Otwarty dostęp

Improving optical and morphological properties of Mn-doped ZnO via Ar ion sputtering followed by high-temperature UHV annealing


Zacytuj

Introduction

Due to its exceptional electronic and optical properties, including a substantial exciton binding energy of 60 meV and a wide band gap energy of 3.37 eV at room temperature, zinc oxide (ZnO) has garnered significant interest as a transparent conductor oxide (TCO). This versatile material finds applications in various fields, such as gas sensors [1], catalysts [26], photovoltaic solar cells [7], supercapacitors [8], and flexible piezoelectric nanogenerators [9]. Numerous researchers have dedicated their efforts to exploring the effect of doping on ZnO, aiming to enhance its optical, electrical, and magnetic characteristics. Incorporating minute quantities of dopant metal oxide materials, such as Mn2+, Mg2+, Sb3+, In3+, Cu2+, Co3+, Al3+, and others [1018], represents the primary approach to synthesizing ZnO alloys with widened band gaps in various nanostructures, including nanofibers, nanocônes, nanowires, and more [1921].

Transition metal (TM) ion-doped zinc oxide (ZnO) has gained significant scientific interest due to its remarkable properties and potential for spintronic applications, where the coexistence of electron spin and charge plays a crucial role.The strong correlation between the “3D” orbital of transition metals and the “s” and “p” orbitals of the anion profoundly alters the electronic structure of the ZnO host lattice. Consequently, ZnO-based dilute magnetic semiconductors (DMS) have emerged as a promising class of materials, surpassing conventional semiconductors in their potential applications [22, 23]. Extensive research efforts have focused on incorporating transition metals like Mn, Fe, Co, Cu, Ni, Cr, and V as substitutional ions in the ZnO lattice to achieve ferromagnetic properties at or above room temperature.

Manganese (Mn2+) is distinguished among the various TM elements because of its notable thermal solubility in ZnO [24]. With its half-filled 3D orbitals housing five spins, Mn2+ exhibits a maximum moment value of 5μB/Mn (where μB represents the Bohr magneton) [25]. Despite the difference in ionic radii between Mn2+ (0.083 nm) and Zn2+ (0.074 nm) and the limited solubility of approximately 13% in the ZnO matrix [25], extensive studies confirm that manganese ions can effectively substitute for zinc ions without inducing distortion in the ZnO lattice. At low concentrations (1%–5%), this substitution process provides favorable properties such as homogenous grain shape, solid texture, and improved crystallinity [6, 2628]. Consequently, ZnMnO thin films prove to be an ideal material for short-wavelength magneto-optical applications.Currently, a variety of techniques are employed for the synthesis of ZnO nanomaterials, including hydrothermal methods [2931], solution processing [32], pulsed laser deposition [33], and pyrolysis sputtering [10, 13, 34]. Among these approaches, pyrolysis is favored in this study because of its convenience of use, nonvacuum, good homogeneity, and cost-effectiveness. Moreover, as a critical phase in the development of reliable devices, the ZnMnO samples undergo Ar+ ion sputtering [35], followed by a precisely controlled thermal annealing process within the ultra-high vacuum (UHV) system [36]. Prior to treatment, the ZnO samples are investigated using XRD, XPS, AFM, and PL techniques, while the ZnMnO samples are examined both before and after treatment.

The aim of this research is to explore the impact of a 4% Mn dopant and investigate how cleaning and thermal annealing processes (Ar+ sputtering followed by high-temperature UHV annealing) affect the structure, morphology, and optoelectronic properties of ZnMnO film. Furthermore, establish a correlation between the electronic band gaps, band transitions, and the structural and morphological characteristics of these thin films.

