Next Article in Journal
Breast Cancer Vaccine Containing a Novel Toll-like Receptor 7 Agonist and an Aluminum Adjuvant Exerts Antitumor Effects
Next Article in Special Issue
Myxococcus xanthus Encapsulin as a Promising Platform for Intracellular Protein Delivery
Previous Article in Journal
Non-Metal-Doped Porous Carbon Nitride Nanostructures for Photocatalytic Green Hydrogen Production
Previous Article in Special Issue
Pt-Based Nanostructures for Electrochemical Oxidation of CO: Unveiling the Effect of Shapes and Electrolytes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Precise Sn-Doping Modulation for Optimizing CdWO4 Nanorod Photoluminescence

1
Department of Physics, National Dong Hwa University, Hualien 97401, Taiwan
2
National Synchrotron Radiation Research Center, Hsinchu 30076, Taiwan
3
Department of Chemistry, University College of Science, Osmania University, Hyderabad 500007, Telangana, India
4
Department of Material Science and Engineering, National Yang Ming Chiao Tung University, Hsinchu 30010, Taiwan
5
Department of Electrical Engineering, National Tsing Hua University, Hsinchu 30013, Taiwan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(23), 15123; https://doi.org/10.3390/ijms232315123
Submission received: 4 November 2022 / Revised: 25 November 2022 / Accepted: 29 November 2022 / Published: 1 December 2022
(This article belongs to the Special Issue Recent Advances in Nanomaterials Science)

Abstract

:
The cadmium tungstate rods have been given much attention due to their potential for usage in numerous luminescent applications. We have prepared single crystalline Sn-doped Cd1−xSnxWO4 (where x = 0, 1, 3, and 5%) nanorods (NRDs) and characterized them using refined X-ray diffraction and TEM analysis, revealing a monoclinic phase and a crystallite size that decreased from 62 to 38 nm as Sn concentration increased. Precise Sn doping modulation in CdWO4 NRDs causes surface recombination of electrons and holes, which causes the PL intensity to decrease as the Sn content rises. The chromaticity diagram shows that an increase in the Sn content caused a change in the emission color from sky blue to light green, which was attributed to the increased defect density. The photoluminescence time decay curve of all samples fit well with double-order exponential decay, and the average decay lifetime was found to be 1.11, 0.93, and 1.16 ns for Cd1−xSnxWO4, x = 0, 1, and 5%, respectively. This work provides an understanding of the behavior of Sn-doped CdWO4 NRDs during electron transitions and the physical nature of emission that could be used in bio-imaging, light sources, displays, and other applications.

1. Introduction

Metal tungstates have attracted much attention recently due to their intriguing physical and chemical characteristics and future uses in various sectors such as photochromism, magnetic devices, photonics, photoluminescence, sensors, scintillators, etc. [1,2,3,4]. The tungstate compounds have several advantages, including non-hygroscopicity, thermal and chemical stability, and non-toxicity [5,6]. Self-activating phosphor CdWO4, one of the tungstate compounds, can crystallize in either a scheelite tetra agonal structure or a monoclinic wolframite structure. CdWO4 is a good candidate host material for phosphors [7]. The intrinsic emission of CdWO4 nanorods (NRDs) appears as a broad band in the blue and green regions. This is primarily because of the charge transfer transition between the oxygen ion’s 2p and tungsten ion’s 5d orbitals and self-trapped excitons at the WO66− complex [8]. Because of characteristics such as a high X-ray absorption coefficient, less radiation damage, a high average refractive index, and a broadband gap, the monoclinic structure is a multifunctional material [2,9,10,11]. Oxytungstate with a second-group metal ion is an appealing substance and has drawn tremendous study interest because of its fascinating luminescence and structural characteristics [12,13].
Numerous studies on the luminescence characteristics of doped CdWO4 have been published to date [14,15,16,17,18,19,20,21], including those on surface flaws and their effects on the structural and photoluminescence characteristics of Eu3+-doped CdWO4 nanocrystals [9], photoluminescence studies and a core shell model approach for rare earth doped CdWO4 nano phosphors [22], the fluorescence properties for divalent and trivalent europium ions in CdWO4 single crystals produced by the Bridgman method [7], the photoluminescence properties of CdWO4 doped with Sm3+ concentration and annealing effects [23], and photoluminescence characteristics of Ce3+ substituted CdWO4 nano phosphors [24]. Recently, a variety of Cu- and Bi-doped CdWO4 NRDs synthesized by our research team using a hydrothermal approach was successfully used to demonstrate improved photocatalytic performance when exposed to sunlight irradiation [25,26]. As a result, the absorption edge moves into the visible spectrum and captures the photogenerated electrons, reducing the rate of electron-hole recombination. Furthermore, Sn-doping has been shown to improve photoluminescence characteristics compared to the pristine compound [27,28,29,30,31]. To the best of the author’s knowledge, Sn-doped CdWO4 has not yet been reported, based on observations of prior studies.
Therefore, in the present study, Sn-doped CdWO4 NRDs were prepared using the hydrothermal method. To understand the structural properties and crystalline phase of the samples were studied by using synchrotron-based powder X-ray diffraction (PXRD), microstructural properties of all samples were analyzed by using field-emission scanning electron microscopy (FE-SEM) and field-emission transmission electron microscopy (FE-TEM). Vibrational bands were studied by using Raman spectra, and we have also investigated the effect of Sn on photoluminescence properties of CdWO4 NRDs at low-temperature and at room temperature. Our in-depth material characterization correlates microstructural, structural, Raman, and photoluminescence properties, confirming Sn doping into the CdWO4.

