Substrate temperature dependence of growth mode, microstructure and optical properties of highly oriented zinc oxide films deposited by reactive sputtering Sukhvinder Singh a, Tapas Ganguli b, Ravi Kumar b, R.S. Srinivasa c, S.S. Major a,⁎ a b c Department of Physics, Indian Institute of Technology Bombay, Mumbai-400076, India Semiconductor Laser Section, RRCAT, Indore-452013, India Department of Metallurgical Engineering & Materials Science, Indian Institute of Technology Bombay, Mumbai-400076, India a b s t r a c t Keywords: ZnO Reactive sputtering Microstructure Optical properties X-ray diffraction Transmission electron microscopy Polycrystalline ZnO films were deposited on quartz substrates by reactive sputtering of zinc target. X-ray powder diffraction, pole figure analysis and high resolution measurements along with transmission electron microscopy, Raman and photoluminescence studies were carried out to study the microstructure, crystallinity and optically active defects in the films. All the films deposited in the substrate temperature range from room temperature to 600 °C exhibited strong c-axis preferred orientation. The changes in preferred orientation of crystallites with substrate temperature were attributed to its being determined by preferential nucleation at lower temperatures and surface diffusion at higher temperatures. A detailed microstructural analysis showed that with increase in substrate temperature from 300 °C to 600 °C, a significant reduction in micro-strain to ∼ 10− 3 takes place, along with a marginal increase in crystallite size. Raman and photoluminescence studies have shown that the films deposited below 300 °C possessed poor crystalline quality. The film deposited at 600 °C yielded the most intense and narrow (∼102 meV) band edge luminescence at room temperature, though it did not exhibit the strongest c-axis orientation of crystallites. This is attributed to its superior crystalline quality and absence of oxygen-deficiency related defects. 1. Introduction ZnO is a wide band gap semiconductor with large exciton binding energy (60 meV), making it one of the most potential materials to realize the next generation optoelectronic devices, operating in short wavelength region [1–3]. Several growth techniques, such as, sputtering, pulsed laser deposition (PLD), metal-organic chemical vapour deposition (MOCVD), Sol–gel process and spray pyrolysis have been used for the deposition of polycrystalline ZnO films. In recent years, high quality epitaxial ZnO films have also been grown on crystalline substrates. Most of this work has been reviewed in Refs. [1,2]. As a deposition technique, sputtering offers several advantages due to its simplicity, versatility and scalability, and has been extensively used for the deposition of ZnO films [4]. Epitaxial ZnO films [5–9] as well as high quality polycrystalline ZnO films on noncrystalline substrates [10–12] have been deposited by various forms of sputtering. In a recent work [13] on reactively sputtered ZnO films on quartz substrates, it has been shown that all the films deposited in the substrate temperature range of room temperature to 600 °C exhibited a single low order XRD peak due to (0002) reflection, except for the film deposited ∼ 200 °C, which exhibited multiple peaks with a dominant (0002) reflection. Thus all the ZnO films were c-axis preferentially oriented. However, the intensity variation of (0002) peak of the films deposited at different substrate temperatures showed an interesting variation. The film deposited at 300 °C showed the strongest (0002) peak. In contrast, all the films deposited at lower temperatures showed much less intense (0002) peaks (by nearly two orders), while those deposited at higher temperatures showed a small and steady decrease in the intensity of (0002) peak with increase of substrate temperature. The films deposited below 300 °C possessed uniform strain (N10− 3), which became negligible in the films deposited above 300 °C. Most interestingly, it was found that the ZnO film deposited at 600 °C exhibited the sharpest absorption edge and a strong and narrow (∼100 meV) room temperature band edge photoluminescence (PL), though it exhibited a lesser degree of c-axis preferred orientation than the film deposited at 300 °C. In order to understand the above mentioned features, an extensive microstructural investigation of ZnO films deposited on quartz substrates at different temperatures has been undertaken, using Xray pole figure analysis, high resolution X-ray diffraction (HRXRD) and transmission electron microscopy (TEM). Raman spectroscopy and photoluminescence measurements were used to obtain additional information on microstructure, crystallinity and defects in the films. A 662 comprehensive attempt has thus been made to establish the role of microstructural parameters in determining the optical quality of polycrystalline ZnO films deposited by reactive sputtering of a zinc target. 