Facile and Ultrafast Green Approach to Synthesize Biobased

Sep 7, 2017 - The improved electrical conductivity of RCD is attributed to the synchronized restoration of sp2 carbon networks and the elimination of ...
0 downloads 14 Views 3MB Size
Subscriber access provided by UNIVERSITY OF THE SUNSHINE COAST

Article

Facile and ultrafast green approach to synthesize bio-based luminescent reduced carbon nanodot: an efficient photocatalyst Rituparna Duarah, and Niranjan Karak ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.7b02590 • Publication Date (Web): 07 Sep 2017 Downloaded from http://pubs.acs.org on September 12, 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 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

Facile and ultrafast green approach to synthesize bio-based luminescent reduced carbon nanodot: an efficient photocatalyst Rituparna Duarah# and Niranjan Karak#* #

Advanced Polymer and Nanomaterial Laboratory, Center for Polymer Science and Technology,

Department of Chemical Sciences, Tezpur University, Napaam, Tezpur-784028, Assam, India. *E-mail: [email protected].; Tel: + 91-3712-267009; Fax: +91-3712-267006.

ABSTRACT: Rising awareness pertaining to global waste management and environmental issues challenges development of efficient metal-free photocatalytic system for visible light assisted degradation of organic contaminants. We, herein, report the synthesis of bio-based luminescent reduced carbon nanodot (RCD) (average size 3 nm) by green reduction of starch based carbon nanodot (CD) using aqueous extracts of Calocasia esculenta leaf, Mesua ferrea Linn leaf, tea leaf and flower bud of Syzygium aromaticum. The reduction was found to be ultrafast (3 min) under sonication using Calocasia esculenta leaf extract in presence of Fe3+ ions at room temperature. The synthesized RCD is an efficient photocatalyst for the degradations of model dirt like methylene blue, methylene orange and their mixture as well as toxic chemical like bisphenol A, under normal sunlight. These degradations followed the pseudo-first-order kinetics model. The catalytic efficiency of RCD was significantly higher than CD. The structure of RCD was characterized by UV-visible, Fourier Transform Infrared, energy dispersive X-ray and Raman spectroscopic analyses; X-ray diffraction and transmission electron microscopic

ACS Paragon Plus Environment

1

ACS Sustainable Chemistry & Engineering

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

Page 2 of 39

studies. The photoluminescence characteristic of RCD was analyzed by fluorescence spectroscopy. The results showed that exploration of sustainable-resource based RCD may offer a novel scope in resolving environmental and ecological problems. KEYWORDS: Reduced carbon nanodot; carbon nanodot; phytoextract; photocatalyst; photodegradation.

INTRODUCTION Chemical transformation using green solvent and nontoxic chemical under ambient reaction conditions by harnessing the abundant energy of solar radiations has emerged as one of the major thrusts in scientific research in order to combat with the adverse impact of industrial effluents on environment and human health.1 Inspite of the significant role of the chemical based industries in society, widespread industrial and anthropogenic ventures have introduced huge amount of chemicals such as surfactants, pesticides, dyes etc. into the environment which destroys the ecosystem.2 The versatility and complexity of such chemicals in use make it difficult to find a common treatment method that completely covers the efficient elimination of all categories of organic pollutants. Although certain physical techniques such as reverse osmosis, flocculation, and adsorption are not harmful, they transport the contaminants to other media which results in secondary pollution.3 Hence there is an imperative demand in the fields of green chemistry and material science to develop newer method with improved performance for destruction of organic contaminants. The exploration of various suitable semiconductor photocatalysts like TiO2, ZnO, Fe2O3 etc. for organic pollutant degradation is one of the major strategies to resolving environmental pollution and energy crisis.3,4 However, they need higher energy UV and/or short wavelength visible radiations. Therefore, exploiting a photocatalyst that can be driven by solar

ACS Paragon Plus Environment

2

Page 3 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

radiations in order to realize the photo-assisted degradation of organic compounds and reduction of inorganic ions with high efficiency is the order of days.4,5 Furthermore, sunlight based photocatalysis is a sustainable and economical practice owing to the use of solar energy which is a nonpolluting, inexpensive and endlessly renewable source of clean energy.6,7 It is pertinent to mention that near infrared (NIR) and infrared (IR) light radiations are still not completely exploited by reported photocatalysts. In this milieu, an emerging nanomaterial, carbon nanodot (CD) has inspired intensive research owing to their excellent biocompatibility, photoluminescent, photostability and nanoscale dispersibility in water.8,9 CD exhibits a definite “optical” energy gap which apparently depends upon its surface texture, size and shape. These guide photo-induced electron transfer capability, down and up-converted PL and excitation-energy-dependent photoluminescence (PL) of CD.10 As CD normally absorb in the UV, green and blue regions, their utilizations in the region of relatively longer wavelength is limited. Even dimensional modification, heteroatom doping and surface passivation etc. proved ineffective to utilize CD in the region of red and near-infrared (NIR) light.11 In addition, the occurrence of up-conversion by CD system is disputed in several studies.12 However, the effective photocatalysis accomplished through down-conversion is also cited in literature.12,13 Keeping this in mind, we tried to focus the development of reduced CD (RCD) with a strong visible to NIR light absorption capacity with enhanced fluorescence. The unique structure and PL properties of RCD may open a host of possibilities for efficient use of this photocatalyst.9,10 Literature cites the use of various conventional reducing agents like sodium borohydride, sodium citrate, ascorbic acid etc. to reduce CD without significant enhancement in luminescence.14-16 Further, these chemical routes are unfavorable due to environmental issues. In contrast, no report is found on reduction of CD by using naturally renewable low cost, non toxic

ACS Paragon Plus Environment

3

ACS Sustainable Chemistry & Engineering

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

Page 4 of 39

reducing agent such as phytoextracts which contain different polyphenolic compounds with high reducing capability, though the global concern with respect to the dwindling petro-based resources has instigated the utility of sustainable feedstock in the synthesis of industrially important materials.6,9,10,17 Therefore, we employed a sustainable, efficient and harmless approach for the synthesis of bio-based RCD by reducing starch based CD using aqueous phytoextracts. Thus, the developed bio-based RCD has the potential to tackle widespread challenges of environmental conservation and sustainability.6 Again, advanced oxidation processes (AOPs) including ozonation, photolysis, and photocatalysis are used to mineralize organic compounds into CO2 and H2O with light, oxidants and semiconductors. In this process, the photocatalysts absorb photons (UV) to produce electronhole (e-/h+) pairs where the e- and h+ react with oxygen and water molecules to give hydroxyl radical (●OH) and superoxide radical (●O2-), respectively. These active radicals subsequently oxidizes pollutants to CO2 and H2O.2,7,10 Furthermore, reduction of CD may result in optimum level of peripheral polar functional groups which in turn interact with the organic pollutants and help to adhere on the surface which assists in better interaction with the active superoxide radicals. Therefore, RCD may photocatalytically degrade the organic dyes and other contaminants under sunlight. In this context, previous literature cites the use of reduced graphene oxide (RGO) in photocatalytic degradation of different organic compounds like rhodamine B and methylene blue.18,19 However, the method for the preparation of graphene oxide is very tedious and has its own drawbacks with respect to safety and environmental issues. Therefore, we employed abundant and cost-effective carbon resources to synthesize CD, while RCD was obtained through an environment friendly green technique from this CD. Again, bisphenol A [2, 2-bis (4-hydroxyphenyl) propane, BPA] is a non-degradable, anthropogenic chemical

