Nanocomposite of MoS2-RGO as Facile, Heterogeneous, Recyclable

Aug 15, 2017 - A nanocomposite comprised of MoS2-RGO having unique structural features was developed by using a facile preparation strategy and demons...
2 downloads 17 Views 8MB Size
Subscriber access provided by University of Rochester | River Campus & Miner Libraries

Article

Nanocomposite of MoS2-RGO as facile, heterogeneous, recyclable and highly efficient Green catalyst for one-pot synthesis of indole alkaloids Ashish Bahuguna, Suneel Kumar, Vipul Sharma, Kumbam Lingeshwar Reddy, Kaustava Bhattacharyya, Ponneri Chandrababu Ravikumar, and Venkata Krishnan ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.7b00648 • Publication Date (Web): 15 Aug 2017 Downloaded from http://pubs.acs.org on August 15, 2017

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 free 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 accessible to all readers and 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.

ACS Sustainable Chemistry & Engineering 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 52

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

Nanocomposite of MoS2-RGO as facile, heterogeneous, recyclable and highly efficient Green catalyst for one-pot synthesis of indole alkaloids Ashish Bahuguna,a Suneel Kumar,a Vipul Sharma,a Kumbam Lingeshwar Reddy,a Kaustava Bhattacharyya,b P. C. Ravikumara,c* and Venkata Krishnana* a

School of Basic Sciences and Advanced Materials Research Center, Indian Institute of Technology Mandi, Kamand, Mandi 175005, Himachal Pradesh, India.

b

Chemistry Division, Bhabha Atomic Research Centre, Mumbai 400085, Maharashtra, India.

c

National Institute of Science Education and Research, Bhubaneswar752050, Odisha, India. Email: [email protected]; [email protected]

ABSTRACT: A nanocomposite comprising of MoS2-RGO having unique structural features was developed by using a facile preparation strategy and demonstrated to be a highly efficient heterogeneous catalyst for the synthesis of indole alkaloids in water. The catalyst could be recycled six times without significant loss of its activity. Green chemistry matrices calculations for the reaction showed high atom economy (A.E. = 94.7%) and small E-factor (0.089). Using this nanocomposite as catalyst, four naturally occurring indole alkaloid, namely Arundine, Vibrindole A, Turbomycin B and Trisindole, were synthesized along with their other derivatives in excellent yields.

KEYWORDS:

Nanocomposite,

heterogeneous

catalyst,

Green

chemistry,

organic

transformation, indole alkaloids

Page 1 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 2 of 52

INTRODUCTION Organic transformations using heterogeneous catalysts has played a crucial role in synthetic chemistry for the past few decades [1]. Heterogeneous catalysis has many advantages over homogeneous catalysis [2, 3], such as easy recyclability, can be operated at elevated temperature and pressure ranges and high industrial relevance (i.e. about 85% of all industrial catalytic processes are catalyzed by heterogeneous catalysts) due to which it has become preferred choice of chemists across the world. Nanocomposites of metals with graphene-based materials are some examples of heterogeneous catalysts, which have demonstrated their wide applicability [4]. Nanocomposite mediated organic transformation has recently emerged as an important and growing field in chemistry and materials science, and has attracted the attention of several researchers working in this area. But most of the organic reactions using nanocomposites have been performed in non-green and toxic organic solvents [5]. In addition, use of adverse reaction conditions, such as high temperature and pressure as well as longer reaction time and low percentage of yields has prompted researchers to develop catalysts, which could overcome these problems. Hence there is a need to develop a new kind of catalyst, which could resolve these issues pertaining to sustainability. Use of Green solvents and ambient reaction conditions are the prerequisite need of Green and sustainable chemistry. Green chemistry refers to the practice of designing environment friendly reactions using safe and less hazardous chemicals, reducing waste, less energy consumption and producing maximum yields [6]. Oxides and sulfides of several metals, such as Zn, Cd, Mo, Cu, Fe, Ru etc., have been used as an active catalyst for various organic transformations[7-11]. There are also few reports in literature [12-14] where researchers have developed various heterogeneous catalysts using nanocomposites of various metal oxides and sulfides with graphene or reduced graphene oxide, which could minimize the use of toxic organic solvents and performed reactions at ambient conditions. Recently Rawat et al. developed a protocol for the synthesis of 3-substituted indole by using ZnO-RGO nanocomposite as a heterogeneous catalyst [12]. The same group later on developed another methodology for the synthesis of aminoindolizines using CuI-carbon spheres based nanocomposites under Green conditions [13]. Verma et al. performed Ritter and other Page 2 of 52 ACS Paragon Plus Environment

Page 3 of 52

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

multicomponent reactions using iron oxide-supported sulfonic acid nanocatalysts [14]. In addition to this chitosan supported ruthenium was also used as a heterogeneous catalyst for hydration of nitriles [15]. Molybdenum metal is considered relatively safe for living being [16] in comparison to other heavy metals and no toxicity has been reported so far [17]. In fact molybdenum has been found essential trace element for the proper growth of plants and animals [18]. Sulfides of molybdenum (e.g. MoS2 etc.) has been employed as efficient catalysts for various reactions such as hydrogen evolution [19], desulfurization in petrochemistry [20], reductive alkylation [21], hydrogenolysis [22] and so on. Recently nanocomposite of MoS2 with reduced graphene oxide (RGO) has attracted attention of researchers from various research areas such as photocatalysis [23], dye sensitized solar cells [24], biosensing [25, 26], photovoltaics [24], supercapacitors [27] etc. Organic transformations using nanocomposites has been emerging as a thrust area of chemistry and nanomaterials. MoS2-RGO and CdS nanocomposite was found to be an efficient photocatalyst for the reduction of 4-nitrophenol [28]. Two dimensional MoS2graphene composites were also used as a supports for Pt electrocatalysts in methanol oxidation [29]. Although MoS2-RGO nanocomposites have been used as catalysts for different reactions, in various other fields of materials but to the best of our knowledge, there exists no report on its use for organic transformation reactions. MoS2 possesses two-dimensional (2D) structure where Mo atom sandwiched between two layers of S atoms which are held together by Van der Waal forces[30]. These MoS2 nanosheets supported over RGO nanosheets results in 2D-2D contact in nanocomposite with high surface area [31] and 2D nature of one atom thick carbon sheet of graphene makes it an ideal catalyst for various applications, including organic transformations. Indole based organic structures have been found in various natural products and drug molecules [32-38]. Among them bis-indole alkaloids have been found to show a large range of medicinal activities such as anti-bacterial [39], anti-malarial [40], anti-fungal [41], antileishmanial [42], anti-cancer [43], and anti-inflammatory [44], etc. Considering the drawbacks mentioned above with regard to performing reactions in organic solvents, we ventured to develop a heterogeneous catalyst composed of non-toxic and non-noble materials, which can perform organic reactions in water (a Green solvent) at ambient

Page 3 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 4 of 52

conditions with good yields even at short reaction times. In this regard, considering the unique properties of graphene-based materials and the 2D transition metal dichalcogenides, we have prepared MoS2-RGO nanocomposite, a novel high surface area 2D-2D hybrid material, using a facile hydrothermal method and have demonstrated it as a highly efficient heterogeneous Green catalyst for the synthesis of indole alkaloids, which are pharmaceutically relevant natural products. We have also performed detailed investigations using multiple substrates and varying the reaction parameters to better understand the activity of the prepared catalyst and its mechanism of action. In addition, we have also demonstrated the recyclability of the catalyst by taking a model reaction. Finally two different gram scale reactions were performed to show the potential and industrial applicability of the developed reaction. Based on the results obtained and insights gained from this work, we have finally proposed some strategies for the development of heterogeneous catalysts for performing organic transformation reactions under Green conditions.

RESULTS AND DISCUSSION Synthesis and Characterization of the Catalyst The crystal phase of the as-prepared nanocomposite and MoS2 precursor was analyzed by powder X-ray diffraction (XRD) and the data are presented in Figure 1. MoS2 shows diffraction peaks at 2θ = 14o, 32o and 57o, which can be assigned to the (002), (100) and (110) reflection planes, respectively. The (002) reflection plane correspond to the d-spacing of about 0.62 nm and the broader nature of this diffraction peak indicates the presence of layered MoS2 having lamellar structure. However MoS2-RGO nanocomposite also display the same diffraction peaks as that of pure MoS2, in addition to these peaks the characteristics diffraction peak due to RGO around 2θ = 24o can also be seen confirming the presence of RGO. Raman spectrum of the as-prepared MoS2-RGO nanocomposite is presented in Figure 2 along with the spectrum of MoS2 precursor. MoS2 shows the presence of prominent Raman peaks around 280 cm-1, 332 cm-1 and 376 cm-1, which could be assigned to E1g, A1 and E2g1 vibrational mode of S atoms with respect to Mo atom, respectively. The Raman peak around 454 cm-1 is a typical peak of MoS2 and assigned as 2LA (M). The MoS2-RGO nanocomposite also Page 4 of 52 ACS Paragon Plus Environment

Page 5 of 52

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

exhibit the same characteristic Raman peaks as that of MoS2 and in addition to these peaks, this composite also displays Raman peaks around 1344 cm-1 and 1584 cm-1, which originates due to D-band and G-band of RGO, respectively. The presence of D-band and G-band confirms the presence of reduced graphene oxide (RGO) in MoS2-RGO nanocomposite. Thus Raman results also confirm the successful fabrication of MoS2-RGO nanocomposite.

Figure 1. XRD patterns of MoS2 and MoS2-RGO nanocomposite.

Page 5 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 6 of 52

Figure 2. Raman spectra of MoS2 and MoS2-RGO nanocomposite.

The morphology of bare GO, MoS2 and MoS2-RGO nanocomposite was investigated by scanning electron microscopy (SEM) and the obtained images are presented in Figure 3. The SEM image of GO sheets (Figure 3a) shows wrinkled morphology due to the presence of various oxygen containing functional groups on its surface, while MoS2 micrograph (Figure 3b) shows sheet-like morphology. In the SEM image MoS2-RGO nanocomposite (Figure 3c), layered sheetlike MoS2 can be seen on RGO nanosheets (GO reduced to RGO during hydrothermal synthesis).. It can be evidenced from the SEM images that the layered MoS2 is well adhered to RGO nanosheets to form nanocomposite as confirmed by other characterization techniques as well. The presence of all constituent elements (Mo, S, C and O) in MoS2-RGO nanocomposite has been confirmed by energy dispersive x-ray spectrum EDAX analysis (Figure 3d).

Page 6 of 52 ACS Paragon Plus Environment

Page 7 of 52

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

Furthermore, the transmission electron microscopy (TEM) analysis was also performed on GO, MoS2 and MoS2-RGO nanocomposite to evaluate their surface morphology at nm-level and the obtained data are presented in Figure 4. Herein, Figure 4a and 4b shows GO nanosheets and layered MoS2, respectively. In MoS2-RGO nanocomposite, layered MoS2 well adhered on the surface of RGO nanosheets can be seen clearly in Figure 4c. The presence of all constituent elements Mo, S, C and O was further confirmed by the EDAX analysis (Figure 4d), which reveals the successful formation of MoS2-RGO nanocomposite.

Figure 3. SEM Images (a) GO sheets, (b) MoS2, (c) MoS2-RGO nanocomposite and (d) EDAX data of MoS2-RGO nanocomposite.

Page 7 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 8 of 52

Figure 4. TEM Images (a) GO sheets, (b), MoS2, (c) MoS2-RGO nanocomposite and (d) EDAX data of MoS2-RGO nanocomposite.

X-ray photoelectron spectroscopy (XPS) studies were performed with MoS2-RGO nanocomposite in order to investigate the surface chemical composition and valence state in nanocomposite and the obtained data are presented in Figure 5. The survey spectrum (Figure 5a) of MoS2-RGO nanocomposite confirms the existence of C-1s, O-1s, Mo-3d and S-2p in nanocomposite. The deconvoluted C-1s spectrum (Figure 5b) displays the binding energy peaks at 284.4 eV and 287.9 eV, which could be assigned to C-C bonds in RGO and C-O-C bonds, respectively [45]. It is noteworthy to mention here that the characteristic binding peak for C=O (287.41 eV) almost diminishes here signifying the reduction of GO to RGO during the hydrothermal synthesis process. Therefore, it could be concluded that the formation of the Page 8 of 52 ACS Paragon Plus Environment

Page 9 of 52

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

RGO from the GO during the hydrothermal synthesis. The deconvoluted O-1s spectrum (Figure 5c) shows the presence of two peaks at 531.0 eV and 529.1 eV, wherein the peak at 531.0 eV could be attributed to lattice oxygen contribution from RGO and the peak at 529.1 eV corresponds to the non-stoichiometric peak due to O. Figure 5d shows the Mo-3d spectrum, wherein two binding energy can be observed at 231.7 eV and 228.2 eV, which corresponds to the Mo 3d3/2 and Mo 3d5/2, respectively. The S-2p spectrum in Figure 5e presents two binding peaks at 162.4 eV and 161.1 eV. These binding energy peaks in case of Mo-3d and S-2p signify the presence of Mo4+ and S2- ions in MoS2-RGO nanocomposite. The two peaks of the S2- could be due to the S from the MoS2 and the other S of MoS2 linked with the RGO, i.e. possessing an electronic interaction with the O-of the RGO. The O being more electronegative than the S will draw the electron clouds towards itself, thereby in all probabilities the 161.1 eV peak could be assigned the S possessing interaction with RGO. Mo is having two oxidation states here and the binding energy of the Mo2+ (MoS2) is at 228.2 eV, which is in line with the previous literatures [45]. The other oxidation sate of Mo is that of Mo5+, suggests that either during the hydrothermal synthesis a part of the MoS2 is oxidized to the of Mo2O5 or that the Mo also has a very strong interaction with that of the O of RGO [46]. The O-1s peak is that of the O of RGO, which is at the 530.0 eV. However, the non-stoichiometry in the O could be result of the O being attached to the Mo partially i.e., suggesting a Mo-RGO interaction along with that of the S-RGO interaction.

