Advanced Porous Materials for Sensing, Capture and Detoxification of

Triazine Functionalized Porous Covalent Organic Framework for Photo-organocatalytic E–Z Isomerization of Olefins. Journal of the American Chemical S...
6 downloads 0 Views 3MB Size
Subscriber access provided by ALBRIGHT COLLEGE

Perspective

Advanced Porous Materials for Sensing, Capture and Detoxification of Organic Pollutants towards Water Remediation Partha Samanta, Aamod V. Desai, Sumanta Let, and Sujit K. Ghosh ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.9b00155 • Publication Date (Web): 20 Mar 2019 Downloaded from http://pubs.acs.org on March 21, 2019

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

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

Page 1 of 87 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

Advanced Porous Materials for Sensing, Capture and Detoxification of Organic Pollutants towards Water Remediation Partha Samanta,†,‡ Aamod V. Desai,†,‡ Sumanta Let† and Sujit K. Ghosh*,†, † Department

of Chemistry, Indian Institute of Science Education and Research, Dr. Homi

Bhabha Road, Pashan, Pune 411008, India. Phone: +91 20 2590 8076; E-mail: [email protected]

Centre for Energy Science, IISER Pune, Dr. Homi Bhabha Road, Pashan, Pune 411008,

India. KEYWORDS: Water pollution; Organic pollutants; Metal-organic frameworks; porous organic materials; Sensing, Sequestration.

ABSTRACT. Environmental pollution and its after effects has become a huge global concern in recent years. Among all type of pollutions, water pollution has been considered as one of the major threat in recent year and this problem will be even worse in near future. In the 21st century, water recycling has become a crucial issue as more

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 87

and more regions across the globe are witnessing water crisis. Detoxification of wastewater via adsorptive removal or photocatalytic degradation for achieving purified and drinkable water has come forth as environmentally benign, energy economic and cost-effective methods. In this context, porous materials have gained enormous research consideration in the recent past for detection of contaminants in effluents and for treatment of wastewater. A combination of high porosity, robustness and structural tunibility makes these nanosorbents stand out materials in environmental remediation. Metal-organic frameworks(MOFs), Covalent-organic frameworks (COFs) and Porousorganic solids are among the recently emerged versatile porous materials that are extensively investigated toward clean environment application. This review intends to provide a summarized compilation of recent research progress in sensing and sequestration of organic pollutants in advanced porous materials.

Introduction The quality of human life is directly influenced by the nature of the local environment. Access to fresh water has been recognized as an essential component to fully realize human rights.[1] The United Nations estimates suggest the global population to reach close to 10 billion by the year 2050.[2] Thus the requirement of fresh water is expected to cause tremendous stress on the existing sources. In addition, the growing industrialization and agricultural activities across the globe bestow significant stress by

ACS Paragon Plus Environment

2

Page 3 of 87 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

virtue of emitting large amount of wastewater. Thus the treatment and safe disposal of effluents is a pressing concern and necessitates urgent addressing. The contamination of water bodies by toxic pollutants also affects the balance of aquatic ecosystem and hence detoxification of aquatic streams is vital for the maintenance of ecological symbiosis. Water pollutants can be perceived as contaminants which are present in higher concentration than the permissible limits which cause adverse biological effects.[3] The prominent sources of contaminants include industrial effluents, oil spills, nuclear waste products, pesticides and mining activities. Several methods have been proposed for tackling this issue, including precipitation, coagulation, physical separation, but on account of cost considerations and higher sophistication, newer methods such as adsorption are seeking greater attention and relevance.[4] As a preliminary step, identification of pollutants holds significant relevance towards complete treatment. Hence, greater emphasis has also been bestowed towards developing approaches for the purpose of detecting species causing adverse biological impacts. Apart from capture and recognition, detoxification is a key step for complete remediation and hence materials or methods focused towards degradation of pollutants are important. The key aspect of dealing with remediation of environmental pollution understands the types of pollutant and its activity with biological species. Broadly, pollutants can be categorized into inorganic and organic based on the chemical constituents. Inorganic pollutants are typically heavy metals, halides and radioactive wastes etc.[5] On the other

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 87

hand, the class of organic pollutants covers wide array of pollutants including agrochemicals, pharmaceuticals, dyes, polyaromatic hydrocarbons etc.[6] Regulating agencies have identified persistent organic pollutants (POPs) as a class of compounds which accumulate in different environmental streams and cause serious noxious effects.[7] POPs in general, are artificially produced compounds which also include industrial by-products. Due to their ability to propagate across different streams, they transport over long range and accumulate or deposit at far off places from the point of origin. Pollutants in general are present as complex mixtures in wastewaters and hence outlining remediation protocols is not trivial. In addition to the knowledge of source, the pathway of activity is necessary to concentrate or destruct specific pollutants. The activity of conventionally observed pollutants is known in the literature but recent trends have resulted in several new pollutants which are termed as Emerging Contaminants (ECs).[8] ECs are typically compounds which were conventionally not identified as contaminants but have been found to cause adverse effects upon entering environment. This class of compound includes personal care products, pharmaceutical drugs, veterinary compounds, food additives, molecules released as industrial by-products and artificially synthesized nanomaterials. As lesser studies have been reported for understanding this class of compounds as pollutants, there is greater need to develop materials/methods to tackle the propagation of these compounds.

ACS Paragon Plus Environment

4

Page 5 of 87 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

Among the different methods tested for the removal of organic pollutants, adsorption based protocols have received affirmative attention on account to cost, simplicity and energy considerations. Development of novel materials or composites as sorbents for these applications has attracted remarkable research attention in the recent years. In particular, porous materials have been frontrunners in the research in this regard, owing to access to high surface areas, higher capacities and inclusion of specific functional sites.[9] In addition, affordance of several transduction pathways renders porous solids as superior materials for the detection of pollutants after entrapping them. Conventionally, the research in the domain of porous solids was limited to zeolites and activated carbons. Since the advent of crystalline coordination polymers and the subsequent emergence of metal-organic frameworks (MOFs), the scope and span of porous solids has enlarged substantially. MOFs are porous crystalline solids built from organic linkers and held by metal ions/clusters periodically and which bear potential voids.[10] Research has been devoted to the development of MOFs as they score over conventional porous materials in terms of tunability, structure-property correlation, presence of various signaling pathways for host-guest interactions and myriad combinations of building blocks. The feasibility of tuning pore architectures and modulating functionalities renders MOFs as suitable sorbents for the screening of specific analytes. On account of such benefits, MOFs have been used for several applications.[11-14] More recently, especially with the advent of water-stable MOFs,[15] this class of compounds has been

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 87

demonstrated for the sensing, capture and detoxification of a wide range of pollutants, present in different forms.[16] Scheme 1. Schematic representation of remediation of organic pollutants with advanced porous materials. Structure corresponding for COFs has been reproduced from the ref [114] with the permission of American Chemical Society.

Apart from MOFs, another class of crystalline porous materials viz. covalent organic frameworks (COFs) has emerged over the last decade or so.[17] Built on covalent linkages, in general, COFs are found to be more stable than MOFs due to the presence of covalent bonds. The ability to obtain crystalline powders ensures the characteristic

ACS Paragon Plus Environment

6

Page 7 of 87 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

advantage of MOFs viz. structure-property correlation, ability to tune pore surfaces and transduction mechanisms are retained. In the recent years, newer synthesis approaches and design strategies have been proposed to ease the synthesis of these materials and yield scalable materials, aimed at practical applications.[18] The prominent features in COFs of stability and extremely high surface areas have been tapped for diverse applications. Another classification of polymers based on organic linkages is that of porous organic polymers (POPs).[19] Unlike MOFs and COFs, which are crystalline solids, POPs are amorphous materials wherein direct attribution of structure-property correlation is non-trivial. Yet POPs, which are typically microporous, present high stability and score over MOFs and COFs in terms of processability. Alike previously described systems, employing functionalized building blocks can afford structural diversity and yield compounds suited for specific applications. Like its congeners, POPs too have been found to be applicable for the capture and sensing of several toxic species.[20] Among these materials MOFs have been applied extensively for the purpose of water remediation till date. But recent advancements have witnessed the emergence of crystalline COFs and amorphous POPs to scavenge organic pollutants from water. Aided with tunable surface area, and better structure property correlation has made MOFs one of the suitable candidate for sensing and detoxification of water. But, weak chemical stability in most of the cases has hindered their applications for the aforementioned apllications. On the other hand, high chemical stability of COFs and

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 87

POPs have been found to play an important role in the emergence of them for this type of specific applications. Apart from polymeric compounds, other porous materials such as porous molecular solids have been proposed which may present another feasible choice of materials as sorbents towards remediation of environmental pollutants.

Table 1. Table showing representative examples of different class of organic pollutants

ACS Paragon Plus Environment

8

Page 9 of 87 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

Advanced Porous materials for the sensing and capture of organic pollutants.

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 87

MOFs and porous organic materials have already been emerged as one of the suitable candidate for the remediation of the environmental pollution (Scheme 1). Among the features possessed by artificially porous solids, the ability to modulate and achieve specific electronic traits leads to this class of materials being able to function as host matrices for read-out of targeted/specific guest molecules. The signal transduction occurs via different modes such as colorimetric, luminescence, magnetic etc. Owing to lack of sophistication and sensitivity, luminescence-based processes have found remarkable attention. The luminescence features can be modulated and is controlled by the choice of building blocks, both in case of MOFs and COFs; although the underlying luminescence features are operational via different pathways.[14] Owing to such facile access, these materials are well suited for being constructed for specific guest recognition. Here, in this section we will discuss about the recent advancement in sensing and removal of various organic pollutants from water with metal-organic frameworks and porous organic materials like COFs, porous organic polymers etc. (Table 1). Different organic dyes. Organic dyes are known for their wide range of applications in various areas like textile, paper, paints, photographic, printing, food, plastics, cosmetics, pharmaceuticals etc. As a consequence of such huge applications water pollution due to organic dyes has become a global concern. It has been estimated that more than 10,000 organic dyes are available in the market commercially. Further estimation revealed that

ACS Paragon Plus Environment

10

Page 11 of 87 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

approximately 9,00,000 metric tons dyes are being produced every year and among them 17-20% dyes are directly contributing to the water pollution.[21] Based on the charge over the organic dyes, they are classified into: neutral, cationic and anionic dyes. Cationic dyes. There are several cationic dyes commercially available which includes methylene blue (MB), crystal violet (CV), rhodamine B (RhB), rhodamine 6G etc. Xu and coworkers reported three MOFs with formula of Cd(L)(bpy) (1), Cd(L)(bpy)∙4H2O∙2.5DMF (2), and Cd(L)(bpy)∙4.5H2O·3DMF (3) synthesized from Cd-ion, 4,4ˊ-bipyridine (bpy) and ligand L (2-amino-1,4-benzenedicarboxylic acid).[22] Third MOF was found to be mesoporous in nature and further utilized in the size exclusion liquid chromatographic technique to separate rhodamine 6G (smaller than the pore of the MOF) and Brilliant blue R-250 (larger than the pore). According to the authors, this was first report where MOF was used as stationary phase in liquid chromatography to separate dye molecules. In another report, Jia et al. reported two MOFs, namely JUC-101 [(In3O)(TDCPB)(3H2O)∙(xGuest)] and JUC-102 [(Mn3OH)(TDCPB)(3H2O)∙(xGuest); where, TDCPB: 1,3,5-tris(3,5-di(4-carboxy-phenyl-1-yl)phenyl-1-yl)benzene] with cationic and anionic framework charge respectively.[23] Owing to the anionic framework charge, JUC102 was employed for the capture of catonic dyes from solution. Capture of Azure B from DMF solution was observed with JUC-102 and later the MOF was regenerated with a saturated solution of NaCl in DMF medium. This regeneration of the MOF was only possible in presence of NaCl, whereas only DMF could not regenerated the MOF.

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 87

In another study by Zhu et al., two anionic MOFs were synthesized from Co- and Ni-ion with molecular formula of [(CH3)2NH2]6[M(H2O)6]3{M6(η6-TATAT)4(H2O)12}∙xH2O; for M: Co x: 6 and for M: Ni x: 11, H6TATAT: 5,5ˊ,5ˊˊ-(1,3,5-triazine-2,4,6-triyltriimino)tri1,3-benzenedicarboxylic acid].[24] Both the MOFs were found to be with rht topology. Owing to the presence of extra framework [(CH3)2NH2+ cation, Co-MOF and Ni-MOF both were employed for the selective removal of cationic dyes. Both the compounds were found to capture cationic dyes like Methylene Blue (MB), Basic Red 2 (BR2), Crystal Violet (CV) and Malachite Green (MG) from water. Due to the different size of the respective dye molecules, removal rate was observed to be MB > MG > CV > BR2. In addition, selectivity of MOFs was also checked and in this regard, both anionic dyes (Methyl Orange (MO) and Orange II) and neutral dye (Methyl Red (MR)) were taken into account. In both cases negligible uptake of the respective dyes was obtained. For real time applications, binary mixtures of dyes were performed and it was evident that even from binary mixtures selectively cationic dyes were adsorbed by the MOFs.

ACS Paragon Plus Environment

12

Page 13 of 87 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 1. (a) View of InIII-MOF along b-axis; UV-Vis absorbance spectra of dye solutions in DMF solvent in presence of InIII-MOF with time for (b) MB, (c) MB-MO, (d) MB-SD and (e) MB-CV. (Inset: respective picture of color of dye solutions before and after treatment with MOF); (f) recyclability experiment of the MOF with MB. Reproduced with permission from Ref. [25] Copyright 2015 Royal Society of Chemistry.

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 87

Recyclability of both compounds were also checked with saturated NaCl solution. Furthermore, Liu et al. reported another anionic framework, namely InIII-MOF with molecular formula of [(CH3)2NH2][In(L)]∙CH3CH2OH, where H4L: 2,3ˊ,5,5ˊ-biphenyl tetracarboxylic acid (Fig. 1a).[25] This InIII-MOF was consist of two different kinds of pore along b-axis, (i) 5.22x6.85 Å2 and (ii) 9.17x14.43 Å2. Selective dye capture experiment was carried out with a set of dyes which included cationic dyes (Methylene Blue and Crystal Violet), neutral dye (Sudan I (SD)) and anionic dye (MO). Since the size of MB, SD and MO were found to be similar, in a typical experiments InIII-MOF was soaked in respective dye solutions in DMF. UV-Visible spectra revealed the capture of cationic dye in InIIIMOF, whereas remaining two dyes were found to be remain in the solution. Further, the effect of the size of respective dyes were investigated by taking MB and CV, where rapid capture of MB was observed but on the other hand negligible change in concentration of CV was obtained (bigger in size as compare to MB). In addition, binary mixture studies revealed selective removal of MB was found even in presence of MO, SD and CV (Fig. 1). Enthused from such performance of InIII-MOF, a chromatographic column was prepared which showed separation of MB from aforementioned binary mixtures of dyes. Recyclability of the InIII-MOF was obtained with a saturated solution of NaCl in DMF medium (Fig. 1f). Apart from MOFs, different kind of porous organic materials were also utilized from the removal of cationic dyes from solutions. In a recent report, Ghosh and

ACS Paragon Plus Environment

14

Page 15 of 87 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

coworkers demonstrated cationic dyes capture with a post-synthetically modified hypercross-linked polymer (HCP), HCP-91@Na.[26] Cost-effective and chemically stable HCP91was prepard via Friedel-Crafts reaction in presence of lewis acid (FeCl3). Further, postsynthetic modification was carried out in HCP-91 to incorporate free cation inside the network through NaOH treatment. HCP-91@Na was employed for the capture of MB, CV and RhB from water medium and monitored with UV-Vis spectroscopy (Fig. 2). Because of the size of respective cationic dyes time taken for MB was least whereas bigger RhB took maximum time for the complete removal from water. Selectivity of the material was checked with an anionic dye, Methyl Orange which showed negligible change in absorbance in presence of HCP-91@Na. Binary mixture of MB-MO and RhBMO also revealed selective removal of MB and RhB from the respective solutions.