ACS Paragon Plus Environment

4

Page 5 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

contaminant, both in water and the soil causing endocrine disruption.20,21 Thus, scientists have developed efficient remediation methods

to degrade BPA, including ultrasonic, Fenton

oxidation, H2O2 oxidation and photocatalytic methods.22 However, there are very few reports on BPA degradation using nano photocatalyst. These photocatalysts have very poor photocatalytic efficiencies even under UV light.20-22 In this context, the synthesized RCD with improved photocatalytic activity under sunlight may efficiently degrade BPA. Therefore, we wish to report the synthesis of RCD from bio-based precursors by using C. esculenta leaf extract by a green one-step approach for the first time. Most importantly, this RCD has a greater extent of sustainability in comparison to one that follows a chemical route. Further, the well characterized sustainable resource-based luminescent RCD was attempted to use like an effective nano photocatalyst to degrade organic contaminants such as toxic BPA, methylene blue (MB), methyl orange (MO) and MB/MO mixture under normal sunlight.

EXPERIMENTAL SECTION Materials. Materials. Soluble starch (Sigma-Aldrich, Germany, Mn = 342.30 g mol-1), ethanol (Merck, India) and bisphenol-A (BPA, Sisco Research Laboratories Pvt. Ltd, India) were used as received. Citrus limon (C. limon) fruits and S. aromaticum (Cloves) flower buds were purchased from the local market at Tezpur, India. Calocasia esculenta (C. esculenta), Mesua ferrea Linn. (M. ferrea Linn.) and tea leaves were collected from neighboring area. Phytoextract preparation. About 2 g of washed C. esculenta, M. ferrea Linn. and tea leaves; and S. aromaticum flower buds were coarsely grounded and subsequently stirred for about 30 min in 50 mL water at 60 °C. The aqueous extract was then filtered (Whatman no. 1 filter paper) under ambient condition.

ACS Paragon Plus Environment

5

ACS Sustainable Chemistry & Engineering

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

Page 6 of 39

RCD synthesis. At first synthesis of CD was conducted by a facile green one-step hydrothermal synthetic route with starch and C. limon extract, without base, as reported in our previous study.23 Briefly, starch solution (5 g in 40 mL distilled water) with a few drops of C. limon extract and 20 mL ethanol were taken in an autoclave and heated at 150 °C for 5 h to obtain a brown water soluble CD with a yield of 54 mg mL-1. Subsequently, RCD was synthesized by reducing CD via a single step green facile process using various phytoextracts. In a typical process, 10 mL of aqueous extract of C. esculenta leaf and Fe3+ ions (10 mL of 0.01 M) were slowly added drop wise to 100 mg of CD solution at room temperature under continuous stirring. Other aqueous phytoextracts such as M. Ferrea Linn. leaf, tea leaf and S. aromacticum flower bud with Fe3+ ions as well as C. esculenta with other metal ions Cu2+, Ni2+ and Cr2+ ions were also used separately for the reduction of CD under the same conditions. The change of color from light brown was taken as completion of the reduction process and was supported by UV analysis. The product was washed thrice with distilled water and separated by centrifugation for 10 min at 8000 rpm. This process removed water soluble components and finally the residue was redispersed in water by ultrasonication. The preparation of RCD was also carried out under refluxed condition using the same technique and reducing agents for comparison purposes. Characterization. The structures and chemical compositions of CD and RCD were characterized by FTIR, XRD, TEM, EDX, UV-visible spectroscopic analyses. Optical and electrical properties were analyzed by photoluminescent set up and conductivity meter. The details of these techniques are given in Electronic Supporting Information (ESI). Photocatalytic activity. Organic dyes such as MB, MO, MB/MO mixture and organic contaminants (OCs) like BPA were used to study the photodegradation activity of RCD and CD.

ACS Paragon Plus Environment

6

Page 7 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

Typically, 50 mg of RCD was taken in four separate flasks containing 100 mL aqueous solution of MO (10 mg/L), MB (10 mg/L) and MB/MO mixture (10 mg/L each, separately) and BPA (50 mg/L). The solutions were made to stir under normal sunlight (60000-80000 lux). In addition, the experiment was conducted using same amount of CD for comparison purposes. The changes in concentrations of MB, MO and BPA were monitored from the UV absorbance intensity at wavelengths of 657 nm, 459 nm and 273 nm, respectively under specific duration of time.19,20 The catalytic efficiency of RCD was calculated from the change of concentration rate of the OC. The same experiment was also conducted using same amount of CD for comparison purpose. The amount of the degraded pollutants was obtained from the equation as given below.10 Degradation (%) = [(C0-C)/C0 × 100]…………………………………………. (i) Where C0 = initial concentration and C = concentration after photocatalytic degradation of the OC.

RESULTS AND DISCUSSION RCD synthesis and characterization. CD was synthesized by hydrothermal acid hydrolysis of the aqueous ethanolic solution of starch and citric acid as the bio-based carbon precursors using the same reported method.23 The details of synthesis and most of the characterizations of CD are provided in the ESI. The times required for catalytic reduction of CD under reflux conditions as well as at room temperature are given in Table S1 of ESI. We focus on reduced CD in order to achieve a strong visible to NIR light absorption capacity with enhanced fluorescence of it which is ultimately used as an efficient solar light based photocatalyst. There are reports on reduced graphene oxide (RGO) as a photocatalyst for degradation of various OCs like rhodamine B and MB.24 However, the method for the preparation of graphene oxide is very tedious and has its own

