Design of Novel Ytterbium Molybdate Nano-flakes Anchored Carbon

PTL under visible light irradiation. Furthermore ... For remediation progress, heterogeneous photocatalysis is a simple, green and cost- effective tec...
0 downloads 0 Views 4MB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

Article

Design of Novel Ytterbium Molybdate Nano-flakes Anchored Carbon Nanofibers: A Challenging Sustainable Catalyst for the Detection and Degradation of Assassination Weapon (Paraoxon-Ethyl) Raj Karthik, Jeyaraj Vinoth Kumar, Shen-Ming Chen, Thangavelu Kokulnathan, Han-Yu Yang, and Velluchamy Muthuraj ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b00936 • Publication Date (Web): 29 May 2018 Downloaded from http://pubs.acs.org on May 29, 2018

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

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

Page 1 of 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

Design of Novel Ytterbium Molybdate Nano-flakes Anchored Carbon Nanofibers: A Challenging Sustainable Catalyst for

the Detection and Degradation of

Assassination Weapon (Paraoxon-Ethyl) Raj Karthika, Jeyaraj Vinoth Kumarb, Shen-Ming Chena*, Thangavelu Kokulnathana, Han-Yu Yanga, Velluchamy Muthurajb a

Electroanalysis and Bioelectrochemistry Lab, Department of Chemical Engineering and

Biotechnology, National Taipei University of Technology, No. 1, Section 3, Chung-Hsiao East Road, Taipei 106, Taiwan, ROC b

Department of Chemistry, VHNSN College, 3/151-1, College Road, Virudhunagar, Tamil Nadu

626001, India

Corresponding Author *E-mail: [email protected] (S-M Chen), Phone: +886-2270-17147. Fax: +886-2270-25238.

KEYWORDS: Chemical weapon, Organophosphate, Paraoxon-ethyl, Detection, Degradation.

1 ACS Paragon Plus Environment

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

ABSTRACT Design of resourceful and sustainable catalyst for the trace level identification as well as detoxification of toxic pollutants into the environment is major concern to the researchers. In view of that, we developed novel flake-like ytterbium molybdate (YbMoO4; YbM) anchored on carbon nanofibers (YbM/f-CNFs) composite via simple wet-chemical route followed by sonication process. The physicochemical properties of as-prepared YbM/f-CNFs were carried out by several spectroscopic techniques. The YbM/f-CNFs composites exhibited excellent electrocatalyst as well as photocatalyst for the detection and detoxification of chemical warfare agent paraoxon-ethyl (PTL). Interestingly, the electrochemical results illustrated that the YbM/fCNFs nanocomposite exhibited an excellent electrocatalytic activity in terms of enhanced cathodic peak current and lower peak potential when compared with other modified electrodes. Furthermore, the YbM/f-CNFs modified electrode showed more extended linear response ranges (0.01-12 and 14-406 µM), lower detection limit (2 nM), good sensitivity (2.8 µAµM−1cm−2), and excellent selectivity for the PTL sensing. Besides, the YbM/f-CNFs catalyst had good recovery to PTL in soil and water sample analysis. In addition, the YbM/f-CNFs composite possesses remarkable photocatalytic activity and stability towards the degradation and mineralization of PTL under visible light irradiation. Furthermore, a possible detection and degradation mechanism was proposed towards PTL. This study provides a novel idea for the design of proficient and stable bifunctional catalyst for the real-time identification and remediation of lethal pollutants.

2 ACS Paragon Plus Environment

Page 2 of 39

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

INTRODUCTION Nowadays, the invention of diseases and demises increased everyday due to the longterm occurrence of pesticides and insecticides in the food, fruits, agri-food, aquatic and soil environment. Among them, organophosphate compounds are widely used as a chemical weapon during the world war and also utilized as pesticides to control the pests/insects in the agricultural and household areas1. In particular, paraoxon-ethyl (PTL) is widely utilized organophosphate pesticide in the cultivation of cotton, rice, garlic, beans, wheat and fruits2. As a result, a large quantity of PTL can enter into the air, soil and aquatic origins, and finally migrates into the food chain. Moreover, PTL can easily penetrate with human body via inhalation to create dangerous health risks including tremor, diarrhea, vomiting, poor visualization, headaches, lung-edema and unconsciousness. Furthermore, it has great capability to restrict the acetylcholinesterase enzyme and leads to damage of the central nervous system3, 4. In addition, PTL has nearly 70% similar chemical properties of nerve agent sarin and, therefore, U.S Environment Protection Agency (EPA) stated that PTL is extremely toxic substance and produces chronic neurological diseases to the both animal and human beings. Furthermore, PTL can persist in the environment prolonged time owing to its high resistance to self-generated hydrolysis nature5-7. For these reasons, the trace level identification and complete detoxification of PTL from the food, water and soil samples is a challenging task to the researchers. Until now, different analytical methods including liquid chromatography, gas chromatography, fluorescence and voltammetric techniques have been successfully developed and utilized for the accurate detection of PTL. Among them, electrochemical (voltammetric) techniques could offer many advantages such as simplicity, low-cost, rapid response, high selectivity and good sensitivity, when compared to the other analytical methods8. Recently, many researchers have been developed electrochemical biosensor for the selective detection of PTL by using enzyme immobilized on the electrode surface. However, the short-term stability of enzyme suffers its practical environment applications9. To overcome these problems, is an essential to develop a non-enzymatic electrochemical sensor for the detection of PTL in the practical applications. In addition to the detection, detoxification of PTL is another primary concern to protect living things and ecosystem. For remediation progress, heterogeneous photocatalysis is a simple, green and costeffective technique, which converts toxic pollutants into non-toxic compounds like CO2, H2O,

3 ACS Paragon Plus Environment

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

H3PO4, etc.,10, 11. Therefore, the discovery of novel and efficient nanomaterials for both detection and detoxification of hazardous PTL is reasonably crucial. In up-to-date chemistry and material science, the physical and chemical properties of functional materials containing either inorganic/organic hybrids or inorganic compounds are fundamentally associated with their dimensionality, shape and size12-15. Hence, the controlled synthesis of well-defined shape, small size and good dimensionality of the material has become a significant research issue in modern year16. Until now, many research endeavors have concentrated on the different ways to control the dimensionality, shape and size of nanomaterials. Hitherto, a wide variety of inorganic nanomaterials such as tungstates, hydrates, molybdates, metals, metal oxides and borates have been successfully developed with controlled shape and size using various methodologies, and the materials were applied to various important applications including catalysis, energy storage devices, sensors, photocatalysis and water treatments due to their extraordinary physicochemical properties17-23. Though, the simple and sensible synthetic method for the controlled synthesis of nanostructured material through chemical self-assembly is quiet drastic and hot research topic. Among the various nanomaterials, metal molybdates are essential inorganic materials that have been widely investigated in the field of catalysis24, photoluminescence25, humidity sensor26, lasers27, illumination28, heavy metal disposal29 and phytoremediation30. However, scheelite-type (ABO4) compounds that have been widely utilized in scintillation crystal, photoluminescence, Raman lasers, laser-host materials, ion conductors, catalyst, medical fields, energy storage devices and numerous technological applications because of their excellent optical, electrical, luminescence properties and large Xray absorption nature31-42. Especially, the scheelite-type molybdates (AMoO4, A= divalent cation containing element) have been broadly investigated owing to their superior chemical and thermal stability, excellent strength and high decomposition temperature which makes them to the wide variety of applications including scintillators in medical devices, sensors, solid state lasers, fiber optic communications, anode material for the Li-ion batteries and photocatalysis43-49. However, ytterbium molybdate (YbMoO4; YbM) with scheelite-type structure has received slight consideration. Recently, Wu et al., reported that the co-precipitation synthesis of 0–100 mol % Er3+ doped YbMoO4 for the investigation on temperature quenching and sensing properties50. Followed by, Volkov et al., described the growth, structure and evaluation of laser properties based on lithium doped Yb(MoO4)2 single crystal51. Until now, there are no reports available on 4 ACS Paragon Plus Environment

Page 4 of 39

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

the electrocatalytic and photocatalytic studies of YbM and YbM based composite due to the difficult to synthesize scheelite-type phase. Therefore, the development of simple and efficient technique to synthesize the size and shape controlled scheelite-type YbM and its composite is quite essential. On the other hand, carbonaceous materials including carbon nanofibers (CNFs), mesoporous carbon, carbon black, graphene and carbon nanotubes (CNTs) have drawn considerable attention because of their exclusive physical properties and wide range of potential applications52-59. Among them, CNFs can be considered as an excellent supporting matrix for the incorporation of nanomaterials on its surface due to their high conductivity, high mechanical strength, chemical stability, adsorption capacity, extremely high surface area and porosity. Therefore, CNFs are potentially applied to the various applications such as sensors, photocatalyst, low-temperature fuel cells, secondary batteries, supercapacitors and solar cells6062

. However, the pristine CNFs (insoluble in water) have inert surfaces and, so, a pre-treatment is

required to succeed desirable interaction with metals/nanomaterials on the CNFs surface. Hence, the functionalization of CNFs (ƒ-CNFs) surface is a major concern to improve the hydrophilicity and active surface area. Moreover, the optical, electrical and mechanical properties of ƒ-CNFs are quite different from the pristine CNFs. Therefore, we have chosen ƒ-CNFs as a supporting matrix for the flakes-like YbM to boost the electro- and photocatalytic activities. In this strategy, we sought an attempt for the fabrication of novel flake-like YbM anchored ƒ-CNFs composite through a simple wet-chemical route followed by sonication process and their physicochemical properties were characterized by various spectroscopic techniques. To the best of our knowledge, there are no reports available on the design of flakelike YbM anchored ƒ-CNFs and their potential dual-functional catalytic applications for the detection and detoxification of PTL. Interestingly, the YbM/f-CNFs composite modified glassy carbon electrodes (GCE) demonstrate superior electrocatalytic performances towards PTL detection with high sensitivity and excellent selectivity. The YbM/f-CNFs composite displays excellent photodegradation and mineralization performances for the detoxification of PTL with good stability. The electrocatalytic and photocatalytic mechanisms were discussed in detail. EXPERIMENTAL SECTION Materials and Apparatus. Ytterbium (III) nitrate pentahydrate (Yb(NO3)3.5H2O), carbon nanofibers (CNFs, D × L = 100 nm × 20–200 µm, average pore volume = 0.075 cm3/g), sodium molybdate (Na2MoO4), paraoxon-ethyl (O2NC6H4OP(O)(OC2H5)2) and all other chemicals were 5 ACS Paragon Plus Environment

