Vanadium Doping Induced Structural and Optical Modifications in TiO2 Thin Films Arshad S. Bhatti1, Awais Ali1, I Ruzybayev2, Emre Yassitepe2 and S. I. Shah2,3 1 Centre for Micro and Nano Devices, Department of Physics, Park Road, COMSATS Institute of Information Technology, Islamabad 44000, Pakistan. 2 Department of Physics and Astronomy, University of Delaware, Newark DE, 19716, United States. 3 Department of Material Science and Engineering, University of Delaware, Newark DE, 19716, United States. *Corresponding author: [email protected] In this work, we present the customized structure and morphology of vanadium (V) doped TiO2 thin films synthesized on glass substrates a growth rate of ~ 0.6Å/s at 500oC. The sputtering targets of pure and V doped TiO2 with three concentrations of V (1.0, 1.5 and 2.0 atomic percentage (at.%)) were prepared from powders. XRD patterns confirmed the grown TiO2 films had the anatase phase. In the doped TiO2 films, the crystallite size reduced by almost half when V concentration increased from 0 to 2 at.% systematically. Incorporation of V in the TiO2 crystal led to the enhanced growth of (211) planes, which significantly modified the grain geometry from the faceted to the elongated as observed in the SEM images. This was simulated by the use of VESTA code. Raman spectroscopy showed phonon confinement, increased nonstoichiometry and asymmetry in bonding with increased V concentration. The XPS spectra confirmed an increase in the nonstoichiometry in TiO2 due to V substitution in the structure. It is suggested that the difference in the valance states of Ti and V produced a clear difference in the ionic radii of the two, which resulted in the augmentation of non-equilibrium (211) planes and hence resulted in the emergence of modified morphology of the synthesized TiO2 thin films. Photoluminescence spectroscopy also revealed the effect of vanadium states on the band gap states. Pure TiO2 did not show any luminescence in the visible, how ever, incorporation of vanadium with different percentages clearly demonstrated its incorporation in V3+, V4+ and V5+ states, which were analysed quantitatly and related to structural modifications. The modified films would be useful for photocatalysis not only structurally but for their improved optical properties. Key Words: TiO2, Vanadium, doping, nonstoichiometry, oxygen vacancies 1 1. Introduction: In recent years, titanium dioxide (TiO2), a wide band gap metal oxide semiconductor, has become one of the intensely studied material due to its use in hydrogen production from water [1], environmental cleaning, e.g., air and water [2–4], self cleansing and non-spotting glass coatings [5], self-sterilizing coatings [2], dye synthesized solar cells [6], etc. TiO2 naturally exists in three polymorphs: anatase, rutile and brookite. Due to its wide band gap (anatse: 3.2 eV and rutile: 3.0 eV), TiO2 absorbs light in the ultraviolet (UV) region of the spectrum which is only 3-5 % of solar spectrum. Thus, the efficient utilization of the solar spectrum (mainly a good portion of the visible light (43-47 %)) is one of the important subjects for developing the future generation of TiO2 based photocatalysts. This essentially requires band gap tailoring of the TiO2 which is achieved by modifying its electronic band structure. Doping of TiO2 is one of the most promising strategies for sensitizing TiO2 to visible light by forming impurity levels within forbidden gap [7]. The doping of TiO2 with 3d transition metals (V, Cr, Mn, Fe, Co, and Ni) is particularly considered as one of the best approaches to narrow the band gap or to define energy levels within the band gap, which significantly enhances the absorption of the visible part of the spectrum [8–12]. Vanadium (V) is a transition metal with multiple characteristics, which has shown to improve absorption of light by TiO2 on doping. For example, V – doped TiO2 with different valance states of V (V0,1+,2+,3+,4+,5+) and Ti (Ti1+,2+,3+,4+) exhibited a difference in oxidation activity [13]. This difference in the oxidation state of V can change the structure and morphology of TiO2. Doping with small concentration of transition metals (around 1-2 at.