You are on page 1of 7

Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 128 (2014) 218–224

Contents lists available at ScienceDirect

Spectrochimica Acta Part A: Molecular and


Biomolecular Spectroscopy
journal homepage: www.elsevier.com/locate/saa

Preparation, structural and morphological studies of Ni doped titania


nanoparticles
B. Rajamannan a,⇑, S. Mugundan b, G. Viruthagiri b, N. Shanmugam b, R. Gobi b, P. Praveen b
a
Department of Engineering Physics (FEAT), Annamalai University, Annamalainagar 608002, India
b
Department of Physics, Annamalai University, Annamalainagar 608002, India

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 No change in anatase phase of bare


and Ni doped TiO2 was predicted even
after calcinations at 500 °C.
 No alkoxy groups were identified in
FTIR spectra.
 The peaks which were existing in bare
TiO2, were blue shifted as result of
doping.
 Particles are almost in spherical
shape with uniform size distribution.
 TEM image that the particle size is 8–
16 nm, which are in good agreement
with the XRD results.

a r t i c l e i n f o a b s t r a c t

Article history: TiO2 nanoparticles doped with different weight percentages (4%, 8%, 12% and 16%) of nickel contents were
Received 16 October 2013 prepared by a modified sol–gel method using Titanium tetra iso propoxide and nickel nitrate as precur-
Received in revised form 4 February 2014 sors and 2-propanol as a solvent. X-ray diffraction studies show that the as prepared and annealed prod-
Accepted 19 February 2014
ucts show anatase structure with average particle sizes running between of 8 and 16 nm. FTIR results
Available online 12 March 2014
demonstrate the presence of strong chemical bonding at the interface of TiO2 nanoparticles. The optical
properties of bare and doped samples were carried out using UV-DRS and photoluminescence measure-
Keywords:
ments. The surface morphology and the element constitution of the nickel doped TiO2 nanoparticles were
TiO2 nanoparticles
Crystalline size
studied by scanning electron microscope attached with energy dispersive X-ray spectrometer arrange-
FT-IR ment. The non linear optical properties of the products were confirmed by Kurtz second harmonic
Optical properties generation (SHG) test and the output power generated by the nanoparticle was compared with that of
Spherical uniform structure potassium di hydrogen phosphate (KDP).
SHG Efficiency. Ó 2014 Elsevier B.V. All rights reserved.

Introduction high performance, photostability (i.e., resistance to photo corro-


sion), low cost, nontoxicity and availability. It is one of the most
Titanium dioxide (TiO2) has been extensively studied in the past promising photocatalysts for environmental cleanup, photo gener-
few years owing to its technological applications. It has been well ation of hydrogen from water, and solar energy utilization [1–3].
recognized that among the semiconductors employed in TiO2 crystallizes in three different structures: brookite, anatase,
photocatalysis, TiO2 is the most effective photocatalyst due to its and rutile, among them, the anatase phase has the highest photo-
catalytic activity compared with the rutile and brookite phases
[4,5]. However, the rutile structure is the most stable form at high
⇑ Corresponding author. Tel.: +91 9443283241.
temperature, and anatase and brookite both can be converted to
E-mail address: rajamannanb@gmail.com (B. Rajamannan).

http://dx.doi.org/10.1016/j.saa.2014.02.116
1386-1425/Ó 2014 Elsevier B.V. All rights reserved.
B. Rajamannan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 128 (2014) 218–224 219