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

purchased from Sigma-Aldrich and used without further purification. The phosphate buffer solution (PB solution, 0.05 M) was prepared by mixing of monobasic sodium phosphate (NaH2PO4) and sodium hydrogen phosphate (Na2HPO4). All the reagents and solvents were of analytical grade and used without further purification. All the required solutions were prepared by using double-distilled (DD) water. The powder X-ray diffraction analysis was investigated in a XRD, XPERT-PRO spectrometer (PANalytical B.V., The Netherlands) with Cu-Kα radiation (λ = 1.5406 Å). A Raman spectrum was measured using an NT-MDT, NTEGRA SPECTRA spectrometer. The scanning electron microscopy (SEM) and Energy dispersive X-ray (EDS) results were observed using Hitachi S-3000 H microscope (SEM Tech Solutions, USA) attached with HORIBA EMAX X-ACT. The TEM images were examined on a Shimadzu JEM-1200 EX with an accelerating voltage of 100 kV. The X-ray photoelectron spectroscopy (XPS) results were collected from Thermo ESCALAB 250 instrument. The Micromeritics, ASAP 2020M instrument was used to determine the specific surface area and pore size distribution of the material. All the electrochemical measurements were performed (Cyclic voltammetry (CV) and amperometric (it) using CHI 405a electrochemical workstation (CH Instruments Company, made in U.S.A) with a conventional three electrode cell system comprised of an glassy carbon electrode and rotating disc glassy carbon electrode (GCE and RDGCE) as working electrode (working area = 0.07 cm2 (GCE); 0.2 cm2 (RDGCE)), platinum wire as auxiliary electrode and Ag/AgCl (saturated KCl) as reference electrode, respectively. Synthesis of Flake-Like Ytterbium Molybdate. The synthesis procedure of YbM was followed by that described in our previous article with slight modification10. In brief, each 0.4 M of Na2MoO4 and 0.2 M Yb(NO3)3 were dissolved in 70 mL of DD water separately and the homogeneous solution was mixed together in a 250 mL beaker and allowed it for continuous stirring. Afterwards, 0.5 g (in 10 mL H2O) of urea was added into the above suspension and stirred for an hour. The obtained products were washed with copious amount of water/ethanol and dried at 80 oC for overnight. Finally, the collected products were calcined at 650 oC for 4 h in air. Functionalization of CNFs. The surface functionalization process of CNFs was followed by route proposed in the previously reported article with slight modification63. In a typical experiment, 1 g of CNFs were added to the freshly prepared 50 mL HNO3/H2SO4 (1:3 ratio) acid 6 ACS Paragon Plus Environment

Page 6 of 39

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

mixture and it was allowed to reflux for 10 h at 50 oC. After that, the resultant ƒ-CNFs was centrifuged and washed with copious amount of DD water until the pH value reach 7.0. Finally, the obtained ƒ-CNFs were dried at 60 oC for overnight and used for composite preparation. Synthesis of YbM/ƒ-CNFs Nanocomposite. In a typical recipe, 20 mg of ƒ-CNFs and 50 mg of flakes-like YbM were redispersed in 100 mL DD water and ultrasonicated for 120 min. Finally, the collected products were dried at 60 oC for 12 h and it denoted as YbM/ƒ-CNFs nanocomposite. The overall synthesis procedure and the applications of YbM/ƒ-CNFs nanocomposite are shown in Scheme 1.

Scheme 1 The overall synthesis procedure and the applications of YbM/ƒ-CNFs nanocomposite. Fabrication of YbM/ƒ-CNFs/GCE Structure. Prior to GCE surface modification, the GCE was well polished with 0.05 µm alumina slurry and washed with DD water and ethanol to remove the alumina particles from the GCE surface. The as-prepared YbM/ƒ-CNFs nanocomposite was redispersed in DD water at a concentration of 5 mg/mL and then sonicated for 30 min to get homogeneous

suspension.

About

8

µL

(optimized

concentration)

of

YbM/ƒ-CNFs

nanocomposite suspension was drop coated on the mirror polished GCE surface and it was allowed to dry at room temperature. After that, the dried YbM/ƒ-CNFs modified GCE was gently washed with DD water to remove the loosely attached composite molecules on the GCE surface. The obtained YbM/ƒ-CNFs/GCE was used for further electrochemical measurements. Photocatalytic Activity. The photocatalytic performances of the as-prepared YbM and YbM/ƒCNFs nanocomposite were determined by the degradation of PTL aqueous suspension under visible light irradiation. In a typical recipe, 0.5g catalyst was dispersed in 100 L of PTL solution

7 ACS Paragon Plus Environment

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

Page 8 of 39

(10 mg/L) and stirred for 1 h in dark condition to ensure the adsorption-desorption equilibrium of the reactants, and the initial concentration of PTL was noted. Afterwards, the above suspension was placed in a photoreactor (Heber Company) for photodegradation experiments. A 300 W tungsten incandescent lamp was applied as a visible light source. At every 10 min, 5 mL of the aliquot was collected and centrifuged to remove the catalyst particles. Then, the concentration changes of PTL were observed by Jasco V-770 UV-Visible spectrophotometer. The degradation percentage of PTL was determined by the following equation, D(%) = (C0-C/C0) x 100

(1)

Whereas, D is the photodegradation percentage, C0 and C represent the intensity values of the PTL before visible light irradiation (t =0) and at selected time interval, respectively.

Fig.1 (A) XRD and (B) Raman spectra of the as-synthesized YbM/ƒ-CNFs nanocomposite. RESULTS AND DISCUSSION Characterization of YbM/ƒ-CNFs Nanocomposite. The crystallographic properties of the asprepared YbM, ƒ-CNFs and YbM/ƒ-CNFs were determined by XRD analysis, as shown in Fig.1A. The distinctive sharp peaks (curve a) at 2θ angles equal to 19.16, 29.63, 32.9, 35.13, 48.88, 50.3, 56.62, 59.5 and 61.5˚ were attributed to the (101), (112), (004), (200), (204), (220), (116), (132) and (224) planes of tetragonal YbMoO4, respectively, with the space group of l41/a (88) [JCPDS No. 35-1471]50. From the curve (b), the broad intense peak at 2θ = 25˚ is related to the (002) plane of ƒ-CNFs64. The XRD pattern of YbM/ƒ-CNFs (curve c) portrayed the appearance of both YbM and ƒ-CNFs diffraction peaks without any other extra peaks. The results suggested the successful fabrication of YbM/ƒ-CNFs nanocomposite. Furthermore, Raman spectroscopy is a valuable tool to probe the structural information about the carbon based 8 ACS Paragon Plus Environment

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

composite materials. In Fig.1B (curve a), the pristine YbMoO4 portrayed the low intensity peaks in the ranges of 400-700 cm-1 were ascribed to the MoO4 tetrahedrons and the high intense peaks (700-1000 cm-1) were related to the stretched MoOn polyhedra vibrations which was aroused from the appearance of several distorted and independent MoO4 tetrahedrons65-67. In brief, the Raman modes at 458 and 547 cm-1 were ascribed to the Mo=O bending modes and Mo-O-Mo symmetric stretching modes, respectively. The high intense Raman peak centered at 795 cm-1 was correspond to the asymmetric stretching vibrations of MoO42- tetrahedrons in YbMoO468. The Raman spectrum of ƒ-CNFs in curve (b) displays the two peaks at 1343 and 1584 cm-1, attributed to the D band disordered structure of sp3 hybridized carbon and G band ordered sp2 hybridized carbon, respectively69. As similar to the XRD pattern, both the YbM and ƒ-CNFs Raman modes were presented in the YbM/ƒ-CNFs nanocomposite (curve c), which undoubtedly confirmed the embedment of YbM on the ƒ-CNFs surfaces.

9 ACS Paragon Plus Environment

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

Fig.2 SEM micrographs of the (A) pristine YbM, (B) ƒ-CNFs, (C) YbM/ƒ-CNFs; TEM images of (D) pristine YbM, (E) ƒ-CNFs, (F) YbM/ƒ-CNFs; EDS spectra of (G) pristine YbM, (H) ƒCNFs and (I) YbM/ƒ-CNFs nanocomposite.

Fig.3 (A) XPS survey of YbM/ƒ-CNFs nanocomposite; High-resolution XPS spectra of (B) Yb 4d, (C) C 1s, (D) Mo 3d and (E) O1s. The surface microstructures of as-prepared YbM, ƒ-CNFs and YbM/ƒ-CNFs nanocomposite were evaluated by SEM and TEM analysis, and the results can be seen in Fig.2. The SEM micrographs in Fig.2A displays the large amount of narrow sized aggregated flake-like structure of YbM70,71 and the corresponding EDS spectrum (Fig.2G) indicates the presence of Yb, Mo and O elements. From the Fig.2B, it is evident that the pristine ƒ-CNFs demonstrates that the randomly attached fiber-like microstructure and the relevant EDS (Fig.2H) spectrum shows the presence of C and O elements, which confirms the successful functionalization of CNFs. The SEM image of YbM/ƒ-CNFs shown in Fig.2C confirms that the YbM flake-like structures randomly covered the CNFs surface. Furthermore, the TEM studies could provide exhaustive structural information about the as-prepared materials. The TEM image of YbM, as displayed in Fig.2D, comprised еруirregular flakes with the diameter of ~70 nm, whereas f-CNFs in Fig.2E portrays network like structure of nanofibers with relatively smooth and porous 10 ACS Paragon Plus Environment