The elemental composition of MoS2-RGO nanocomposite was also

determined from XPS studies and the obtained results are presented in Table 1. The relative percentage of the constituent elements present in MoS2-RGO nanocomposite matches well with the expected stoichiometry. Table 1. Elemental composition of MoS2-RGO nanocomposite as determined from XPS Catalyst MoS2-RGO

Element C-1s O-1s S-2p Mo-3d

Peak positions (eV) 284.4 and 287.9 531.0 and 529.1 161.1 and 162. 4 228.2 and 231.7

Atom % 38.55 28.51 21.65 11.29

Page 9 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 10 of 52

Figure 5. X-ray photoelectron spectra (XPS) for the MoS2-RGO nanocomposite: (a) Survey spectrum, (b) C-1s, (c) O-1s, (d) Mo-3d and (e) S-2p. Page 10 of 52 ACS Paragon Plus Environment

Page 11 of 52

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

ACS Sustainable Chemistry & Engineering

Figure 6. TGA curves of MoS2 and MoS2-RGO nanocomposite.

The thermal stability of MoS2 and MoS2-RGO nanocomposite were studied by thermogravimetric analyses (TGA) in N2 atmosphere and the obtained data have been presented in Figure 6. MoS2 shows initial weight loss around 100oC, which is attributed to the evaporation of adsorbed water molecules. Furthermore, MoS2 shows weight loss around 300oC, which could be attributed to oxidation of MoS2 into MoO3. MoS2-RGO nanocomposite shows continuous weight loss from initial temperature due to loss of co-intercalated water molecules. Sudden weight loss at 500oC occurs due to oxidation of MoS2 and removal of organic groups. Beyond 700oC, stability of nanocomposite decreases due to presence RGO. Based on a literature procedure [47], the relative content of the constituent compounds of the nanocomposite could be estimated from the weight loss observed in the TGA curves. In the case of MoS2-RGO (1:1) nanocomposite, the corresponding values determined were 44.35% MoS2 and 55.65% of RGO. Hydrophilicity study of ZnO-RGO nanocomposite was carried out in an earlier study [12] by contact angle measurement, in which ZnO-RGO nanocomposite showed 74° water contact angle. Inspired from the above study we measured the contact angle for MoS2-RGO nanocomposite and it was found to be 67° (Figure 7), which indicates slightly more Page 11 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 12 of 52

hydrophilic nature of MoS2-RGO nanocomposite in comparison to previously reported ZnO-RGO nanocomposite [12]. Hydrophilicity favors the water molecules to get attached on the surface of the catalyst and thereby enhance the reactivity.

Figure 7. Water contact angle on the surface of MoS2-RGO nanocomposite.

Catalytic Activity Studies Scheme 1. MoS2-RGO nanocomposite catalyzed synthesis of bis and tris indolylmethanes

The catalytic potential of MoS2-RGO nanocomposite has been explored for the Green synthesis of various bis indolylmethanes (BIM) and trisindolylmethanes (TIM), as these molecules are of high significance and are found in the skeleton of various natural products and drugs [48-50]. The catalytic activity of MoS2-RGO nanocomposite was examined on indole and

Page 12 of 52 ACS Paragon Plus Environment

Page 13 of 52

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

benzaldehyde as an initial model reaction by using 5 wt% of catalyst to obtain a phenyl bis indolylmethane at room temperature in water as reaction medium (Scheme 1). NMR based confirmation of the product was also verified by the single crystal x-ray data of 3a (CCDC No. 1517905). Optimization of reaction conditions are shown in Table 2, wherein it can be evidenced that reaction was optimized with GO, MoS2 and MoS2-RGO nanocomposite in water at room temperature and varying different concentration of water. Table 2. Optimization MoS2-RGO nanocomposite catalyzed synthesis of bisindolylmethanes

Entry

R

Solvent

Catalyst

Time

Conc. (M)

(5 wt%)

(in h)

Temperature

% Yield

1

H

0.50

GO

24

RT

10

2

H

0.50

MoS2

6

RT

52

3

H

0.50

MoS2

12

RT

63

4

H

0.50

MoS2-RGO

2

RT

97

2

RT

85

2

RT

70

2

RT

82

4

RT

93

(1:1) 5

H

0.25

MoS2-RGO (1:1)

6

H

0.10

MoS2-RGO (1:1)

7

Me

0.50

MoS2-RGO (1:1)

8

Me

0.50

MoS2-RGO (1:1)

Page 13 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

9

Me

0.25

MoS2-RGO

Page 14 of 52

2

RT

74

2

RT

61

2

RT

87

2

RT

74

2

RT

72

2

RT

62

2

RT

54

2

RT

45

48

RT

0

(1:1) 10

Me

0.10

MoS2-RGO (1:1)

11

H

0.50

MoS2-RGO (1:2)

12

Me

0.50

MoS2-RGO (1:2)

13

H

0.50

MoS2-RGO (1:3)

14

Me

0.50

MoS2-RGO (1:3)

15

H

0.50

MoS2-RGO (1:4)

16

Me

0.50

MoS2-RGO (1:4)

17

H

0.50

No catalyst

Due to slightly acidic nature of GO [51] reaction proceeded with simple GO also but percentage yields of isolated was very poor (Table 2, entry 1). Then the reaction was carried out with MoS2 alone as a catalyst, which also led to formation of product up to a significant yield (52%) in 6 h (Table 2, entry 2), but further continuation of reaction for a period of 12 h, (Table 2, entry 3) did not lead to significant improvement in the yield (63%). Furthermore, some unknown side products were formed before consumption of total starting material. Surprisingly, when the same reaction was performed with MoS2-RGO (1:1) nanocomposite, it was completed in just two hours (Table 2, entry 4) with 97% isolated yield. Subsequently, various other combination of solvent concentration were examined (Table 2, entries 5- 10) and better yields were found with 0.5 M concentration of the solvent for both indole and N-methyl indole substrates (Table 2, entries 4 & 8). In order to ascertain the effect of RGO content on the Page 14 of 52 ACS Paragon Plus Environment

Page 15 of 52

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

catalytic activity of the MoS2-RGO nanocomposite, three additional nanocomposites of MoS2RGO in weight ratios 1:2, 1:3 and 1:4 were prepared and their catalytic activity were examined (Table 2, entries 11-16). It could be evidenced that the best yields for the synthesis of bisindolylmethanes were obtained when MoS2-RGO (1:1) nanocomposite was used as catalyst. Moreover, the amount of catalyst optimized for the reaction was found to be 5 wt%, and it could be recycled up to 6 cycles without significant loss in its activity (Figure 8). Recycling of catalyst was congenial due to magnetic properties of MoS2 [52], which helps the catalyst to stick to the magnetic stir bar slowly as the reaction proceeds. To ascertain this, we measured the magnetization versus magnetic field (M-H) curves for MoS2 and MoS2-RGO (1:1) nanocomposite and the obtained data are shown in Figure S1 (supporting information), which reveals the room temperature magnetic behavior of the catalyst. Furthermore, to explore the generality of this reaction, various aromatic, aliphatic and heterocyclic aldehydes were examined (Table 3) and most of the aldehydes, resulted in good to excellent yields.

Figure 8. Recyclability study of MoS2-RGO nanocomposite catalyst for the synthesis of 3a.

Page 15 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 16 of 52

Figure 9. Progress of the reaction of Indole with aldehydes to form bisindolylmethanes: (a) reaction of indole with aromatic aldehydes and (b) reaction of indole with aliphatic aldehydes.

Page 16 of 52 ACS Paragon Plus Environment

Page 17 of 52

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

ACS Sustainable Chemistry & Engineering

Table 3. Substrate scope with various aldehydes

N N Me 3b, 91% Me

NH

HN 3a, 97%

N Et

3c, 90%

N

N Et

Allyl

N n-Hex

N n-Bu

N n-Hex

3f, 88%

CH3

OH

NH

HN

NH

N Me

NH 3m, 92%

HN

3l, 95% Cl

3e, 91%

N n-Pr

OMe

N 3j, 88% Et

N

N 3i, 88% Me

Et

F

F

HN

3k, 97%

N Me

NH 3h, 89%

HN

3g, 82%

n-Pr

OMe

OMe

N n-Bu

N

N 3d, 92% Allyl

F

N

N Me

Et

3n, 93%

3o, 92%

N Et

No2

CN

O Cl

NH

HN

NH

HN 3p, 85%

NH

HN

3q, 93%

HN

N H 3t, 85%

N H

NH 3s, 96%

3r, 92%

OMe NH

S

N H

N H 3u, 82%

OMe

MeO

3v, 79%

MeO

N Et

N

NH

HN

NH

HN

OMe

MeO

Et

N Me

3x, 92%

3w, 92%

OMe

N Me

3y, 86%

OMe MeO

OMe

HN N Et

HN

NH

N 3z, 83% Et

N

NH 3zb, 94%

3za, 98%

N

N 3zc, 91%

N

3zd, 89%

OMe Cl F N

N Allyl

Allyl

N Bn

3ze, 87%

Br

N Bn

N Bn

3zf, 98%

Br

3zg, 99%

NC

CN

HN

N Bn

O2N

NH

NH 3zj, 93%

HN

NH 3zk, 99%

HN

NH 3zi, 95%

3zh, 91%

NO2 H N

HN

Cl

F

HN

H N

H N N

NH 3zl, 97%

H N

3zm, 0%

N

3zn, 0%

Page 17 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 18 of 52

Furthermore, the electron donating substituent on indole ring also favored the reaction to proceed well and produced good yields. Electron withdrawing atom inside indole ring structure (e.g. 7-azaindole) disfavored the reaction. But interestingly, electron withdrawing substituent (such as F, Cl, Br, CN and NO2) outside of the indole ring structure has no effect on the reactivity for the synthesis of bisindolylmethanes. All the indoles with electron withdrawing group were found to give the desired product in excellent yields (Table 3, 3zh- 3zl). Moreover, there was no significant effect on reactivity due to the presence of electron donating or withdrawing group present on the aldehyde part unlike indole ring. In addition to this, the reaction proceeded equally good even in the presence of various alkyl groups on the indole nitrogen atom. To our delight, we synthesized four indole alkaloids Turbomycin B reductant [53] (3a), Arundine [54] (3y) Vibrindole A [55] (3z) and trisindole (3x ) [2, 56, 57] by this method in very good yields. Similarly, various other derivatives of these indole alkaloids were also synthesized (3b-3g and 3y-3za) by our developed protocol. Progress of the reaction (Figure 9) was monitored as a function of time and it was found that aromatic aldehydes react more efficiently with indole than the corresponding aliphatic aldehydes. Furthermore, we tried several other benzo-fused electron rich systems such as 3methylindole, benzimidazole, indazole and some five membered electron rich heterocycles such as imidazole, furan and thiophene, but none of them led to the formation of desired products. In the case of 3-methylindole, position-3 which is more nucleophilic in nature is substituted by a methyl group, and the position-2 being less nucleophilic could not facilitate the reaction. Similarly, in the case of imidazole and benzimidazole, the carbon atom is present between the two nitrogen atoms at position-2, hence nucleophilicity might be suppressed by the electron withdrawing effect of nitrogen at position-3. The specific reason for the other five membered heterocycles not yielding the desired product could not be unambiguously ascertained.

Page 18 of 52 ACS Paragon Plus Environment

Page 19 of 52

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

ACS Sustainable Chemistry & Engineering

Scheme 2. Synthesis of unsymmetrical bis- indolylmethanes

Table 4. Optimization MoS2-RGO nanocomposite catalyzed synthesis of unsymmetrical bisindolylmethanes O

MeO N H

+

water, RT

0.5 mmol

0.5 mmol

1aa

1a

OMe

Cat. (wt%)

+ N H

0.5 mmol

Entry

H

+ 3a + 3w

4h NH

HN

2a

4aa

Solvent

Catalyst

Time

Conc. (M)

(wt%)

(in h)

1

0.50

4

4

2

0.25

4

3

0.50

4

Temperature

% Yield

% Yield

( 4aa)

(3a+3w)

RT

51

48

4

RT

58

40

5

4

RT

52

46

0.25

5

4

RT

61

37

4

0.50

6

4

RT

54

44

5

0.25

6

4

RT

63

36

6

0.25

6

6

RT

55

43

7

0.25

8

4

RT

52

47

8

0.25

10

4

RT

48

51

9

0.25

20

4

RT

40

59 Page 19 of 52

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

10

0.25

No catalyst

24

RT

Page 20 of 52

0

0

To extend the scope of the developed catalytic system, we tried to synthesize unsymmetrical bisindolylmethanes. After careful optimization of reaction conditions we were able to synthesize unsymmetrical bisindoles having electron withdrawing, electron donating as well as halogen substituent in the product (Scheme 2, Table 4). The isolated yields for unsymmetrical bisindolylmethanes were found to be moderate to good, ranging from 61-72 %. Finally we examined our catalytic system for the development of ketone derived bisindolylmethanes. After several permutations and combinations of amount of catalyst and reaction temperature, we were able to optimize the reaction conditions for the synthesis of ketone derived bisindoles. Slow reactivity of ketones was observed in comparison to the aldehydes for the synthesis of bisindolylmethanes (Scheme 3, Table 5). Due to slow and low reactivity of ketones, reaction was optimized at 45oC. Prolonged reaction at elevated temperature led to the formation of a series of side products. Among ketones, cyclohexanone, cyclopentanone and acetone reacted faster and produced moderate to good yields. But the reactivity of 9-fluorenone and acetophenone was sluggish in comparison to aliphatic ketones. Surprisingly, no reactivity was observed with benzophenone.

Page 20 of 52 ACS Paragon Plus Environment

Page 21 of 52

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

ACS Sustainable Chemistry & Engineering

Scheme 3: Ketone derived synthesis of bis- indolylmethanes

Table 5. Optimization of ketone derived synthesis of bisindolylmethanes

Entry

Solvent

Catalyst

Time (in h)

Temperature

% Yield

Conc. (M)

(wt%)

1

0.50

5

4

RT

20

2

0.25

5

4

RT

32

3

0.50

6

4

RT

23

4

0.25

6

4

RT

35

5

0.25

8

4

RT

43

6

0.25

10

4

RT

48

7

0.25

10

6

RT

54

8

0.25

10

4

45oC

72

9

0.25

10

4

60oC

57 Page 21 of 52

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

10

0.25

20

4

Page 22 of 52

RT

44

Industrial utility of the developed strategy was demonstrated by carrying out a gram scale reaction (Scheme 2). Starting from 1 g of indole, 0.99 g (95%) of Arundine and 1.02 g (92%) of Vibrindole A natural products were synthesized. These results illustrate the potential of the developed catalyst for industrial applications to perform large scale synthesis of these molecules of high significance at ambient conditions using water as solvent. For a standard Green chemistry reaction atom economy/ reaction catalysts [58-62] for mass efficiency should be high and environmental factor as well as process mass intensity should be low. When Green chemistry matrices were calculated for our optimized reaction, we found high atom economy (A.E. = 94.7%)/ reaction mass efficiency (R.M.E = 70.9%) and small E- factor (= 0.089)/ Process mass intensity (P.M.E. = 1.089) (refer supporting information section for the metrics calculations). The comparison of the activity of the catalyst with various other reported support based catalyst for the synthesis of 3,3'-(phenylmethylene)bis(1H-indole) is shown in the Table 6. The comparison indicates that our developed protocol is mild, efficient, recyclable and is equally competitive to other reported methods.