%) in TiO2 has been demonstrated to successfully reduce the recombination processes by introducing traps for electrons and/or holes [10]. Previous reports on anatase agree on the origin of the visible photoluminescence (PL) in TiO2, attributed to the radiative recombination of self-trapped excitons or surface radiative recombination [14-18]. But still a lot needs to be understood. In this paper, we report the effect of the varied concentration of the doped V and the growth conditions on the structure, morphology and optical properties of TiO2 thin films. The pure and doped TiO2 films were RF sputter deposited on ITO coated glass substrates at a very slow growth rate and at high substrate temperature. It was demonstrated that incorporation of V resulted in the reduction of crystallite size and enhanced growth of a certain plane at the 2 employed growth conditions. Furthermore, the films showed phonon confinement of certain modes along with red shift due to anti-symmetry of the bonding. The photoluminescence spectroscopy also confirmed creation of “V” defect states in the band gap, which showed strong relation with the content of V and the oxygen vacancies in TiO2. The coexistence of multiple valance states of V with reduced Ti confirmed the substitution of V in TiO2, which altered the stoichiometry, structure, morphology and optical properties of the grown films. 2. Experiment: TiO2 thin films were deposited by RF magnetron sputtering using 5 cm diameter custom prepared targets prepared by mixing TiO2 powder (5N) and V2O3 powder (2N) with 0.0, 1.0, 1.5 and 2.0 at.% V in TiO2 by means of standard solid state reaction technique. ITO coated ( ~100nm thick, 4-8 Ω-cm conductivity) microscope glass substrates were cleaned by sonication in various solutions, such as detergent, isopropyl alcohol and acetone in a proper sequence and then thoroughly rinsed in DI water. The ITO coated glass slides were then dried in dry nitrogen before placing in the chamber. After achieving the base pressure of ~ 2 × 10-6 torr, the chamber was back filled with Ar and brought to the working pressure of 10 mtorr. All depositions were done at this pressure and the substrate temperature was kept at 500o C. The films were deposited at a very slow deposition rate of ~ 0.6 Å/s for four hours at 150 Watt. Structure and phase analyses were performed by Rigaku D-Max B X-ray Diffractometer using Cu – Kα radiation (λ = 0.154 nm). The 2θ scans were made in the range from 20° to 80°. Room temperature Raman spectroscopy was carried out by Bruker’s Micro-Raman spectrometer and spectra were collected with an excitation energy of 532 nm from Nd:YAG laser at an incident power of 20mW. The surface and cross-sectional microscopy was performed by JEOL JSM-7400F field emission scanning electron microscopy (FESEM) and the EDX spectra was taken by Oxford Instruments’ PentaFET-6900 energy dispersive X-ray (EDX) spectrometry system installed in the FESEM. XPS was performed by using an incident beam of non-chromatic Al X-ray (1486.5 eV) operating at 10 kV, 10 mA for all scans (survey and high resolution). Measured peaks were then charge corrected to C-1s peak position at 284.6 eV. Photoluminescence (PL) spectroscopy in the band gap of TiO2 was studied at room temperature by Lab Ram ІІІ from DongWoo Optron with Ar+ laser emitting at 488 nm. 3. Result and Discussion: 3 The scanning electron microscopy of the synthesized films revealed interesting surface morphology of the pure and 2 at.% V doped TiO2 thin films as shown in Figures 1(a) and 1(b), respectively. The thickness of the deposited thin film was 850 ± 5 nm as seen in Figure 1 (c), which shows the cross-sectional image of the sample showing all three layers, glass substrate, ITO, and TiO2. The presence of V was confirmed using EDX as shown in Figure 1 (d) of the 2 at.