You are on page 1of 8

Journal of Materials Science: Materials in Electronics (2018) 29:3657–3664

https://doi.org/10.1007/s10854-017-8296-1

Improved photodegradation activity of ­SnO2 nanopowder


against methyl orange dye through Ag doping
N. Manjula1,2 · G. Selvan2 · A. R. Balu3

Received: 26 October 2017 / Accepted: 17 November 2017 / Published online: 21 November 2017
© Springer Science+Business Media, LLC, part of Springer Nature 2017

Abstract
Silver doped tin oxide ­(SnO2:Ag) nanopowders were synthesized by a simple soft chemical route with 0, 5, 10 and 15 wt%
concentrations of Ag. The structural, morphological, optical, photoluminescence and photocatalytic properties of the syn-
thesized samples were studied and the results obtained are reported in this paper. XRD studies confirm the polycrystalline
nature of the synthesized samples. The undoped and doped samples exhibit a strong (1 0 1) preferential growth. Decreased
crystallite size is observed with Ag doping. Nanosized grains were observed for the doped samples. Peak related to Sn–O–Sn
lattice vibration is observed for both the undoped and doped samples in the FTIR spectra. Peaks related to oxygen vacancies
were observed at 362 and 499 nm for all the samples in the PL spectra. Enhanced photocatalytic activity was observed for
the doped samples and the ­SnO2:Ag nanopowder with 10 wt% Ag doping concentration exhibited maximum photodegrada-
tion efficiency against the degradation of methyl orange dye.

1 Introduction pairs in semiconductor particles under illumination by vis-


ible/UV light. However, a quick recombination of the pho-
In recent years, metal oxide semiconductor photocatalysts togenerated electron and hole prevents augmentation of the
have attracted considerable attention due to their potential photocatalytic efficiency. To counter this challenge many
applications in hazardous waste remediation and energy strategies such as semiconductors combination, transition
problems with abundant solar light [1]. Among the metal metal doping and metal deposition has been adopted [5].
oxide semiconductor photocatalysts, tin oxide (­ SnO2) is a Eventhough, ­SnO2 seems to be a promising semiconduc-
wide band gap n-type semiconductor which has high opti- tor material for photovoltaic applications, its photocatalytic
cal transparency, electrical conductivity, high photosensi- efficiency is often restricted because of poor quantum yield
tivity and chemical sensitivity [2]. Due to its good optical caused by the fast recombination rate and ineffective utiliza-
properties, chemical and mechanical stability, ­SnO2 is used tion of photo-generated electron–hole pairs [6]. To improve
in many applications such as oxidation catalysis, gas sen- the photocatalytic efficiency of S­ nO2, doping has been per-
sors, solar cells, transparent conducting electrodes, lithium formed as the dopant ions could introduce lattice defects
ion batteries, etc [3]. In the field of photocatalysis, ­SnO2 which could lower its effective optical band gap, thus leading
seems to be an interesting material due to its different mor- to a better photoactivity [7]. Silver is noble metal of group
phologies, high photochemical stability, strong oxidizing I which possess high electrical conductivity which facili-
power and non-toxic nature [4]. The photocatalytic activ- tates fast electron transfer and low work function favoring
ity mainly relies on the generation of electron–hole (­ e−/h+) the formation of good band alignment and Ag doped/loaded
photocatalyst showed enhanced photocatalytic activity due
* G. Selvan to the surface plasmon resonance phenomenon under light
gselvan96@gmail.com irradiation [8]. Improved photocatalytic activity has been
reported earlier for Ag-doped CdS thin films [9] and Ag-
1
PG and Research Department of Physics, Bharathiyar doped ­TiO2 nanoparticles [10]. Motivated by these results,
University, Coimbatore, Tamilnadu, India
in the present work Ag-doped ­SnO2 ­(SnO2:Ag) nanopow-
2
PG and Research Department of Physics, Thanthai Hans ders were synthesized by a cost effective chemical method
Roever College, Perambalur, Tamilnadu, India
with different concentrations of Ag (0, 5, 10 and 15 wt%).
3
PG and Research Department of Physics, AVVM Sri Among the various chemical methods used to synthesize
Pushpam College, Poondi, Tamilnadu, India

13
Vol.:(0123456789)

3658 Journal of Materials Science: Materials in Electronics (2018) 29:3657–3664

undoped and doped nanoparticles the cost effective simple


chemical method adopted here has several advantages such
as low cost, rapid synthesis, ease to dope, low preparation
temperature, etc [11]. The synthesized ­SnO2:Ag nanopow-
ders was characterized by techniques like XRD, SEM, TEM,
UV–Vis–NIR, FTIR and PL. The photocatalytic activity was
tested against the photodegradation of methyl orange dye
under visible light irradiation and the results obtained are
reported.