rutile upon heating [6]. These three phases can be commonly de- and 30 mA. The functional groups were idenified by AVATAR 330
scribed as constituted by arrangements of the same building Fourier-transform infrared spectrometer in which the IR spectra
block-Ti–O6 octahedron in which Ti atom is surrounded by six oxy- were recorded by diluting the mixed powders in KBr and in the
gen atoms situated at the corners of a distorted octahedron. Inspite wavelength between 4000 and 400 cm1. The band-gap energy
of the similarities in building blocks of Ti–O6 octahedra for these and the particle size were measured at wavelength in the range
polymorphs, the electronic structures are significantly different. of 200–2500 nm by UV–Vis–NIR spectrophotometer (Varian/Carry
Titanium dioxide (TiO2) offers a great attention due to its 5000) equipped with an integrating sphere and the baseline cor-
unique physical and chemical properties such as wide energy gap rection was performed using a calibrated reference sample of pow-
(3.2 eV), high electrochemical stability, refractive index, and dered barium sulphate (BaSO4). The photoluminescence (PL)
dielectric constant [7,8]. spectra of the samples were recorded with a Spectroflurometer
Titanium dioxide (TiO2) is a well known semiconductor photo- (Jobin Yvon, FLUOROLOG – FL3-11). The surface morphology of
catalytic material that has attracted much attention due to its pure and Ni–TiO2 was observed by a scanning electron microscope
biological and chemical inertness, strong photo oxidizing power, (SEM: Hitachi, S-3400 N). The dispersion of titanium, oxygen and
cost effectiveness, and long term stability against photo and chem- nickel in the products was characterized by energy dispersive
ical corrosion [9]. Among the semiconductors, titanium dioxide is X-ray elemental analysis (EDX: Thermo Super Dry II) equipped to
the most active photocatalyst and commonly used in organic com- the scanning electron microscope. Transmission electron
pounds degradation. However, the widespread technological use of microscope (TEM) images were taken using a technai t20 operated
TiO2 is impaired by its wide band gap (3.2 eV for crystalline at voltage of 200 kV. The NLO property of the materials was
anatase phase), the fast charge-carrier recombination, and the confirmed by the Kurtz powder second harmonic generation
low interfacial charge-transfer rates [10]. (SHG) test. The materials were illuminated using Spectra physics
Nanosized TiO2 has been fabricated using sol–gel, sputtering, Quanta Ray DHS2. The SHG radiations of 532 nm green light were
combustion flame, and thermal plasma [11,12]. Although the sol– collected by a photomultiplier tube (PMT-Hamasu R2059) after
gel method is considered as a suitable method to synthesize being monochromated (monochromator – Czerny-Turner) to col-
ultra-fine particles, this method needs a large quantity of solution, lect only the 532 nm radiation. A Q-switched Nd–YAG laser whose
longer processing time and heat treatment for crystallization since output was filtered through 1064 nm narrow pass filter was used
amorphous TiO2 has a very little photocatalytic activity. In this for this purpose. The input power of the laser beam was measured
work, we report on the sol–gel synthesis and characterization of to be 4.5 mJ/pulse.
TiO2 nanomaterials doped with nickel in different atomic weight
percentages (4%, 8%, 12% and 16%) capable to have enhanced opti- Results and discussion
cal properties. Since, Titanium tetra iso-propoxide is non toxic and
ecofriendly, in the present in experiment it is used as a titanium Powder X-ray diffraction study (XRD)
source. Also, the structural, functional and morphological analyses
of the nanopowders were performed. The NLO activity of the TiO2 Fig. 1 shows X-ray powder diffraction patterns of pure and Ni
nanostructures was investigated for the first time. doped TiO2 nanoparticles calcined at 500 °C. All observed peaks
could be corresponding to the anatase phase tetragonal structure
Sample preparation of TiO2 (JCPDS card No. 21-1272). There is no indication of rutile
and brookite phase formation on 500 °C of sintering. The peaks at
Synthesis of bare and Ni doped TiO2 nanoparticles 2h = 25.5° (1 0 1) and 48° (2 0 0) were in anatase phase of TiO2
[13]. However, few peaks related to nickel were identified at
To synthesize Ni doped TiO2, 90 ml of 2-propanol (C3H8O) and 2h = 33.10° and 36.88° indicate the presence of nickel [14]. The Dif-
nickel nitrate [Ni(NO3)2] in 10 ml aqueous solution with different fraction Plane (1 0 1) shows a sharp and higher intensity for un-
concentrations (4%, 8%, 12% and 16%) were mixed drop by drop. doped and Ni doped samples due to nature TiO2 nanoparticles.
The mixture was stirred magnetically at room temperature until From the obtained peak the average nanocrystalline sizes was
a homogenous solution was obtained. Then 10 ml of Titanium tetra measured according to Debye Scherrer formula (Eq. 1) and were
iso-propoxide was added drop by drop to the above mixture. The presented in Table 1.
entire was continuously stirred for 5 h using a magnetic stirrer. Kk
After stirring the solution was filtered by using whatman filter pa- D¼ Å ð1Þ
bCos h
per and washed several times using deionized water. The precipi-
tate formed was dried at 80 °C for 5 h to evaporate organic
residues. Then the dry gel was calcined at 500 °C and 800 °C for
2 h to obtain desired anatase and rutile TiO2 nanoparticles. Then
the calcined powders were grained in an agate mortar to avoid
agglomeration.
The bare TiO2 nanoparticles were synthesized by following the
same procedure without doping material.