2. Experimental details ZnO films were deposited on quartz substrates by reactive rf magnetron sputtering using argon–oxygen mixture. A 99.9% pure Zn target of 3-in. diameter was used. The target to substrate distance was 55 mm. The base pressure was 1 × 10−3 Pa. The flow rates of argon (24 SCCM) and oxygen (6 SCCM) were controlled by mass flow controllers. Deposition was carried out at a working pressure of ∼1 Pa, after presputtering with argon for 10 min. The sputtering power was maintained at 400 W during deposition. The depositions were carried out in the temperature range of room temperature to 600 °C. X-ray diffraction (XRD) studies were performed with a PANalytical X'Pert PRO powder diffractometer using Cu Kα radiation. X-ray pole figures were measured using a PANalytical MRD four-circle diffractometer using CuKα radiation. HRXRD measurements were carried out in ω and ω–2θ scan geometries using PANalytical X'Pert MRD system. The incident beam optics had a 4-bounce hybrid monochromator, which ensured CuKα1 (1.54056A°) output collimated to about 20 arc sec in the plane of scattering. A 1/2° slit was placed at the output before the detector. Transmission electron microscopy (TEM) studies were carried out using a PHILIPS Model CM200 Supertwin Microscope operated at 200 kV. Raman spectra of the films were recorded at room temperature in back scattering geometry, using JOBIN YVON HORIBA HR-800 confocal micro-Raman spectrometer equipped with a 20 mW Ar+ laser (514.5 nm). Photoluminescence (PL) measurements were carried out at room temperature using a 325 nm He–Cd laser and a JOBIN YVON HR-460 monochromator. Ambios XP-2 surface profilometer was used for thickness measurements. 3. Results and discussion 3.1. Microstructural studies All the studies were carried out on films of approximately the same thickness (650 ± 50 nm). The deposition rate was nearly the same (∼ 1 μm/h) for all the films. Typical XRD patterns of the films deposited at 100 °C, 150 °C, 200 °C, 250 °C, 300 °C and 600 °C are shown in Fig. 1. As reported earlier [13], the films deposited upto 100 °C showed a single low order peak due to (0002) reflection, hence only a typical case is shown in Fig. 1. No XRD peaks corresponding to reflections other than (0002) and (0004) planes of hexagonal ZnO were observed in these cases. However, all the films deposited in the range of 150–250 °C, clearly showed the presence of low intensity ̄ and (1011) ̄ reflections, with a dominant (0002) peaks due to (1010) reflection. The film deposited at 200 °C showed relatively higher ̄ peaks. In contrast, the films deposited intensity of (101̄ 0) and (1011) at higher substrate temperatures again showed only (0002) peak with much higher intensity, though the intensity decreased monotonically with increase of substrate temperature from 300 °C to 600 °C. Typical cases of the films deposited at 300 °C and 600 °C are shown in Fig. 1. Thus all the ZnO films exhibit a strong c-axis preferential orientation of crystallites, though the extent of preferred orientation appears to be the highest in the film deposited at 300 °C. It is also seen from Fig. 1 that the position of (0002) peak shifts monotonically towards the bulk value with increase of substrate temperature and the uniform strain in the films deposited above 300 °C is practically negligible, as reported earlier [13]. Fig. 2 shows the pole figures of (0002) reflection for ZnO films deposited on quartz substrates at different temperatures. All the films show a very pronounced (0002) texture, with rotational symmetry. The films deposited at room temperature and 100 °C show broad pole Fig. 1. XRD patterns of ZnO films deposited at different substrate temperatures (as indicated in the figure). All the films were deposited at rf power of 400 W with 20% oxygen in the sputtering atmosphere. figures, which become sharper with initial increase in substrate temperature. The films deposited at 300 °C and 400 °C thus show very sharp and high intensity pole figures, confirming the presence of a strong and nearly complete c-axis orientation of crystallites. Interestingly, at higher substrate temperatures of 500 °C and 600 °C, the pole figures again become broader, indicating enhanced mis-orientation of (0002) planes of crystallites with respect to the film surface. This explains the steady decrease in the intensity of (0002) XRD peak with increase in substrate temperatures above 300 °C, as mentioned above and reported earlier [13]. TEM studies were carried out on typical ZnO films, which were lifted-off from quartz substrates. The film on quartz substrate was immersed in dilute HF for few secs, which resulted in etching at the film-substrate interface and thinning of the film. Subsequently, the film on substrate was carefully transferred to a bath of de-ionized water, which resulted in peeling-off of the film and floating on water surface. The film was transferred onto a 200 mesh carbon coated 663 Fig. 2. (0002) peak pole figures of ZnO films deposited at different substrate temperatures (as indicated in the figure). copper grid for TEM studies. Both bright and dark field images were examined in all the cases. Fig. 3 presents typical TEM results in the form of bright and dark field images, along with the corresponding diffraction patterns and crystallite size distributions for the films deposited at room temperature, 300 °C and 600 °C. Electron diffraction data obtained for all