Dual Functional Ta-Doped Electrospun TiO2 Nanofibers with

Aug 4, 2017 - Lili YangYusi PengYong YangJianjun LiuZhiyuan LiYunfeng MaZhang ZhangYuquan WeiShuai LiZhengren HuangNguyen Viet Long...
0 downloads 0 Views 1MB Size
Subscriber access provided by Warwick University Library

Article

Dual functional Ta doped electrospun TiO2 nanofibers with enhanced photocatalysis and SERS detection for organic compounds Narendra Singh, Jai Prakash, Mrinmoy Misra, Ashutosh Sharma, and Raju Kumar Gupta ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.7b07571 • Publication Date (Web): 04 Aug 2017 Downloaded from http://pubs.acs.org on August 4, 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 Applied Materials & Interfaces 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 32

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

Dual Functional Ta Doped Electrospun TiO2 Nanofibers with Enhanced Photocatalysis and SERS Detection for Organic Compounds Narendra Singh,a Jai Prakash,a,b Mrinmoy Misra,a Ashutosh Sharmaa,c and Raju Kumar Gupta*a,c a Department of Chemical Engineering and Center for Nanosciences, Indian Institute of Technology Kanpur, Kanpur-208016, UP, India b Department of Physics, University of the Free State, Bloemfontein-9300, South Africa c Center for Environmental Science and Engineering, Indian Institute of Technology Kanpur, Kanpur-208016, UP, India * Corresponding author. Tel: +91-5122596972; Fax: +91-5122590104. E-mail address: [email protected] Abstract There is a growing interest in multifunctional nanomaterials for the detection as well as degradation of organic contaminants in the water. In this work, we report on the development of dual functional TiO2 nanofibers (TNF) with different tantalum (Ta) doping (1-10 mol%) by a simple electrospinning technique. As-prepared TNF show mesoporous dominant structure, which are favourable for photocatalytic activity due to the presence of catalytic spots. Ta doping decreases the crystalline size within TiO2 matrix due to the incorporation of Ta5+ ions and restricts the phase transformation from anatase to rutile. Ta doping slightly enhances the visible light absorption due to the Ti3+ defects sites created upon Ta5+ doping. The effect of Ta doping within TiO2 matrix was systematically studied for the degradation of methylene blue (MB) dye under ultraviolet (UV) and solar light irradiation. The 5% doped Ta TNF were found to be optimal and showed 5.1 and 2.2 times higher photocatalytic activity as compared to TNF under UV and solar light irradiation, respectively. The effect of Ta doping for the detection of MB molecules was also studied by surface enhanced Raman scattering (SERS). It was observed that 5% Ta doped TNF exhibit higher photocatalytic activity and enhanced SERS signals of adsorbed MB molecules as compared to the TNF. The enhanced photocatalytic and SERS activities can be explained as combined effects of enhanced visible light absorption, lower crystalline size and slightly higher surface area. The observed results show that Ta doping induces new energy levels below the conduction band of TiO2 due to Ti3+ defects, which inhibit the photogenerated charge recombination acting as electron traps and promote charge transfer mechanism acting as an intermediate state for TiO2 to MB molecule electron transfer, and are mainly responsible for the enhanced photocatalytic and SERS activities, respectively. 1 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

Keywords: Electrospinning, Nanofibers, Doping, Degradation, Photocatalysis, SERS. 1.

Introduction Among metal oxide semiconductors, titanium dioxide (TiO2) has received a greater

attention because of its fascinating properties such as low chemical toxicity, resistance to chemical breakdown, higher stability and low cost.1-3 It shows multifunctional applications in various fields such as photocatalysis, sensing, solar cell and ultra-trace detection of organic pollutants/biomolecules using surface enhanced Raman scattering (SERS).4-7 Due to the modern industrialization and frequent water pollution problems, there is a growing demand to produce cost effective techniques and multifunctional nanomaterials for the detection as well as degradation of organic contaminants in the wastewater. SERS is an ultrasensitive technique, which gives molecular level information of surface adsorbed molecules. It has the ability to detect even a single molecule and has shown great promises in recent years to detect organic pollutants such as dye molecules.5, 8 Recent literature shows that TiO2 and TiO2 based nanomaterials have gained special attention for such environmental issues because of their great potential to be applied as photocatalyst as well as SERS substrates.6, 9 TiO2 is an active nanomaterial mainly in ultraviolet (UV) light due to its wide band gap (energy band gap for rutile - 3.0 eV and for anatase - 3.2 eV).10 In order to maximize the effective utilization of TiO2 as a photocatalyst and in other fields, its optical property needs to be extended in visible light range. Effective strategies for reducing the energy band gap of TiO2 remain to be of high interest by the scientific community for its variable applications. There are several approaches to improve the visible light absorption such as metal nanoparticles deposition over TiO2, doping with anions/transition metals, sensitization by quantum dots and formation of hybrid composites, etc.10-11 These treatments modify the energy band gap of TiO2 by introducing some defects or sub-energy levels between conduction and valance bands, thus facilitating separation of photogenerated charge carriers. Furthermore, the optoelectronic properties are enhanced providing different aspect for multifunctional applications. Metal nanoparticles deposition over TiO2 reduces excitons recombination significantly by creating Schottky barrier acting as an electron trap and improves the photocatalytic activity of the TiO2.12-13 Moreover, metal nanoparticles deposited over TiO2 forming nanocomposites also extend its functionality for the ultratrace detection of organic molecules serving as an active SERS substrate.6, 14 Quantum dots sensitization over metal oxides (TiO2, ZnO, etc.) enhances their optical absorption and charge separation at the 2 ACS Paragon Plus Environment

Page 3 of 32

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

interface (e.g. TiO2/CdS creates excitons upon solar illumination and photogenerated electrons from CdS transferred to TiO2 and simultaneously quantum dots show tunable optical property with size).15 Anion doping (e.g. N, S, C, etc.) decreases its band gap by some extent due to creation of new discrete energy level above the valance band.11, 16-19 Similarly, doping of transition metals (e.g. Cr, W, Ni and Fe, etc.) into TiO2 creates new energy level between conduction and valance bands, which enhances visible light absorption.20-23 Moreover, several dopants such as Mn, Co, Zn etc. in TiO2 have shown enhanced photocatalytic properties due to the improved optical and charge separation ability of the doped TiO2 through the production of surface defects. Additionally, these doped TiO2 materials show the enhanced SERS functionality towards the organic molecules.24-26 Similarly, Ni and Zn co-doping in TiO2 has shown to create abundant surface states and metal doping levels band gap, which facilitates the charge transfer from TiO2 to the probe molecule and enhances the SERS sensitivity.27 Other dopants such as Nb, Ta, etc. have been extensively used in improving the properties of TiO2 for various applications ranging from photocatalysis to photovoltaic devices. However, Ta doping has advantage over Nb doping owing to its higher equilibrium solubility within TiO2 network because of lower energy requirement.28 The Ta ions are soluble in TiO2 matrix due to ionic radius of Ta5+ ions (0.064 nm) which is slightly larger than the Ti4+ ions (0.061 nm).29 There are several studies which report on enhanced photocatalytic properties of Ta functionalized TiO2 nanomaterials using different techniques. For example: Wang et al. prepared three dimensional macroporous TiO2/Ta2O5 mixed oxide structures which showed that photocatalytic activity was increased by 2.6 times than TiO2 under UV light.30 Znad et al. prepared Ta/TiO2 mixed oxide by impregnation method which exhibited decreased band gap and enhanced photocatalytic activity under solar light.31 Bawaked et al. prepared undoped and Ta doped TiO2 thin films by aerosol assisted chemical vapour deposition method and found that doped TiO2 thin films showed good transport properties and hence better photocatalytic activity.32 Similarly, there are few studies demonstrating the application of TiO2 based nanomaterials in SERS. For example: three-dimensional TiO2 nanorods decorated with Ag nanoparticles

33

and Ag embedded TiO2 nanotubes34 exhibited enhanced SERS

properties. Recently, Ta/N co-doped TiO2 nanorods have been reported for their excellent photocatalytic activities.35 However, there are only few studies showing multifunctional applications based on one dimensional TiO2 nanostructures. For instance: recently, Kumar et al., have shown 3 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

multifunctional Ag-TiO2 nanorods with enhanced photocatalytic degradation and SERS detection of dye molecules.9 However, these composite materials involves multistep synthesis procedure including synthesis of noble metal nanoparticles, their characterizations and then assembly with TiO2, while TiO2 itself does not play any role in SERS activity.6,

9

On the

contrary, as discussed above, doped TiO2 nanostructures have shown their capability to be implemented in both the field of photocatalysis as well as SERS for the ultra-detection of the molecules and have easy synthesis process. To enhance the performance of TiO2, one-dimensional structures such as nanotubes, nanowires and nanofibers have been synthesized to provide unidirectional paths for photogenerated electrons and to minimize its recombination process.36-37 One-dimensional nanostructures can be prepared by hydrothermal, electrochemical, and electrospinning etc.3841

Among these techniques, electrospinning is a simple and preferred route for the synthesis

of one dimensional nanofibers at room temperature with high aspect ratio, controlled porosity and high chemical reactivity. Furthermore, electrospinning technique provide more flexibility in synthesizing different composites and compounds.42 It is thus interesting to fabricate and study Ta doped one-dimensional TiO2 nanostructures and their multifunctional applications in various directions. This is the first study exploring the dual functional application of Ta doped TiO2 nanofibers in the field of photocatalysis and SERS for the photo-degradation and ultra-detection of dye molecules, respectively. In this work, one dimensional Ta doped TiO2 nanofibers have been synthesized via electrospinning technique. We have explored the effects of Ta doping on energy band gap and visible light absorption for the Ta doped TiO2 electrospun nanofibers. Furthermore, the photocatalytic activity of doped and undoped TiO2 nanofibers has been studied to degrade MB (model pollutant) under the irradiation of UV and solar light to evaluate the effect of doping. In addition, these nanofibers have also been explored to be used as SERS substrate for the detection of MB dye molecules. Finally, mechanisms have been proposed for the observed photocatalytic and SERS enhancements. 2. 2.1

Experimental details Materials required Titanium tetraisopropoxide (TTIP, 97%) and tantalum ethoxide were purchased from

Sigma Aldrich Pvt. Ltd. Polyvinylpyrrolidone (PVP, MW= 1,300,000 g mol-1) and MB were purchased from Alfa Aesar. Deionized water (DI) was used for preparation of MB solution.

