Sutapa
Dey
a,
Anusmita
Chakravorty
b,
Shashi Bhusan
Mishra
c,
Nasima
Khatun
a,
Arnab
Hazra
d,
Birabar Ranjit Kumar
Nanda
c,
Chandran
Sudakar
e,
Debdulal
Kabiraj
b and
Somnath C.
Roy
*a
aSemiconducting Oxide Materials, Nanostructures and Tailored Heterojunction (SOMNaTH) Lab, Functional Oxide Research Group (FORG), Department of Physics, Indian Institute of Technology Madras, Chennai 600036, India. E-mail: somnath@iitm.ac.in
bInter-University Accelerator Centre, Aruna Asaf Ali Marg, New Delhi 110067, India
cCondensed Matter Theory and Computational Lab, Department of Physics and Center for Atomistic Modelling and Materials Design (CAMMD), Indian Institute of Technology Madras, Chennai 600036, India
dDepartment of Electrical & Electronics Engineering, Birla Institute of Technology & Science-Pilani, Pilani Campus, Pilani-333031, Rajasthan, India
eMultifunctional Materials Laboratory, Department of Physics, Indian Institute of Technology Madras, Chennai 600036, India
First published on 15th November 2021
Irradiation of materials by high energy (∼MeV) ions causes intense electronic excitations through inelastic transfer of energy that significantly modifies physicochemical properties. We report the effect of 100 MeV Ag ion irradiation and resultant localized (∼few nm) thermal spike on vertically oriented TiO2 nanorods (∼100 nm width) towards tailoring their structural and electronic properties. Rapid quenching of the thermal spike induced molten state within ∼0.5 picosecond results in a distortion in the crystalline structure that increases with increasing fluences (ions per cm2). Microstructural investigations reveal ion track formation along with a corrugated surface of the nanorods. The thermal spike simulation validates the experimental observation of the ion track dimension (∼10 nm diameter) and melting of the nanorods. The optical absorption study shows direct bandgap values of 3.11 eV (pristine) and 3.23 eV (5 × 1012 ions per cm2) and an indirect bandgap value of 3.10 eV for the highest fluence (5 × 1013 ions per cm2). First principles electronic structure calculations corroborate the direct-to-indirect transition that is attributed to the structural distortion at the highest fluence. This work presents a unique technique to selectively tune the properties of nanorods for versatile applications.
Titanium dioxide (TiO2) remains one of the most prominent members of the metal oxide family owing to its high chemical and thermal stability, excellent photoactivity, biocompatibility and availability in the form of three different polymorphs (anatase, brookite and rutile).8,9 There are a few reports, which have discussed SHI irradiation-induced structural changes in TiO2, for example, from amorphous to anatase,10 rutile,11,12 and mixed-phase;12 and anatase to amorphous,13 rutile,14 and mixed-phase15 transformations. SHI irradiation induced effects have also been studied on doped TiO2. For example, Gautam et al.16 reported 120 MeV Ag and 130 MeV Ni ion induced phase transformation in undoped and niobium doped anatase TiO2 composite thin films. Thakurdesai et al. observed a SHI-induced phase transition of a pulsed laser deposited TiO2 thin film from the rutile to the anatase phase.17 SHI-induced nano or microlithography18–21 has been reported on single crystal rutile TiO2. Further, formation of the ion track under SHI irradiation has been investigated in single crystal rutile TiO2 with (001),22,23 (110),24–26 and (100)27 orientations. Latent tracks of different morphologies such as cylindrical, dumbbell-shaped or sandglass-like have been observed with variation of ion path length as a consequence of molten phase outflow and recrystallization.28 In the case of thin films, 120 MeV Ag ion irradiation has been performed on a rutile TiO2 film deposited by ion beam sputtering on silicon to investigate the influence of SHI irradiation on the dynamics of phonon interactions.29 Rath et al. reported 79 MeV Br ion induced modifications in a 200 nm thick rutile TiO2 thin film prepared by DC magnetron sputtering on a Si (100) substrate.30 The sample crystallinity was retained up to a fluence of 1 × 1013 ions per cm2, even though the Se value exceeded the threshold for amorphization along the ion track in TiO2.30 Some of the important published studies on Ag ion irradiation effects on TiO2 are listed in Table T1 (ESI†). It is therefore apparent that several studies have been reported on SHI irradiation of TiO2 single-crystal and thin films; however, such studies on TiO2 nanostructures remain unexplored.