% V doped TiO2 film. It was also confirmed that the prepared thin films had the one to one correspondence of the dopant composition with the prepared targets. Two clear observations were drawn from the micrographs, first, the transformation of morphology, i.e., from the faceted in the pure TiO2 film to the elongated grains in the TiO2:V (2 % at.) doped film, and second, the decrease in the grain size in the doped TiO2 film. The average grain size dropped to almost one third of the average grain size of the pure TiO2 film, i.e., from 160 ± 10 nm to 45 ± 5 nm and the density increased by almost two times. With the increase in the concentration from 0 to 2 at.% of vanadium. The modification in the morphology was ascribed to the dopant incorporation in the host lattice [19] and the decrease in the grain size was due to the replacement of Ti by V in the host lattice. As a consequence of doping, the induced stress due to difference in the two ionic radii hindered the grain growth. The cross-sectional image of the synthesized films also revealed that films had uniform columnar structure which rendered crystal growth under saturated regime [20]. The increased surface area due to morphological modification is always desirable for better efficiency in the photocatalysis. XRD patterns as shown in Figure 2 (a) confirmed the growth of TiO2 anatase phase. The XRD patterns of the grown ITO (bottom), pure TiO2 (middle) and TiO2:V (2 at %) doped film (top) are shown. The unidentified peaks in the TiO2 films emerged from the ITO substrate. . The diffraction peaks were due to anatase (101), (004) and (211) according to the JCPDS card # 020387 as labeled in the Figure 2 (a). This also confirmed the good quality of the synthesized films with no preferred oriented growth in the pure TiO2 films. On the other hand, when doped with V, TiO2 films showed preferential growth along (211) plane as the at.% of V was increased. This is shown in the top XRD pattern and is marked by a dotted circle to show the difference in the intensities of (211) diffraction plane in the two films. A downshift in the diffraction angle was also observed with the increase in V doping and this was considered a signature of dopant incorporation in the host lattice, which created the stress. The shift to lower 2θ value in (211) peak was ~ 1% from pure to 2 at.% V concentration. 4 Figure 1: SEM micrographs of (a) the pure TiO2, (b) 2% at. V doped TiO2 thin films. (c) Crosssectional ciew of the film showing the thickness of the grown film and various layers. (d) EDX spectrum of the 2% at. V doped TiO2 film. The average crystallite size of synthesized films decreased by an almost 1.8 times from the pure to 2 at.% V doped TiO2 films as determined by Debye Scherrer’s formula. This variation was consistent with the SEM findings. This explained the inhibited growth of grain and crystallite sizes was the result of V doping. This was ascribed to the difference in the oxidation state of Ti and V, which had different ionic radii and produced stress in the host lattice, which caused the growth of non equilibrium (211) planes. It was also assumed that during the grain growth, there was a competition between growth of equilibrium (101) plane and nonequilibrium (211) plane and the increase of V concentration led to the pronounced growth of non-equilibrium plane (211). The change in the growth orientation caused the change of morphology as observed in the SEM micrographs and XRD results. This was further confirmed by using structural visualization (VESTA software). The substitution of Ti by V caused the change in the bond lengths as well, which was studied and confirmed by Raman spectroscopy. 5 The substitution of V also caused introduction of structural defects and dopant related defects, which was further studied by photoluminescence spectroscopy. Figure 2: (a) X-ray diffraction patterns of the ITO film (bottom), pure TiO2 film (centre) and TiO2:V (2 % at.) film (top). ITO diffraction pattern was included to distinguish TiO2 films from the substrate film diffraction peaks. (b) Room temperature Raman spectra as obtained from ITO film (bottom), pure TiO2 film (center) and TiO2:V (2 % at.) film (top). No Raman modes from the ITO substrate are observed. Figure 2 (b) represents the room temperature Raman spectra of the ITO film (bottom), pure TiO2 film (middle) and TiO2:V 2.0 at.