the films confirmed the presence of polycrystalline hexagonal ZnO. The lattice constants (‘a’ and ‘c’) obtained from the diffraction patterns are listed in Table 1 for typical cases. The value of ‘c’, decreases with increasing substrate temperature and becomes close to the bulk value for the films deposited at 300 °C or higher temperatures. These observations are in agreement with the above described XRD results. The lattice constant ‘c’ however, remains close to the bulk value, irrespective of the substrate temperature. The average lateral crystallite size of the films was obtained from dark field TEM images. The values are listed in Table 1. The results are in good agreement with the corresponding bright field images. The crystallite size of the film deposited at room temperature is (21 ± 14) nm, which increases slightly to (25 ± 20) nm for the film deposited at 300 °C. However, for the film deposited at 600 °C, the crystallite size increases substantially to (43 ± 30) nm. The above described microstructural features of ZnO films, especially the pattern of change in preferred orientation of crystallites with substrate temperature is unusual and interesting. The unusual reduction in the extent of c-axis preferred orientation in sputtered ZnO films at intermediate temperatures and its subsequent enhancement at higher temperatures has been reported earlier [14], though without much explanation. There are also some reports on ZnO films deposited by sputtering [10,15] and MOCVD [16], showing a strong caxis preferred orientation near room temperature, followed by its reduction with increase in substrate temperature to the range of 150– 300 °C. In a recent review, Kajikawa [17] has extensively analyzed the issue of preferred orientation in sputtered ZnO films by compiling the relevant experimental observations and explanations put forth for the preferred orientation behaviour. A comprehensive understanding of the complex dependence of c-axis preferred orientation of sputtered ZnO films on processes such as, preferential nucleation, preferential 664 [17,20,21] driven by minimization of surface energy. The preferred orientation can also be significantly affected at the growth stage by two processes, namely, sticking and surface diffusion of ad atoms. Fujimira et al. [21] have proposed that the differences in the sticking probabilities of growing species on different planes determine preferred orientation. For example, in the case of sputtered ZnO films [22,23], the existence of Zn and O in the sputtering gas causes aaxis orientation, while the higher sticking probability of Zn–O species causes c-axis preferred orientation. In the present work, it has been noticed that the preferred orientation effects show a strong dependence on substrate temperature, but are practically independent of rf power (in the range of 300–400 W) and oxygen content (in the range of 10–30%) of the sputtering atmosphere. Hence, effects due to sticking coefficients of growing species have been ignored, while explaining the preferred orientation behaviour. The preferred orientation behaviour of sputtered ZnO films seen in the present work can be explained in the light of the above discussion. As shown in Fig. 1, the films deposited up to 100 °C, exhibit a single XRD peak indicating (0002) preferred orientation, though the peak intensity in these cases are nearly two orders of magnitude smaller, compared to the strongly (0002) oriented film deposited at 300 °C. Further, Fig. 2 shows that the corresponding pole figures for the films deposited up to 100 °C are very broad, indicating significant misorientation of the crystallites. TEM studies of these films have also indicated that the lateral crystallite size is relatively small (∼ 20 nm). On the other hand, in these films, the crystallite size along the growth direction (c-axis) is relatively larger (∼ 100 nm), as obtained from the (0002) peak width. The broad and low intensity pole figures and small lateral crystallite size indicate that although these films exhibit preferred c-axis orientation, the degree of crystallinity and overall structural order in these films is not very high. The presence of disordered component in the films can also not be ruled out. This can be explained by assuming negligible surface diffusion at lower substrate temperatures. Thus, it is primarily due to preferential (0002) nucleation that the c-axis oriented crystallites grow along with some poorer crystallinity material. As the substrate temperature is increased to the range of 150– 250 °C, the XRD patterns (Fig. 1) of the corresponding films show multiple peaks. Though in all these cases, the (0002) peak remains dominant, but it shows a significant increase in its half width. For the film deposited at 200 °C, the (0002) peak shows the largest width corresponding to a value of ∼ 50 nm for the crystallite size along growth direction (c-axis). This behaviour can be explained by considering that at these temperatures, the preferred orientation of crystallites is no longer determined by preferred nucleation. Instead, it is significantly affected by the growth process, controlled essentially by enhanced surface diffusion due to