One-dimensional nanostructures such as nanowires and nanorods, offer the advantage of unidirectional charge transport along with high specific surface area.31 Particularly, highly aligned and vertically oriented TiO2 nanorods hydrothermally grown on a fluorine doped tin oxide (FTO) coated glass substrate have shown strong potential in applications such as electrode materials with a transparent conducting window for solar cells,32 water splitting,33 electrocatalysis,34 bio-photoelectrochemical reactions,35 UV photodetectors,36 electron transport layers in perovskite solar cells,37 electron injection layers in DSSCs,38 sensors,39etc. In addition, the single-crystalline nature of each nanorod provides a long-range periodicity throughout the individual nanorod structure, which is a suitable platform for the study of irradiation effects. Moreover, a small lateral dimension (a few tens of nanometers) of the nanorods provides limited space for thermal energy transfer and atomic displacement in the radial direction. Therefore, it is interesting to explore the effect of SHI irradiation-induced localized energy transfer in rutile TiO2 nanorods. Here, we present the crystalline phase, morphology and electronic structure transformation of the SHI (100 MeV Ag ion) irradiated TiO2 nanorods, and their effect on the optical absorption. The experimental data and observation are correlated with the thermal spike model-based analysis and density functional theory (DFT) calculated band structure to gain an insight into ion beam-induced disorder and defects in these nanorods.
At a low fluence of 1 × 1012 ions per cm2, no significant changes are observed (Fig. 1(b)). However, the nanorod structures show slight bending at a few locations (Fig. 1(c)) after irradiation at a fluence of 5 × 1012 ions per cm2 (indicated by yellow ellipses), followed by an increase in bending and morphological distortion at a fluence of 1 × 1013 ions per cm2 (Fig. 1(d)). Further increase of fluence to 5 × 1013 ions per cm2 causes a significant structural distortion and bending of individual nanorods with the appearance of a semi-transparent layer connecting the adjacent rods (Fig. 1(e)). The interconnected adjacent nanorods are indicated by red ellipses in Fig. 1(e). It appears that the energy transferred from an incident ion to a nanorod induces intense local heating leading to melting and bending of the TiO2 nanorods. Diffusion of the molten material between adjacent nanorods results in the formation of such interconnecting layers. Here, the individual nanorods having lateral dimensions of ∼100 nm provide limited space for heat dissipation and molten state flow. Therefore, as a result of thermal spike induced melting, significant morphological changes are clearly observed from the FESEM images. However, in the case of polycrystalline films even if a thermal spike induced change in crystallinity occurs, it would not be so clearly evident through morphological changes. The digital photographs of actual samples irradiated at different fluences are shown in the ESI (Fig. S1†).
To investigate the structural changes induced by SHI irradiation, XRD analysis was carried out for pristine and irradiated samples, and the results are presented in Fig. 2(a). The XRD pattern of the pristine sample shows peaks at 36.10°, 41.33°, 54.59°, 62.93° and 68.99°, which correspond to (101), (111), (211), (002) and (301) of rutile TiO2 (ICSD reference code 01-073-1765) respectively. The additional peaks are attributed to SnO2 (ICSD reference code 01-077-0452) from the FTO coating of the substrate. After irradiation, there is a gradual reduction in crystallinity as indicated by decreasing intensities for (101), (211), and (002) peaks. To analyze the shift in the position of the peaks, a high-resolution X-ray scan was carried out for the (101) and (211) planes and the data are presented in Fig. 2(b) and (c), respectively. A clear shift to a lower 2θ value is observed, indicating an increase in the inter-planner spacing as a result of ion irradiation.
When high-energy ions pass through a material, an enormous amount of energy is transferred within a short period. As a result, in some cases, the amorphous material turns crystalline,10 whereas the crystalline material becomes amorphous13 due to thermal spike-induced heating, local melting, and ultrafast re-solidification. The excessive local heating along the ion trajectories creates latent ion tracks that are often distinct in density or crystallinity compared to the surrounding material. In the present case, the gradual decrease in crystallinity can be attributed to the increase in disorder resulting from the SHI-induced ion tracks that start overlapping with increasing ion fluence. Further, the shift of (101) and (211) peaks to lower diffraction angles after irradiation at a fluence of 1 × 1012 ions per cm2 can be attributed to the ion irradiation-induced stresses in the TiO2 lattice. The passage of energetic Ag ions causes atomic movements and vacancy creation in the lattice resulting in tensile stress leading to an expansion in the inter-planner spacing (d). Such an expansion causes a consequent shift of the XRD peak to a lower 2θ value.