Scheme 4. Gram scale synthesis of Arundine (3w) and Vibrindole A (3x)

Table 6. Comparison of catalytic activity of MoS2-RGO nanocomposite with other reported support based catalysts for the synthesis of 3,3'-(phenylmethylene)bis(1H-indole) Entry

Catalyst

Time

Temp.

%Yield

Reference Page 22 of 52

ACS Paragon Plus Environment

Page 23 of 52

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

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9

Sulfonated carbon/nano metal-oxide composite Mesoporous benzene silica Pd-silica cellulose complex Carbohydrate based tosylsulfonyl hydrazide Poly(ammonium methane sulfonate) Cu-tmtppa based catalyst PAN supported ionic liquid TPPMS/CBr4 MoS2-RGO

4h

100°C

90

[63]

20 min 3h 12 h

60°C 100°C RT

95 87 84

[64] [65] [66]

4h

RT

96

[67]

2h 4h 4h 2h

90°C RT RT RT

92 96 95 97

[68] [69] [70] This work

The proposed catalytic cycle (Scheme 5) shows that the initial aldehyde moiety becomes more electrophilic, due to in the presence of MoS2 on the surface of RGO and gets attacked by highly nucleophilic 3-position of indole via a Friedel–Craft pathway and forms a hydroxyalkyated intermediate 4, which gives off an hydroxyl ion to form an intermediate 5, which is known as an enamine [71-73] intermediate. The enamine intermediate 5, being highly electrophilic in nature act similar to a Michael acceptor and get attacked by another indole molecule to form a bis-indolylmethane moiety with the removal of a water molecule and the catalyst was set free for the next cycle.

Page 23 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 24 of 52

Scheme 5. Mechanism of catalytic activity of MoS2-RGO nanocomposite RGO

MX

RGO

R2

MX

O H

H

O

MX R2

- H+

N R1

R2

O

N R1

2

N R1

1

RGO

H+

3

RGO

R1

MX N

RGO

MX

R1 N

8

R2

R2

O

H

4

N R1

7

- H+

R1

H2O

- OH-

N H R1 N H

R1 N

5

6

R1 O MX RGO

R2

N

H O

R1 N

R2

MX

H O

R2

RGO

MX RGO

CONCLUSIONS In summary, we report the use of MoS2-RGO nanocomposite as an environmentally benign, highly efficient heterogeneous catalyst for the synthesis of indole alkaloids in water in a one-pot reaction. The nanocomposite catalyst was prepared using a facile hydrothermal method and has been characterized for its structure, morphology and surface properties using x-ray diffraction, Raman spectroscopy, SEM, TEM, EDAX, XPS, TGA and contact angle. Green chemistry matrices calculations showed high atom economy (A.E. = 94.7%) and small E-factor (0.089) for the reactions. Potential industrial use of the catalyst was validated by performing a gram-scale synthesis of natural products. In addition, the recyclability of the catalyst, without Page 24 of 52 ACS Paragon Plus Environment

Page 25 of 52

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

any significant loss of activity, was also demonstrated. The developed nanocomposite perhaps can also be used to catalyze several other industrially relevant organic transformation reactions in the future.

EXPERIMENTAL SECTION Instrumentation X-ray diffraction (XRD) measurements were performed using the Rigaku Smart Lab 9kW rotating anode x-ray diffractometer with Ni-filtered Cu Kα irradiation (λ = 0.1542 nm) at 45 kV and 100 mA in 2θ ranging from 10o - 80o with a scan rate of 2o per minute with stepping size of 0.02o. Raman spectroscopic measurements were done using Horiba LabRAM high resolution UV-VIS-NIR instrument using 633 nm laser excitation. Morphology of the samples was investigated by using scanning electron microscope (SEM), FEI Nova Nano SEM-450 and transmission electron microscope (TEM), FEI Tecnai G2 20 S-twin microscope operating at 200 kV. Energy dispersive x-ray spectra (EDAX) were obtained using the same SEM and TEM instruments. X-ray photoelectron spectroscopic (XPS) measurements were done using the SPECS instrument with a PHOBIOS 100/150 delay line detector (DLD) with 385W, 13.85 kV and 138.6 nA (sample current). We have used Al Kα (1486.6eV) dual anode as the source. The XPS was taken with pass energy of 50 eV. As an internal reference for the absolute binding energy, the C-1s peak (284.5 eV) was used. The data obtained from the instrument was processes using the CASA software. The survey peak of the specimens was initially corrected for the C-1s peak and post that the quantitative estimate for the elemental analysis was performed. The XPS peaks were deconvoluted using a combination of Gaussian and Lorentzian curve [GL(30)]- and the baseline correction for each spectra was initially carried out using a Shirley background correction using the CASA software. Thermogravimetric (TGA) analyses were carried out by using a Perkin Elmer Pyris 1 instrument. The samples were heated from room temperature to 1100°C at a heating rate of 10oC min−1 under nitrogen atmosphere with a flow rate of 20 mL min−1. A small amount of the sample (2-3 mg) was kept in a standard platinum crucible and an empty crucible was used as reference. Contact angle measurements were performed using the Phoenix 300 contact angle instrument. Magnetic properties (M-H curve) were measured

Page 25 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 26 of 52

using a Quantum design Inc. magnetic property measurement system (MPMS). Melting points were determined using open capillary method. NMR spectra were recorded on a JEOL-USA (JNMECX500) spectrometer in CDCl3-d1 taking TMS (tetramethyl silane) as an internal standard. The NMR chemical shift was reported in ppm relative to 7.26 ppm and 77.00 ppm of CDCl3 solvent as the standards for 1H and 13C spectra, respectively. 1H spectra were recorded in 500 MHz frequencies and

13

C NMR spectra were recorded in 125 MHz frequencies. Coupling

constant ‘J’ was calculated in Hz. FT-IR spectra were acquired on a Perkin-Elmer Spectrum two spectrometer. Mass spectra were recorded on an advance Bruker Daltonics (impact HD) UHRQqTOF (Ultra-High Resolution Qq-Time-Of-Flight) mass spectrometer. Deionized water (18.2 MΩ-cm) used in synthesis of catalyst was obtained from double stage water purifier (ELGA PURELAB Option-R7). Single crystal X-ray data was obtained by using Agilent Supernova SuperNova E(Dual) Diffactometer System.

Chemicals All the required chemicals were purchased from Aldrich, Fluka, Loba, Merck and TCI suppliers and some of them were purified by column chromatography prior to use. For GO synthesis, graphite powder (crystalline, -300 mesh, 99%) was purchased from Alfa Aesar, whereas sodium nitrate (NaNO3), sulphuric acid (H2SO4), potassium permanganate (KMnO4) and hydrogen peroxide (H2O2) were purchased from Merck. Sodium molybdate dihydrate (Na2MoO4 .2H2O) was supplied by Merck. L-cysteine (98%) and hydrochloric acid (HCl) used in synthesis were purchased from Alfa Aesar and Fischer Scientific, respectively.

Synthesis of Graphene Oxide Graphene oxide (GO) was synthesized from natural graphite flakes using modified Hummers’ method [74]. In brief, 1.0 g of graphite powder and 0.5 g of NaNO3 was stirred in 23 mL of concentrated H2SO4 in an ice bath to maintain the reaction temperature below 10oC. This was followed by the slow addition of 3.0 g of KMnO4 to the reaction mixture with continuous stirring. Subsequently, the reaction mixture was stirred in an oil bath at 35oC until a brown colored paste was formed in about 4 h. Now, the reaction was terminated by slow addition of

Page 26 of 52 ACS Paragon Plus Environment

Page 27 of 52

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

deionized water (90 mL), which increased the temperature to 95-98oC and resulting suspension was maintained at this temperature for 15-20 min; subsequently, the suspension was then diluted to about 250 mL by addition of deionized water. This is followed by the addition of 10 mL H2O2 to remove unreacted KMnO4 in the reaction mixture. In order to remove the ions of oxidant origin, the mixture was washed with 10% HCl and then with deionized water until pH value of the filtrate was neutral. Obtained graphite oxide was exfoliated by sonication (1mg/mL in aqueous solution) and then centrifuged at 4500 rpm for 15 min. The supernatant was decanted and dried with rotary evaporator at 40oC followed by vacuum drying at same temperature to obtain GO powder.

Synthesis of MoS2 Nanosheets MoS2 nanosheets were prepared by hydrothermal method using a reported procedure [75]. In brief, 0.20 g of Na2MoO4.2H2O and L-cysteine (0.4 g) were mixed and kept for stirring for about 1 h. Then the reaction mixture was transferred into a Teflon-lined stainless steel autoclave and maintained at 180oC for 24 h. The obtained product was recovered by centrifugation and washed with deionized water thrice. The final material was dried at 80oC overnight to obtain MoS2-RGO nanocomposite.

Synthesis of MoS2 -RGO Nanocomposite For the preparation of MoS2-RGO nanocomposite, same hydrothermal strategy was adopted as for the preparation of MoS2 nanosheets. Initially, 0.2 g of as-prepared GO was well dispersed in deionized water with the help of ultra-sonication, which was followed by addition of 0.10 g of NaMoO4.2H2O and L-cysteine (0.2 g). This reaction mixture was maintained at 180oC for 24 h in a teflon-lined stainless steel autoclave. The obtained product was recovered by centrifugation and washed with deionized water and ethanol thrice. The final product was dried at 80oC overnight to obtain MoS2-RGO nanocomposite.

Typical Procedure for the Synthesis of Bisindolylmethanes (3a-3zg)

Page 27 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 28 of 52

A mixture of indole (1, 0.5 mmol), aldehydes (2, 0.25 mmol), MoS2-RGO catalyst (5 wt%), and water (1.1 mL) were mixed in a 5 mL round-bottomed flask and stirred at ambient temperature. Progress of the reaction was monitored using thin layer chromatography (TLC). After completion of reaction, mixture was decanted to remove water, and ethanol was added to the reaction mixture. Solid catalyst was separated by centrifugation. The organic layer was dried over Na2SO4, and the solvent was removed under reduced pressure using rotary evaporator. Purification of the crude reaction mixture was done by column chromatography using a suitable eluent mixture. Isolated yields for each synthesized molecule have been given below in the characterization part. It is noteworthy to mention that, during the process of establishing this Green synthesis strategy, we developed an economical and less solvent consuming method to purify the crude reaction mixture. In this method, the crude reaction mixture is triturated with hexane (2 x 3 mL) followed by washing with 5:95 ratio mixture of EtOAc and hexane (2 x 2 mL). All the new compounds were characterized by MP, NMR, IR, and HRMS, while all the known compounds were confirmed by 1H and 13C NMR only.

Compound Characterization (1)

3-((1H-indol-3-yl)(phenyl)methyl)-1H-indole (Turbomycin B reductant) [76-78]

White solid; 74 mg (97%, Rf = 0.57, 30:70 EtOAc/Hex) m.p. 125-127 °C; 1H NMR (CDCl3, 500 MHz) δ 7.90 (br s, 2H)), 7.39-7.33 (m, 6H), 7.29-7.25 (m, 2H), 7.22-7.15 (m, 3H), 7.00 (t, 2H, J = 7.5 Hz), 6.65 (d, 2H, J = 1.4 Hz), 5.88 (s, 1H); 13C NMR (CDCl3, 125 MHz) δ 143.9, 136.6, 128.7, 128.2, 127.0, 126.1, 123.6, 121.9, 120.0, 119.7, 119.2, 111.0, 40.1.

(2)

1-methyl-3-((1-methyl-1H-indol-3-yl)(phenyl)methyl)-1H-indole [76, 79]

Whitish pink solid, 66.5 mg (91%, Rf = 0.72, 20:80 EtOAc/Hex) m.p.182-185 °C; 1H NMR (CDCl3, 500 MHz) δ 7.37 (d, 2H, J = 7.5 Hz), 7.34 (d, 2H, J = 7.5 Hz), 7.29-7.24 (m, 4H), 7.21-7.17 (m, 3H), 6.98 (d, 2H, J = 6.8 Hz), 6.52 (s, 2H), 5.87 (s, 1H), 3.67 (s, 6H); 13C NMR (CDCl3, 125 MHz) δ 144.4, 137.4, 128.6, 128.2, 128.1, 127.4, 125.9, 121.4, 120.0, 118.6, 118.2, 109.0, 40.0, 32.7.

Page 28 of 52 ACS Paragon Plus Environment

Page 29 of 52

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

ACS Sustainable Chemistry & Engineering

(3)

1-ethyl-3-((1-ethyl-1H-indol-3-yl)(phenyl)methyl)-1H-indole [80]

White solid, 67 mg (90%, Rf = 0.76, 20:80 EtOAc/Hex) m. p 162-163°C; 1H NMR (CDCl3, 500 MHz): δ = 7.37-7.30 (m, 6H), 7.28-7.25 (m, 2 H), 7.23-7.15 (m, 3H), 6.97 (t, 1H, J = 7.5 Hz), 6.57 (s, 2H), 4.06 (q, 4H, J = 7.5 Hz), 1.37 (t, 3H, J = 7.5 Hz). 13C NMR (125 MHz, CDCl3): δ = 144.4, 136.3, 128.7, 128.1, 127.6,126.6, 125.9, 121.1, 120.2, 118.5, 118.2, 109.1, 40.8, 40.2, 15.5.