ACS Paragon Plus Environment

7

ACS Sustainable Chemistry & Engineering

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

Page 8 of 39

drawbacks with respect to safety and environmental issues. Therefore, we focused on the green reduction of CD by employing easily available and cheap carbon resources through an environment friendly greener technique for the first time. As mentioned in the literature CD contains a large number of polar peripheral groups which imparts exceptional nano state aqueous dispersibility, intriguing optical properties and biocompatibility. However, reduction of CD using renewable resources like phytochemicals results in decrease of oxygen content in it by removing some of the oxygeneous functional groups. This optimum level of polar functionalization on RCD helps in better interaction with the polar groups of the organic pollutants and adheres on their surfaces, thus assisting in interaction with the active superoxide radicals. Also upon reduction, there is generation of graphitized aromatic structure in RCD which helps to delocalize the photo generated electrons and stabilized them. This consecutively delayed the recombination of e-/h+ pairs. Further, RCD has lower band gap than CD, and thus generates more electron/hole pairs upon absorption of light energy. Therefore, we focus on reduced CD (RCD) in order to get improved photocatalytic activity of OCs under sunlight. The results showed that reduction was fastest in case of C. esculenta leaf extract among the used phytochemicals. Further, the reduction was quicker under refluxed condition than at room temperature in all the cases because at higher temperature the activation energy for CD reduction was attained at a much faster rate. We observed that the time for reduction was unaffected by the amount of phytoextract. Further, the reduction of CD by phytochemicals with different metal ions (Fe3+, Cu2+, Ni2+ and Cr2+ ions) under ambient condition revealed the effective reduction only with Fe3+ ions.24 The reduction times for different phytoextracts in the presence of Fe3+ ions, with and without sonication clearly demonstrated the efficiency of C. esculenta leaf extract (Table S2). It took 3

ACS Paragon Plus Environment

8

Page 9 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

min and 10 min with and without sonication, respectively (Table S2). This is a significant accomplishment as the reduction time was drastically reduced by this approach (Table S1 and Table S2). This may be due to complex formation between Fe3+ and the polyphenol groups of the phytoextract as supported by UV-visible absorbance data.17,25 The schematic representation of the synthesis of RCD using C. esculenta leaf extract in presence of Fe3+ ions is shown in Scheme 1. Literature cites C. esculenta leaf extract contains polyphenolic compounds like apigenin, pectins, luteolin, flavonoids, ascorbic acid and various flavones which showed strong potential for reduction of graphene oxide as reported earlier.17,26 The reduction potential of this phytoextract was also reported to be enhanced by addition of Fe3+ ions.27 These polyphenolic compounds contain different polar functional groups such as hydroxyl, carboxy etc. which can form complexes with Fe3+ ions (Scheme 2).17,25 The absorption at 277 nm may be due to the formation of such complexes as depicted in the given reaction.24 6ArOH + FeCl3

H3[Fe(OAr)6] + 3HCl ……………….. (ii) Iron polyphenolic complex

The reduction time was dependent on the complex formation rate, as Fe3+ ions might form a stronger and stable complex with polyphenol compounds, present in the phytoextract as compared to other studied metal ions, thereby exhibiting quicker reduction. During complexation, a large number of H+ ions were released, which brought about a change in the pH of the system.25 The pH of metal ion containing C. esculenta leaf extract and phytoextract were measured to be around 3 and 6, respectively. For the better understanding of the result of pH, we also conducted the reduction by regulating the pH of the mixture at 3. However, even after 1 h, no effective reduction was achieved which implies that pH has no influence in the course of reduction. After completion of reduction, the UV-visible spectrum of RCD shows a red shift of

ACS Paragon Plus Environment

9

ACS Sustainable Chemistry & Engineering

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

Page 10 of 39

the characteristic peak at 269 nm, suggesting that electronic conjugation was restored.25,28 Further, the optical absorption of RCD was found to be higher than that of CD (Figure 1a). To clarify the role of Fe3+ ions in reduction process of CD, an additional experiment was also conducted where we observed that CD was not reduced with only Fe3+ ions even after prolonged time. There was no red shift of the characteristic peak of CD in the UV visible spectrum of the solution containing CD and Fe3+ as shown in Figure S1. This clearly indicates that Fe3+ is not a reductant in its bare state. However, as Fe3+ helps in complexation with the polyphenol compound and enhances the rate of reduction of CD, it is a promoter in this catalytic process. Further, literature supports the formation of H+ ions and faster release of electrons during complexation of C. esculenta and Fe3+ ions, which have a strong influence in this reduction process.24 From the above results it was observed that the preparation of RCD was the fastest and most effective using aqueous C. esculenta leaf extract with Fe3+ ions, under normal atmospheric condition, so this process was used for further study. Scheme 2 shows the plausible mechanism of electron transfer for CD reduction through electron transfer and nucleophilic attack (SN2). The FTIR absorbance bands (ν max/cm−1) like C=O (1731), C-O-C (1420), C-O (1266) and epoxy group (917) in CD (Figure S2a) were diminished in RCD (Figure 1b (i)).25 The FTIR spectrum of RCD was also compared with CD reduced by sodium borohydride spectrum (Figure 1b (ii)) to understand the reduction. The XRD patterns show that the broad peak of CD near 2θ ~ 21° was disappeared while peak near 32° was slightly sharpened in RCD (Figure 1c), thus indicating the formation of a more graphitic structure.29 The characteristic bands such as G band (1590 cm-1), D band (1361 cm-1) and 2D band (2912 cm-1) in Raman spectra of RCD (Figure 1d) clearly demonstrated the increase of ID/IG, as compared to CD (Figure S2b), thus indicating multilayer graphitization of RCD. This is due to restoration of sp2 carbon with decrease in its

ACS Paragon Plus Environment

10

Page 11 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

domain size upon reduction of CD, as well as owing to the unrepaired defects that remained even subsequent to the removal of oxygeneous groups.30,31 This value of ID/IG ratio is consistent with majority of the literature on chemical reduction of RGO.24,32,33 In addition there is an increase in carbon to oxygen ratio (C/O) of RCD (11.2) in EDX analysis (Figure 1e) as compared to CD (1.59) (Figure S2c). HRTEM images confirmed the spherical shapes of CD (Figure S3) and RCD (Figure 2). Further, the average size 3 nm and lattice spacing 0.20 nm were found for RCD. Poor crystalline nature of RCD was also supported by selected area electron diffraction (SAED) pattern (as shown inset of Figure 2b). It is pertinent to mention that we also used the Bragg’s equation to calculate the inter-planar distance of RCD using the following equation: d = nλ/2sinθ……………………………………………………….. (iii) Where n = integer (1), λ = wavelength of the incident x-ray beam (0.154 nm) at certain angle of incidence (theta, θ = 32°). Using the above equation, the lattice spacing of RCD was found to be 0.15 nm, for the peak 32° whereas from the TEM the lattice fringe was calculated to be 0.20. Thus the difference is not very significant. However, this little difference may be explained as follows. During TEM analysis, very dilute solution of the sample was taken and the lattice fringes were determined for a particular particle of RCD where the inter-planar spacing between the two planes was found to be 0.20 nm. On the other hand during XRD analysis the sample was taken in powdered state due to which agglomeration take place which generates XRD peaks of many particles of different characters. Thus, the lattice fringe of RCD determined from TEM and XRD analyses may differs slightly. Optical property. The UV-visible absorption spectrum for CD displayed absorption peaks at around 220 nm, attributed to the π-π* transition band of the conjugated double bond and at around 280 nm (through a tail continuing into the visible range), credited to the band of n-π*