Experimental section
Preparation process

On a silicon (Si) substrate, thin films of pure ZnO and ZnMnO (4at% Mn) were prepared using the ultrasonic spray pyrolysis technique. Zinc acetate dehydrates (ZnC4H6O4_2H2O) (99%, Sigma Aldrich) were dissolved in 100 ml of methanol at a molar concentration of 0.4 mol/l to produce ZnO. Manganese chloride tetrahydrate (MnCl2 4H2O) (Sigma Aldrich, 98%) is used as a dopant to create ZnMnO. The precursor solution contains 4% of Mn as the molar ratio concentration |Mn|/(|Mn|+|Zn|). To make sure the two solutions are homogeneous, they are each stirred for one hour at 40°C. The elaboration procedure was succeeded by spraying the precursor solutions over the heated substrates for 60 min at a temperature of 400°C.

Treatment process

Using a Scientaomicron UHV system with a preparation chamber and an analysis chamber, thin films of manganese-doped ZnO at 4% Mn were created. The samples were transported into the preparation or analysis chamber by a magnetically coupled probe transporter, and each chamber was pumped separately using manual or electropneumatic gate valves. When it comes to cleaning, an ion gun with a pulverized source is used to bombard the samples’ surfaces in order to get rid of different contaminants like carbon and oxygen that are on the surface layer. The cathodic sputtering ion source’s low energy is sufficient to produce flawless cleaning. It should be noted that the pressure of Ar+ ion gas released by a leaky valve becomes 2.10−6 mbar. With a low ion current of 3 A, the energy is fixed at 1.2 keV for 15 min. Using a standard manipulator and a pyrolytic boron nitride (PBN)-resistive device, the heating process was conducted in steps of increasing temperature (from 350°C up to 800°C) for 20 min at a time.

Characterization details

To study the structural properties, the EMPYREAN Diffractometer System was employed for conducting X-ray diffraction (XRD) analysis on ZnO and ZnMnO thin films. Operating at 45 kV and 40 mA, the system utilizes CuKα radiation (λ = 1.54060 Å) to capture the XRD spectrum within the angular range of 10° to 66°, employing a step size of 2θ° = 0.02°. To obtain X-ray photoelectron spectra (XPS) of the samples, a dual-anode X-ray source (DAR400) [31] was utilized. The XPS spectra were acquired by exciting the samples with monochromatic Mg-Kα (1253.6 eV) radiation. The anode voltage was set to 15 kV, and the emission current was maintained at 15 mA, corresponding to a power of 225 W. Casa-XPS software was used to calculate the atomic concentrations (at.%) of the chemical elements present on the surface. For the characterization of surface morphology and roughness, the Flex-Axiom Nanosurf atomic force microscope (AFM) was employed. Gwyddion data analysis software was utilized to process the topographic images obtained from the AFM measurements. For the purpose of investigating the presence of defects,photoluminescence (PL) measurements were carried out at room temperature using the HORIBA iHR-550 spectrometer. The measurements were conducted within the wavelength range of 340–1040 nm, and a power of 30 mW was carefully chosen as the excitation source to ensure its suitability for effectively exciting wide bandgap semiconductors.

Results and discussion
AFM result

Figure 1 exhibits the line profiles of ZnO and ZnMnO, accompanied by the 2D and 3D AFM images, denoting pure ZnO (a), untreated ZnMnO (b), and treated ZnMnO (c), respectively. The surface morphology of ZnMnO differs significantly from that of ZnO, necessitating a precisely controlled treatment procedure involving Argon ion (Ar+) beam sputtering [37], followed by thermal annealing [38]. The purpose of this treatment is to enhance surface roughness and diminish grain size. Notably, Figure 1c illustrates the outcome of this process. Roughness, a parameter associated with geometric defects on the actual surface, serves as an informative indicator of the adhesion and surface conditions of the deposited layers. ArzuÇolak, Zandvliet, and Poelsema [39] previously conducted comprehensive assessments of adhesion force in relation to surface morphology and power spectral density on both smooth and etched surfaces. Additionally, this technique facilitates the determination of particle size and the distance between different grain boundaries [40]. The AFM images are substantiated quantitatively in terms of Ra (average roughness), Rms (root mean square), maximum height (Zmax), and median height (Zmed), all acquired from the topographic information rendered by the atomic force microscope maps. A summary of these findings is presented in Table 1. Manganese has a significant effect on the atomic arrangement, forming small, well-resolved crystal grains on the ZnMnO surface. In addition, UHV treatment significantly improves the morphology of ZnMnO, as inferred from the mean roughness Sa calculated using Gwydion software (43.03 nm and 13.5 nm for pure ZnO and ZnMnO, respectively). Importantly, the maximum value of 288.84 observed in ZnMnO grains surpasses that of pure ZnO by threefold. This discovery holds profound significance since a rough surface optimizes the conversion of light, thereby heightening the efficiency of solar cells [41].