increase in substrate temperature. However, at relatively low substrate temperatures, surface diffusion may be limited to being within the grains, i.e., among planes only. According to Kajikawa et al. [17,20], when surface diffusion takes place among planes within a grain, higher surface energy planes may become preferred orientation planes, because surface diffusion essentially occurs to conceal the plane with higher surface energy. Table 1 Characteristic parameters of ZnO films deposited at room temperature, 300 °C and 600 °C on quartz substrates Fig. 3. Bright field and dark field TEM images of ZnO films deposited at different substrate temperatures (as indicated in the figure). The diffraction pattern and the analysis of the crystallite sizes are shown as insets. crystallization, sticking, surface diffusion and grain growth appears to be lacking. However, it is generally believed (though, this has also been debated [18,19]) to originate from preferential (0002) nucleation Substrate temperature Room temp. 300 °C 600 °C From TEM From HRXRD Lateral crystallite size (nm) ‘c’ (Å) (± 0.01) ‘a’ (Å) (± 0.01) Tilt (degree) Vertical coherence length (nm) Micro-strain 21 ± 14 25 ± 20 43 ± 31 5.25 5.21 5.20 3.26 3.25 3.26 – 2.6 11.1 – 140 ± 15 65 ± 15 – 2.0 × 10− 3 1.1 × 10− 3 665 Such a preferred orientation behaviour has been recently reported [24] and discussed in detail for sputtered hexagonal GaN films. The ̄ ̄ appearance of (1010) and (1011) reflections along with (0002) reflections in the XRD patterns of these films, indicating the presence of differently oriented crystallites with respect to the film surface, is thus attributed to the surface diffusion among planes within the grains. It is noteworthy that the AFM studies of these films reported in Ref. [13], showed that the surface morphological features of two vastly different sizes were present in the film deposited at 200 °C, unlike the films deposited at other substrate temperatures. It is also noticed from Fig. 2 that the pole figure for the film deposited at 200 °C is relatively sharp, which indicates that there is an increased tendency of the (0002) oriented crystallites to align with the growth direction. These features, together with the decrease of crystallite size along the growth direction to ∼ 50 nm indicate that a constrained growth of caxis oriented crystallites along with crystallites with other orientations takes place in this temperature range. At higher substrate temperatures (300 °C and above), the significantly increased surface diffusion coefficient is expected to take the growth process to the regime where surface diffusion among grains becomes dominant. In this temperature regime, the grains with higher energy surface shrink and those with lower energy surface grow laterally to determine the nature of preferred orientation. Consequently, the lowest surface energy plane of ZnO, i.e., (0002) acquires a preferential orientation, as has been observed in the present case at substrate temperatures of 300 °C or higher. The dominance of surface diffusion and tendency of lateral growth in these films is also evident from the increase in lateral crystallite size (Fig. 3) and the large surface features seen in their AFM images, reported earlier [13]. However, an interesting feature of the films deposited at and above 300 °C is the monotonic decrease in the intensity of (0002) peak (Fig. 1) and the broadening of corresponding pole figures (Fig. 2) with increase in substrate temperature. The films deposited at 300 °C and 600 °C were analyzed by high resolution X-ray diffraction (HRXRD) studies, which have provided additional insight into this behaviour. The films deposited below 300 °C could not be studied, owing to poor signal intensity. The HRXRD results are discussed below. A strongly oriented polycrystalline film on amorphous substrates can be considered to consist of crystallites, with certain mean vertical and lateral dimensions. For materials having hexagonal structure, usually the crystallites are oriented with their b0002N axis along the growth direction and are rotated randomly about this direction. Information about lateral and vertical coherence lengths (crystallite sizes), crystallite tilt and micro-strain can be estimated from full width at half maximum (FWHM) of the omega (ω) and ω–2θ scans of (000 l) reflections, using Williamson–Hall plots [25]. The terms vertical and lateral refer respectively, to the growth direction and a direction perpendicular to it, in the plane of the film. The finite crystallite size, tilt and micro-strain cause the broadening of reciprocal space points [26]. The broadening due to finite size is invariant in reciprocal space with higher orders of reflection (increasing scattering vector, q), but the broadening due to both tilt and micro-strain increase with increasing magnitude of q [26]. The finite vertical coherence length and micro-strain cause the broadening, Δqz of (000l) reflections in ω– 2θ scans, which can be respectively, estimated from the intercept and slope of the linear Williamson–Hall plots of Δqz (FWHM along qz direction) vs. q. Here, q is the magnitude of the position of (000 l) point in the reciprocal