Page 10 of 39

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

surface. As for YbM/ƒ-CNFs in Fig.2F unambiguously revealed that the YbM still maintained its flake-like structure and uniformly occupied on the carbon nanofiber matrix, which leads to the establishment of interlinked network between YbM and ƒ-CNFs. The EDS spectrum in Fig.2I revealed the presence of Yb, Mo, O and C in the YbM/ƒ-CNFs nanocomposite without discernible components of impurity elements. In addition, the surface elemental compositions and their accurate oxidation states were gathered by using XPS measurements. The survey XPS spectrum of YbM/f-CNFs depicts the signals of C 1s, Yb 4d, Mo 3d and O 1s elements (Fig.3A), which perfectly matched with the EDS report. The enlarged view of Yb 4d XPS spectrum displays (Fig.3B) the doublet peaks at the binding energies of 182.5 and 191.31 eV attributed to the Yb 4d5/2 and Yb 4d3/2 spin-orbit components, which is characteristic to the divalent state of Yb in the YbM/f-CNFs composite72. From the Fig.3C, the high-resolution C 1s spectrum showed the high intense peak at 286.8 eV which was indexed to epoxy or hydroxyl carbon (C-O) and the Gaussian-fittings at the binding energies of 284.8 and 287.8 eV correspondent to amorphous carbon and carboxyl carbon (OC=O) species, respectively73. The high-resolution Mo 3d (Fig.3D) core-level XPS spectrum depicts the peaks at 232.3 and 235.6 eV, which were attributed to the Mo 3d5/2 and Mo 3d3/2 of Mo6+ state74-77. The O 1s band (Fig.3E) is a superposition of three fitted peaks with the binding energies 530.6, 531.2 and 532.1 eV, and the components were assigned to lattice oxygen, chemisorbed oxygen and adsorbed water molecules on the surface of YbM/f-CNFs nanocomposite, respectively78. The electrocatalytic and photocatalytic performances of a catalyst are strongly dependent on specific surface area and pore-size volume. Therefore, the textural assessment of as-prepared YbM and YbM/f-CNFs composite was evaluated using N2 adsorption/desorption isotherm analysis. In Fig.4, the B.E.T isotherm linear plot and the inset demonstrate their corresponding pore-size distribution value (which was gained by Barrett-Joyner-Halenda (BJH) pattern) of YbM and YbM/f-CNFs nanocomposite. As seen in Fig.4 (A, B), the determined specific surface area of YbM/f-CNFs was 47m2g-1 which is higher than that of pristine YbM (21 m2g-1). Moreover, the pore-size distribution of YbM and YbM/f-CNFs (inset Fig.4 (A, B)) showed mesoporous ranges of 24 and 37 nm, respectively. Thus, the specific surface area and pore volume were increased with the introduction of f-CNFs, which helps to generate the large number of reactive sites for the electrocatalytic and photocatalytic systems. DRS UV-Visible 11 ACS Paragon Plus Environment

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

spectroscopy is an essential tool to identify the suitable kind of light sources (UV or visible) which need for the YbM and YbM/f-CNFs nanocomposite. As found in the Fig.4C, the YbM/fCNFs nanocomposite demonstrated a red shift in the visible light region, when compared to pristine YbM. The appearance of red shift in YbM/f-CNFs nanocomposite might be owing to the powerful electronic interfacial contact between the YbM and f-CNFs, which greatly induces the diffusion of intense photons. The optical band gap was determined by Tauc’s formula11 and the obtained results are shown in Fig 4D. The estimated band gaps of YbM and YbM/f-CNFs are 2.90 eV and 2.83 eV, respectively.

Fig.4 N2 adsorption/desorption isotherms of (A) YbM, (B) YbM/ƒ-CNFs, (C) DRS UV-Vis diffuse reflectance spectra and (D) Energy gap spectra of YbM and YbM/ƒ-CNFs nanocomposite. 12 ACS Paragon Plus Environment

Page 12 of 39

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

Fig.5 The CVs response of (a) bare GCE, (b) CNFs/GCE, (c) YbM/GCE (d) ƒ-CNFs/GCE and (e) YbM/ƒ-CNFs/GCE in 5mM [Fe(CN)6]3-/4- containing 0.1M KCl at scan rate of 50 mVs-1 (A). The CVs response of YbM/ƒ-CNFs/GCE in 5mM [Fe(CN)6]3-/4- containing 0.1M KCl by varying scan rate from 10 to 100 mV s-1 (B) and their corresponding linear calibration plot (C). Different catalyst amount of YbM/ƒ-CNFs loading (4 to 10 µL) on the GCE surface (D).

Electrocatalytic Activity of YbM/ƒ-CNFs Nanocomposite. To evaluate the electrochemical properties of different modified electrodes and the synergistic effects between flakes-like YbM and ƒCNFs on the GCE surface, the CV were implemented, as it can be seen in Fig.5A. From the Fig.5A, a pair of broad redox couple was clearly observed as related to the electrochemical response of [Fe(CN)6]3−/4− with the bare GCE (a). Later, CNFs (b), YbM (c), ƒ-CNFs (d) and YbM/ƒ-CNFs were modified on the bare GCE surface, and the redox peak current response 13 ACS Paragon Plus Environment

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

Page 14 of 39

became greater with a smaller peak-to-peak separation (∆EP) such as 0.356 V (bare GCE), 0.117 V (CNFs), 0.128 V (YbM), 0.139 V (ƒ-CNFs) and 0.111 V (YbM/ƒ-CNFs nanocomposite), respectively. This phenomenon might be a distinct evidence of the fact that YbM can significantly enhance the electron transfer when composite with ƒ-CNFs. Moreover, the prepared flake-like YbM/ƒ-CNFs exhibited fast electron transfer (lower peak-to-peak separation) and higher redox peak current, which clearly suggest the unique electronic structures and electrochemical properties of flake-like YbM/ƒ-CNFs that enhanced and improved electrochemical activity by higher electroactive area when compared to other modified and unmodified electrodes. Furthermore, the electroactive surface area of bare GCE, CNFs/GCE, YbM/GCE, ƒCNFs/GCE and YbM/ƒ-CNFs nanocomposite was estimated by using CV results. Fig.5B shows the CVs response of flake-like YbM/ƒ-CNFs modified GCE containing 5 mM [Fe(CN)6]3−/4− in 0.1 M KCl with various potential scan rate from 10 to 100 mV/s. The electroactive surface area (A) was calculated using Randles–Sevick equation as follows79. Ip=2.69 x 105ACn3/2D1/2v1/2

(2)

where, Ip is the obtained cathodic and anodic peak current, v is the potential scan rate, D is the diffusion coefficient of Fe(CN)63-/4- (7.6 × 10-6 cm2 s-1), n is the number of involved electrons and C is the concentration of Fe(CN)63-/4- (5 mM). The linear plot for square root of scan rate (v1/2) vs. peak current is shown in Fig.5C. The obtained electroactive surface of YbM/ƒ-CNFs/GCE was 0.112 cm2, which was larger than that of other modified electrodes such as bare GCE (0.034 cm2), CNFs/GCE (0.049 cm2), YbM/GCE (0.082 cm2), and ƒ-CNFs/GCE (0.091 cm2). Optimization of Amount of the Modifier. The electrochemical response and sensitivity of the modified electrode for the detection of PTL were affected by the loading of catalyst concentration on the electrode surface. Hence, the optimization of the loading level of YbM/ƒCNFs on the GCE surface has a major role on the voltammetric responses. The different amounts of YbM/ƒ-CNFs suspension such as 4, 6, 8 and 10 µL were drop casted on the GCE surface and their electrochemical responses (calibration plot) of PTL are shown in Fig.5D. It observed that the reduction peak current increases gradually while increasing the loading amount from 4 to 8 µL of YbM/ƒ-CNFs suspension on the GCE surface. On further increasing the concentration (above 8 µL), the reduction peak current of PTL decreased, suggesting that the limited amount of YbM/ƒ-CNFs suspension on the GCE surface can effectively detect the PTL. Notably, the higher

14 ACS Paragon Plus Environment

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

composite suspension amount on the GCE surface may block (agglomeration of composite) the active area on the electrode surface and, thus, the peak current and sensitivity of the electrode

15 ACS Paragon Plus Environment

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

Fig.6 The CVs of (a) bare GCE, (b) flake-like YbM/GCE, (c) CNFs/GCE, (d) ƒ-CNFs/GCE and YbM/ƒ-CNFs/GCE in the absence (A; a-e) and presence (B; f-j) of 400 µM PTL containing 0.05 M PB solution at a scan rate of 50 mVs-1. (C) The CV curves of YbM/ƒ-CNFs/GCE in 0.05 M PB solution containing 400 µM PTL for the different scan rates ranging from 10 to 100 mVs-1 and (D) the corresponding linear plot of cathodic peak current vs. scan rate. (E) The CV responses of the reduction of PTL (400 µM) at YbM/ƒ-CNFs/GCE in various pHs ranging from 3.0 to 11.0 and (F) the corresponding calibration plot of pH vs. cathodic peak current.

were decreased. In this case, the maximal reduction peak current of PTL was observed in 8 µL of YbM/ƒ-CNFs suspension. Therefore, the 8 µL of YbM/ƒ-CNFs/GCE suspension was chosen as the optimum modification for the PTL sensing. Electrochemical Behavior of the Flake-like YbM/ƒ-CNFs Nanocomposite Based PTL Sensor. The electrochemical properties of PTL on different modified electrodes were evaluated using CV. Fig.6A, B shows the electrochemical performance on bare GCE (a), flake-like YbM/GCE (b), CNFs/GCE (c), ƒ-CNFs/GCE (d), and YbM/ƒ-CNFs nanocomposite modified GC electrodes in the absence (Fig.6A= a,b,c,d,e) and presence (Fig.6B= f,g,h,i,j) of 400 µM PTL containing 0.05 M PB solution at a scan rate of 50 mVs-1. The PTL absence is obvious in Fig. 6A. There is no oxidation/reduction peak observed for the modified and unmodified electrodes, whereas a sharp, enhanced and irreversible cathodic peak current (R1) was observed in the presence of 400 µM PTL (Fig.6B (j)). The observed cathodic peak (R1) is related to the direct reduction of PTL to phenylhydroxylamine with four electrons (e-) and four protons (H+) transfer process (system I). Furthermore, two more peaks (redox couple) were observed and they were denoted as Q1 and Q2. The observed Q1 and Q2 (reversible) peaks are related to the formation of hydroxylamine to nitroso group with two electrons (e-) and two protons (H+) transfer process (system II)80. Besides, the obtained cathodic peak (R1) current of PTL is much higher when compared with Q2, which designates the PTL reduction at flakes-like YbM/ƒ-CNFs/GCE is more favored to form a phenylhydroxylamine in neutral medium or more alkaline medium. The overall electrochemical mechanism of PTL was well documented previously and it can be seen in scheme 2. Conversely, the obtained cathodic peak potential at flake-like YbM/ƒ-CNFs/GCE (0.62 V) is shifted to more positive side when compared to other modified electrodes such as bare GCE (-0.73 V), flake-like YbM/GCE (-0.67 V) and CNFs/GCE (-0.64V). In addition, the 16 ACS Paragon Plus Environment

Page 16 of 39

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

observed cathodic peak current of PTL at YbM/ƒ-CNFs/GCE (j) is 9.7 (bare GCE (f)), 8.5 (YbM/GCE (g)), 2.09 (CNFs/GCE (h)) and 2.7 (ƒ-CNFs/GCE (i)) fold higher when compared to the aforementioned modified electrodes. These results suggest that the decrease in lower potential of flake-like YbM/ƒ-CNFs nanocomposite modified GCE toward PTL is attributed to the strong interaction between the PTL and ƒ-CNFs and more active sites of flake-like YbM. The overall result demonstrates that proposed YbM/ƒ-CNFs nanocomposite acts as an excellent electron mediator between the electrolyte and electrode surface and it could offer fabulous platform for the PTL sensing by contributing surplus electroactive sites and availability of large surface area on the surfaces of YbM and ƒ-CNFs. Thus, it improves the electrocatalytic activity for the PTL detection.