% V (top). ITO showed no Raman mode in the region of interest, however, the commonly known six Raman modes of the TiO2 anatase phase were observed, i.e., three Eg modes at 144, 197, and 639 cm-1, one B1g mode at 399 cm-1, and one A1g or B1g mode at 519 cm-1 [21] and labeled in Figure 2 (b). The Raman spectra showed a remarkable drop of 67% in intensity of the Eg mode (@144 cm-1) with the increased V 6 concentration from 0 to 2 at.%. In addition, FWHM of the mode increased from 7.8 to 9.8 cm-1 as V concentration increased in the same range. This was ascribed to the relaxation of Raman selection rules near the centre of the Brillouin zone and a range of q (i.e., q+∆q) wavevectors became accessible, where ∆q~1/L, and L is the crystallite size [22]. So, the systematic drop in intensity and increase in FWHM of Eg mode (@ 144 cm-1) with the addition of V in the TiO2 lattice was a clear indication of the change in the bond lengths and nonstoichiometry in the grown films. Another observation was the red shift of Eg mode (@144 cm-1) by 3.32 cm-1. Which was a typical indication of the phonon confinement effect associated to the reduction of the crystallite size on V doping. It is already established fact that the Eg mode is produced due to the symmetric stretching vibration of O-Ti-O bonds in TiO2 and the B1g mode is produced due symmetric bending vibration of O-Ti-O results in the B1g mode, whereas the anti-symmetric bending vibration of OTi-O produces the A1g mode [23]. It was observed that the ratio of Eg to A1g mode also decreased as the V doping concentration increased. It further confirmed that the asymmetry in the host lattice was produced by both, i.e., the dopant substitution and creation of oxygen vacancies as was observed in the XPS measurements. Visualization of electrical and structure analysis (VESTA) software [24] was employed to determine the crystal structure of the pure and doped TiO2. VESTA employes the periodic bond chain (PBC) theory to visualize materials in 3D. VESTA confirmed the faceted morphology of the anatase phase with the incorporation of (101) plane and (105) plane. However, elongated morphology of the anatase phase was observed with the inclusion of (211) plane in addition to (101) and (105) planes for the 2 at.% V doped TiO2 films. It must be noted that the d-spacing of (211) plane (1.66 Å) is almost half the d-spacing of the (101) plane (3.51 Å). Thus, for the pure TiO2, growth of stable planes was preferred and the growth rate along (101) was more than twice the growth rate along (211). However, the doping of “V” led to the enhanced growth rate of unstable (211) plane compared to (101) plane and this resulted into small elongated grains with reduced grain size as observed in the SEM micrographs. Figure 3 shows a high resolution scan of V-2p3/2 region, which confirmed the substitution of V in multiple valance states in TiO2:V (1 % at. and 2 % at.) doped films. The fit confirmed the existence of valance states of V and O-1s satellite peak between V-2p1/2 and V-2p3/2. This was 7 a typical signature of V in the oxide state (V1+/2+/3+/4+/5+) [25]. This confirmed the incorporation of V in TiO2 lattice substituting Ti and bonding with O. With the increase in V doping concentration the area of V3+ at 515.3 eV decreased and V5+ at 516.8 eV increased. The details of integrated area of each valance states were determined. For 1.0 at% V doping, only 20 % of V was in 5+ states but this number increased to almost 54 % for 2.0 at.% V concentration. it is interesting to note that V5+ in comparison with Ti4+ has 8.7 % smaller ionic radius, which was good enough to produce stress in the TiO2 lattice responsible for reduction in crystallite/grain size, change in morphology and enhanced growth along (211) plane. The coexistence of V3+ and V5+ and the presence of satellite peaks of O-1s also confirmed the nonstoichiometry in TiO2. Figure 3: High resolution XPS spectra of TiO2:V (1 % at. (bottom) and 2 % at. (top)) to demonstrate the contribution of various ionized V substitution in TiO2 host lattice. The spectra have been fitted with Gaussian functions to determine the contribution of V+3, V+4, and V+5 states. TiO2 in both phases, i.e., anatase and rutile