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 87

Figure 2. (a) Schematic representation for the synthesis of HCP-91@Na; UV-Vis absorbance spectra with different time intervals and respective photograph of dye solutions after the addition of HCP-91@Na for (b) methylene blue, (c) crystal violet and

ACS Paragon Plus Environment

16

Page 17 of 87 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

(d) rhodamine B. Reproduced with permission from Ref. [26] Copyright 2017 Royal Society of Chemistry. Anionic dyes. Apart from cationic dyes, anionic dyes are another major and important class of compounds in the family of organic dyes. In a recent , Ghosh and coworkers reported a water stable cationic framework which was even stable in base medium and further utilized for dye capture study.[27] In this report a Ni-based cationic framework, namely IPM-MOF-201 with formula of [{Ni(L)2}∙(BPSA)∙xG]n (where L: tris-(1H-imidazol)4-phenylamine, G: guest and BPSA: 4,4ˊ-biphenyldisulfonic acid), was synthesized via solvothermal method. IPM-MOF-201 was found to be excellent base resistant compound which was extremely rare in literature in case of cationic framework. Due to the presence of free anions inside the network, IPM-MOF-201 was employed for the anionic dye capture from water medium. This MOF showed excellent performance to remove anionic dyes like methyl orange, indigo carmine (IC) and alizarin red S (ARS). Selectivity was studied in presence of MB, IPM-MOF-201 showed selectivity for the anionic dyes over cationic dye. In addition to the dye capture in water medium, this MOF was further explored for the anion exchange in various pH-range (Fig. 3). For real time applicability IPM-MOF-201 was used in column chromatography and even in this case the MOF showed retention of its performance. In another report, Chen et al. demonstrated anionic dye capture with a cage-based cationic MOF in MeOH medium.[28] This cationic framework with formula of {[(Cu4Cl)(CPT)4-

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 87

(H2O)4]∙3NO3∙5NMP∙3.5H2O}n (where, NMP: N-methyl-2-pyrrolidone) was synthesized from Cu-ion and 4-(4-carboxyphenyl)-1,2,4-triazole (HCPT) ligand (Fig. 4a). Owing to the presence of free NO3ˉ ion inside the network, Cu-MOF was utilized for the anionic dye capture experiments. Upon capture of MO from MeOH, blue colored MOF turned into yellow green in color. Selectivity of the MOF was tested by a set of cationic dyes (MB and rhodamine B) and neutral dye (dimethyl yellow) (Fig. 4b-e). In addition, size selectivity was affirmed by the negligible uptake of bigger anionic dye molecule congo red (CR). Recyclability was achieved by the saturated solution of NaCl in methanol where facile release of MO was obtained. In another report, Song et al. synthesized a threedimensional In-based cationic framework with formula of [In(OH)L]5(NO3)5∙33H2O∙14DMF (where, H2L+Clˉ: 1,3-bis(4-carboxyphenyl)imidazolium chloride).[29] On account of the presence of free nitrate anions, In-MOF was chosen for the ion

ACS Paragon Plus Environment

18

Page 19 of 87 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. (a) Schematic representation of the metal-ligand coordination in IPM-MOF201; (b) Space-filling view of packing in IPM-MOF-201; (c) Representation of anionic dye capture at various pH-range with IPM-MOF-201. Reproduced with permission from Ref. [27] Copyright 2018 Cell Press.

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 87

exchange study for organic dye compounds. In this context, two sets of dye were chosen: (a) dyes with same size and different charges (methylene blue, sudan I, acid orange 7, new coccine and orange G) and (b) dyes with different sizes and same anionic charges (orange G, Crocein scarlet 3B, methyl blue and acid black 1). For the first set studies, UV-Vis absorption spectroscopy experiments revealed the decrease in concentration for the anionic dyes like acid orange 7, new coccine and orange G, whereas the no change was observed in absorbance for the cationic methylene blue and neutral sudan I. In case of second set of dyes, the experiment revealed role of size in anion exchange process where the order of size follows: orange G < Crocein scarlet 3B < acid black 1 < methyl blue. It was found that time taken for dye removal directly followed the order of size, i.e., bigger dyes took higher time and vice-versa. Furthermore, Sun and coworkers demonstrated rapid separation of anionic methyl orange with Mn-based MOF with molecular formula of Mn(H2L)2(H2O)2(H2bibp), where H4L: thiophene-2-phosphonic acid, bibp: 4,4ˊ-bis(1-imidazolyl)biphenyl.[30] This three dimensional Mn-MOF was fund to be consist of square channel with aperture of 11.6x11.6 Å2. In order to evaluate the ion exchange study, a set of dyes, which included safranine O (SO), CV, MB, rhodamine B (RB), rhodamine 6G (R6G) and MO, was chosen. A rapid capture of MO was observed with Mn-MOF. In addition, selectivity study was performed with binary mixture of MO/RB, MO/SO, MO/CV, MO/MB, and MO/R6G, where the MOF was found to be selective for the methyl orange compound. This study

ACS Paragon Plus Environment

20

Page 21 of 87 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

was further repeated in column chromatographic technique which was filled with the MOF and here also clear separation of MO was observed. Moreover, till now MOF based capture of anionic dyes have been discussed, whereas, porous organic materials have also been employed for the capture of anionic dyes in recent years. In a recent report, Jiang and co-workers demonstrated removal of methyl orange with crystalline cationic covalent organic framework (COF). The dye adsorption studies were carried out even in wide range pH-solutions where the performance was found to be maintained. Recyclability of the COF showed the efficiency of the material for real time applications.[31] In another report by Yu et al., a viologen based cationic COF was utilized for the capture of anionic dyes from solutions.[32]

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 87

Figure 4. (a) Schematic representation for the synthesis of MOF; UV-Vis spectra of methanolic dye solutions in presence of MOF at different time intervals for (b) MO and (c) MO-MB. Reproduced with permission from Ref. [28] Copyright 2015 Royal Society of Chemistry. Apart from the above mentioned examples, several other reports are available in literature for the dye capture with MOFs and porous organic materials.[33-45] Pharmaceutical products and personal care products. Water pollution due to pharmaceutical wastes has also become another type of major concern in recent years. Pharmaceutical products represent a widely used medicinal products used by human and other veterinary. Pharmaceutical wastes have been considered as emerging water pollutants globally. Reports showed that in 2003 in Italy approximately 1500 tons of pharmaceuticals were used and upto 95% of the doses were excreted through stool or urine to the wastewater.[46] There are various types of pharmaceutical product known, among them antibiotics, antinflammatory, antimicrobials etc. are well known. Furthermore, antibiotics have been considered as one of the major water pollutants among pharmaceutical products, as it has been found to be used in large quantity for several decades in the treatment of both human and animals. More than 250 different types of antibiotics have been registered till now for therapeutic treatments.[47] Huge production of such antibiotics and consequently release in various sewages have already made them one of the emerging water pollutants. Recent reports revealed that in last

ACS Paragon Plus Environment

22

Page 23 of 87 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

ten years the usage of antibiotics was found to be increased by 30% and in future the scenario will be even worst due to increase in population and industrializations.[48] In addition, estimation showed that over 1,62,000 tons of antibiotics was used in China in 2013 and among this over 50,000 tons was discharged to wastewater which contributed to the water pollution.[49] Water pollution with antibiotics finally leads to the accumulation of these pharmaceutical products in human body which further leads to the growing resistance of different bacteria towards antibiotics. As a result of reduced efficacy of antibiotics, approximately 25,000 people has been found to die every year in European Union and globally the number has reached to 5,00,000 per year.[50] Due to such threatening concerns over the water pollution with pharmaceutical products, in recent years a lot of researches have been devoted to the sensing and capture of these hazardous wastes from water with porous materials. In a recent study, Wang et al. demonstrated sensing and removal of antibiotics with two fluorescent MOFs, namely BUT-12 and BUT-13 (where, BUT: Beijing University of Technology).[51] Both the three dimensional frameworks of BUT-12 (Zr6O4(OH)8(H2O)4(CTTA)8/3) and BUT-13(Zr6O4(OH)8(H2O)4(TTNA)8/3) constructed from Zr-ion and respective organic linkers [H3CCTA: 5′-(4-carboxyphenyl)-2′,4′,6′-trimethyl[1,1′:3′,1″-terphenyl]-4,4″-dicarboxylic acid for BUT-12 and H3TTNA: 6,6′,6″-(2,4,6trimethylbenzene-1,3,5-triyl)tris(2-naphthoic acid) for BUT-13] were found to be stable in water medium (Fig. 5a). In this study, total twelve antibiotics of different five classes

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 87

were employed in aqueous phase, viz. nitroimidazole (NMs) [ronidazole (RDZ), metronidazole (MDZ), dimetridazole (DTZ), ornidazole (ODZ)], nitrofurans (NFs) [furazolidone (FZD), nitrofurazone (NZF), nitrofurantoin (NFT)], chloramphenicols (CPs) [chloramphenicol (CAP), thiamphenicol (THI)], sulfonamides (SAs) [sulfadiazine (SDZ), sulfamethazine (SMZ)] and β-lactams [penicillin (PCL)]. Both BUT-12 and BUT-13 showed emission maxima at 381 nm and 399 nm respectively, which were attributed to the emission of the respective organic linkers. This emission of both MOFs got quenched upon addition of the aforementioned antibiotics and for the nitro-derivatives high quenching was observed (Fig. 5b-5c). Maximum quenching was observed for NZF and NFT in case of both BUT-12 (92% for NZF and 91% for NFT) and BUT-13 (95% for NZF and 94% for NFT). The rapid quenching of the MOFs also reflected in the Stern-Volmer (Ksv) constant, where Ksv values for NZF and NFT were found to be 1.1×105 and 3.8×104 M-1 for BUT-12 and 7.5×104 and 6.0×104 M-1 BUT-13 respectively with detection limit (LOD) of 58 and 90 ppb for NZF with BUT-12 and BUT-13 respectively. Further, reversibility was also checked in this case and was found that up to six cycle both BUT12 and BUT-13 showed almost unchanged sensing performance. Apart from detection of antibiotics, both the MOFs were further employed for the adsorption of NMs, NFs, SAM and CAP. In another study, Hou and co-workers showed determination and adsorption of sulphonamide (SA) derivatives based on solid phase extraction (SPE) with MIL-101(Cr) MOF (MIL: Materials Institute Lavoisier) and coupled with UPLC-MS/MS.[52]

ACS Paragon Plus Environment

24

Page 25 of 87 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

This MOF based SPE technique was used as sorbent for the different SA-antibiotics like sulfadiazine (SDA), sulfamethoxazole (SMX), sulfachloropyridazine (SCP) and sulfamethazine (SMZ) from various environmental water samples.

Figure 5. (a) Structures of BUT-12; (b) Emission spectra of BUT-12 in presence of NZF; (c) Extent of fluorescence quenching of BUT-12 (blue) BUT-13 (pink) in presence of different antibiotics. ( color code: C: black; O: red; and Zr: green; H atoms are omitted for clarity; the large pink and yellow spheres represent cage void regions inside the MOFs) Reproduced with permission from Ref. [51] Copyright 2016 American Chemical Society.

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 87

Interaction between host porous materials and guest molecules also played crucial role in removal of antibiotics from aquatic systems. Jin et al. reported two threedimensional amide-based MOFs constructed from Zn- and Cu-ion, namely PCN-124-stu ([M2(PDBAD)(H2O)]n where, H4PDBAD: 5,5’-(pyridine-3,5-dicarbonyl)bis(azanediyl)diisophthalic acid, and M= Zn and Cu).[49] Owing to high porosity and stability, PCN-124stu(Cu) was used further for the capture of fluoroquinolones based antibiotics from water. Fluoroquinolones accounted as the third highest category of the synthetic antibiotics and in this work ofloxacin (OFL), enrofloxacin (ENR) and norfloxacin (NOR) were used as representative example of this category. PCN-124-stu(Cu) showed excellent removal performance and the capacities were calculated as 198 mg g-1, 292 mg g-1 and 354 mg g-1 for ENR, OFL and NOR respectively. Theoretical studies revealed that such high performance was accounted for the open metal sites present in the PCN124-stu(Cu) which interacted with the guest antibiotic molecules. Recyclability of the MOF for the capture of fluoroquinolones was checked and up to four cycles the respective antibiotics. Furthermore, in an earlier study Jhung and co-workers showed the role of H-bonding in the sequestration of pharmaceuticals and personal care products from water with MOFs.[53] In this context, MIL-101(Cr) (Cr3O(F/OH)(H2O)2[C6H4(CO2)2]) and its different functionalized derivatives (MIL-101-OH, MIL-101-(OH)2, MIL-101-NH2, MIL-101-NO2) were employed for the capture of naproxen (anti-inflammatory drug), ibuprofen (anti-inflammatory drug) and oxybenzone (used in sunscreen lotions) from

ACS Paragon Plus Environment

26

Page 27 of 87 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

water. Although surface area of the MOF was found to get reduced upon insertion of various functionalities, but due to H-bonding ability of hydroxy- and aminefunctionalized MOFs, MIL-101-OH(Cr), MIL-101-(OH)2(Cr) and MIL-101-NH2(Cr) performed better over parent MIL-101(Cr); especially, hydroxy-functionalized MIL101(Cr) showed excellent performance. On the other MIL-101-NO2(Cr) registered least performance as it lacked H-donor group and also pore got blocked due to functional groups. Moreover, Jhung and co-workers contributed few more works on capture of pharmaceuticals and personal care products from aquatic systems with MOFs.[39-41, 54-56]