Scheme 2 The overall electrochemical reduction/oxidation mechanism of PTL on the YbM/ƒCNFs nanocomposite Influence of Scan Rate. The effect of scan rate on flake-like YbM/ƒ-CNFs nanocomposite modified GCE for the detection of PTL was investigated by CV at various scan rates ranging over 10-100 mVs-1 (Fig.6C) in order to study the electron-transfer process (selected molecules; PTL) between the electrolyte solution and electrode surface. In Fig.6C, it can be clearly seen that 17 ACS Paragon Plus Environment

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

Page 18 of 39

the cathodic peak (R1) and redox peak (Q1 and Q2) currents increased linearly with increasing the scan rate, and it offered a linear relationship between scan rate and cathodic peak current (R1) of PTL (Fig.6D). The linear regression equation was obtained as follows: Ipc= -1.62 (mV) – 64.07 (R2 = 0.990) (for PTL reduction (R1))

(3)

indicated that the reduction of PTL was an typically adsorption-controlled process. Influence of pH. The electrochemical response of PTL must be affected by changing the pH solution (values) due to the contribution of protons in the overall electrode reaction. The effect of pH (pH values = 3.0, 4.0, 7.0, 9.0 and 11.0) on flake-likeYbM/ƒ-CNFs nanocomposite modified GCE was investigated in the presence of 400 µM PTL at a scan rate of 50 mVs-1 by CV. As it can be seen in Fig.6E, the cathodic peak current of PTL gradually increases with increasing the pH values from 3.0 to 7.0. When the pH value is above 7.0, the cathodic peak current of PTL was gradually decreased. The maximal cathodic peak current was observed at pH 7.0 (Fig.6F) and, therefore, the pH 7.0 was chosen as an optimal level for PTL detection. Besides, with the pH increasing from 3.0 to 11.0 in solution, the cathodic peak potential (Epc) is shifted towards more negative potential. The linear plot was obtained between the pH vs. Epc as it shown in Fig.S1. From the linear plot, the linear regression equation can be expressed as follows: Epc (V) = -0.047 pH - 0.277 (R2 = 0.971)

(4)

According to Nernstian equation: dEpc/dpH = (2.303mRT)/(nF)

(5)

where, n is the number of electron, m is the number of proton and respectively and other symbols such as R, T and F were consistent with their usual meanings. Accordingly, the proton transfer number (m) was calculated to be 2, indicating that the electrochemical reduction process of PTL on the flake-likeYbM/ƒ-CNFs/GCE is a transfer process with equal number of protons and electrons. Amperometric Determination of PTL. In Fig. 7A, the amperometric (i-t) curve is shown for the flake-like YbM/ƒ-CNFs nanocomposite modified RDGCE with consecutive additions of PTL (0.01 - 1576 µM) in 0.05 M PB solution (pH 7.0) at an applied potential of -0.62 V (vs. Ag/AgCl) under stirring (1200 rpm). It is seen that YbM/ƒ-CNFs/RDGCE exhibited tremendous amperometric current response with increasing PTL concentration. The linear relationship was obtained between the PTL concentration and current signal for YbM/ƒ-CNFs/RDGCE, as depicted in Fig.7B. As a result, two linear ranges were observed in Fig.7B. The first linear 18 ACS Paragon Plus Environment

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

response is found over the PTL concentration range of 0.01-12 µM with linear regression equation of Ip (µA) = 0.39 + 0.20, (R2 = 0.997). The second linear response is over the PTL concentration range of 14 - 406 µM with the linear regression equation of Ip (µA) = 0.13 + 8.30, (R2 = 0.985). The enlarged view of amperometric current response of PTL at low concentrations is shown in Fig. 7A (inset). The sensitivity (2.87 µAµM-1cm-2) and limit of detection (LOD = 2 nM) were calculated from the lower linear response range of PTL with the signal-to-noise ratio of three (S/N=3). A comparison of LOD, sensitivity and wide linear response ranges for YbM/ƒCNFs/RDGCE and other previously reported PTL sensor can be observed in Table 1. The obtained analytical parameters (linear range, LOD and sensitivity) for YbM/ƒ-CNFs/RDGCE are very comparable and even superior those obtained for various electrodes reported on recently.

Fig.7 (A) Amperometric (i-t) response at YbM/ƒ-CNFs/RDGCE for the addition of different concentrations (0.01 to 1576 µM) of PTL into the frequently stirred 0.05 M PB solution (pH 7.0); 19 ACS Paragon Plus Environment

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

applied potential = -0.62 V; rotation speed = 1200 rpm; inset-enlarged view of lower concentration and current response of PTL. (B) A linear plot for concentrations of PTL vs. cathodic current response. (C) Amperometric (i-t) response of the electrochemical sensor for the addition of PTL (a) malathion (b), carbofuran (c), diuron (d), chlorpyrifos (e) glucose (f), uric acid (g), ascorbic acid (h), dopamine (i) metronidazole (j), nitrofurantoin (k), fenitrothion (l), flutamide (m), 4-nitrophenol (n), methyl parathion (o) and nitrobenzene (p) into the frequently stirred 0.05 M PB solution pH 7.0. (D) Steady-state current response at YbM/ƒ-CNFs/RDGCE PTL sensor for the addition of 50 µM PTL in 0.05 M PB solution (pH 7.0) up to 1400 s.

Interference Studies. The anti-interference ability of the YbM/ƒ-CNFs/RDGC electrode was investigated for the detection of PTL in the existence of nitro-aromatic containing drugs, physiological compounds and pesticides using amperometric (i-t) technique. Fig.7C reveals the amperometric current response for PTL and other interfering compounds at YbM/ƒCNFs/RDGCE in the presence of 50 fold excess concentrations of pesticides malathion (b), carbofuran (c), diuron (d), chlorpyrifos (e); 50 fold excess of concentrations of physiological compounds glucose (f), uric acid (g), ascorbic acid (h), dopamine (i); and 20 fold excess concentrations of nitro-aromatic containing drugs metronidazole (j), nitrofurantoin (k), fenitrothion (l), flutamide (m), 4-nitrophenol (n), methyl parathion (o) and nitrobenzene (p) in a constantly stirred 0.05 M PB solution (pH 7.0). The well-defined peak current response was observed for the addition of 50 µM PTL (a) into the PB solution and further injection of 50 µM PTL in the next step with a sample interval of 50 s, the PTL current response was increased and steady state current reached within 5s. Moreover, there is no considerable peak current response observed for the presence of pesticides, physiological compounds and nitro-aromatic containing drugs (b-p) in the same solution. However, addition of 50 µM PTL induces the same current response (curve (a)) even in the presence of aforementioned co-interfering compounds. These results clearly suggest that the determination of PTL is possible even in the presence of 50, 50, and 20-fold excess concentrations of pesticides, physiological compounds and nitro-aromatic containing drugs, respectively. Stability, Repeatability and Reproducibility. To study the operational stability of the PTL sensor, the amperometric current response of flake-like YbM/ƒ-CNFs/RDGCE to 50 µM PTL was investigated at an applied potential of -0.62 V over a period of 1400 s (Fig.7D). The 20 ACS Paragon Plus Environment

Page 20 of 39

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

reduction current still remained about 92 % of the initial response after 1400s, indicating the good operational stability of the sensor. To demonstrate the accuracy and feasibility of the proposed method, the repeatability of YbM/ƒ-CNFs/GCE was evaluated by measuring the 50 µM PTL solution for 5 successive times with the same modified electrode. The obtained results suggested that the modified electrode holds an acceptable repeatability with relative standard deviation (RSD) of 3.43 %. The reproducibility of the sensor was investigated using three independent YbM/ƒ-CNFs modified electrodes with (50 µM PTL) RSD of 3.54 %, suggested that the proposed sensor has good reproducibility. Table 1 Comparison of the analytical parameters of PTL detection with other reports Electrode

Linear range (µM)

LOD (µM)

Sensitivity

Refs

(µAµM-1cm-2) CP/OPH

4.6-46.0

0.9

0.0014

81

GCE/MWCNTs/OPH

0.5-2.0

0.31

0.0259

82

GCE/SWCNTs/OPH

0.25-4.0

0.15

0.025

83

GCE/CB/MC/OPH

0.2-8.0

0.12

0.1980

84

CP/OPH

0.02-18

0.02

0.012

85

GCE/SWCNTs/OPH

0.5-8.5

0.01

0.002

86

SPE/PBNPs/BuChE

0.007-0.01

0.004

-

87

GCE/PPy-CNTs/AChE

-

0.003

-

88

SPE/AChE

-

1.8

-

89

0.01-12

0.002

2.87

This

GCE/YbM/ƒ-CNFs

14 - 406

work

CP Carbon paste; OPH Organophosphorus hydrolase; GCE Glassy carbon electrode; CNTs Carbon nanotubes; SWCNTs Single wall carbon nanotubes; CB-carbon black; MC Mesoporous carbon; SPE Screen printed electrodes; PBNPs Prussian blue nanoparticles; BuChE Butyrylcholinesterase; PPy-poly(pyrrole); AChE Acetylcholinesterase. Practical Application. To evaluate the practical viability of the proposed method for the PTL determination in soil and water samples, these two samples were analyzed under optimized (amperometric) conditions. The water and soil samples were collected from around Taipei, Taiwan. For the analysis, soil and water samples was directly used. However, no PTL was detected in soil and water samples. Therefore, the known concentration of PTL was added to the 21 ACS Paragon Plus Environment