ACS Paragon Plus Environment

11

ACS Sustainable Chemistry & Engineering

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

Page 12 of 39

transition of the carbonyl bond (Figure 1a (i)).9,23 However, upon reduction of CD, the band near 280 nm was disappeared and a peak near 220 nm undergoes a red shift to 269 nm which indicates the restoration of conjugation and resulting in the formation of an aromatized multilayered structure (Figure 1a).24 Further, RCD exhibited a broad absorption band over a wide range of wavelength from 250 to 800 nm, indicating effective photo-absorption which would be useful for its photocatalytic activity under visible light.34 This broad absorption band over a wide range of wavelength is due to graphitic structure of RCD which has long π-π conjugation. The aqueous solutions of CD and RCD were found to be brown and dark brown in color, respectively, under normal light however exhibited green and blue fluorescence under UV light at various wavelengths as shown in Scheme 1. Upon reduction of CD, the quantum yield increased from 9.6 to 18.5% (excited at 340 nm, using Quinine sulfate as the reference). The size dependent photoluminescence is the characteristic feature of the synthesized CD and RCD. The PL spectra of CD and RCD at same concentration (10 mg/mL) shows that the intensity depends on their respective concentrations and excitation wavelengths (Figure 3).6,9 With an increase in the excitation wavelength, the emission peaks of both CD and RCD shifted to a higher wavelength, reaching fluorescence maximum while excited at wavelength 340 nm but subsequently decreases with further increase in it (Figure 3a and b). The highest PL intensity of CD and RCD was attributed to the nature of their surfaces and the fact that maximum number of particles was excited at that particular wavelength. The polar functional groups present on the surfaces of CD and RCD might consequence in the emissive traps in-between π and π* transition of C=C. The emission is dominated by a surface energy trap when CD and RCD are illuminated at a definite excitation wavelength.9 The intensity of the PL spectra increases with increase in the concentration of CD and RCD. This is due to the increase of interaction among polar surface

ACS Paragon Plus Environment

12

Page 13 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

groups at high concentration (Figure 3c and d). The difference in size of the nanoparticles leads to the variation in the position of emission peak. Similar to semiconductor quantum dots, the energy gap of CD and RCD increases with the decrease in size because of quantum effect. Therefore, particles with larger size get excited at higher wavelengths, whereas the particles with smaller size get excited at lower wavelength.6,35 The enhanced green emission of the RCD is due to the zig-zag sites of its surfaces.14,36 The following equation was used to calculate the optical band gaps of CD and RCD from the UV-visible spectra.21,24 α = C(hv – Ebulk)1/2/hv………………………………………………..(iv) Where α = absorption coefficient, C = constant, h = plank’s constant, ν = is the frequency and Ebulk = ‘band gap’. The band gaps were calculated by drawing the tangent on the plot hv versus (αhv)2 that cut at the X-axis (Figure 4a and b). The band gaps of CD and RCD were found to be 3.64 and 3.18 eV, respectively. The result indicates that the reduction of CD results in restoration of the π conjugated system. These band gap values are similar to the reported band gap of RGO and GO system.25,37 Further, contrary to the commercial TiO2, CD and RCD belong to a class of carbonaceous organic semiconductors which provide a sustainable and green preparative route by using environmentally benign solvents, nontoxic chemicals and renewable precursors. Compared to traditional semiconductor catalysts like TiO2, photoluminescent CD and RCD are superior in terms

of

functionalization,

resistance

to

photo

bleaching,

toxicity

and

profound

biocompatibility.6,16 In this regard, RCD may display an efficient photocatalytic degradation of OCs under sunlight, due to its strong visible light absorption band and enhanced fluorescence.

ACS Paragon Plus Environment

13

ACS Sustainable Chemistry & Engineering

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

Page 14 of 39

Electrical conductivity. Electrical conductivities of CD and RCD were measured from their current-voltage (I-V) characteristics as obtained by a four probe setup (Figure 4c). RCD exhibited a linear I-V relationship with the change of voltage in the range of -10 - +10 V. The slope of I-V plot for CD was found to be almost equal to zero. This clearly indicates that, prior to reduction, CD behaved like an insulating material.

This is due to the presence of large

oxygeneous groups. The structure of CD is mainly amorphous owing to distortions from high amount of sp3-carbon. The random distribution separates the sp2-hybridized aromatized rings from the sp3-hybridized rings which lead to the insulating nature of CD. Nevertheless, the I-V slope of RCD considerably increased after reduction, indicating high electrical conductivity. Further, the linear behavior of the I-V curve of RCD relates to the Ohmic contact of the graphitic structure with the electrode. The improved electrical conductivity of RCD is attributed to the synchronized restoration of sp2 carbon networks and the elimination of oxygeneous groups upon reduction. The conductivities of CD and RCD were found to be 2.54 x 10-7 Sm-1 and 2.53 x 10-6 Sm-1, respectively. The Raman results and the I-V measurement data revealed that with increasing sp2 carbon content and ID/IG value, the conductivity of CD increased. The results follow a similar trend as reported by Lopez et al. where they showed that the chemical vapor deposition-graphene oxide (CVD-GO) demonstrated a linear increase in electrical conductivity with the increase of ID/IG value.38 It is pertinent to mention that this conductivity is not due to the presence of Fe3+ ions, as no traces of iron was present in RCD; confirmed from EDX analysis (no peak was observed for iron at ~6.4 keV) (Figure 1e). Further, the qualitative test of aqueous ammonium thiocyanate (NH4SCN) solution resulted in no color for aqueous dispersion of RCD, whereas it produces a blood-red color for the solution containing CD, phytoextract and Fe3+ ions.

ACS Paragon Plus Environment

14

Page 15 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

Thermal behavior. CD exhibited a two step thermal degradation pattern, whereas more steps were observed for RCD as shown in Figure 5. In both the cases, initial (2-4)% weight loss near (110-112) °C may be attributed to loss of entrapped water molecules between the nanoparticles due to their surface polar functional groups.25 The definite preliminary degradation for CD and RCD commences near 200 °C due to the loss of labile oxygeneous groups (hydroxyl, epoxy etc.). RCD exhibited a steady loss of only 18 wt% up to 280 °C, which was much lower than CD where loss of 32 wt% up to 250 °C was observed. These results indicate a notable decrease in amount of oxygeneous groups in RCD. Further RCD showed minimum weight loss up to the temperature of 250 °C and it exhibited a weight loss of 38% between (500-800) °C. This may be attributed to the presence of phytoextract (PE) bound to the surface of RCD. This can be confirmed from the TGA curve of pure PE which shows the same trend.39 Therefore, RCD experiences (48-50)% less weight loss as compared to CD in the temperature range of (300-800) °C, which is also an indication of elimination of oxygeneous groups by reduction and higher thermal stability of RCD compared to CD. The residual weights obtained for CD and RCD at 800 °C was about 12% and 62%, respectively. Photocatalytic activity. The pathways for photocatalytic degradation of organic contaminants (OCs) such as MB, MO, MB/MO mixture and BPA were studied by CD and RCD under normal sunlight as shown in Scheme 3. Actually, when RCD/CD nanoparticles were exposed to the OC solution, they absorbed visible light along with near infrared light which led to the production of e-/h+ pairs in RCD/CD by excitation of valence band (VB) electrons.40 The generation of e-/h+ pairs in the RCD/CD photocatalyst is shown in equation (v).41 Photon activation: RCD/CD + hν

e- + h+ ………………….. (v)