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 87

Figure 6. (a) Dye exchange in the 3D framework of the MOF; (b) Emission spectra of the dye encapsulated MOF in presence of NZF; (c) Quenching efficiency of different nitro-derived antibiotics with dye encapsulated MOF. Reproduced with permission from Ref. [60] Copyright 2018 Royal Society of Chemistry. Further, Han et al. demonstrated sensing of ornidazole (ODZ) (another well known antibiotic) in water with a luminescent sodium-europium cluster based MOF, viz. CTGU-7 (CTGU: China three gorges university) with formula of [NaEu2(TATAB)2(DMF)3]·OH (H3TATAB: 4,4′,4′′-s-triazine-1,3,5-triyltri-m-aminobenzoic acid, DMF: N,N′-dimethylformamide).[57] A set of twelve commonly used antibiotics were employed and among them ODZ showed a drastic quenching in the emission profile of the MOF. The LOD of CTGU-7 was calculated to be 0.8 μM for ODZ and one of the reason behind this quenching behavior was explained in terms of energy-transfer process. Liu et al. reported further tetraphenylethene (TPE) core based fluorescent ZnMOF [Zn4O(BCTPE)3, H2BCTPE: 4,4'-(1,2-diphenylethene-1,2-diyl)dibenzoic acid] where aggregation-induced emission (AIE) was observed.[58] This Zn-MOF showed emission at 499 nm upon excitation at 365 nm and incremental addition of nitro-functionalized antibiotics (nitrofurazone and metronidazole) led to the quenching of luminescence. 50 ppm of nitrofurazone and metronidazole led to the 93% and 50% quenching of emission of Zn-MOF. In addition, LOD of the MOF for metronidazole and nitrofurazone were calculated to be 0.6 and 0.1 ppm respectively. Photon-induced electron transfer

ACS Paragon Plus Environment

28

Page 29 of 87 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

(PET) was accounted as the main mechanism behind the quenching of luminescence of the MOF. In another report, sensing and removal of tetracycline antibiotic was demonstrated by Zhou et al. with Zr-MOF (PCN-128Y).[59] PCN-128Y showed drastic quenching in presence of tetracycline with a high Ksv value of 9.84×105 M−1. Furthermore, not only MOFs but also guests inside the frameworks can also play crucial role in sensing of antibiotics. In this context, Fu et al. reported detection of antibiotics with a dye encapsulated MOF in water medium.[60] Through anion exchange method HPTS (8-hydroxy-1,3,6-pyrenetrisulfonicacid trisodium salt) dye was incorporated to the cationic framework of MOF with formula of [Zn(TIPA)(NO3ˉ)2(H2O)]∙5H2O [where, TIPA: (tri(4-imidazolylphenyl)amine)] (Fig. 6a). HPTS encapsulated MOF showed dual emission at around 375–450 and 500–575 nm which was caused by pyrene moiety of the dye. Upon 60 ppm addition of nitrofurazone (NZF), furazolidone (FZD) and dimetronidazole (DTZ) to aqueous medium of dye encapsulated MOF dispersed phases around 90% quenching was observed in each cases (Fig. 6b-6c). The quenching constants for NZF, DTZ and nitrofurantoin (NFT) were found to be 1.72×104, 1.72×104 and 1.01×104 M-1 respectively. Photo-induced electron transfer (PET) as well as fluorescence resonance energy transfer (FRET) both were accounted for the luminescence quenching in presence of antibiotics. Apart from antibiotics, anti-inflammatory compounds are another class of important and widely used pharmaceuticals. In a recent report, Yun and co-workers demonstrated removal of nonsteroidal anti-inflammatory drugs (ibuprofen, naproxen,

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 87

furosemide, salicylic acid, ketoprofen, acetophenone, indomethacin and amitriptyline) from water with Zr-MOFs.[61] UiO-66 [Zr6O4(OH)4(BDC)5; where, BDC: benzene-1,4dicarboxylic acid], MOF-808 [Zr6O4(OH)4(BTC)2; where, BTC: benzene-1,3,5-tricarboxylic acid] and MOF-802 [Zr6O4(OH)4(PZDC)5; where, PZDC: 3,5-pyrazoledicarboxylic acid] (UiO: University of Oslo) were employed for the capture of anti-inflammatory pharmaceuticals from water (Fig. 7a). Among these three MOFs UiO-66 and MOF-808 performed well, especially UiO-66 showed excellent performance towards antiinflammatory drugs (Fig. 7b). Defect sites inside the MOFs due to incompletecoordination on Zr-based cluster and π-π interactions were accounted as the reason behind such high removal performances with UiO-66 and MOF-808.

ACS Paragon Plus Environment

30

Page 31 of 87 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 7. (a) Structure of UiO-66, MOF-808 and MOF-802 (Color code: cyan: Zr; red: O; blue: N; gray: C; white: H; Terminal −OH/−OH2 groups and some H-atoms are omitted for clarity. The green and blue balls indicate the space in the framework); (b) Bar diagram represents adsorption capacities of anti-inflammatory pharmaceuticals in UiO66, MOF-808, and MOF-802. Reproduced with permission from Ref. [61] Copyright 2018 American Chemical Society. Apart from MOFs, porous organic materials also recently employed for the removal of pharmaceutical products from water. Mellah et al. recently reported capture of ibuprofen with a fluorine rich covalent organic framework (COF), viz. TpBD-(CF3)2.[62]

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 87

This COF was synthesized from 3,3ˊ-bis(trifluoromethyl)benzidine (BD-(CF3)2) and triformylphloroglucinol (Tp) via Schiff base condensation reaction. TpBD-(CF3)2 was found to capture ibuprofen efficiently with a high capacity value of 119 mg g-1 and further adsorbent was found to be recycled for several cycles (Fig. 8a). In addition, pHdependent studies revealed that TpBD-(CF3)2 able to scavenge ibuprofen even at lower pH due to the facile interaction between the drug molecules and COF, while at alkaline pH the performance dropped down drastically. In another report, Liu et al. reported a magnetic hypercross-linked polymers (MHCPs) were employed for the capture of antibiotics from water (Fig. 8b).[63] Removal of chloramphenicol (CAP) and tetracycline hydrochloride (TC) antibiotics (known to inhibit Gram-positive and Gram-negative bacteria) were studied and adsorption capacities for TC and CAP 212.77 and 114.94 mg g-1 respectively. MHCPs were used in column chromatography for breakthrough separation of antibiotics from water and further regenerated with diluted HCl. Moreover, apart from the aforementioned studies a few more works on sensing and sequestration of pharmaceuticals and personal care products with MOFs and porous organic compounds were also reported in literature.[39-41, 64-67]

ACS Paragon Plus Environment

32

Page 33 of 87 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 8. (a) Structure of the TpBD-(CF3)2 COF and the respective pharmaceutical molecules were studied; (b) Schematic representation for the synthesis of magnetic HCPs (MHCPs). For a) Reproduced with permission from Ref. [62] Copyright 2018 John Wiley & Sons; for b) Reproduced with permission from Ref. [63] Copyright 2018 American Chemical Society.

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 87

Toxic agrochemicals. Increasing global population has led to the enhanced demand for the food and consequently demand for improved agriculture has got increased. Recent statistics showed that the modern agriculture almost feed over 6000 million people.[68] To improve the productivity and quality of the agriculture, different types of agrochemicals are being used in huge quantity. When these untreated agrochemicals are reaching various water bodies through water sewages, they are causing immense threat to the living system. Different types of herbicide or pesticide are being considered as one of the major and emerging toxic water pollutants. Sensing and sequestration of such hazardous agrochemical waste with MOFs has attracted much concern in worldwide. 2,4-Dichlorophenoxyacetic acid (2,4-D) is a widely used acidic herbicide and it is known for its high toxicity and carcinogenic nature. Because of its high toxicity, the world health organization (WHO) suggested the allowed limit of 2,4-D in drinking water to be 70 g L-1. Jhung and coworkers studied the removal of the toxic 2,4-D from water with a well known Cr-terephthalic acid based MOF, i.e., MIL-53(Cr) with formula of Cr(OH)(C6H4(CO2)2)∙nH2O.[69] MIL-53 showed excellent removal performance of 2,4-D and it was found to be better than activated carbon and zeolite USY. Adsorption capacity of MIL-53 was found to be 556 mg g-1, whereas capacities of activated carbon and USY were found to be 286 and 256 mg g-1 respectively. MIL-53 showed rapid capture of toxic 2,4-D even at low concentration which known to be useful for the real

ACS Paragon Plus Environment

34

Page 35 of 87 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

time applications. In addition, the porous material was recycled for three times without losing any significant efficiency. The adsorption mechanism was explained in terms of electrostatic interactions which was based on the zeta potentials of the MOF material. The same research group further explored the adsorptive removal study of another highly toxic acidic herbicide, namely methylchlorophenoxypropionic acid (MCPP).[70] Zrbenzenedicarboxylate MOF, namely UiO-66, was used here as the adsorbent material and the efficiency of the MOF was compared with activated carbon. Efficient removal of MCPP was obtained with UiO-66 even at low concentration of the analyte and the adsorption isotherm followed pseudo-second-order non-linear kinetic model. Adsorption capacity for MCPP with UiO-66 was observed to be 370 mg g-1 which was higher than that of with activated carbon (303 mg g-1). In this study also, the adsorption mechanism was explained in terms of electrostatic interaction (in case of lower pH) as well as π-π interaction (in case of higher pH). The reusability of UiO-66 was obtained by washing with ethanol-water and later used for the next batch adsorption study of MCPP.

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 87

Figure 9. (a) Crystal structure of NKU-101; (b) UV-Vis spectra of methyl viologen solution in presence of NKU-101 at different time intervals. Reproduced with permission from Ref. [72] Copyright 2015 Royal Society of Chemistry. Furthermore, dimethyl viologen (MV) or paraquat (PQ) and diquat (DQ) are also well known cationic toxic herbicides and also causing threat to the environment due to immense water pollutions by these kind of pollutants. In this regard, Hill and coworkers investigated molecular sieving effect of Ni-PTC MOF ([Ni3(μ3-O)(H2O)3(PTC)1.5]∙xH2O; where PTC: perylene-3,4,9,10-tetracarboxylate) to capture toxic methyl viologen.[71] Moreover, Bu and coworkers recently reported an anionic framework based Zn-MOF, viz. NKU-101 (NKU: Nankai University) with molecular formula of [(CH3CH2)2NH2]1/2[Zn(BTC)2/3(PyC)1/4]∙solvent (H2PyC: 4-pyrazolecarboxylic acid; H3BTC: 1,3,5-benzenetricarboxylic acid), and further utilized for the capture of herbicide (Fig. 9a).[72] Owing to the anionic framework, NKU-101 was found to contain free [(CH3CH2)2NH2+ cation inside the network to balance the charge of overall material. Since free organic cation was present inside the pore of NKU-101, cation exchange phenomenon was employed here for the sequestration of cationic herbicides (PQ and DQ) in methanol medium (Fig. 9b). Again, sorption capacity of the NKU-101 compound was calculated to be 160 and 200 mg g-1 for PQ and DQ respectively and it was much higher as compare to activated carbon and zeolite materials. After adsorption of the herbicides, NKU-101 was again regenerated with NaCl in N,Nˊ-dimethylformamide

ACS Paragon Plus Environment

36

Page 37 of 87 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

(DMF) medium. Furthermore, in a recent study Du et al. demonstrated sensing of toxic methyl viologen in aqueous medium with a Tb-MOF ([(CH3)2NH2]9{Tb6(η6TATAT)4(H2O)12}∙3Cl∙DMA∙7H2O; where H6TATAT: 5,5ˊ,5ˊˊ-(1,3,5-triazine-2,4,6triyltriimino)tri-1,3-benzenedicarboxylic acid).[73] The Tb-MOF showed emission peaks at 489, 544, 583 and 621 nm upon excitation at 346 nm. Rapid quenching of luminescence of the Tb-MOF was observed in presence of MV2+ and detection of the paraquat analyte was obtained as low as 10-8 M concentration. Again, organophosphorus pesticieds (OPs) are another class of known highly hazardous and widely used organic compounds and acute toxicity of its can lead to organ failure and paralysis. MOFs based capture of such OPs attracted much attention in recent years. Recently, Zhu et al. demonstrated Zr-BPDC (BPDC: 4,4ˊbiphenyldicarboxylic acid) MOF, viz. UiO-67, based removal study of glyphosate (GP) and glufosinate (GF) which were taken as representative examples of OPs.[74] Rapid decrease in concentration of both GP and GF with time was observed in presence of UiO-67 which was measured with the help of ICP-AES. In addition, adsorption capacity of GF and GP with UiO-67 was found to be 360 and 537 mg g-1 respectively. In another study, Zhang and coworkers demonstrated capture of toxic MCPP compound with the help MOF-polymer composite monolith for better applicability.[75] Ice-templated two chitosan-UiO-66 composite monoliths were prepared by varying ratio of polymer and UiO-66, namely Chitosan/UiO-66-1 (1:1 weight ratio) and Chitosan/UiO-66-2 (1:2 weight

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 87

ratio of chitosan to MOF) (Fig. 10). Again, both the monoliths along with chitosan were soaked in cyclohexane and then air-dried, after this treatment all the materials were denoted as Chitosan-W, Chitosan/UiO-66-1-W and Chitosan/UiO-66-2-W. Both the composite monoliths were then employed for the decontamination of MCPP from aqueous medium. Chitosan/UiO-66-2-W represented a better efficiency as compare to the Chitosan-W and Chitosan/UiO-66-1-W. Adsorption capacity for Chitosan/UiO-66-2W was found to be 337.83 mg g-1and this was reused for even after 3rd cycle. Furthermore, along with these aforementioned studies, a few more examples also reported in literature.[39-41, 76]

Figure 10. (a) Photographs of Chitosan-W, Chitosan/UiO-66-1-W, Chitosan/UiO-66-2-W monoliths; (b) and (c) removal study of MCPP with Chitosan-W, Chitosan/UiO-66-1-W,

ACS Paragon Plus Environment

38

Page 39 of 87 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

Chitosan/UiO-66-2-W and UiO-66 (from left to right). Reproduced with permission from Ref. [75] Copyright 2017 American Chemical Society.