(4)

1-allyl-3-((1-allyl-1H-indol-3-yl)(phenyl)methyl)-1H-indol [50]

White solid; 68 mg (92 %, Rf = 0.80, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.37 (d, 2H, J = 8.2 Hz), 7.34 (d, 2H, J = 6.9 Hz), 7.28-7.24 (m, 4H), 7.21-7.14 (m, 3H), 6.97 (t, 2H, J = 6.8 Hz), 6.56 (s, 2H), 5.96-5.89 (m, 1H), 5.84 (s, 1H), 5.12 (dd, 2H, J = 10.3 & 1.3 Hz), 5.01 (dd, 2H, J = 17.1 & 1.3 Hz), 4.61 (d, 4H, J = 4.8 Hz); 13C NMR (CDCl3, 125 MHz) δ 144.2, 136.8, 133.7, 128.7, 128.2, 127.6, 127.4, 126.0, 121.4, 120.1, 118.7, 118.5, 116.8, 109.5, 48.7, 40.1.

(5)

1-methyl-3-(1-(1-methyl-1H-indol-3-yl)ethyl)-1H-indole[50]

Colorless oil; 59 mg (91 %, Rf = 0.80, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.36-7.25 (m, 8H), 7.21-7.14 (m, 3H), 6.96 (t, 2H, J = 5.3 Hz), 6.55 (s, 2H), 5.86 (s, 1H), 3.96 (t, 4H, J = 6.9 Hz), 1.78 (sext, 4H, J = 6.8 Hz), 0.86 (t, 6H, J = 6.8 Hz); 13C NMR (CDCl3, 125 MHz) δ 144.4, 136.6, 128.7, 128.1, 127.5, 127.4, 125.9, 121.1, 120.2, 118.4, 118.0,109.2, 47.8, 40.1, 23.5, 11.5.

(6)

1-butyl-3-((1-butyl-1H-indol-3-yl)(phenyl)methyl)-1H-indole [50]

Colorless oil; 58 mg (88 %, Rf = 0.86, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.36-7.25 (m, 8H), 7.22-7.14 (m, 3H), 6.95 (t, 2H, J = 6.8 Hz), 6.54 (s, 2H), 5.86 (s, 1H), 3.98 (t, 4H, J = 6.8 Hz), 1.72 (sext, 4H, J = 7.5Hz), 1.26 (sext, 4H, J = 7.5Hz), 0.88 (t, 6H, J = 7.5 Hz); 13C NMR (CDCl3, 125 MHz) δ 144.4, 136.6, 128.7, 128.1, 127.5, 127.4,125.9, 121.1, 120.1, 118.4, 118.0, 109.2, 45.9, 40.2, 32.3, 20.1, 13.7.

(7)

1-hexyl-3-((1-hexyl-1H-indol-3-yl)(phenyl)methyl)-1H-indole [50]

Colorless oil; 50 mg (82 %, Rf = 0.90, 15:85 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.36-7.26 (m, 8H), 7.21-7.14 (m, 3H), 6.96 (t, 2H, J= 7.5Hz), 6.54 (s, 2H, J = 6.9 Hz), 5.86 (s, 1H), 3.99 (t, 4H,

Page 29 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 30 of 52

J = 6.9 Hz), 1.74 (t, 4H, J = 6.2 Hz), 1.24 (br s, 12H), 0.84 (t, 6H, J = 6.2 Hz); 13C NMR (CDCl3, 125 MHz) δ 144.5, 136.6, 128.7, 128.1, 127.5, 127.4, 125.9, 121.1, 120.1, 118.4, 118.0, 109.2, 46.2, 40.1,31.3, 30.1, 26.6, 22.5, 14.0.

(8)

3-((1H-indol-3-yl)(4-methoxyphenyl)methyl)-1H-indole

Red solid, 60 mg (89 %, Rf = 0.5, 30:70 EtOAc/Hex) m.p.197-198 °C; 1H NMR (CDCl3, 500 MHz) δ 7.90 (br s, 2H), 7.38 (d, 2H, J = 8.2 Hz), 7.35 (d, 2H, J = 7.5 Hz), 7.24 (d, 2H, J = 7.5 Hz), 7.16 (t, 2H, J = 6.8 Hz), 6.99 (t, 2H, 7.5 Hz), 6.81 (d, 2H, J = 9 Hz), 6.64 (s, 2H), 5.83 (s, 1H), 3.77 (s, 3H); 13

C NMR (CDCl3, 125 MHz) δ 157.8, 136.6, 136.2, 129.6, 127.0, 123.5, 121.8, 120.0, 119.9, 119.1,

113.5, 110.9, 55.2, 39.3.

(9)

3-((4-methoxyphenyl)(1-methyl-1H-indol-3-yl)methyl)-1-methyl-1H-indole [50]

Brown solid, 62 mg (89 %, Rf = 0.58, 25:75 EtOAc/Hex) m.p.220-222 °C; 1H NMR (CDCl3, 500 MHz) δ 7.37 (d, 2H, J = 8 Hz), 7.29-7.23 (m, 4H), 7.19 (t, 2H, J = 8 Hz), 6.98 (t, 2H, J = 8 Hz), 6.81 (d, 2H, J = 8.6 Hz), 6.51 (s, 2H), 5.83 (s, 1H), 3.78 (s, 3H), 3.68 (s, 6H); 13C NMR (CDCl3, 125 MHz) δ 157.8, 137.3, 136.6, 129.5, 128.2, 127.4, 121.3, 120.0, 118.5, 113.5, 109.0, 55.2, 39.2, 32.6.

(10)

1-ethyl-3-((1-ethyl-1H-indol-3-yl)(4-methoxyphenyl)methyl)-1H-indole [50]

Yellowish orange solid, 63 mg (88 %, Rf = 0.65, 25:75 EtOAc/Hex) m.p. 191-192 °C; 1H NMR (CDCl3, 500 MHz) δ 7.36 (d, 2H, J = 8.2 Hz), 7.31 (d, 2H, J = 8.2 Hz), 7.24 (d, 2H, J = 8.2 Hz), 7.17 (t, 2H, J = 6.8 Hz), 6.97 (t, 2H, J = 6.8 Hz), 6.81 (d, 2H, J = 8.2 Hz), 6.56 (s, 2H), 5.82 (s, 1H), 4.05 (q, 4H, J = 7.6 Hz), 3.78 (s, 3H), 1.36 (t, 6H, J = 7.6 Hz); 13C NMR (CDCl3, 125 MHz) δ 157.7, 136.7, 136.3, 129.6, 127.6, 126.6, 121.1, 120.2, 118.5, 118.4, 113.4, 109.1, 55.1, 40.8, 39.3,15.5.

(11)

4-(di(1H-indol-3-yl)methyl)phenol [81-83]

Red solid; 85.7 mg (97 %, Rf = 0.60, 40:60 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.88 (br s, 2H), δ 7.37 (d, 2H, J = 8.2 Hz), 7.33 (d, 2H, J = 8.2 Hz), 7.18-7.14 (m, 4H), 6.99 (t, 2H, J= 8.2 Hz),

Page 30 of 52 ACS Paragon Plus Environment

Page 31 of 52

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

6.71 (d, 2H, J = 8.2 Hz), 6.61 (d, 2H, J= 2.0 Hz), 5.81 (s, 1H); 13C NMR (CDCl3, 125 MHz) δ 136.6, 136.1, 129.7, 126.9, 123.5, 121.8, 119.9, 119.1, 115.0, 111.0, 39.2.

(12)

3,3'-(p-tolylmethylene)bis(1H-indole) [70, 84, 85]

White solid; 83.4 mg (95 %, Rf = 0.40, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.88 (br s, 2H), δ 7.39 (d, 2H, J = 7.5 Hz), 7.34 (d, 2H, J = 8.2 Hz), 7.22 (d, 2H, J= 8.2 Hz), 7.15 (t, 2H, J= 7.6 Hz), 7.07 (t, 2H, J= 8.2 Hz), 6.99 (t, 2H, J = 7.5 Hz), 6.65 (d, 2H, J= 1.4 Hz), 5.84 (s, 1H), 2.31(s, 3H) ; 13C NMR (CDCl3, 125 MHz) δ 140.9,136.6, 135.4, 128.8, 128.5, 127.0, 123.5, 121.8, 119.9, 119.8, 119.2, 110.9, 39.7, 21.1.

(13)

3-((4-fluorophenyl)(1H-indol-3-yl)methyl)-1H-indole [80, 86]

White solid, 49.5 mg (92 %, Rf = 0.58, 30:70 EtOAc/Hex) m.p.73-75°C; 1H NMR (CDCl3, 500 MHz) δ 7.93 (s, 2H), 7.36-7.38 (d, 4H, J = 8.2 Hz), 7.31-7.28 (m, 2H), 7.18 (t, 2H, J = 7.5 Hz), 7.01 (t, 2H, J = 7.5 Hz), 6.96 (t, 2H, J = 8.2 Hz), 6.64 (s, 2H), 5.87 (s, 1H); 13C NMR (CDCl3, 125 MHz) δ 162.2, 160.4, 139.6, 136.6, 130.1, 130.0, 126.9, 123.5, 122.0, 119.8, 119.5, 119.2, 115, 114.8, 111.1, 39.4.

(14)

3-((4-fluorophenyl)(1-methyl-1H-indol-3-yl)methyl)-1-methyl-1H-indole [76]

Brown solid, 57 mg (93 %, Rf = 0.71, 20:80 EtOAc/Hex) m.p. 202-204°C; 1H NMR (CDCl3, 500 MHz) δ 7.36 (d, 2H, J = 7.5 Hz), 7.31-7.26 (m, 4H), 7.21 (t, 2H, J = 7.5 Hz), 7.01 (t, 2H, J = 8.2 Hz), 6.96 (t, 2H, J = 8.2 Hz), 6.51 (s, 2H), 5.87 (s, 1H), 3.69 (s, 6H); 13C NMR (CDCl3, 125 MHz) δ 162.3, 160.3, 140.1, 137.4, 130.0, 129.9, 128.2, 127.2, 121.5, 119.9, 118.7, 118.1, 115,114.8,109.1, 39.3, 32.7

(15)

1-ethyl-3-((1-ethyl-1H-indol-3-yl)(4-fluorophenyl)methyl)-1H-indole [50]

Whitish Pink solid, 59.5 mg (92 %, Rf = 0.76, 20:80 EtOAc/Hex) 197-198 °C; 1H NMR (CDCl3, 500 MHz) δ 7.35-7.25 (m, 6H), 7.18 (t, 2H, J = 7.5 Hz), 6.99-6.93 (m, 4H), 6.55 (s, 2H), 5.87 (s, 1H), 4.06 (q, 4H, J = 7.6 Hz), 1.37 (t, 6H, J = 7.6 Hz); 13C NMR (CDCl3, 125 MHz) δ 162.2, 160.3, 140.1,

Page 31 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 32 of 52

136.4, 130.1,130.0, 127.4, 126.6, 121.3, 120.1, 118.6, 118.1, 114.9, 114.8, 109.2, 40.8, 39.5,15.5.

(16)

3,3'-((2-chlorophenyl)methylene)bis(1H-indole) [70, 85, 87, 88]

White solid; 79 mg (85 %, Rf = 0.70, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.91 (br s, 2H), δ 7.40-7.34 (m, 4H), 7.25-7.01 (m, 6H), 6.63 (s, 2H ), 6.34 (s, 1H); 13C NMR (CDCl3, 125 MHz) δ 141.2, 136.7, 133.9, 130.3, 129.5, 127.5, 127.0, 126.6, 123.7, 122.0, 119.8, 119.3, 118.4, 111.0, 36.6.

(17)

3,3'-((4-chlorophenyl)methylene)bis(1H-indole) [70, 87, 88]

White solid; 86.4 mg (93 %, Rf = 0.70, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.90 (br s, 2H), δ 7.35 (d, 4H, J = 8.2 Hz), 7.27-7.16 (m, 6H), 7.01 (t, 2H, J= 8.2 Hz), 6.63 (s, 2H ), 5.85 (s, 1H); 13

C NMR (CDCl3, 125 MHz) δ 142.5, 136.6, 131.7, 130.0, 128.3, 126.8, 123.5, 122.0, 119.8, 119.3,

119.1, 111.0, 39.6.

(18)

4-(di(1H-indol-3-yl)methyl)benzonitrile [78, 89]

White solid; 83.3 mg (92 %, Rf = 0.70, 40:60 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.99 (br s, 2H), δ 7.56 (d, 2H, J = 8.2 Hz), 7.44 (d, 2H, J = 8.2 Hz), 7.37 (d, 2H, J= 8.2 Hz), 7.32 (d, 2H, J= 7.5 Hz), 7.19 (t, 2H, J= 8.2 Hz), 7.02 (t, 2H, J = 6.9 Hz), 6.65 (d, 2H, J= 2.1 Hz), 5.93 (s, 1H); 13C NMR (CDCl3, 125 MHz) δ 149.6,146.9, 144.4, 136.6, 132.2, 129.5, 126.6, 125.3, 123.6, 122.3, 119.5, 118.2,111.2, 40.3.

(19)

3,3'-((4-nitrophenyl)methylene)bis(1H-indole) [47, 75, 84]

Yellow solid; 91.6 mg (96 %, Rf = 0.80, 40:60 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 8.14(d, 2H, J= 8.2 Hz), 8.02 (br s, 2H), δ 7.50 (d, 2H, J = 8.2 Hz), 7.39 (d, 2H, J = 8.2 Hz), 7.33 (d, 2H, J= 8.2 Hz), 7.20 (t, 2H, J= 8.2 Hz), 7.02 (t, 2H, J= 8.2 Hz), 6.68 (d, 2H, J= 2.7 Hz), 5.99 (s, 1H); 13C NMR (CDCl3, 125 MHz) δ 151.9,136.7, 129.6, 126.7, 123.7, 122.4, 119.7, 119.6,118.2, 111.3, 40.3.

(20)

3,3'-(furan-2-ylmethylene)bis(1H-indole) [88, 90-92]

Page 32 of 52 ACS Paragon Plus Environment

Page 33 of 52

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

Dark Red solid; 66.3 mg (85 %, Rf = 0.80, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.94 (br s, 7.47(d, 2H, J= 7.5 Hz), 7.35 (d, 3H, J = 7.5 Hz), 7.17 (t, 3H, J = 8.2 Hz), 7.03 (t, 2H, J = 7.5 Hz), 6.87 (d, 2H, J = 2.7 Hz), 6.29 (t, 1H, J = 2.0 Hz), 6.05 (d, 1H, J = 3.4 Hz), 5.94 (s, 1H); 13C NMR (CDCl3, 125 MHz) δ 157.0, 141.2, 136.5, 126.7,123.0,121.9,119.7,119.3,117.2, 111.0, 110.1, 106.6, 34.0.