ACS Paragon Plus Environment

15

ACS Sustainable Chemistry & Engineering

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

Page 16 of 39

Consequently, the photo-generated electrons in RCD/CD easily react with the oxygen molecules (O2) present in the aqueous medium to produce superoxide radicals (●O2-) whereas the photo-generated holes in the VB react with water molecules to produce hydroxyl radicals (●OH). These reactive oxidative species, ●O2- and ●OH degrade the organic molecules through an oxidative pathway, as described below in equation (vi)-(vii).40 Adsorbed oxygen: (O2)ads + e-



Water/moisture present, H2O + h+

O2- ……………………….. (vi) ●

OH + H ………………………… (vii)

Accordingly, the amounts of ●OH radical depend on the quantity of h+ generated in RCD/CD. Furthermore, the quantity of h+ also determines the ability of photocatalytic degradation. Subsequently, the organic pollutants are transformed into their degraded products by these active oxygen radicals into CO2 and H2O via photocatalytic pathways as shown in equation (viii)-(ix).40,41 R-H + ●OH ●

R + h+



● +

R

R + H2O …………………………….. (viii) CO2 + H2O ………………….. (ix)

It is pertinent to mention that RCD has the ability to absorb more solar light in comparison to CD, as revealed from their absorption spectra (degradation results are shown in Figure 6-8). This assists in increasing the photocatalytic efficiency of RCD under sunlight. RCD was finely ground by using a mortar and a pestle in order to increase its surface area. UV absorbance was used to monitor the concentration changes of the OCs with time. Figure 6 and 7 show the plots of optical absorbance against wavelength for photodegradation of OCs at various time intervals for both RCD and CD.

The photocatalytic activity of a semiconductor results from the excited e-

produced under UV light in the conduction band together with the equivalent h+ in the valence band, which react with the pollutants adsorbed on the surface of the photocatalyst.10 The

ACS Paragon Plus Environment

16

Page 17 of 39

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

ACS Sustainable Chemistry & Engineering

photocatalytic efficiency of CD is reduced by the relatively larger band gap of CD owing to low absorption capability of visible light and high recombination rate of photo-generated e-/h+ pairs that are produced during the photocatalytic process. Due to the lower band gap of RCD (3.18 eV) than CD (3.64 eV), RCD has higher ability of absorbance towards sunlight than CD and thus the energy required for photo-generation of e-/h+ pair is relatively low.40 Thus, the effective number of electrons generated for utilization in the degradation of OCs is higher in RCD compared to those in CD. Although recombination of e-s and h+s in RCD takes place, the neat number of e-/h+ pairs required for generating ●O2- and ●OH to degrade the OCs is higher than CD. Also RCD contains delocalized π electrons as it gets stability through the π conjugation in its aromatized structure which helps in reduction of e-/h+ recombination. Thus, RCD possesses higher photocatalytic efficiency due to its greater surface adsorption capacity to hydroxyl groups and lower charge carrier recombination rate as compared to CD. The degradations of different dyes and BPA are cited in literatures. The degradation rate of methylene blue (MB), methyl orange (MO) and MB/MO mixture with time for RCD and CD under solar light are shown in Figure 6 and 8. Wang et al. reported that 50 mg of Zn (II) based catalyst in 10 ppm solution of MB and 10 ppm solution of MO showed effective degradation of 90.9% and 91.7%, respectively under UV light irradiation, within 120 min.42 Tang et al. showed that using 0.3 g CaIn2O4 photocatalyst in 100 mL aqueous solution of MB at 47.8 µmol/L concentration, nearly 80% degradation was observed under visible light within 120 min.43 Literature also reports that 0.1 g of TiO2/ZnO photocatalyst can degrade 97% MO at 20 mg/L concentration 30 min of UV light irradiation.44 Conversely, 50 mg RCD can degrade 100% MB (10 ppm) within just 45 min and 96% MO (10 ppm) within 60 min under sunlight (Figure 6a and c). The dye degradation using CD was also tested to study the effect of loading and the

ACS Paragon Plus Environment

17

ACS Sustainable Chemistry & Engineering

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

Page 18 of 39

advantage of RCD compared to CD as shown in Figure 6 b, d and f. CD took longer time for the same degradation of organic contaminants which was completed by RCD in a shorter time. This is due to their ability to act as effective solar light assisted photocatalysts and not by the sensitization of dye molecules. The photocatalytic degradation activity of CD and RCD was also studied on the MB/MO mixture. It was observed that after 15 min of exposure under sunlight, the intensity of UV absorption peak of MO decreased sharply in case of RCD, while the intensity for MB still remained high level as shown in Figure 6e. After subsequent exposure under sunlight up to 90 min, MB underwent rapid degradation by RCD. However, both CD and RCD exhibited a complete higher degradation rate for MO in comparison to MB in the MB/MO mixture (Figure 6e and f). The difference implies that RCD and CD exhibit better degradation of MO in the MB/MO mixture. Lower rate of degradation of MB in MB/MO mixture as compared to individual MB solution is primarily attributed to the presence of N=N (which makes MO more reactive), -CH3 group in MB (which makes it resistant to photodegradation) and the occurrence of competitive consumption of the oxidizing species in MO.39 Literature sites similar preferences in the photocatalytic degradation between MB/MO mixture for mesoporous ZnO/ZnAl2O4 powder45 and NiO-Bi2O3 nanocomposite as reported earlier.46 For the same reasons as stated above, RCD demonstrated superiority over CD when the selfcleaning property was studied by films coated with silica for photocatalytic degradation of MB as model dirt was studied by photocatalytic degradation of MB under sunlight by visual means as shown in Figure 6g. Both the nanomaterials, RCD and CD decolorize the solution of MB. For an improved comparison, glass slides with silica and silica with RCD and CD films were dipped in dye solution and exposed to sunlight. The change in color of the films was recorded at different

