You are on page 1of 41

Journal Pre-proof

Piezo-phototronic and plasmonic effect coupled Ag-NaNbO3 nanocomposite for


enhanced photocatalytic and photoelectrochemical water splitting activity

Dheeraj Kumar, Surbhi Sharma, Neeraj Khare

PII: S0960-1481(20)31556-1
DOI: https://doi.org/10.1016/j.renene.2020.09.132
Reference: RENE 14281

To appear in: Renewable Energy

Received Date: 27 July 2020


Revised Date: 25 September 2020
Accepted Date: 28 September 2020

Please cite this article as: Kumar D, Sharma S, Khare N, Piezo-phototronic and plasmonic effect
coupled Ag-NaNbO3 nanocomposite for enhanced photocatalytic and photoelectrochemical water
splitting activity, Renewable Energy (2020), doi: https://doi.org/10.1016/j.renene.2020.09.132.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.


Credit authorship contribution statement

Dheeraj Kumar: Conceptualization, Methodology, Formal Analysis, Validation

Investigation, Writing – original draft. Surbhi Sharma: Validation, Investigation, Writing –

review & editing. Neeraj Khare: Conceptualization, Methodology, Validation, Supervision,

Writing – review & editing.

of
ro
-p
re
lP
na
ur
Jo
Graphical Abstract

of
ro
-p
re
lP
na
ur
Jo
1 Piezo-Phototronic and Plasmonic Effect Coupled Ag-NaNbO3 Nanocomposite for

2 Enhanced Photocatalytic and Photoelectrochemical Water Splitting Activity

3 Dheeraj Kumar, Surbhi Sharma and Neeraj Khare*

4 Physics Department, Indian Institute of Technology Delhi, Haus Khaz, New Delhi-110016,

5 India.

6 Abstract

7 Silver (Ag) nanoparticles decorated NaNbO3 nanorods (Ag-NaNbO3) based nanocomposite

8 have been successfully synthesized by simple chemical solution method with the aim to

of
9 couple the plasmonic and piezo-phototronic effect. The Ag-NaNbO3 nanocomposite showed

ro
10 much enhanced photocatalytic and photoelectrochemical water splitting properties as

11
-p
compared to bare NaNbO3. A ~10 fold enhancement in the photodecomposition of organic
re
12 MB dye and a ~9 fold increment in the photocurrent density of photoelectrochemical water
lP

13 splitting was observed as compared to bare NaNbO3, which has been attributed to the
na

14 combined plasmonic and piezo-phototronic effect. The plasmonic effect due to the presence

15 of Ag nanoparticles on the NaNbO3 surface resulted in enhanced absorption of visible light


ur

16 and piezo-photoelectric effect resulted in better separation of the photogenerated charges due
Jo

17 to the built-in electric field. This approach demonstrates a novel strategy for enhancing the

18 performance of silver decorated semiconducting/piezoelectric material for achieving efficient

19 photocatalytic dye degradation and PEC water splitting.

20 Keywords: Ag-NaNbO3; Plasmonic effect; Piezo-phototronic effect; Piezo-photocatalytic

21 effect; PEC Water Splitting.

22 *Corresponding Author’s E-mail: nkhare@physics.iitd.ernet.in

23

24

25

1
26 1. Introduction

27 Semiconductor photocatalysts have shown potentiality towards environmental remediation

28 and hydrogen generation [1-3]. Significant research activities have been focused on exploring

29 new photocatalysts materials for photodegradation of organic dyes and photoelectrochemical

30 (PEC) water splitting [4, 5]. Although, several nanostructured oxide semiconductors such as

31 Fe2O3, WO3, CuO, MnO2, and IrO2 have been used as photocatalyst material for organic

32 compound degradation and PEC water splitting due to appropriate absorbance of light and

33 efficient separation of the photogenerated charge carriers [6-15]. But most of these oxide

of
34 semiconductor materials are not effective because of the faster recombination rate of charge

ro
35 carriers and inactively in the visible region.

36
-p
An interesting phenomenon “Piezo-phototronic effect” firstly introduced in 2010 has
re
37 attracted enormous research interests [16]. Piezo-phototronic effect mainly focuses on the
lP

38 coupling of optical, electrical and mechanical properties in ferro or piezo-electric


na

39 semiconductor materials [17]. It has been proven that mechanical force induced localized

40 polarization charges can effectively control or modulate the charge carrier generation,
ur

41 separation and migration by the local electric field distribution in the area of homojunction or
Jo

42 heterojunction systems [18, 19]. Thus, by virtue of the piezo-phototronic effect faster

43 recombination rate of charge carriers can be inhibited. The capability of the piezo-phototronic

44 effect has recently been reported for achieving enhanced performance in various applications,

45 for instance, in solar-cell [20], electroluminescence [21], photoelectrochemistry [22],

46 photodetectors [23] and optoelectronic devices [24].

47 Among other strategies, the coupling of semiconductor with various noble-metals can

48 also be an effective method to enhance the photocatalytic performance in the visible light

49 region. The presence of a noble-metals such as gold (Au), silver (Ag), and platinum (Pt)

50 metals on the surface of semiconductor can help in absorbing a wide spectrum of light

2
51 (visible range) due to resonant oscillation of allowed surface plasmon mediated electrons

52 induced by light, termed as plasmonic effect [25-27].

53 In the search of new material for photocatalytic and PEC application, sodium niobate

54 (NaNbO3) having a chemical formula ABO3 (where, A = Sr, Ba, Bi, Na and B = Ti, Nb, Fe)

55 can be a good alternative to conventional ZnO and BaTiO3 materials due to its non-toxicity,

56 low-cost, excellent stability in the aqueous solution and suitable band edge positions [28, 29].

57 Due to the unique non-centrosymmetrical crystalline structure, NaNbO3 semiconductor also

58 possesses piezoelectric properties. Recently, it has been demonstrated that the piezo-electric

of
59 field of NaNbO3 lead to efficient separation of the photogenerated electron-hole pairs and

ro
60 thereby decreases the recombination rate of the charge carriers [30]. However, NaNbO3 with

61
-p
wide band gap energy (Eg ~3.35 eV) can absorb only ultraviolet portion of the solar spectra
re
62 range [31], thus it is necessary to enhance the range of optical absorption.
lP

63 In this work, we report the synthesis of Ag-NaNbO3 nanocomposite and have


na

64 demonstrated that by coupling both plasmonic effect and piezo-phototronic effect

65 significantly improved photocatalytic activity and photoelectrochemical water splitting


ur

66 activity. To the best of our knowledge, the coupling of the piezo-phototronic and plasmonic
Jo

67 effect of Ag-NaNbO3 nanocomposite have not been studied so far for photocatalytic and PEC

68 water splitting application.

69 2. Experimental details

70 2.1 Material Synthesis

71 2.1.1. Synthesis of NaNbO3 nanorods

72 Niobium pentaoxide (Nb2O5) and sodium hydroxide (NaOH) were used as precursors to

73 synthesize NaNbO3 nanorods using hydrothermal method. Initially, 12 M NaOH was mixed

74 in 50 ml deionized (D.I) water and stirred for 30 min. Afterward, 60 mM Nb2O5 was added to

75 the mixture and stirred for 4 h using a magnetic stirrer. In the end, the mixture was

3
76 transferred to a Teflon lined stainless steel autoclave (100 ml) and the autoclave was kept at

77 160 °C in a hot air oven for 4 h. The final powder product was washed with deionized water

78 and dried at 80 °C for 12 h. The dried product was heat treated in a furnace at 550 °C for 6 h.

79 2.1.2. Synthesis of Ag-NaNbO3 nanocomposite

80 For the synthesis of Ag decorated NaNbO3 nanocomposite, a facile chemical route was

81 employed using tri-sodium citrate, silver nitrate (AgNO3) and sodium borohydride (NaBH4)

82 as precursors. Firstly, well synthesize NaNbO3 nanorods (200 mg) was mixed in 50 ml

83 deionised water and stirred for 15 min and afterwards an aqueous solution of AgNO3 (25

of
84 mM) was added to the above solution dropwise and stirred for 30 min. An aqueous solution

ro
85 of tri-sodium citrate (25 mM) was further added to the mixture. A 4 ml solution of reducing

86
-p
agent sodium tetrahydridoborate (NaBH4) was added dropwise into the above solution and
re
87 stirred for 2 h. After stirring the mixture, the final product was filtered and washed with
lP

88 deionised water and finally dried at 80 °C in a hot air oven for 4 h.


na

89 2.2. Fabrication of photoelectrodes

90 The photoelectrodes comprising of bare NaNbO3 nanorods and Ag-NaNbO3 nanocomposite


ur

91 for photoelectrochemical (PEC) measurements, were prepared by depositing the NaNbO3


Jo

92 nanorods and Ag-NaNbO3 nanocomposite onto fluorine doped tin oxide (FTO) glass

93 substrates using spray coating technique. A 20 mg of NaNbO3 or Ag-NaNbO3 nanocomposite

94 was dispersed in 10 ml of 2-isopropanol for the deposition of films. The as-prepared

95 photoelectrodes were coated with non-conducting epoxy, leaving the working portion (area ~

96 0.95 x 0.95 cm2) for PEC measurements.