(21)

3,3'-(thiophen-2-ylmethylene)bis(1H-indole) [76, 85]

Pink solid; 66.5 mg (82 %, Rf = 0.80, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ7.91(br s, 2H), 7.46 d, 2H, J= 8.2 Hz), δ 7.35 (d, 2H, J = 8.2 Hz), 7.18-7.14 (m, 3H), 7.03 (t, 2H, J= 8.2 Hz), 6.926.89(m, 2H), 6.83 (d, 2H, J= 2.0Hz), 6.16 (s, 1H); 13C NMR (CDCl3, 125 MHz) δ 148.6, 136.5, 126.7,126.4,125.1,123.6,123.1,122, 119.7, 119.6, 119.3, 111.1, 35.3.

(22)

tri(1H-indol-3-yl)methane [76]

White solid; 71.4 mg (79 %, Rf = 0.70, 60:40 EtOAc/Hex) 1H NMR (DMSO, 500 MHz) δ 10.70 (s, 3H), 7.37 (d, 3H, J = 8.2 Hz), 7.31 (d, 3H, J = 8.2 Hz), 7.00 (t, 3H, J= 7.5 Hz), 6.92 (d, 3H, J = 1.3 Hz), 6.83 (t, 3H, J= 7.5 Hz), 6.03 (s, 1H); 13C NMR (DMSO, 125 MHz) δ 137.0, 127.2,123.7, 121.1, 119.7, 118.7, 118.4, 111.8,.

(23)

3,3'-(phenylmethylene)bis(5-methoxy-1H-indole)

Pink solid; 68 mg (93 %, Rf = 0.75, 25:75 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.84 (br s, 2H), δ 7.34 (d, 2H, J = 6.9 Hz), 7.29-7.13 (m, 6H), 7.22 (t, 1H, J= 7.5 Hz), 6.83(d, 1H, J=2.0 Hz), 6.81(d, 1H, J=2.0 Hz), 6.79(d, 2H, J=2.1 Hz), 6.66 (d, 2H, J=2.1 Hz), 5.77 (s, 1H ), 3.68 (s, 3H); 13C NMR (CDCl3, 125 MHz) δ 153.6 143.8, 131.8, 128.7, 128.2, 127.5, 126.1, 124.4, 119.3, 111.9, 111.6, 101.9, 55.8, 40.2.

(24)

3,3'-(phenylmethylene)bis(1-ethyl-5-methoxy-1H-indole)

Pink solid; 100.8 mg (92 %, Rf = 0.70, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.34 (d, 2H, J = 7.5 Hz), 7.28 (d, 3H, J = 7.5 Hz), 7.20 (t, 3H, J = 8.9 Hz), 6.84 (d, 1H, J = 2.0 Hz), 6.82 (d, 1H, J = 2.0 Hz), 6.77 (d, 2H, J = 2.1 Hz), 6.57 (s, 2H ), 5.74 (s, 1H), 4.03(q, 4H, J = 6.8 Hz), 3.67(s, 6H),

Page 33 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 34 of 52

1.35 (d, 6H, J = 6.9 Hz) ; 13C NMR (CDCl3, 125 MHz) δ 153.3,144.4, 131.7,128.7, 128.1, 127.9, 127.2, 125.9, 117.6, 111.3, 109.8, 102.1, 55.9,41.0, 40.4,15.6.

(25)

5-methoxy-3-((5-methoxy-1-methyl-1H-indol-3-yl)(4-methoxyphenyl)methyl)-1-

methyl- 1H-indole [50] Pink solid, 59.5 mg (86 %, Rf = 0.71, 25:75 EtOAc/Hex) m.p.171.5-172.5 °C; 1H NMR (CDCl3, 500 MHz) δ 7.24 (d, 2H), 7.17 (d, 2H, J = 8.9 Hz), 6.79-6.86 (m, 6H), 6.49 (s, 2H), 5.70 (s, 1H), 3.78 (s, 3H), 3.70 (s, 6H), 3.65 (s, 6H); 13C NMR (CDCl3, 125 MHz) δ 157.7, 153.3, 136.5, 132.8, 129.5,128.8, 127.7, 117.8, 113.5, 111.3, 109.7, 102.1, 55.9, 55.2, 39.2, 32.8.

(26)

1-ethyl-3-((1-ethyl-5-methoxy-1H-indol-3-yl)(4-methoxyphenyl)methyl)-5-methoxy-1H-

indole [50] Pink solid, 60 mg (83 %, Rf = 0.77, 20:80 EtOAc/Hex) m.p.125.5-126.5 °C; 1H NMR (CDCl3, 500 MHz) δ 7.25 (d, 2H, J = 8.9 Hz), 7.21 (d, 2H, J = 8.9 Hz), 6.81-6.84 (m, 4H), 6.78 (d, 2H, J = 2Hz), 6.55 (s, 2H), 5.70 (s, 1H), 4.03 (q, 4H, J = 6.8Hz), 3.79 (s, 3H), 3.68 (s, 6H), 1.35 (t, 6H, J = 6.8 Hz); 13

C NMR (CDCl3, 125 MHz) δ 157.7, 153.2, 136.6, 131.7, 129.6, 127.9, 127.2, 117.9, 113.5, 111.2,

109.8, 102.2, 55.9, 55.2, 40.9, 39.5, 15.6.

(27)

di(1H-indol-3-yl)methane [93]

White solid, 60.3 mg (98%, Rf = 0.55, 15:85 EtOAc/Hex) m. p 162-163°C; 1H NMR (CDCl3, 500 MHz): δ 7.90 (br s, 2H), δ 7.62 (d, 2H, J = 8.2 Hz), 7.36 (d, 2H, J = 7.5 Hz), 7.19 (t, 2H, J= 8.2 Hz), 7.09 (t, 2H, J= 7.5 Hz), 6.94 (d, 2H, J = 2.0 Hz), 4.24 (s, 2H) ; 13C NMR (CDCl3, 125 MHz) δ 136.4, 127.5, 122.1, 121.8, 119.2, 115.6, 111.0, 21.2.

(28)

3,3'-(ethane-1,1-diyl)bis(1H-indole) [93]

Colorless oil; 61.1 mg (94%, Rf = 0.67, 20:80 EtOAc/Hex); 1H NMR (CDCl3, 500 MHz) δ 7.84 ( br s,2H), δ 7.59 (d, 2H, J = 7.5 Hz), 7.34 (d, 2H, J = 8.2 Hz), 7.18 (t, 2H, J= 6.9 Hz), 7.06 (t, 2H, J= 8.2 Hz), 6.91 (d, 2H, J = 2.0 Hz), 4.69 (q, 1H, J= 6.8 Hz), 1.82 (d, 3H, J = 7.5 Hz) ; 13C NMR (CDCl3, 125 MHz) δ 136.6, 126.8, 121.7, 121.6, 121.2, 119.7, 118.9, 111.0, 28.1, 21.7.

Page 34 of 52 ACS Paragon Plus Environment

Page 35 of 52

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

(29)

1-methyl-3-(1-(1-methyl-1H-indol-3-yl)ethyl)-1H-indole[94]

Colorless oil; 65.6 mg (91 %, Rf = 0.62, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 7.57 (d, 2H, J = 8.2 Hz), 7.26 (d, 2H, J = 8.2 Hz), 7.18 (t, 2H, J = 6.8 Hz), 7.02 (t, 2H, J = 6.8 Hz), 6.76 (s, 2H), 4.66 (q, 1H, J = 6.8 Hz), 3.67 (s, 6H), 1.78 (d, 3H, J = 6.8 Hz); 13C NMR (CDCl3, 125 MHz) δ 137.2, 126.0, 125.9, 121.2, 120.2, 119.8, 119.7, 118.3, 109.0, 32.6, 27.9, 21.1.

(30)

1-ethyl-3-(1-(1-ethyl-1H-indol-3-yl)ethyl)-1H-indole[95]

Colorless oil; 70.4 mg (89 %, Rf = 0.67, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) 7.59 (d, 2H, J = 8.2 Hz), 7.32 (d, 2H, J = 8.2 Hz),7.19(t, 2H, J= 6.8 ) 7.04 (t, 2H, J = 6.9 Hz), 6.86 (s, 2H), 4.67 (q, 1H, J = 7.5 Hz), 4.03-3.99 (m, 4H), 1.80 (d, 3H, J = 7.6 Hz), 1.42 (t, 6H, J = 7.5 Hz); 13C NMR (CDCl3, 125 MHz) δ 136.3, 127.4, 124.3, 121.0, 120.3, 120.0, 118.2, 109.1, 40.7, 28.1, 22.1, 15.5.

(31)

3,3'-(ethane-1,1-diyl)bis(1-allyl-1H-indole)[50]

Colorless oil; 74 mg (87 %, Rf = 0.67, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) 7.59 (d, 2H, J = 8.2 Hz), 7.30 (d, 2H, J = 7.5 Hz), 7.19 (t, 2H, J= 7.5Hz), 7.05 (t, 2H, J = 7.5 Hz), 6.86 (s, 2H), 6.025.94 (m, 2H), 5.18 (d, 2H, J = 10.3 Hz), 5.08 (d, 2H, J =16.5 Hz), 4.72-4.67 (m, 5H ), 1.82 (d, 3H, J = 6.8 Hz); 13C NMR (CDCl3, 125 MHz) δ 136.7, 133.8, 133,7, 127.5, 125.0, 124.9, 121.2, 120.5, 119.9, 118.5, 116.8, 109.4, 48.6, 28.1, 22.0.

(32)

3,3'-(phenylmethylene)bis(1-benzyl-1H-indole)

White solid; 123.1 mg (98 %, Rf = 0.50, 5:95 EtOAc/Hex) m.p.137.5-138.5; 1H NMR (CDCl3, 500 MHz) δ 7.40 (d, 2H, J= 8.2 Hz), δ 7.37 (d, 2H, J = 7.5 Hz), 7.29-7.20 (m, 11H), 7.12 (t, 2H, J= 6.8 Hz), 7.03 (d, 2H, J= 7.6 Hz), 6.98 (t, 2H, J= 7.5Hz), 5.93 (s, 1H), 5.22 (s, 4H); 13C NMR (CDCl3, 125 MHz) δ 144.0, 137.8, 136.9,128.7,128.6,128.2,127.9,127.7, 127.3, 126.4, 126.1, 121.6, 120.1, 118.8, 118.7, 109.6, 49.9, 40.22. HRMS (ESI): m/z [M + Na]+ calcd for C37H30N2Na: 529.2307; found: 529.2311.

(33)

3,3'-((4-methoxyphenyl)methylene)bis(1-benzyl-1H-indole)

Page 35 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 36 of 52

White solid; 131.8 mg (99 %, Rf = 0.50, 7:93 EtOAc/Hex) m.p.136-137 °C; 1H NMR (CDCl3, 500 MHz) δ 7.40 (d, 2H, J= 7.5 Hz), δ 7.37, 7.28-7.19 (m, 10H), 7.11 (t, 2H, J= 7.5 Hz), 7.03 (d, 4H, J= 7.5 Hz), 6.79 (t, 2H, J= 6.8Hz), 6.82(d, 2H, J = 8.2 Hz), 6.64 (s, 2H), 5.88 (s, 1H), 5.21(s, 3H), 3.77 (s, 3H); 13C NMR (CDCl3, 125 MHz) δ 157.8,137.8, 137.0,136.3, 129.6, 128.6, 127.8, 127.7, 127.3, 126.4, 121.6, 120.2, 119.1, 118.8, 113.5, 109.6, 55.2, 49.9, 39.4. HRMS (ESI): m/z [M + Na]+ calcd for C38H32N2NaO: 555.2412; found: 555.2418.

(34) 3,3'-(phenylmethylene)bis(6-fluoro-1H-indole) [49] Red solid; 79.7 mg (91 %, Rf = 0.50, 20:80 EtOAc/Hex) m.p.131.5-132.5; 1H NMR (CDCl3, 500 MHz) δ 7.88(brs, 2H), 7.31-7.19( m, 7H), 7.00( dd, 2H, J= 9.6 & 2.0 Hz), 6.74(dt, 2H, J= 9.6 & 2.0 Hz), 6.59( d, 2H, J= 1.3 Hz), 5.78( s, 1H) ); 13C NMR (CDCl3, 125 MHz) δ 160.8, 158.9, 143.5, 136.5, 136.4, 133.7, 130.1, 128.5, 128.4, 128.2, 126.3, 123.7, 123.5, 120.5,120.4, 119.5, 108.1, 108.0, 40.1.

(35) 3,3'-(phenylmethylene)bis(5-chloro-1H-indole)[49] White solid; 92.93 mg (95 %, Rf = 0.50, 25:75 EtOAc/Hex) m.p.222-224; 1H NMR (CDCl3, 500 MHz) δ 7.95(brs, 2H ), 7.31-7.23( m, 9H), 7.11(d, 2H, J= 8.2Hz),6.66(s, 2H), 5.75(s, 1H) ); 13C NMR (CDCl3, 125 MHz) δ 143.0, 135.0, 128.5, 128.4, 128.0, 126.5, 125.0, 124.9, 122.4, 119.2, 119.1, 112.1, 39.9.

(36) 3,3'-(phenylmethylene)bis(5-bromo-1H-indole) [49] Red solid; 111.6 mg (93 %, Rf = 0.50, 20:80 EtOAc/Hex) m.p.233-234; 1H NMR (CDCl3, 500 MHz) δ 11.08(s, 2H), 7.42(d, 2H, J= 1.4 Hz), 7.34-7.31(m, 4H), 7.29(t, 2H, J= 7.5Hz), 7.19(t, 1H, J= 7.5Hz), 7.14(dd, 2H, J= 8.2 & 2.1Hz), 6.88(d, 2H, J= 2.0Hz), 5.85(s, 1H)); 13C NMR (CDCl3, 125 MHz) δ 144.3, 135.2, 128.3, 128.2, 126.0, 125.2, 123.4, 121.2, 117.6, 113.5, 110.9.