Veterinary feed additives and artificial sweeteners. Recent studies have revealed the presence of different types of artificial sweetener in groundwater as well as in various wastewater sewages which are causing several environmental issues. Due to wide range of applications of such artificial sweeteners in foods or beverage industries etc. for several decades, these organic compounds are causing water pollution. On the other hand, veterinary feed additives are another emerging class of organic pollutants which includes roxarsone, p-arsanilic acid (ASA), phenylarsonic acid (PAA) etc. Upon degradation, such feed additives lead to the formation of toxic and hazardous arsenic pollutants in water medium. As a consequence of this, removal of such artificial sweeteners and veterinary feed additives has drawn special attention. Jhung and coworkers reported capture of p-arsanilic acid (ASA) from water with a well known MOF, zeolitic imidazolate framework-8 (ZIF-8).[77] ASA used worldwide as a feed additive for broiler feeds to promote the growth, improve the feed efficiency, prevent dysentery etc. Mesoporous ZIF-8 showed very high adsorption capacity for ASA from water, i.e., 791.1 mg g-1 with a pseudo second order kinetic constant of 9.6x10-3 g mg-1 h-1. This MOF was found to be recycled up to three cycles without losing any

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 87

efficiency of its. In another report by the same group, removal of organic arsenic acids (OAAs) like ASA and phenylarsonic acid (PAA) was demonstrated via non-covalent interactions between MOFs and guest molecules.[78] To perform the targeted study, hydroxyl groups were inserted in a Cr-MOF, MIL-101 to synthesize MIL-101(OH) and MIL-101(OH)3 via aminoalcohol grafting. Removal performance for both the pollutants at equilibrium was found to be following the order, MIL-101(OH)3 > activated carbon > MIL-101(OH) > MIL-101. MIL-101-(OH)3 showed excellent performance and capacity values for ASA and PAA were found to be as high as 238 and 139 mg g-1 respectively. Hbonding interactions between MOF and the guest organo arsenic acids was accounted as the plausible mechanism behind such high adsorption. Further, Jhung and coworkers investigated removal of sweeteners from water by using the same Cr-based MOF, MIL101.[79] As similar to the previous study, MIL-101 was post-synthetically modified to incorporate urea or melamine via grafting to the open metal sites of the MIL-101. Removal of artificial sweeteners (ASWs) like saccharin (SAC), cyclamate (CYC) and acesulfame (ACE) from water was studied with these urea or melamine grafted MOFs and compared with activated carbon, parent MOF and NO2- group modified MOF, i.e., O2N-MIL-101. Although surface area of urea-MIL-101 and melamine-MIL-101 reduced as compare to the parent MOF, but capture performance of both the compounds were found to be much better as compare to parent MIL-101. The order of performance for the saccharin removal from water was found to follow the order, urea-MIL-101 >

ACS Paragon Plus Environment

40

Page 41 of 87 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

melamine-MIL-101 > MIL-101 > activated carbon > O2N-MIL-101. This high performance of urea or melamine grafted MIL-101 was accounted for the possible Hbonding interactions between the NH2-group and artificial sweetener molecules (Fig. 11a). As similar to the saccharin, removal of cyclamate and acesulfame with modified MIL-101 also followed the similar trend. Reusability of adsorbents were checked and up to third cycles it was found that removal efficiency was almost retained.

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 87

Figure 11. (a) Plausible mechanism for the adsorption of respective sweeteners saccharin, acesulfame, and cyclamate over urea grafted MIL-101 (from left to right). (b) Adsorption mechanism of p-ASA in defect sites of UiO-67-NH2. Fig. 11(a) Reproduced with permission from Ref. [79] Copyright 2016 American Chemical Society. Fig. 11(b) Reproduced with permission from Ref. [80] Copyright 2018 American Chemical Society. In another study, Tian et al. reported capture of ASA from water medium with amine-modified MOF.[80] UiO-67, UiO-67-NH2 and UiO-67-(NH2)2 were used for the capture feed additive ASA compound from aquatic medium (Fig. 11b). Further studies revealed that for the adsorption of ASA with UiO-67 and its derivatives three possible mechanistic contributors were proposed and the order was found to be as As−O−Zr coordination > π−π stacking > H-bonding. The capture performance was found to be recycled even up to four cycles. Furthermore, Gu and coworkers demonstrated the role defect in MOFs for the removal studies of such organic feed additives.[81] Roxarsone (ROX) is another well known organic arsenic compound used widely as feed additive. UiO-66 MOF was utilized to scavenge ROX from water via defect engineering inn MOF. Different amount of defects were created in the UiO-66, depending on equivalent of benzoic acid used during synthesis UiO-66 MOF was denoted as UiO-66-2, UiO-66-10 and UiO-66-20 here (2, 10 and 20 molar equivalent of benzoic acid was used respectively). Adsorption of ROX on UiO-66 series was observed to follow pseudo second order kinetics. Adsorption capacity of the ROX removal with UiO-66-20 was

ACS Paragon Plus Environment

42

Page 43 of 87 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

calculated as high as 730 mg g-1. This work represented the effect of defect engineering in a MOF for the removal of pollutants from water. Apart from these aforementioned examples a few more studied was carried out in this domain.[82]

Tackling oil spillage. Marine oil spillage has evolved as another great concern for the environment. Due to release of various toxic and harmful oil constituents into different water bodies, water pollution by oil spillage have attracted tremendous attention. In a recent report, Ghosh and co-workers reported an ultra-hydrophobic MOF which was synthesized from Cu-ion and bis(trifluoromethyl)-based carboxylate linker (Fig. 12a).[83] This MOF, namely UHMOF-100, was found to be very much hydrophobic in nature due to the presence of -CF3 groups inside the network. The water contact angle was found to be as high as 176 which was in the range of ultra-hydrophobic region and one of the highest value in the domain of MOFs. Further, UHMOF-100 was employed for the separation of oil from oil/water mixture. A composite membrane was again fabricated with the UHMOF-100 for the potential practical applications, viz., UHMOF-100/PDMS/PP (Fig. 12b). This membrane was further tested with different well known constituent of oil such as crude oil, hexadecane, biodiesel, CCl4 and toluene. In addition the performance was found to be recyclable even after 10 cycles with UHMOF-100/PDMS/PP composite material. In another work, Rao et al. demonstrated fabrication of bimetallic porous coordination polymers (PCPs) based high temperature stable super-hydrophobic

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 87

composite.[84] Here, the composite material was obtained by de Novo synthesis approach. Moreover, Omary and coworkers reported two new hydrophobic MOFs, viz., FMOF-1 and FMOF-2 (FMOF: fluorous metal-organic framework), which were synthesized from Ag(I)-ion and 3,5-bis(trifluoromethyl)-1,2,4-triazolate linker.[85] Spill oil clean up was further showed with these MOFs. In another work, Jiang et al. reported a MOF coated sponge material for the easy separation of oil from oil/water mixture.[86] In this study, a fluorinated MOF, namely USTC-6 (USTC: University of Science and Technology of China) was synthesized from a -CF3 functionalized linker. The coating of MOF on graphene oxide modified sponge produced the composite sorbent material (USTC-6@GO@sponge) which was then used as the sorbent for oil. Another MOFcomposite based oil/water separation was demonstrated by Jayaramulu et al. where composite of ZIF-8 and highly fluorinated graphene oxide (HFGO) was fabricated.[87] The water contact angle for this composite material was found to be 162 whereas only ZIF8 and HFGO showed water contact angle of 56 and 125 respectively. Further, this composite coated sponge material was fabricated which showed excellent performance for the separation of oil.

ACS Paragon Plus Environment

44

Page 45 of 87 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 12. (a) Orthographic view of as-made UHMOF-100 along a-axis; (b) Pictorial representation (top) of PP (white) and UHMOF-100/PDMS/PP (green) and drop-cast of water droplets on UHMOF-100/PDMS/PP composite (bottom). Reproduced with permission from Ref. [83] Copyright 2016 John Wiley & Sons.

Apart from aforementioned examples, macro-porous materials based separation of oil-water are also demonstrated in few recent reports. Wan et al. demonstrated superhydrophobicity and superoleophilicity in a MoS2 coated melamine-formaldehyde sponge which was fabricated with room temperature vulcanized (RTV) silicone rubber. This MoS2@RTV sponge was found adsorb a wide variety of organic solvents, oils etc. with high selectivity and further the composite material showed excellent recyclability and high mechanical durability. In another report by Gao et al., superhydrophobic and superoleophilic MoS2 nano-sheet sponge (SMS) was employed for the oil-water separation application. Owing to the high porosity and superhydrophobicity, this

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 87

compound was found to be promising material for oil recovery and tackling marine oil spillage. Apart from only oil-water separation, MoS2 coated composite materials were also employed for organic pollutants degradation. In a recent report by Wan et al., MoS2 coated melamine-formaldehyde sponges were found to be efficient scavenger of water soluble dyes along with oil-water separation applications. This bifunctional material showed excellent adsorption capacities for different oils and fast removal of well known methylene blue and methyl orange dyes. Furthermore, in another report Li and coworkers demonstrated fabrication of superhydrophobic Cu-BiOBr surface which was used for the oil-water separation. In addition, this material was found to be efficient candidate for the dye degradation in water. Rhodamine B dye was degraded utilizing this material and even after five cycles of dye degradation, the material was found to be stable. Therefore, this type of advanced multifunctional porous materials are also very important for the decontamination of water pollutants.

Other organic contaminants. Apart from the aforementioned major classes of organic pollutants, there are many types of known organic pollutant mostly generated from different industries, household usages etc. Bisphenol A, different phenol derivatives (phenol,

2,4,6-trinitrophenol

etc.),

surfactants

(perfluorooctanesulfonic

acid,

perfluorooctanoic acid etc.), phthalic acid and its anhydride derivatives etc. are well known organic contaminants. In a recent report, Han et al. demonstrated capture of

ACS Paragon Plus Environment

46

Page 47 of 87 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

phenol from aqueous medium with MOF-based composite material.[88] Carbon nanotubes (CNTs) loaded CNT@MIL-68(Al) composites were prepared with different loading amount of CNT (0.15%, 0.75%, 2.5%, 3.5%, 7% and15%) and further CNT@MIL68(Al) were used for the adsorption of phenol. The 0.75% CNT@MIL-68(Al) composite showed enhanced adsorption capacity of 188.7% in comparison to the pristine MOF. In addition, the composite was found to be recyclable even up to five cycles. In another report, removal of another emerging organic contaminants (EOCs) phthalic acid and diethyl phthalate was shown with ZIF-8 by Jhung and co-workers.[89] Phthalates are well known endocrine disruption compounds and extensive use of such compounds as plasticizers in industries and domestic usage have led increase in water pollution. In this work, ZIF-8 showed an excellent performance for the respective capture studies with high capacity values which was found to be much higher as compare to activated carbon, UiO-66 and H2N-UiO-66. Furthermore, aromatic nitro-compounds like 2,4,6trinitrophenol (TNP), 2,4,6-trinitrotoluene (TNT), nitrobenzene (NB) etc. are well known explosive materials and environmental pollutants. Since TNP has high water solubility, consequently pollution in the aquatic system due to such nitro-derivatives has evolved as serious concern. Our group demonstrated sensing of TNP with different luminescent MOFs both in organic medium as well as water medium.[90-92]

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 87

Figure 13. (a) View of the SCU-8 structure along c-axis; (b) Top (above) and side (bottom) of the SCU-8 and PFOS for the simulation study. Reproduced with permission from Ref. [93] Copyright 2017 Springer Nature. Furthermore, various surfactant molecules are considered as one of the major EOCs and among them water pollution due to perfluorooctanesulfonic acid (PFOS) and perfluorooctanoic acid (PFOA) have attracted huge attention in recent days. Wang and co-workers reported a cationic Th-based MOF, namely SCU-8 with molecular formula of [Th3(bptc)3O(H2O)3.78]Cl·(C5H14N3Cl)·8H2O

(where,

H3bptc:

[1,1′-biphenyl]-3,4′,5-

tricarboxylicacid).[93] Owing to the presence of extra framework anions, SCU-8 was employed for the exchange of anions from water medium.(Fig. 13a) Capture of PFOS with SCU-8 was found to be having fast kinetics and high sorption capacity of 44.79 mg gˉ1. In addition, computational studies revealed the capture of PFOS with SCU-8 was

ACS Paragon Plus Environment

48

Page 49 of 87 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

driven by non-covalent interactions like hydrophobic interaction, hydrogen bonding, electrostatic interactions and van der Waals interactions at different stages of the incorporation. (Fig. 13b) In another study, Dichtel and co-workers reported amino functionalized crystalline covalent organic frameworks ([NH2]−COF) for the removal of per- and polyfluorinated alkyl substances (PFAS).[94] By varying loading amount of amine groups, a series of COFs were employed for the targeted application to check the role amine moieties. Amine loading of 20% and 28% to [NH2]−COF showed excellent efficiency for the capture of PFAS from water. Apart from these examples, MOFs and porous organic materials based capture of such organic hazardous compounds have been reported in the literature.[ 39-41, 95-96]

Degradation of organic pollutants with advanced porous materials. One of the initial reports documenting the photo-catalytic potential in a MOF was with MOF-5.[97] This cubic framework was crafted with Zn4O clusters located at the corner which were further connected orthogonally via terepthalate linkers. MOF-5 showed broad absorption band ranging from 500-840 nm that are attributed to the delocalized electrons having lifetime in microsecond time scale. The band gap value was calculated to be 3.4 eV. This MOF showed a comparable degradation activity of phenol in water medium to that of commercially available TiO2. The probable mechanism proposed by the authors suggest a similar mechanism as in the case of TiO2 where photocatalytic

ACS Paragon Plus Environment

49

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 50 of 87

degradation of phenol might proceed through chain reactions that involves a radical cation formation initially via electron transfer from phenol to the MOF-5 hole or formation of oxygen active species generated from the reaction of photoejected electrons and oxygen. Furthermore, this MOF showed an interesting observation of reverse shape-selectivity with various organic compounds. Garcia and co-workers found that larger phenolic molecules that were unable to diffuse freely into the micropores degraded at a significantly faster rate compared to the smaller ones having access to the inside cavity of MOF-5. They investigated the photo degradation of 2,6-di-tertbutylphenol (DTBP) and phenol (P) where DTBP was much bulkier compared to P molecule. Upon irradiating the both compounds individually under the UV light along with MOF-5 , DTBP and P degraded at a similar initial rate with a k(DTBP)pure/k(P)pure ratio of 1.1. However, when the same experiment was carried out in a solution containing both DTBP and P, DTBP degraded fairly quickly with a rate constant ratio i.e. k(DTBP)mix/k(P)mix value) 4.42 fold higher than that of P which implied around 82% selective degradation toward bulkier DTBP with respect to P. Decomposition of around 50% phenol was achieved after 180 min of illumination while 100% DTBP degraded after the same time. These findings were explained considering the size factor where smaller P molecule could easily diffuse inside MOF-5 accounting for lower degradation rate while larger DTBP remained on the surface of MOF-5 resulting in a higher degradation rate. Taking a cue from the findings of MOF-5 as an effective photo catalyst, Chen and

ACS Paragon Plus Environment

50

Page 51 of 87 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

co-workers synthesized a novel semiconducting doubly interpenetrated porous MOF Zn4O(2,6-ndc)3(DMF)1.5(H2O)0.5.4DMF.7.5H2O(UTSA-38),having high density of Zn4O sites in the framework (Fig. 14a-b).[98] The semiconducting MOF (band gap = 2.85 eV) displayed photocatalytic potential of degrading methyl orange (MO) in water. The authors noted that when irradiated with visible light, the absorption peak maxima at 464 nm of MO in aqueous solution gradually diminished with increasing time, indicating degradation of MO in the presence of the UTSA-38. Furthermore, when UV light was employed to initiate the photocatalytic reaction, the degradation rate of MO became remarkably faster. Under UV light irradiation for 120 minutes MO was entirely degraded to colorless small molecules illustrating the better efficiency of the UV light for photocatalytic degradation of MO compared to visible light (Fig. 14c-d). Additionally, post photocatalytic reaction UTSA-38 MOF catalyst could be easily retrieved through simple filtration with no obvious drop in the catalytic efficacy even after 7 cycles. The authors also speculated the mechanism for dye degradation included formation of electron–hole pairs within the UTSA-38.