ACS Paragon Plus Environment

18

Page 19 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

times (Figure 6g). The photographs demonstrate that RCD exhibits better decolorization of MB within 70 min in comparison to CD. Thus, RCD can be used as a material for designing selfcleaning surfaces. Among various OCs, BPA is known mostly difficult to be degraded under visible light and a number of reports have shown the degradation of BPA by inorganic semiconductor based photocatalysts under UV irradiation. Bechambi et al. reported that 1 gL-1 C-doped ZnO in 50 mg L-1 BPA with H2O2 at pH 8 effectively attained 100% degradation of BPA after 24 h of UV light irradiaion.21 Wang et al. reported that 1% immobilized TiO2 can degrade 97% BPA at 10 ppm concentration within 6 h upon UV radiation.2 However, only a few reports on the photodegradation of BPA are found under visible light. Qu et al. showed that 10 mg CNT can photodegrade 74.8% BPA at 10 ppm concentration upon exposure of solar light for 180 min.19 On the other hand, 50 mg of RCD can degrade 100% BPA at 50 mg/L concentration within 3.5 h under sunlight (Figure 7). The present study demonstrated the degradation of MB, MO, MB/MO mixture and BPA using RCD under solar irradiation for the first time. In all the above cases of photocatalytic degradation of OCs, the photocatalytic activity of RCD is significantly faster than CD. This may be due to the optimum level of polar functional groups that may interact with the OCs and help to anchor on the surfaces. This helps in better interaction of OCs with the active oxygen radicals. Thus, inspite of using an equal amount of nanoparticles in both the cases, the same degree of degradation was not observed during the same period of exposure. Further, to clarify the doubt of self-degradation of the dye a blank experiment was designed without using CD or RCD. No change in dye intensity even after 5 h of sunlight exposure, as measured by UV-visible spectroscopy, clearly indicates that the dye degradation is not self-degradation (Figure S4a). Further, no color change in the dye solution was observed even after exposure of 5 h sunlight (Figure S5). In this context, another experiment of dye

ACS Paragon Plus Environment

19

ACS Sustainable Chemistry & Engineering

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

Page 20 of 39

degradation with CD and RCD was performed under dark condition. In this case also no change in dye intensity was observed even after prolonged time, as measured by UV-visible spectroscopy (Figure S4b and c). Thus, this result eliminates the possibility of adsorption of dye molecule by the nanomaterial. The photo-degradation behavior of the nanomaterials was described by the pseudo-first order kinetics model equation as given below (Figure 7 and 8): -dC/dt = K1t……………………………………………………………………… (x) Where C = concentration of the OC at specific time t and K1 = apparent rate constant. After integrating equation (x), we obtain the following equation: ln(C/C0) = -K1t……………………………………………………………………... (xi) Where C0 = initial concentration (at t = 0) of OC. The fitting plots of ln(C/C0) versus time are shown in Figure 7d and 8d-f, which reveal that the OC degradation is well fitted by pseudo-firstorder kinetics model with the fitting coefficients over 0.8, representing a regular photocatalytic degradation behavior. Although CD degrades OCs, their photocatalytic efficiencies are very poor due to their low absorbance of visible light and high recombination rate of photo generated e−/h+ pairs. All those above results are clearly indicating superior photocatalytic activity of RCD as compared to CD.

CONCLUSION Thus, from this study, it can be concluded that bio-based carbon nanodots (CD) can be effectively reduced through a simple benign one pot ultrafast route using aqueous phytoextract of C. esculenta in the presence of Fe3+ ions. This reduced CD (RCD) exhibited excellent excitation wavelength dependent fluorescence. This novel zero dimensional carbon based nanomaterial

ACS Paragon Plus Environment

20

Page 21 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

efficiently able to degrade organic pollutants like methyl orange, methylene blue and methyl orange/methylene blue mixture and bisphenol A under normal solar radiation. The photocatalytic degradation efficiency of RCD is found to be more compared to CD. These degradation processes follow the pseudo-first-order kinetics model. Therefore, this electron transfer mechanism of RCD may be broadened for its prospective applications in the field of solar cells and photo-catalysis. In a nutshell, environment friendly RCD keep promises as a photocatalyst towards economical and efficient to make the eco-system safe and sustainable.

ASSOCIATED CONTENT Supporting Information Characterization techniques; synthesis and characterization of CD; FTIR spectrum, Raman spectrum, HRTEM images and EDX map of CD; UV-vis spectra of CD, CD + Fe3+, Fe3+; and dye under different conditions and images for dye at different times of sunlight exposure of CD; Table for the reduction time of CD using different phytoextracts; and Table for the reduction time of CD using different phytoextracts with Fe3+ ions with and without sonication.

AUTHOR INFORMATION Corresponding Author *Email: [email protected].; Tel: +91-3712-267009, Fax: +91-3712-267006 Author Contributions All authors contributed equally and approved the final version of the manuscript.

ACS Paragon Plus Environment

21

ACS Sustainable Chemistry & Engineering

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

Page 22 of 39

ACKNOWLEDGEMENT The authors acknowledge SAIC, Tezpur University, Tezpur and SAIF, NEHU, Shillong, India for HRTEM analyses.

REFERENCES (1) Li, H.; Liu, R.; Lian, S.; Liu, Y.; Huang, H.; Kang, Z. Near-infrared light controlled photocatalytic activity of carbon quantum dots for highly selective oxidation reaction. Nanoscale 2013, 5(8), 3289-3297. (2) Wang, R.; Ren D.; Xia S.; Zhang, Y.; Zhao, J. Photocatalytic degradation of Bisphenol A (BPA) using immobilized TiO2 and UV illumination in a horizontal circulating bed photocatalytic reactor (HCBPR). J. Hazard. Mater. 2009, 169(1-3), 926-932. (3) Ameta, R.; Benjamin, S.; Ameta, A.; Ameta, S. C. Photocatalytic degradation of organic pollutants: A review article. Mater. Sci. Forum 2013, 734, 247-272. (4) Khanchandani, S.; Kumar, S.; Ganguli A. K. Comparative study of TiO2/CuS Core/Shell and composite nanostructures for efficient visible light photocatalysis. ACS Sustainable Chem. Eng. 2016, 4(3), 1487-1499. (5) Wang, W. W; Zhu, Y. J; Yang, L. X. ZnO-SnO2 hollow spheres and hierarchical nanosheets: hydrothermal preparation, formation mechanism, and photocatalytic properties. Adv. Funct. Mater. 2007, 17, 59-64. (6) Gogoi, S.; Kumar, M.; Mandal, B. B.; Karak, N. High performance luminescent thermosetting waterborne hyperbranched polyurethane/carbon quantum dot nanocomposite with in vitro cytocompatibility. Compos. Sci. Technol. 2015, 118, 39-46.