97 2.3. Characterization

98 X-ray diffraction measurements were performed to identify the crystal structure of the

99 synthesized samples by using X-ray diffractometer (Rigaku Ultima IV, wavelength, λ = 1.54

100 Å) in the range of 2θ angle for 20 to 70° at a scanning speed of 2°/min. Transmission-electron

4
101 microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) was

102 employed to observe the morphology of the samples by using transmission-electron

103 microscopy, FEI-Tecnai. The elemental analysis and mapping were done by scanning

104 electron microscope TM3000, Hitachi equipped with energy dispersive X-ray spectroscopy

105 (EDX), SwiftED3000 Oxford. UV-visible absorption spectra were employed using 1050

106 Perkin Elmer Lambda spectrophotometer. Piezo-response force microscopy (PFM) study was

107 performed, using Pt/Ir coated silicon (Si) probes on an atomic force microscope (AFM,

108 Bruker, Dimension Icon).

of
109 2.4. Photocatalytic and piezo-photocatalytic activity

ro
110 NaNbO3 nanorods or Ag-NaNbO3 nanocomposite (100 mg) were dispersed in a 250 ml of

111
-p
methylene blue (MB) organic dye solution for investigating the photocatalytic performance
re
112 under UV-visible light irradiation. In order to see the piezo-photocatalytic performance,
lP

113 photodegradation of MB dye was studied when ultra-sonic vibration (40 kHz, 100 W) was
na

114 also present along with UV-visible light illumination. The photodegradation performance was

115 studied in six different conditions: (i) NaNbO3 in the organic dye solution under dark, (ii)
ur

116 NaNbO3 in the organic dye solution under light without ultra-sonic vibration, (iii) NaNbO3 in
Jo

117 the organic dye solution under light with ultra-sonic vibration, (iv) Ag-NaNbO3 in the organic

118 dye solution under dark, (v) Ag-NaNbO3 in the organic dye solution under light without

119 ultra-sonic vibration, and (vi) Ag-NaNbO3 in the organic dye solution under light with ultra-

120 sonic vibration.

121 2.5. Photoelectrochemical and piezo-photoelectrochemical activity

122 The photoelectrochemical and piezo-photoelectrochemical measurements of the photoanodes

123 were performed by utilizing Zahner Zennium potentiostat, PP 211, with a halogen light

124 source (intensity ~100 mW/cm2). A standard three electrode assembly with photoelectrodes

125 comprising of bare NaNbO3 nanorods or Ag-NaNbO3 nanocomposite as a working electrode,

5
126 Ag/AgCl with 3 M NaCl as a reference electrode, and a platinum wire as a counter electrode

127 were employed. For observing the piezo-photoelectrochemical activity, the three electrode

128 cell assembly was exposed to an ultra-sonic vibration (frequency ~40 kHz, vibrating power

129 ~100 W). Electrolyte solution of 0.5 M sodium hydroxide (NaOH) was used in the PEC cell.

130 For a typical current density (J) - potential (V) measurements, a potential scan speed of 0.01

131 V/s was used and EIS measurements were performed for 0.1 Hz to 1000 Hz frequency range.

132 3. Results and discussions

133 3.1. Structural and morphological studies

of
134 Fig. 1 shows the X-ray diffraction patterns of bare NaNbO3 and Ag-NaNbO3

ro
135 nanocomposite. All XRD peaks of NaNbO3 match with the single orthorhombic phase of

136
-p
NaNbO3 (JCPDS file no. 820606). Similarly, the peaks of Ag nanoparticles matched to a
re
137 single cubic phase of Ag (JCPDS file no. 870597). The XRD pattern of Ag nanoparticles is
lP

138 shown in the inset of Fig. 1. The peaks observed in Ag-NaNbO3 composite exhibit peaks
na

139 corresponding to both NaNbO3 and Ag.


ur
Jo

140

141 Fig. 1. XRD patterns of bare NaNbO3 nanorods, Ag nanoparticle (inset of the image) and
142 Ag-NaNbO3 nanocomposite.

6
143 Fig. 2 shows the transmission electron microscopy (TEM) and high resolution

144 transmission electron microscopy (HRTEM) images of NaNbO3 and Ag-NaNbO3

145 nanocomposite. TEM images confirm rod like structure of NaNbO3 and loading of Ag

146 nanoparticles onto NaNbO3 in Ag-NaNbO3 nanocomposite as shown in Fig. 2a and b. The

147 average size of Ag nanoparticles is observed as ~20 nm. The HRTEM images of NaNbO3 and

148 Ag-NaNbO3 nanocomposite show clear lattice-fringes with interplanar spacing (d) ~0.39 nm

149 and ~0.32 nm, which is indexed to (020) and (100) planes, respectively as shown in Fig. 2c

150 and d.

of
ro
-p
re
lP
na
ur
Jo

151
152
153 Fig. 2. TEM images of (a) bare NaNbO3 nanorods and (b) Ag-NaNbO3 nanocomposite and
154 HRTEM images of (c) bare NaNbO3 and (d) Ag-NaNbO3 nanocomposite.

7
155 Fig. 3 show elemental mapping distribution and energy dispersive X-ray

156 spectroscopy (EDX) spectrum images obtained for Ag-NaNbO3 nanocomposite. The

157 elemental analysis of Ag-NaNbO3 was performed in the specific area of SEM image

158 shown in Fig. 3a and revealed the presence of sodium (Na), niobium (Nb), oxygen (O) and

159 silver (Ag) signals (Fig. 3b-e). No other peak of impurity was observed as shown in Fig.

160 3f. The elemental mapping also demonstrated that Ag gets decorated onto NaNbO3

161 nanorods.

of
ro
-p
re
lP
na
ur

162
Jo

163 Fig. 3. Scanning electron microscope (SEM) image and elemental mapping images of Ag-
164 NaNbO3 nanocomposite showing the distribution of sodium (Na), niobium (Nb),
165 oxygen (O) and silver (Ag) (a-e) and (f) a typical EDX spectrum obtained for Ag-
166 NaNbO3 nanocomposite.
167 3.2. Optical, ferroelectric and piezoresponse force microscopy studies

168 Fig. 4a and b show the UV-vis absorbance spectra and corresponding tauc plots of bare

169 NaNbO3 nanorods and Ag-NaNbO3 nanocomposite, respectively. The optical band gap

170 energy of bare NaNbO3 nanorods and Ag-NaNbO3 nanocomposite are observed as ~3.34 eV

171 and ~ 2.44 eV corresponds to the absorption edge ~371 nm and ~508 nm, respectively by

172 using the (αhν)2 vs. hν plot as shown in Fig. 4a and b. As compared to NaNbO3, Ag-NaNbO3

173 nanocomposite showed absorbance in the visible range also, the shift of absorbance edge

174 towards higher wavelength is attributed to the surface plasmonic effect of Ag nanoparticles

8
175 which are decorated over NaNbO3 nanorods. Fig. 4c and d show the piezoresponse force

176 microscopy (PFM) phase hysteresis loop (measures the polarization direction) and amplitude

177 loop (measures the polarization amplitude), respectively to investigate the ferroelectric and

178 piezoelectric properties in NaNbO3 nanorods and Ag-NaNbO3 nanocomposite. The

179 measurements have been done on the NaNbO3 and Ag-NaNbO3 nanocomposite films

180 deposited on fluorine doped tin oxide (FTO) layer coated glass substrate by locally applying

181 DC bias voltage. The bias voltage is applied to the upper surface of the film through the tip of

182 the AFM which is directly in contact with the upper surface of the film while the lower

of
183 surface of the film is grounded. Both NaNbO3 and Ag-NaNbO3 nanocomposite samples are

ro
184 clearly revealed ferroelectric behaviour as demonstrated by the well-defined phase hysteresis

185
-p
loop and butterfly loop as shown in Fig. 4c and d. NaNbO3 and Ag-NaNbO3 nanocomposite
re
186 show sharp domain change by 180° with the DC applied voltage and it is observed from the
lP

187 curve that the coercive voltage of Ag-NaNbO3 nanocomposite film is lower as compared to
na

188 NaNbO3 film. These curves suggest that a lower coercive voltage is required to change the

189 NaNbO3 domain in the presence of Ag nanoparticles. The phase hysteresis loop response is
ur

190 directly related to the alignment of the electric dipole moments. Therefore, the switching of
Jo

191 NaNbO3 dipoles are forced to align along the direction of the applied electric field. This 180°

192 phase switching shows the evidence of ferroelectric property in the pristine NaNbO3

193 nanorods and Ag-NaNbO3 nanocomposite [32]. On the other hand, Fig. 4d shows the

194 amplitude loop, which is correlated to the local strain produced in the films due to applied

195 DC voltage experienced by the cantilever. Upon applying the electric field, a classical

196 butterfly loop is observed for both NaNbO3 and Ag-NaNbO3 films which is an indication of

197 their piezoelectric property. From the slope of the curve, the piezoelectric coefficient (d33)