Characterizations

The crystalline phase and particle size of pure and Ni–TiO2


nanoparticles were analyzed by X-ray diffraction (XRD) measure-
ment which was carried out at room temperature by using
XPERT-PRO diffractometer system (scan step of 0.05° (2h),
counting time of 10.16 s per data point) equipped with a Cu tube
for generating Cu Ka radiation (k = 1.5406 Å) as an incident beam
in the 2-theta mode over the range of 20–80°, operated at 40 kV Fig. 1. X-ray diffraction pattern for bare TiO2 and Ni doped TiO2 nanoparticle.
220 B. Rajamannan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 128 (2014) 218–224

where D is the crystallite size; K is the shape factor; k is the


0.154 nm; b is the full width at half maximum and h is reflection
angle.
The mean crystallite size of bare TiO2 is 15.31 nm, whereas on
different levels (4%, 8%, 12% and 16%) of Ni doping the estimated
sizes are 8.77, 16.04, 16.82 and 11.77 nm respectively.
When compared to bare TiO2 all the Ni doped products show in-
creased size as a result of improved crystallinity (except 4% and
16%). The reduction of particle size from 15.31 to 8.77 nm on
4 wt.% of Ni doping reveals its optimum level.

Fourier Transform Infrared Spectroscopy (FTIR)

Fig. 2 shows the FTIR spectra of the obtained bare and Ni doped
TiO2 nanoparticles after calcined at 500 °C for 5 h. The broad band
around 3400 cm1 is attributed to O–H stretching, and the peak
near 1600 cm1 is due to O–H bending which are related to phys-
ically absorbed moisture. The IR band observed from 400 to
900 cm1 corresponds to the Ti–O stretching vibrations [15,16].
The band at 1634 cm1 was attributed to stretching vibrations of
OH. The peaks in between 2924 and 2843 cm1 were assigned to Fig. 2. FTIR spectra pattern for bare TiO2 and Ni doped TiO2 nanoparticle.
C–H stretching vibrations of alkane groups. The alkane and carbox-
ylic groups come from Titanium tetra isopropoxide and 2-Propanol
(precursor material), when we used in the synthesis process. In
addition, a broad absorption band between 500 and 1000 cm1 is
ascribed to the vibration absorption of the Ti–O–Ti linkages in
TiO2 nanoparticles [17]. However, the vibrational bands between
1300 and 4000 cm1 are mainly assigned to the chemisorbed and
physisorbed H2O and CO2 molecules on the surface of the com-
pound. In addition, the band centered at 1371 cm1 is assigned
to bending vibrations of C–H bond in the species linking
–Ti–O–Ti– structural network [18,19]. When metal ions are doped
to the surface of TiO2, the absorption band significantly transforms
and simultaneously new absorption band appears.

Ultra violet-diffuse reflectance spectra (UV-DRS)