4 ACS Paragon Plus Environment

Page 5 of 32

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

Ethanol (emsure 98%) and acetic acid (HPLC grade) were purchased from Merck Chemicals. All chemicals were used without further purification. 2.2

Synthesis of undoped and Ta doped TNF To prepare electrospinnable solution, 0.45 g of PVP was dissolved in 7.5 mL ethanol at

55 oC under continuous magnetic stirring for 20-30 min. In another reagent bottle, 1.5 g of TTIP was dissolved in 6 mL of mixed acetic acid and ethanol (volume ratio = 1:1) under continuous stirring at room temperature for 20-30 min. Ethanolic solution of PVP was poured into TTIP solution at room temperature and stirred continuously for another 10 h to get homogeneous solution for electrospinning. For Ta doped TNF, Ta precursor was mixed with TiO2 precursor in the proper molar ratio for different Ta doping. Ta doped TNF with different molar ratios of 1%, 2%, 5% and 10% Ta are named as 1% Ta TNF, 2% Ta TNF, 5% Ta TNF and 10% Ta TNF, respectively. The prepared electrospinnable solution was loaded into a plastic syringe of 10 mL volume. The voltage applied at the needle was 13 kV and rotary drum collector was grounded. The solution was pushed through a syringe pump at a rate of 20 μL min-1. The Rotary drum collector was placed at 10 cm from the tip of needle. The drum was wrapped with aluminium foil to collect nanofibers mat. The nanofibers mat collected on aluminium foil was placed in the oven at 60 oC to peel off from aluminium foil. The collected freestanding mat was placed in a furnace at 500 oC for 2 h under air to remove polymer and solvent. Then, the sample was allowed to cool naturally inside the furnace to room temperature for further use. 2.3

Photocatalytic experiments Prior to the photocatalysis experiments, adsorption-desorption experiment was carried out

for 2 h in dark with constant stirring. 20 μM and 40 μM MB solutions were prepared for the photocatalytic process under UV and solar light, respectively. Photocatalytic study was carried out in a petridish by dispersing 10 mg of TNF photocatalyst (both separately doped and undoped) in 10 mL of 20 μM MB dye and then the solution was exposed to UV lamp (12 W, 365 nm). To study the visible light photocatalytic activity, the photo degradation study was performed under simulated light source (1 Sun, OAI TriSOL). For this experiment, 10 mg of each photocatalyst was dispersed in 50 mL (40 μM) of MB solution and exposed to solar light. Reusability test of 5% Ta TNF was carried out for 5 cycles under illumination of UV and solar light following the same experimental conditions as above. In order to monitor the time dependent photocatalytic degradation process, sample (MB) was collected at regular

5 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

time intervals and absorbance of MB dye was recorded using Agilent technologies Cary 7000 UV-vis-NIR absorption spectrophotometer. 2.4

SERS experiments For SERS experiment, first of all, an aqueous solution of 20 µM of MB was prepared,

then 5 mg of TNF and Ta doped TNF were dispersed in 5 mL of 20 µM MB aqueous solution separately. The solution was stirred for 2 h. Then an aliquot of the solution was dropped on a glass slide and kept it overnight for drying at the room temperature. These glass slides were used as a sample for SERS measurements. The Raman spectra were taken at many places and average of Raman spectra was used for the analysis.

2.5 Characterization techniques The surface morphology of TNF and Ta doped TNF were investigated through field emission scanning electron microscope (FESEM, Quanta 200, Zeiss) and transmission electron microscope (TEM, Tecnai 20G2). Element composition of nanofibers was determined by energy dispersive X-ray spectroscopy (EDX linked to FESEM, Oxford Instrument). Crystal structure of prepared TNF and Ta doped TNF was investigated by XRay Diffraction (XRD, X’Pert Pro, PANanalytical) using Cu Kα as X-ray source (λ=1.5406 Ao). Raman spectra of all samples were collected by Raman spectrometer (WiTec, using laser light of λ = 532 nm). The surface area and pore size of synthesized sample were investigated using Brunauer–Emmett–Teller instrument (BET, Quantachrome Instruments). Thermal behaviour of as-collected electrospun TiO2 mats was determined through thermo gravimetric analysis (TGA) using TA instrument. Electronic structure of TNF and Ta doped TNF were investigated by X-ray photoelectron spectroscopy (XPS, PHI 5000 Versa Probe II, FEI Inc.) and XPS spectra was analysed using the XPSPEAK 4.1 software. UV-vis absorption spectra of all the samples were taken using UV-vis-NIR spectrophotometer (Agilent technologies, Cary 7000). Fourier transform infrared spectra (FT-IR, Perkin Elmer) were recorded using KBr pellet method. Electron paramagnetic resonance (EPR) spectra were collected using Bruker-EMX spectrophotometer. Surface potential of undoped and 5% Ta TNF were investigated using Kelvin probe force microscopy (KPFM, Asylum MFP-3D AFM).

6 ACS Paragon Plus Environment

Page 7 of 32

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

3.

Results and discussion

3.1 Characterization of TNF and Ta doped TNF Figure 1a shows XRD patterns of TNF and Ta doped TNF in the range of 20-80o. XRD pattern of TNF is well matched with anatase phase of TiO2 and consistent with standard card (JCPDS, No. 01-084-1285). It also shows small content of rutile phase (JCPDS, No. 01-0770440). The main characteristic peak of anatase TiO2 is observed at 25.30o that corresponds to (101) plane. The XRD patterns do not contain any peak corresponding to carbon or polymer which confirms that the polymer and solvent were removed completely during heat treatment. XRD patterns of 1%, 2% and 5% Ta doped TNF are matched with anatase phase of TiO2 and no other peaks corresponding to Ta or Ta2O5 are observed. It indicates the well distribution of doped Ta ions within the TiO2 matrix. However, XRD pattern of 10% Ta TNF shows small peak at 22.94o that may be due to the segregation of Ta2O5 (JCPDS No. 01-073-0005). Ta doping restricts the phase transformation from anatase to rutile in Ta doped TNF even at higher temperature as reported previously.43 Interestingly, it can be observed that a peak corresponding to anatase (101) plane is slightly shifting to lower 2θ with an increase in Ta doping concentration (Figure 1b). The peak corresponding to plane (101) also gets broadened with an increase in doping, which is a signature of decreasing crystalline size. This peak is shifted from 25.30o to 25.21o for TNF to 10% Ta TNF which is attributed to the incorporation of Ta5+ at the expense of Ti4+ into anatase phase of TiO2 matrix. The crystalline size of anatase (101) plane for TNF and Ta doped TNF samples was calculated using Debye Scherrer formula, as given below:44

D=

Kλ βcosθ

(1)

where, K is 0.89 taken as a shape factor, λ is the wavelength of X-Ray used, β is full width at half maximum for (101) plane of anatase phase of TiO2 and θ is the diffraction angle. The calculated crystalline sizes of (101) plane for TNF, 1% Ta TNF, 2% Ta TNF, 5% Ta TNF and 10% Ta TNF are found to be 14.66, 10.98, 10.63, 10.51 and 10.19 nm, respectively using equation (1).

7 ACS Paragon Plus Environment

TNF

Page 8 of 32

TNF 1% Ta TNF 2% Ta TNF 5% Ta TNF 10% Ta TNF

b)

A204

A105 A211

A200

R101 A004 R111

R110

a)

2% Ta TNF

5% Ta TNF 10% Ta TNF

* 20

Intensity (a.u.)

1% Ta TNF

Intensity (a.u.)