The changes in crystallinity were also studied by Raman spectroscopy, which is highly sensitive to the different crystalline phases of TiO2. Fig. 3 presents the Raman spectra of the pristine and irradiated TiO2 nanorods with different fluences. The spectra for the pristine sample show three first-order Raman peaks at 143.91 cm−1 (B1g), 444.85 cm−1 (Eg), and 608.62 cm−1 (A1g) and a broad peak attributed to second-order scattering at 237.16 cm−1, which correspond to the rutile TiO2.43,44 In all the Raman active modes, the Ti atom in the O–Ti–O bond remains at rest and vibration takes place at the oxygen atoms. The Eg mode is caused by the asymmetric bending of the O–Ti–O bonds in the {001} plane, where O atoms move in the opposite direction across the O–Ti–O bond and vibrate along the c-axis.45 In contrast, the B1g mode occurs as a result of the asymmetric bending of the O–Ti–O bonds in the {001}, {110} and {-110} planes. On the other hand, the A1g mode arises due to the symmetric stretching of the O–Ti–O bonds in the {110} plane caused by the opposite directional movement of O atoms in the O–Ti–O bonds. In B1g and A1g modes, the O atoms move perpendicular to the c-axis with respect to the stationary Ti atom. The vibrations of oxygen atoms corresponding to each Raman mode are schematically45 shown in Fig. S2 in the ESI.† It is observed that the peak intensities decrease gradually with the increase in ion fluences. This indicates a gradual loss of crystallinity in the TiO2 lattice. In addition, there is a slight shift for B1g at a fluence of 1 × 1012 ions per cm2. As the ion fluence increases, the peak corresponding to A1g mode shifts towards a higher frequency, while those of Eg and B1g modes shift to lower frequencies. This results from the localized disorder and distortion of chemical bonds due to the oxygen vacancies in the TiO2 octahedra,46 which was also observed in a previous study.30 In particular, compressive (or tensile) stress is known to cause a blue (or redshift) in the Raman peaks.47 In our case, the blue shift of the A1g peak indicates compressive stress perpendicular to the c-axis, which is caused by the shortening of Ti–O bonds. Gautam et al.29 reported that the high density of oxygen vacancies in {110} planes results in tensile strain, which elongates the equatorial Ti–O bond and shortens the apical Ti–O bond. The shortening of apical Ti–O bond makes the bond stronger that leads to the blue shift of the A1g peak.29 On the other hand, the redshifts of B1g and Eg modes suggest the development of tensile stress. The generation of compressive or tensile stresses is attributed to the non-uniform oxygen deficiency in the TiO2 lattice caused by the energetic ion beam.47 Further, it is observed that the relative intensity of Eg reduces compared to that of A1g. The vertically aligned TiO2 nanorods grow along the [001] direction.48 Therefore, ion irradiation on the nanorods leads to a higher impact on the {001} planes compared to the others. Hence, the intensity of the Eg mode decreases rapidly as compared to that of the A1g mode. Finally, at a fluence of 5 × 1013 ions per cm2, the peaks almost disappear. This is correlated with the large density of defects and structural disorder created by the ion beam.
To further investigate the changes in the microstructure and crystallinity of the irradiated nanorods, we performed TEM analysis. The micrographs along with the selected area electron diffraction (SAED) patterns shown in insets are presented in Fig. 4. The TEM image of a single pristine nanorod showing an almost perfect tetragonal shape is presented in Fig. 4(a). The single-crystalline nature of this nanorod is confirmed from a clear spot-type SAED pattern. The microstructure of the nanorods irradiated at a fluence of 5 × 1012 ions per cm2 is presented in Fig. 4(b). In the SAED pattern, we observe diffused spots indicating an interruption in the long-range periodicity of the lattice. A closer look at the microstructure reveals striking features of parallel trajectories (indicated by yellow arrows) attributed to the track formation by SHI. The observed width of the tracks is ∼10 nm, which is corroborated by the radial dimension of the simulated ion track obtained from the thermal spike model presented in the next section. Further, the smooth surface of the nanorod becomes corrugated (indicated by the yellow ellipse) at the sites of ion beam impingement. When the samples are irradiated at a fluence of 5 × 1013 ions per cm2 (Fig. 4(c)), significant microstructural deformation takes place, as indicated by the ring-type SAED pattern. Here, we observe that the nanorods having rounded edges, which may be attributed to partial melting caused by a high amount of thermal energy resulting from the interaction with the ion beam.