space. Similarly, the finite lateral coherence length and tilt cause broadening, Δqx in the qx–qy plane, which can be measured by the spread of (000l) reflections in ω-scan. The lateral coherence length and tilt can be estimated respectively, from the intercept and slope of the linear Williamson–Hall plots of Δqx (FWHM along qx direction) vs. q. ω–2θ scans of symmetric (0002), (0004) and (0006) reflections from ZnO films deposited at substrate temperatures of 300 °C and 600 °C are shown respectively, in Fig. 4(a). For all the cases, the values of FWHM in reciprocal space (Δqz) are also indicated. As discussed above, the finite vertical crystallite size and micro-strain were calculated from Williamson–Hall plots (Δqz vs. q), shown in Fig. 4(b). The values of these parameters are also listed in Table 1. It may be noted that the vertical crystallite size of the film deposited at 300 °C is ∼140 nm, as compared to the value of ∼65 nm for the film deposited at 600 °C. The micro-strain (non-uniform strain) is found to be 1.1 × 10− 3 in the film deposited at 600 °C, which is nearly half the value of 2.0×10− 3, measured in the film deposited at 300 °C. Fig. 4. (a) ω–2θ scans for ZnO films deposited at substrate temperatures of 300 °C and 600 °C for (0002), (0004) and (0006) reflections. (b) Broadening in the reciprocal space as obtained from the symmetric scans along ω–2θ-axis (qz) for (0002), (0004) and (0006) reflections. 666 Fig. 5(a) shows the ω scans of symmetric (0002), (0004) and (0006) reflections from ZnO films deposited at 300 °C and 600 °C. For all the cases, the values of FWHM in reciprocal space (Δqx) are also indicated. The ω scans of the film deposited at 600 °C were found to be much broader compared to the film deposited at 300 °C. The corresponding Williamson–Hall plots of Δqx vs. q are shown in Fig. 5(b), for both the films. The linear fits for the two films show significantly different slopes, from which the values of tilt have been estimated and listed in Table 1. The average tilt value of ∼ 2.6° for the film deposited at 300 °C is significantly smaller than the value of ∼11.1° for the film deposited at 600 °C. These results are in tune with the qualitative information obtained from the corresponding pole figures (Fig. 2). The intercepts of Δqx vs. q plots however, possess large uncertainties owing to the broad ω curves, especially in case of the film deposited at 600 °C. This implies the presence of a large error in the estimation of lateral coherence length, hence the lateral crystallite size in these films has been taken as measured by TEM, and listed in Table 1. The microstructural features of the films deposited at 300 °C and 600 °C can now be compared. The average vertical (along growth direction) and lateral crystallite sizes of the film deposited at 300°C are ∼140 nm and ∼25 nm, respectively. This film also exhibits a very sharp (0002) pole figure (Fig. 2) and correspondingly a small value (∼2.6°) of tilt. It clearly possesses the highest quality in terms of c-axis crystallite orientation, in comparison to all the other films. As the substrate temperature is increased, the pole figures become broader, indicating that the mis-orientation of (0002) planes of crystallites with respect to the film surface increases. Consequently, the film deposited at 600 °C shows a much larger value of tilt (∼11.1°). Further, the average vertical and lateral crystallite sizes for this film are ∼65 nm and ∼43 nm, respectively, indicating that with increase in substrate temperature, the vertical crystallite size decreases, but the lateral crystallite size increases. It can thus be inferred that the film deposited at 600 °C consists of relatively equiaxed crystallites with slightly larger overall crystallite volume compared to the film deposited at 300 °C. These observations can be explained in accordance with the structure-zone model, proposed originally by Thornton [27]. According to this model [28], surface diffusion controlled columnar growth usually takes place at intermediate temperatures. At higher substrate temperatures, lattice and grain boundary diffusion processes begin to dominate, leading to formation of equiaxed re-crystallized grains. The results presented above show that a transition from columnar crystallites to equiaxed crystallites takes place with increase of substrate temperature. The increase in tilt seen in the films deposited above 300 °C is thus indicative of misorientation amongst equiaxed crystallites, which is attributed to the absence of columnar growth in sputtered films deposited on amorphous substrates, at relatively higher temperatures. The HRXRD studies also show that the micro-strain present in the film deposited at 600 °C is nearly half of that in the film deposited at 300 °C. The higher micro-strain in the film deposited at 300 °C is indicative of a larger presence of point defects in this film. It may be recalled from earlier work [13], that the uniform strain in the films deposited below 300 °C was considered intrinsic to the growth process. The intrinsic strain was attributed to the O− ions formed at the target, which may have sufficient energies to bombard the growing film [29], causing implantation, displacement or removal of surface atoms. The relatively large micro-strain present in the film deposited at 300 °C is also attributed to bombardment due to O− ions. The significant decrease in micro-strain seen in the film deposited at 600 °C is attributed to annealing effects, leading to reduction in point defects. 