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

Page 22 of 39

both samples and they were taken for the analysis using standard addition method. The obtained recoveries are displayed in Table 2 and the values were quite satisfactory and acceptable. Table 2 Results of determination of PTL in soil and water samples Sample Soil

Water

Added (µM) 0.0 10.0 20.0 0.0 10.0 20.0

Found (µM) 9.75 19.86 9.97 19.91

Recovery (%) 97.5 99.3 99.7 99.5

Photocatalytic Activity. The photocatalytic study of as-prepared flake-like YbM and YbM/ƒCNFs nanocomposite was evaluated for the degradation of as detected hazardous pesticide PTL under visible light irradiation. The UV-visible absorption spectroscopy was used to monitor the time-dependent concentration changes of PTL aqueous suspension in the presence of pristine YbM (Fig.8A) and YbM/ƒ-CNFs nanocomposite (Fig.8B) under visible light irradiation. As seen in Fig.8B, the major characteristic absorption peak of PTL at 273 nm was diminished continuously with increasing the irradiation time. After visible light exposure for 100 min, the intensity of the major peak decreases to nearly zero, which confirms the complete PTL decomposition by YbM/ƒ-CNFs nanocomposite. There were no extra peaks corresponding to the intermediates, and this indicated that YbM/ƒ-CNFs nanocomposite does not affect the photodegradation pathway of PTL. Before the photodegradation experiments, the blank (presence of YbM and YbM/ƒ-CNFs alone) and dark (presence of light alone) tests were performed for the PTL removal. As evident in Fig.8C, there was only 6%, 9% and 14% degradation observed in the presence of light alone, pristine YbM alone and YbM/ƒ-CNFs nanocomposite alone, respectively, which suggests that the catalyst/light combination is needed for the efficient removal of PTL from the environment. Moreover, the as-prepared YbM/ƒ-CNFs nanocomposite exhibited an excellent photocatalytic performance towards PTL, could degrades above 98% after 100 min visible light irradiation while pristine YbM demonstrates only 52%. The enhanced degradation efficiency of YbM/ƒ-CNFs was attributed to larger surface area and interfacial interaction between flake-like YbM and ƒ-CNFs, which helps to ameliorate the rapid electron transfer and decelerate the photogenerated charge carrier recombination.

22 ACS Paragon Plus Environment

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

Furthermore, the photodegradation kinetic rate of PTL using YbM and YbM/ƒ-CNFs nanocomposite was elucidated by pseudo-first-order kinetic equation as follows, ln C/C0 = -kt

(6)

Where, C0 denotes the initial concentration of PTL (t=0), C is the concentration of PTL at various reaction time, k is the rate constant. Fig.8D depicts the kinetic rate constant of YbM/ƒCNFs composite is to be 0.17369 min-1, which is 2.9 times higher than that of pristine YbM (k= 0.05988 min-1).

Fig.8 Time-dependent absorption spectrum of PTL in the presence of (A) YbM and (B) YbM/ƒCNFs composite under visible light irradiation. (C) Effect of different catalysts on the photodegradation of PTL. (D) Kinetic curves of PTL photodegradation using YbM and YbM/ƒCNFs nanocomposite.

23 ACS Paragon Plus Environment

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

Fig.9 (A) Effects of different amount of YbM/ƒ-CNFs dosage. (B) Effect of various scavengers on photodegradation of PTL over YbM/ƒ-CNFs. (C) TOC removal and (D) Reusability study of YbM/ƒ-CNFs on the photodegradation of PTL.

The amount of catalyst loading plays a key role in the photodegradation of organic pollutants. The optimized amount of YbM/ƒ-CNFs catalyst dosage is an important concern for the effective degradation of PTL and it also avoids the catalyst wastage. Therefore, we carried out the influence of different amount of catalyst dosage ranging from 0.1 to 0.7g on the photodegradation of PTL, whilst, other parameters are kept constant (light sources and PTL concentration). As shown in Fig.9A, the photodegradation efficiency gradually increases with increasing the catalyst amount up to 0.5g, which might be due to the propagations of sufficient level of photogenerated charge carriers needed for the PTL degradation. At higher dosage 24 ACS Paragon Plus Environment

Page 24 of 39

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

(beyond 0.5g), the catalyst leads to aggregation (particle-particle interaction) which prevents the visible light diffusion into the reaction solution. As a result, the intimate contact between the PTL solution and the photogenerated active species from the catalyst surface was diminished; therefore, the degradation efficiency was decreased. The results suggested that 0.5g catalyst is much enough for the proficient degradation of PTL. The photodegradation reaction was mainly conducted by the reactive oxidative species (ROS) such as holes (h+), superoxide radical anions (O2•−) and hydroxyl radicals (•OH)90-93. The major contributions of ROS for the PTL degradation were identified by the radical trapping experiments using ethylenediaminetetraacetic acid (EDTA), acryl amide (AA), and isopropyl alcohol as a scavengers for h+ ,O2•− and •OH, respectively. The control experiment (without scavenger) was performed while catalyst amount, initial PTL concentration and light source were kept constant. The impacts of different scavengers on the photodegradation of PTL over YbM/ƒCNFs nanocomposite are displayed in Fig.9B. From Fig.9B, the introduction of EDTA or AA into the photodegradation system provides inhibitory results of only 9 and 16% degradation of PTL, respectively, implying the minor involvement of h+ and O2•− radicals. Conversely, 76% degradation efficiency was induced by the IPA addition, suggesting that the PTL degradation was mostly persuaded by the •OH radicals. Furthermore, minor involvement of O2•− and major role of •OH radicals during the photodegradation reaction were determined by using nitro blue tetrazolium (NBT) and terepthalic acid (TA) tests, respectively94-96. Fig.S2 portrays the photoluminescence (PL) intensity peak of TAOH at 426 nm greatly increases with increasing the irradiation time, suggests the generation of large amount of •OH radicals during the reaction. Conversely, the slight decrement of NBT absorption spectrum in Fig.S3 were observed which designate that the production of minimum amount of O2•− radicals in the photodegradation of PTL. The results indicate that the •OH radicals generated on the YbM/ƒ-CNFs nanocomposite surface play a significant role in the PTL photodegradation. Based on the trapping experiment results and enthused from previous report97, we have provided a plausible degradation mechanism of PTL, as shown in scheme 3. It can be clearly seen that the strong electrophilic •OH radicals can cleave the P-O-C, P-O, C-C bonds and oxidizes the P and C atoms of PTL. Therefore, the structure of PTL saturated into paranitrophenol, diethyl phosphoric acid ester and acetaldehyde. Finally, these compounds were further decomposed into phosphoric acid, carbon dioxide and water. 25 ACS Paragon Plus Environment

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

Scheme 3 A plausible photodegradation mechanism of PTL through the •OH radicals

Furthermore, the mineralization efficiency of YbM/ƒ-CNFs nanocomposite towards the PTL degradation was determined by TOC analysis under optimized conditions. Fig.9C represents that the TOC removal efficiency increases with the irradiation time increasing. After 100 min, the TOC removal efficiency reaches about 71%, indicating the superior mineralization performance of YbM/ƒ-CNFs nanocomposite. Nevertheless, TOC removal efficiency was 28% lesser than the degradation efficiency which might belongs to the existence of recalcitrant and stable salts in the solution. For practical implementation of the catalyst, the reusability and stability are essential factors. Consequently, the stability of YbM/ƒ-CNFs nanocomposite 26 ACS Paragon Plus Environment

Page 26 of 39

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

catalyst was evaluated by repeating the PTL degradation process for six times and the results are given in Fig.9D. At each degradation process end, the catalyst was recovered by filtration and dried, and used in the next cycle. As shown in Fig.9D, the YbM/ƒ-CNFs nanocomposite catalyst still maintained 92% degradation efficiency even after five repeated cycles that evidently verifies superior stability and reusability. CONCLUSIONS To sum up, the novel and efficient flake-like YbM anchored on carbon nanofibers nanocomposite was successfully prepared by simple wet-chemical route followed by sonication process and the product was characterized in detail. The electrocatalytic and photocatalytic performances of the as-prepared YbM/f-CNFs nanocomposite were scrutinized towards the detection and degradation of environmentally toxic PTL pesticide. Interestingly, the flake-like YbM/f-CNFs nanocomposite exhibited a tremendous electrocatalytic activity as effective and excellent electron mediator for the PTL detection. Also, the flake-like YbM/f-CNFs nanocomposite modified GCE showed lower LOD, well sensitivity, wide linear response ranges and good selectivity even in the existence of potentially co-interfering compounds. Besides, YbM/f-CNFs/GCE attained acceptable recoveries to determine PTL in soil and water samples. As a photocatalyst, YbM/ƒ-CNFs nanocomposite could degrades above 98% and mineralize 71% of PTL under visible light irradiation. For those admirable bifunctional activities of flake-like YbM decorated on the ƒ-CNFs surface, it could be applied for the fabrication of electrochemical sensor and degradation of PTL to the contaminated soil and waste water treatment.

27 ACS Paragon Plus Environment

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

■ ASSOCIATED CONTENT Supporting Information The linear plot for pH vs. cathodic peak potential, PL spectrum of YbM/f-CNFs and Determination of O2•− radical using nitro blue tetrazolium (NBT)

■ AUTHOR INFORMATION Corresponding Author *E-mail: [email protected]. (S-M Chen) Phone: +886-2270-17147. Fax: +886-227025238. ORCID Shen-Ming Chen: 0000-0002-8605-643X Notes The authors declare no competing financial interest.

■ ACKNOWLEDGMENTS This project was supported by the Ministry of Science and Technology (MOST 106-2113-M027-003 and MOST 106-2811-M-027-004), Taiwan, ROC. The author's Jeyaraj Vinoth Kumar and Velluchamy Muthuraj are sincere thanks to the College managing board, Principal and Head of the Department, VHNSN College for the providing necessary research facilities.