2 Experimental details

Ag-doped ­SnO2 ­(SnO2:Ag) nanopowders were prepared by


a simple soft chemical route with 0, 5, 10 and 15 wt% Ag
doping concentrations. Undoped S ­ nO2 nanopowder was pre-
pared using tin(II) chloride [­ SnCl2·2H2O] as the precursor
salt. Tin(II) chloride with 0.1 M was dissolved in a mixture
of HCl and de-ionized water in the volume ratio of 1:13 in
a total volume of 140 ml. The pH value of this solution was
raised to 10 by adding 10 ml liquid ammonia. The resultant
solution was stirred well for 2 h and kept undisturbed for Fig. 1  XRD patterns of ­SnO2:Ag nanopowders
4 h. The settled precipitates were cleaned two times with
de-ionized water and then calcined at 200 °C for 1 h in a
muffle furnace. The residual product was then crushed using indicating that silver ions were well incorporated into the
an agate mortar to get grey colored S ­ nO2 nanopowders. To ­SnO2 lattice. This is in contrast with the results reported by
achieve Ag doping, silver nitrate [Ag(NO3)2] with concentra- Mouchaal et al. [12] where they detected A ­ g2O in the XRD
tions 5, 10 and 15 wt% was added to the precursor solution pattern of 4 wt% Ag, in codoped S ­ nO2 thin films, which
and the same procedure used to synthesize pure S ­ nO2 nano- confirms that the chemical route adopted here is very effec-
powders was adopted to get ­SnO2:Ag nanopowders. X-ray tive in synthesizing homogenous S ­ nO2:Ag nanopowders. It
diffractometer (PANanalytical-PW 340/60 X’pert PRO) with was also found that the major (1 0 1) peak of pure ­SnO2
CuKα radiation (λ = 1.5406 Å) is used to analyse the crys- shifted towards lower 2θ angle with Ag doping which might
tal structure of the samples. Scanning electron microscope be due to the expansion of PbS lattice due to the substitution
(HITACHI S-3000 H) and transmission electron microscope of smaller sized S­ n2+ ions (0.93 Å) with larger sized A ­ g+
(200 kV Tecnai-20 G2) are used to examine the surface mor- ions (1.22 Å). Similar result has been reported earlier by
phology of the samples. The functional groups present in Narasimman et al. [13] for Ag-doped CdS thin films. The
the samples were identified using Perkin Elmer RX-1 FTIR lattice constants (a and c) of pure S
­ nO2 (Table 1) increases
spectrophotometer. Optical studies were performed using with increase in Ag doping concentration. The calculated
Perkin Elmer UV–Vis–NIR double beam spectrophotometer lattice parameter values of the ­SnO2:Ag nanopowders are
(LAMBDA 35). Photocatalytic activity of the samples was found to be higher than the bulk (a = 4.737 Å, c = 3.186 Å)
evaluated by the photodegradation of methyl orange (MO) suggesting that the samples are under strain with Ag doping.
dye under visible light, irradiated from a 100 W incandes- The crystallite size (D) and strain (ε) values of the
cent lamp. ­SnO2:Ag nanopowders are calculated using the formulae
[14]:
0.9 𝜆
3 Results and discussion D= (1)
𝛽 cos 𝜃
Figure 1 shows the XRD patterns of the S ­ nO2:Ag nanopo-
wders. Polycrystalline nature is observed for all the sam- 𝛽 cos 𝜃
𝜀= (2)
ples and the diffraction peaks fit well with tetragonal ­SnO2 4
(JCPDS Card No. 88-0287). All the samples exhibit a strong where β is the full width at half maximum value of the (1
(1 0 1) preferential growth. No peaks related to metallic Ag 0 1) peak in radians, λ is the wavelength of the X-ray used
or ­Ag2O were detected within the instrumental sensitivity,

13
Journal of Materials Science: Materials in Electronics (2018) 29:3657–3664 3659

Table 1  Crystallite size, Ag doping con- Crystallite Strain ‘ε’ × 10−3 Lattice constant Photodegradation efficiency (%)
strain, lattice parameter and centration (wt%) size ‘D’ (nm)
photodegradation efficiency a (Å) c (Å) 60 min 120 min 180 min
values of the ­SnO2:Ag
nanopowders 0 27.62 0.758 4.768 3.210 41.67 56.67 75.04
5 25.14 0.864 4.842 3.287 53.85 71.53 88.22
10 23.04 0.957 4.912 3.363 68.85 84.43 93.44
15 23.91 0.904 4.978 3.411 62.50 81.25 90.18