ACS Paragon Plus Environment

22

Page 23 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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) Yoon, T. P.; Ischay M. A.; Du, J. Visible light photocatalysis as a greener approach to photochemical synthesis. Nat. Chem. 2010, 2, 527-532. (8) Zacharakis, A.; Chatzisymeon, E.; Binas, V.; Frontistis, Z.; Venieri, D.; Mantzavinos D. Solar photocatalytic degradation of bisphenol A on immobilized ZnO or TiO2. Int. J. Photoenergy 2013, 2013, 1-9. (9) De, B.; Karak, N. A green and facile approach for the synthesis of water soluble fluorescent carbon dots from banana juice. RSC Adv. 2013, 3, 8286-8290. (10) Hazarika, D.; Karak, N. Photocatalytic degradation of organic contaminants under solar light using carbon dot/titanium dioxide nanohybrid, obtained through a facile approach. App. Surf. Sci. 2016, 376, 276-285. (11) Peng, H.; Travas-Sejdic, J. Simple aqueous solution route to luminescent carbogenic dots from carbohydrates. Chem. Mater. 2009, 21(23), 5563-5565. (12) Jelinek, R. Carbon Quantum Dots: Synthesis, Properties and Applications: Carbon Nanostructures; Springer International Publishing AG, Switzerland, 2016. (13) Zhuo, S.; Shao, M.; Lee, S. T. Upconversion and downconversion fluorescent graphene quantum dots: ultrasonic preparation and photocatalysis. ACS Nano 2012, 6(2), 1059-1064. (14) Zheng, H.; Wang, Q.; Long, Y.; Zhang, H.; Huang, X. Zhu R. Enhancing the luminescence of carbon dots with a reduction pathway. Chem. Commun. 2011, 47, 10650-10652. (15) Zhuo, Y.; Zhong, D.; Miao, H.; Yang, X. Reduced carbon dots employed for synthesizing metal nanoclusters and nanoparticles. RSC Adv. 2015, 5, 32669-32674. (16) Lim, S.Y.; Shen, W.; Gao, Z. Carbon quantum dots and their applications. Chem. Soc. Rev. 2015, 44, 362-381.

ACS Paragon Plus Environment

23

ACS Sustainable Chemistry & Engineering

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

Page 24 of 39

(17) Thakur, S.; Karak, N. Alternative methods and nature-based reagents for the reduction of graphene oxide: A review. Carbon 2015, 94, 224-242. (18) Ghavami, M.; Mohammadi, R.; Koohi, M.; Kassaee, M. Z. Visible light photocatalytic activity of reduced graphene oxide synergistically enhanced by successive inclusion of γ-Fe2O3, TiO2, and Ag nanoparticles. Mater. Sci. Semicond. Process 2014, 26, 69-78. (19) Fan, F.; Wang, X.; Ma, Y.; Fu, K.; Yang, Y. Enhanced photocatalytic degradation of dye wastewater using ZnO/reduced graphene oxide hybrids. Fuller Nanotub. Car. N 2015, 23(11), 917-921. (20) Chiang, K.; Lim, T. M.; Tsen, L.; Lee, C. C. Photocatalytic degradation and mineralization of bisphenol A by TiO2 and platinized TiO2. Appl. Catal. A Gen 2004, 261, 225237. (21) Bechambi, O.; Sayadi, S.; Najjar, W. Photocatalytic degradation of bisphenol A in the presence of C-doped ZnO: Effect of operational parameters and photodegradation mechanism. J. Ind. Eng. Chem. 2015, 32, 201-210. (22) Qu, J.; Cong, Q.; Luo, C.; Yuan, X. Adsorption and photocatalytic degradation of bisphenol A by low-cost carbon nanotubes synthesized using fallen leaves of poplar. RSC Adv. 2013, 3, 961-965. (23) Duarah, R.; Singh, Y. P.; Gupta, P.; Mandal, B. B.; Karak, N. High performance biobased hyperbranched polyurethane/carbon dot-silver nanocomposite: a rapid self-expandable stent. Biofabrication 2016, 8(4), 045013. (24) Thakur, S.; Karak, N. Multi-stimuli responsive smart elastomeric hyperbranched polyurethane/reduced graphene oxide nanocomposites, J. Mater. Chem. A 2014, 2(36), 1486714875.

ACS Paragon Plus Environment

24

Page 25 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

(25) Thakur, S.; Karak, N. Green reduction of graphene oxide by aqueous phytoextracts. Carbon 2012, 50(14), 5331-5339. (26) Chakraborty, P.; Deb, P.; Chakraborty, S.; Chatterjee, B.; Abraham, J. Cytotoxicity and antimicrobial activity of Colocasia esculenta. J. Chem. Pharm. Res. 2015, 7(12), 627-635. (27) Markova, Z.; Novak, P.; Kaslik, J.; Plachtova, P.; Brazdova, M.; Jancula, D.; Siskova, K. M.; Machala, L.; Marsalek, B.; Zboril, R.; Varma, R. Iron(II,III)-polyphenol complex nanoparticles derived from green tea with remarkable ecotoxicological impact. ACS Sustainable Chem. Eng. 2014, 2(7), 1674-1680. (28) Dey, R. S.; Hajra, S.; Sahu, R. K.; Raj, C. R.; Panigrah, M. K. A rapid room temperature chemical route for the synthesis of graphene: metal-mediated reduction of graphene oxide. Chem Commun. 2012, 48, 1787-1789. (29) Stobinski, L.; Lesiak, B.; Malolepszy, A.; Mazurkiewicz, M.; Mierzwa, B.; Zemek, J.; Jiricek, P.; Bieloshapka, I. Graphene oxide and reduced graphene oxide studied by the XRD, TEM and electron spectroscopy methods. J. Electron Spectrosc. Relat. Phenom. 2014, 195, 145154. (30) Cui, P.; Lee, J.; Hwang, E.; Lee, H. One-pot reduction of graphene oxide at subzero temperature. Chem. Commun. 2011, 47, 12370-31272. (31) Stankovich, S.; Dikin, D. A.; Piner, R. D.; Kohlhaas, K. A.; Kleinhammes, A.; Jia, Y.; Wu, Y.; Nguyen, S. T.; Ruoff, R. S. Synthesis of graphene-based nanosheets via chemical reduction of exfoliated graphite oxide. Carbon 2007, 45(7), 1558-1565. (32) Zhou, Y.; Bao, Q.; Tang, L. A. L.; Zhong, Y.; Loh, K. P. Hydrothermal dehydration for the “green” reduction of exfoliated graphene oxide to graphene and demonstration of tunable optical properties. Chem. Mater. 2009, 21(13), 2950-2956.