ACS Paragon Plus Environment

51

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 52 of 87

Figure 14. (a) Structure of UTSA-38 representing coordination environment around Znion; (b) Representation of double interpenetrating cubic nets in UTSA-38; (c) UV-Vis profile of MO degradation by UTSA-38 at different time intervals; (d) Decolorisation of MO solution under UV light for 0 min, 20 min, 40 min, 60 min, 80 min, 100 min and 120 min (from left to right, respectively). Reproduced with permission from Ref. [98] Copyright 2011 Royal Society of Chemistry.

Li and co-workers reported three new metal–organic coordination polymers, [Cd3(5-NO2-bdc)2(5-NO2-bdcH)2(bpyo)2]n (1), [Mn(5-NO2-bdc)(bbim)]n (2) and {[Gd(5NO2-bdc)(5-NO2-bdcH)](bpyo)0.5.2H2O}n (3) (where (5-NO2-bdcH2 = 5-nitro-1,3benzenedicarboxylic acid; bpyo = 4,4' -bipyridine-N,N' -dioxide; bbim = 1,1' -(1,4butanediyl)bis(benzimidazole)) built from non-rigid

bridges

and probed the

ACS Paragon Plus Environment

52

Page 53 of 87 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

photocatalytic activity of compound 3 using X3B as model pollutant in water medium.[99] The Gd based compound 3 displayed photocatalytic degradation of X3B both under UV and visible light while better performance was observed for UV irradiation. Additionally control experiments were performed (in dark, without catalyst) to validate the role of light and active MOF catalyst for the degradation process. The authors speculated the underlying mechanism to be based on HOMO-LUMO approach accompanied by the formation of hydroxyl radical (∙OH) as the active species in the photocatalytic process for the degradation of X3B effectively. This study marks one of the rare reports in literature of exploiting a lanthanide based coordination polymer as a photocatalyst to degrade organic pollutant. Li and co-workers developed a novel bimetallic copper (I)/copper (II)salen CP, {[CuII(SalImCy)](CuII)2∙DMF}n(1, wherein SalImCy = N,Nˊ-bis-[(imidazol-4yl)methylene]cyclohexane-1,2-diamine)

that

was

utilized

for

visible

light-driven

photocatalytic degradation of MB, RhB and MO respectively in water.[100] The authors noted that 1 could degrade 65% of MB even without visible-light irradiation which they attributed to the catalytic nature of Cu (II) ions. However, up to 96% of MB degradation was achieved when the reaction was carried out in presence of visible-light which was speculated as a consequence of cooperative decomposition through photoactive Cu (I) sites present in the CP. Additionally >95% photocatalytic degradation was obtained for RhB and MO

upon

~50 min of light irradiation. The catalyst was recyclable and

maintained structural integrity after photocataysis. Yi and co-workers reported a rare 3

ACS Paragon Plus Environment

53

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

D

porous

molybdophosphate-based

FeII,III

{Na6(H2O)12[FeII2]2[FeIII4(PO4)][FeII(Mo6O15)2(PO4)8]2}(OH)3·33H2O(1)

Page 54 of 87

containing for

MOF selective

photocatalytic degradation of Rhodamine B (RhB).[101] 1 was composed of sandwichtype FeII[P4Mo6O31]2 clusters further connected to a large SBU and extended to porous 3 D network by FeII2 dimers having cross-shaped channels. Under visible-light illumination 1 was found to degrade 100% of RhB while only 10% photodegradation was achieved in case of MO reflecting the excellent selectivity in photocatalysis for FeMoP-MOF (1). Further, the catalyst was efficient even after four cycles and did not undergo any structural change after the catalysis. In a fascinating work Zhou and co-workers designed a unique and versatile High Valence Metathesis and Oxidation (HVMO) strategy to obtain a series of photoactive Ti containing MOFs having known topology and structural features.[102] The synthesized novel MOFs namely, PCN-333(Sc)-Ti, MIL-100(Sc)–Ti, MOF-74(Zn)–Ti and MOF-74(Mg)– Ti were examined for phocatalytic degradation of MB. When irradiated with a 300 W Xe lamp (9 minutes) all the Ti-MOFs exhibited better catalytic efficiency than TiO2 while MOF-74(Zn)–Ti and MOF-74(Mg)–Ti displayed an outstanding degradation up to 98% in just three minutes (Fig.15a). The authors attributed the superior performance of Ti-MOF74 to its ability to absorb a broader range of irradiated light. Additionally a plausible mechanism was proposed where Ti-MOF-74 behaved like a chromophore and electron transfer takes place from DOBDC linker to the Ti-oxo cluster upon excitation.

ACS Paragon Plus Environment

54

Page 55 of 87 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

Consequently Ti4+ gets reduced to Ti3+ separating the electron-hole couple while DOBDC upon oxidation is stabilized by generation of a benzoquinone moiety via sequential electron transfer. Finally the catalytic pathway is completed with the reduction of the benzoquinone species accompanied by the oxidation of Ti3+ by O2 (Fig. 15b). Wang and co-workers synthesized a MOF based photocatalyst NNU-36, [Zn2(BPEA)(BPDC)2]·2DMF (where BPEA = 9,10-bis(4-pyridylethynyl)-anthracene) and BPDC = 4,4′-biphenyldicarboxylate) constructed using a pillaring ligand.[103] Apart from efficient photocatalytic reduction of Cr (VI) NNU-36 was found to be an active photocatalyst for RhB degradation. However only 46.6% RhB degradation was observed after 70 minutes of visible-light illumination. The authors speculated this inefficiency as a result of very fast charge carrier recombination. Further H2O2 was added to facilitate the catalytic reaction and cut down the recombination process. Highest RhB degradation of 96.2% was achieved within 70 min upon addition of 1.0 M H2O2 solution. The H2O2 involved advanced oxidation process (AOP) was verified as a suppressed degrading efficiency (85.8% in 70 min) was observed with addition of hydroxyl radical scavenger tert-butyl alcohol (TBA). Also NNU-36 was found to achieve a H2O2 aided degrading efficiency of 94.2% (80 min) and 93.5% (90 min) for MB and R6G respectively. Following a different approach recently Lin and co-workers synthesized two unique bicontinuous donor-acceptor hybrid heterostructures namely [K6Ni2(BCEbpy)4(H2O)10(Pb9I30)] (1),[Co(BCEbpy)(H2O)4]∙[Pb3I8](2) (where, BCEbpy = N,N'-

ACS Paragon Plus Environment

55

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 56 of 87

bis(carboxyethyl)-4,4'-bipyridinium) that showed semiconductive and photocatalytic bahavior towards RhB degradation.[104] Viologen moiety acted as the acceptor while an electron rich anionic infinite 1D iodoplumbate nanowire (Pb9I30)n 12n– housed as the guest motif. The metal Viologen frameworks had a narrow band gap value of 1.89 and 1.83 eV for Ni and Co respectively. Both the heterostructures exhibited efficient photocatalytic activity in degrading RhB under irradiation of 300 W xenon lamp where nearly 100% of RhB degradation was achieved within 20 min for Ni compound while Co one needed 25 min for complete degradation. Jin and co-workers explored the photocatalytic properties of a biocompatible MOF {[Zn2(fer)2]∙0.5H2O}n (1)(fer = ferulic acid) using MV and RhB as model pollutants.[105] It was observed that upon illuminating with a Hg lamp (250 W) for 100 min 54% and 88% of degradation was attained for MV and RhB respectively. Further, in absence of the MOF catalyst only 8.3% and 5.8% conversion took place for MV and RhB. The catalyst was found to maintain its catalytic efficiency over five cycles while a probable mechanism was proposed by calculating the density of states (DOS).

ACS Paragon Plus Environment

56

Page 57 of 87 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 15. (a) MB degradation without catalyst (blank), TiO2, PCN-333(Sc)–Ti, MIL100(Sc)–Ti, MOF-74(Zn)–Ti and MOF-74(Mg)–Ti in presence of 300 W xenon light ; (b) plausible mechanism for MB degradation in air with Ti-MOF-74. Reproduced with permission from Ref. [102] Copyright 2016 Royal Society of Chemistry. Very recently Stylianou and co-workers came up with a “waste to wealth” strategy with photocatalytic H2 production via degradation of RhB simultaneously.[106] A series of co-catalyst/MIL-125-NH2 was systemically explored for photocatalytic H2 production via simple mixing procedure with inexpensive transition metal co-catalysts (Ni2P, NiO, CoP,

ACS Paragon Plus Environment

57

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

Page 58 of 87

Co3O4, Fe2O3, and CuO). The authors found that when using RhB (conc. 1.2 ppm) as an electron donor a maximum of 335 μmol hˉ1 gˉ1 H2 evolution could be achieved with Ni2P/MIL-125-NH2 catalyst. Also no significant change in RhB concentration was noticed when either Ni2P/MIL-125-NH2 or 1.2 ppm RhB was kept in absence of light or only 1.2 ppm RhB was irradiated in absence of MOF catalyst. Additionally, trapping experiments validated the role of RhB in photocatalytic H2 production where addition of tert-butanol (radical scavanger) and TEOA (hole scavanger) resulted in 11.2% and 18.7% degradation of RhB respectively compared to 45% with no scavenger being added. Notably this was the first report on MOF-based dual-functional photocatalyst that produces H2 and simultaneously degrades an organic pollutant in a single concurrent process. In another work Yang and co-workers reported a highly stable Cu (I) triazolate MOF (CuTz-1) crafted from 3,5-diphenyltriazolate for efficient photocatalytic degradation of organic dyes.[107] In presence of H2O2 and upon photoirradiation CuTz-1 could completely degrade MLB, MO, RhB and MB in 8, 15, 22 and 24 min respectively. A first-order kinetics regarding the photocatalytic degradation with rate constants (k) of 0.42, 0.17, 0.19 and 0.13 min-1 for MLB, MO, RhB and MB respectively was obtained. Additionally simultaneous degradation of a four component mixture of dyes was also possible with CuTz-1. The high efficacy of this MOF catalyst was attributed to its high efficiency in generating hydroxyl radicals (∙OH) that plays the most significant role in dye degradation. Also, CuTz-1 was found to be stable and efficient event after four catalytic

ACS Paragon Plus Environment

58

Page 59 of 87 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

cycles. Very recently Sicard, Ricoux and co-workers utilized an innovative approach for biodegradation of pollutant dye molecules by encapsulating a mini enzyme Microperoxidase-8 (MP8) into a mesoporous MIL-101 (Cr) MOF.[108] The authors found that the biocatalyst with immobilized with peroxidase-type enzyme displayed enhanced catalytic degradation for MO when compared to free MP8. This was attributed to the electrostatic attractive interactions between MP8@MIL-101 and negatively charged MO while repulsive interactions were expected for MB. Additionally, MP8 embedded within MOF matrix showed greater resistance toward harsh conditions due to the confinement effect. Notably this work demonstrated the first example of synergy of a MOF matrix and an enzyme for enhanced catalytic performance which was attributed to the preconcentration of reactants based on charge matching. To mitigate water contaminants, Wang and co-workers employed a water stable MOF, bio-MOF-11-Co for the catalytic degradation of pharmaceuticals and personal care products (PPCPs).[109] The authors studied the degradation of sulfachloropyradazine (SCP) and para-hydroxybenzoic acid (p-HBA) as model compounds of pharmaceuticals and personal care products in presence of peroxymonosulfate (PMS) via AOP aided catalytic degradation. 100% degradation of both SCP and p-HBA was observed only within 20-30 min for bio-MOF11-Co in presence of PMS while 70% of SCP was removed using only PMS in 40 min. Further, other oxidative agents such as H2O2 (40%) and PDS (25%) showed lesser degrading efficiency of SCP. Different concentrations (15-60 mg/L) of SCP and p-HBA

ACS Paragon Plus Environment

59

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 60 of 87

was used to test the catalytic efficiency of bio-MOF-11-Co where the catalyst (50 mg/L) was able to remove 100% of p-HBA (50 mg/L) in 45 min while it only took 15 min for complete SCP removal. Also 100 % removal of SCP and p-HBA was achieved in just 5 min at 45 °C. A mechanistic insight involving the generation of both SO42-∙ and ∙OH radicals was also proposed. Han and co-workers demonstrated the utility of sulfidation process in MOFs for photodegradation of organic pollutants.[110] They employed MIL68-In as a template to obtain In2S3 nanorods which was treated for catalytic degradation of water contaminant tetracycline hydrochloride (TC) and methyl orange (MO). With a band gap of 2.54 eV the In2S3 exhibited the capability of harvesting visible light and act act as an active photocatalyst. Following a pseudo-first-order kinetics In2S3–8h displayed best photocatalytic degradation of MO with 97% degradation being attained within 120 min. Also 66% degradation of 200 mL TC (50mg/L) was observed in In2S3–8h (60 mg) upon illuminating visible light for 120 min. Moreover the catalyst was found to be stable and efficient over three catalytic cycles. Additionally a plausible mechanism operating through radical generation was also discussed. Redel and co-workers documented a novel synthetic stratigy by successfully fabricating Bi2O3@HKUST-1 porous thin films that was illustrated for photodegradation of NFR dye molecules.[111] The semiconducting Bi2O3 nanoparticles (diameter 1-3 nm) embedded within HKUST-1 SURMOF demonstrated enhanced photo-efficiencies compared to conventionally synthesized NPs. Upon UV-light irradiation of 255 nm the

ACS Paragon Plus Environment

60

Page 61 of 87 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

hybrid-photocatalyst exhibited 100% degradation of nuclear fast red (NFR) dye while pure HKUST-1 and Bi2O3 could achieve 44% and 74% degradation respectively under similar conditions. Moreover the NP@SURMOF was stable, reusable and did not lose its catalytic efficacy after at least over four catalytic cycles. The authors speculated the superior activity of the composite on the account of higher surface area of the NP@MOF and suggested that confinement effect in pores of the HKUST-1 MOF was able to further stabilize the small Bi2O3 NPs while inhibiting sintering/clustering of the NPs. The efficient charge-separation was suspected to be taking place in a similar way already established for QD@MOF systems. Zhang and co-workers reported a series of well dispersed and highly stable conjugated microporous nanoparticles that displayed photocatalytic RhB degradation.[112] Combined with of high porosity and increased solution dispersibility these metal free catalysts exhibited good efficiency in degradation of RhB using 23 W household energy-saving bulbs as a light source. Among the CMP NPs, B-BT3-b(constructed from 1,3,5-Triethynylbenzene and 4,7-dibromobenzo[c]-1,2,5thiadiazole) having the lowest optical band gap ( 1.76 eV) displayed highest degradation rate of RhB (80%) in just 25 min while B-BT3-a(constructed from 1,3,5-phenyltriboronic acid and 4,7-dibromobenzo[c]-1,2,5-thiadiazole) with band gap of 1.96 eV could only convert about 50% of RhB after the same period of time. Interestingly the authors noted that B-BPh3-b (constructed from 1,3,5-Triethynylbenzene and 4,4’-dibromobiphenyl) irrespective of its larger band gap (2.36 eV) showed slightly better efficiency in the

ACS Paragon Plus Environment

61

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 62 of 87

starting minutes than B-BT3-a which they ascribed to high BET surface area (512 m2 g-1) leading to large reactive interface. Additionally, control experiments with various scavengers were performed in order to establish the role of superoxide (*O2 -), singlet oxygen (1O2), .OH and the photogenerated hole (h+) in the photocatalytic degradation process. Cai and co-workers further contributed to the field with a series of randomly copolymerized n-type porous conjugated polymers using 2,5,8,11-tetrabromo perylene diimide (PDI–Br4) and 1,2,4,5-tetrabromobenzene with linear linkers 4,4′-diethynyl-1,1′biphenyl (bph) and 5,5′-diethynyl-2,2′- bipyridine (bpy) and explored their photocatalytic behavior towards degradation of methylene blue (MB) dye.[113] The band gap energy of these PDI (Perylene diimides) based PCPs were reported to vary from 1.54 to 2.25 ev. Photocatalytic studies disclosed PCP2-100%PDI to be the best performing catalyst while PCP1-100%PDI also demonstrated good activity whereas other PCPs exhibited negligible towards MB degradation. The authors noted that dye degradation rate enhances with decrease in the HOMO energy levels of the PCPs implying the photodegradation might be occurring through the photogenerated holes in the PCPs. Conclusions and future outlook. The pressing need to sequester environmental pollutants has called for development of novel materials or methods. In particular, the growing accumulation of organic pollutants in different water streams which cause adverse ecological impacts has caused a global concern. Porous solids having high surface areas, functioning as sorbents have emerged as frontrunners in this regard.