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

A101

ACS Applied Materials & Interfaces

30

40

50

60

70

80

24

Angle (2θ)

25

26

27

Angle (2θ)

Figure 1. (a) XRD patterns and (b) Magnified XRD patterns in the region between 24o and 27o of TNF and Ta doped TNF. To acquire more information about TNF and Ta doped TNF, Raman spectra of TNF and Ta doped TNF (different doping %) were recorded at room temperature upon excitation with 532 nm laser beam. It is already specified that anatase TiO2 contains six Raman active characteristics peaks [Eg(1), Eg(2), B1g

(1),

(A1g+B1g(2)) and Eg(2)] corresponding to different

modes as shown in the Figure 2a. Figure 2a shows Raman spectra of TNF and Ta doped TNF. Raman spectrum of TNF shows major peaks at 144, 397, 517 and 640 cm-1 correspond to Eg(1), B1g(1), (A1g+B1g(2)) and Eg(2) modes of anatase TiO2, respectively.45 A very small peak at 196 cm-1 is also observed corresponds to Eg(2) mode. Raman spectra of all Ta doped TNF are also harmonised with Raman active modes of anatase TiO2. Furthermore, Raman spectra do not show any other detectable peak corresponding to any other form of Ti, TiO2 or Ta compounds. Thus, Raman spectra of Ta doped TNF confirm the presence of anatase phase, which is in good agreement with XRD results. It is also found that with an increase in Ta doping concentration, Raman spectra of doped TNF show a peak shift of B1g(1) (397 cm-1) mode to a lower wavenumber (Figure 2b). It confirms that Ti ions are substituted by Ta ions because B1g(1) peak at 397 cm-1 is a signature of the Ti-O stretching mode.46

8 ACS Paragon Plus Environment

Page 9 of 32

a)

B1g

Eg(2)

100

200

0% Ta doped TiO2 solid nanofiber 1% Ta doped TiO2 solid nanofiber 2% Ta doped TiO2 solid nanofiber 5% Ta doped TiO2 solid nanofiber 10% Ta doped TiO2 solid nanofiber

300

400

A1g B1g

500

Normalized Raman Intensiy

Eg(1)

Normalized Raman Intensiy

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

Eg(2)

600

700

800

900

300

TNF 1% Ta TNF 2% Ta TNF 5% Ta TNF 10% Ta TNF

b)

320

340

-1

Raman Shift (cm )

360

380

400

420

440

-1

Raman Shift (cm )

Figure 2. (a) Raman spectra and (b) Magnified Raman pattern in the region of 300-450 cm-1 of TNF and Ta doped TNF. In order to investigate the optical absorption behaviour of TNF and Ta doped TNF, the UV-vis absorption study was carried out in the range of 200-800 nm as shown in Figure 3a. Ta doping slightly alter the light absorption properties as observed in Figure 3a. The absorption curves indicate that with an increase in Ta doping concentration, absorption increases up to 5% Ta TNF and then decreases. Figures 3b-c and Figure S1 show transformed Kubelka Munk function vs. energy of light, indicating the energy band gap alteration with Ta doping. The energy band gap of TNF and Ta doped TNF was estimated using the following equation (2):47

(F(R) × hν) ∝ (hν − E g ) 2

(2)

where Eg is the band gap of the material, ν is the frequency of light, F(R) is the Kubelka Munk function and h is the planks constant. The band gap of TNF, 1% Ta TNF, 2% Ta TNF, 5% Ta TNF, 10% Ta TNF are found to be 3.11, 3.01, 2.99, 2.96 and 3.00 eV, respectively, as shown in the Figures 3b-c and Figure S1. An increase in Ta doping concentration results in the decrease in the energy band gap slightly up to 5% Ta TNF along with visible light absorption. The possible reason for the observed decrease in the energy band gap may be due to the presence of Ti3+ states within TiO2 matrix upon Ta5+ doping, as evidenced by XPS (discussed later), which can provide d-d transitions.32, 48 The absorbance study reveals that 5% Ta doping is the optimum doping in TNF.

9 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

TNF 1% Ta TNF 2% Ta TNF 5% Ta TNF 10% Ta TNF

2.5 2.0 1.5 1.0

3.0

0.5 0.0 200

2.5

3.5

b)

TNF

2.0 1.5 1.0 0.5

300

400

Wavelength (nm)

500

3.0

Energy Band gap = 3.11 eV

(F(R)*hν)0.5

3.0

3.5

a)

(F(R)*hν)0.5

3.5

F(R)

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

2.5

c)

5% Ta TNF Energy Band gap = 2.96 eV

2.0 1.5 1.0 0.5

0.0 0.0 2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0 2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0 E (eV) E (eV)

Figure 3. (a) UV-vis diffuse reflectance spectra (DRS) of TNF and Ta doped TNF; Plot of transformed Kubelka–Munk function vs. energy for (b) TNF and (c) 5% Ta TNF. Figures 4a-e show typical FESEM micrographs of TNF and Ta doped TNF. It has been observed that all doped and undoped samples show nanofibers morphology, which are continuous in nature. The diameter distribution and low magnified images of all the samples are shown in Figure S2. The diameters of all the samples are distributed in the range of 20100 nm whereas; average diameters of different doped and undoped matrix are in the range of 40-60 nm. All nanofibers are of several μm in length as observed by magnified FESEM images shown in Figures 4a-e and Figure S2. In order to investigate the doping concentration of the Ta in the nanofibers matrix, EDX study was performed and results are presented in Figure S3. As the Ta doping concentration is increased in primary electrospun solution, the atomic concentration of the Ta within TiO2 matrix is also increased, as confirmed by EDX (Figure S3). The surface of nanofibers both doped and undoped TNF is rough in nature, which suggests the porous nature of nanofibers. To further investigate the morphology in detail, TNF and 5% Ta TNF samples were studied using TEM. Figures 4f-g show TEM micrographs of undoped and 5% Ta doped solid nanofibers. TEM analysis also shows that all TNF are solid and porous in nature which is in good agreement with FESEM results. TEM analysis reveals that Ta doping does not affect the morphology of the TNF after doping. Also, digital images of TNF and 5% Ta TNF does not show any significant difference in colour (Figure 4h).

10 ACS Paragon Plus Environment

Page 11 of 32

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

a)

b)

c)

d)

e)

f)

g)

h) TNF

5% Ta TNF

Figure 4. FESEM micrographs of (a) TNF, (b) 1% Ta TNF, (c) 2% Ta TNF, (d) 5% Ta TNF, and (e) 10% Ta TNF; TEM micrographs of (f) TNF and (g) 5% Ta TNF; (h) digital image of TNF and 5% Ta TNF. As discussed in absorbance spectra analysis, the absorbance of Ta doped TNF is increased and band gap is decreased with an increase in Ta doping concentration up to 5% Ta TNF. Therefore, the XPS of TNF and 5% Ta TNF was performed to study the electronic structure and chemical changes due to Ta doping as shown in Figure 5. The high resolution XPS spectrum of Ti for TNF sample shows a 2p3/2 characteristic peak corresponding to Ti4+ oxidation state at a binding energy of 458.6 eV (Figure 5a).49-50 Moreover, the other peak at 464.3 eV (2p1/2) corresponds to Ti4+ oxidation state with separation of 5.7 eV. High resolution XPS spectra of Ti and Ta for 5% Ta TNF are shown in Figures 5b-c. It can be observed that the Ti has a doublet characteristic peak at 458.6 and 464.3 eV corresponding to 2p3/2 and 2p1/2 of Ti4+ state as shown in the Figure 5b. A detectable peak can also be observed at lower binding energy than the peak appeared at 458.6 eV. This additional peak at 457.1 eV appears after the Ta doping which refers to Ti3+ oxidation state due to incorporation of Ta5+ ions within the crystal structure. High resolution XPS spectrum of Ta shows two 4f7/2 and 4f5/2 peaks corresponding to Ta5+ oxidation state, appeared at 26.0 and 27.9 eV, respectively (as shown in Figure 5c) with a separation of 1.9 eV.51 The XPS study reveals that Ta doping creates Ti3+ oxidation states within the TiO2 matrix in doped TNF which serve as defect sites and are responsible for the observed changes in optical properties.

11 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

Ti(IV) 2p1/2

Intensity (a.u.)

Ti2p

b)

c)

Ti(IV) 2p3/2

Intensity (a.u.)

Ti(IV) 2p3/2

a) Intensity (a.u.)

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

Ti2p

Ti(IV) 2p1/2

Page 12 of 32

Ta4f

Ta4f7/2 Ta4f5/2

Ti(III) 2p3/2

468 466 464 462 460 458 456 454 468 466 464 462 460 458 456 454 30

B.E.(eV)

B.E.(eV)

29

28

27

26

25

24

B.E.(eV)

Figure 5. XPS profiles of (a) Ti2p spectra of TNF; (b) Ti2p spectra and (c) Ta4f spectra of 5% Ta TNF. Furthermore, generation of Ti3+ states is also confirmed by EPR analysis as shown in Figure S4. It shows a room temperature EPR signal that lacks the signature of diamagnetic TiO2. A characteristic signal for paramagnetic Ti3+ was observed at average g = 1.97, which clearly suggests its paramagnetic nature and that is from the Ti3+. We have explored the Fermi level shift of undoped and 5% Ta TNF. We conducted Kelvin probe force microscopy (KPFM) of the undoped and 5% Ta TNF. KPFM is a technique which measures the distributed contact potential difference (CPD) between the tip and sample and provides surface potential using equation VCPD =

φ tip − φ sample e

(3)