The main statistical parameters calculated through the AFM images in Figure 1 using Gwyddion software

Sample Maximum height (Zmax) (nm) Median height (Zmed) (nm) Mean roughness (Sa) (nm) Number of grains (N)
ZnO 98.84 60.22 13.50 115
Untreated ZnMnO 198.0 106.46 27.05 152
Treated ZnMnO 288.84 132.52 43.03 126

Fig. 1.

2D and 3D topographic AFM images accompanied by curves of line profiles corresponding to pure ZnO (a), untreated ZnMnO (b), and treated ZnMnO (c)

XRD results

Figure 2 illustrates the XRD patterns obtained for the two samples. The diffraction peaks of pure ZnO and treated ZnMnO exhibit well-matched indexing with the würtzite crystal structure (space group P63mc) [JCPDF card 36-1451], with preferable orientation along (002). This suggests that the synthesized ZnO crystals are predominantly aligned along the c-axis. Moreover, it is worth mentioning that the c/a values closely resemble those of the ideal würtzite structure (ca=83)$({c \over a} = \sqrt {{8 \over 3}} )$.

Fig. 2.

XRD spectra of undoped ZnO and treated ZnMnO With six pics (100), (002), (101), (102), (110) and (103)

Figure 2 shows fluctuating disturbance halos appearing on the diffraction spectrum above (hump) and below (hollow) the line representing the noise contained in the diffraction. The spectra of the two ZnO and ZnMnO samples are located in the two regions: 2θ (30°–32.5°) and (37.5°–47.5°), repectively. This appearance is caused, as indicated by other authors [42, 43], by amorphous materials emerging during heat treatment undergone by the two samples during the processes of their successive elaborations. Such treatment forms ultrathin layers of silicon oxide. or silicate for the as-prepared ZnO. Moreover, it can be seen that these halos become more pronounced and wider for the ZnMnO samples. This increase in dimensions is-due to the addition of an ultra-thin layer of manganese oxide. MnO or Mn2O3 on the surface of the manganese-doped ZnO sample.

The XRD analysis reveals that the ZnMnO crystal is free from any additional phases or impurities. Interestingly, the major peaks observed in the XRD pattern of ZnMnO show a slight shift towards lower angles compared to pure ZnO. This shift suggests a minor expansion of the unit cell in the ZnMnO sample. Consequently, the calculated lattice parameters for ZnMnO are slightly larger than those for undoped ZnO. These findings strongly suggest the successful incorporation of Mn2+ ions (with an ionic radius of 0.83 Å) into the Zn2+ lattice sites (with an ionic radius of 0.74 Å). The findings of the XRD analysis are summarized in Table 2, which provides the position peaks (2θ°), corresponding diffraction planes, and the corresponding dhkl interplanar spacing. Notably, these results demonstrate a high level of consistency with similar XRD studies conducted on Mn-doped ZnO films synthesized through spray pyrolysis [4449].