Chemical compositions of pristine and irradiated nanorods were investigated by XPS. Fig. 5 presents (a) Ti 2p and (b) O 1s spectra of pristine and irradiated (5 × 1012 and 5 × 1013 ions per cm2) samples. The peaks located at 457.47 eV and 463.06 eV (in Fig. 5(a)) are attributed to Ti 2p3/2 and Ti 2p1/2, respectively, and correspond to the Ti4+ state.49Fig. 5(a) shows no significant change in the peak position and shape after Ag ion irradiation. However, the peak intensities decrease compared to that of pristine nanorods, which indicates a reduction in Ti4+ content in the sample.50 The O 1s peak in Fig. 5(c) shows three components attributed to the lattice oxygen of TiO2 (528.06 eV), oxygen vacancies (529.67 eV), and surface adsorbed oxygen (530.58 eV).51Fig. 5(b) shows that the contribution from oxygen vacancies becomes stronger in irradiated samples. After irradiation at fluences of 5 × 1012 ions per cm2 and 5 × 1013 ions per cm2, the peaks shift to the higher binding energies observed from Fig. 5(d) and (e). The peak positions in terms of binding energies (in eV) of Ti and O are presented in Table 1. The observed shift is attributed to the irradiation-induced changes in the electronic structure, as also observed in the case of ZnO nanorods.52
Chemical state of elements | Pristine | 5 × 1012 ions per cm2 | 5 × 1013 ions per cm2 |
---|---|---|---|
Ti 2p3/2 | 457.47 | 457.34 | 457.34 |
Ti 2p1/2 | 463.06 | 463.06 | 463.06 |
O 1s (lattice oxygen) | 528.06 | 528.40 | 528.54 |
O 1s (oxygen vacancy) | 529.67 | 529.96 | 530.27 |
O 1s (adsorbed oxygen) | 530.58 | 531.05 | 531.33 |
![]() | (1) |
![]() | (2) |
Fig. 6(a) shows that the electronic subsystem retains the energy deposited by the incident ion for a very short period (∼10−15 s) and then it is transferred to the lattice subsystem via electron–phonon coupling controlled by the parameter g. It is observed from Fig. 6(b) that the lattice temperature reaches the melting point of TiO2 (2130 K)27 at 5.5 × 10−15 s and the boiling point (3200 K)27 at 2.3 × 10−14 s. Fig. 6(b) also shows that the temperature within a radius of 5.5 nm exceeds the melting point of 2130 K. This observation is in agreement with the lateral dimension (∼10 nm) of the track as observed from high-resolution TEM images (Fig. 4(b)). According to the thermal spike model, the ion track appears as a consequence of the re-solidification during the rapid quenching of the molten state along the ion trajectory. Fig. 6(b) shows that the quenching from the molten state starts at 5.2 × 10−11 s, enabling us to ascribe the SHI-induced track formation to the rapid quenching of the thermal spike. Further, melting of the material as predicted by the i-TS model explains the observations in the SEM images (Fig. 1), which show a thin layer of material connecting the adjacent nanorods.
![]() | ||
Fig. 6 (a) Electronic and (b) lattice temperature profiles along the radial distance from the center of the ion trajectory with the evolution of time. |
The interaction of the 100 MeV Ag ion beam with TiO2 nanorod arrays is schematically shown in Fig. 7. At a relatively lower fluence (5 × 1012 ions per cm2), tracks are formed inside the material preserving the tetragonal shape of each nanorod. However, at the highest fluence (5 × 1013 ions per cm2), individual ion tracks overlap that result in the overall melting of the nanorods. Being in a molten state, the nanorods start to bend and interconnect with adjacent nanorods through the diffusion of molten layers. Due to the rapid quenching of the molten state, significant distortions of the crystalline structure and the tetragonal shape occur along with the appearance of interconnected nanostructure morphology. Initially, TRIM simulation42 (Fig. S3(c)†) was performed to quantify the disorder created inside the TiO2 lattice. However, the TRIM code42 does not account for the thermal spike phenomenon and therefore provides an under-estimation of the disorder compared to the experimental observations. In the present study, the disorderliness is highly pronounced as a consequence of thermal spike and rapid quenching.