198 value for NaNbO3 and Ag-NaNbO3 nanocomposite films has been estimated as ~53.3 pm/V

199 and ~57.1 pm/V. It is noticeable that the butterfly loop becomes irregular for Ag-NaNbO3

9
200 with the presence of Ag nanoparticles and this irregularity is attributed to the electrode self-

201 poling effect appearing at the interface among Ag nanoparticles and NaNbO3 [33]. These

202 observations of the phase hysteresis changing and the amplitude loops confirm that both

203 NaNbO3 and Ag-NaNbO3 nanocomposite possess ferroelectric and piezoelectric properties.

of
ro
-p
re
lP
na
ur
Jo

204
205 Fig. 4. (a) UV-Vis absorbance spectra and (b) plot of (αhν)2 vs. hν of pristine NaNbO3 and
206 Ag-NaNbO3 nanocomposite. (c) and (d) represent piezoresponse force microscopy
207 (PFM) hysteresis loop (phase) and butterfly loop (amplitude) of NaNbO3 nanorods
208 and Ag-NaNbO3, respectively.
209
210 3.3. Photocatalytic performance

211 In order to demonstrate the piezo-phototronic and plasmonic effect on the

212 photocatalytic performance, we have performed the measurements under six different

213 conditions: (i) NaNbO3 in the organic dye solution under dark, (ii) NaNbO3 in the organic

214 dye solution under light without ultra-sonic vibration, (iii) NaNbO3 in the organic dye

10
215 solution under light with ultra-sonic vibration, (iv) Ag-NaNbO3 in the organic dye solution

216 under dark, (v) Ag-NaNbO3 in the organic dye solution under light without ultra-sonic

217 vibration, and (vi) Ag-NaNbO3 in the organic dye solution under light with ultra-sonic

218 vibration (Fig. 5). Under dark condition, both NaNbO3 and Ag-NaNbO3 shows negligible

219 degradation of organic MB dye (Fig. 5a and d) and when it is exposed to light illumination,

220 the rate of degradation of dye using Ag-NaNbO3 nanocomposite is greater than NaNbO3

221 nanorods (Fig. 5b and e). The utilization of solar visible-light range for NaNbO3 is negligible

222 due to the large band gap while Ag nanoparticles decorated NaNbO3 exhibit significantly

of
223 improved absorption range of solar light due to the plasmonic effect. Under the combined

ro
224 effect of ultra-sonic vibration and light illumination, the degradation rate of organic

225
-p
methylene blue for NaNbO3 and Ag-NaNbO3 nanocomposite is significantly improved as
re
226 compared to photocatalytic performance with only UV-visible light irradiation as shown in
lP

227 Fig. 5c and f. This enhancement is attributed to the coupling of plasmonic and piezo-
na

228 phototronic effect in Ag-NaNbO3 nanocomposite. The presence of ultra-sonic vibrations

229 induces piezo-voltage at the interface helps in separating the photogenerated electron-hole
ur

230 pairs and resulted in the enhancement of the photocatalytic activity. Fig. 5g shows the relative
Jo

231 concentration plots of MB organic dye ( ) after complete reaction with NaNbO3 and Ag-

232 NaNbO3 photocatalyst [34];

233 ln = kt (1)

234 where, c, c , t and k is the concentration of MB reactant, instantaneous concentration, time

235 interval and apparent rate constant, respectively. Under dark condition, the rate constant of

236 NaNbO3 and Ag-NaNbO3 are estimated to be 0.0002 min-1 and 0.0004 min-1, respectively.

237 After the decoration of Ag nanoparticles and under light irradiation, the rate constant is

238 increased to 0.002 min-1 for Ag-NaNbO3 from 0.001 min-1 for NaNbO3 (under light without

239 ultra-sonic vibration). Furthermore, under the piezo-phototronic effect, the rate constant is

11
240 increased further to 0.004 min-1 for bare NaNbO3 (under light with ultra-sonic vibration)

241 whereas for Ag-NaNbO3 the rate constant is 0.011 min-1 (Fig. 5i). The plasmonic Ag-

242 NaNbO3 nanocomposite has shown ~2 fold improvement in the decomposition of organic

243 MB dye as compared to bare NaNbO3 (under light without ultra-sonic vibration) whereas,

244 under the coupling of plasmonic and piezo-phototronic effect, ~10 fold enhancement is

245 achieved for Ag-NaNbO3 (under light with ultra-sonic vibration) as compared to bare

246 NaNbO3. The piezoelectric potential is induced in Ag-NaNbO3 due to the periodic ultra-sonic

247 vibration and this induced piezo potential enhanced the separation of photogenerated e--h+

of
248 pairs at the interfaces. These photogenerated charge carriers transferred to the external

ro
249 surface site of Ag-NaNbO3 where they participate in the photocatalytic reactions for dye

250
-p
degradation. The normalized instantaneous concentration and apparent rate constant plots
re
251 (Fig. 5g and h) evidently show significant improvement in the photodegradation of MB dye
lP

252 using Ag-NaNbO3 and NaNbO3 (under light with ultra-sonic vibration) as compared to Ag-
na

253 NaNbO3 and NaNbO3 (under light without ultra-sonic vibration).


ur
Jo

12
of
ro
-p
re
lP
na

254
255 Fig. 5. Absorbance spectra of photocatalytic decomposition activity of MB dye with 30 min
256 intervals for 3 h under six different conditions: (a) NaNbO3 (dark), (b) NaNbO3
ur

257 (under light without ultra-sonic vibration), (c) NaNbO3 (under light with ultra-sonic
258 vibration), (d) Ag-NaNbO3 (dark), (e) Ag-NaNbO3 (under light without ultra-sonic
Jo

259 vibration), (f) Ag-NaNbO3 (under light with ultra-sonic vibration), (g) comparison of
260 the photocatalytic degradation proficiency done in various test conditions, (h)
261 demonstrate kinetics fit to the information, and (i) comparison of the rate constant (k)
262 values for the decomposition activity of MB dye for different test conditions.
263 To understand the mechanism of the photocatalytic degradation of MB dye using Ag-

264 NaNbO3 nanocomposite under the combined effect of piezo-phototronic and plasmonic effect

265 simultaneously, different scavengers were used to identify the participation of the active

266 species in the photocatalytic degradation experiment. Various scavengers such as silver

267 nitrate (AgNO3), ammonium oxalate (AO), a benzoquinone (BQ) and t-BuOH were

268 employed to elucidate the role of photogenerated electrons, holes, superoxide anion radicals

269 (O2• ─) and hydroxyl radicals (OH•), respectively in the photocatalytic degradation process.

13
270 We have added a known amount of these scavengers into the dye solution before the addition

271 of Ag-NaNbO3 nanocomposite sample and performed the photodegradation test under the

272 combined effect of light irradiation and ultra-sonic vibration. As shown in the Fig. 6, the

273 photocatalytic degradation efficiency of MB dye over Ag-NaNbO3 nanocomposite was

274 decreased from 98.4% to 28.2% when AgNO3 was added into the dye solution. This result

275 indicates the essential role of the photogenerated electrons in driving the photodegradation of

276 MB by combined effect of light irradiation and piezo potential induced by ultra-sonic

277 vibration. The decrement in photocatalytic efficiency of MB dye degradation in the presence

of
278 of AO from 98.40% to 91.80% revealed the insignificant role of holes in the

ro
279 photodegradation of MB. Moreover, the photocatalytic efficiency was noticeably decreased

280
-p
from 98.40% to 13.60% in the presence of BQ, whereas the efficiency was drastically
re
281 reduced to 9.20% in the presence of t-BuOH. These results clearly showed the important
lP

282 roles of electrons, OH• and O2• ─ which are responsible for the photodegradation of MB.
na
ur
Jo

283

284 Fig. 6. Effect of different active species scavengers for the photodegradation of MB over Ag-
285 NaNbO3 nanocomposite under the combined effect of both light irradiation and ultra-
286 sonic vibration, simultaneously for 180 min.
287

14
288 The recyclability and stability test of the photocatalyst during the photocatalytic

289 process is very essential for their sustainable utilization in the environmental applications.

290 Therefore, the recyclability of the as-prepared Ag-NaNbO3 nanocomposite has been explored

291 by using the same photocatalyst for executing several repeated runs of MB degradation

292 experiments under the combined effect of light irradiation and ultra-sonic vibration,

293 simultaneously (see supporting information, Fig. S1a). It is evident that the Ag-NaNbO3

294 nanocomposite showed excellent photocatalytic dye degradation performance even after three

295 consecutive runs of photocatalysis. Furthermore, the XRD measurement has been carried out

of
296 for the Ag-NaNbO3 nanocomposite sample after the three consecutive runs of photocatalytic

ro
297 performance and compared with the XRD pattern of the fresh Ag-NaNbO3 nanocomposite.