2. Results and Discussion

Morphological characterization. FE-SEM was used to examine the morphologies of pure and Sn-doped CdWO4; the high and low magnification images are shown in Figure 1a–d and Figure S1 (Supporting Information), respectively. The FE-SEM images of CdWO4 and Sn-doped CdWO4 clearly show the rod-like morphology, which is consistent with the literature [18]. The mean diameter d SEM of rods is obtained by fitting a log-normal distribution function:
f ( d ) = 1 d σ 2 π exp [ ( lnd d SEM ) 2 2 σ 2 ]
to the histogram obtained from the FE-SEM images, as shown in Figure 1e–h, where the value of σ represents a standard deviation of the fitted function [18]. The nanorod length histogram distributions were also shown in Figure S2 (Supporting Information). All obtained fitted parameters were summarized in Table S1 (Supporting Information). An increase in Sn content from 0 to 5% causes an increase in aggregation, a decrease in diameter from 65 to 34 nm, and a decrease in length from 317 to 115 nm, resulting in an astonishing ~31% reduction in the aspect ratio, as shown in Figure S3 (Supporting Information). According to the above analysis of the FE-SEM images, the reduction in CdWO4 crystals is caused by Sn doping. The above findings are in line with studies that have been published on Cu-, Fe-, La-, and Eu-doped CdWO4 and Cd-doped ZnWO4 [1,28,32,33,34,35,36]. Therefore, doping changes the CdWO4 lattice structure regardless of the type of dopant used, and it may even prevent crystal formation by changing the surface charge [37]. According to the SEM pictures, the mean size reduction at the increased Sn content causes high surface energy, which causes aggregation (Figure S1, Supporting Information). These results show that aggregation occurred in the reaction mixture and that Sn content can effectively control it. Figure S4a–d (Supporting Information) shows the EDS analysis of Sn-free and Sn-doped CdWO4, respectively. According to the EDS analysis, there are no impurities in the synthesized samples. Furthermore, the energy dispersive absorption X-ray spectroscopy confirmed the previously reported uniform distribution of all constituent elements [26].
Figure 2a–c depict the TEM images of 1% Sn, 3% Sn, and 5% Sn doping CdWO4 NRDs, revealing a decreased diameter of the nanorods with increased Sn concentration. Figure 2d shows the high-resolution TEM image of pure 1% Sn-doped CdWO4 NRDs, yielding lattice spacing of 0.3078 and 0.3049 nm for (−1 1 1) and (1 1 1) planes, respectively. Similarly, 3% and 5% Sn-doped CdWO4 NRDs yield the lattice spacing [0.3184, 0.3171 nm] and [0.3259, 0.3200 nm], corresponding to the [(−1 1 1), (1 1 1)] planes, respectively (Figure 2e,f) [38]. The mean value of d-spacing obtained from high-resolution images is shown in Figure S5 (Supporting Information). The obtained lattice spacing values corresponding to the (1 1 1) plane well agrees with the reported value elsewhere [38]. The 1, 3, and 5% Sn-doped CdWO4 oriented along the [0 1 −1] zone axis and the SAED patterns of 1%, 3%, and 5% Sn-doped CdWO4 NRDs confirm the monoclinic structure (Figure 2g–i). The Fourier transform generated from HRTEM images (inset of Figure 2d–f) are in agreement with the SAED pattern acquired from a single NRD. The SAED patterns of 1%, 3%, and 5% Sn-doped CdWO4 NRDs confirm the monoclinic structure (Figure 2g–i). The TEM analysis results agree with the XRD results and confirm that the 1%, 3%, and 5% Sn-doped CdWO4 NRDs are purely monoclinic crystal structures.
Structural Characterizations. The refined PXRD patterns of Cd1−xSnxO4 with x = 0 (CdWO4), 1, 3, and 5% are shown in Figure 3a–d. The monoclinic structure of the JCPDS-14-0676 file [39], which has a space group of P2/C, is matched by the diffraction peaks seen in the XRD patterns, confirming the single phase. The Rietveld refinement approach was used to confirm the phase purity firmly. The refined diffraction patterns are displayed in Figure 3a–d, where there is a slight variation, nearly zero, between the calculated (red) and absolute intensities (black) [40]. The dopant Sn ions may have been successfully incorporated into the lattice because there are no additional peaks in the pattern and any