Fig. 3 shows UV-DRS spectra of bare and Ni doped TiO2 nano-


particles obtained after calcined at 500 °C for 5 h. The TiO2 material
showed an intense absorption in the UV region and the absorption
edge of titania can be easily discerned. Compared with bulk TiO2 a
shift of the reflectance spectra of the synthesized TiO2 nanocrystals
towards the lower wavelength region were observed. This result is
Fig. 3. UV-DRS spectra for bare TiO2 and Ni doped TiO2 nanoparticle-direct
identical with the reported one [20]. transition.
The band gap energy (Eg) of pure TiO2 was obtained from the
wavelength value corresponding to the intersection point of the orbitals of the oxide anions) to the conduction band (mainly
vertical and horizontal part of the spectrum, using (Eq. (2)) formed by 3dt2g orbitals of the Ti4+ cations) [21]. However, on dop-
ing the absorption edge was blue shifted. Therefore, the increase of
hc 1240
Eg ¼ eV; Eg ¼ eV ð2Þ incorporating Ni content, the band-gap energy of the TiO2 nanopar-
k k ticles increased systematically.
where Eg is the band gap energy (eV); h is the Planck’s constant At the same time on a higher level of doping (8–12%) in addition
(6.626  1034 Js); c is the light velocity (3  108 m/s) and k is the to UV band visible bands are also predicted. These bands are posi-
wavelength (nm). tioned at 571,600 and 585 nm respectively for 8%, 12% and 16% of
The absorption edge of nano TiO2 was noted at 345.49 nm, cor- Ni doping.
responding to a band gap of 3.5891 eV, which is usually ascribed to The obtained values of Eg are 3.58, 3.65, 3.61, 3.83 and 4.0 eV for
charge-transfer from the valence band (mainly formed by 2p the TiO2 and TiO2 doped with 4%, 8%, 12% and 16% nickel as shown
in Table 2 respectively, suggesting the change the band gap energy
of TiO2 on Ni doping.
Table 1
Crystallite sizes for the bare and Ni doped TiO2 nanoparticle.

Samples Crystallite size (nm) Photoluminiscence (PL)


Bare TiO2 15.31
4% Ni doped TiO2 8.77 The room temperature PL spectra were recorded for bare and 8%
8% Ni doped TiO2 16.04 of nickel doped TiO2 with an excitation wavelength of 325 nm
12% Ni doped TiO2 16.82
(Fig. 4). From the PL spectra, it is predicted that the bare TiO2
16% Ni doped TiO2 11.77
exhibits three emission bands, one at UV region (377 nm) and
B. Rajamannan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 128 (2014) 218–224 221

Table 2 drastically decreased with the incorporation of Ni in the TiO2 nano-


Bandgap energy for bare and Ni doped TiO2 nanoparticle. particles. The broad band emission from the PL spectrum at room
Samples Bandgap (eV) temperature indicates that the TiO2 and Ni–TiO2 nanoparticle have
Bare TiO2 3.58 good luminescence quality [24]. The intensity of the PL peaks is an
4% Ni doped TiO2 3.65 indicator for the electrons/hole recombination rate, low intensity
8% Ni doped TiO2 3.61 indicates low recombination and vice versa. As shown in the figure,
12% Ni doped TiO2 3.83 incorporation of Ni–TiO2 nanoparticle leads to decrease the inten-
16% Ni doped TiO2 4.0
sity which preliminarily claims good photoactivity for the
Ni-doped TiO2 nanoparticles compared with the undoped ones.
On doping these bands are shifted to lower wavelength regions
with reduced intensity. However, the broadness of the bands has
been enhanced on doping; the broadness may be open up an
enhanced luminescent property of the doped product.

Scanning electron microscope (SEM with EDX)

Fig. 5(a and b) shows the scanning electron microscope (SEM)


image of bare TiO2 nanoparticles. As shown in figure most of the
particles are almost in spherical shape with uniform size distribu-
tion and they consists of either some single particles or cluster of
particles. But, on (8 wt.%) Ni doping multi sized particles are pre-
dicted with non uniform distribution (Fig. 6(a and b)). This result
is in accordance with the fact that Ni doping can alter the TiO2 par-
ticles growth [25].
The energy dispersive X-ray (EDX) spectrum is used to analyze
the Ni doped TiO2 nanoparticles, and the results are shown in
Figs. 5(c) and 6(c). Peaks of Ti, O and Ni are obviously found in
Fig. 4. PL spectra for bare TiO2 and Ni doped TiO2 nanoparticle. the spectra, confirming the formation of Ni doped TiO2 nanoparti-
cles [26].
the other two at visible regions (431 and 587 nm). A band at
377 nm is attributed to the first vibronic fluorescence band. The Transmission electron microscope (TEM with SAED)
obvious PL peak at about 431 nm may be due to band edge free
excitons and this surface emission is attributed to indirect transi- Fig. 7(a and b) shows the transmission electron microscope
tion X1a to C1b and linked to exciton recombination in shallow (TEM) micrographs of bare TiO2 nanoparticles. As shown in figure,
trapped surface state [22,23]. most of the particles are get agglomerated and showing a cross
The PL spectrum of doped TiO2 is dominated by the broad band linked chain structure with sizes in the range of 15.31 nm. The
emission in the range of 580–620 nm. The PL intensity was SAED pattern of bare TiO2 nanoparticles reveals the formation of