(1.5406 Å) and θ is the Bragg angle. The calculated D and 10 and 15 wt% Ag concentrations were found to be equal to
𝜖 values compiled in Table 1, show opposite trend behav- 3.27, 3.35, 3.45 and 3.4 eV, respectively. Increased band gap
ior. Reduction in crystallite size may be due to the reduc- values were observed for the doped samples which might
tion in diffusion rate which makes the nanopowders intact be due to size induced quantum confinement present in the
and due to increased strain in S ­ nO2 lattice with Ag doping. Ag-doped SnO2 nanopowders [22]. Similar enhancement
Decreased crystallite sizes result in a drastical decrement in Eg value has been reported earlier for Mn-doped ­SnO2
of photoinduced charges’ recombination, which efficiently nanoparticles [23]. Increased band gap values are impor-
promote photocatalytic reactions [15]. tant in preventing the electron–hole recombination which
Figres 2 and 3 shows the SEM and TEM images of (a) 0, (b) can enhance the photocatalytic activity of Ag-doped ­SnO2
5, (c) 10 and (d) 15 wt% Ag-doped S ­ nO2 nanopowders. All the nanopowders [24].
surfaces appear to be composed of grains with different sizes The room temperature PL spectra of the S ­ nO2:Ag nano-
and shapes. Nanosized grains are also visible. All the samples powders are shown in Fig. 6. The peaks at 499 and 362 nm
possessed quite large surface area, which was believed to be are mainly due to oxygen vacancies which form donor
an important factor in photocatalytic reactions, in accordance levels in S­ nO2 [25]. The blue emission peaks at 470 and
with small crystallite sizes as evinced from XRD studies. 459 nm are due to the electronic transitions from the tin
Figure 4 shows the FTIR spectra of (a) undoped and (b) Ag- interstitials and/or oxygen vacancies [26]. The peak arising
doped ­SnO2 nanopowders. The peaks observed in the range from the recombination of the electrons from the conduc-
3150–1500 cm−1 for both the undoped and doped samples are tion band and holes in the valence band is observed at
attributed to O–H stretching vibrations [16]. The peak cor- 441 nm [27]. The near band edge (NBE) peak due to free
responding to –CH3 bending mode is observed at 1402 cm−1 exciton recombination is observed at 408 nm [28]. The
for both the undoped and doped samples [17]. This might be peak attributed to higher level excitonic emission related
due to the addition of liquid ammonia to the precursor solu- to quantum confinement is observed at 324 nm [29].
tion while synthesizing the samples. The peaks at 1136 and The photocatalytic activity of the S ­ nO2:Ag photocata-
1023 cm−1 observed for the undoped sample is attributed to lysts was examined by studying the oxidative degradation
Sn–O–Sn lattice vibrations [18]. The peak at 1136 cm−1 got of methyl orange dye under visible light irradiation. The
shifted to 1121 cm−1 with ­Ag+ doping. The peak at 780 cm−1 photocatalytic experiment was carried out by preparing
observed for the undoped sample corresponds to N–O stretch- reaction suspensions with 6 mg of the photocatalysts into
ing vibration [19]. The peaks observed in the wavenumber 100 ml of 0.25 M MO solution. The dye solution with
range 520–620 cm−1 are related to metal oxygen (MO) bond the photocatalysts was stirred in dark for 1 h to maintain
which corresponds to the stretching vibrations of Sn–O [20]. absorption/desorption equilibrium and then illuminated
Figure 5a shows the absorbance spectra of the S ­ nO2:Ag under visible light at different time intervals (0, 60, 120
nanopowders with 0, 5, 10 and 15 wt% Ag concentrations. and 180 min). After every 1 h interval, 6 ml solution was
It can be seen that the absorption edge of pure S ­ nO2 shifts withdrawn from the suspension and the catalysts were
towards lower wavelengths with Ag doping. The optical band separated from it using a centrifuge. The concentration of
gap ­(Eg) of the S
­ nO2:Ag nanopowders was evaluated from the MO was analyzed by using Perkin Elmer UV–Visible–NIR
reflection spectra using the Kubelka–Munk function F(R) [21]: double beam spectrophotometer at λ = 470 nm. The photo-
degradation efficiency of the ­SnO2:Ag photocatalysts was
(1 − R)2 calculated using the relation [30]:
F(R) = (3)
2R ( )
C
where R represents the percentage reflectance. The direct 𝜂 = 1− × 100 (4)
C0
band gap values was calculated by extrapolating the linear
part of the plots of [F(R) × hυ]2 versus hυ (Fig. 5b) at α = 0. where C is the concentration of MO after irradiation and C0
The band gap values of the S ­ nO2:Ag nanopowders with 0, 5, is the concentration before light irradiation. The calculated