(37) 3,3'-(phenylmethylene)bis(5-nitro-1H-indole) [93] Yellow solid; 101 mg (98 %, Rf = 0.50, 50:50 EtOAc/Hex) m.p.277-278; 1H NMR (CDCl3, 500 MHz) δ 11.68( s, 2H)8.32(d,2H, J= 2.0Hz), 7.97(dd, 2H, J= 8.9& 2.7Hz), 7.53(d, 2H, J= 8.9Hz),

Page 36 of 52 ACS Paragon Plus Environment

Page 37 of 52

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

7.39( d, 2H, J= 6.8Hz), 7.32(t, 2H, J= 7.5Hz), 7.22( t, 2H, J=7.5Hz), 7.14( d, 2H, J= 7.5 Hz), 6.20(s, 1H) ); 13C NMR (CDCl3, 125 MHz) δ 143.8, 140.2, 139.8, 128.5, 128.2, 127.6, 126.4, 125.8, 120.5, 116.6, 116.2, 112.1, 38.4.

(38) 3,3'-(phenylmethylene)bis(1H-indole-5-carbonitrile)[96] 0range solid; 93.2 mg (99 %, Rf = 0.50, 35:65 EtOAc/Hex) m.p.142-143; 1H NMR (CDCl3, 500 MHz) δ 11.5( s, 2H), 7.53(d, 2H, J= 8.2Hz), 7.39(t, 4H, J= 8.2Hz), 7.30(t, 2H, J= 6.8Hz), 7.20(t, 1H, J= 6.8Hz), 7.13(s, 2H), 6.03(s, 1H) ); 13C NMR (CDCl3, 125 MHz) δ 144.0, 138.2, 128.3, 128.2, 126.7, 124.5, 123.7, 120.7, 118.8, 112.9, 100.3, 38.5.

(39) 3-((1H-indol-3-yl)(phenyl)methyl)-5-methoxy-1H-indole Pink solid; 55.5 mg (63 %, Rf = 0.45, 20:80 EtOAc/Hex) m.p.162-163; 1H NMR (CDCl3, 500 MHz) δ 7.92(s, 1H), 7.81( s,1H), 7.38-7.33( m, 4H), 7.29-7.23 (m, 3H), 7.20(t, 1H, J = 7.5Hz), 7.16(t, 1H, J =7.5Hz), 7.00(t, 1H, J = 7.5Hz), 6.83-6.81( m, 2H), 6.65(dd, 2H, J = 16.5 & 1.4 Hz), 5.82( s, 1H), 3.69(s, 3H) ); 13C NMR (CDCl3, 125 MHz) δ 153.7, 143.9, 136.6, 131.7, 128.7, 128.2, 127.4, 127.0, 126.1, 124.4, 123.6, 121.8, 120.0, 119.5, 119.4, 119.2, 111.9, 111.6, 111.0, 101.8, 55.9, 40.2. HRMS (ESI): m/z [M]+ calcd for C24H20N2O: 352.1576; found: 352.1574

(40) 3-((1H-indol-3-yl)(phenyl)methyl)-5-nitro-1H-indole Yellow solid; 62.5mg (68 %, Rf = 0.40, 30:70 EtOAc/Hex) m.p.201-202; 1H NMR (CDCl3, 500 MHz) δ 11.62 (s, 1H), 10.8( s, 1H), 8.24(d,1H, J =2.1Hz), 7.95(dd, 1H, J = 8.9, 2.0 Hz), 7.52(d,1H, J = 8.9Hz), 7.38-7.34(m,3H), 7.30-7.28( m, 3H), 7.19(t, 1H, J = 6.8Hz), 7.09( S, 1H), 7.04( t, 1H, J = 7.5Hz), 6.88-6.84(m, 2H), 5.98(S,1H) ); 13C NMR (CDCl3, 125 MHz) δ 144.3, 140.0, 139.8, 136.6, 128.3, 127.5, 126.4, 126.1, 125.9, 123.6, 121.1, 121.0, 119.1, 118.3, 117.5, 116.5, 116.2, 112.0, 111.5. HRMS (ESI): m/z [M + Na]+ calcd for C23H17N3O2Na: 390.1218; found: 390.1221

(41) 3-((1H-indol-3-yl)(phenyl)methyl)-5-chloro-1H-indole Pink solid,; 64.1 mg (72 %, Rf = 0.50, 20:80 EtOAc/Hex) 1H NMR (CDCl3, 500 MHz) δ 11.03( s,1H), 10.81( s, 1H), 7.36-7.32(m, 4H), 7.28- 7.24(m, 3H), 7.21 (d,1H, J = 2.0Hz), 7.17( t, 1H, J=

Page 37 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 38 of 52

6.9Hz),7.04-7.01 (m, 2H), 6.88 ( m, 2H), 6.85(t, 1H, J= 8.2Hz), 6.80(d, 1H, J= 2.0Hz), 5.80 (s, 1H) ); 13

C NMR (CDCl3, 125 MHz) δ 144.6, 136.5, 135.2, 128.3, 128.1, 126.5, 125.9, 125.2, 123.5, 123.3,

121.2, 120.9, 119.1, 118.2, 117.8, 117.7, 113.5, 111.5, 110.8. HRMS (ESI): m/z [M]+ calcd for C23H17N2Cl: 356.1080; found: 356.1084.

(42) 3-((1H-indol-3-yl)(phenyl)methyl)-5-bromo-1H-indole [97] Red solid; 64.2 mg (64 %, Rf = 0.50, 20:80 EtOAc/Hex) m.p.138-140 ; 1H NMR (CDCl3, 500 MHz) δ 11.06(s,1H), 10.8( s, 1H), 7.40(s, 1H), 7.35- 7.26(m, 7H), 7.19-7.13(m,2H), 7.03( t, 1H, J= 7.5Hz), 6.88-6.81( m, 3H), 5.83(s, 1H) ); 13C NMR (CDCl3, 125 MHz) δ 144.6, 136.5, 135.2, 128.3, 128.1, 126.5, 125.9, 125.2, 123.5, 123.3, 121.2, 120.9, 119.1, 118.2, 117.8, 117.7, 113.5, 111.5, 110.8. HRMS (ESI): m/z [M]+ calcd for C23H17N2Br: 400.0575; found: 400.0579.

(43) 3,3'-(cyclohexane-1,1-diyl)bis(1H-indole)[78, 81, 89] White solid; 56.6 mg (72 %, Rf = 0.40, 20:80 EtOAc/Hex) m.p.117-118; 1H NMR (CDCl3, 500 MHz) δ 7.91( brs, 2H), 7.55( d, 2H, J = 8.2Hz), 7.28 (d, 2H, J = 8.2Hz), 7.09 (d, 2H, J =2.0Hz), 7.05( t, 2H), 6.88 (t, 2H, J = 8.2Hz), 2.54(t, 4H, J =5.5Hz), 1.64( t, 4H, J = 4.8Hz), 1.58 (t, 2H, J= 4.8 Hz) ); 13

C NMR (CDCl3, 125 MHz) δ 137.0, 126.3, 123.7, 121.9, 121.4, 121.9, 118.5, 111.0, 39.5, 36.8,

26.8, 22.9.

(44) 3,3'-(cyclopentane-1,1-diyl)bis(1H-indole) [78, 81, 89] White solid; 47.3 mg (63 %, Rf = 0.50,20:80 EtOAc/Hex) m.p.71-72; 1H NMR (CDCl3, 500 MHz) δ 7.86( brs, 2H), 7.50(d, 2H, j= 8.2Hz), 7.28( d, 2H, J= 8.2Hz), 7.09- 7.05( m, 4H), 6.90( t, 2H, J = 7.5Hz), 2.51( t, 4H, J = 6.9Hz), 1.82( t, 4H, J= 6.9Hz) ); 13C NMR (CDCl3, 125 MHz) δ 131.1, 126.5, 123.4, 121.3, 121.1, 120.9, 118.5, 110.1, 38.6, 29.6, 23.9.

(45) 3,3'-((9H-fluoren-9-ylidene)methylene)bis(1H-indole) White solid; 45.6 mg (46 %, Rf = 0.50, 20:85 EtOAc/Hex) m.p.237-239; 1H NMR (CDCl3, 500 MHz) δ 7.88 (brs, 2H), 7.82( d, 2H, J = 7.5 Hz), 7.63 (d, 2H, J = 8.2 Hz), 7.34 (t, 2H, J = 7.5 Hz), 7.30 (d, 2H, J = 7.5 Hz), 7.17 (t, 2H, J =7.5Hz), 7.09-7-03( m, 4H), 6.86-6.82( m, 4H) ; 13C NMR

Page 38 of 52 ACS Paragon Plus Environment

Page 39 of 52

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

(CDCl3, 125 MHz) δ 144.8.0. 136.8, 135.3, 128.5, 128.4, 128.0, 126.7, 126.2, 125.6, 123.7, 123.0,121.3, 121.2, 119.3, 118.5, 118.4, 118.2, 118.0, 113.4,111.8,119.0, 111.0. HRMS (ESI): m/z [M]+ calcd for C29H20N2: 390.1626; found: 390.1621

(46) 3,3'-(propane-2,2-diyl)bis(1H-indole) [82, 83] White solid; 37.1 mg (54 %, Rf = 0.40, 15:85 EtOAc/Hex); 1H NMR (CDCl3, 500 MHz) δ 7.82 (brs, 2H), 7.44 (d, 2H, J =8.2 Hz), 7.30( d, 2H, J = 8.2Hz), 7.09 (t, 2H, J= 8.2Hz), 7.02( d, 2H, J= 1.4Hz), 6.90( d, 2H, J= 8.2Hz), 1.93 (s, 6H ); 13C NMR (CDCl3, 125 MHz) δ 137.0, 126,2, 125.4, 121.3, 121.2, 120.4, 118.6, 111.0, 34.8, 29.9.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI:10.1021/acssuschemeng.xxxxxx. Crystallographic Data for 3a, Green Chemistry matrix calculations, M-H curves of MoS2 and MoS2-RGO nanocomposite and 1H and 13C NMR data of the synthesized compounds.

AUTHOR INFORMATION Corresponding Author *Email: [email protected] ORCID Ashish Bahuguna: 0000-0001-5542-1708 Kumbam Lingeshwar Reddy: 0000-0001-6885-1328 Suneel Kumar: 0000-0002-5259-1792 Vipul Sharma: 0000-0002-4460-4610 Venkata Krishnan: 0000-0002-4453-0914 Notes The authors declare no competing financial interest.

Page 39 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 40 of 52

ACKNOWLEDGEMENTS Advanced Materials Research Center (AMRC), Indian Institute of Technology Mandi is gratefully acknowledged for providing all the necessary laboratory and instrumentation facilities. VK acknowledges the financial support from Department of Science and Technology, India under Young Scientist Scheme (YSS/2014/000456). PCR acknowledges SERB, India for the financial support (CS-185/2011). We acknowledge Mr. D. Rambabu for his help in solving crystal structure. AB and SK acknowledge CSIR and UGC, India for senior research fellowships, respectively. VS and KLR thank MHRD, India for senior research fellowship. We would like to thank the reviewers for their very helpful suggestions and comments concerning the work.

REFERENCES 1.

National Environmental Policy Act of 1969. In U.S.C., 1994; Vol. 42, pp 102-105.

2.

Astruc, D.; Lu, F.; Aranzaes, J. R., Nanoparticles as recyclable catalysts: the frontier between homogeneous and heterogeneous catalysis. Angewandte Chemie International Edition 2005, 44, (48), 7852-7872. DOI: 10.1002/anie.200500766

3.

Corma, A.; Garcia, H., Lewis acids: from conventional homogeneous to green homogeneous and heterogeneous catalysis. Chemical reviews 2003, 103, (11), 43074366. DOI: 10.1021/cr030680z.

4.

Machado, B. F.; Serp, P., Graphene-based materials for catalysis. Catalysis Science & Technology 2012, 2, (1), 54-75. DOI: 10.1039/C1CY00361E.

5.

Kalidindi, S. B.; Jagirdar, B. R., Nanocatalysis and prospects of green chemistry. ChemSusChem 2012, 5, (1), 65-75. DOI: 10.1002/cssc.201100377

6.

Anastas, P. T.; Kirchhoff, M. M., Origins, current status, and future challenges of green chemistry.

Accounts

of

chemical

research

2002,

35,

(9),

686-694.

DOI: 10.1021/ar010065m. 7.

Ghosh, P. P.; Das, A. R., Nanocrystalline and Reusable ZnO Catalyst for the Assembly of Densely Functionalized 4 H-Chromenes in Aqueous Medium via One-Pot Three Component Reactions: A Greener “NOSE” Approach. The Journal of Organic Chemistry 2013, 78, (12), 6170-6181. DOI: 10.1021/jo400763z.

Page 40 of 52 ACS Paragon Plus Environment

Page 41 of 52

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

8.

Kumar, A.; Saxena, D.; Gupta, M. K., Nanoparticle catalyzed reaction (NPCR): ZnO-NP catalyzed Ugi-reaction in aqueous medium. Green Chemistry 2013, 15, (10), 2699-2703. DOI:10.1039/C3GC41101J.

9.

Chen, Z.; Liu, S.; Yang, M.-Q.; Xu, Y.-J., Synthesis of uniform CdS nanospheres/graphene hybrid nanocomposites and their application as visible light photocatalyst for selective reduction of nitro organics in water. ACS applied materials & interfaces 2013, 5, (10), 4309-4319. DOI: 10.1021/am4010286.

10.

Liu, S.; Chen, Z.; Zhang, N.; Tang, Z.-R.; Xu, Y.-J., An Efficient Self-Assembly of CdS Nanowires–Reduced Graphene Oxide Nanocomposites for Selective Reduction of Nitro Organics under Visible Light Irradiation. The Journal of Physical Chemistry C 2013, 117, (16), 8251-8261. DOI: 10.1021/jp400550t.

11.

Guo, X.; Hao, C.; Jin, G.; Zhu, H. Y.; Guo, X. Y., Copper nanoparticles on graphene support: an efficient photocatalyst for coupling of nitroaromatics in visible light. Angewandte

Chemie

International

Edition

2014,

53,

(7),

1973-1977.

DOI:

10.1002/anie.201309482 12.