ACS Paragon Plus Environment

62

Page 63 of 87 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

MOFs, COFs and porous organic polymers have scored over conventional porous compounds owing to feasible accessibility to tuning pore size, shape and functionalities. In particular, the emergence of water stable MOFs has actuated research towards practical applications, as there is greater understanding of the strategies which lead to formation of stable compounds. Apart from hydrolytic stability it is essential for MOFs to present stability in diverse chemical environments for seeking practical realization. While COFs and POPs are relatively less explored, their superior stability should propel the research more rapidly. In addition to capture, the access to transduction pathways has resulted in these compounds being developed as sensory materials to read-out the inclusion of targeted species. The ability to tune electronic features has resulted in these materials being substrates for the degradation of bulky organic molecules. In addition, several materials have been fabricated employing eco-friendly surface modification processes that have found extensive applications in wastewater treatment and various other domains. Especially, polymer-grafted membranes as well as coating serves as an excellent methodology mostly for oil-water separation. Although the different class of compounds has shown affirmative responses, there remain challenges of bulk-scale implementation. These include preparation of workable forms, scalability, cost effectiveness, environmental toxicity and ability to perform over several cycles. Significant research attention has been devoted to address these hurdles and

ACS Paragon Plus Environment

63

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

Page 64 of 87

consequently the applicability towards capture and remediation of environmental pollutants is of high relevance for future research. AUTHOR INFORMATION Corresponding Author * Dr. Sujit K. Ghosh, Indian Institute of Science Education and Research (IISER), Dr. Homi Bhabha Road, Pashan, Pune-411008, India. E-mail: [email protected]. Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. ‡These

authors contributed equally.

ACKNOWLEDGMENT P.S. and S.L. acknowledges UGC (India) and CSIR (India) for fellowship respectively. A.D. is thankful to IISER Pune for fellowship. S.K.G. acknowledges funding supports from DST-SERB project (EMR/2016/000410) and DST Nanomission Thematic Unit, Govt. of India.

BIOGRAPHIES

ACS Paragon Plus Environment

64

Page 65 of 87 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

Partha Samanta did his BSc from the University of Calcutta in the year 2011 and then went to pursue his MSc from IIT-Kharagpur. He joined the research group of Dr. Sujit K. Ghosh in the year 2013 and since then he has been exploring various porous organic materials toward energy and wide variety of environmental applications.

Dr. Aamod V. Desai completed his BSc from University of Pune (2011). Thereafter he joined IISER, Pune as an integrated-PhD student. After completing MS from IISER (2013), he joined the research group of Dr. Sujit K. Ghosh to pursue his PhD degree which was completed in 2018. Dr. Desai's thesis has involved the development and exploration of functional MOFs for recognition and capture of environmental pollutants.

ACS Paragon Plus Environment

65

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 66 of 87

Sumanta Let received BSc from University of Calcutta (2014), after which he enrolled at the Indian Institute of Technology (ISM), Dhanbad for MSc. Upon completion of MSc (2016), he enrolled in the doctoral program at IISER Pune under the supervision of Dr. Sujit K. Ghosh. Sumanta's research revolves around the design and development of function-led porous materials for remediation of pollutants present in air and water.

Dr. Sujit K. Ghosh completed his PhD from the Indian Institute of Technology (IIT), Kanpur in 2006 under the supervision of Prof. Parimal K. Bharadwaj after which he went to Kyoto University, Japan as a JSPS and CREST postdoctoral fellow (Host: Prof. Susumu Kitagawa). He is presently an Associate Professor in the Department of Chemistry at

ACS Paragon Plus Environment

66

Page 67 of 87 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

IISER, Pune. Currently, his research work is primarily focused upon the functional advanced porous materials (such as MOFs, MOPs, POPs etc.) aimed at environmental applications.

REFERENCES 1.

Resolution A/RES/64/292. United Nations General Assembly, July 2010.

2.

World Population Prospects: The 2015 Revision, Methodology of the United

Nations Population Estimates and Projections, Working Paper No. ESA/P/WP.242. Department of Economic and Social Affairs, Population Division, United Nations, 2015. 3.

Chapman, P. M. Determining when contamination is pollution - weight of

evidence determinations for sediments and effluents. Environ. Int. 2007, 33, 492-501, DOI 10.1016/j.envint.2006.09.001. 4.

Shannon, M. A.; Bohn, P. W.; Elimelech, M.; Georgiadis, J. G.; Marinas, B. J.; Mayes,

A. M. Science and technology for water purification in the coming decades. Nature 2008, 452, 301-310, DOI 10.1038/nature06599. 5.

Skougstad, M. W.; Fishman, M. J. Inorganic pollution in water—the analyst's

viewpoint. Toxic. Environ. Chem. Rev. 1978, 2, 219-236, DOI 10.1080/02772247809356927. 6.

Grandclement, C.; Seyssiecq, I.; Piram, A.; Wong-Wah-Chung, P.; Vanot, G.;

Tiliacos, N.; Roche, N.; Doumenq, P. From the conventional biological wastewater

ACS Paragon Plus Environment

67

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

Page 68 of 87

treatment to hybrid processes, the evaluation of organic micropollutant removal: A review. Water Res. 2017, 111, 297-317, DOI 10.1016/j.watres.2017.01.005. 7.

Jones, K. C.; Voogt, P. de, Persistent organic pollutants (POPs): state of the

science. Environ. Pollution 1999, 100, 209-221, DOI 10.1016/S0269-7491(99)00098-6. 8.

Bolong, N.; Ismail, A. F.; Salim, M. R., Matsuura, T. A review of the effects of

emerging contaminants in wastewater and options for their removal. Desalination 2009, 239, 229-246, DOI 10.1016/j.desal.2008.03.020. 9.

Slater, A. G.; Cooper, A. I. Function-led design of new porous materials. Science

2015, 348, aaa8075, DOI 10.1126/science.aaa8075. 10.

Batten, S. R.; Champness, N. R.; Chen, X.-M.; Garcia-Martinez, J.; Kitagawa, S.;

Ohrstrom, L.; O’Keeffe, M.; Suh, M. P.; Reedijk, J. Terminology of metal–organic frameworks and coordination polymers. Pure. Appl. Chem. 2013, 85, 1715-1724, DOI 10.1351/PAC-REC-12-11-20. 11.

Lin, Y.; Kong, C.; Zhang, Q.; Chen, L. Metal-Organic Frameworks for Carbon

Dioxide Capture and Methane Storage. Adv. Energy Mater. 2017, 7, 1601296, DOI 10.1002/aenm.201601296. 12.

Shimizu, G. K. H.; Taylor, J. M.; Kim, S.-R. Proton conduction with metal-organic

frameworks. Science 2013, 341, 354-355, DOI 10.1126/science.1239872.

ACS Paragon Plus Environment

68

Page 69 of 87 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

13.

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

applications of metal-organic frameworks. Science 2013, 30, 1230444, DOI 10.1126/science.1230444. 14.

Lustig, W. P.; Mukherjee, S.; Rudd, N. D.; Desai, A. V.; Li, J.; Ghosh, S. K. Metal-

organic frameworks: functional luminescent and photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46, 3242-3285, DOI 10.1039/C6CS00930A. 15.

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

Zhang, P.; Wang, Q.; Zou, L.; Zhang, Y.; Zhang, L.; Fang, Y.; Li, J.; Zhou, H.-C. Stable Metal– Organic Frameworks: Design, Synthesis, and Applications. Adv. Mater. 2018, 30, 1704303, DOI 10.1002/adma.201704303. 16.

Gao, Q.; Xu, J.; Bu, X.-H. Recent advances about metal–organic frameworks in the

removal of pollutants from wastewater. Coord. Chem. Rev. 2019, 378, 17-31, DOI 10.1016/j.ccr.2018.03.015. 17.

Lohse, M. S.; Ben, T. Covalent organic frameworks: structures, synthesis and

applications. Adv. Funct. Mater. 2018, 28, 1705553, DOI 10.1002/adfm.201705553. 18.

Zhao, W.; Xia, L.; Liu, X. Covalent organic frameworks (COFs): perspectives of

industrialization. CrystEngComm 2018, 20, 1613-1634, DOI 10.1039/C7CE02079A. 19.

Kaur, P.; Hupp, J. T.; Nguyen, S. T. Porous organic polymers in catalysis:

opportunities and challenges. ACS Catal. 2011, 1, 819-835, DOI 10.1021/cs200131g.

ACS Paragon Plus Environment

69

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

20.

Page 70 of 87

Wu, J.; Xu, F.; Li, S.; Ma, P.; Zhang, X.; Liu, Q.; Fu, R.; Wu, D. Porous Polymers as

Multifunctional Material Platforms toward Task‐Specific Applications. Adv. Mater. 2018, doi: 10.1002/adma.201802922. 21.

Rawat, D.; Mishra, V.; Sharma, R. S. Detoxification of azo dyes in the context of

environmental processes. Chemosphere 2016, 155, 591-605, DOI 10.1016/j.chemosphere.2016.04.068. 22.

Jiang, H. L.; Tatsu, Y.; Lu, Z. H.; Xu, Q. Non-, micro-, and mesoporous

metal−organic framework isomers: reversible transformation, fluorescence sensing, and large molecule separation. J. Am. Chem. Soc. 2010, 132, 5586-5587, DOI 10.1021/ja101541s. 23.

Jia, J.; Sun, F.; Borjigin, T.; Ren, H.; Zhang, T.; Bian, Z.; Zhu, G. Highly porous and

robust ionic MOFs with nia topology constructed by connecting an octahedral ligand and a trigonal prismatic metal cluster. Chem. Commun. 2012, 48, 6010-6012, DOI 10.1039/C2CC32347H. 24.

Zhu, Z.; Bai, Y. L.; Zhang, L.; Sun, D.; Fang, J.; Zhu, S. Two nanocage anionic metal–

organic frameworks with rht topology and {[M(H2O) 6] 6}12+ charge aggregation for rapid and selective adsorption of cationic dyes. Chem. Commun. 2014, 50, 14674-14677, DOI 10.1039/C4CC07365G. 25.

Liu, L.; Zhang, X. -N.; Han, Z. -B.; Gao, M. -L.; Cao, X. -M.; Wang, S. -M. An InIII-

based anionic metal–organic framework: sensitization of lanthanide (III) ions and

ACS Paragon Plus Environment

70

Page 71 of 87 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

selectivenabsorption and separation of cationic dyes. J. Mater. Chem. A 2015, 3, 14157– 14164, 10.1039/C5TA00986C. 26.

Samanta, P.; Chandra, P.; Desai, A. V.; Ghosh, S. K. Chemically stable microporous

hyper-cross-linked polymer (HCP): an efficient selective cationic dye scavenger from an aqueous medium. Mater. Chem. Front. 2017, 1, 1384-1388, DOI 10.1039/C6QM00362A. 27.

Desai, A. V.; Roy, A.; Samanta, P.; Manna, B.; Ghosh, S. K. Base Resistant Ionic

Metal-Organic Framework as a Porous Ion-Exchange Sorbent. iScience 2018, 3, 21-30, DOI 10.1016/j.isci.2018.04.004. 28.

Chen, D.; Shi, W.; Cheng, P. A cage-based cationic body-centered tetragonal

metal–organic framework: single-crystal to single-crystal transformation and selective uptake of organic dyes. Chem. Commun. 2015, 51, 370-372, DOI 10.1039/C4CC07357F. 29.

Song, B. -Q.; Wang, X. -L.; Zhang, Y. -T.; Wu, X. -S.; Liu, H. -S.; Shao, K. -Z.; Su, Z. -

M. Periodic tiling of triangular and square nanotubes in a cationic metal–organic framework for selective anion exchange. Chem. Commun. 2015, 51, 9515-9518, DOI 10.1039/C5CC02639C. 30.

Gao, C. -Y.; Yang, Y.; Ai, J.; Tian, H. -R.; Li, L. -J.; Yang, W.; Dang, S.; Sun, Z. -M. A

Multifunctional MnII Phosphonate for Rapid Separation of Methyl Orange and ElectronTransfer Photochromism. Chem. Eur. J. 2016, 22, 11652–11659, DOI 10.1002/chem.201601318. 31.

Huang, N.; Wang, P.; Addicoat, M. A.; Heine, T.; Jiang, D. Ionic Covalent Organic

Frameworks: Design of a Charged Interface Aligned on 1D Channel Walls and Its

ACS Paragon Plus Environment

71

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

Page 72 of 87

Unusual Electrostatic Functions. Angew. Chem. Int. Ed. 2017, 56, 4982-4986, DOI 10.1002/anie.201611542. 32.

Yu, S. -B.; Lyu, H.; Tian, J.; Wang, H.; Zhang, D. -W.; Liu, Y.; Li, Z. -T. A polycationic

covalent organic framework: a robust adsorbent for anionic dye pollutants. Polym. Chem. 2016, 7, 3392–3397, DOI 10.1039/C6PY00281A. 33.

Dey, A.; Konavarapu, S. K.; Sasmal, H. S.; Biradha, K. Porous Coordination Polymers

Containing Pyridine-3,5-Bis(5-azabenzimidazole): Exploration of Water Sorption, Selective Dye Adsorption, and Luminescent Properties. Cryst. Growth Des. 2016, 16, 5976–5984, DOI 10.1021/acs.cgd.6b01028. 34.