Structural parameters of ZnO and treated ZnMnO: tabulated XRD spectroscopy results

Sample Peak position (2θ °) hkl FWHM (2θ°) d-spacing (Å) Average grain size D (nm) Cell parameters (Å) c/a Strain (ε)
Scherrer equation W-H Plot a = b c Scherrer equation W-H Plot
ZnO(0%) 31.84 1 00 0.58 2.80 25.20 28.88 3.242 5.195 1.602 0.0014 0.00085
34.50 002 0.33 2.60
36.32 1 0 1 0.30 2.47
47.65 1 02 0.33 1.90
56.68 1 1 0 0.36 1.62
63.02 1 03 0.83 1.47
ZnMnO(4%) 31.70 1 00 0.56 2.82 12.41 12.38 3.256 5.212 1.600 0.0025 0.00122
34.38 002 0.67 2.60
36.17 1 0 1 0.36 2.48
47.56 1 02 0.29 1.91
56.44 1 1 0 0.21 1.63
62.87 1 03 0.30 1.47

Moreover, the average grain size (D) is determined through the XRD spectroscopy study by calculating it as a function of the broadening of the highest intensity peak corresponding to the diffraction plane (002). This calculation utilizes Debye Scherer’s formula, as given in Eq. 1 [50]. D=0.9λβ(rd)cosθ $$D = {{0.9\lambda } \over {\beta (rd)\cos \theta }}$$ Where λ is the X-ray wavelength (Cu-Kα: 1.5406 A°), β (rd) is the full width-half-maximum (FWHM), and θ (rd) is the peak position. The lattice parameters (a, c, and c/a) were calculated using formula (2): 1dhkl2=43(h2+hk+k2a2)+l2c2 $${1 \over {d_{hkl}^2}} = {4 \over 3}({{{h^2} + hk + {k^2}} \over {{a^2}}}) + {{{l^2}} \over {{c^2}}}$$ Where dhkl is the interplanar spacing determined by Bragg’s law and h, k, and l are the Miller indices. the strain (ε) in the film is calculated using the following formula: ε=β(rd)cosθ4 $$\varepsilon = {{\beta (rd)\cos \theta } \over 4}$$

The particle size of pure ZnO measures 25.20 nm, whereas the crystallite size of ZnMnO demonstrates a significant reduction to 12.41 nm. This decrease can be attributed to the enhanced solubility of Mn in ZnO, as observed in the study by Staumal et al. [51]. Additionally, it is important to note that the material experiences an increase in strain. This change arises from multiple factors, including the exchange of ions with different ionic radii, shifts in lattice constants, the introduction of impurities or defects, and the thermal extinction coefficient between the substrate and the film [52].

Figure 3 depicts the variance in βT cos(θ) vs. 4Sin(θ) (Williamson–Hall analysis) [53]. Equation (4) depicts (in linear form) y=mx + c, where m = strain and c = Kλ/D, such that the slope of the linear plot of cos(θ) versus Sin(θ) is lattice strain ε, and the intercept is Kλ/D. βTcosθ=ε(4sinθ)+KλD $${\beta _T}\cos \theta = \varepsilon (4\sin \theta ) + {{K\lambda } \over D}$$

Fig. 3.

Size-strain analysis (W-H plot) for ZnO and ZnMnO

The W-H analysis for grain size look comparable; however, the results for strain appear to be widely off. These findings are consistent with the results obtained from the AFM study, providing further support for the notion that the diminished crystallite size predominantly stems from the distortion of the ZnO host lattice caused by the presence of Mn impurities. Consequently, this distortion facilitates accelerated crystal growth and the amalgamation of crystal grains.

XPS results

This study utilizes X-ray photoelectron spectra (XPS) to investigate the impact of Mn incorporation and UHV treatment on the chemical composition of ZnO thin films. Representative XPS results of clean ZnO and ZnMnO(untreated, sputtered, and annealed) are showcased in Figure 4(a). Notably, all photoelectron peaks (Zn 2p, Mn 2p, O 1s, C 1s, Zn 3s, Zn 3p, and Zn 3d) and Auger transition peaks (C-KLL, O KLL, and Zn LMM) associated with ZnO or ZnMnO materials are accurately identified. For data correction, the C 1s peak, located at 284.6 eV, is utilized as a reference. The atomic percentages (at%) of oxygen, manganese, and zinc elements are derived from the O 1s, Mn 2p3/2, and Zn 2p3/2 signals and are presented in Table 3. Figure 4(b) reveals a noteworthy observation that the C1s carbon signal is more pronounced in ZnMnO-compared to clean ZnO. Furthermore, the untreated ZnMnO exhibits a more significant C1s carbon signal than the annealed ZnMnO.