have a band gap in the ultraviolet (UV) region with no visible emission expected from it. However, a few papers have recently highlighted the influence of defects, particularly structural, on the luminescence characteristics in the visible and IR from the pure and modified TiO2 [14-18]. It was anticipated that three types of defects would emanate in the TiO2 films; due to oxygen vacancies, structural defects (e.g., Ti3+ or grain boundaries) and defects introduced by the dopant “V”. The PL spectra of the pure TiO2 and TiO2:V (2 % at.) doped films obtained at the room temperature in the visible region (band gap luminescence) exhibited very interesting features as shown in Figure 4 (a, b). Figure 4 (a) shows 8 two sharp bands centered one at 530 nm due to oxygen vacancies (SV), other at 600 nm due to surface oxygen vacancies (SOV), and a broad band at 750 nm due to structural defects (SD). The structural defects were produced by Ti. However, substitution of the dopant “V” in the host lattice resulted in the appearance of another strong band in the visible and in the infra-red (IR) region overlapped with the SD band. The PL bands were resolved mainly with Gaussian functions, which enabled determination of the contribution of each component. XPS spectra has already confirmed that the stoichiometry of TiO2 was quite sensitive to the amount of V. In the following paragraph, the origin and variation of observed bands in the PL spectra are discussed. Figure 4: Room temperature band gap photoluminescence spectra of (a) the pure TiO2 film, and (b) TiO2:V (2 % at.) doped film. The contribution from various responsible defects is marked in (b) as discussed in the text. Structural defect and Vanadium defects contribution is pronounced in the doped film. The PL peak observed at around 530 nm was attributed to the recombination of free electrons with the shallow trapped holes usually originated from oxygen vacancies (OV). The orange – red emission broader peak between 590 and 650 nm was from the recombination of trapped electrons with the valence band free holes. Such a PL was attributed to the presence of uncoordinated Ti3+ only when electrons occupied it [14]. In pure TiO2 film, the major contribution to the PL spectra was due to oxygen vacancies related transitions; however a minute contribution from Ti3+ was also observed. The PL spectra collected from all doped samples also 9 exhibited the presence of same bands but with appearance of strong side bands of the band due to structural defects (SD). The contribution of new bands in the modified SD band was determined as a function of “V” concentration and summarized in Table I. Interestingly, energy position of new bands remained insensitive to the V concentration, however their intensities varied drastically.. Interestingly, the intensity of the band at 750 nm increased while intensity dropped for the band at 800 nm when “V” concentration was increased. Thus, new PL bands emerged at 750 nm and 800 nm were due to the presence of V5+ and V3+ states as previously confirmed by the XPS. The intensity behavior of all PL bands is summarized in Table I. It was observed that the PL due to oxygen vacancies and “V” defects increased drastically, although it was greater for “V” defects. The variation in the PL intensity due to OV and VD was quite consistent with the XPS observation. Table I: Table of contribution of various defects in the band gap luminescence of the pure TiO2 and V doped TiO2 with varying V concentration. Integrated PL Intensity V at. % in TiO2 Oxygen Vacancies Vanadium Defects (OV) Surface Oxygen Vacancies (SOV) Titanium Structural Defects (TSD) 0.0 15.25 20.18 09.81 00.00 1.0 10.16 00.63 07.57 37.56 1.5 14.90 01.20 00.47 41.53 2.0 16.82 00.52 09.05 57.20 (VD) 4. Conclusions In summary, V doped TiO2 thin films doped with varied V concentration from 0.0 to 2.0 % at. were synthesized on ITO coated glass slides by RF sputtering. The incorporation of V systematically modified the structure, morphology and optical properties of the grown TiO2 films. The growth orientation changed from (101) to (211) in pure to 2 % at. doped TiO2 films as confirmed by XRD