecr.7b01164 (112) Chen, D.; Liang, F.; Feng, D.; Xian, M.; Zhang, H.; Liu, H.; Du, F., An efficient route from reproducible glucose to 5-hydroxymethylfurfural catalyzed by porous coordination polymer heterogeneous catalysts. Chem. Eng. J. 2016, 300 177-184. https://doi.org/10.1016/j.cej.2016.04.039 (113) Chen, J.; Li, K.; Chen, L.; Liu, R.; Huang, X.; Ye, D., Conversion of fructose into 5hydroxymethylfurfural catalyzed by recyclable sulfonic acid-functionalized metal-organic frameworks. Green Chem. 2014, 16 (5), 2490-2499. https://doi.org/10.1039/C3GC42414F (114) Hu, Z.; Peng, Y.; Gao, Y.; Qian, Y.; Ying, S.; Yuan, D.; Horike, S.; Ogiwara, N.; Babarao, R.; Wang, Y.; Yan, N.; Zhao, D., Direct synthesis of hierarchically porous metal– organic frameworks with high stability and strong Brønsted acidity: The decisive role of hafnium in efficient and selective fructose dehydration. Chem. Mater. 2016, 28 (8), 2659-2667. https://doi.org/10.1021/acs.chemmater.6b00139 (115) Zhang, Y.; Degirmenci, V.; Li, C.; Hensen, E. J. M., Phosphotungstic acid encapsulated in metal–organic framework as catalysts for carbohydrate dehydration to 5hydroxymethylfurfural. Chemsuschem 2011, 4 (1), 59-64. https://doi.org/10.1002/cssc.201000284 (116) Wettstein, S. G.; Alonso, D. M.; Gürbüz, E. I.; Dumesic, J. A., A roadmap for conversion of lignocellulosic biomass to chemicals and fuels. Curr. Opin. Chem. Eng. 2012, 1 (3), 218-224. https://doi.org/10.1016/j.coche.2012.04.002 (117) Dubkov, K. A.; Panov, G. I.; Starokon, E. V.; Parmon, V. N., Non-catalytic liquid phase oxidation of alkenes with nitrous oxide. 2. Oxidation of cyclopentene to cyclopentanone. React. kinet. Catal. Lett. 2002, 77 (1), 197-205. https://doi.org/10.1023/A:1020372726494 (118) Yang, Y.; Du, Z.; Huang, Y.; Lu, F.; Wang, F.; Gao, J.; Xu, J., Conversion of furfural into cyclopentanone over Ni-Cu bimetallic catalysts. Green Chem. 2013, 15 (7), 1932-1940. https://doi.org/10.1039/C3GC37133F (119) Fang, R.; Liu, H.; Luque, R.; Li, Y., Efficient and selective hydrogenation of biomassderived furfural to cyclopentanone using Ru catalysts. Green Chem. 2015, 17 (8), 4183-4188. https://doi.org/10.1039/C5GC01462J (120) Chen, Z.; Chen, J.; Li, Y., Metal–organic-framework-based catalysts for hydrogenation reactions. Chin. J. Catal. 2017, 38 (7), 1108-1126. https://doi.org/10.1016/S18722067(17)62852-3 (121) Su, Y. P.; Chen, C.; Zhu, X. G.; Zhang, Y.; Gong, W. B.; Zhang, H. M.; Zhao, H. J.; Wang, G. Z., Carbon-embedded Ni nanocatalysts derived from MOFs by a sacrificial template method for efficient hydrogenation of furfural to tetrahydrofurfuryl alcohol. Dalton Trans 2017, 46 (19), 6358-6365. https://doi.org/10.1039/C7DT00628D (122) Ning, L. M.; Liao, S. Y.; Liu, X. G.; Yu, L. H.; Zhuang, X. L.; Tong, X. L., Selective transformation of renewable furfural catalyzed by diverse active species derived from 2D Cobased metal-organic frameworks. J. Catal. 2017, 352 480-490. https://doi.org/10.1016/j.jcat.2017.06.017

ACS Paragon Plus Environment

45

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 48

(123) Lin, K. Y. A.; Lai, H. K.; Chen, Z. Y., Selective generation of vanillin from catalytic oxidation of a lignin model compound using ZIF-derived carbon-supported cobalt nanocomposite. J. Taiwan Inst. Chem. Eng. 2017, 78 337-343. https://doi.org/10.1016/j.jtice.2017.06.029 (124) Stavila, V.; Parthasarathi, R.; Davis, R. W.; El Gabaly, F.; Sale, K. L.; Simmons, B. A.; Singh, S.; Allendorf, M. D., MOF-based catalysts for selective hydrogenolysis of carbon–oxygen ether bonds. ACS Catal. 2016, 6 (1), 55-59. https://doi.org/10.1021/acscatal.5b02061 (125) Zhang, F.; Jin, Y.; Fu, Y.; Zhong, Y.; Zhu, W.; Ibrahim, A. A.; El-Shall, M. S., Palladium nanoparticles incorporated within sulfonic acid-functionalized MIL-101(Cr) for efficient catalytic conversion of vanillin. J. Mater. Chem. A 2015, 3 (33), 17008-17015. https://doi.org/10.1039/C5TA03524D (126) Zhang, F.; Zheng, S.; Xiao, Q.; Zhong, Y.; Zhu, W.; Lin, A.; Samy El-Shall, M., Synergetic catalysis of palladium nanoparticles encaged within amine-functionalized UiO-66 in the hydrodeoxygenation of vanillin in water. Green Chem. 2016, 18 (9), 2900-2908. https://doi.org/10.1039/C5GC02615F

ACS Paragon Plus Environment

46

Page 47 of 48 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

For Table of Contents Use Only

Valorization of lignocellulosic biomass into fine chemical over a series of metal-organic frameworks-derived (MOF-derived) solid catalysts.

ACS Paragon Plus Environment

47

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 48 of 48

Photographs and Biographies Dr. Yu-Te Liao received his MS and Ph.D. under Prof. Kevin C.-W. Wu in chemical engineering from National Taiwan University (NTU). He works as postdoctoral researcher under the guidance of Prof. Kevin C.-W. Wu and his research is focused on the synthesis of metal embedded, metal–organic frameworks derived materials for variety of catalytic reactions.

Dr. Babasaheb M. Matsagar received his Ph.D. degree from CSIRNational Chemical Laboratory (NCL), India in 2016. Most of his research work is focused on the hydrolysis and dehydration reactions of saccharides, using acidic ionic liquids as a catalyst, and developing sustainable biomass conversion methods for chemicals and energy applications. At present, he is working as a postdoctoral researcher in the department of chemical engineering at National Taiwan University (NTU). Kevin C.-W. Wu obtained his Ph.D. degree in 2005 from The University of Tokyo and then worked as postdoctoral researcher at Waseda Univ. (Japan) and Iowa State Univ. (USA) before starting his own research group at the National Taiwan University in 2008. His research interests include the design and synthesis of functional nanoporous materials for energy and biomedical applications.

ACS Paragon Plus Environment

48