3.2. Raman spectroscopy studies Fig. 5. (a) ω-scan for ZnO films deposited at substrate temperatures of 300 °C and 600 °C for (0002), (0004) and (0006) reflections. (b) Broadening in the reciprocal space as obtained from the symmetric scans along ω-axis (qx) for (0002), (0004) and (0006) reflections. Fig. 6 shows the Raman spectra of ZnO films deposited at different substrate temperatures. The spectrum of the quartz substrate is also shown for comparison. All the ZnO films deposited in the substrate temperature range of room temperature to 250 °C show two very broad peaks centered ∼400 cm− 1 and ∼ 570 cm− 1. The broad peak ∼400 cm− 1 is attributed to a combination of A1(TO) and E1(TO) modes (TO denotes transverse optical) of ZnO, which are reported at 380 cm− 1 and 407 cm− 1, respectively [30,31]. This broad peak has a shoulder on the higher frequency side at ∼440 cm− 1, which is assigned 667 compared to the film deposited at 300 °C, shows a much larger tilt (mis-orientation of crystallites). This result clearly shows that the mis-orientation of crystallites does not significantly affect the intrinsic quality of crystallites in terms of point defects and strain. Raman results have thus shown that the mere appearance of a single (0002) XRD peak in ‘a film’, which is usually taken as indicative of a strong c-axis orientation of crystallites and hence its ‘high’ crystalline quality, can be misleading. It is clear that additional data is required to make such assertions. 3.3. Photoluminescence studies Absorption studies on ZnO films have shown [13] that all the films exhibited band gaps ∼3.3 eV. It was however, observed that with increase in substrate temperature, the absorption edge became sharper and sub-band gap absorption was significantly reduced. It was also shown that the ZnO film deposited at 600 °C exhibited a strong and narrow band edge photoluminescence peak. In this section, a detailed study of the near band edge and defect level luminescence at room temperature is presented for the ZnO films deposited at different substrate temperatures. Fig. 6. Raman spectra of ZnO films deposited at different substrate temperatures (as indicated in the figure), along with the spectrum of the quartz substrate. to E2-high mode [30,31]. These features are rather unexpected, since in a c-axis oriented ZnO film in backscattering geometry, only a strong E2-high mode and a weak A1(LO) mode (LO denotes longitudinal optical) are expected to be present [1,30]. The broad peak ∼ 570 cm− 1 is assigned to a combination of A1(LO) and E1(LO) modes, which are reported at 574 cm− 1 and 583 cm− 1, respectively [30,31]. These features of Raman spectra of ZnO films are indicative of poor crystallinity, arising from lattice strain, structural defects and disorder [32–34]. These observations, especially the presence of weak E2-high mode as a shoulder, indicate that though most of the films deposited below 300 °C showed [13] a single and intense (0002) peak, indicating c-axis preferred orientation of crystallites, these films actually possess poor crystalline quality. The features seen in the Raman spectra are thus consistent with the significantly low intensity of (0002) XRD peaks (Fig. 1), broad pole figures (Fig. 2), small lateral crystallite size (Fig. 3) and the high uniform strain (N10− 3) exhibited by these films (reported in Ref. [13]). The film deposited at 300 °C however, shows drastically different spectral features. An enhancement of E2-high intensity is clearly seen, though a shoulder due to TO modes and a broad band due to LO modes continue to exist. The increase in the prominence of E2-high peak corroborates well with the significant improvement in c-axis orientation of crystallites, as also seen from the corresponding XRD patterns and pole figures. The films deposited at higher substrate temperatures show a monotonic increase in the intensity and sharpness of the E2-high peak and decrease in the intensity of TO and LO mode peaks. Interestingly, the film deposited at 600 °C shows only an intense E2-high peak along with a very weak and broad A1 (LO) mode at ∼ 575 cm− 1. As mentioned above, such features are characteristic of strong c-axis orientated ZnO films [1,31]. The small shift of LO modes towards higher frequencies and the reduction in their intensity are attributed to decrease in point defects [31]. The decrease in point defects correlates well with the significant decrease in micro-strain and the small increase in crystallite size, for the film deposited at 600 °C. It is interesting to note that such features are seen in the ZnO film deposited at 600 °C, which, as Fig. 7. PL spectra of ZnO films deposited at different substrate temperatures (as indicated in the figure). 