other impurity phases [41]. Table S2 (Supporting Information) contains an occupancy of Sn-doped CdWO4 derived by Rietveld refinement. Using the Fullprof software, the lattice parameters and volumes have been determined and summarized in Table 1. Because of the difference in ionic radius between Sn2+ (0.69 Å) and Cd2+ ions (0.97 Å), the lattice parameters varied a small amount as Sn concentration increased (Figure 3e) [42]. Figure 3f shows a small clear shift in diffraction peaks (−1 1 1) with increasing Sn concentrations. This shifting could be caused by differences in the nuclear radius of the Sn and Cd atoms, which results in a small variation in lattice constant value, as shown in Table 1. The Sm3+-doped CdWO4 has exhibited similar behavior, according to recent reports [23]. Unit cell volume decreases with increases in Sn2+ concentration due to the ionic radius of Cd2+ ions (0.97 Å) and ions greater than the Sn2+ (0.69 Å) ions, as can be seen in Table 1. Figure S6a–d (Supporting Information) displays the fitting of the Gaussian model to the (−1 1 1) diffraction peak [43]. It can be seen that the entire FWHM value of the fitting peaks increased with rising Sn concentrations. The Williamson–Hall (WH) plot method (Equation (2)) was used to calculate the crystallite size and microstrain [44,45]:
ε = ε size + ε strain = 1 cos θ ( k λ d XRD + 4 η sin θ )
where k is a constant, β(hkl) is the full width half maximum (FWHM), λ is the wavelength of the incident X-ray, and ε represents FWHM. The WH plot is displayed in Figure S7 (Supporting Information), and the obtained mean crystallite size is shown in Table 1. The calculated crystallite size from the WH method agrees with the reported value for Cu-CdWO4 NRDs [32]. As shown in Figure 3g, the observed microstrain (%) value increases with increasing Sn content, indicating an increase in [MO6] (M = Cd and Sn) cluster deformation brought on by the incorporation of Sn2+ into the CdWO4 [34]. By increasing Sn concentration the diffraction peak may broaden, which may be associated with a reduction in crystallite size and an increase in strain, showing that the Sn ions have successfully entered the CdWO4 lattice [46]. This outcome makes it evident that the dopant Sn ions were successfully incorporated into the lattice.
Raman scattering. Group theory simulations show that tungstates with monoclinic or triclinic structures exhibit 36 distinct Raman vibrational modes (8Ag + 10Bg + 8Au + 10Bu), 18 of which are anticipated to be active modes (8Ag + 10Bg). In this study, Raman spectra were collected at wavenumber ranging from 80 to 1000 cm−1. The Raman spectra of Cd1−xSnxWO4 NRDs (x = 0, 1, 3, and 5%) are displayed in Figure 4a. All samples had a vital vibration peak at 898 cm−1. Several weak vibrations were at 772, 707, 687, 548, 516, 387, 352, 307, 269, 248 230, 178, 148, 134, 117, and 99 cm−1 [32,47]. In this study, 17 active Raman modes were observed, which is completely in line with the findings reported by Ross-Medgaarden and Wachs [36]. According to previously reported works, one more Raman mode will be observed at 77cm−1. Thus, in the current work, we observed only 17 active Raman modes because we studied the Raman spectra from 80 to 1000 cm−1, confirming the monoclinic structure in our samples [32,36]. In general, external (600 cm−1) and internal (>600 cm−1) vibrational modes found in the Raman spectra of tungstates may be divided into two groups, according to the reported literature [32,48]. The vibrational bands in the 150–300 cm−1 region were attributed to the Cd–O stretching modes. The symmetric stretching of [CdO6] octahedra is responsible for the vibrational band at 307 cm−1. The vibration bands in the 500–600 cm−1 region are attributed to symmetric W–O–W stretching modes. The vibrational bands 687 and 772 cm−1 modes involve [WO6] octahedra motions against the Cd2+. Normal WO vibration of [WO6] octahedra corresponds to the 387 and 898 cm−1 modes. The line width broadening concerning Sn content is observed in all the peaks. A Raman shift of ~2 cm−1 was observed when Sn doping concentration increased from 0 to 5% (230 to 228 cm−1 and 898 to 896 cm−1 vibration modes, as shown in Table S3, Supporting Information). With