298
-p
There are no significant changes in the XRD patterns of the fresh and recovered Ag-NaNbO3
re
299 nanocomposite even after three times runs of the photocatalytic dye degradation experiment
lP

300 (see supporting information, Fig. S1b). Thus, it can be clearly seen that the as-prepared Ag-
na

301 NaNbO3 nanocomposite sample is efficient, stable and durable for photocatalysis.

302 Fig. 7 illustrates a schematic of the possible mechanism of Ag-NaNbO3


ur

303 nanocomposite for higher decomposition rate of MB dye under the combined effect of both
Jo

304 light irradiation and ultra-sonic vibration simultaneously. The conduction band (CB) position

305 of NaNbO3, valence band (VB) position of NaNbO3 and Fermi level position of Ag is at

306 −0.60 eV, +2.75 eV and +0.4 eV (vs. NHE), respectively [35, 36]. Under UV-visible light

307 irradiation, the e--h+ pairs are generated in the conduction and valence band of NaNbO3 and

308 plasmon meditated hot electrons are produced in Ag nanoparticles due to the surface

309 plasmonic effect. The generated electrons are transferred from the excited states of Ag

310 nanoparticles to the conduction band of bare NaNbO3 which contribute in the reactions at the

311 surface sites of NaNbO3 and the holes transfer in the reverse direction from the valence band

312 of NaNbO3 to Ag. The photogenerated electrons present in the conduction band of NaNbO3

15

313 can react with absorbed O2 to generate superoxide anion O2• as the conduction band

314 position is more negative than the standard oxidation potential for the conversion of O2 to O2•

315

(E0 (O2/O2• ─) = −0.046 eV vs. NHE [37]). Simultaneously, the photogenerated holes are

316 migrated from the valence band of NaNbO3 and generate hydroxyl (OH•) radicals as the

317 valence band potential +2.75 eV is more positive than the standard potential of OH─/OH•

318 (+1.99 eV vs. NHE [38]). Furthermore, in Ag-NaNbO3 nanocomposite (under periodic ultra-

319 sonic vibration), a significant polarization voltage is induced and due to the effect of piezo-

320 potential, separation of e--h+ pairs and the recombination rate of photogenerated e--h+ pairs

of
321 are reduced. In addition, the Ag-NaNbO3 nanocomposite interface will also facilitate the

ro
322 charge carrier migration and separation of the photogenerated e--h+ pairs which will further

323
-p
result in the reduction of the rate of recombination of photogenerated charge carriers and
re

324 improve the photocatalytic activity. These generated O2• and OH• radicals finally
lP

325 decompose the MB dye. The possible chemical reactions in this process are as follows [39-
na

326 42]:

327 Ag-NaNbO3 + hν → Ag (h+VB) ─ NaNbO3 (e-CB) (2)


ur

2 e-CB + O2 → O2• ─ (3)


Jo

328

329 O2• ─ + H2O → OOH• + OH─ (4)

330 OOH• + H2O → H2O2 + OH• (5)

331 H2O2 → OH• + OH• (6)

332 h+VB + OH─→ OH• (7)

333 OH• + MB → CO2 +H2O (8)

334 The generated piezo potential significantly improves the separation of photogenerated charge

335 carriers, whereas the plasmonic effect enhances the light absorption resulting in generation of

336 more electron-hole pairs.

16
of
ro
337
338
-p
re
339 Fig. 7. Schematic of the possible working mechanism of Ag-NaNbO3 nanocomposite
340 showing photogenerated charge carrier separation resulting in efficient degradation of
lP

341 organic dye.


342 3.4. Photoelectrochemical studies
na

343 For the photoelectrochemical (PEC) measurements, NaNbO3 nanorods and Ag-
ur

344 NaNbO3 nanocomposite photoanodes were tested in a three-electrode photoelectrochemical


Jo

345 configured with Ag/AgCl (3 M NaCl) and platinum (Pt) wire as a reference and the counter

346 electrode, respectively. A 0.5 M sodium hydroxide (NaOH) was utilized as an electrolyte

347 solution. The photoanodes were prepared by depositing thin films of NaNbO3 nanorods and

348 Ag-NaNbO3 nanocomposite on conducting fluorine doped tin oxide (FTO) layer coated glass

349 substrates by using the spray coating technique. We applied an external bias, so that

350 randomly oriented dipoles present in NaNbO3 and Ag-NaNbO3 films become aligned and the

351 bands bend upward, which favours the oxidation evolution reaction at the

352 semiconductor/electrolyte interface. Fig. 8a indicates the current density (J) - potential (V)

353 curves of NaNbO3 nanorods and Ag-NaNbO3 nanocomposite photoanodes with dark and

354 light illumination (intensity ~100 mW/cm2) in the potential range of 0 V to +1.0 V (vs.

17
355 Ag/AgCl) and at the scan rate of 0.01 V/s under eight different conditions: (i) NaNbO3 under

356 dark without ultra-sonic vibration, (ii) Ag-NaNbO3 under dark without ultra-sonic vibration,

357 (iii) NaNbO3 under dark with ultra-sonic vibration, (iv) Ag-NaNbO3 under dark with ultra-

358 sonic vibration, (v) NaNbO3 under light without ultra-sonic vibration, (vi) Ag-NaNbO3 under

359 light without ultra-sonic vibration, (vii) NaNbO3 under light with ultra-sonic vibration, and

360 (viii) Ag-NaNbO3 under light with ultra-sonic vibration. Due to the plasmonic effect, the

361 photocurrent density of Ag-NaNbO3 photoanode is found to increase to 4.15 mA/cm2 from

362 1.52 mA/cm2 (at 1V vs. Ag/AgCl) for bare NaNbO3 photoanode (under light without ultra-

of
363 sonic vibration). Under the piezo-phototronic effect, the photocurrent density is observed to

ro
364 increase to 4.83 mA/cm2 from 1.52 mA/cm2 for bare NaNbO3 when the photoanode was

365
-p
exposed to ultra-sonic vibrations. The Ag-NaNbO3 (under light with ultra-sonic vibration)
re
366 exhibited the highest photocurrent density of 9.65 mA/cm2 (at 1V vs. Ag/AgCl). The
lP

367 plasmonic Ag-NaNbO3 (under light without ultra-sonic vibration) has shown ~3 fold
na

368 enhancement in the photocurrent density as compared to bare NaNbO3 (under light without

369 ultra-sonic vibration), whereas under the presence of both plasmonic and piezo-phototronic
ur

370 effect, ~9 fold enhancement is achieved for Ag-NaNbO3 (under light with ultra-sonic
Jo

371 vibration) as compared to bare NaNbO3 (under light without ultra-sonic vibration). This

372 enhancement is attributed to the coupling of plasmonic and piezo-phototronic effect. Under

373 ultra-sonic vibration, a slight shift in the onset potential of current densities was also

374 observed which show a modification in the band position alignment at the interface of Ag-

375 NaNbO3/electrolyte resulting in the enhancement of the photocurrent densities [43].

376 Moreover, the as-prepared photoanodes have showed excellent photostability as confirmed

377 from the J-V measurements executed up to eight repeated cycles (see supporting information,

378 Fig. S2). Fig. 8b shows the transient photocurrent measurements of all the as-prepared

379 photoelectrodes with an applied potential of 1 V with respect to Ag/AgCl during light ON-

18
380 OFF cycles. The duration of light ON-OFF cycles is ~20 seconds. It can be clearly seen from

381 the plot that Ag-NaNbO3 (under light with ultra-sonic vibration) photoanode possesses

382 highest current density as compared to all other photoelectrodes and these results are in

383 agreement with the current density (J) - potential (V) measurements which further confirms

384 that Ag-NaNbO3 (under light with ultra-sonic vibration) photoanode shows the best

385 photoresponse in the PEC activity.

386 In order to understand the effect of plasmonic and piezo-phototronic effect on the

387 charge transfer phenomenon at the electrode/electrolyte interface, we measured the

of
388 electrochemical impedance spectrum (EIS), which is an effective method for evaluating the

ro
389 charge-transfer resistance (Rct) [44]. Fig. 8c shows the Nyquist curves for NaNbO3 (without

390
-p
ultra-sonic vibration), Ag-NaNbO3 (without ultra-sonic vibration), NaNbO3 (with ultra-sonic
re
391 vibration) and Ag-NaNbO3 (with ultra-sonic vibration) photoanodes under light illumination.
lP

392 The semicircles of the Nyquist plots are fitted using Z-View software and revealed the
na

393 charge-transfer resistance (Rct) at the semiconductor/electrolyte interface. The equivalent

394 circuit consisting of a series resistance of (Rs) and charge-transfer resistance (Rct) are shown
ur

395 in the insets of Fig. 8c. The charge-transfer resistance (Rct) are obtained as 6.8 kΩ, 6.0 kΩ,
Jo

396 4.5 kΩ and 3.6 kΩ for the NaNbO3 (without ultra-sonic vibration), Ag-NaNbO3 (without

397 ultra-sonic vibration), NaNbO3 (with ultra-sonic vibration), and Ag-NaNbO3 (with ultra-sonic

398 vibration), respectively. Significantly lower value of the charge transfer resistance of Ag-

399 NaNbO3 photoanode as compared to NaNbO3 photoanode clearly indicates that there is a

400 substantial enhancement in the charge transfer through the Ag-NaNbO3/electrolyte interface

401 which explains the observed improvement in the current-density of Ag-NaNbO3 photoanode.