patterns. The substitution of V in place of Ti created stress and modified the crystallite and grain sizes as the ionic radii of V was considerably different than Ti. In addition, the oxidation state of V was also found to be different. Raman spectroscopy confirmed the 10 increase in phonon confinement, and Ti – O bond asymmetry with the increase in V concentration. XPS showed also confirmed the increase in the nonstoichiometry. This was attributed to the substitution of V in the V3+ and V5+ states. The band gap PL spectra showed strong contribution from the states associated with V5+ and V3+ and dominated the spectra on the OV and defect associated PL. It was therefore concluded that the substitution of V in an oxidation state different than Ti oxidation state was the main cause of the change in the morphology, non-stoichiometry and appearance of additional emission bands in the photoluminescence of TiO2. The research work was funded by Higher Education Commission (HEC) of Pakistan though the NRPU grant 1770. One of the author AA is also thankful to HEC for its IRSIP program and indigenous PhD scholarship. 5. References: [1] A. Fujishima, K. Honda, Nature 238 (1972) 37. [2] A. Fujishima, X. Zhang, D. Tryk, Surface Science Reports 63 (2008) 515. [3] O. Carp, C. L. Huisman, A. Reller, Progress in Solid State Chemistry 32 (2004) 33. [4] A. L. Linsebigler, G. Lu, J. T. Yates, Chemical Reviews 95 (1995) 735. [5] A. Mills, J. Wang, M. Crow, Chemosphere 64 (2006) 1032. [6] B. O’Regan, M. Grätzel, Nature 353 (1991) 737. [7] R. Dholam, N. Patel, M. Adami, a. Miotello, International Journal of Hydrogen Energy 34 (2009) 5337. [8] M. Ni, M.K.H. Leung, D.Y.C. Leung, K. Sumathy, Renewable and Sustainable Energy Reviews 11 (2007) 401. [9] W. Choi, A. Termin, M.R. Hoffmann, The Journal of Physical Chemistry 98 (1994) 13669. [10] M. I. Litter, J. A. Navfo, Journal of Photochemistry and Photobiology A: Chemistry 98 ( 1996) 171. [11] M. Khan, S. Woo, O. Yang, International Journal of Hydrogen Energy 33 (2008) 5345. [12] J. Zhu, Z. Deng, F. Chen, J. Zhang, H. Chen, M. Anpo, J. Huang, L. Zhang, Applied Catalysis B: Environmental 62 (2006) 329. 11 [13] T. D. Nguyen-Phan, M. B. Song, H. Yun, E. J. Kim, E. S. Oh, E. W. Shin, Applied Surface Science 257 (2011) 2024. [14] Candy C. Mercado, Fritz J. Knorr, Jeanne L. McHale, Shirin M. Usmani, Andrew S. Ichimura and Laxmikant V. Saraf, Journal of Physical Chemistry C 116 (2012) 10796. [15] Riley E. Rex, Fritz J. Knorr, and Jeanne L. McHale, Journal of Physical Chemistry C. 117 (2013) 7949. [16] Ana Stevanovic, Michael Buttner, Zhen Zhang and John T. Yates Jr, Journal of American Chemical Society 134 (2012) 324. [17] Xiuli Wang, Zhaochi Feng, Jianying Shi, Guoqing Jia, Shuai Shen, Jun Zhouab and Can Li, Physical Chemistry Chemical Physics 12 (2010) 7083. [18] Lucia Cavigli, Franco Bogani, Anna Vinattieri, Valentina Faso and Giovanni Baldi, Journal of Applied Physics 106 (2009) 053516. [19] A. Ali, E. Yassitepe, I. Ruzybayev, S. Ismat Shah, A. S. Bhatti, Journal of Applied Physics 112 (2012) 113505. [20] A. Borras, J.R. Sanchez-valencia, R. Widmer, V.J. Rico, A. Justo, A.R. Gonzalez- elipe, Crystal Growth & Design 9 (2009) 2868. [21] F. Rossella, P. Galinetto, M.C. Mozzati, L. Malavasi, Y. Diaz Fernandez, G. Drera, L. Sangaletti, Journal of Raman Spectroscopy 41 (2009) 558. [22] I. H. Campbell and P.M. Fauchet, Solid State Communications 58 (1986) 739. [23] F. Tian, Y. Zhang, J. Zhang, C. Pan, Journal of Physical Chemistry C 16 (2012) 7515 [24] K. Momma, F. Izumi, Journal of Applied Crystallography 44 (2011) 1272. [25] G. Silversmit, D. Depla, H. Poelman, G. B. Marin, R. De Gryse, Journal of Electron Spectroscopy and Related Phenomena 135 (2004) 167. [26] Y. J. Kao and N. M. Haegel, Physical Review B 48 (1993) 4433. [27] H. Y. Chen, O. Zahraa, M. Bouchy, F. Thomas, J. Y. Bottero, Journal of Photochemistry and Photobiology A: Chemistry 85 (1995) 179. [28] X. Liang, S. Zhu, Y. Zhong, J. Zhu, P. Yuan, H. He, J. Zhang, Applied Catalysis B: Environmental 97 (2010) 151. [29] J.A. Cusumano, M. J. D. Low, The Journal of Physical Chemistry 74 (1970) 792. 12
© Copyright 2025 Paperzz