28 ACS Paragon Plus Environment

Page 28 of 39

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

■ REFERENCES 1. Post, 1998. Organophosphate (Post note 12). Parliamentary Office of Science and Technology, London, UK 2. http://portal.anvisa.gov.br/documents/111215/117782/P03%2B%2BParationamet%25C3%25ADlica.pdf/7edd5934-0e95-44e4-bf70-596b2a884621, accessed in March 2017. 3. U.S. EPA, Summary of the Risks and Uses of Organophosphate Parathion-methyl. 2000 4. Marx, S.; Zaltsman, A.; Turyan, I.; Mandler, D. Parathion Sensor Based on Molecularly Imprinted Sol–Gel Films. Anal. Chem. 2004, 76, 120–126. 5. Shi, H.; Wang, R.; Yang, J.; Ren, H.; Liu, S.; Guo, T. Novel Imprinted Nanocapsule with Highly Enhanced Hydrolytic Activity for Organophosphorus Pesticide Degradation and Elimination. Eur. Polym. J. 2015, 72, 190-201. 6. "Parathion". Integrated Risk Information System. U. S. Environmental Protection Agency. 26 January 2007. 7. Cross, Glenn (2017). Dirty War: Rhodesia and Chemical Biological Warfare, 1975–1980. Solihull, UK: Helion & Company. ISBN 978-1-911512-12-7. 8. Stoytcheva, M.; Zlatev, R.; Montero, G.; Velkova, Z.; Gochev, V. Nanostructured Platform for the Sensitive Determination of Paraoxon by Using an Electrode Modified with a Film of Graphite-Immobilized Bismuth. Microchim. Acta 2017, 184, 2707–2714. 9. Arduini, F.; Guidonea, S.; Amine, A.; Palleschi, G.; Moscone, D. Acetylcholinesterase Biosensor Based on Self-Assembled Monolayer-Modified Gold-Screen Printed Electrodes for Organophosphorus Insecticide Detection. Sens. Actuator B-chem. 2013, 179, 201–208. 10. Kumar, J. V.; Karthik, R.; Chen, S. M.; Muthuraj, V.; Karuppiah, C. Fabrication of PotatoLike Silver Molybdate Microstructures for Photocatalytic Degradation of Chronic Toxicity Ciprofloxacin and Highly Selective Electrochemical Detection of H2O2. Sci. rep. 2016, 6, 34149, 1-13. 11. Karthik, R.; Kumar, J. V.; Chen, S. M.; Kumar. S.; Karuppiah, C.; Chen, T. W.; Muthuraj, V. Investigation on the Electrocatalytic Determination and Photocatalytic Degradation of Neurotoxicity Drug Clioquinol by Sn(Moo4)2 Nanoplates. ACS Appl. Mater. Interfaces 2017, 9, 26582–26592

29 ACS Paragon Plus Environment

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

12. Chen, A.; Peng, X.; Koczkur, K.; Miller, B. Super-hydrophobic Tin Oxide Nanoflowers. Chem. Commun. 2004, 17, 1964-1965. 13. Golovnev, N. N.; Molokeev, M. S.; Lesnikov, M. K.; Atuchin, V. V. Two Salts and the Salt Co-crystal of Ciprofloxacin with Thiobarbituric and Barbituric Acids: The Structure and Properties. J. Phys. Org. Chem. 2018, 31, 3773. 14. Lim, C. S.; Atuchin, V.V. Cyclic MAM Synthesis of SPION/BaMoO4:Er3+,Yb3+ Composite and Its Optical Properties. Proc. SPIE., 2013, 8771, 877110. 15. Dorozhkin, K. V.; Dunaevsky, G. E.; Sarkisov, S. Y.; Suslyaev, V. I.; Tolbanov, O. P.; Zhuravlev, V. A.; Sarkisov, Y. S.; Kuznetsov, V. L.; Moseenkov, S. I.; Semikolenova, N. V.; Zakharov, V. A.; Atuchin, V. V. Terahertz Dielectric Properties of Multiwalled Carbon Nanotube/Polyethylene Composites. Mater. Res. Express, 2017, 4, 106201. 16. Dong, H.; Li, X. J.; Peng, Q.; Wang, X.; Chen, J. P.; Li, Y. D. Monodisperse Magnetic Single‐Crystal Ferrite Microspheres. Angew. Chem., Int. Ed. 2005, 44, 2842-2845. 17. Ding, Y.; Yu, S. H.; Liu, C.; Zang, Z. A. 3D Architectures of Iron Molybdate: Phase Selective Synthesis, Growth Mechanism, and Magnetic Properties. Chem. Eur. J. 2007, 13, 746-753. 18. Yang, J.; Lin, C. K.; Wang, Z. L.; Lin, J. In(OH)3 and In2O3 Nanorod Bundles and Spheres: Microemulsion-Mediated Hydrothermal Synthesis and Luminescence Properties. Inorg. Chem. 2006, 45, 8973-8979. 19. Zhang, N.; Bu, W. B.; Xu, Y. P.; Jiang, D. Y.; Shi, J. L. Self-assembled Flowerlike Europium-Doped Lanthanide Molybdate Microarchitectures and Their Photoluminescence Properties. J. Phys. Chem. C 2007, 111, 5014-5019. 20. Jiang, C. L.; Zhang, W. Q.; Liu, Y. K.; Qian, Y. T. Self-assembled Copper Nanowalls into Microstructures with Different Shapes: A Facile Aqueous Approach. Cryst. Growth Des. 2006, 6, 2603-2606. 21. Bai, P.; Wu, P. P.; Yan, Z. F.; Zhou, J. L.; Zhao, X. S. Self-assembly of Clewlike ZnO Superstructures in the Presence of Copolymer. J. Phys. Chem. C 2007, 111, 9729-9733. 22. Zhang, J.; Lin, J. Vaterite-type YBO3: Eu3+ Crystals: Hydrothermal Synthesis, Morphology and Photoluminescence Properties. J. Cryst. Growth 2004, 271, 207-215.

30 ACS Paragon Plus Environment

Page 30 of 39

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

23. Wang, X.; Li, Y. Rare‐earth‐compound Nanowires, Nanotubes, and Fullerene‐like Nanoparticles: Synthesis, Characterization, and Properties. Chem. Eur. J. 2003, 9, 56275635. 24. Pan, L.; Li, L.; Chen, Y. J. Synthesis and Electrocatalytic Properties of Microsized Ag2WO4 and Nanoscaled MWO4 (M= Co, Mn). J. Sol-Gel Sci. Technol. 2013, 66, 330–336. 25. Longo, E.; Voanti, D. P.; Longo, V. M.; Gracia, L.; Nogueira, I. C.; Almeida, M. A. P.; Pinheiro, A. N.; Ferrer, M. M.; Cavalcante, L. S.; Andres, J. Toward an Understanding of the Growth of Ag Filaments on α-Ag2WO4 and Their Photoluminescent Properties: A Combined Experimental and Theoretical Study. J. Phys. Chem. C 2014, 118, 1229–1239. 26. Sundaram, R.; Nagaraja, K. S. Solid State Electrical Conductivity and Humidity Sensing Studies on Metal Molybdate–Molybdenum Trioxide Composites (M= Ni2+, Cu2+ and Pb2+). Sens. Actuator B-chem. 2004, 101, 353–360. 27. Bradler, M.; Baum, P.; Riedle, E. Femtosecond Continuum Generation in Bulk Laser Host Materials with Sub-µJ Pump Pulses. Appl. Phys. B: Lasers Opt. 2009, 97, 561–574. 28. Kumar, A.; Kumar, J. Perspective on Europium Activated Fine-Grained Metal Molybdate Phosphors for Solid State Illumination. J. Mater. Chem. 2011, 21, 3788–3795. 29. Al-Othman, Z. A.; Naushad, M. Forward (M2+− H+) and Reverse (H+− M2+) Ion Exchange Kinetics of the Heavy Metals on Polyaniline Ce (IV) Molybdate: A Simple Practical Approach for the Determination of Regeneration and Separation Capability of Ion Exchanger. Chem. Eng. J. 2011, 171, 456–463. 30. Qu, J.; Wang, L.; Yuan, X.; Cong, Q.; Guan, S. S. Effects of Ammonium Molybdate on Phytoremediation by Alfalfa Plants and (Im) Mobilization of Toxic Metals in Soils. Environ. Earth Sci. 2011, 64, 2175–2182. 31. Mikhailik, V. B.; Kraus, H. Performance of Scintillation Materials at Cryogenic Temperatures. Phys. Status Solidi B, 2010, 247, 1583-1599. 32. Longo, V. M.; Cavalcante, L. S.; Paris, E. C.; Sczancoski, J. C.; Pizani, P. S.; Li, M. S.; Andres, J.; Longo, E.; Varela, J. A. Hierarchical Assembly of CaMoO4 Nano-Octahedrons and Their Photoluminescence Properties. J. Phys. Chem. C, 2011, 115, 5207-5219. 33. Luo, Y.S.; Zhang, W.D.; Dai, X.J.; Yang, Y.; Fu, S.Y. Facile Synthesis and Luminescent Properties of Novel Flowerlike BaMoO4 Nanostructures by a Simple Hydrothermal Route. J. Phys. Chem. C, 2009, 113, 4856-4861. 31 ACS Paragon Plus Environment