Here, e is the electronic charge and ϕtip and ϕsample are the work functions of the tip and the sample, respectively. 5% Ta TNF sample showed CPD value of around -40 mV in comparison to undoped TNF showing CPD value of ~90 mV. The ϕtip has value of 5.0 eV, thus, the Fermi level of sample can be calculated using equation 3. The calculated value for undoped TNF sample is 4.9 eV, whereas for 5% Ta TNF sample, it is ~5.1 eV. Thus, this study reveals that Ta doping shifts the CB and Fermi level downward, which is in agreement with the UV-vis results. The N2 adsorption-desorption isotherms and pore size distributions of TNF and 5% Ta TNF have been shown in Figure 6. From N2 adsorption-desorption isotherms (Figures 6a and c) of TNF and 5% Ta TNF, it can be observed that N2 adsorption-desorption isotherms of both samples are a typical type IV isotherm (according to IUPAC classification) with a hysteresis loop (capillary condensation).52 Figures 6b and d show pore size distributions of TNF and 5% Ta TNF as calculated by Barrett-Joyner-Halenda (BJH) method. The BET surface area of TNF and 5% Ta TNF has been estimated as 36.2 and 46.1 m2 g-1, respectively. The enhanced surface area of 5% Ta TNF is due to the lower crystalline size as estimated by 12 ACS Paragon Plus Environment

Page 13 of 32

XRD results. The total pore volumes of TNF and 5% Ta doped TNF are 0.110 and 0.163 cm3 g-1, respectively. TNF contains both mesopores and macropores with mesopores in dominance (90% of the total pore volume). On the other hand, 5% Ta TNF contains all types of pores (micropores, mesopores and macropores) with mesopores in dominance (about 80% of the total pore volume). 80 70

0.40

a)

0.25

dV(logd)

40 30

0.20 0.15 0.10

20

0.05

10

0.00 -0.05

0 0.0

0.2

0.4

0.6

0.8

1

1.0

P/Po

120

100

0.25

c)

d)

0.20

Adsorption Desorption

80

10

Diameter (nm)

0.15

dV(logd)

Volume (cm3/g)

0.30

50

100

b)

0.35

Adsorption Desorption

60

Volume (cm3/g)

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

60 40

0.10 0.05

20 0.00

0 0.0

0.2

0.4

0.6

0.8

1.0

1

10

100

1000

Diameter (nm)

P/Po

Figure 6. (a) N2 adsorption-desorption isotherm and (b) Pore size distribution of TNF; (c) N2 adsorption-desorption isotherm and (d) Pore size distribution of 5% Ta TNF. FT-IR spectra of TNF and Ta doped TNF (2% and 10%) are shown in Figure S5. FTIR spectra of TNF, 2% Ta TNF and 10% Ta TNF show two absorption bands at 3420 cm-1 and 1635 cm-1 due to O-H stretching and O-H bending modes of adsorbed water respectively.53 All samples show absorption band at 2346 cm-1 due to the presence of atmospheric CO2 during measurement.53-54 TNF exhibits broad absorption bands in the range of 400-800 cm-1 due to the presence of vibrational modes of anatase TiO2. Two absorption bands at 670 cm-1 and 470 cm-1 are assigned to the Ti-O and Ti-O-Ti vibrational modes of TiO2.53,

55

Ta2O5

shows two absorption bands at 540 cm-1 and 670 cm-1 and these vibrational mode are due to the presence of Ta-O and Ta-O-Ta.56 For 2% Ta TNF and 10% Ta TNF, the absorption band corresponding to 670 cm-1 becomes stronger as compared to absorption band at 470 cm-1 due to overlapping of Ti-O and Ta-O vibrational modes. 13 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

Figure S6 shows the TGA graph of TNF and 5% Ta TNF. The complete removal of solvent and PVP can be observed during calcination of spun nanofibers. TNF and 5% Ta TNF follow the same trend due to the removal of solvent and polymer. It is found that the weight loss occurred during the temperature increment from room temperature to 200 oC due to desorption of water and removal of solvent. A further rise in temperature degrads the polymer to volatile compound and it removes completely around 480 oC. An additional rise in temperature does not cause any weight loss, which confirms the formation of TNF and Ta doped TNF. The estimated weight loss during the 100-500 oC for TNF and 5% Ta TNF (same volume of solvent and the same quantity of PVP) is around 47.2 % and 45.8 %, respectively. The weight loss is less for doped TNF as compared to undoped one due to the incorporation of Ta ions in the TiO2 crystal lattice. The TGA analysis supports XRD as well as EDX results, which show the absence of carbon in calcined nanofibers. 3.2

Photocatalytic activity of TNF and Ta doped TNF Photocatalytic tests were performed under UV and simulated solar light irradiation,

separately. These controlled experiments were performed in dark conditions in the presence of photocatalysts as shown in Figure S7. Analysis shows that MB molecules were adsorbed onto the surface of TNF and doped TNF photocatalysts due to the presence of surface area availability as well as porosity existing on the photocatalyst material. The MB adsorption over TNF, 1% Ta TNF, 2% Ta TNF, 5% Ta TNF, 10% Ta TNF was 18.6%, 16.3%, 18.9%, 21.8% and 28.9%, respectively after 2 h under continuous stirring (Figure S7). The MB molecules show its characteristic absorption peak at 664 nm, which diminishes gradually upon illumination in the presence of photocatalyst material (Figure S8). The blue colour of the dye also disappears with time and becomes colourless at the end as shown in the Figure 7b inset. During photocatalysis process, MB molecules are initially adsorbed on the photocatalyst surface, further degraded into intermediates via opening of central aromatic ring, followed by degradation of intermediates into CO2, SO42-, NH4+, and NO3-.57 When MB molecules are exposed to UV light irradiation in the presence of photocatalysts, it is observed that the doping improves the photocatalytic activity in comparison to undoped one. The concentration ratio (C/Co) vs. time graph for all the photocatalysts was obtained using time dependent study of MB degradation under UV light as shown in Figure 7a. According to Langmuir-Hinshelwood model with approximation, the kinetic rate constant with different photocatalyst was determined by following equation (4): 58

14 ACS Paragon Plus Environment

Page 15 of 32

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

C  ln o  = K app t C

(4)

where, Co is the initial concentration of MB, C is the concentration of MB with time, Kapp is the apparent kinetic rate constant and t is exposure time. The kinetic rate constants were calculated using ln(Co/C) vs. time graph, as shown in Figure 7c and tabulated in Table 1.

Table 1. Photocatalytic rate constant of different photocatalyst under UV light. Photocatalyst (1 mg of photocatalyst/mL of 20 μM MB)

-1

Rate constant (min ) (Under UV light)

TNF

0.00991

1% Ta TNF

0.01308

2% Ta TNF

0.02993

5% Ta TNF

0.05095

10% Ta TNF

0.02105

Interestingly, the photocatalytic activity of 5% Ta TNF sample is found to be increased by 5.1 times than the TNF. 5% Ta TNF show the best photocatalytic activity as compared to undoped and other doped TNF as summarized in Table 1. The other Ta doped TNF samples (less than 5% Ta doping) also show better photocatalytic activity as compared to the undoped sample. The photocatalytic activity is enhanced by 1.3 and 2.9 times for 1% Ta TNF and 2% Ta TNF, respectively as compared to TNF. When the Ta doping concentration is increased to 10%, the photocatalytic activity of the 10% Ta TNF sample is decreased slightly in comparison to the 5% Ta TNF sample. However, it show 2.1 times higher photocatalytic activity than the undoped TNF. The decrement of photocatalytic activity at higher Ta doping concentration (10% Ta TNF) may be due to the segregation of Ta2O5 as observed in XRD pattern (Figure 1a).

15 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

Table 2. Photocatalytic activity of different photocatalyst under solar light. -1

Photocatalyst (1 mg of photocatalyst/

Rate constant (min )

5 mL of 40 μM MB)

(Under solar light )

TNF

0.00502

1% Ta TNF

0.00569

2% Ta TNF

0.00983

5% Ta TNF

0.01121

10% Ta TNF

0.00761

1.2

TNF 1% Ta TNF 2% Ta TNF 5% Ta TNF 10% Ta TNF

0.8

0.6 0.4

0.6 0.4 0.2

0.2 0

20

40

60

0.0

Time (min)

4.0

c)

3.5 3.0 2.5

0

50

TNF 1% Ta TNF 2% Ta TNF 5% Ta TNF 10% Ta TNF

d)

2.5

2.0 1.5

100

150

200

250

200

250

Time (min)

3.0

ln(Co/C)

0.0

b)

1.0

C/Co

0.8

C/Co

1.2

TNF 1% Ta TNF 2% Ta TNF 5% Ta TNF 10% Ta TNF

a)

1.0

ln (Co/C)

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

TNF 1% Ta TNF 2% Ta TNF 5% Ta TNF 10% Ta TNF

2.0 1.5 1.0

1.0

0.5

0.5 0.0

0

20

40

Time (min)

60

0.0

0

50

100

150

Time (min)

Figure 7. Photocatalytic activity of TNF and Ta doped TNF photocatalyst for degradation of MB (a) under UV light (photocatalyst loading = 1 mg/mL of 20 μM MB), (b) under solar light (photocatalyst loading = 0.2 mg/mL of 40 μM MB); (c) and (d) Photocatalytic reaction kinetics of MB degradation under UV and solar light, respectively; Inset of (b) shows digital image of MB solution before and after degradation in the presence of 5% Ta TNF.