402 Fig. 8d shows the incident photon to current conversion efficiency (IPCE)

403 photoresponses of as fabricated photoelectrodes under four conditions: (i) NaNbO3 under

404 light without ultra-sonic vibration, (ii) Ag-NaNbO3 under light without ultra-sonic vibration,

19
405 (iii) NaNbO3 under light with ultra-sonic vibration, and (iv) Ag-NaNbO3 under light with

406 ultra-sonic vibration. The IPCE measurement was obtained under monochromatic light

407 illumination from 400 to 650 nm with respect to Ag/AgCl reference electrode, according to

408 the formula; IPCE (%) = [1240 x J (mAcm-2)]/[λ (nm) x P (mWcm-2)]. where J is the

409 measured photocurrent density, λ is the wavelength of incident monochromatic light and P is

410 the incident power density. It can be seen that Ag-NaNbO3 photoanode possess the highest

411 IPCE value ~29.6% at a wavelength of 430 nm when subjected to combined effect of light

412 illumination and ultra-sonic vibration. A ~5 fold enhancement in the IPCE value is observed

of
413 as compared to bare NaNbO3 (under light without ultra-sonic vibration) photoanode. The

ro
414 IPCE value of NaNbO3 under light without ultra-sonic vibration, Ag-NaNbO3 under light

415
-p
with ultra-sonic vibration, NaNbO3 under light with ultra-sonic vibration is observed to be
re
416 6.1%, 13.8% and 18.2%, respectively. This clearly shows that the combined piezo-
lP

417 phototronic and plasmonic effect significantly enhanced the ability of light absorption and
na

418 improved the separation and transfer efficiency of the charges, thereby enhancing the PEC

419 performance.
ur
Jo

20
of
ro
-p
re
lP

420
421 Fig. 8. (a) Current density (J) – potential (V) scans of NaNbO3 and Ag-NaNbO3 phototanode
422 under dark and light conditions (scan speed 0.01 V/s): (i) NaNbO3 and (ii) Ag-
na

423 NaNbO3 without ultra-sonic vibration and (iii) NaNbO3 and (iv) Ag-NaNbO3 with
424 ultra-sonic vibration, (b) transient photocurrent curves for NaNbO3 and Ag-NaNbO3
ur

425 (without ultra-sonic vibration), and NaNbO3 and Ag-NaNbO3 (with ultra-sonic
426 vibration) photoelectrodes at 1 V with 20 sec ON/OFF cycles, (c) EIS Nyquist plots
Jo

427 of NaNbO3 and AgNaNbO3 (without ultra-sonic vibration) and NaNbO3 and Ag-
428 NaNbO3 (with ultra-sonic vibration) photoanodes under light illumination. (Solid dots
429 (•) denote experimental points, whereas solid lines (─) denote the simulated curves
430 using the equivalent circuit shown in the inset) and (d) IPCE photoresponse of
431 NaNbO3 and Ag-NaNbO3 (without ultra-sonic vibration) and NaNbO3 and Ag-
432 NaNbO3 (with ultra-sonic vibration) photoanodes.
433 To determine the type of conductivity, flatband potential (Vfb) and charge carrier
434 density (Nd) of all the as-prepared photoelectrodes, Mott-Schottky measurements have been
435 carried out at different frequencies (1 kHz, 2 kHz and 3 kHz) under dark condition. The Mott-
436 Schottky relation is given as [45]:
1 2 k(T
= $ V && − V' − $ (9)
C qε ε N# q
437 where C is the capacitance of the semiconductor/electrolyte interface, q is the electron
438 charge, ε0 is the permittivity in a vacuum, εs is the dielectric constant of NaNbO3 (100) [46],
439 Vapp is the applied voltage, Vfb is the flatband potential, Nd is the charge carrier density, kB is

21
440 the Boltzmann’s constant and T is the absolute temperature. As shown in Fig. 9a-d, the
441 NaNbO3 (under dark without ultra-sonic vibration), Ag-NaNbO3 (under dark without ultra-
442 sonic vibration), NaNbO3 (under dark with ultra-sonic vibration) and Ag-NaNbO3 (under
443 dark with ultra-sonic vibration) photoanodes under dark condition with different frequencies
444 exhibited a positive slopes, indicating typical n-type semiconductor behaviour. The flat band
445 potential of all as-prepared photoanodes is obtained by extrapolating the straight line of the
446 curve to the x-axis. The value of flatband potential (Vfb) is estimated as −0.42 V vs. Ag/AgCl
447 for NaNbO3 (under dark without ultra-sonic vibration), −0.41 V vs. Ag/AgCl for Ag-
448 NaNbO3 (under dark without ultra-sonic vibration), −0.45 V vs. Ag/AgCl for NaNbO3 (under
449 dark with ultra-sonic vibration) and −0.53 V vs. Ag/AgCl for Ag-NaNbO3 (under dark with

of
450 ultra-sonic vibration) photoanodes. The value of the Vfb is same at different frequencies for

ro
451 all the as-prepared photoelectodes. The slightly negative shift in the Vfb value of the Ag-
452 NaNbO3 (under dark without ultra-sonic vibration) is attributed to the higher separation of
453
-p
charge carrier due to the shifting of Fermi level of Ag and NaNbO3. Furthermore, after
re
454 piezophototronic effect (ultra-sonic vibration), the Vfb value of NaNbO3 (under dark with
455 ultra-sonic vibration) and Ag-NaNbO3 (under dark with ultra-sonic vibration) further shifts to
lP

456 more negative as compared to NaNbO3 (under dark without ultra-sonic vibration) suggesting
na

457 the upward shifting of the respective Fermi levels with respect to redox potential of water.
458 The more upward shifting of Fermi level for Ag-NaNbO3 (under dark with ultra-sonic
ur

459 vibration) photoanode resulted in an increase in the band bending at the semiconductor
460 photoelectrode/electrolyte interface which increased the separation efficiency and enhanced
Jo

461 the transfer of photogenerated charge carriers. The Nd value of all the as-prepared
462 photoelectrodes at 1 kHz frequency is also estimated from the Mott-Schottky plot. The value
463 of Nd is observed as ~0.4 x1018 cm-3, ~1.2 x1018 cm-3, ~1.7 x 1018 cm-3 and ~2.2 x 1018 cm-3
464 for NaNbO3 (under dark without ultra-sonic vibration), Ag-NaNbO3 (under dark without
465 ultra-sonic vibration), NaNbO3 (under dark with ultra-sonic vibration) and Ag-NaNbO3
466 (under dark with ultra-sonic vibration) photoanodes, respectively. The increment in the value
467 of Nd of Ag-NaNbO3 (under dark with ultra-sonic vibration) photoanode is attributed to the
468 transfer of charge carriers (i.e. electrons) from Ag to NaNbO3 due to the plasmonic effect of
469 Ag nanoparticles and more band bending of NaNbO3 due to the piezo-phototronic effect
470 resulting in enhanced mobility and separation of the charge carriers and suppression the
471 recombination of photogenerated charge carriers.

22
472 In order to further understand the charge transfer dynamics of the as-prepared
473 photoelectrodes, the charge injection efficiency (η , ) and charge separation efficiency (η -& )

474 measurements were carried out to by adding 0.5 M of hydrogen peroxide (H2O2) as a hole
475 scavenger in the 0.5 M of NaOH electrolyte solution [47]. The significant assumption of this
476 method is that the oxidation kinetics of H2O2 are very rapid and its charge injection efficiency
477 is 100% or the rate of recombination of the photogenerated charges at the surface is supposed
478 to be removed [48]. The η , and η -& can be estimated by following equations [49];
./0
479 η , = .12 32
(10)

.12 32
480 η -& = (11)
.567

of
where J92 :2 is the photocurrent density in the presence of 0.5 M H2O2 + 0.5 M NaOH, J;

ro
481 < is
482 the theoretical maximum photocurrent and J&= is the photocurrent density in the presence of
483
-p
0.5 M NaOH. Fig. 9e and 9f illustrate the charge injection efficiency (η , ) and charge
re
484 separation efficiency (η -& ), respectively. As shown in Fig. 9e, the η , value of NaNbO3
lP

485 (under light without ultra-sonic vibration) is observed to be 74.1% at 1 V with respect to
486 Ag/AgCl, while the η , of Ag-NaNbO3 (under light without ultra-sonic vibration), NaNbO3
na

487 (under light with ultra-sonic vibration), and Ag-NaNbO3 (under light with ultra-sonic
488 vibration) photoanodes has been improved up to 87.4%, 92.1% and 96.6% at 1V with respect
ur

489 to Ag/AgCl, respectively. The highest η , for Ag-NaNbO3 (under light with ultra-sonic
Jo

490 vibration) photoanode is signifying inhibited recombination rate of the photogenerated


491 charges at the interface which expedites the oxidation evolution reaction and enhances the
492 PEC performance as compared to other photoanodes. In addition, the value of η -& of
493 NaNbO3 (under light without ultra-sonic vibration), Ag-NaNbO3 (under light without ultra-
494 sonic vibration), NaNbO3 (under light with ultra-sonic vibration), and Ag-NaNbO3 (under
495 light with ultra-sonic vibration) photoanodes is observed to be 26.3%, 47.5%, 54.7%, and
496 65.7% at 1 V with respect to Ag/AgCl, respectively (Fig. 9f). A significant increment in the
497 η -& value of Ag-NaNbO3 (under light with ultra-sonic vibration) is attributed to the
498 combined effect of piezo-phototronic and plasmonic effect which results in an increase of the
499 separation of charges, increase the mobility of the transfer of photogenerated charges and
500 efficiently reduce the recombination of charges . These results also reveal that the coupling of
501 the piezo-phototronic and plasmonic effect not only support the separation of charges, but
502 also promote the water oxidation reaction on the surface of the photoelectrode.