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

34. Gong, Q.; Qian, X. F.; Ma, X. D.; Zhu, Z. K. Large-Scale Fabrication of Novel Hierarchical 3D CaMoO4 and SrMoO4 Mesocrystals via a Microemulsion-Mediated Route. Cryst. Growth Des., 2006, 6, 1821-1825. 35. Nikl, M.; Bohacek, P.; Mihokowa, N; Kobayashi, M.; Ishii, M.; Usuki, Y.; Babin, V.; Stolovich, A.; Zazubowich, S.; Bacci, M. Excitonic Emission of Scheelite Tungstates AWO4 (A= Pb, Ca, Ba, Sr). J. Lumin., 2000, 1136, 87-89. 36. Nikl, M.; Bohacek, P.; Mihokowa, N; Solovieva, N.; Vedda, A.; Martini, M.; Pazzi, G. P.; Fabeni, P.; Kobayashi, M.; Ishii, M. Enhanced Efficiency of PbWO4:Mo,Nb Scintillator. J. Appl. Phys. 2002, 91, 5041-5044. 37. Brenier, A.; Jia, G.; Tu, C. Raman Lasers at 1.171 and 1.517µm with Self-frequency Conversion in SrWO4: Nd3+ Crystal. J. Phys. Condens. Matter., 2004, 16, 9103. 38. Takai, S.; Sugiura, K.; Esaka, T. Ionic Conduction Properties of Pb1− xMxWO4+ δ (M= Pr, Tb). Mater. Res. Bull., 1999, 34, 193-202. 39. Hanuza, J.; Trzebiatowska, B. J.; Oganowski, W. The Oxygen Bond in Mo(V)/Mo(Vi) Active Centres and its Role in the Oxidizing Dehydrogenation Process. J.Mol.Catal., 1978, 4, 271-287. 40. Guzik, M.; Tomaszewicz, E.; Guyot, Y.; Legendziewicz, J.; Boulon, G. Structural and Spectroscopic Characterizations of New Cd1−3xNd2x-xMoO4 Scheelite-type Molybdates with Vacancies as Potential Optical Materials. J. Mater. Chem. C, 2015, 3, 4057-4069. 41. Faure, N.; Borel, C.; Couchaud, M.; Basset, G.; Templier, R.; Wyon, C. Optical Properties and Laser Performance of Neodymium Doped Scheelites CaWO4 and NaGd(WO4)2 Appl. Phys. B, 1996, 63, 593-598. 42. Kim, S. S.; Ogura, S.; Ikuta, H.; Uchimoto, Y.; Wakihara, M. Reaction Mechanisms of MnMoO4 for High Capacity Anode Material of Li Secondary Battery. Solid State Ionics, 2002, 146, 249-256. 43. Sleight, A.W. Accurate Cell Dimensions for ABO4 Molybdates and Tungstates. Acta Crystallogr. Sect. B, 1972, 28, 2899–2902. 44. Sharma, N.; Shaju, K. M.; Subba, G. V. R.; Chowdari, B.V.R.; Dong, Z. L.; White, T. J. Carbon-coated Nanophase CaMoO4 as Anode Material for Li-ion Batteries. Chem. Mater. 2004, 16, 504–512.

32 ACS Paragon Plus Environment

Page 32 of 39

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

45. Neeraj, S.; Kijima, N.; Cheetham, A. K. Novel Red Phosphors for Solid-State Lighting: The System NaM(WO4)2− x(MoO4)x: Eu3+(M= Gd, Y, Bi). Chem. Phys. Lett. 2004, 387, 2–6. 46. Cavalli, E.; Boutinaud, P.; Mahiou, R.; Bettinelli, M.; Dorenbos, P. Luminescence Dynamics in Tb3+-doped CaWO4 and CaMoO4 Crystals. Inorg. Chem. 2010, 49, 4916–4921. 47. Marques, V. S.; Cavalcante, L. S.; Sczancoski, J. C.; Alcantara, A. F. P.; Orlandi, M. O.; Moraes, E.; Longo, E.; Varela, J. A.; Li, M. S.; Santos, M. R. M. C. Effect of Different Solvent Ratios (water/ethylene glycol) on the Growth Process of CaMoO4 Crystals and Their Optical Properties. Cryst. Growth Des. 2010, 10, 4752–4768. 48. Singh, B. P.; Parchur, A. K.; Ningthoujam, R. S.; Ansari, A. A.; Singh, P.; Rai, S. B.; Influence of Gd3+ Co-Doping on Structural Property of CaMoO4: Eu Nanoparticles. Dalton Trans. 2014, 43, 4770–4778. 49. Karthik, R.; Karikalan, N.; Chen, S. M.; Kumar, J. V.; Karuppiah, C.; Muthuraj, V. Assessment of Divergent Functional Properties of Seed-Like Strontium Molybdate for the Photocatalysis and Electrocatalysis of the Postharvest Scald Inhibitor Diphenylamine. J. Catal. 2017, 352, 606–616. 50. Wu, J. L.; Cao, B. S.; Lin, F.; Chen, B. J.; Sun, J. S.; Dong, B. A New Molybdate Host Material: Synthesis, Upconversion, Temperature Quenching and Sensing Properties. Ceram. Int. 2016, 42, 18666–18673 51. Volkov, V.; Cascales, C.; Kling, A.; Zaldo, C. Growth, Structure, and Evaluation of Laser Properties of LiYb(MoO4)2 Single Crystal. Chem. Mater. 2005, 17, 291-300. 52. Park, S. H.; Kim, B. K.; Lee, W. J. Electrospun Activated Carbon Nanofibers with Hollow Core/Highly Mesoporous Shell Structure as Counter Electrodes for Dye-Sensitized Solar Cells. J. Power Sources 2013, 239, 122-127.27. 53. Sebastian, D.; Baglio, V.; Girolamo, M.; Moliner, R.; Lazaro, M.; Arico, A.S. Carbon Nanofiber-Based Counter Electrodes for Low Cost Dye-Sensitized Solar Cells. J. Power Sources. 2014, 250, 242-249. 54. Ramasamy, E.; Lee, J. Large-pore Sized Mesoporous Carbon Electrocatalyst for Efficient Dye-Sensitized Solar Cells. Chem. Commun. 2010, 46, 2136-2138. 55. Murakami, T. N.; Ito, S.; Wang, Q.; Nazeeruddin, M. K.; Bessho, T.; Cesar, I.; Liska, P.; Baker, R. H.; Comte, P.; Pechy, P. Highly Efficient Dye-Sensitized Solar Cells Based on Carbon Black Counter Electrodes. J. Electrochem. Soc. 2006, 153, 2255-2261. 33 ACS Paragon Plus Environment

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

56. Zhu, G.; Pan, L.; Lu, T.; Xu, T.; Sun, Z. Electrophoretic Deposition of Reduced GrapheneCarbon Nanotubes Composite Films as Counter Electrodes of Dye-Sensitized Solar Cells. J. Mater. Chem. 2011, 21, 14869-14875. 57. Zhou, Z.; Sigdel, S.; Gong, J.; Vaagensmith, B.; Elbohy, H.; Yang, H.; Krishnan, S.; Wu, X. F.; Qiao, Q. Graphene-beaded Carbon Nanofibers with Incorporated Ni Nanoparticles as Efficient Counter-Electrode for Dye-Sensitized Solar Cells. Nano Energy 2016, 22, 558-563. 58. Gong, H. H.; Park, S.H.; Lee, S. S.; Hong, S. C. Facile and Scalable Fabrication of Transparent and High Performance Pt/reduced Graphene Oxide Hybrid Counter Electrode for Dye-Sensitized Solar Cells. Int. J. Precis. Eng. Manuf. 2014, 15, 1193-1199. 59. Karthik, R.; Sasikumar, R.; Chen, S. M.; Kumar, J. V.; Elangovan, A.; Muthuraj, V.; Muthukrishnan, P.; Al-Hemaid, F. M.; Ali, M. A.; Elshikh, M. S. A highly Sensitive and Selective Electrochemical Determination of Non-Steroidal Prostate Anti-Cancer Drug Nilutamide based on f-MWCNT in Tablet and Human Blood Serum Sample. J. Colloid. Interface Sci. 2017, 487, 289-296. 60. Joh, H.; Ha, H. Y.; Prabhuram, J.; Jo, S. M.; Moon, S. H. Synthesis of Branched Carbon Nanotubes by Carbonization of Solid Polyvinylidene Fluoride Fibers. Carbon 2011, 49, 4601−4603. 61. Mao, X. W.; Yang, X. Q.; Rutledge, G. C.; Hatton, T. A. Ultra-Wide-Range Electrochemical Sensing using Continuous Electrospun Carbon Nanofibers with High Densities of States. ACS Appl. Mater. Interfaces, 2014, 6, 3394−3405. 62. Liu, J.; Kuo, Y. T.; Klabunde, Y. T.; Rochford, C.; Wu, J.; Li, J. Novel Dye-Sensitized Solar Cell Architecture using TiO2-coated Vertically Aligned Carbon Nanofiber Arrays ACS Appl. Mater. Interfaces 2009, 1, 1645-1649. 63. Ramaraj, S.; Mani, S.; Chen, S. M.; Kokulnathan, T.; Lou, B. S.; Ali, M. A.; Hatamleh, A. A.; Al-Hemaid, F. M. A. Synthesis and Application of Bismuth Ferrite Nanosheets Supported Functionalized Carbon Nanofiber for Enhanced Electrochemical Detection of Toxic Organic Compound in Water Samples. J. Colliod. Interface. Sci. 2018, 514, 59-69. 64. Huang, Y.; Miao, Y.; Ji, S.; Tjiu, W. W.; Liu, T. Electrospun Carbon Nanofibers Decorated with Ag−Pt Bimetallic Nanoparticles for Selective Detection of Dopamine. ACS Appl. Mater. Interfaces, 2014, 6, 12449−12456.