16 ACS Paragon Plus Environment

Page 17 of 32

In order to explore the effectiveness of TNF and Ta doped TNF on dye degradation, the photocatalytic degradation studies were also performed at low photocatalyst material loading (10 mg of photocatalyst /50 mL of 40 μM MB) under simulated solar light irradiation. The concentration ratio (C/Co) vs. time was obtained using time dependent MB degradation results (Figure 7b). The kinetic rate constants for different photocatalysts were calculated using ln(Co/C) vs. time graph (Figure 7d). The kinetic rate constants of different photocatalysts follow the order: 5% Ta TNF > 2% Ta TNF > 10% Ta TNF > 1% Ta TNF > TNF, as shown in Table 2. The photocatalytic degradation of MB is found to be enhanced by 2.2 times for 5% Ta TNF sample in comparison to undoped one. 100

80 94.7 %

a) 91.5 % 77.5 %

60

74.7 %

69.6 %

40 20 0

2nd 3rd 4th Number of cycle

5th

b) 80

96.2 % 80.7 % 73.1 %

60

71.0 %

68.2 %

40 20 0

1st

2nd 3rd 4th Number of cycle

e)

d)

20

5th

Before reaction After reaction under UV light After reaction under solar light

Intensity (a.u.)

c)

1st

Dye Degradation (%)

100 Dye Degradation (%)

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

30

40

50

60

70

80

Angle (2θ)

Figure 8. Reusability test of 5% Ta TNF for the degradation of MB (a) under UV light (reaction time = 1 h), (b) under solar light (reaction time = 4 h); FESEM image of 5% Ta TNF after 5 cycles of reuse, (c) under UV light, (d) under solar light; (e) XRD pattern of 5% Ta TNF before and after 5 cycles of reuse.

Figure 8 shows reusability test of the 5% Ta TNF towards the degradation of MB under UV and solar light over 5 cycles. All experiments were performed under the same conditions as mentioned above. After each cycle, the photocatalyst was removed by centrifugation and again dispersed in the same volume of dye for the next cycle. In Figure 8a, It can be observed 17 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

that there is not much loss in the photocatalytic activity in the first 2 cycles under UV light (reaction time = 1 h). However, a considerable loss is observed in subsequent 3rd cycle, while 4th and 5th cycle do not show a significant loss in photocatalytic activity. While a considerable loss in photocatalytic efficiency can be observed in 2nd, 3rd, 4th and 5th cycle under solar light (reaction time = 4 h) (Figure 8b). This observed loss in photocatalytic activity may be due to the lost catalyst during recovery by centrifugation and reduced catalytic surface area.59-60 Figures 8c-d show FESEM image of 5% Ta TNF after 5 cycles of the reusability test under UV and solar light. It is observed that there is no significant change in morphology of the nanofibers even after 5 repetitive usages. Figure 8e shows the XRD of 5% Ta TNF after reusability test which does not show any significant change in its crystal structure even after 5 cycles of reuse. FT-IR and Raman spectra of the 5% Ta TNF sample after 5 cycles of reuse have also been shown in Figure S9. FT-IR spectrum shows a few characteristics peaks (1316 cm-1, 1395 cm-1, 1496 cm-1) of MB (Figure S9a).61 Also, Raman spectra show the characteristics peaks of adsorbed MB molecules (Figure S9b) (discussed in section 3.3), which reduce the available catalytic sites for other incoming molecules. It can be concluded that MB molecules adsorb on the surface of photocatalyst and reduce the catalytic surface area for the other molecules, which decreases the photocatalytic activity with increase in the number of cycles.

18 ACS Paragon Plus Environment

Page 19 of 32

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

3.3

SERS detection of MB molecules SERS detection of MB molecules in 20 μM aqueous solution was performed using TNF

and 5% Ta TNF as 5% Ta doping is the optimum doping in TNF (as revealed by UV-vis absorption study) and also shows the best photocatalytic activity as discussed above. All Raman and SERS spectra were recorded using 532 nm laser, which rules out any contribution of resonant Raman Effect of MB because the absorbance peak of MB falls above 550 nm with characteristic absorbance peak at 664 nm as shown in Figure S8. Figure 9a shows the normal Raman spectra of the solid MB in powder form and MB aqueous solution along with the chemical structure of the MB. It is clear that the Raman signals of MB in aqueous solution are not so prominent because of the low concentration of MB molecules. The Raman spectrum of solid MB shows several Raman signals with characteristic peaks at 1153, 1396 and 1622 cm-1 assigned to symmetric in plane C-H bending, symmetric C-N stretching (and in plane C-H ring deformation) and C-C (and C-N-C) ring stretching mode, respectively.62-65 SERS activity of TNF and 5% Ta TNF was studied by Raman spectra of MB (aqueous solution) adsorbed on these TNF. Raman spectra below 1000 cm-1 have not been shown to avoid any overlapping with TiO2 Raman modes. Figure 9b shows the SERS spectra of MB molecules adsorbed on the TNF and 5% Ta TNF, where each peak corresponds to a characteristic SERS signals of the adsorbed MB molecules. The SERS and adsorption of MB molecules can be understood by shift in the Raman peak position and enhancement in the Raman intensity. Moreover, it can be observed from the SERS spectra (Figure 9b) that there is a shift in Raman peak positions as well as change in their intensities as compared to the normal Raman spectra of solid MB (Figure 9a). As can be compared in Figure 9, that the main characteristic Raman peaks at 1396 and 1622 cm-1 in normal Raman spectra are shifted to 1390 and 1634 cm-1 in SERS spectra. This indicates SERS enhancement of MB molecules adsorbed on TNF and 5% Ta TNF. The major Raman shifts in normal Raman and SERS spectra with proper assignments of the different Raman modes of MB molecule have been given in Table 3.

19 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

Figure 9. (a) Normal Raman spectra of solid MB powder and MB aqueous solution, and (b) SERS spectra of MB molecules adsorbed on TNF and 5% Ta TNF.

Table 3. Normal and SERS Raman shifts and their assignments for MB molecules.62-65 Normal (cm-1)

Raman peak assignments

SERS (cm-1)

1035

1040

in plane C-H bending

1070

1084

symmetric C-N stretching

1153

1157

in plane C-H bending

1301

1295

C-C ring stretching

1396

1390

symmetric C-N stretching

1436

asymmetric C-N stretching

1467

C-C ring stretching

1528

asymmetric C-C stretching

1634

C-C (and C-N-C) ring stretching

1469

1622

20 ACS Paragon Plus Environment

Page 21 of 32

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

Interestingly, very strong and high intensity characteristic Raman peaks (1040, 1153, 1390 and 1634 cm-1) of adsorbed MB molecules on 5% Ta TNF can observed in addition to some new Raman peaks (1335, 1436 and 1528 cm-1). It demonstrates the enhanced SERS activity of 5% Ta TNF as compared to TNF. These results indicate that Ta doping plays an important role in the SERS enhancement of MB adsorbed of TNF. SERS enhancement is generally described as either chemical enhancement or electromagnetic enhancement,6,

8, 66

which are responsible for the shift in Raman peak position and change in intensity as observed in the present case of MB adsorbed on TNF and 5% Ta TNF. The electromagnetic SERS enhancement takes place as a result of coupling of surface plasmon resonance (SPR) of noble metals with excitation wavelength. Therefore, the possibility of electromagnetic enhancement is ruled out in case of TiO2, however, in such metal oxide semiconductors, the surface plasmon resonant frequency falls in IR region, while the excitation source used is 532 nm.67 The chemical SERS enhancement mainly takes place due to the charge transfer (CT) between the adsorbed molecule and semiconductor surfaces.67 In the present case, shift in Raman peak positions, intensity improvement and splitting of peak (1469 cm-1) are therefore resulted due to chemisorption of the MB molecules on semiconductor based TNF and 5% Ta TNF. These observations are also consistent with other results for MB adsorbed on TiO2 surfaces and are explained on the basis of chemical SERS enhancement due to CT mechanism from TiO2 to MB molecules62-63 as the detailed mechanism is given in the next section. 3.4 Mechanism of enhanced photocatalytic performance and SERS detection The photocatalytic property as well as the SERS activity of TiO2 based nanomaterials are determined by several material properties such as energy band gap, specific surface area, crystallinity and their surface structure. These properties can be modified and improved for the particular applications by many ways. In the present work, we have reported the improvement in such properties of TNF by Ta doping and observed their enhanced photocatalytic degradation and SERS performance towards MB molecules. As observed in XRD analysis, the crystalline size of (101) plane is decreased with the Ta doping, thus helping in photocatalytic activity. BET surface area of 5% Ta doped TNF sample is slightly higher than the undoped one, thus offering more MB molecule to adsorb on the surface and hence enhancing photocatalytic and SERS activity. Therefore, it can be predicted that combined effect of these enhanced properties of the materials (Ta doped TNF) is responsible for the enhancement photocatalytic and SERS activity. As the Ta5+ ions have similar ionic 21 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

radius to Ti4+ ions, therefore, it can easily replace few Ti4+, which causes increased electron concentration within TNF matrix along with replacement of few Ti4+ with Ti3+ sites. The created Ti3+ defects introduce new energy levels below the conduction band, which enhance the visible light absorption.68 These electrons can also moderately fill the surface traps, thereby shifting the Fermi level to more negative potential. The enhanced visible light absorption and photocatalytic activity is due to Ti3+ sites present in TNF matrix. Increasing doping concentration of Ta increases the Ti3+ defects leads to the stronger optical absorption and thus higher photocatalytic activity. However, higher doping can create higher defects, which can act as charge carrier recombination sites.31

Figure 10. Schematic of mechanism of enhanced photocatalytic degradation and SERS performance of Ta doped TNF.