Fig. 4.

XPS spectra of pure ZnO (i), initial ZnMnO (ii), ZnMnO post Ar ion sputtering (iii), and ZnMnO post Ar annealing at 800°C (iv). (a): full scan spectra. (b), (c), (d), and (e): narrow scan spectra corresponding to C 1s, doubled Zn 2p, O 1s, and doubled Mn 2p, respectively

Atomic percentages of Zn, O, and Mn elements corresponding to pure ZnO and treated ZnMnO. (a) atomic ratio of the Mn atoms taking the Zn sites. (b) stoichiometry factor of O to Zn and Mn

Sample at(%) Comments
Zn O Mn (a): [Mn][Mn+Zn]×100${{[Mn]} \over {[Mn + Zn]}} \times 100$ (b): [O][Mn+Zn]${{[O]} \over {[Mn + Zn]}}$
ZnO 34.00 66.00 0.00 0.00 1.94
ZnMnO@800°C 22.30 74.04 3.65 14.07 2.85

In Figure 4(c), the presence of the Zn 2p core level, which represents the divalent state of zinc ions in ZnO, is confirmed. The Zn 2p3/2 (1021.90 eV) and Zn 2p1/2 (1044.91 eV) photoelectron peaks exhibit a double peak energy separation of 23.01 eV, consistent with previous studies [54, 55]. However, when 4% Mn doping is introduced, the ZnMnO signals of Zn 2p3/2 and Zn 2p1/2 become less resolved due to the presence of a carbon contamination layer on the surface [56, 57]. Notably, following UHV treatment of ZnMnO, there is a substantial increase in the intensity of the Zn 2p signals, accompanied by a shift towards lower binding energies (approximately 0.45 eV) [58]. This observed shift could be attributed to morphological changes, as illustrated by the AFM images.

Figure 4(e) presents the high-resolution Mn 2p spectra, revealing prominent peaks at binding energy values of 641.8 eV and 653.9 eV, which correspond to Mn 2p3/2 and Mn 2p1/2, respectively. Interestingly, the Mn 2p signals exhibit negligible intensity in the contaminated state. However, upon subjecting the samples to UHV treatment, a notable improvement in peak shapes was observed.

Gaussian deconvolution was employed to explore the distribution of the Mn element within the thin films. As shown in Figure 5(a), the presence of distinct peaks at 641.10 eV and 653.0 eV provides unequivocal evidence for the existence of Mn2+ ions. These findings not only corroborate the Mn2+ oxidation state but also affirm the successful substitution of Mn2+ ions into the Zn site positions within the ZnO lattice. Furthermore, higher binding energy peaks at 642.55 eV and 654.60 eV are unequivocally assigned to Mn3+ ions present in the ZnMnO thin film. The coexistence of two charge states, Mn+2 and Mn+3, in Mn-doped ZnO has also been reported in previous studies [59, 60], further validating the observed Mn3+ peaks. Additionally, the presence of two satellite peaks, located around 645.32 eV and 657.15 eV, can be attributed to the electron shake-up of Mn2+.

Fig. 5.

Gaussian deconvolution of high-resolution XPS spectrum of: (a) doubled Mn 2p corresponding to treated ZnMnO(final state); and (b) O 1s corresponding to ZnMnO (untreated, annealed)