668 Fig. 7 shows the room temperature PL spectra of ZnO films deposited at different substrate temperatures. All the films show strong emission at ∼ 376 nm, attributed to near band edge luminescence of ZnO [35]. Weak and broad bands are also seen in visible region, which are known to arise from oxygen vacancies and zinc interstitials [36,37]. The intensity of band edge luminescence is found to increase with substrate temperature, accompanied by narrowing of the peak, as reported earlier for sputtered ZnO films on Si [38]. The films deposited at room temperature, 100 °C and 200 °C showed very weak band edge emission and negligible luminescence in the visible region. This is attributed to poor crystallinity and large disorder in the films, as seen earlier from UV–visible absorption studies [13], which showed large sub-band gap absorption and is in tune with the Raman results. It has also been reported that ZnO films with smaller crystallites show poorer luminescence, which has been attributed to non-radiative relaxation through surface states [38,39]. With increase in substrate temperature, the intensity of band edge emission begins to increase together with the defect related luminescence in the visible region. The film deposited at substrate temperature of 400 °C thus showed comparable intensities of band edge and defect luminescence. The increase in both band edge as well as defect luminescence together, can be attributed to decrease in non-radiative relaxation processes. With further increase in substrate temperature to 500 °C, the intensity of band edge emission begins to increase faster than the defect related luminescence. The band edge luminescence intensity however, shows a drastic increase for the film deposited at 600 °C, while the defect related luminescence bands remain practically unchanged and weak. The half width of the band edge luminescence peak is also significantly reduced to ∼ 102 meV. The sharp and intense band edge PL from the film deposited at 600 °C is in line with the microstructural and Raman studies, which indicated reduction of point defects in this film. This is further supported by the presence of weak defect related luminescence in the visible region. It has been reported [36] that the visible emission in ZnO originates from oxygen vacancy or zinc interstitial related defects, and gets significantly reduced by the improvement in stoichiometry of ZnO films. It may be pointed out that the band edge PL peak width of 102 meV at room temperature is quite comparable to the reported values for ZnO films deposited by MBE [40], MOVPE [35,41] and PLD [37], on single crystalline substrates. It is also interesting to note that the most intense and narrow PL does not come from the ZnO film deposited at 300 °C, which exhibited the strongest and nearly complete c-axis preferred orientation of crystallites. This result is important and interesting, as it indicates that the main cause of the enhancement and narrowing of band edge luminescence in polycrystalline ZnO films is the decrease of micro-strain, which results from the reduction in point defects, such as oxygen vacancies and zinc interstitials. It may also be mentioned that the film deposited at 600 °C possessed equiaxed crystallites with a slightly larger overall size, as compared to the film deposited at 300 °C. This may also partly contribute to its superior optical quality. 4. Conclusions Polycrystalline ZnO films were deposited on quartz substrates by reactive sputtering of zinc target. The films deposited below 300 °C showed preferred c-axis orientation of crystallites but Raman and PL studies indicated their poor crystallinity and optical quality. The films deposited up to substrate temperatures of 100 °C, show c-axis preferred orientation due to preferential nucleation. The appearance of multiple orientations in the substrate temperature range of 150– 250 °C is attributed to surface diffusion between planes within the grains. The film deposited at 300 °C and higher temperatures showed a strong c-axis orientation of crystallites due to the dominance of surface diffusion between grains. However, with increase in substrate temperature from 300 °C to 600 °C, the c-axis orientation was again found to steadily decrease, owing to transition from columnar structure to equiaxed structure of crystallites. The micro-strain present in the film deposited at 600 °C was nearly half of that in the film deposited at 300 °C, which is attributed to lesser point defects. Raman and PL studies also show that the film deposited at 600 °C possesses lesser oxygen-deficiency related point defects, which correlates well with the observed decrease in its micro-strain to ∼ 10− 3. As a consequence of reduction in optically active point defects, the ZnO film deposited at 600 °C showed a high intensity and narrow band edge luminescence with FWHM of ∼ 102 meV at room temperature. Polycrystalline ZnO films deposited by reactive sputtering of a metallic zinc target have not been usually reported to possess such structural and optical quality. The reduction in defects is attributed to the presence of energetic zinc ad atoms having high surface mobility as well as annealing effects. It is also inferred that the micro-strain present within the crystallites plays the most crucial role compared to other