Fig. 5. SEM micrograph of TiO2 nanoparticles (a and b) and corresponding to EDX diagram (c).
222 B. Rajamannan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 128 (2014) 218–224

Fig. 6. SEM micrograph of 8% Ni doped TiO2 nanoparticles (a and b) and corresponding to EDX diagram (c).

Fig. 7. TEM image of TiO2 nanoparticles (a and b) and corresponding to SAED pattern (c).
B. Rajamannan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 128 (2014) 218–224 223

Fig. 8. TEM image of 8% Ni doped TiO2 nanoparticles (a and b) and corresponding to SAED pattern (c).

Table 3 green light (532 nm) emission from both the materials (bare & Ni
Comparison of SHG efficiency of bare and 8% Ni doped TiO2 nanoparticles. doped TiO2) and KDP. Thus the efficiency results show an output
Samples Input (mJ/ Output KDP SHG of 8.0 mV for pure TiO2, when compared to 17.6 mV of KDP with
pulse) (mV) (mV) efficiency same input laser power of 2.6 m J/pulse. The relative SHG conver-
Bare TiO2 4.5 8.0 10.2 0.78 sion efficiency of bare TiO2 is found o be about 0.78 times that of
Ni doped TiO2 2.6 5.6 17.6 0.31 KDP. However, the SHG efficiency of Ni doped TiO2 is 0.31 times
less than that of KDP respectively. The SHG/NLO measurement re-
sults of bare and doped materials are presented in Table 3.
well separated concentrate rings indicating its higher crystalline
nature (Fig. 7c).
Further, on Ni (8 wt.%) doping well separated individual parti- Conclusion
cles are noted in the TEM images (Fig. 8(a and b)). However, the
sizes of the particles are in the range of 8–16 nm. The SAED pattern The bare and Ni (4%, 8%, 12% and 16%) doped TiO2 nanoparticles
of Ni doped TiO2 exhibiting well isolated concentric rings, reveal- were successfully synthesized by sol–gel method at room temper-
ing the improved crystalline nature (Fig. 8(c)). ature and then annealed at 500 °C and 800 °C for getting anatase
and rutile phases, respectively. The XRD analyses reveal that the
prepared products were attributed to the tetragonal system and
Nonlinear optical studies anatase phase was found to be present in all the samples. When
compared to bare TiO2, the presences of hydroxyl ions were con-
The oven dried powders of pure and 8% Ni doped TiO2 were kept firmed by FTIR analysis in doped TiO2. However, no alkoxyl groups
into muffle furnace and annealed at 800 °C for 4 h to obtain rutile are present in the samples. The optical cut-off wavelengths of Ni
phase. In order to confirm the NLO property, the specimens was doped TiO2 were blue shifted towards the lower wavelength side,
subjected to a Kurtz powder test [27] using a Q-switched, mode resulting larger band gap energies compared to pure TiO2. PL Spec-
locked Nd: YAG laser of 1064 nm and a pulse width of 8 ns. The in- tra observed a broad and sharp range of emission spectra. The
put laser beam was directed on the synthesized powders to get peaks which were existing in bare TiO2 were blue shifted as a re-
maximum powder SHG then the emitted light passed through an sult of doping. The morphological analysis of 8% Ni–TiO2 nanopar-
IR filter was measured by means of a photomultiplier tube and ticles indicates that the uniform distribution of spherical shaped
oscilloscope assembly. The SHG efficiency of the materials was particles compared to pure TiO2. As can be seen from the TEM im-
evaluated by taking the crystalline powder of KDP as the reference age of 8 wt.% of Ni doped TiO2, the particles are having sizes in the
material. The SHG behavior was confirmed by the output of intense range of 8–16 nm, which are in good agreement with the crystallite
224 B. Rajamannan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 128 (2014) 218–224