Figure 4(d) presents the high-resolution oxygen 1s signal, showcasing a binding energy of 530.65 eV for the annealed ZnMnO sample. Notably, the O 1s peak undergoes asymmetry upon Mn doping but regains its symmetrical shape following UHV treatment. Deconvolution of the O 1s peak in untreated ZnMnO, as illustrated in Figure 5(b), reveals three distinct Gaussian peaks labeled as a, b, and c. The low binding energy peak Oa, located at 530.89 eV, corresponds to O2− ions participating in Zn-O and/or Mn-O bonds. The intermediate binding energy peak Ob, observed at 532.53 eV, is associated with Zn-OH species. Additionally, the high binding energy peak Oc, appearing at 533.75 eV, is attributed to chemisorbed oxygen contributing to the formation of H2O, CO2, and/or oxygen defects such as Vo or Oi species, as previously reported [61, 62]. Upon the removal of contaminated layers containing OH, COx, and H2O species through Ar ion sputtering and high-temperature annealing treatment, the Gaussian fit of the O 1s peak reveals only two sub-peaks, Oa and Ob, positioned at 530.37 eV and 531.48 eV, respectively [10, 14], showing considerable reduction in Vo or/and Zni concentrations.

Photoluminescence results

In this study, photoluminescence (PL) spectroscopy is utilized to investigate the impact of Mn incorporation and UHV treatment on the presence of defects. These defects subsequently alter the band gap and luminescence characteristics of zinc oxide [10, 19, 63].

Figure 6 shows the spectrum (i) related to pure ZnO. It exhibits two distinct emission bands. Firstly, a sharp ultraviolet (UV) emission at 3.27 eV (λ = 380 nm) is attributed to near band-edge emission (NBE) induced by free excitonic (Fx) radiative recombination. This emission is highly dependent on the crystalline quality [64]. Secondly, a broad visible emission ranging from 2.8 eV to 1.4 eV (λ = 442.86–885.70 nm), with a center at 2.26 eV (λ = 550 nm), is associated with intrinsic structural defects present in the optical band gap of ZnO [65]. These defects include zinc vacancies (VZn), oxygen vacancies (Vo), zinc interstitials (Zni), oxygen interstitials (Oi), and oxygen antisites (OZn).

Fig. 6.

PL spectra of pure ZnO (i), untreated ZnMnO (ii), and treated ZnMnO (iii)

Furthermore, Figure 6 (ii) illustrates three significant changes resulting from the incorporation of 4% Mn into the ZnO matrix: (1) The UV emission peak becomes more intense and broader compared to the undoped film. (2) The visible emission peak decreases. (3) There is a notable increase in the NIR (near infrared) emission peak at 1.65 eV. It is worth mentioning that the broadening of the NBE peak observed in spectrum (ii) aligns with the XRD findings, which indicate a broader FWHM for Mn-doped ZnO due to its inferior crystalline nature. The analysis of the PL spectrum further confirms the successful incorporation of Mn into the lattice.

Moreover, upon comparing spectra (ii) and (iii), it is evident that the UHV treatment provides several changes. Firstly, it facilitates the removal of the primary visible peak by completely eliminating the structural defects associated with Oi states and Vo, as discussed in the XPS results of O 1s. Secondly, it leads to the broadening of the UV emission peak, accompanied by a slight redshift of 0.013 eV. This shift is attributed to the band gap narrowing resulting from the strong interactions between the “s-p” electrons of Zn+2 ions and the “d” electrons of Mn+2 ions. Thirdly, it effectively eliminates the shoulder of the near-infrared emission peak caused by the presence of hydrogen impurity (Hi) at 1.65 eV (750 nm) [66]. Based on the aforementioned observations, we have established a connection between ZnO doping with Mn, UHV treatment, and their influence on the structural and photoelectronic properties. Furthermore, this study highlights the informative relationship between the XRD and XPS analysis techniques and the PL measurements.

For more clarification of modifications due to Mn doping and UHV treatment, Figure 7 gives a Gaussian deconvolution of the PL spectra corresponding to pure ZnO (i), untreated Zn-MnO(ii), and treated ZnMnO(iii), accompanied by schematic band diagrams with the approximate positions of defect levels and luminescence mechanisms as suggested by other authors.

Fig. 7.