Rajesh, U. C.; Wang, J.; Prescott, S.; Tsuzuki, T.; Rawat, D. S., RGO/ZnO nanocomposite: An efficient, sustainable, heterogeneous, amphiphilic catalyst for synthesis of 3substituted indoles in water. ACS Sustainable Chemistry & Engineering 2014, 3, (1), 9-18. DOI: 10.1021/sc500594w.

13.

Rajesh, U. C.; Purohit, G.; Rawat, D. S., One-pot synthesis of aminoindolizines and chalcones using CuI/CSP nanocomposites with anomalous selectivity under green conditions. ACS Sustainable Chemistry & Engineering 2015, 3, (10), 2397-2404. DOI: 10.1021/acssuschemeng.5b00701.

14.

Gawande, M. B.; Rathi, A. K.; Nogueira, I. D.; Varma, R. S.; Branco, P. S., Magnetitesupported sulfonic acid: a retrievable nanocatalyst for the Ritter reaction and multicomponent reactions. Green Chemistry 2013, 15, (7), 1895-1899. DOI: 10.1039/C3GC40457A.

Page 41 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

15.

Page 42 of 52

Baig, R. N.; Nadagouda, M. N.; Varma, R. S., Ruthenium on chitosan: a recyclable heterogeneous catalyst for aqueous hydration of nitriles to amides. Green Chemistry 2014, 16, (4), 2122-2127. DOI: 10.1039/C3GC42004C.

16.

Pais, I.; Jones Jr, J. B., The handbook of trace elements. CRC Press: 1997.

17.

Smith, K. S.; Huyck, H. L., An overview of the abundance, relative mobility, bioavailability, and human toxicity of metals. The environmental geochemistry of mineral deposits 1999, 6, 29-70.

18.

Underwood, E., Trace elements in human and animal nutrition 4e. Elsevier: 2012.

19.

Hinnemann, B.; Moses, P. G.; Bonde, J.; Jørgensen, K. P.; Nielsen, J. H.; Horch, S.; Chorkendorff, I.; Nørskov, J. K., Biomimetic hydrogen evolution: MoS2 nanoparticles as catalyst for hydrogen evolution. Journal of the American Chemical Society 2005, 127, (15), 5308-5309. DOI: 10.1021/ja0504690.

20.

Trakarnpruk, W.; Seentrakoon, B., Hydrodesulfurization activity of MoS2 and bimetallic catalysts prepared by in situ decomposition of thiosalt. Industrial & Engineering Chemistry Research 2007, 46, (7), 1874-1882. DOI: 10.1021/ie061176y.

21.

Dovell, F. S.; Greenfield, H., Base-metal sulfides as reductive alkylation catalysts. The Journal of Organic Chemistry 1964, 29, (5), 1265-1267. DOI: 10.1021/jo01028a511.

22.

Nishimura, S., Handbook of heterogeneous catalytic hydrogenation for organic synthesis. Wiley New York etc: 2001.

23.

Min, S.; Lu, G., Sites for high efficient photocatalytic hydrogen evolution on a limitedlayered MoS2 cocatalyst confined on graphene sheets―the role of graphene. The Journal of Physical Chemistry C 2012, 116, (48), 25415-25424. DOI: 10.1021/jp3093786.

24.

Liu, C.-J.; Tai, S.-Y.; Chou, S.-W.; Yu, Y.-C.; Chang, K.-D.; Wang, S.; Chien, F. S.-S.; Lin, J.-Y.; Lin, T.-W., Facile synthesis of MoS 2/graphene nanocomposite with high catalytic activity toward triiodide reduction in dye-sensitized solar cells. Journal of Materials Chemistry 2012, 22, (39), 21057-21064. DOI: 10.1039/C2JM33679K.

25.

Wang, T.; Zhu, H.; Zhuo, J.; Zhu, Z.; Papakonstantinou, P.; Lubarsky, G.; Lin, J.; Li, M., Biosensor based on ultrasmall MoS2 nanoparticles for electrochemical detection of

Page 42 of 52 ACS Paragon Plus Environment

Page 43 of 52

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

H2O2 released by cells at the nanomolar level. Analytical chemistry 2013, 85, (21), 10289-10295. DOI: 10.1021/ac402114c. 26.

Chou, S. S.; De, M.; Kim, J.; Byun, S.; Dykstra, C.; Yu, J.; Huang, J.; Dravid, V. P., Ligand conjugation of chemically exfoliated MoS2. Journal of the American Chemical Society 2013, 135, (12), 4584-4587. DOI: 10.1021/ja310929s.

27.

da Silveira Firmiano, E. G.; Rabelo, A. C.; Dalmaschio, C. J.; Pinheiro, A. N.; Pereira, E. C.; Schreiner, W. H.; Leite, E. R., Supercapacitor electrodes obtained by directly bonding 2D MoS2 on reduced graphene oxide. Advanced Energy Materials 2014, 4, (6). DOI: 10.1002/aenm.201301380

28.

Peng, W.-c.; Chen, Y.; Li, X.-y., MoS 2/reduced graphene oxide hybrid with CdS nanoparticles as a visible light-driven photocatalyst for the reduction of 4-nitrophenol. Journal

of

hazardous

309,

2016,

materials

173-179.

DOI:

10.1016/j.jhazmat.2016.02.021. 29.

Zhai, C.; Zhu, M.; Bin, D.; Ren, F.; Wang, C.; Yang, P.; Du, Y., Two dimensional MoS 2/graphene composites as promising supports for Pt electrocatalysts towards methanol oxidation.

Journal

of

Power

Sources

2015,

275,

483-488.

DOI:

10.1016/j.jpowsour.2014.11.030. 30.

Lee, Y. H.; Zhang, X. Q.; Zhang, W.; Chang, M. T.; Lin, C. T.; Chang, K. D.; Yu, Y. C.; Wang, J. T. W.; Chang, C. S.; Li, L. J., Synthesis of Large-Area MoS2 Atomic Layers with Chemical Vapor

Deposition.

Advanced

Materials

2012,

24,

(17),

2320-2325.

DOI:

10.1002/adma.201104798 31.

Liang, Y. T.; Vijayan, B. K.; Lyandres, O.; Gray, K. A.; Hersam, M. C., Effect of dimensionality on the photocatalytic behavior of carbon–titania nanosheet composites: charge transfer at nanomaterial interfaces. The journal of physical chemistry letters 2012, 3, (13), 1760-1765. DOI: 10.1021/jz300491s.

32.

McClelland, D. J.; Motagamwala, A. H.; Li, Y.; Rover, M. R.; Wittrig, A. M.; Wu, C.; Buchanan, J. S.; Brown, R. C.; Ralph, J.; Dumesic, J. A., Functionality and molecular weight distribution of red oak lignin before and after pyrolysis and hydrogenation. Green Chemistry 2017, 19, (5), 1378-1389. DOI: 10.1039/C6GC03515A.

Page 43 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

33.

Page 44 of 52

Karanjkar, P. U.; Burt, S. P.; Chen, X.; Barnett, K. J.; Ball, M. R.; Kumbhalkar, M. D.; Wang, X.; Miller, J. B.; Hermans, I.; Dumesic, J. A., Effect of carbon supports on RhRe bifunctional catalysts for selective hydrogenolysis of tetrahydropyran-2-methanol. Catalysis Science & Technology 2016, 6, (21), 7841-7851. DOI: 10.1039/C6CY01763K.

34.

Bhan, A.; Gounder, R.; Macht, J.; Iglesia, E., Entropy considerations in monomolecular cracking of alkanes on acidic zeolites. Journal of Catalysis 2008, 253, (1), 221-224. DOI: 10.1016/j.jcat.2007.11.003.

35.

Romano, P. N.; de Almeida, J. M.; Carvalho, Y.; Priecel, P.; Falabella Sousa-Aguiar, E.; Lopez-Sanchez, J. A., Back Cover: Microwave-Assisted Selective Hydrogenation of Furfural to Furfuryl Alcohol Employing a Green and Noble Metal-Free Copper Catalyst (ChemSusChem

24/2016).

ChemSusChem

2016,

9,

(24),

3528-3528.

DOI:

10.1002/cssc.201601672 36.

Maneffa, A.; Priecel, P.; Lopez-Sanchez, J. A., Biomass-Derived Renewable Aromatics: Selective Routes and Outlook for p-Xylene Commercialisation. ChemSusChem 2016, 9, (19), 2736-2748. DOI: 10.1002/cssc.201600605

37.

Sukumaran, R. K.; Singhania, R. R.; Mathew, G. M.; Pandey, A., Cellulase production using biomass feed stock and its application in lignocellulose saccharification for bioethanol

production.

Renewable

Energy

2009,

34,

(2),

421-424.

DOI:

10.1016/j.renene.2008.05.008. 38.

John, R. P.; Anisha, G.; Nampoothiri, K. M.; Pandey, A., Micro and macroalgal biomass: a renewable source for bioethanol. Bioresource technology 2011, 102, (1), 186-193. DOI: 10.1016/j.biortech.2010.06.139.

39.

Xu, W.; Miller, S. J.; Agrawal, P. K.; Jones, C. W., Depolymerization and hydrodeoxygenation of switchgrass lignin with formic acid. ChemSusChem 2012, 5, (4), 667-675. DOI: 10.1002/cssc.201100695

40.

Barrett, J. A.; Gao, Y.; Bernt, C. M.; Chui, M.; Tran, A. T.; Foston, M. B.; Ford, P. C., Enhancing Aromatic Production from Reductive Lignin Disassembly: in Situ OMethylation of Phenolic Intermediates. ACS Sustainable Chemistry & Engineering 2016, 4, (12), 6877-6886. DOI: 10.1021/acssuschemeng.6b01827.

Page 44 of 52 ACS Paragon Plus Environment

Page 45 of 52

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

41.

Bernt, C. M.; Bottari, G.; Barrett, J. A.; Scott, S. L.; Barta, K.; Ford, P. C., Mapping reactivities of aromatic models with a lignin disassembly catalyst. Steps toward controlling product selectivity. Catalysis Science & Technology 2016, 6, (9), 2984-2994. DOI: 10.1039/C5CY01555C.

42.

Barta, K.; Ford, P. C., Catalytic conversion of nonfood woody biomass solids to organic liquids.

Accounts

of

chemical

2014,

research

47,

(5),

1503-1512.

DOI: 10.1021/ar4002894. 43.

Tan, S.; Gil, L. B.; Subramanian, N.; Sholl, D. S.; Nair, S.; Jones, C. W.; Moore, J. S.; Liu, Y.; Dixit, R. S.; Pendergast, J. G., Catalytic propane dehydrogenation over In 2 O 3–Ga 2 O 3 mixed

oxides.

Applied

Catalysis

A:

General

2015,

498,

167-175.

DOI:

10.1016/j.apcata.2015.03.020.

44.

Hong, D.-Y.; Miller, S. J.; Agrawal, P. K.; Jones, C. W., Hydrodeoxygenation and coupling of aqueous phenolics over bifunctional zeolite-supported metal catalysts. Chemical Communications 2010, 46, (7), 1038-1040. DOI: 10.1039/B918209H.

45.

Kumar, S.; Sharma, V.; Bhattacharyya, K.; Krishnan, V., Synergetic effect of MoS 2–RGO doping to enhance the photocatalytic performance of ZnO nanoparticles. New Journal of Chemistry 2016, 40, 5185--5197. DOI: 10.1039/C5NJ03595C.

46.

Qin, P.; Fang, G.; Ke, W.; Cheng, F.; Zheng, Q.; Wan, J.; Lei, H.; Zhao, X., In situ growth of double-layer MoO 3/MoS 2 film from MoS 2 for hole-transport layers in organic solar cell.

Journal

of

Materials

Chemistry

A

2014,

2,

(8),

2742-2756.

DOI:

10.1039/C3TA13579A. 47.

Illathvalappil, R.; Unni, S. M.; Kurungot, S., Layer-separated MoS 2 bearing reduced graphene oxide formed by an in situ intercalation-cum-anchoring route mediated by Co (OH) 2 as a Pt-free electrocatalyst for oxygen reduction. Nanoscale 2015, 7, (40), 1672916736. DOI: 10.1039/C5NR04415D.

48.

Carrier, M.; Loppinet-Serani, A.; Denux, D.; Lasnier, J.-M.; Ham-Pichavant, F.; Cansell, F.; Aymonier, C., Thermogravimetric analysis as a new method to determine the lignocellulosic composition of biomass. Biomass and Bioenergy 2011, 35, (1), 298-307. DOI: 10.1016/j.biombioe.2010.08.067.

Page 45 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

49.

Page 46 of 52

Wang, P.; Wang, J.; Au, C. T.; Qiu, R.; Xu, X.; Yin, S. F., Air-stable Organobismuth (V) Bisperfluorooctanesulfonate as an Efficient Catalyst for the Synthesis of N-Containing Compounds. Advanced Synthesis & Catalysis 2016, 358, (8), 1302-1308.

50.

Bahuguna, A.; Sharma, R.; Sagara, P.; Ravikumar, P., Exploration of Aberrant Behaviour of Grignard Reagents with Indole-3-carboxaldehyde: Application to the Synthesis of Turbomycin B and Vibrindole A Derivatives. Synlett 2017, 28, (01), 117-121. DOI: 10.1055/s-0036-1588885.

51.

Naik, S. N.; Goud, V. V.; Rout, P. K.; Dalai, A. K., Production of first and second generation biofuels: a comprehensive review. Renewable and Sustainable Energy Reviews 2010, 14, (2), 578-597. DOI: 10.1016/j.rser.2009.10.003.

52.

Liang, C.; Li, Z.; Dai, S., Mesoporous carbon materials: synthesis and modification. Angewandte Chemie International Edition 2008, 47, (20), 3696-3717. DOI: 10.1002/anie.200702046

53.

Alonso, D. M.; Bond, J. Q.; Dumesic, J. A., Catalytic conversion of biomass to biofuels. Green Chemistry 2010, 12, (9), 1493-1513. DOI: 10.1039/C004654J.

54.

Xu, C.; Arancon, R. A. D.; Labidi, J.; Luque, R., Lignin depolymerisation strategies: towards valuable chemicals and fuels. Chemical Society Reviews 2014, 43, (22), 74857500. DOI: 10.1039/C4CS00235K.

55.