size obtained from XRD results. When compared with doped prod- [14] X. Li, J. Yao, F. Liu, H. He, M. Zhou, N. Mao, P. Xiao, Y. Zhang, Bio. Sens. Actuator.
B. 181 (2013) 501–508.
uct bare TiO2 shows more than two fold NLO efficiency.
[15] N. Riaz, F.K. Chong, B.K. Dutta, Z. Man, M.S. Khan, E. Nurlaela, Chem. Eng. J.
185-186 (2012) 108–119.
References [16] X. Yan, J. He, D.G. Evans, Y. Zhu, X. Duan, J. Por. Mater. 11 (2004) 131–139.
[17] T.X. Liu, F.B. Li, X.Z. Li, J. Hazard. Mater. 152 (2008) 347–355.
[1] A. Mills, S.L. Hunte, J. Photochem. Photobiol. A 108 (1997) 1–35. [18] K. Karthik, S.K. Pandian, N.V. Jaya, Appl. Surf. Sci. 256 (2010).
[2] J.M. Herrmann, Catal. Today 53 (1999) 115–129. [19] J.A. Wang, R.L. Ballesteros, T. Lopez, A. Moreno, R. Gomez, O. Novaro, X.
[3] M.A. Ahmed, J. Photochem. Photobiol. A 238 (2012) 63–70. Bokhimi, J. Phys. Chem. B 105 (2001) 9692.
[4] J. Ding, K.S. Kim, J. Chem. Eng. 29 (2012) 54–58. [20] H.J. Zhai, L.S. Wang, J. Am. Chem. Soc. 129 (2007) 3022–3026.
[5] L. Znaidi, R. Seraphimov, J.F. Bocquet, Mater. Res. Bull. 36 (2001) 1–6. [21] L. Ravichandran, K. Selvam, B. Krishnakumar, M. Swaminathan, J. Hazard.
[6] X. Chen, S.S. Mao, Chem. Rev. 107 (2007) 2891–2959. Mater. 167 (2009) 763–769.
[7] L.A. Patil, D.N. Suryawanshi, I.G. Pathan, D.M. Patil, Sens. Actuator, B 176 (2013) [22] S.K. Parayil, H.S. Kibombo, C.M. Wu, R. Peng, J. Baltrusaitis, R.T. Koodali, Int. J.
514–521. Hydrogen Energy 37 (2012) 8257–8267.
[8] M.A. More, J.L. Gunjkar, C.D. Lokhande, R.S. Mane, Micron 38 (2007) 500–504. [23] I. Shown, M. Ujihara, T. Imae, J. Colloid Interface Sci. 352 (2010) 232–237.
[9] D. Zhang, R.J. Phys. Chem. A 86 (2012) 498–503. [24] R. Nirmala, H.Y. Kim, C. Yi, R. Navamathavan, M.E. Newehy, Int. J. Hydrogen.
[10] W. Balcerski, S.Y. Ryu, M.R. Hoffmann, J. Phys. Chem. C 111 (2007) 15357– Energy. 37 (2012) 10036–10045.
15362. [25] T. Sun, J. Fan, E. Liu, L. Liu, Y. Wang, H. Dai, Y. Yang, W. Hou, X. Hu, Z. Jiang,
[11] J. Zhang, K. Xu, Appl. Surf. Sci. 221 (2004) 1–3. Powder Technol. 228 (2012) 210–218.
[12] G.S. Vicente, A. Morales, M.T. Gutierrez, Thin Solid Films 391 (2001) 33–137. [26] H.H. Tseng, M. C Wei, S.F. Hsiung, C.W. Chiou, Chem. Eng. J. 150 (2009) 160–
[13] Y. Liu, Z. Wang, W. Fan, Z. Geng, Libang Feng, Ceram. Int. (2013), http:// 167.
dx.doi.org/10.1016/j.ceramint.2013.08.030. [27] S.K. Kurtz, T.T. Perry, J. Appl. Phys. 39 (1968) 3798–3814.

You might also like