Moreover, it can be explained as Ta doping induces new energy levels below the conduction band of TiO2 due to Ti3+ defects, which inhibit the photoexcited charge recombination acting as electron traps and also promote CT mechanism acting as an 22 ACS Paragon Plus Environment

Page 23 of 32

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

intermediate state for TiO2 to MB molecule electron transfer, and are mainly responsible for the enhanced photocatalytic and SERS activities, respectively as shown schematically in Figure 10. Since, TiO2 has wide bandgap and excitation with 532 nm laser would not excite the valance band electrons to be transferred to conduction band. As in the case of noble metal nanostructures, such excitation is important for resonance condition and for electromagnetic SERS enhancement which is governed by surface plasmon resonance on noble metal nanostructures. However, in case of semiconductors (i.e TiO2), such resonance condition is not essential for the SERS enhancement, on the other hand, chemical enhancement as a result of charge interaction between the TiO2 surface and the adsorbed molecules is responsible for the SERS activity. For example: Leon et al.69 reported that photo-induced CT processes mainly occurred between the adsorbed molecules and the semiconductor substrates and SERS enhancement can be taken place without the resonance conditions. They performed the SERS experiments at different excitation wavelengths (532 and 632.8 nm) and observed that the SERS signals of a N719 dye molecules adsorbed on TiO2 nanoparticle could be obtained without the resonance condition at the excitation wavelength of 632.8 nm. In another experiment work by Yang et al.14 SERS effect of TiO2 and a series of silver-deposited TiO2 nanoparticles were studied using a Raman with excitation wavelength of 514.5 nm. It was observed that pure TiO2 nanoparticles showed SERS signals of 4-MBA molecules adsorbed on its surface which was attributed to the intrinsic TiO2-to-molecule CT mechanism. Furthermore, these SERS signals could be enhanced on Ag-TiO2 nanocomposite nanoparticles as compared to those on pure TiO2 nanoparticles due to the synergic contribution of both CT and surface plasmon resonance of Ag nanoparticles. Similarly, the same group reported that metal ion doping energy levels could be created between the CB and VB of the TiO2 nanoparticles by metal ion doping which facilitated the CT from TiO2 to molecules resulting in the enhanced SERS signals of the adsorbed molecules.70

In the present case, it can be explained that the SERS spectra of a MB molecule adsorbed on undoped TNF and Ta doped TNF have been obtained without the resonance condition and by means of chemical SERS enhancement through CT mechanism. In case of Ta doped TNF, the energy levels are introduced below the conduction band of TiO2 due to the Ti3+ defects created by Ta doping. Under the laser excitation, the excited electrons can be transferred to these Ti3+ defect energy levels and then transit to the lowest unoccupied molecular orbital (LUMO) level of the chemisorbed MB molecules on doped TNF surface thus facilitating the 23 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

CT process. This CT mechanism provides an additional Ta doped TNF to molecule CT besides the undoped TNF-to-molecule CT, and is responsible for the enhanced SERS enhancement. The present work motivates for synthesizing one dimensional multifunctional novel nanomaterials.

4.

Conclusions

TNF and Ta doped TNF were successfully prepared by electrospinning technique. The prepared nanofibers were continuous, solid and uniform throughout the matrix. Undoped and Ta doped TNF sample were mesoporous in nature indicating positive effect for photocatalysis application. Ta doping in TNF slightly decreased its crystalline size due to the incorporation of Ta5+ within TiO2 matrix which restricted crystal growth and hence enhanced photocatalytic activity. Ta doping showed enhancement of visible light absorption due to the presence of Ti3+ states created by Ta5+ ions doping. 5% Ta doped showed higher photocatalytic activity among all the photocatalysts due to the combined effect of low crystalline size, higher surface area and enhanced visible light absorption. The 5% Ta TNF showed 5.1 and 2.2 times higher photocatalytic activity in comparison to TNF under UV and solar light irradiation, respectively. The 5% Ta TNF also showed the enhanced SERS functionality as compared to the TNF in detection of MB molecules. The observed results are explained on the basis of the Ta doping induced new energy levels below the conduction band of TiO2 due to Ti3+ defects. These Ti3+ defects energy levels play an important role in improving the photoexcited charge separation acting as electron traps and promote CT mechanism acting as an intermediate state for TiO2 to MB electron transfer, and are mainly responsible for the enhanced photocatalytic and SERS activities, respectively. The material synthesized here is also likely to be useful in other applications such as water splitting, solar cell and sensors etc.

ASSOCIATED CONTENT Supporting Information Kubelka–Munk function vs. energy graph for 1%, 2% and 10% Ta TNF, Low magnified FESEM micrographs and diameter distribution of doped and undoped TNF, EDX of Ta doped and undoped TNF, TEM, FT-IR, adsorption study with different photocatalysts, FT-IR spectrum and Raman spectra of 5% Ta TNF before and after 5 cycles of reuse (under UV and solar light irradiation). This material is available free of charge.

24 ACS Paragon Plus Environment

Page 25 of 32

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

AUTHOR INFORMATION Corresponding Author *E-mail: [email protected]

Notes The authors declare no competing financial interest. Acknowledgment RKG acknowledges financial assistance from Department of Science and Technology (DST), India, through the INSPIRE Faculty Award (Project No. IFA-13 ENG-57) and Grant No. DST/TM/WTI/2K16/23(G). JP acknowledges DST, India for the prestigious INSPIRE faculty award (INSPIRE/04/2015/002452). DST support to the Center for Nanosciences is acknowledged.

References 1.

Konstantinou, I. K.; Albanis, T. A. Photocatalytic Transformation of Pesticides in

Aqueous Titanium Dioxide Suspensions using Artificial and Solar Light: Intermediates and Degradation Pathways. Appl. Catal. B: Environ. 2003, 42, 319-335. 2.

Chong, M. N.; Jin, B.; Chow, C. W. K.; Saint, C. Recent Developments in

Photocatalytic Water Treatment Technology: A Review. Water Res. 2010, 44, 2997-3027. 3.

Kazuhito, H.; Hiroshi, I.; Akira, F. TiO2 Photocatalysis: A Historical Overview and

Future Prospects. Jpn. J. Appl. Phys. 2005, 44, 8269. 4.

Khan, S. U. M.; Al-Shahry, M.; Ingler, W. B. Efficient Photochemical Water Splitting

by a Chemically Modified n-TiO2. Science 2002, 297, 2243-2245. 5.

Ben-Jaber, S.; Peveler, W. J.; Quesada-Cabrera, R.; Cortés, E.; Sotelo-Vazquez, C.;

Abdul-Karim, N.; Maier, S. A.; Parkin, I. P. Photo-Induced Enhanced Raman Spectroscopy for Universal Ultra-Trace Detection of Explosives, Pollutants and Biomolecules. Nat. Commun. 2016, 7, 12189. 6.

Jai, P.; Promod, K.; Harris, R. A.; Chantel, S.; Neethling, J. H.; Vuuren, A. J. v.;

Swart, H. C. Synthesis, Characterization and Multifunctional Properties of Plasmonic Ag– TiO2 Nanocomposites. Nanotechnology 2016, 27, 355707.

25 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

7.

Page 26 of 32

Tyagi, A.; Tripathi, K. M.; Singh, N.; Choudhary, S.; Gupta, R. K. Green Synthesis of

Carbon Quantum Dots from Lemon Peel Waste: Applications in Sensing and Photocatalysis. RSC Adv. 2016, 6, 72423-72432. 8.

Prakash, J.; Harris, R. A.; Swart, H. C. Embedded Plasmonic Nanostructures:

Synthesis, Fundamental Aspects and Their Surface Enhanced Raman Scattering Applications. Int. Rev. Phys. Chem. 2016, 35, 353-398. 9.

Kumar, S.; Lodhi, D. K.; Singh, J. P. Highly Sensitive Multifunctional Recyclable

Ag-TiO2 Nanorod SERS Substrates for Photocatalytic Degradation and Detection of Dye Molecules. RSC Adv. 2016, 6, 45120-45126. 10.

Kumar, S. G.; Devi, L. G. Review on Modified TiO2 Photocatalysis under UV/Visible

Light: Selected Results and Related Mechanisms on Interfacial Charge Carrier Transfer Dynamics. J. Phys. Chem. A 2011, 115, 13211-13241. 11.

Pelaez, M.; Nolan, N. T.; Pillai, S. C.; Seery, M. K.; Falaras, P.; Kontos, A. G.;

Dunlop, P. S. M.; Hamilton, J. W. J.; Byrne, J. A.; O'Shea, K.; Entezari, M. H.; Dionysiou, D. D. A Review on the Visible Light Active Titanium Dioxide Photocatalysts for Environmental Applications. Appl. Catal. B: Environ. 2012, 125, 331-349. 12.

Ramacharyulu, P. V. R. K.; Praveen kumar, J.; Prasad, G. K.; Srivastava, A. R.

Synthesis, Characterization and Photocatalytic Activity of Ag-TiO2 Nanoparticulate Film. RSC Adv. 2015, 5, 1309-1314. 13.