23
of
503

ro
504 Fig. 9. Mott-Schottky curves of (a) NaNbO3 (under dark without ultra-sonic vibration), (b)
505
506
-p
Ag-NaNbO3 (under dark without ultra-sonic vibration), (c) NaNbO3 (under dark with
ultra-sonic vibration), (d) Ag-NaNbO3 (under dark with ultra-sonic vibration)
re
507 photoanodes at different frequencies in the dark condition, (e) Variation of the charge
508 injection efficiency (η , ) and (f) charge separation efficiency (η -& ) with an applied
lP

509 bias for NaNbO3 (under light without ultra-sonic vibration), Ag-NaNbO3 (under light
510 without ultra-sonic vibration), NaNbO3 (under light with ultra-sonic vibration) and
511 Ag-NaNbO3 (under light with ultra-sonic vibration) photoanodes.
na

512
513 Fig. 10 shows a schematic of the possible charge carrier transfer mechanism at the
ur

514 interface of Ag-NaNbO3/electrolyte for the enhanced PEC water splitting activity under the
Jo

515 combined effect of ultra-sonic vibration and UV-visible light illumination. Under light

516 illumination, the photoexcited charge carriers (e--h+) will be generated in the CB and VB of

517 NaNbO3. Simultaneously, the photoexcited charges (i.e. surface plasmon mediated hot

518 electrons) will be generated on the surface of Ag nanoparticles. Due to proper band

519 alignment, these generated hot electrons will transfer from the excited state of Ag

520 nanoparticles to CB of NaNbO3 overcoming the Schottky barrier. Furthermore, on applying

521 the strain (i.e. with ultra-sonic vibrations), a piezo potential will eventually generate on the

522 surface of NaNbO3 nanorods by virtue of its piezoelectric property which raises the valence

523 band of NaNbO3 upward and it reaches to nearer to the water oxidation potential resulting in

524 the transfer of more h+ to contribute in the O2 photoelectrochemical evolution reactions at the

24
525 photoanode. The raised up conduction band level of NaNbO3 at the electrolyte interface due

526 to the strain effect increases the mobility of transport of e- towards the FTO substrate and

527 contribute in the H2 photoelectrochemical evolution reaction at the counter (Pt) electrode.

528 The possible photoreactions that take place at the Ag-NaNbO3/electrolyte interface and the Pt

529 electrode are as follows [50]:

530 Ag-NaNbO3 + hν → Ag (h+VB) ─ NaNbO3 (e-CB) (12)

531 h+VB + H2O → OH• + H+ (13)

532 OH• + OH• → H2O2 (14)

of
533 2H2O2 → 2H2O + O2 (15)

ro
534 2h+VB + H2O → ½ O2 + 2H+ (16)

535 2H+ + 2e-CB → H2


-p (17)
re
536 The observed enhanced photoelectrochemical activity of the fabricated Ag-NaNbO3
lP

537 photoelectrode is attributed to the combined effect of plasmonic Ag nanoparticles and piezo-
na

538 phototronic effect of NaNbO3 nanorods.


ur
Jo

539
540
541 Fig. 10. Schematic diagram of PEC three electrode assembly (left) and energy band diagram
542 of Ag-NaNbO3/electrolyte interface with and without strain (right).
543 4. Conclusions

25
544 In summary, piezoelectric NaNbO3 nanorods have been coupled to Ag nanoparticles to form

545 Ag-NaNbO3 nanocomposite by chemical solution method. The Ag-NaNbO3 nanocomposite

546 exhibited significantly higher photocatalytic degradation rate and PEC water splitting

547 performance. As compared with bare NaNbO3 (under light without ultra-sonic vibration), Ag-

548 NaNbO3 (under light without ultra-sonic vibration) exhibited ~2 fold improvement in the

549 photocatalytic decomposition of MB dye and Ag-NaNbO3 (under light with ultra-sonic

550 vibration) shows ~10 fold improvement in the photocatalytic decomposition of MB dye as

551 compared to bare NaNbO3 (under light without ultra-sonic vibration). Whereas ~3 fold

of
552 enhancement in photocurrent density for Ag-NaNbO3 (under light without ultra-sonic

ro
553 vibration) photoelectrode as compared to bare NaNbO3 (under light without ultra-sonic

554
-p
vibration) and ~9 fold enhancement in photocurrent density for Ag-NaNbO3 (under light with
re
555 ultra-sonic vibration) photoelectrode as compared to bare NaNbO3 (under light without ultra-
lP

556 sonic vibration) is achieved in photoelectrochemical water splitting activity, respectively. The
na

557 enhancement in the photocatalytic and photoelectrochemical activities can be attributed to the

558 coupling of piezo-phototronic and plasmonic effect together. The surface plasmonic effect
ur

559 due to presence of Ag nanoparticles expands the visible-light absorption part and the piezo-
Jo

560 phototronic effect of NaNbO3 material enhances the drift and separation of the

561 photogenerated charge carriers.

562 Conflict of Interest

563 The authors declare that they have no conflict of interest.

564 Credit authorship contribution statement

565 Dheeraj Kumar: Conceptualization, Methodology, Formal Analysis, Validation

566 Investigation, Writing – original draft. Surbhi Sharma: Validation, Investigation, Writing –

567 review & editing. Neeraj Khare: Conceptualization, Methodology, Validation, Supervision,

568 Writing – review & editing.

26
569 Acknowledgements

570 We gratefully acknowledge the financial assistance from Indian Institute of Technology

571 Delhi (IITD), Ministry of Electronics and Information Technology (MeitY), Department of

572 Science and Technology in the NNetRA Project (no- RP03530, MI01756). We also

573 acknowledge Nanoscale Research Facility (NRF) and Central Research Facility (CRF), IIT

574 Delhi for giving us an opportunity to use the characterization facilities.

575

576

of
577

ro
578

579
-p
re
580
lP

581 References
na

582 [1] J. Joy, J. Mathew, S. C. George, Nanomaterials for photoelectrochemical water

583 splitting – review, Int. J. Hydrogen Energy. 43 (2018) 4804–4817


ur

584 [2] Z. Boukhemikhem, R. Brahimi, G. Rekhila, G. Fortas, L. Boudjellal, M. Trari, The


Jo

585 photocatalytic hydrogen formation and NO2- oxidation on the hetero-junction

586 Ag/NiFe2O4 prepared by chemical route, Renew. Energy. 145 (2020) 2615–2620.

587 [3] S. Sharma, S. Singh, N. Khare, Enhanced photosensitization of zinc oxide nanorods

588 using polyaniline for efficient photocatalytic and photoelectrochemical water

589 splitting, Int. J. Hydrogen Energy 41 (2016) 21088–21098.

590 [4] S. Xu, J. Jiang, W. Ren, H. Wang, R. Zhang, Y. Xie, Y. Chen, Construction of

591 ZnO/CdS three-dimensional hierarchical photoelectrode for improved

592 photoelectrochemical performance, Renew. Energy. 153 (2020) 241–248.

27
593 [5] S. Sharma, D. Kumar, N. Khare, Hierarchical PANI/ZnO Nanocomposite: Synthesis

594 and Synergistic effect of Shape-Selective ZnO Nanoflowers and Polyaniline

595 Sensitization for Efficient Photocatalytic dye degradation and Photoelectrochemical

596 Water Splitting, Nanotechnology 31 (2020) 465402–465420.

597 [6] Y. Yang, L. Liu, Q. Qi, F. Chen, M. Qiu, F. Gao, J. Chen, A low-cost and stable

598 Fe2O3/C-TiO2 system design for highly efficient photocatalytic H2 production from

599 seawater Catal. Commun. 143 (2020) 106047–106050.