ACS Paragon Plus Environment

25

ACS Sustainable Chemistry & Engineering

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

Page 26 of 39

(33) Dong, Y.; Niu, X.; Song, W.; Wang, D.; Chen, L.; Yuan, F.; Zhu, Y. Facile synthesis of vanadium oxide/reduced graphene oxide composite catalysts for enhanced hydroxylation of benzene to phenol. Catalysts 2016, 6(5), DOI 10.3390/catal6050074. (34) Zhu, S.; Song, Y.; Zhao, X.; Shao, J.; Zhang, J.; Yang, B. The photoluminescence mechanism in carbon dots (graphene quantum dots, carbon nanodots, and polymer dots): current state and future perspective. Nano Res. 2015, 8(2), 355-381. (35) Sahu, S.; Behera, B.; Maiti, T. K.; Mohapatra, S. Simple one-step synthesis of highly luminescent carbon dots from orange juice: application as excellent bio-imaging agents. Chem. Commun. 2012, 48, 8835-8837. (36) Xu, Y.; Wu, M.; Feng, X. Z.; Yin, X. B.; He, X. W.; Zhang, Y. K. Reduced carbon dots versus oxidized carbon dots: photo- and electrochemiluminescence investigations. Chem. Eur. J. 2013, 19(20), 6282-6288. (37) Loh, K. P.; Bao, Q.; Eda, G.; Chhowalla, M. Graphene oxide as a chemically tunable platform for optical applications. Nat. Chem. 2010, 2, 1015-1024. (38) López, V.; Sundaram, R. S.; Gómez-Navarro, C.; Olea, D.; Burghard, M.; GómezHerrero, J.; Zamora, F.; Kern, K. Chemical vapor deposition repair of graphene oxide: a route to highly-conductive graphene monolayers. Adv. Mater. 2009, 21(46), 4683-4686. (39) Khan, M.; Al-Marri, A. H.; Khan, M.; Mohri, N.; Adil, S. F.; Al-Warthan, A.; Siddiqui, M. R. H.; Alkhathlan, H. Z; Berger, R.; Tremel, W; Tahir, M. N. Pulicaria glutinosa plant extract: A green and eco-friendly reducing agent for the preparation of highly reduced graphene oxide. RSC Adv. 2014, 4, 24119-24125.

ACS Paragon Plus Environment

26

Page 27 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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

(40) Wong, C. P. P.; Lai, C. W.; Lee, K. M.; Hamid, S. B. A. Advanced chemical reduction of reduced graphene oxide and its photocatalytic activity in degrading reactive black 5. Materials 2015, 8(10), 7118-7128. (41) Colmenares, J. C.; Luque, R. Heterogeneous photocatalytic nanomaterials: prospects and challenges in selective transformations of biomass-derived compounds. Chem Soc Rev. 2014, 43(3), 765-78. (42) Trandafilović, L. V.; Jovanović, D. J.; Zhang X.; Ptasińska, S.; Dramićanina, M. D. Enhanced photocatalytic degradation of methylene blue and methyl orange by ZnO:Eu nanoparticles. Appl. Catal., B 2017, 203, 740-752. (43) Tang, J.; Zou, Z.; Yin, J.; Ye, J. Photocatalytic degradation of methylene blue on CaIn2O4 under visible light irradiation. Chemical Physics Letters 2003, 382(1-2), 175-179. (44) Zha, R.; Nadimicherla. R.; Guo, X. Ultraviolet photocatalytic degradation of methyl orange by nanostructured TiO2/ZnO Heterojunctions. J. Mater. Chem. A 2015, 3, 6565-6574. (45) Huo, R.; Kuang, Y.; Zhao, Z.; Zhang, F.; Xu, S. Enhanced photocatalytic performances of hierarchical ZnO/ZnAl2O4 microsphere derived from layered double hydroxide precursor spraydried microsphere. J. Colloid Interface Sci. 2013, 407, 17-21. (46) Hameed, A.; Gombac, V.; Montini, T.; Graziani, M.; Fornasiero, P. Synthesis, characterization and photocatalytic activity of NiO-Bi2O3 nanocomposites. Chem. Phys. Lett. 2009, 472, 212-216.

ACS Paragon Plus Environment

27

ACS Sustainable Chemistry & Engineering

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

Page 28 of 39

Figure 1. (a) UV-visible spectra of (i) CD, (ii) M. ferrea Linn. reduced CD, (iii) tea reduced CD, (iv) S. aromaticum reduced CD and (v) C. esculenta reduced CD; (b) FTIR spectra of (i) RCD and (ii) sodium borohydride reduced CD (used for comparison purposes); (c) XRD patterns of (i) CD and (ii) RCD; (d) Raman spectrum of RCD; and EDX map of RCD.

ACS Paragon Plus Environment

28

Page 29 of 39

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

ACS Sustainable Chemistry & Engineering

Figure 2. (a) HRTEM image of RCD at 2 nm magnification; (b) particle distribution of RCD at 100 nm magnification with its SAED pattern as inset; FFT images of RCD phase (c) before masking, (d) after masking; and (e) IFFT of RCD phase showing the lattice fringes.

ACS Paragon Plus Environment

29

ACS Sustainable Chemistry & Engineering

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

Page 30 of 39

Figure 3. PL spectra with variation of excitation wavelength 300-400 nm of (a) CD and (b) RCD; and PL spectra with variation of concentration of (c) CD and (d) RCD.

ACS Paragon Plus Environment

30

Page 31 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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 4. (αhν)1/2 vs hν plots of (a) CD and (b) RCD; and (c) I-V curves for (i) RCD and (ii) CD.

ACS Paragon Plus Environment

31

ACS Sustainable Chemistry & Engineering

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

Page 32 of 39

Figure 5. TGA thermograms of (i) RCD and (ii) CD.

ACS Paragon Plus Environment

32

Page 33 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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. Plots of UV absorbance against wavelength at different times for the degradation of MB in the presence of (a) RCD and (b) CD; plots of UV absorbance against wavelength for the degradation of MO in the presence of (c) RCD and (d) CD; and plots of UV absorbance against wavelength for the degradation of MB/MO mixture in the presence of (e) RCD and (f) CD; and (g) images showing the decolorization of MB by films of (i) silica gel, (ii) silica gel with RCD and (iii) silica gel with CD.

ACS Paragon Plus Environment

33

ACS Sustainable Chemistry & Engineering

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

Page 34 of 39

Figure 7. Plots of UV absorbance against wavelength at different times for the degradation of BPA in the presence of (a) RCD and (b) CD; (c) degradation curves of aqueous solution of BPA; and (d) fitting degradation kinetic curves for BPA by CD and RCD.

ACS Paragon Plus Environment

34

Page 35 of 39

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

ACS Sustainable Chemistry & Engineering

Figure 8. Degradation curves of aqueous solutions of (a) MB, (b) MO and (c) MB/MO mixture; and fitting degradation kinetic curves for (d) MB, (e) MO and (f) MB/MO mixture using RCD and CD.

ACS Paragon Plus Environment

35

ACS Sustainable Chemistry & Engineering

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

Page 36 of 39

Scheme 1 Synthesis of RCD by C. esculenta leaf extract in presence of Fe3+ ions

ACS Paragon Plus Environment

36

Page 37 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 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 Plausible electron transfer mechanism for the reduction of CD

ACS Paragon Plus Environment

37

ACS Sustainable Chemistry & Engineering

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

Page 38 of 39

Scheme 3 Proposed photocatalytic mechanism for degradation of organic pollutants by RCD and CD

ACS Paragon Plus Environment

38

Page 39 of 39

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

ACS Sustainable Chemistry & Engineering

For Table of Contents Use Only

Synopsis Facile and ultrafast green approach to synthesize sustainable-resource based luminescent reduced carbon nanodot for removal of organic pollutants.

ACS Paragon Plus Environment

39