Dey, K.; Pal, M.; Rout, K. C.; Kunjattu H, S.; Das, A.; Mukherjee, R.; Kharul, U. K.;

Banerjee, R. Selective Molecular Separation by Interfacially Crystallized Covalent Organic Framework Thin Films. J. Am. Chem. Soc. 2017, 139, 13083−13091, DOI 10.1021/jacs.7b06640. 35.

Dey, A.; Bairagi, D.; Biradha, K. MOFs with PCU Topology for the Inclusion of One-

Dimensional Water Cages: Selective Sorption of Water Vapor, CO2, and Dyes and Luminescence Properties. Cryst. Growth Des. 2017, 17, 3885–3892, DOI 10.1021/acs.cgd.7b00502. 36.

Kumar, S.; Verma, G.; Gao, W. -Y.; Niu, Z.; Wojtas, L.; Ma, S. Anionic Metal–Organic

Framework for Selective Dye Removal and CO2 Fixation. Eur. J. Inorg. Chem. 2016, 4373– 4377, DOI 10.1002/ejic.201600218. 37.

Konavarapu, S. K.; Biradha, K. Luminescent Triazene-Based Covalent Organic

Frameworks Functionalized with Imine and Azine: N2 and H2 Sorption and Efficient Removal of Organic Dye Pollutants. Cryst. Growth Des. 2019, 19, 362−368, DOI 10.1021/acs.cgd.8b01458. 38.

Zhao, X.; Bu, X.; Wu, T.; Zheng, S. -T.; Wang, L.; Feng, P. Selective anion exchange

with nanogated isoreticular positive metal-organic frameworks. Nat. Commun. 2013, DOI: 10.1038/ncomms3344.

ACS Paragon Plus Environment

72

Page 73 of 87 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

39.

Mon, M.; Bruno, R.; Ferrando-Soria, J.; Armentano, D.; Pardo, E. Metal–organic

framework technologies for water remediation: towards a sustainable ecosystem. J. Mater. Chem. A 2018, 6, 4912–4947, DOI 10.1039/C8TA00264A. 40.

Gao. Q.; Jian Xu, J.; Bu, X. -H. Recent advances about metal–organic frameworks in

the removal of pollutants from wastewater. Coord. Chem. Rev. 2019, 378,17–31, DOI 10.1016/j.ccr.2018.03.015. 41.

Dias, E. M.; Petit, C. Towards the use of metal–organic frameworks for water

reuse: a review of the recent advances in the field of organic pollutants removal and degradation and the next steps in the field. J. Mater. Chem. A 2015, 3, 22484–22506, DOI 10.1039/C5TA05440K. 42.

Van de Voorde, B.; Bueken, B.; Denayer, J.; De Vos, D. Adsorptive separation on

metal–organic frameworks in the liquid phase. Chem. Soc. Rev. 2014, 43, 5766-5788, DOI 10.1039/C4CS00006D. 43.

Wang, C.; Liu, X.; Demir, N. K.; Chen, J. P.; Li, K. Applications of water stable metal–

organic frameworks. Chem. Soc. Rev. 2016, 45, 5107-5134, DOI 10.1039/C6CS00362A. 44.

Wang, C. -C.; Li, J. -R.; Lv, X. -L.; Zhang, Y. -Q.; Guo, G. Photocatalytic organic

pollutants degradation in metal–organic frameworks. Energy Environ. Sci. 2014, 7, 2831– 2867, DOI 10.1039/C4EE01299B. 45.

Ayati, A.; Shahrak, M. N.; Tanhaei, B.; Sillanpää, M. Emerging adsorptive removal

of azo dye by metal-organic frameworks. Chemosphere 2016, 160, 30-44, DOI 10.1016/j.chemosphere.2016.06.065. 46.

Castiglioni, S.; Bagnati, R.; Fanelli, R.; Pomati, F.; Calamari, D.; Zuccato, E. Removal

of Pharmaceuticals in Sewage Treatment Plants in Italy. Environ. Sci. Technol. 2006, 40, 357-363, DOI 10.1021/es050991m.

ACS Paragon Plus Environment

73

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

47.

Page 74 of 87

Ahmed, M. B.; Zhou, J. L.; Ngo, H. H.; Guo, W. Adsorptive removal of antibiotics

from water and wastewater: Progress and challenges. Sci. Total Environ. 2015, 532, 112– 126, DOI 10.1016/j.scitotenv.2015.05.130. 48.

Zhang, Q. -Q.; Ying, G. -G.; Pan, C. -G.; Liu, Y. -S.; Zhao, J. -L. Comprehensive

Evaluation of Antibiotics Emission and Fate in the River Basins of China: Source Analysis, Multimedia Modeling, and Linkage to Bacterial Resistance. Environ. Sci. Technol. 2015, 49, 6772−6782, DOI 10.1021/acs.est.5b00729. 49.

Jin, W. -G.; Chen, W.; Xu, P. -H. ; Lin, X. -W. ; Huang, X. -C.; Chen, G. -H.; Lu, F.;

Chen, X. -M. An Exceptionally Water Stable Metal–Organic Framework with AmideFunctionalized Cages: Selective CO2/CH4 Uptake, Removal of Antibiotics and Dyes from Water. Chem. Eur. J. 2017, 23,13058–13066, DOI 10.1002/chem.201701884. 50.

Stockholm International Water Institute, Antibiotic resistance: the importance of

water, 2018, Issue-sheet-antibiotics-ENG_WEB. 51.

Wang, B.; Lv, X. -L.; Feng, D.; Xie, L. -H.; Zhang, J.; Li, M.; Xie, Y.; Li, J. -R.; Zhou, H. -

C. Highly Stable Zr(IV)-Based Metal−Organic Frameworks for the Detection and Removal of Antibiotics and Organic Explosives in Water. J. Am. Chem. Soc. 2016, 138, 6204−6216, DOI 10.1021/jacs.6b01663. 52.

Dai, X.; Jia, X.; Zhao, P.; Wang, T.; Wang, J.; Huang, P.; He, L.; Hou, X. A combined

experimental/computational study on metal-organic framework MIL-101(Cr) as a SPE sorbent for the determination of sulphonamides in environmental water samples

ACS Paragon Plus Environment

74

Page 75 of 87 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

coupling with UPLC-MS/MS. Talanta 2016, 154, 581–588, DOI 10.1016/j.talanta.2016.03.042. 53.

Seo, P. W.; Bhadra, B. N.; Ahmed, I.; Khan, N. A.; Jhung, S. H. Adsorptive Removal

of Pharmaceuticals and Personal Care Products from Water with Functionalized Metalorganic Frameworks: Remarkable Adsorbents with Hydrogenbonding Abilities. Sci. Rep. 2016, DOI: 10.1038/srep34462. 54.

Song, J. Y.; Jhung, S. H. Adsorption of Pharmaceuticals and Personal Care

Products over Metal-organic frameworks Functionalized with Hydroxyl Groups: Quantitative Analyses of H-bonding in Adsorption. Chem. Eng. J. 2017, 322, 366-374, DOI 10.1016/j.cej.2017.04.036. 55.

Seo, P. W.; Khan, N. A.; Jhung, S. H. Removal of nitroimidazole antibiotics from

water by adsorption over metal–organic frameworks modified with urea or melamine. Chem. Eng. J. 2017, 315, 92-100, DOI 10.1016/j.cej.2017.01.021. 56.

Sarker, M.; Song, J. Y.; Jhung, S. H. Carboxylic-acid-functionalized UiO-66-NH2: A

promising adsorbent for both aqueous- and non-aqueous-phase adsorptions. Chem. Eng. J. 2018, 331, 124-131, DOI 10.1016/j.cej.2017.08.017. 57.

Han, M. -L.; Wen, G. -X.; Dong, W. -W.; Zhou, Z. -H.; Wu, Y. -P.; Zhao, J.; Li, D. -S.;

Ma, L. -F.; Bu. X.; A heterometallic sodium–europium-cluster-based metal–organic framework as a versatile and water-stable chemosensor for antibiotics and explosives. J. Mater. Chem. C 2017, 5, 8469-8474, DOI 10.1039/C7TC02885G.

ACS Paragon Plus Environment

75

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

58.

Page 76 of 87

Liu, X.-G.; Tao, C. -L.; Yu, H. -Q.; Chen, B.; Liu, Z.; Zhu, G. -P.; Zhao, Z.; Shen, L.;

Tang, B. Z. A new luminescent metal–organic framework based on dicarboxylsubstituted tetraphenylethene for efficient detection of nitro-containing explosives and antibiotics in aqueous media. J. Mater. Chem. C 2018, 6, 2983-2988, DOI 10.1039/C7TC05535H. 59.

Zhou, Y.; Yang, Q.; Zhang, D.; Gan, N.; Li, Q.; Cuan, J. Detection and removal of

antibiotic tetracycline in water with ahighly stable luminescent MOF. Sensors and Actuators B 2018, 262, 137–143, DOI 10.1016/j.snb.2018.01.218. 60.

Fu, H. -R.; Yan, L. -B.; Wu, N. -T.; Ma, L. -F.; Zang, S. -Q. Dual-emission MOFdye

sensor for ratiometric fluorescence recognition of RDX and detection of a broad class of nitro-compounds. J. Mater. Chem. A 2018, 6, 9183–9191, DOI 10.1039/C8TA02857E. 61.

Shuo Lin, Yufeng Zhao, Yeoung-Sang Yun, . Highly Effective Removal of

Nonsteroidal Anti-inflammatory Pharmaceuticals from Water by Zr(IV)-Based Metal−Organic Framework: Adsorption Performance and Mechanisms. ACS Appl. Mater. Interfaces 2018, 10, 28076−28085, DOI 10.1021/acsami.8b08596. 62.

Mellah, A.; Fernandes, S. P. S.; Rodríguez, R.; Otero, J.; Paz, J.; Cruces, J.; Medina, D.

D.; Djamila, H.; Espiña, B.; Salonen, L. M. Adsorption of Pharmaceutical Pollutants from Water Using Covalent Organic Frameworks. Chem. Eur. J. 2018, 24,10601–10605, DOI 10.1002/chem.201801649.

ACS Paragon Plus Environment

76

Page 77 of 87 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

63.

Liu, Y.; Fan, X.; Jia, X.; Chen, X.; Zhang, A.; Zhang, B.; Zhang, Q. Preparation of

Magnetic Hyper-Cross-Linked Polymers for the Efficient Removal of Antibiotics from Water. ACS Sustainable Chem. Eng. 2018, 6, 210−222, DOI 10.1021/acssuschemeng.7b02252. 64.

Li, S.; Zhang, X.; Huang, Y. Zeolitic imidazolate framework-8 derived nanoporous

carbon as an effective and recyclable adsorbent for removal of ciprofloxacin antibiotics from water. J. Hazard. Mat. 2017, 321, 711-719, DOI 10.1016/j.jhazmat.2016.09.065. 65.

Chen, M.; Gan, N.; Zhou, Y.; Li, T.; Xu, Q.; Cao, Y.; Chen, Y. A novel aptamer- metal

ions- nanoscale MOF based electrochemicalbiocodes for multiple antibiotics detection and signal amplification. Sensors and Actuators B 2017, 242, 1201–1209, DOI 10.1016/j.snb.2016.08.185. 66.

Liu, G. -N.; Zhao, R. -Y.; Xu, R. -D.; Zhang, X.; Tang, X. -N.; Dan, Q. -J.; Wei, Y. -W.;

Tu, Y. -Y.; Bo, Q. -B.; Li, C. A Novel Tetranuclear Copper(I) Iodide Metal−Organic Cluster [Cu4I4(Ligand)5] with Highly Selective Luminescence Detection of Antibiotic. Cryst. Growth Des. 2018, 18, 5441–5448, DOI 10.1021/acs.cgd.8b00819. 67.

Dhaka, S.; Kumar, R.; Deep,A.; Kurade, M. B.; Ji, S. -W.; Jeon, B. -H. Metal–organic

frameworks (MOFs) for the removal of emerging contaminants from aquatic environments. Coord. Chem. Rev. 2019, 380, 330–352, DOI 10.1016/j.ccr.2018.10.003.

ACS Paragon Plus Environment

77

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

68.

Page 78 of 87

Tilman, D.; Cassman, K. G.; Matson, P. A.; Naylor, R.; Polasky, S. Agricultural

sustainability and intensive production practices. Nature 2002, 418, 671-677, DOI 10.1038/nature01014. 69.

Jung, B. K.; Hasan, Z.; Jhung, S. H. Adsorptive removal of 2,4-

dichlorophenoxyacetic acid (2,4-D) from water with a metal–organic framework. Chem. Eng. J. 2013, 234, 99–105, DOI 10.1016/j.cej.2013.08.110. 70.

Seo, Y. S.; Khan, N. A.; Jhung, S. H. Adsorptive removal of

methylchlorophenoxypropionic acid from water with a metal-organic framework. Chem. Eng. J. 2015, 270, 22–27, DOI 10.1016/j.cej.2015.02.007. 71.

Majumder, M.; Sheath, P.; Mardel, J. I.; Harvey, T. G.; Thornton, A. W.; Gonzago, A.;

Kennedy, D. F.; Madsen, I.; Taylor, J. W.; Turner, D. R.; Hill, M. R. Aqueous Molecular Sieving and Strong Gas Adsorption in Highly Porous MOFs with a Facile Synthesis. Chem. Mater. 2012, 24, 4647−4652, DOI 10.1021/cm301726k. 72.

Jia, Y. -Y.; Zhang, Y. -H.; Xu, J.; Feng, R.; Zhang, M. -S.; Bu, X. -H. A high-

performance ‘‘sweeper’’ for toxic cationic herbicides: an anionic metal–organic framework with a tetrapodal cage. Chem. Commun. 2015, 51, 17439-17442, DOI 10.1039/C5CC07249B. 73.

Du, W.; Zhu, Z.; Bai, Y. -L.; Yang, Z.; Zhu, S.; Xu, J.; Xie, Z.; Fang, J. An anionic sod-

type terbium-MOF with extralarge cavities for effective anthocyanin extraction and

ACS Paragon Plus Environment

78

Page 79 of 87 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

methyl viologen detection. Chem. Commun. 2018, 54, 5972-5975, DOI 10.1039/C8CC02193G. 74.

Zhu, X.; Li, B.; Yang, J.; Li, Y.; Zhao, W.; Shi, J.; Gu, J. Effective Adsorption and

Enhanced Removal of Organophosphorus Pesticides from Aqueous Solution by ZrBased MOFs of UiO-67. ACS Appl. Mater. Interfaces 2015, 7, 223−231, DOI 10.1021/am5059074. 75.

Fu, Q.; Wen, L.; Zhang, L.; Chen, X.; Pun, D.; Ahmed, A.; Yang, Y.; Zhang, H.

Preparation of Ice-Templated MOF−Polymer Composite Monoliths and Their Application for Wastewater Treatment with High Capacity and Easy Recycling. ACS Appl. Mater. Interfaces 2017, 9, 33979−33988, DOI 10.1021/acsami.7b10872. 76.

Cao, X.; Liu, G.; She, Y.; Jiang, Z.; Jin, F.; Jin, M.; Du, P.; Zhao, F.; Zhang, Y.; Wang, J.