600 [7] Y. J Chen, L. Y. Chen, The study of carrier transfer mechanism for nanostructural

of
601 hematite photoanode for solar water splitting, Appl. Energy. 164 (2016) 924–933.

ro
602 [8] J. Lee, S. K. Kim, Y. Sohn, Understanding photocatalytic coupled-dye degradation,

603
-p
and photoelectrocatalytic water splitting and CO2 reduction over WO3/MoO3 hybrid
re
604 nanostructures, J. Ind. Eng. Chem. 62 (2018) 362–374.
lP

605 [9] Z. Liu, J. Wu, J. Zhang, Quantum dots and plasmonic Ag decorated WO3 nanorod
na

606 photoanodes with enhanced photoelectrochemical performances, Int. J. Hydrogen

607 Energy. 41 (2016) 20529–20535.


ur

608 [10] Y. Wang, M. Zhou, Y. He, Z. Zhou, Z. Sun, In situ loading CuO quantum dots on
Jo

609 TiO2 nanosheets as cocatalyst for improved photocatalytic water splitting, J. Alloys.

610 Compd. 813 (2020) 152184–152190.

611 [11] C. Y. Chian, Y. Shin, S. Ehrman, Dopant effects on conductivity in copper oxide

612 photoelectrochemical cells, Appl. Energy. 164 (2016) 1039–1042.

613 [12] J. Zhao, Z. Zhao, N. Li, J. Nan, R. Yu, J. Du, Visible-light-driven photocatalytic

614 degradation of ciprofloxacin by a ternary Mn2O3/Mn3O4/MnO2 valence state

615 heterojunction, Chem. Eng. Sci. 353 (2018) 805–813.

616 [13] Y. H. Su, S. H. Huang, P. Y. Kung, T. W. Shen, W. L. Wang, Hydrogen Generation

617 of Cu2O Nanoparticles/MnO−MnO2 Nanorods Heterojunction Supported on

28
618 Sonochemical-Assisted Synthesized Few-Layer Graphene in Water-Splitting

619 Photocathode, ACS Sustainable Chem. Eng. 3 (2015) 1965–1973.

620 [14] D. Wang, W. Wang, Q. Wang, Z. Guo, W. Yuan, Spatial separation of Pt and IrO2

621 cocatalysts on SiC surface for enhanced photocatalysis, Mater. Lett. 201 (2017) 114–

622 117.

623 [15] R. Badam, M. Hara, H. H. Huang, M. Yoshimura, Synthesis and electrochemical

624 analysis of novel IrO2 nanoparticle catalysts supported on carbon nanotube for oxygen

625 evolution reaction, Int. J. Hydrogen Energy. 43 (2018) 18095–18104.

of
626 [16] Y. F. Hu, Y. L. Chang, P. Fei, R.L. Snyder, Z. L. Wang, Designing the electric

ro
627 transport characteristics of ZnO micro/nanowire devices by coupling piezoelectric and

628
-p
photoexcitation effects, ACS Nano. 4 (2010) 1234–1240.
re
629 [17] Z. Liang, C. Yan, S. Rtimi, J. Bandara, Piezoelectric materials for
lP

630 catalytic/photocatalytic removal of pollutants : Recent advances and outlook Appl.


na

631 Catal. B: Environ. 241 (2019) 256–269.

632 [18] K. S. Hong, H. Xu, H. Konishi, X. Li, Direct Water Splitting Through Vibrating
ur

633 Piezoelectric Microfibers in Water, J. Phys. Chem. Lett. 1 (2010) 997–1002.


Jo

634 [19] Z. Zhanga, Q. Liaoa, Y. Yu, X. Wang, Y. Zhang, Enhanced photoresponse of ZnO

635 nanorods-based self-powered photodetector by piezotronic interface engineering, Nano

636 Energy. 9 (2014) 237–244.

637 [20] X. N. Wen, W. Z. Wu, Z. L. Wang, Effective piezophototronic enhancement of solar

638 cell performance by tuning material properties, Nano Energy. 2 (2013) 1093–1110.

639 [21] Y. H. Hu, Y. Zhang, L. Lin, Y. Ding, G. Zhu, Z. L. Wang, Piezo-phototronic effect on

640 electroluminescence properties of p-type GaN thin films, Nano Lett. 12 (2012) 3851–

641 3856.

642 [22] X. Guo, Y. Fu, D. Hong, B. Yu, H. He, Q. Wang, L. Xing, X. Xue, High-efficiency

29
643 sono-solar-induced degradation of organic dye by the piezophototronic/photocatalytic

644 coupling effect of FeS/ZnO nanoarrays, Nanotechnology. 27 (2016) 375704–373714.

645 [23] F. Zhang, Y. Ding, Y. Zhang, X. L. Zhang, Z. L. Wang, Piezo-phototronic effect

646 enhanced visible and ultraviolet photodetection using a ZnO/CdS core/shell micro/

647 nanowire, ACS Nano. 6 (2012) 9229–9236.

648 [24] Q. Yang, Y. Wu, Y. Liu, C. Pana, Z. L. Wang, Features of the piezo-phototronic effect

649 on optoelectronic devices based on wurtzite semiconductor nanowires, Phys. Chem.

650 Chem. Phys. 16 (2014) 2790–2800.

of
651 [25] Q. Liu, Y. Chai, L. Zhang, J. Ren, W. L. Dai, Highly efficient Pt/NaNbO3 nanowire

ro
652 photocatalyst : Its morphology effect and application in water purification and H2

653
-p
production, Appl. Catal. B: Environ. 205 (2017) 505–513.
re
654 [26] Y. Li, Z. Liu, Z. Guo, M. Ruan, X. Li, Y. Liu, Efficient WO3 Photoanode Modified by
lP

655 Pt Layer and Plasmonic Ag for Enhanced Charge Separation and Transfer To Promote
na

656 Photoelectrochemical Performances, ACS Sustainable Chem. Eng. 7 (2019) 12582–

657 12590.
ur

658 [27] X. Chen, Y. Li, X. Pan, D. Cortie, X. Huang, Z. Yi, Photocatalytic oxidation of
Jo

659 methane over silver decorated zinc oxide nanocatalysts, Nat. Commun. 7 (2016) 1–8.

660 [28] S. Chen, Y. Hu, L. Ji, X. Jiang, X. Fu, Preparation and characterization of direct Z-

661 scheme photocatalyst Bi2O3/NaNbO3 and its reaction mechanism, Appl. Surf. Sci. 292

662 (2014) 357–366.

663 [29] D. Kumar, S. Singh, N. Khare, Plasmonic Ag nanoparticles decorated NaNbO3

664 nanorods for efficient photoelectrochemical water splitting Int. J. Hydrogen Energy.

665 43 (2018) 8198–8205.

666 [30] S. Singh, N. Khare, Coupling of piezoelectric, semiconducting and photoexcitation

667 properties in NaNbO3 nanostructures for controlling electrical transport : Realizing an

30
668 efficient piezo-photoanode and piezo-photocatalyst, Nano Energy. 38 (2017) 335–

669 341.

670 [31] X. Li, G. Li, S. Wu, X. Chen, W. Zhang, Preparation and photocatalytic properties of

671 platelike NaNbO3 based photocatalysts, J. Phys. Chem. Solids. 75 (2014) 491–494.

672 [32] S. Z. Ji, H. Liu, Y. H. Sang, W. Liu, G. W. Yu, Y. H. Leng, Synthesis, Structure, and

673 Piezoelectric Property of Ferroelectric and Antiferroelectric NaNbO3 Nanostructures,

674 CrystEngComm. 16 (2014) 7598–7606.

675 [33] C. H. Liow, X. Lu, C. F. Tan, K. H. Chan, K. Zeng, S. Li, Ghim Wei Ho, Spatially

of
676 Probed Plasmonic Photothermic Nanoheater Enhanced Hybrid Polymeric–Metallic

ro
677 PVDF-Ag Nanogenerator, Small. 14 (2018) 1702268–1702276.

678 [34]
-p
S. Sharma, N. Khare, Sensitization of narrow band gap Bi2S3 hierarchical
re
679 nanostructures with polyaniline for its enhanced visible-light photocatalytic
lP

680 performance, Colloid Polym. Sci. 296 (2018) 14791489.


na

681 [35] D. Kumar, S. Sharma, N. Khare, Enhanced photoelectrochemical performance of

682 plasmonic Ag nanoparticles grafted ternary Ag/PaNi/NaNbO3 nanocomposite


ur

683 photoanode for photoelectrochemical water splitting, Renew. Energy. 156 (2020)
Jo

684 173–182.

685 [36] F. Dong, Q. Li, Y. Zhou, Y. Sun, H. Zhang, Z. Wu, In situ decoration of plasmonic

686 Ag nanocrystals on the surface of (BiO)2CO3 hierarchical microspheres for enhanced

687 visible light photocatalysis, Dalton Trans. 43 (2014) 9468–9480.

688 [37] N. K. Veldurthi, R. R. Jitta, G. Ravi, R. Guje, R. Velchuri, P. Venkataswamy, M.

689 Vithal, Fabrication and Visible – light induced Photocatalytic Activity of NaNbO3

690 Oriented Composite Photocatalyst, Chemistry.Select. 1 (2016) 2783–2791.

31
691 [38] R. Marschall, Semiconductor Composites : Strategies for Enhancing Charge Carrier

692 Separation to Improve Photocatalytic Activity, Adv. Funct. Mater. 24 (2014) 2421–

693 2440.

694 [39] S. Chengjie, F. Mingshan, H. Bo, C. Tianjun, W. Liping, S. Weidong, Synthesis of a

695 g-C3N4-sensitized and NaNbO3-substrated II-type heterojunction with enhanced

696 photocatalytic degradation activity, CrystEngComm. 17 (2015) 4575–4583.