Misra, M.; Singh, N.; Gupta, R. K. Enhanced Visible-Light-Driven Photocatalytic

Activity of Au@Ag Core-Shell Bimetallic Nanoparticles Immobilized on Electrospun TiO2 Nanofibers for Degradation of Organic Compounds. Catal. Sci. Technol. 2017, 7, 570-580. 14.

Yang, L.; Jiang, X.; Ruan, W.; Yang, J.; Zhao, B.; Xu, W.; Lombardi, J. R. Charge-

Transfer-Induced Surface-Enhanced Raman Scattering on Ag−TiO2 Nanocomposites. J. Phys. Chem. C 2009, 113, 16226-16231. 15.

Singh, N.; Mondal, K.; Misra, M.; Sharma, A.; Gupta, R. K. Quantum Dot Sensitized

Electrospun Mesoporous Titanium Dioxide Hollow Nanofibers for Photocatalytic Applications. RSC Adv. 2016, 6, 48109-48119. 16.

Nolan, N. T.; Synnott, D. W.; Seery, M. K.; Hinder, S. J.; Van Wassenhoven, A.;

Pillai, S. C. Effect of N-Doping on the Photocatalytic Activity of Sol–Gel TiO2. J. Hazard. Mater. 2012, 211–212, 88-94. 17.

Etacheri, V.; Seery, M. K.; Hinder, S. J.; Pillai, S. C. Nanostructured Ti1-xSxO2-yNy

Heterojunctions for Efficient Visible-Light-Induced Photocatalysis. Inorg. Chem. 2012, 51, 7164-7173. 26 ACS Paragon Plus Environment

Page 27 of 32

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

18.

Yang, G.; Yan, Z.; Xiao, T. Low-Temperature Solvothermal Synthesis of Visible-

Light-Responsive S-Doped TiO2 Nanocrystal. Appl. Surf. Sci. 2012, 258, 4016-4022. 19.

Etacheri, V.; Michlits, G.; Seery, M. K.; Hinder, S. J.; Pillai, S. C. A Highly Efficient

TiO2–xCx Nano-Heterojunction Photocatalyst for Visible Light Induced Antibacterial Applications. ACS Appl. Mater. Interfaces 2013, 5, 1663-1672. 20.

Su, R.; Bechstein, R.; Kibsgaard, J.; Vang, R. T.; Besenbacher, F. High-Quality Fe-

Doped TiO2 Films with Superior Visible-Light Performance. J. Mater. Chem. 2012, 22, 23755-23758. 21.

Khakpash, N.; Simchi, A.; Jafari, T. Adsorption and Solar Light Activity of

Transition-Metal Doped TiO2 Nanoparticles as Semiconductor Photocatalyst. J. Mater. Sci.: Mater. Electron. 2012, 23, 659-667. 22.

Nirmala, R.; Kim, H. Y.; Yi, C.; Barakat, N. A. M.; Navamathavan, R.; El-Newehy,

M. Electrospun Nickel Doped Titanium Dioxide Nanofibers as an Effective Photocatalyst for the Hydrolytic Dehydrogenation of Ammonia Borane. Int. J. Hydrogen 2012, 37, 1003610045. 23.

Sajjad, A. K. L.; Shamaila, S.; Zhang, J. Study of New States in Visible Light Active

W, N Co-Doped TiO2 Photo catalyst. Mater. Res. Bull. 2012, 47, 3083-3089. 24.

Wang, Y. J.; Yang, M.; Ren, L. Z.; Zhou, W. Q.; Yang, K. G.; Yu, F. M.; Meng, M.;

Wu, S. X.; Li, S. W. Enhanced Raman Scattering in Copper-doped TiO2 films. Thin Solid Films 2016, 598, 311-314. 25.

Yang, L.; Zhang, Y.; Ruan, W.; Zhao, B.; Xu, W.; Lombardi, J. R. Improved Surface-

Enhanced Raman Scattering Properties of TiO2 Nanoparticles by Zn Dopant. J. Raman Spectrosc. 2010, 41, 721-726. 26.

Xue, X.; Ji, W.; Mao, Z.; Li, Z.; Ruan, W.; Zhao, B.; Lombardi, J. R. Effects of Mn

Doping on Surface Enhanced Raman Scattering Properties of TiO2 Nanoparticles. Spectrochim. Acta Mol. Biomol. Spectrosc. 2012, 95, 213-217. 27.

Jiang, X.; Song, K.; Li, X.; Yang, M.; Han, X.; Yang, L.; Zhao, B. Double Metal Co–

Doping of TiO2 Nanoparticles for Improvement of their SERS Activity and Ultrasensitive Detection of Enrofloxacin: Regulation Strategy of Energy Levels. ChemistrySelect 2017, 2, 3099-3105. 28.

Anh Huy, H.; Aradi, B.; Frauenheim, T.; Deák, P. Comparison of Nb- and Ta-Doping

of Anatase TiO2 for Transparent Conductor Applications. J. Appl. Phys. 2012, 112, 016103.

27 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

29.

Page 28 of 32

Taro, H.; Yutaka, F.; Atsuki, U.; Kinnosuke, I.; Kazuhisa, I.; Yasushi, H.; Go, K.;

Yukio, Y.; Toshihiro, S.; Tetsuya, H. Ta-Doped Anatase TiO2 Epitaxial Film as Transparent Conducting Oxide. Jpn. J. Appl. Phys. 2005, 44, L1063. 30.

Wang, C.; Geng, A.; Guo, Y.; Jiang, S.; Qu, X.; Li, L. A Novel Preparation of Three-

Dimensionally Ordered Macroporous M/Ti (M = Zr or Ta) Mixed Oxide Nanoparticles with Enhanced Photocatalytic Activity. J. Colloid Interface Sci. 2006, 301, 236-247. 31.

Znad, H.; Ang, M. H.; Tade, M. O. Ta/TiO2-and Nb/TiO2-Mixed Oxides as Efficient

Solar Photocatalysts: Preparation, Characterization, and Photocatalytic Activity. Int. J. Photoenergy 2012, 2012, 548158. 32.

Bawaked, S. M.; Sathasivam, S.; Bhachu, D. S.; Chadwick, N.; Obaid, A. Y.; Al-

Thabaiti, S.; Basahel, S. N.; Carmalt, C. J.; Parkin, I. P. Aerosol Assisted Chemical Vapor Deposition of Conductive and Photocatalytically Active Tantalum Doped Titanium Dioxide Films. J. Mater. Chem. A 2014, 2, 12849-12856. 33.

Fang, H.; Zhang, C. X.; Liu, L.; Zhao, Y. M.; Xu, H. J. Recyclable Three-

Dimensional Ag Nanoparticle-Decorated TiO2 Nanorod Arrays for Surface-Enhanced Raman Scattering. Biosens. Bioelectron. 2015, 64, 434-441. 34.

Ling, Y.; Zhuo, Y.; Huang, L.; Mao, D. Using Ag-Embedded TiO2 Nanotubes Array

as Recyclable SERS Substrate. Appl. Surf. Sci. 2016, 388, Part A, 169-173. 35.

Nakada, A.; Nishioka, S.; Vequizo, J. J. M.; Muraoka, K.; Kanazawa, T.; Yamakata,

A.; Nozawa, S.; Kumagai, H.; Adachi, S.-i.; Ishitani, O.; Maeda, K. Solar-Driven Z-Scheme Water Splitting using Tantalum/Nitrogen Co-Doped Rutile Titania Nanorod as an Oxygen Evolution Photocatalyst. J. Mater. Chem. A 2017, 5, 11710-11719. 36.

Lee, K.; Mazare, A.; Schmuki, P. One-Dimensional Titanium Dioxide Nanomaterials:

Nanotubes. Chem. Rev. 2014, 114, 9385-9454. 37.

Wang, X.; Li, Z.; Shi, J.; Yu, Y. One-Dimensional Titanium Dioxide Nanomaterials:

Nanowires, Nanorods, and Nanobelts. Chem. Rev. 2014, 114, 9346-9384. 38.

Wu, W.-Q.; Lei, B.-X.; Rao, H.-S.; Xu, Y.-F.; Wang, Y.-F.; Su, C.-Y.; Kuang, D.-B.

Hydrothermal Fabrication of Hierarchically Anatase TiO2 Nanowire arrays on FTO Glass for Dye-sensitized Solar Cells. Sci. Rep. 2013, 3, 1352. 39.

Si, P.; Ding, S.; Yuan, J.; Lou, X. W.; Kim, D.-H. Hierarchically Structured One-

Dimensional TiO2 for Protein Immobilization, Direct Electrochemistry, and Mediator-Free Glucose Sensing. ACS Nano 2011, 5, 7617-7626.

28 ACS Paragon Plus Environment

Page 29 of 32

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

40.

Zhang, Z.; Zhang, L.; Hedhili, M. N.; Zhang, H.; Wang, P. Plasmonic Gold

Nanocrystals Coupled with Photonic Crystal Seamlessly on TiO2 Nanotube Photoelectrodes for Efficient Visible Light Photoelectrochemical Water Splitting. Nano Lett. 2013, 13, 14-20. 41.

Chuangchote, S.; Jitputti, J.; Sagawa, T.; Yoshikawa, S. Photocatalytic Activity for

Hydrogen Evolution of Electrospun TiO2 Nanofibers. ACS Appl. Mater. Interfaces 2009, 1, 1140-1143. 42.