microstructural parameters, in determining the optical quality of polycrystalline ZnO films. This work has shown that the mere appearance of a single and intense (0002) peak in ‘a film’ is not a sufficient indicator of its ‘high’ crystalline quality. This is particularly relevant to the case of polycrystalline ZnO films, which under most conditions of deposition exhibit a single or intense (0002) XRD peak. It is also clear that in general, there is a need to examine other aspects of microstructure, such as, crystallite size, crystallite tilt and uniform as well as nonuniform strain (micro-strain) and also independently assess the presence of optically active defects, before drawing meaningful conclusions about the ‘quality’ of polycrystalline semiconductor films. Acknowledgements The financial support from MHRD (Govt. of India) for this work is gratefully acknowledged. Sukhvinder Singh is thankful to CSIR, New Delhi (India) for Senior Research Fellowship. Tapas Ganguli and Ravi Kumar are thankful to Dr. S. M. Oak for support and encouragement. The National OIM Texture Facility of IIT Bombay is gratefully acknowledged for X-ray pole figure measurements. SAIF and CRNTS, IIT Bombay are respectively acknowledged, for providing TEM and Raman facilities. References [1] U. Ozgur, Ya.I Alivov, C. Liu, A. Teke, M.A. Reshchikov, S. Dogan, V. Avrutin, S.J. Cho, H. Morkoc, J. Appl. Phys. 98 (2005) 041301 [and references therein]. [2] S.J. Pearton, D.P. Norton, K. Ip, Y.W. Heo, T. Steiner, J. Vac. Sci. Technol. B 22 (2004) 932 [and references therein]. [3] C. Klingshirn, R. Hauschild, H. Priller, M. Decker, J. Zeller, H. Kalt, Superlattices Microstruct. 38 (2005) 209 [and references therein]. [4] K. Ellmer, J. Phys. D: Appl. Phys. 33 (2000) R17 [and references therein]. [5] K.-K. Kim, J.-H. Song, H.-J. Jung, W.-K. Choi, S.-J. Park, J.-H. Song, J. Appl. Phys. 87 (2000) 3573. [6] K.C. Ruthe, D.J. Cohen, S.A. Barnett, J. Vac. Sci. Technol. A 2 (2004) 2446. [7] I.S. Kim, S.H. Jeong, B.T. Lee, Semicond. Sci. Technol. 22 (2007) 683. [8] T. Koyama, S.F. Chichibu, J. Appl. Phys. 95 (2004) 7856. [9] J.W. Park, Y. Park, J.W. Park, M. Jeon, J.K. Lee, J. Vac. Sci. Technol. A 23 (2005) 1. [10] R. Ondo-Ndong, F. Pascal-Delannoy, A. Boyer, A. Giani, A. Foucaran, Mater. Sci. Eng. B 97 (2003) 68. [11] T. Inukai, M. Matsuoka, K. Ono, Thin Solid Films 257 (1995) 22. [12] S. Zhu, C.H. Su, S.L. Lehoczky, P. Peters, M.A. George, J. Cryst. Growth 211 (2000) 106. [13] S. Singh, R.S. Srivivasa, S.S. Major, Thin Solid Films 515 (2007) 8718. [14] L.J. Meng, M. Andritschky, M.P. dos Santos, Vacuum 45 (1994) 19. [15] K. Wasa, S. Hayakawa, Thin Solid Films 7 (1971) 135. [16] X.L. Chen, X.H. Geng, J.M. Xue, D.K. Zhang, G.F. Hou, Y. Zhao, J. Cryst. Growth 296 (2006) 43. [17] Y. Kajikawa, J. Cryst. Growth 289 (2006) 387. [18] A. Wander, N.M. Harrison, Surf. Sci. 468 (2000) L851. [19] A. Wander, F. Schedin, P. Steadman, A. Norris, R. McGrath, T.S. Turner, G. Thornton, N.M. Harrison, Phys. Rev. Lett. 86 (2001) 3811. [20] Y. Kajikawa, S. Noda, H. Komiyama, J. Vac. Sci. Technol. A, 21 (2003) 1943. [21] N. Fujimura, T. Nishihara, S. Goto, J. Xu, T. Ito, J. Cryst. Growth 130 (1993) 269. [22] C.R. Aita, A.J. Purdes, R.J. Lad, P.D. Funkenbusch, J. Appl. Phys. 51 (1980) 5533. 669 [23] C.R. Aita, M.E. Marhic, J. Appl. Phys. 52 (1981) 6584. [24] B.S. Yadav, S.S. Major, R.S. Srinivasa, J. Appl. Phys. 102 (2007) 073516. [25] R. Chierchia, T. Böttcher, H. Heinke, S. Einfeldt, S. Figge, D. Hommel, J. Appl. Phys. 93 (2003) 8918. [26] M.E. Vickers, M.J. Kappers, R. Datta, C. McAleese, T.M. Smeeton, F.D.G. Rayment, C.J. Humphreys, J. Phys. D: Appl. Phys. 38 (2005) A99. [27] J.A. Thornton, Ann. Rev. Mater. Sci 7 (1977) 239. [28] M. Ohring, Materials Science of Thin Films, Academic Press, San Diego, 2002. [29] O. Kappertz, R. Drese, M. Wuttig, J. Vac. Sci. Technol. A 20 (2002) 2084. [30] T.C. Damen, S.P.S. Porto, B. Tell, Phys. Rev. 142 (1966) 570. [31] N. Ashkenov, B.N. Mbenkum, C. Bundesmann, V. Riede, M. Lorenz, D. Spemann, E.M. Kaidashev, A. Kasic, M. Schubert, M. Grundmann, J. Appl. Phys. 93 (2003) 126. [32] Y.W. Chen, Y.C. Liu, S.X. Lu, C.S. Xu, C.L. Shao, C. Wang, J.Y. Zhang, Y.M. Lu, D.Z. Shen, X.W. Fan, J. Chem. Phys. 123 (2005) 134701. [33] J.Z. Wang, M. Peres, J. Soares, O. Gorochov, N.P. Barradas, E. Alves, J.E. Lewis, E. Fortunato, A. Neves, T. Monteiro, J. Phys.: Condens. Matter 17 (2005) 1719. [34] M. Gomi, N. Oohira, K. Ozaki, M. Kayano, Jpn. J. Appl. Phys. 42 (2003) 481. [35] Y. Ma, G.T. Du, T.P. Yang, D.L. Qiu, X. Zhang, H.J. Yang, Y.T. Zhang, B.J. Zhao, X.T. Yang, D.L. Liu, J. Cryst. Growth 255 (2003) 303. [36] S.-H. Jeong, B.-S. Kim, B.-T. Lee, Appl. Phys. Lett. 82 (2003) 2625. [37] X.M. Fan, J.S. Lian, Z.X. Guo, H.J. Lu, Appl. Surf. Sci. 239 (2005) 176. [38] S.H. Jeong, J.K. Kim, B.T. Lee, J. Phys. D: Appl. Phys. 36 (2003) 2017. [39] T. Matsumoto, H. Kato, K. Miyamoto, M. Sano, E.A. Zhukov, Appl. Phys. Lett. 81 (2002) 1231. [40] Y. Chen, D.M. Bagnall, H.J. Koh, K.T. Park, K. Hiraga, Z. Zhu, T. Yao, J. Appl. Phys. 84 (1998) 3912. [41] X. Wang, S. Yang, J. Wang, M. Li, X. Jiang, G. Du, X. Liu, R.P.H. Chang, J. Cryst. Growth 226 (2001) 123.
© Copyright 2025 Paperzz