Gaussian fitting of PL spectra accompanied by schematic band diagram corresponding to pure ZnO (i), untreated ZnMnO (ii), and treated ZnMnO (iii)

The NBE peak has been reported to include three sub-peaks. The first one at 3.275 eV (378 nm) concerns the free excitons (FX) from the state energy of 0.06 eV under the bottom of the conduction band minimum to the valence band maximum (FXVBM) [67]. The second peak appears at 3.16 eV (392 nm), corresponding to the transitions from Zni states located at 0.15 eV under the bottom of the conduction band minimum (CBM) to the valence band maximum (ZniVBM) [68]. The last peak is much less intense than the previous ones, at 3.03 eV (409 nm), corresponding to the transition from CBM to VZn-bound states located 0.27 eV above VBM (CBM → VZn) [69]. It should be noted that this peak becomes more obvious with Mn-doped ZnO.

The broad visible emission was found to be fitted with four Gaussian subpeaks positioned at 2.50, 2.28, 2.05, and 1.80 eV for pure ZnO. Figure 7 shows a noticeable shift in the position peaks and clear variations in the relative intensities when Mn substitutes Zn in the ZnO film. The blue luminescence centered at 2.50 eV (496 nm) originates from the transition of electrons between VZn and the CBM (CBM → VZn) [70] and disappears completely for ZnMnO. The green and yellow luminescence centered at 2.28 (544) and 2.05 eV (605 nm), respectively, originates from the transition of electrons between CBM and the Oi (CBM → Oi) [71] and decreases notably for ZnMnO before disappearing completely after the UHV treatment. Finally, the transition from the state extended-Zni at 0.6 eV from the CBM to the state VO at 0.9 eV above the VBM (ex-Zni→ VO) is associated with the orange luminescence centered at 1.80 eV (689 nm) [72].

The near-IR luminescence shows two sub-peaks at 1.65 and 1.5 eV, respectively; the peak at 1.5 eV due to the hydrogen impurity (Hi) (Hi → VO) [73] is completely erased by the UHV treatment. This implies that the hydrogen impurity (Hi) has been effectively removed; however, the other peak caused by the VO defect is located in the middle of the bandgap (CBM → VO, VO → VBM) [71, 74]. Controlled Mn doping of ZnO, according to the PL study, can improve its optical characteristics for uses beyond spintronics, such as optoelectronics.

Conclusion

The objective of this paper was to investigate the effects of low-level Mn substitution and the treatment process on the chemical, structural, morphological, and optoelectronic properties of ZnO and ZnMnO thin films with 3.6 at% Mn. The thin films were synthesized using the ultrasonic pyrolysis technique on Si (100) substrates, following a standardized deposition procedure. Analysis of the AFM images revealed significant changes in surface morphology attributed to the incorporation of Mn, leading to noticeable decreases in grain size, and increases in height,and average roughness. XRD analysis confirmed the presence of the würtzite crystal structure of ZnO, with the dominant (002) plane governing the growth process across all films. Furthermore, slight reductions and shifts towards lower diffraction angles in the XRD peaks, accompanied by minor variations in lattice parameters and an increase in micro strain, provided compelling evidence supporting the successful integration of Mn2+ ions through Zn2+ substitution.

To gain deeper insights into the optical properties, XPS and PL spectroscopy were employed. The observed near band-edge emission (NBE) at 3.27 eV within the ultraviolet (UV) range, suppression of the yellow-orange, green emission (~2.6-1.8 eV), and the emergence of a distinct emission peak at 1.65 eV were identified as characteristic signatures associated with the presence of Mn atoms. Therefore, the above-mentioned results, in combination with the reduced grain size, improved micro-stain deformations, and decreased ionization potential, facilitated the accurate reproduction of the diagrams representing both ZnO and ZnMnO compounds. This is both before and after UHV treatment and annealing. It is worth noting that the structural defects responsible for visible emissions were effectively eliminated in ZnMnO, resulting in well-defined emissions confined solely to the nearinfrared and ultraviolet ranges.

eISSN:
2083-134X
Język:
Angielski
Częstotliwość wydawania:
4 razy w roku
Dziedziny czasopisma:
Materials Sciences, other, Nanomaterials, Functional and Smart Materials, Materials Characterization and Properties