697 [40] S. Sharma, S. Singh, N. Khare, Synthesis of polyaniline/CdS (nanoflowers and

698 nanorods) nanocomposites : a comparative study towards enhanced photocatalytic

of
699 activity for degradation of organic dye, Colloid Polym. Sci. 294 (2016) 917–926.

ro
700 [41] S. Kumar, R. Parthasarathy, A. P. Singh, B. Wickman, M. Thirumald, A. K. Ganguli,

701
-p
Dominant {100} facet selectivity for enhanced photocatalytic activity of NaNbO3 in
re
702 NaNbO3/CdS core/shell heterostructures, Catal. Sci. Technol. 7 (2017) 481–495.
lP

703 [42] S. Sharma, N. Khare, Hierarchical Bi2S3 nanoflowers: A novel photocatalyst for
na

704 enhanced photocatalytic degradation of binary mixture of Rhodamine B and

705 Methylene blue dyes and degradation of mixture of p-nitrophenol and p-chlorophenol
ur

706 Adv. Powder. Technol. 29 (2018) 3336–3347.


Jo

707 [43] H. Li, Y. Yu, M. B. Starr, Z. Li, X. Wang, Piezotronic-Enhanced

708 Photoelectrochemical Reactions in Ni(OH)2‑Decorated ZnO Photoanodes, J. Phys.

709 Chem. Lett. 6 (2015) 3410–3416.

710 [44] D. Kumar, S Sharma, N. Khare, Enhanced photoelectrochemical performance of

711 NaNbO3 nanofiber photoanodes coupled with visible light active g-C3N4 nanosheets

712 for water splitting, Nanotechnology. 31 (2020) 135402–135414.

713 [45] Y. Bai, H. Bai, K. Qu, F. Wang, P. Guan, D. Xu, W. Fan, W. Shi, In-situ approach to

714 fabricate BiOI photocathode with oxygen vacancies: Understanding the N2 reduced

715 behavior in photoelectrochemical system, Chem. Eng. J. 362 (2019) 349–356.

32
716 [46] V. Lingwal, B. Semwal, N. Panwar, Dielectric properties of Na1–xKxNbO3 in

717 orthorhombic phase, Bull. Mater. Sci. 26 (2003) 619-625.

718 [47] W. Cui, J. Shang, H. Bai, J. Hu, D. Xu, J. Ding, W. Fan, W. Shi, In-situ implantation

719 of plasmonic Ag into metal-organic frameworks for constructing efficient Ag/NH2-

720 MIL-125/TiO2 photoanode, Chem. Eng. J. 388 (2020) 124206-124215.

721 [48] A. G. Tamirat, W. N. Su, A. A. Dubale, H. Chen, B. J. Hwang, Photoelectrochemical

722 water splitting at low applied potential using NiOOH coated codoped (Sn, Zr) α-

723 Fe2O3 photoanode, J. Mater. Chem. A. 3 (2015) 5949-5961.

of
724 [49] P. Guan, H. Bai, F. Wang, H. Yu, D. Xu, W. Fan, W. Shi, In-situ anchoring Ag

ro
725 through organic polymer for configuring efficient plasmonic BiVO4 photoanode,

726 Chem. Eng. J. 358 (2019) 658–665.


-p
re
727 [50] S. Sharma, D. Kumar, N. Khare, Plasmonic Ag nanoparticles decorated Bi2S3
lP

728 nanorods and nanoflowers : Their comparative assessment for photoelectrochemical


na

729 water splitting, Int. J. Hydrogen Energy. 44 (2019) 3538–3552.

730
ur

731
Jo

732

733

734

735

736

737

738

739

740

33
741

742

743

744

745

746

747

748

of
749

ro
750

751
-p
re
752 List of figures captions
lP

753 Fig. 1 XRD patterns of bare NaNbO3 nanorods, Ag nanoparticle (inset of the image) and
na

754 Ag-NaNbO3 nanocomposite.

755 Fig. 2 TEM images of (a) bare NaNbO3 nanorods and (b) Ag-NaNbO3 nanocomposite and
ur

756 HRTEM images of (c) bare NaNbO3 and (d) Ag-NaNbO3 nanocomposite.
Jo

757 Fig. 3 Scanning electron microscope (SEM) image and elemental mapping images of Ag-

758 NaNbO3 nanocomposite showing the distribution of sodium (Na), niobium (Nb),

759 oxygen (O) and silver (Ag) (a-e) and (f) a typical EDX spectrum obtained for Ag-

760 NaNbO3 nanocomposite.

761 Fig. 4 (a) UV-Vis absorbance spectra and (b) plot of (αhν)2 vs. hν of pristine NaNbO3 and

762 Ag-NaNbO3 nanocomposite. (c) and (d) represent piezoresponse force microscopy

763 (PFM) hysteresis loop (phase) and butterfly loop (amplitude) of NaNbO3 nanorods

764 and Ag-NaNbO3, respectively.

34
765 Fig. 5 Absorbance spectra of photocatalytic decomposition activity of MB dye with 30 min

766 intervals for 3 h under six different conditions: (a) NaNbO3 (dark), (b) NaNbO3

767 (under light without ultra-sonic vibration), (c) NaNbO3 (under light with ultra-sonic

768 vibration), (d) Ag-NaNbO3 (dark), (e) Ag-NaNbO3 (under light without ultra-sonic

769 vibration), (f) Ag-NaNbO3 (under light with ultra-sonic vibration), (g) comparison of

770 the photocatalytic degradation proficiency done in various test conditions, (h)

771 demonstrate kinetics fit to the information, and (i) comparison of the rate constant (k)

772 values for the decomposition activity of MB dye for different test conditions.

of
773 Fig. 6. Effect of different active species scavengers for the photodegradation of MB over Ag-

ro
774 NaNbO3 nanocomposite under the combined effect of both light irradiation and ultra-

775
-p
sonic vibration, simultaneously for 180 min.
re
776 Fig. 7. Schematic of the possible working mechanism of Ag-NaNbO3 nanocomposite
lP

777 showing photogenerated charge carrier separation resulting in efficient degradation of


na

778 organic dye.

779 Fig. 8. (a) Current density (J) – potential (V) scans of NaNbO3 and Ag-NaNbO3 phototanode
ur

780 under dark and light conditions (scan speed 0.01 V/s): (i) NaNbO3 and (ii) Ag-
Jo

781 NaNbO3 without ultra-sonic vibration and (iii) NaNbO3 and (iv) Ag-NaNbO3 with

782 ultra-sonic vibration, (b) transient photocurrent curves for NaNbO3 and Ag-NaNbO3

783 (without ultra-sonic vibration), and NaNbO3 and Ag-NaNbO3 (with ultra-sonic

784 vibration) photoelectrodes at 1 V with 20 sec ON/OFF cycles, (c) EIS Nyquist plots

785 of NaNbO3 and AgNaNbO3 (without ultra-sonic vibration) and NaNbO3 and Ag-

786 NaNbO3 (with ultra-sonic vibration) photoanodes under light illumination. (solid dots

787 (•) denote experimental points, whereas solid lines (─) denote the simulated curves

788 using the equivalent circuit shown in the inset) and (d) IPCE photoresponse of

35
789 NaNbO3 and Ag-NaNbO3 (without ultra-sonic vibration) and NaNbO3 and Ag-

790 NaNbO3 (with ultra-sonic vibration) photoanodes.

791 Fig. 9. Mott-Schottky curves of (a) NaNbO3 (under dark without ultra-sonic vibration), (b)

792 Ag-NaNbO3 (under dark without ultra-sonic vibration), (c) NaNbO3 (under dark with

793 ultra-sonic vibration), (d) Ag-NaNbO3 (under dark with ultra-sonic vibration)

794 photoanodes at different frequencies in the dark condition, (e) Variation of the charge

795 injection efficiency (η , ) and (f) charge separation efficiency (η -& ) with an applied

796 bias for NaNbO3 (under light without ultra-sonic vibration), Ag-NaNbO3 (under light

of
797 without ultra-sonic vibration), NaNbO3 (under light with ultra-sonic vibration) and

ro
798 Ag-NaNbO3 (under light with ultra-sonic vibration) photoanodes.

799
-p
Fig. 10. Schematic diagram of PEC three electrode assembly (left) and energy band diagram
re
800 of Ag-NaNbO3/electrolyte interface with and without strain (right).
lP
na
ur
Jo

36
Highlights

• Piezoelectric NaNbO3 and Plasmonic Ag-NaNbO3 nanocomposites have been

synthesized.

• Ag-NaNbO3 exhibited higher photocatalytic and PEC activity as compared to

NaNbO3.

• Enhancement is attributed to the coupling of piezo-phototronic and plasmonic effect.

• Mechanism of charge transfer process for Ag-NaNbO3 nanocomposite was proposed.

of
ro
-p
re
lP
na
ur
Jo
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

None

of
ro
-p
re
lP
na
ur
Jo

You might also like