Corma, A.; Iborra, S.; Velty, A., Chemical routes for the transformation of biomass into chemicals. Chemical reviews 2007, 107, (6), 2411-2502. DOI: 10.1021/cr050989d.

56.

Xia, Q.-H.; Ge, H.-Q.; Ye, C.-P.; Liu, Z.-M.; Su, K.-X., Advances in homogeneous and heterogeneous catalytic asymmetric epoxidation. Chemical reviews 2005, 105, (5), 16031662. DOI: 10.1021/cr0406458.

57.

Bartholomew, C. H.; Farrauto, R. J., Fundamentals of industrial catalytic processes. John Wiley & Sons: 2011.

58.

Wang, P.; Huang, B.; Qin, X.; Zhang, X.; Dai, Y.; Wei, J.; Whangbo, M. H., Ag@ AgCl: a highly efficient and stable photocatalyst active under visible light. Angewandte Chemie International Edition 2008, 47, (41), 7931-7933. DOI: 10.1002/anie.200802483

Page 46 of 52 ACS Paragon Plus Environment

Page 47 of 52

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

59.

Zhan, Z.; Zheng, L.; Pan, Y.; Sun, G.; Li, L., Self-powered, visible-light photodetector based on thermally reduced graphene oxide–ZnO (rGO–ZnO) hybrid nanostructure. Journal of Materials Chemistry 2012, 22, (6), 2589-2595. DOI: 10.1039/C1JM13920G.

60.

Liu, S.; Yu, B.; Zhang, H.; Fei, T.; Zhang, T., Enhancing NO 2 gas sensing performances at room temperature based on reduced graphene oxide-ZnO nanoparticles hybrids. Sensors and Actuators B: Chemical 2014, 202, 272-278. DOI: 10.1016/j.snb.2014.05.086.

61.

Kumar, N.; Srivastava, A. K.; Patel, H. S.; Gupta, B. K.; Varma, G. D., Facile Synthesis of ZnO–Reduced Graphene Oxide Nanocomposites for NO2 Gas Sensing Applications. European

Journal

of

Inorganic

Chemistry

2015,

2015,

(11),

1912-1923.

DOI:10.1002/ejic.201403172 62.

Liu, X.; Pan, L.; Zhao, Q.; Lv, T.; Zhu, G.; Chen, T.; Lu, T.; Sun, Z.; Sun, C., UV-assisted photocatalytic synthesis of ZnO–reduced graphene oxide composites with enhanced photocatalytic activity in reduction of Cr (VI). Chemical Engineering Journal 2012, 183, 238-243. DOI: 10.1016/j.cej.2011.12.068.

63.

Haldorai, Y.; Voit, W.; Shim, J.-J., Nano ZnO@ reduced graphene oxide composite for high performance supercapacitor: Green synthesis in supercritical fluid. Electrochimica Acta 2014, 120, 65-72. DOI: 10.1016/j.electacta.2013.12.063.

64.

Huang, X.; Zeng, Z.; Fan, Z.; Liu, J.; Zhang, H., Graphene-based electrodes. Advanced Materials 2012, 24, (45), 5979-6004. DOI: 10.1002/adma.201201587

65.

George, S. M., Introduction: heterogeneous catalysis. Chemical Reviews 1995, 95, (3), 475-476. DOI: 10.1021/cr00035a001.

66.

Rouquerol, J.; Rouquerol, F.; Llewellyn, P.; Maurin, G.; Sing, K. S., Adsorption by powders and porous solids: principles, methodology and applications. Academic press: 2013.

67.

Lilja, J.; Murzin, D. Y.; Salmi, T.; Aumo, J.; Mäki-Arvela, P.; Sundell, M., Esterification of different acids over heterogeneous and homogeneous catalysts and correlation with the Taft equation. Journal of molecular catalysis A: Chemical 2002, 182, 555-563. DOI: 10.1016/S1381-1169(01)00495-2.

68.

Sobhani, S.; Safaei, E.; Hasaninejad, A.-R.; Rezazadeh, S., An eco-friendly procedure for the efficient synthesis of bis (indolyl) methanes in aqueous media. Journal of

Page 47 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Organometallic

2009,

Chemistry

694,

(18),

Page 48 of 52

3027-3031.

DOI:

10.1016/j.jorganchem.2009.05.004. 69.

Shi, X. L.; Lin, H.; Li, P.; Zhang, W., Friedel–Crafts Alkylation of Indoles Exclusively in Water Catalyzed by Ionic Liquid Supported on a Polyacrylonitrile Fiber: A Simple “Release and Catch” Catalyst. ChemCatChem 2014, 6, (10), 2947-2953. DOI: 10.1002/cctc.201402396

70.

Huo, C.; Sun, C.; Wang, C.; Jia, X.; Chang, W., Triphenylphosphine-m-sulfonate/carbon tetrabromide as an efficient and easily recoverable catalyst system for Friedel–Crafts alkylation of indoles with carbonyl compounds or acetals. ACS Sustainable Chemistry & Engineering 2013, 1, (5), 549-553. DOI: 10.1021/sc400033t.

71.

Weitkamp, J.; Hunger, M.; Rymsa, U., Base catalysis on microporous and mesoporous materials: recent progress and perspectives. Microporous and Mesoporous Materials 2001, 48, (1), 255-270. DOI: 10.1016/S1387-1811(01)00366-3.

72.

Auer, E.; Freund, A.; Pietsch, J.; Tacke, T., Carbons as supports for industrial precious metal catalysts. Applied Catalysis A: General 1998, 173, (2), 259-271. DOI: 10.1016/S0926-860X(98)00184-7.

73.

Matter, P. H.; Wang, E.; Ozkan, U. S., Preparation of nanostructured nitrogen-containing carbon catalysts for the oxygen reduction reaction from SiO 2-and MgO-supported metal

particles.

Journal

of

Catalysis

2006,

243,

(2),

395-403.

DOI:

10.1016/j.jcat.2006.07.029. 74.

Li, D.; Mueller, M. B.; Gilje, S.; Kaner, R. B.; Wallace, G. G., Processable aqueous dispersions of graphene nanosheets. Nature nanotechnology 2008, 3, (2), 101-105. DOI: 10.1038/nnano.2007.451.

75.

Chang, K.; Mei, Z.; Wang, T.; Kang, Q.; Ouyang, S.; Ye, J., MoS2/graphene cocatalyst for efficient photocatalytic H2 evolution under visible light irradiation. ACS nano 2014, 8, (7), 7078-7087. DOI: 10.1021/nn5019945.

76.

Xiang, J.; Wang, J.; Wang, M.; Meng, X.; Wu, A., One-pot total synthesis of streptindole, arsindoline B and their congeners through tandem decarboxylative deaminative dual-

Page 48 of 52 ACS Paragon Plus Environment

Page 49 of 52

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 reaction of amino acids with indoles. Organic & biomolecular chemistry 2015, 13, (14), 4240-4247. DOI: 10.1039/C5OB00025D. 77.

Nagarajan, R.; Perumal, P. T., InCl 3 and In (OTf) 3 catalyzed reactions: synthesis of 3acetyl indoles, bis-indolylmethane and indolylquinoline derivatives. Tetrahedron 2002, 58, (6), 1229-1232. DOI: 10.1016/S0040-4020(01)01227-3.

78.

Ganguly, N. C.; Mondal, P.; Barik, S. K., Iodine in aqueous micellar environment: a mild

effective ecofriendly catalytic system for expedient synthesis of bis (indolyl) methanes and 3substituted indolyl ketones. Green Chemistry Letters and Reviews 2012, 5, (1), 73-81. DOI: 10.1080/17518253.2011.581700.

79.

Nair, V.; Abhilash, K.; Vidya, N., Practical synthesis of triaryl-and triheteroarylmethanes by reaction of aldehydes and activated arenes promoted by gold (III) chloride. Organic letters 2005, 7, (26), 5857-5859. DOI: 10.1021/ol052423h.

80.

Gopalaiah, K.; Chandrudu, S. N.; Devi, A., Iron-Catalyzed Oxidative Coupling of Benzylamines and Indoles: Novel Approach for Synthesis of Bis (indolyl) methanes. Synthesis 2015, 47, (12), 1766-1774. DOI: 10.1055/s-0034-1380012.

81.

Wang, L.; Han, J.; Tian, H.; Sheng, J.; Fan, Z.; Tang, X., Rare earth perfluorooctanoate [RE (PFO) 3]-catalyzed condensations of indole with carbonyl compounds. Synlett 2005, 2005, (02), 337-339. DOI: 10.1055/s-2004-837210.

82.

Yadav, J.; Reddy, B. S.; Padmavani, B.; Gupta, M. K., Gallium (III) halide-catalyzed coupling of indoles with phenylacetylene: synthesis of bis (indolyl) phenylethanes. Tetrahedron letters 2004, 45, (41), 7577-7579. DOI: 10.1016/j.tetlet.2004.08.126.

83.

Azizian, J.; Teimouri, F.; Mohammadizadeh, M. R., Ammonium chloride catalyzed onepot

synthesis

of

diindolylmethanes

under

solvent-free

conditions.

Catalysis

Communications 2007, 8, (7), 1117-1121. DOI: 10.1016/j.catcom.2006.06.002. 84.

Das, P. J.; Das, J., Synthesis of aryl/alkyl (2, 2′-bis-3-methylindolyl) methanes and aryl (3, 3′-bis indolyl) methanes promoted by secondary amine based ionic liquids and microwave irradiation. Tetrahedron Letters 2012, 53, (35), 4718-4720. DOI: 10.1016/j.tetlet.2012.06.106.

Page 49 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

85.

Page 50 of 52

Mulla, S. A.; Sudalai, A.; Pathan, M. Y.; Siddique, S. A.; Inamdar, S. M.; Chavan, S. S.; Reddy, R. S., Efficient, rapid synthesis of bis (indolyl) methane using ethyl ammonium nitrate

as

an

ionic

liquid.

RSC

2012,

Advances

2,

(8),

3525-3529.

DOI: 10.1039/C2RA00849A. 86.

Chakraborti, A. K.; Roy, S. R.; Kumar, D.; Chopra, P., Catalytic application of room temperature ionic liquids:[bmim][MeSO 4] as a recyclable catalyst for synthesis of bis (indolyl) methanes. Ion-fishing by MALDI-TOF-TOF MS and MS/MS studies to probe the proposed mechanistic model of catalysis. Green Chemistry 2008, 10, (10), 1111-1118. DOI: 10.1039/B807572G.

87.

Alinezhad, H.; Haghighi, A. H.; Salehian, F., A green method for the synthesis of bisindolylmethanes and 3, 3′-indolyloxindole derivatives using cellulose sulfuric acid under solvent-free conditions. Chinese Chemical Letters 2010, 21, (2), 183-186. DOI: 10.1016/j.cclet.2009.09.001.

88.

Handy, S.; Westbrook, N. M., A mild synthesis of bis (indolyl) methanes using a deep eutectic

solvent.

Tetrahedron

Letters

2014,

55,

(35),

4969-4971.

DOI:

10.1016/j.tetlet.2014.07.024. 89.

Yadav, J. S.; Reddy, B. V. S.; Murthy, C. V.; Kumar, G. M.; Madan, C., Lithium perchlorate catalyzed reactions of indoles: An expeditious synthesis of bis (indolyl) methanes. Synthesis 2001, 2001, (05), 0783-0787. DOI: 10.1055/s-2001-12777.

90.

Mendes, S. R.; Thurow, S.; Penteado, F.; da Silva, M. S.; Gariani, R. A.; Perin, G.; Lenardão, E. J., Synthesis of bis (indolyl) methanes using ammonium niobium oxalate (ANO) as an efficient and recyclable catalyst. Green Chemistry 2015, 17, (8), 4334-4339. DOI: 10.1039/C5GC00932D.

91.

Talukdar, D.; Thakur, A. J., A green synthesis of symmetrical bis (indol-3-yl) methanes

using phosphate-impregnated titania catalyst under solvent free grinding conditions. Green Chemistry Letters and Reviews 2013, 6, (1), 55-61. DOI: 10.1080/17518253.2012.703700.

92.

Halimehjani, A. Z.; Hooshmand, S. E.; Shamiri, E. V., Synthesis and characterization of a tetracationic acidic organic salt and its application in the synthesis of bis (indolyl)

Page 50 of 52 ACS Paragon Plus Environment

Page 51 of 52

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

methanes and protection of carbonyl compounds. RSC Advances 2015, 5, (28), 2177221777. DOI: 10.1039/C5RA01422K. 93.

Guo, X.; Pan, S.; Liu, J.; Li, Z., One-pot synthesis of symmetric and unsymmetric 1, 1-Bisindolylmethanes via tandem iron-catalyzed C− H bond oxidation and C− O bond cleavage.

The

Journal

of

organic

chemistry

2009,

74,

(22),

8848-8851.

DOI: 10.1021/jo902093p. 94.

Niu, T.; Huang, L.; Wu, T.; Zhang, Y., FeCl 3-promoted alkylation of indoles by enamides. Organic & biomolecular chemistry 2011, 9, (1), 273-277. DOI: 10.1039/C0OB00709A.

95.

Badigenchala, S.; Ganapathy, D.; Das, A.; Singh, R.; Sekar, G., Iron (II) Chloride–1, 1′Binaphthyl-2, 2′-diamine (FeCl2–BINAM) Complex Catalyzed Domino Synthesis of Bisindolylmethanes from Indoles and Primary Alcohols. Synthesis 2014, 46, (01), 101109. DOI: 10.1055/s-0033-1340052.

96.

Hikawa, H.; Yokoyama, Y., Pd-catalyzed C–H activation in water: synthesis of bis (indolyl) methanes from indoles and benzyl alcohols. RSC Advances 2013, 3, (4), 1061-1064. DOI: 10.1039/C2RA21887A.

97.

Deb, M. L.; Bhuyan, P. J., A novel and efficient method for the synthesis of unsymmetrical diindolylmethanes and heterocyclic (indolyl) alkanes. Synthesis 2008, 2008, (18), 2891-2898. DOI: 10.1055/s-2008-1067217.

Page 51 of 52 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 52 of 52

Table of content graphic

Nanocomposite of MoS2-RGO has been developed and used as highly efficient catalyst for the synthesis of indole alkaloids in water.

Page 52 of 52 ACS Paragon Plus Environment