13

3660 Journal of Materials Science: Materials in Electronics (2018) 29:3657–3664

Fig. 2  SEM images of a 0, b 5, c 10 and d 15 wt% Ag-doped S


­ nO2 nanopowders

efficiency values are compiled in Table 1. It is observed that doped samples might be due to the effective charge trans-
the photodegradation efficiency of pure ­SnO2 got enhanced fer ability of ­Ag+ ions. This is in accordance to the results
with Ag doping and among the doped samples the 10 wt% reported by Vignesh et al. [31]. Figure 7a shows the absorp-
Ag-doped ­SnO2 photocatalyst exhibited a maximum effi- tion spectrum of 10 wt% Ag-doped ­SnO2 photocatalyst as a
ciency of 93.44%. The increased activity observed for the function of irradiation time. The absorption peak decreases

13
Journal of Materials Science: Materials in Electronics (2018) 29:3657–3664 3661

Fig. 3  TEM images of a 0, b 5, c 10 and d 15 wt% Ag-doped ­SnO2 nanopowders

confirming the fact that MO molecules degrades with irra- Under visible light irradiation, electron–hole ­(e−/h+) pairs
diation time and almost 93.44% of the MO molecules gets are produced from the ­SnO2 photocatalyst by the equation:
degraded after 180 min. The mechanism involved in the pho-
SnO2 + h𝛾 → e− + h+
tocatalytic activity of the S
­ nO2:Ag photocatalyts (Fig. 7b)
is as follows:

13

3662 Journal of Materials Science: Materials in Electronics (2018) 29:3657–3664

Fig. 4  FTIR spectra of S
­ nO2:Ag nanopowders

Fig. 5  a Absorbance spectra of ­SnO2:Ag nanopowders, b plots of [F(R) × hυ]2 versus hυ of ­SnO2:Ag nanopowders

The holes generated reacts with hydroxyl groups and


adsorbed ­H 2O to form hydroxyl radicals (­OH ·) by the
equations:

OH− + h+ → OH⋅

H2 O + h+ → OH⋅ + H+
The ­Ag+ ions act as electron traps generating superoxide
ions ­(O2·−) reaction with ­O2 by the equation:
Ag + e− → Ag−

Ag− + O2 → O⋅−
2
+ Ag
­ H· by
The final product of the reduction may also be O
Fig. 6  PL spectra of S
­ nO2:Ag nanopowders the equation:

13
Journal of Materials Science: Materials in Electronics (2018) 29:3657–3664 3663

­ nO2 photocatalyst, b photocatalytic activity of ­SnO2:Ag nanopowders


Fig. 7  a Absorbance spectrum of 10 wt% Ag-doped S

2O⋅−
2
+ 2H+ → 2OH⋅ References
The hydroxyl radicals and superoxide ions are the reac-
1. K. Yin, M. Shao, Z. Zhang, Z. Lin, Mater. Res. Bull. 47, 3704
tive oxygen species which degrade MO by the equation: (2012)
2. H. Yuan, J.Q. Xu, Int. J. Appl. Chem. 1, 241 (2010)
OH⋅ + O⋅− + MO → CO2 + H2 O (degradation of MO molecules)
2 3. S. Ferrari, L.G. Pampilo, F.D. Saccone, Mater. Chem. Phys.
The decreased photocatalytic activity observed for 177, 206 (2016)
the 15 wt% Ag-doped S ­ nO2 photocatalyst might be due 4. P. Kamaraj, R. Vennila, M. Arthanareeswari, S. Devikala, World
J. Pharm. Sci. 3, 382 (2014)
to increased crystallite size, loss of stoichiometry which 5. Z. Nasir, M. Shakir, R. Wahab, M. Shoeb, P. Alam, R.H. Khan,
leads to quantum tunnelling. M. Mobin, Int. J. Biol. Macromol. 94, 554 (2017)
6. L.R. Zheng, Y.H. Zheng, C.Q. Chen, Y.Y. Zhan, X.Y. Lin, Q.
Zheng, K.M. Wei, J.F. Zhu, Inorg. Chem. 48, 1819 (2009)
7. M.A.M. Al-Hamdi, M. Sillanpaa, J. Duha, J. Mater. Sci. 49,
5151 (2014)
4 Conclusion 8. D.D. Lin, H. Wu, R. Zhang, W. Pan, Chem. Mater. 21, 3479
(2009)
SnO2:Ag nanopowders were successfully synthesized by 9. M. Ristova, M. Ristov, P. Tosev, Thin Solid Films 315, 301
(1998)
a cost effective simple chemical route with 0, 5, 10 and 10. L. Zhang, J.C. Yu, H.Y. Yip, Q. Li, K.W. Kwong, A.W. Xu, P.K.
15 wt% Ag doping concentrations. Structural studies con- Wong, Langmuir 19, 10372 (2003)
firmed the polycrystalline nature of the samples. Undoped 11. M. Suganya, D. Prabha, S. Balamurugan, A.R. Balu, J. Mater. Sci.
and Ag-doped ­S nO 2 nanopowders exhibited tetragonal Mater. Electron. 28, 5344 (2016)
12. Y. Mouchaal, A. Enesca, C. Mihoreanu, A. Khelil, A. Duta, Mater.
crystal structure with a strong (1 0 1) preferential growth. Sci. Eng. B 199, 22 (2015)
Nanosized grains are evinced from the TEM images of the 13. V. Narasimman, V.S. Nagarethinam, K. Usharani, Mater. Res.
synthesized samples. Photocatalytic activity of pure S ­ nO2 Innov. (2016). https://doi.org/10.1080/14328917.2016.1264857
against the degradation of methyl orange got enhanced 14. S. Ravishankar, A.R. Balu, Surf. Eng. 33, 506 (2017)
15. K. Zhang, D. Jing, Q. Chen, L. Guo, Int. J. Hydrogen Energy, 35,
with Ag doping. The enhanced photocatalytic activities 2048 (2010)
observed make ­SnO2:Ag nanopowders suitable for waste 16. M.A. Abbasi, D. Ghanbari, M.S. Niasari, M. Hamadanian, J.
water treatment and other bioremediation processes. Mater. Sci. 27, 4800 (2016)
17. S. Balamurugan, A.R. Balu, K. Usharani, M. Suganya, S. Anitha,
Acknowledgements  The authors thank the Director, STIC Cochin for D. Prabha, S. Ilangovan, Pac. Sci. Rev. A 18, 228 (2016)
the TEM analysis. 18. M.A.A. Dakhel, Powder Technol. 237, 333 (2013)

13

3664 Journal of Materials Science: Materials in Electronics (2018) 29:3657–3664

19. K.K. Onchoke, C.M. Hadad, P.K. Dutta, J. Phys. Chem. A 110, 26. K. Usharani, A.R. Balu, V.S. Nagarethinam, Surf. Eng. 32, 829
76 (2006) (2016)
20. N. Manjula, G. Selvan, J. Mater. Sci. (2017). https://doi. 27. S.J. Gnanamuthu, S.J. Jeyakumar, I.K. Punithavathi, K. Paras-
org/10.1007/s10854-017-7380-x uram, V.S. Nagarethinam, A.R. Balu, Trans. Indian Inst. Met. 70,
21. P. Pascariu, A. Airinei, M. Grigoras, N. Fifere, L. Sacarescu, N. 1503 (2017)
Lupu, L. Stoleriu, J. Alloys Compd. 668, 65 (2016) 28. Y. Jin, Q. Cui, K. Wang, J. Hao, Q. Wang, J. Zhang, J. Appl. Phys.
22. T. Sivaraman, V.S. Nagarethinam, A.R. Balu, Surf. Eng. 32, 596 109, 053521 (2011)
(2016) 29. J. Srivind, V.S. Nagarethinam, A.R. Balu, Mater. Sci. Pol. 34, 393
23. K. Anandan, V. Rajendran, Superlattices Microstruct. 85, 185 (2016)
(2015) 30. M. Suganya, A.R. Balu, D. Prabha, S. Anitha, S. Balamurugan, J.
24. L. Kumaresan, M. Mahalakshmi, M. Palanichamy, V. Murugesan, Mater. Sci. (2017). https://doi.org/10.1007/s10854-017-8007-y
Ind. Eng. Chem. Res. 49, 1480 (2010) 31. K. Vignesh, R. Hariharan, M. Rajarajan, A. Suganthi, Sol. State
25. J. Jeong, S.P. Choi, C.I. Chang, D.C. Shin, J.S. Park, B.T. Lee, Sci. 21, 91 (2013)
Y.J. Park, H.J. Song, Sol. State. Commun. 127, 595 (2003)

13

You might also like