Kumar, P. S.; Sundaramurthy, J.; Sundarrajan, S.; Babu, V. J.; Singh, G.;

Allakhverdiev, S. I.; Ramakrishna, S. Hierarchical Electrospun Nanofibers for Energy Harvesting, Production and Environmental Remediation. Energy Environ. Sci. 2014, 7, 31923222. 43.

Baiju, K. V.; Shajesh, P.; Wunderlich, W.; Mukundan, P.; Kumar, S. R.; Warrier, K.

G. K. Effect of Tantalum Addition on Anatase Phase Stability and Photoactivity of Aqueous Sol–Gel Derived Mesoporous Titania. J. Mol. Catal. A: Chem. 2007, 276, 41-46. 44.

Zhang, Z.; Shao, C.; Zhang, L.; Li, X.; Liu, Y. Electrospun Nanofibers of V-Doped

TiO2 with High Photocatalytic Activity. J. Colloid Interface Sci. 2010, 351, 57-62. 45.

Liu, B.; Peng, L. Facile Formation of Mixed Phase Porous TiO2 Nanotubes and

Enhanced Visible-Light Photocatalytic Activity. J. Alloys Compd. 2013, 571, 145-152. 46.

Yang, M.; Hume, C.; Lee, S.; Son, Y.-H.; Lee, J.-K. Correlation between

Photocatalytic Efficacy and Electronic Band Structure in Hydrothermally Grown TiO2 Nanoparticles. J. Phys. Chem. C 2010, 114, 15292-15297. 47.

López, R.; Gómez, R. Band-Gap Energy Estimation from Diffuse Reflectance

Measurements on Sol–Gel and Commercial TiO2: A Comparative Study. J. Sol-Gel Sci. Technol. 2012, 61, 1-7. 48.

Liu, J.; Yang, H.; Tan, W.; Zhou, X.; Lin, Y. Photovoltaic Performance Improvement

of Dye-Sensitized Solar Cells Based on Tantalum-Doped TiO2 Thin Films. Electrochim. Acta 2010, 56, 396-400. 49.

Fang, W. Q.; Zhou, J. Z.; Liu, J.; Chen, Z. G.; Yang, C.; Sun, C. H.; Qian, G. R.; Zou,

J.; Qiao, S. Z.; Yang, H. G. Hierarchical Structures of Single-Crystalline Anatase TiO2 Nanosheets Dominated by {001} Facets. Chem. Eur. J. 2011, 17, 1423-1427. 50.

Chu, D.; Yuan, X.; Qin, G.; Xu, M.; Zheng, P.; Lu, J.; Zha, L. Efficient Carbon-doped

Nanostructured TiO2 (Anatase) Film for Photoelectrochemical Solar Cells. J. Nanopart. Res. 2008, 10, 357-363.

29 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

51.

Page 30 of 32

Gonçalves, R. V.; Wojcieszak, R.; Uberman, P. M.; Teixeira, S. R.; Rossi, L. M.

Insights into the Active Surface Species Formed on Ta2O5 Nanotubes in the Catalytic Oxidation of CO. PCCP 2014, 16, 5755-5762. 52.

Sing, K. S. W.; Everett, D. H.; Haul, R. A. W.; Moscou, L.; Pierotti, R. A.;

RouqueroL, J.; Siemieniewska, T. Reporting Physisorption Data for Gas, Solid Systems with Special Reference to the Determination of Surface Area and Porosity (recommendations, 1984). Pure Appl. Chem. 1985, 57, 603–619. 53.

Liu, S.; Wang, W.; Chen, J.; Li, J.-G.; Li, X.; Sun, X.; Dong, Y. Foamed Single-

Crystalline Anatase Nanocrystals Exhibiting Enhanced Photocatalytic Activity. J. Mater. Chem. A 2015, 3, 17837-17848. 54.

Moulton, S. E.; Barisci, J. N.; McQuillan, A. J.; Wallace, G. G. ATR-IR

Spectroscopic Studies of the Influence of Phosphate Buffer on Adsorption of Immunoglobulin G to TiO2. "Colloids Surf., A" 2003, 220, 159-167. 55.

Liu, C.; Miao, L.; Zhou, J.; Huang, R.; Tanemura, S. Bottom-up Assembly to Ag

Nanoparticles Embedded Nb-Doped TiO2 Nanobulks with Improved n-type Thermoelectric Properties. J. Mater. Chem. 2012, 22, 14180-14190. 56.

Dharmaraj, N.; Park, H. C.; Kim, C. H.; Viswanathamurthi, P.; Kim, H. Y. Nanometer

Sized Tantalum Pentoxide Fibers Prepared by Electrospinning. Mater. Res. Bull. 2006, 41, 612-619. 57.

Houas, A.; Lachheb, H.; Ksibi, M.; Elaloui, E.; Guillard, C.; Herrmann, J.-M.

Photocatalytic Degradation Pathway of Methylene Blue in Water. Appl. Catal. B: Environ. 2001, 31, 145-157. 58.

Kumar, K. V.; Porkodi, K.; Rocha, F. Langmuir–Hinshelwood Kinetics – A

Theoretical Study. Catal. Commun. 2008, 9, 82-84. 59.

Harish, S.; Navaneethan, M.; Archana, J.; Silambarasan, A.; Ponnusamy, S.;

Muthamizhchelvan, C.; Hayakawa, Y. Controlled Synthesis of Organic Ligand Passivated ZnO Nanostructures and Their Photocatalytic Activity under Visible Light Irradiation. Dalton Trans. 2015, 44, 10490-10498. 60.

Ghosh Chaudhuri, R.; Paria, S. Visible Light Induced Photocatalytic Activity of

Sulfur Doped Hollow TiO2 Nanoparticles, Synthesized via a Novel Route. Dalton Trans. 2014, 43, 5526-5534. 61.

Horvath, E.; Ribic, P. R.; Hashemi, F.; Forro, L.; Magrez, A. Dye Metachromasy on

Titanate Nanowires: Sensing Humidity with Reversible Molecular Dimerization. J. Mater. Chem. 2012, 22, 8778-8784. 30 ACS Paragon Plus Environment

Page 31 of 32

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

62.

Srichan, C.; Ekpanyapong, M.; Horprathum, M.; Eiamchai, P.; Nuntawong, N.;

Phokharatkul, D.; Danvirutai, P.; Bohez, E.; Wisitsoraat, A.; Tuantranont, A. HighlySensitive Surface-Enhanced Raman Spectroscopy (SERS)-based Chemical Sensor using 3D Graphene Foam Decorated with Silver Nanoparticles as SERS substrate. Sci. Rep. 2016, 6, 23733. 63.

Xiao, G.-N.; Man, S.-Q. Surface-Enhanced Raman Scattering of Methylene Blue

Adsorbed on Cap-Shaped Silver Nanoparticles. Chem. Phys. Lett. 2007, 447, 305-309. 64.

Naujok, R. R.; Duevel, R. V.; Corn, R. M. Fluorescence and Fourier Transform

Surface-Enhanced Raman Scattering Measurements of Methylene Blue Adsorbed onto a Sulfur-Modified Gold Electrode. Langmuir 1993, 9, 1771-1774. 65.

Nicolai, S. l. H. d. A.; Rodrigues, P. R. P.; Agostinho, S. M. L.; Rubim, J. C.

Electrochemical and Spectroelectrochemical (SERS) Studies of the Reduction of Methylene Blue on a Silver Electrode. J. Electroanal. Chem. 2002, 527, 103-111. 66.

Prakash, J.; Kumar, V.; Kroon, R. E.; Asokan, K.; Rigato, V.; Chae, K. H.; Gautam,

S.; Swart, H. C. Optical and Surface Enhanced Raman Scattering Properties of Au Nanoparticles Embedded in and Located on a Carbonaceous Matrix. PCCP 2016, 18, 24682480. 67.

Yang, L.; Jiang, X.; Ruan, W.; Zhao, B.; Xu, W.; Lombardi, J. R. Observation of

Enhanced Raman Scattering for Molecules Adsorbed on TiO2 Nanoparticles: ChargeTransfer Contribution. J. Phys. Chem. C 2008, 112, 20095-20098. 68.

Zuo, F.; Bozhilov, K.; Dillon, R. J.; Wang, L.; Smith, P.; Zhao, X.; Bardeen, C.; Feng,

P. Active Facets on Titanium(III)-Doped TiO2: An Effective Strategy to Improve the VisibleLight Photocatalytic Activity. Angew. Chem. Int. Ed. 2012, 51, 6223-6226. 69.

Pérez León, C.; Kador, L.; Peng, B.; Thelakkat, M. Characterization of the Adsorption

of Ru-bpy Dyes on Mesoporous TiO2 Films with UV−Vis, Raman, and FTIR Spectroscopies. J. Phys. Chem. B 2006, 110, 8723-8730. 70.

Yang, L.; Qin, X.; Gong, M.; Jiang, X.; Yang, M.; Li, X.; Li, G. Improving Surface-

Enhanced Raman Scattering Properties of TiO2 Nanoparticles by Metal Co Doping. Spectrochim. Acta Mol. Biomol. Spectrosc. 2014, 123, 224-229.

31 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

TOC

32 ACS Paragon Plus Environment