Preparation of magnetic metal organic framework composites for the extraction of neonicotinoid insecticides from environmental water samples. RSC Adv. 2016, 6, 113144–113151, DOI 10.1039/C6RA23759B. 77.

Jung, B. K. ; Jun, J. W.; Hasan, Z.; Jhung, S. H. Adsorptive removal of p-arsanilic

acid from water using mesoporous zeolitic imidazolate framework-8. Chem. Eng. J. 2015, 267, 9–15, DOI 10.1016/j.cej.2014.12.093. 78.

Jung, B. K. ; Jun, J. W.; Hasan, Z.; Jhung, S. H. Adsorption of organic arsenic acids

from water over functionalizedmetal-organic frameworks. J. Hazard. Mat. 2017, 335, 162–169, DOI 10.1016/j.jhazmat.2017.04.044.

ACS Paragon Plus Environment

79

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

79.

Page 80 of 87

Seo, P. W.; Khan, N. A.; Hasan, Z.; Jhung, S. H. Adsorptive Removal of Artificial

Sweeteners from Water Using Metal−Organic Frameworks Functionalized with Urea or Melamine. ACS Appl. Mater. Interfaces 2016, 8, 29799−29807, DOI 10.1021/acsami.6b11115. 80.

Tian, C.; Zhao, J.; Ou, X.; Wan, J.; Cai, Y.; Lin, Z.; Dang, Z.; Xing, B. Enhanced

Adsorption of p-Arsanilic Acid from Water by Amine-Modified UiO-67 as Examined Using Extended X-ray Absorption Fine Structure, X-ray Photoelectron Spectroscopy, and Density Functional Theory Calculations. Environ. Sci. Technol. 2018, 52, 3466−3475, DOI 10.1021/acs.est.7b05761. 81.

Li, B.; Zhu, X.; Hu, K.; Li, Y.; Feng, J.; Shi, J.; Gu, J. Defect creation in metal-organic

frameworks for rapid andcontrollable decontamination of roxarsone from aqueous solution. J. Hazard. Mat. 2016, 302, 57–64, DOI 10.1016/j.jhazmat.2015.09.040. 82.

Lv, Y.; Zhang, R.; Zeng, S.; Liu, K.; Huang, S.; Liu, Y.; Xu, P.; Lin, C.; Cheng, Y.; Liu, M.

Removal of p-arsanilic acid by an amino-functionalized indium-based metal-organic framework: Adsorption behavior and synergetic mechanism. Chem. Eng. J. 2018, 339, 359-368, DOI 10.1016/j.cej.2018.01.139. 83.

Mukherjee, S.; Kansara, A. M.; Saha, D.; Gonnade, R.; Mullangi, D.; Manna, B.;

Desai, A. V.; Thorat, S. H.; Singh, P. S.; Mukherjee, A.; Ghosh, S. K. An Ultrahydrophobic Fluorous Metal–Organic Framework Derived Recyclable Composite as a Promising Platform to Tackle Marine Oil Spills. Chem. Eur. J. 2016, 22, 10937– 10943, DOI 10.1002/chem.201601724.

ACS Paragon Plus Environment

80

Page 81 of 87 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

84.

Rao, K. P.; Higuchi, M.; Suryachandram, J.; Kitagawa, S. Temperature-Stable

Compelled Composite Superhydrophobic Porous Coordination Polymers Achieved via an Unattainable de Novo Synthetic Method. J. Am. Chem. Soc. 2018, 140, 13786–13792, DOI 10.1021/jacs.8b07577. 85.

Yang, C.; Kaipa, U.; Mather, Q. Z.; Wang, X.; Nesterov, V.; Venero, A. F.; Omary, M.

A. Fluorous Metal-Organic Frameworks with Superior Adsorption and Hydrophobic Properties toward Oil Spill Cleanup and Hydrocarbon Storage. J. Am. Chem. Soc. 2011, 133, 18094–18097, DOI 10.1021/ja208408n. 86.

Jiang, Z. -R.; Ge, J.; Zhou, Y. -X.; Wang, Z. U; Chen, D.; Yu, S. -H.; Jiang, H. -L.

Coating sponge with a hydrophobic porous coordination polymer containing a lowenergy CF3-decorated surface for continuous pumping recovery of an oil spill from water. NPG Asia Mater. 2016, 8, e253; doi:10.1038/am.2016.22. 87.

Jayaramulu, K.; Datta, K. K. R.; Rçsler, C.; Petr, M.; Otyepka, M.; Zboril, R.; Fischer, R.

A. Biomimetic Superhydrophobic/Superoleophilic Highly Fluorinated Graphene Oxide and ZIF-8 Composites for Oil–Water Separation. Angew. Chem. Int. Ed. 2016, 55, 1178 – 1182, DOI 10.1002/anie.201507692. 88.

Han, T.; Xiao, Y.; Tong, M.; Huang, H.; Liu, D.; Wang, L.; Zhong, C. Synthesis of

CNT@MIL-68(Al) composites with improved adsorption capacity for phenol in aqueous solution. Chem. Eng. J. 2015, 275,134–141, DOI 10.1016/j.cej.2015.04.005. 89.

Khan, N. A.; Jung, B. K.; Hasan, Z.; Jhung, S. H. Adsorption and removal of phthalic

acid and diethyl phthalate fromwater with zeolitic imidazolate and metal–organic frameworks. J. Hazard. Mater. 2015, 282,194–200, DOI 10.1016/j.jhazmat.2014.03.047. 90.

Nagarkar, S. S.; Desai, A. V.; Ghosh, S. K. A fluorescent metal–organic framework

for highly selective detection of nitro explosives in the aqueous phase. Chem. Commun. 2014, 50, 8915-8918, DOI 10.1039/C4CC03053B.

ACS Paragon Plus Environment

81

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

91.

Page 82 of 87

Joarder, B.; Desai, A. V.; Samanta, P.; Mukherjee, S.; Ghosh, S. K. Selective and

Sensitive Aqueous‐Phase Detection of 2,4,6‐Trinitrophenol (TNP) by an Amine‐Functionalized Metal–Organic Framework. Chem. Eur. J. 2015, 21, 965– 969, DOI 10.1002/chem.201405167. 92.

Nagarkar, S. S.; Joarder, B.; Chaudhari, A. K.; Mukherjee, S.; Ghosh, S. K. Highly

Selective Detection of Nitro Explosives by a Luminescent Metal–Organic Framework. Angew. Chem. Int. Ed. 2013, 52, 2881 –2885, DOI 10.1002/anie.201208885. 93.

Li, Y.; Yang, Z.; Wang, Y.; Bai, Z.; Zheng, T.; Dai, X.; Liu, S.; Gui, D.; Liu, W.; Chen, M.;

Chen, L.; Diwu, J.; Zhu, L.; Zhou, R.; Chai, Z.; Albrecht-Schmitt, T. E.; Wang, S. A mesoporous cationic thorium-organic framework that rapidly traps anionic persistent organic pollutants. Nat. Commun. 2017, DOI: 10.1038/s41467-017-01208-w. 94.

Ji, W.; Xiao, L.; Ling, Y.; Ching, C.; Matsumoto, M.; Bisbey, R. P.; Helbling, D. E.;

Dichtel, W. R. Removal of GenX and Perfluorinated Alkyl Substances from Water by Amine-Functionalized Covalent Organic Frameworks. J. Am. Chem. Soc. 2018, 140, 12677–12681, DOI 10.1021/jacs.8b06958. 95.

Byun, J.; Patel, H. A.; Thirion, D.; Yavuz, C. T. Charge-specific size-dependent

separation of water-soluble organic molecules by fluorinated nanoporous networks. Nat. Commun. 2016, DOI: 10.1038/ncomms13377. 96.

Ling, Y.; Klemes, M. J.; Xiao, L.; Alsbaiee, A.; Dichtel, W. R.; Helbling, D. E.

Benchmarking Micropollutant Removal by Activated Carbon and Porous β‑Cyclodextrin Polymers under Environmentally Relevant Scenarios. Environ. Sci. Technol. 2017, 51, 7590−7598, DOI 10.1021/acs.est.7b00906. 97.

Llabres i Xamena, F. X.; Corma, A.; Garcia, H. Applications for Metal-Organic

Frameworks (MOFs) as Quantum Dot Semiconductors. J. Phys. Chem. C. 2007, 111, 8085, DOI 10.1021/jp063600e.

ACS Paragon Plus Environment

82

Page 83 of 87 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

98.

Das, M. C.; Xu, H.; Wang, Z.; Srinivas, G.; Zhou, W.; Yue, Y. –F.; Nesterov, V. N.;

Qian, G.; Chen, B.A Zn4O-containing doubly interpenetrated porous metal–organic framework for photocatalytic decomposition of methyl orange. Chem. Commun. 2011, 47, 11715–11717, DOI 10.1039/C1CC12802G. 99.

Wang, D. –E.; Deng, K. –J.; Lv, K. –L.; Wang, C. –G.; Wen, L. –L.; Li, D. –F. Structures,

photoluminescence and photocatalytic properties of three new metal–organic frameworks based on non-rigid long bridges.CrystEngComm.2009, 11, 1442–1450, DOI 10.1039/B821263E. 100.

Hou, Y. –L.; Sun, R. W. –Y.; Zhou, X. –P.; Wang, J. –H.; Li, D.A copper(I)/copper(II)–

salen coordination polymer as a bimetallic catalyst for three-component Strecker reactions and degradation of organic dyes. Chem. Commun.2014, 50, 2295—2297, DOI 10.1039/C3CC47996J. 101.

Zhu, W.; Yang, X. –Y.; Li, Y. –H.; Li, J. –P.; Wu, D.; Gao, Y.; Yi, F. –Y. A novel porous

molybdophosphate-based FeII,III-MOF showing selective dye degradation as a recyclable photocatalyst. Inorg. Chem. Commun.2014, 49, 159-162, DOI 10.1016/j.inoche.2014.09.033. 102.

Zou, L.; Feng, D.; Liu, T. –F.; Chen, Y. –P.; Yuan, S.; Wang, K.; Wang, X.; Fordham, S.;

Zhou, H. –C. A versatile synthetic route for the preparation of titanium metal–organic frameworks. Chem. Sci. 2016, 7, 1063–1069, DOI 10.1039/C5SC03620H.

ACS Paragon Plus Environment

83

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

103.

Page 84 of 87

Zhao H.; Xia, Q.; Xing, H.; Chen, D.; Wang, H. Construction of Pillared-Layer MOF

as Efficient Visible-Light Photocatalysts for Aqueous Cr(VI) Reduction and Dye Degradation. ACS Sustainable Chem. Eng. 2017, 5, 4449−4456, DOI 10.1021/acssuschemeng.7b00641. 104.

Liu, J. –J.; Shan, Y. –B.; Dai, W. –X.; Huang, C. –C.; Lin, M. –J. Assembly of donor–

acceptor hybrid heterostructures based on iodoplumbates and viologen coordination polymers. Dalton Trans.,2017, 46, 11556–11560, DOI 10.1039/C7DT02133J. 105.

Zhou, E. –H.; Li, B. –H.; Chen, W. –X.; Luo, Z.; Liu, J.; Singh, A.; Kumar, A.; Jin, J. –C.

Photocatalytic degradation of organic dyes by a stable and biocompatible Zn(II) MOF having ferulic acid: Experimental findings and theoretical correlation. J. Mol. Struct.2017, 1149, 352-356, DOI 10.1016/j.molstruc.2017.07.105. 106.

Kampouri, S.; Nguyen, T. N.; Spodaryk, M.; Palgrave, R. G.; Züttel, A.; Smit, B.;

Stylianou, K. –C. Concurrent Photocatalytic Hydrogen Generation and Dye Degradation Using MIL‐125‐NH2 under Visible Light Irradiation. Adv. Funct. Mater. 2018, 1806368, DOI 10.1002/adfm.201806368. 107.

Liu, C. –X.; Zhang, W. –H.; Wang, N.; Guo, P.; Muhler, M..; Wang, Y. Dr.; Lin, S.;

Chen, Z.; Yang, G. Highly Efficient Photocatalytic Degradation of Dyes by a Copper– Triazolate Metal–Organic Framework. Chem. Eur. J. 2018, 24, 16804 –16813, DOI 10.1002/chem.201803306.

ACS Paragon Plus Environment

84

Page 85 of 87 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

108.

Gkaniatsou, E.; Sicard, C.; Ricoux, R.; Benahmed, L.; Bourdreux, F.; Zhang, Q.; Serre,

C.; Mahy, J. –P.; Steunou, N. Enzyme Encapsulation in Mesoporous Metal–Organic Frameworks for Selective Biodegradation of Harmful Dye Molecules. Angew. Chem. Int. Ed. 2018, 57, 16141 –16146, DOI 10.1002/anie.201811327. 109.

Azhar, M. R.; Vijay, P.; Tadé, M. O.; Sun, H.; Wang, S. Submicron sized water-stable

metal organic framework (bio-MOF-11) for catalytic degradation of pharmaceuticals and personal care products. Chemosphere. 2018, 196, 105-114, DOI 10.1016/j.chemosphere.2017.12.164. 110.

Fang, Y.; Zhu, S. –R.; Wu, M. –K.; Zhao, W. –N.; Han. L. MOF-derived In2S3

nanorods for photocatalytic removal of dye and antibiotics. J. Solid State Chem. 2018, 266, 205-209, DOI 10.1016/j.jssc.2018.07.026. 111.

Guo, W.; Chen, Z.; Yang, C.; Neumann, T.; Kübel, C.; Wenzel, W.; Welle, A.; Pfleging,

W.; Shekhah, O.; Wӧll, C.; Redel, E. Bi2O3 nanoparticles encapsulated in surface mounted metal–organic framework thin films. Nanoscale. 2016, 8, 6468–6472, DOI 10.1039/C6NR00532B. 112.

Ma, B. C.; Ghasimi, S.; Landfester, K.; Vilela, F.; Zhang, K. A. Conjugated

microporous polymer nanoparticles with enhanced dispersibility and water compatibility for photocatalytic applications. J. Mater. Chem. A. 2015, 3, 16064–16071, DOI 10.1039/C5TA03820K.

ACS Paragon Plus Environment

85

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

113.

Page 86 of 87

Li, L.; Cai, Z. Structure control and photocatalytic performance of porous

conjugated polymers based on perylene diimide. Polym. Chem.2016, 7, 4937–4943, DOI 10.1039/C6PY00972G. 114.

Huang, N.; Zhai, L.; Xu, H.; Jiang, D. Stable Covalent Organic Frameworks for

Exceptional Mercury Removal from Aqueous Solutions. J. Am. Chem. Soc. 2017, 139, 2428−2434, DOI 10.1021/jacs.6b12328.

Table of Content

This review summaries detection and sequestration of toxic and hazardous organic pollutants with advanced porous materials for the sustainable ecosystem.

ACS Paragon Plus Environment

86

Page 87 of 87 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

ACS Paragon Plus Environment

87