34 ACS Paragon Plus Environment

Page 34 of 39

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

65. Atuchin, V. V.; Aleksandrovsky, A. S.; Chimitova, O. D.; Krylov, A. S.; Molokeev, M. S.; Bazarov, B. G.; Bazarova, J. G.; Xia, Z. Synthesis and Spectroscopic Properties of Multiferroic β'-Tb2(MoO4)3. Opt. Mater., 2014, 36, 1631-1635. 66. Atuchin, V. V.; Grossman, V. G.; Adichtchev, S. V.; Surovtsev, N. V.; Gavrilova, T. A.; Bazarov, B. G. Structural and Vibrational Properties of Microcrystalline TlM(MoO4)2 (M = Nd, Pr) Molybdates. Opt. Mater., 2012, 34, 812-816. 67. Atuchin, V. V.; Aleksandrovsky, A. S.; Molokeev, M. S.; Krylov, A. S.; Oreshonkov, A. S.; Zhou, D. Structural and Spectroscopic Properties of Self-Activated Monoclinic Molybdate BaSm2(MoO4)4. J. Alloys Compd., 2017, 729, 843-849. 68. Mestl, G.; Srinivasan, T. K. K. Raman Spectroscopy of Monolayer-Type Catalysts: Supported Molybdenum Oxides. Catal. Rev. Sci. Eng, 1998, 40,451-570. 69. Wu, J.; Qin, X.; Zhang, H.; He, Y. B.; Li, B.;Ke, L.; Lv, W.; Du, H.; Yang, Q. H.; Kang, F. Multilayered Silicon Embedded Porous Carbon/Graphene Hybrid Film as a High Performance. Anode Carbon 2015, 84, 434-443. 70. Lim, C. S.; Aleksandrovsky, A.; Molokeev, M.; Oreshonkov, A.; Atuchin, V. The Modulated Structure and Frequency Up-conversion Properties of CaLa2(Moo4)4:Ho3+/Yb3+ Phosphors Prepared by Microwave Synthesis. Phys. Chem. Chem. Phys., 2015, 17, 19278-19287. 71. Lim, C. S.; Aleksandrovsky, A. S.; Molokeev, M. S.; Oreshonkov, A. S.; Atuchin, V. V. Microwave Synthesis and Spectroscopic Properties of Ternary Scheelite-Type Molybdate Phosphors NaSrLa(MoO4)3:Er3+,Yb3+. J. Alloys Compd., 2017, 713, 156-163. 72. Szytula, A.; Jezierski, A.; Penc, B.; Winiarski, A.; Leithe-Jasper, A.; Kaczorowski, D. Electronic Structure of YbTX Compounds. J. Alloy. Compd., 2003, 360, 41–46. 73. Gopiraman, M.; Babu, S. G.; Khatri, Z.; Wei, K.; Kim, Y. A.; Endo, M.; Karvembu, R.; Kim, I. S. Dry Synthesis of Easily Tunable Nano Ruthenium Supported on Graphene: Novel Nanocatalysts for Aerial Oxidation of Alcohols and Transfer Hydrogenation of Ketones. J. Phys. Chem. C 2013, 117, 23582−23596. 74. Atuchin, V. V.; Khyzhun, O. Y.; Chimitova, O. D.; Molokeev, M. S.; Gavrilova, T. A.; Bazarov, B. G.; Bazarova, J. G. Electronic Structure of β-RbNd(MoO4)2by XPS and XES. J. Phys. Chem. Solids, 2015, 77, 101-108. 75. Atuchin, V. V.; Aleksandrovsky, A. S.; Chimitova, O. D.; Diao, C. P.; Gavrilova, T. A.; Kesler, V. G.; Molokeev, M. S.; Krylov, A. S.; Bazarov, B. G.; Bazarova, J. G.; Lin, Z. 35 ACS Paragon Plus Environment

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

Electronic Structure of β-RbSm(MoO4)2 and Chemical Bonding in Molybdates. Dalton Trans., 2015, 44, 1805-1815. 76. Solodovnikov, S. F.; Atuchin, V. V.; Solodovnikova, Z. A.; Khyzhun, O. Y.; Danylenko, M. I.; Pishchur, D. P.; Plyusnin, P. E.; Pugachev, A. M.; Gavrilova, T. A.; Yelisseyev, A. P.; Reshak, A. H.; Alahmed, Z. A.; Habubi, N. F. Synthesis, Structural, Thermal, and Electronic Properties of Palmierite-Related Double Molybdate α‑Cs2Pb(MoO4)2. Inorg. Chem., 2017, 56, 3276-3286. 77. Ghosh, D.; Giri, S.; Das, C. K. Synthesis, Characterization and Electrochemical Performance of Graphene Decorated with 1D NiMoO4. nH2O nanorods. Nanoscale 2013, 5, 10428-10437. 78. Bielska, D.; Karewicz, A.; Lachowicz, T.; Berent, K.; Szczubialka, K.; Nowakowska, M. Hybrid Photosensitizer based on Halloysite Nanotubes for Phenol-Based Pesticide Photodegradation. Chem. Eng. J. 2015, 262, 125–132. 79. Suea-Ngam, A.; Rattanarat, P.; Chailapakul, O.; Srisa-Art, M. Electrochemical DropletBased Microfluidics using Chip-Based Carbon Paste Electrodes for High-Throughput Analysis in Pharmaceutical Applications. Anal. Chim. Acta, 2015, 883, 45-54. 80. Karthik, R.; Kumar, J. V.; Chen, S. M.; Karuppiah, C.; Cheng, Y. H.; Muthuraj, V. A Study of Electrocatalytic and Photocatalytic Activity of Cerium Molybdate Nanocubes Decorated Graphene Oxide for the Sensing and Degradation of Antibiotic Drug Chloramphenicol. ACS Appl. Mater. Interfaces, 2017, 9, 6547-6559. 81. Alhassan, S. M, Qutubuddin, S.; Schiraldi, D. A. Graphene Arrested in Laponite-Water Colloidal Glass. Langmuir, 2012, 28, 4009–4015. 82. Wang, J.; Chen, L.; Mulchandani, A.; Mulchandani, P.; Chen, W. Remote Biosensor for InSitu Monitoring of Organophosphate Nerve Agents. Electroanalysis, 1999, 11, 866–869. 83. Laothanachareon, T.; Champreda, V.; Sritongkham, P.; Somasundrum, M.; Surareungchai, W. Cross-linked Enzyme Crystals of Organophosphate Hydrolase for Electrochemical Detection of Organophosphorus Compounds. World J MicrobiolBiotechnol. 2008, 24, 3049– 3055. 84. Deo, R. P.; Wang, J.; Block, I.; Mulchandani, A.; Joshic, K.A.; Trojanowicz, M.; Scholz, F.; Chen, W.; Lin, Y. Determination of Organophosphate Pesticides at a Carbon Nanotube/Organophosphorus Hydrolase Electrochemical Biosensor. Anal. Chim. Acta, 2005, 530, 185–189. 36 ACS Paragon Plus Environment

Page 36 of 39

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

85. Lee, J. H.; Park, J. Y.; Min, K.; Cha, H. J.; Choi, S. S.; Yoo, Y. J. A Novel Organophosphorus Hydrolase-Based Biosensor using Mesoporous Carbons and Carbon Black for the Detection of Organophosphate Nerve Agents. Biosens. Bioelectron. 2010, 25, 1566–1570. 86. Chough, S. H.; Mulchandani, A.; Mulchandani, P.; Chen, W.; Wang, J.; Rogers, K. R. Organophosphorus Hydrolase-Based Amperometric Sensor: Modulation of Sensitivity and Substrate Selectivity. Electroanalysis, 2002, 14, 274–276. 87. Pedrosa, V. A.; Paliwal, S.; Balasubramanian, S.; Nepal, D.; Davis, V.; Wild, J.; Ramanculo, E.; Simonian, A. Enhanced Stability of Enzyme Organophosphate Hydrolase Interfaced on the Carbon Nanotubes. Colloids Surf B: Biointerfaces, 2010, 77, 69–74. 88. Arduini, F.; Neagu, D.; Scognamiglio, V.; Patarino, S.; Moscone, D.; Palleschi, G. Automatable Flow System for Paraoxon Detection with an Embedded Screen-Printed Electrode Tailored with Butyrylcholinesterase and Prussian Blue Nanoparticles. Chem Aust. 2015, 3, 129–145. 89. Jha, N., Ramaprabhu, S. Carbon Nanotube-Polymer based Nanocomposite as Electrode Material for the Detection of Paraoxon. J. Nanosci. Nanotechnol. 2010, 10, 2798–2802. 90. Huang, H.; Tu, S.; Zeng, C.; Zhang, T.; Reshak, A. H.; Zhang, Y. Macroscopic Polarization Enhancement Promoting Photo- and Piezoelectric-Induced Charge Separation and Molecular Oxygen Activation. Angew. Chem. Int. Ed., 2017, 56, 11860–11864. 91. Huang, H.; Li, X.; Wang, J.; Dong, F.; Chu, P. K.; Zhang, T.; Zhang, Y. Anionic Group SelfDoping as a Promising Strategy: Band-Gap Engineering and Multi-Functional Applications of High-Performance CO32−-Doped Bi2O2CO3. ACS Catalysis, 2015, 5, 4094-4103. 92. Huang, H.; Xiao, K.; He, Y.; Zhang, T.; Dong, F.; Du, X.; Zhang, Y. In Situ Assembly of BiOI@Bi12O17Cl2 p-n junction: Charge Induced Unique Front-Lateral Surfaces Coupling Heterostructure with High Exposure of BiOI {001} Active Facets for Robust and Nonselective Photocatalysis. Appl. Catal. B: Environ., 2016, 199, 75-86. 93. Huang, H.; He, Y.; Li, X.; Li, M.; Zeng, C.; Dong, F.; Du, X.; Zhang, T.; Zhang, Y. Bi2O2(OH)(NO3) as a Desirable [Bi2O2]2+ Layered Photocatalyst: Strong Intrinsic Polarity, Rational Band Structure and {001} Active Facets Co-Beneficial for Robust Photooxidation Capability. J. Mater. Chem. A, 2015, 3, 24547–24556.

37 ACS Paragon Plus Environment

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

94. 94. Hirakawa, T.; Nosaka, Y. Properties of O2•−and •OH Formed in TiO2 Aqueous Suspensions by Photocatalytic Reaction and the Influence of H2O2 and Some Ions. Langmuir, 2002, 18, 3247– 3254. 95. 95. Uma, S.; Singh, J.; Thakral, V. Facile Room Temperature Ion-Exchange Synthesis of Sn2+ Incorporated Pyrochlore-Type Oxides and Their Photocatalytic Activities. Inorg. Chem., 2009, 48, 11624–11630. 96. 96. Chen, Z.; Li, D.; Zhang, W.; Shao, Y.; Chen, T.; Sun, M.; Fu, X. Photocatalytic Degradation of Dyes by ZnIn2S4 Microspheres under Visible Light Irradiation. J. Phys. Chem. C, 2009, 113, 433–4440. 97. Prasad, G. K.; Ramacharyulu, J. P.; Srivastava, A.R.; Singh, B. Photocatalytic Degradation of Paraoxon-ethyl in Aqueous Solution using Titania Nanoparticulate film. Thin Solid Films 2012, 520, 5597–5601.

38 ACS Paragon Plus Environment

Page 38 of 39

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 Ytterbium molybdate/ƒ-carbon nanofibers nanocomposite exhibited bifunctional catalytic activity for the detection and degradation of chemical warfare agent paraoxon-ethyl

39 ACS Paragon Plus Environment