![]() | ||
Fig. 7 Schematic presentation of the effect of high energy ion irradiation on the TiO2 nanorods at different fluences (5 × 1012 ions per cm2 and 5 × 1013 ions per cm2). |
![]() | ||
Fig. 8 (a) UV-Vis diffuse reflectance spectra, (b) absorbance spectra and (c and d) Tauc plots of pristine and samples irradiated at 5 × 1012 and 5 × 1013 ions per cm2 fluences. |
The luminescence characteristics of the pristine and irradiated nanorods were investigated by PL spectroscopy, and the spectra are presented in Fig. 9. The PL spectrum of the pristine sample exhibits a prominent peak at 408 nm attributed to the near-band-gap emission and a peak at 467 nm, a characteristic peak for the rutile TiO2.56 Other small peaks at 450 nm, 482 nm, and 492 nm may arise from low-intensity radiative transitions between sub-band-gap states. The PL peaks do not show any significant shift after ion irradiation, but a decrease in intensities is observed. The peak at 408 nm is significantly reduced in intensity for the irradiated samples due to the incorporation of lattice defects and disorder build-up with increasing ion fluence. However, the peak at 467 nm suggests that some extent of the rutile structure persists even after the irradiation at the highest fluence of 5 × 1013 ions per cm2.
![]() | ||
Fig. 9 PL spectra of pristine and irradiated samples with 1 × 1012, 5 × 1012, 1 × 1013 and 5 × 1013 ions per cm2 fluences. 320 nm wavelength was used for excitation. |
The calculations have been performed using the plane-wave-pseudopotential approach as implemented in Quantum ESPRESSO.57 The experimentally obtained lattice parameters of the tetragonal structure of the rutile TiO2 estimated by Rietveld refinement of XRD data have been considered for the calculations. The Vanderbilt ultrasoft pseudopotentials are used to describe the electron–ion interactions,58 in which the valence states of Ti include 12 electrons from 3s, 3p, 4s and 3d, and that of O includes 6 electrons from 2s and 2p shells. The exchange-correlation functional is approximated through the PBE-GGA functional.59 The convergence criterion for self-consistent energy is taken to be 10−8 Ry. A k-mesh of 8 × 8 × 8 is used for the Brillouin zone integration. The kinetic energy cut-off for the electron wave functions is set at 30 Ry, and the augmented charge density cut-off is set to be 300 Ry. Our previous analysis on TiO2 tells that with these sets of parameters, the lattice parameters for the case of anatase TiO2 match with that of the theoretical and experimental reports.60,61
The rutile TiO2 crystallizes in a tetragonal structure with P42/mnm (space group #136). Each unit cell contains two formula units of TiO2, in which each Ti atom is coordinated with six O atoms and each O atom to three Ti atoms, as shown in Fig. 10(a). The TiO2 octahedra are slightly distorted and connected at the edge through O atoms. In the present work, we have used the experimental lattice parameters obtained from the XRD data. For pristine TiO2, the observed lattice parameters are a = 4.575 Å and c = 2.936 Å. When the structure is subjected to a fluence of 5 × 1012 ions per cm2, the Ti–O bond undergoes slight distortion (Fig. 10(b)), indicating a structural transformation from the crystalline to the quasidisordered state. At a fluence of 5 × 1013 ions per cm2, the structure undergoes a larger distortion which is clear from the breakdown of TiO6 octahedra as presented in Fig. 10(c). Although the lattice parameter is hardly affected, the distortion of the Ti–O bond is quite significant, and its impact on the band structure is discussed in the following section.
With the GGA exchange-correlation functional, a direct bandgap is observed for pristine TiO2 having a value of 1.68 eV with the band minima at the high symmetry point Γ (Fig. 10(d)). It is well understood that the GGA functional underestimates the bandgap.62 Although the GGA + U can approximate the bandgap to the experimental value;63 however, in the present situation, the GGA functional is sufficient to provide a comparison of the change in the bandgap with the structural deformation. At a fluence of 5 × 1012 ions per cm2, a minor structural distortion takes place as a result of which the band dispersive nature remains almost like that of the pristine structure except for an increase in the bandgap by 0.21 eV (Fig. 10(e)). When the structure is subjected to a fluence of 5 × 1013 ions per cm2, the band dispersion changes due to distortion of the Ti–O bond (Fig. 10(f)). It goes from the direct bandgap to the indirect bandgap system, where the valence band maxima are observed at Γ and M points and the conduction band minimum is at the R point. The intertwined Ti-d – O-p – Ti-d covalent interactions play a major role in the band dispersion of the lower-lying conduction bands. The increase in bond length and further tilting of the Ti–O–Ti axis modify this interaction to create the CBM at R instead of Γ. Such a direct to indirect transition of the bandgap has also been observed in ZnO and Si nanowires.64–66
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1na00666e |
This journal is © The Royal Society of Chemistry 2022 |