increasing Sn2+ concentration, a Raman shift was observed, indicating lattice distortion in the Sn2+-doped CdWO4 nanorod [49]. The most intense Raman band observed around Ag*~898 cm−1 in undoped and Sn-doped samples characterized by the WO6 symmetric stretching vibration (Figure 4b). A similar Raman band has also been reported around 891 cm−1 from NiWO4 nanopowders [50]. The integrated peak intensity decreases with the increase in Sn concentration, as shown in Figure 4c. It should also be noted that a new Raman band at 926 cm−1 was discovered from a 1% Sn sample. The intensity of the new peak decreases as Sn concentrations increase. A similar Raman band was also reported around 932 cm−1 from 1% Sn-doped CdWO4 NRDs ascribed to surface disorder layers [32]. As a result, there is evidence that the Sn dopant caused structural inhomogeneity.
Photoluminescence studies. The emission spectrum of Sn-free CdWO4 and Sn-doped CdWO4 NRDs were obtained at 300 K (red) and 15 K (black), respectively, as shown in Figure 5a–c. The A 1 1 T 3 1 transitions in the [WO6] complex are the only ones responsible for the PL emission in CdWO4 [18]. The excitation spectra of Sn-doped CdWO4 consist of a broad band in the wavelength ranges of 100 to 315 nm, with the most significant intensities at approximately 267, 265, and 265 nm for 0, 1, and 5%, respectively, at 15 K. This broadband can be attributed to the overlap of ligand to metal charge transfer transitions between oxygen to Sn2+ and the filled 2p orbital of oxygen and the empty 5d orbital of W6+ within the WO66− complex [1,46,51]. The most vigorous UV emission (267 nm) was found at x = 0 and gradually weakened with Sn content doping. Regarding the visible region’s emission, Sn-doped CdWO4 nanocrystals showed the weakest emission. One crucial observation is that the UV and visible emissions’ relative PL intensities decreased with the increasing Sn dopant. This leads us to the conclusion that the electron relaxation time of the conduction band is prolonged due to the charge transfer between the conduction band and the impurity levels [52]. CdWO4 NRDs are excited by 267 nm UV light, which results in just one strong blue-green emission spectrum in the range of 350 to 700 nm, with a peak centered at 484 nm near the values that have been reported [18]. The charge transfer transition inside the WO66− complex is also responsible for the broad band of intense wavelengths that was also seen toward the higher wavelength regions in the ranges between 330 and 600 nm, in addition to these peaks around 484 nm attributed to (6H5/2-4F7/2), which agree with the reported values [32,53,54]. The asymmetry is due to the existence of energetically distinct O(1) and O(2) sites such that the decay process of WO 6 6 groups can occur via two slightly different electronic transitions [32]. The emission peak profile did not significantly change as Sn concentrations varied; but, nevertheless, Sn concentrations considerably impacted photoluminescence intensity. There is an efficient energy transfer from the host lattice to dopant Sn2+ because there is significant broadband owing to O2-W6+/Sn2+ as contrasting to the f-f transitions of Sn [51]. Additionally, it was found that the emission intensity reduced as temperature increased from 15 to 300 K, possibly because rising temperatures increase the elimination of luminescence. Photoluminescence spectra also show that broadening of emission and excitation peaks decreases the temperature from 300 to 15 K. There was no shift in emission or excitation peak with increasing temperature. After doping Sn, their interactions grow, eventually leading to cross-relaxation, which causes the emission intensity to drop when the mean distance between them is less than a critical value. More photogenerated electrons and holes can participate in the oxidation and reduction reactions when the PL intensity is lower. Similar characteristics preserved in the Sn-doped samples with a more pronounced quench indicate a decrease in electron and hole recombination rate with increasing Sn content. With Sn doping, the photocatalytic performance can be enhanced from an application standpoint. The temperature-dependent PL measurement was carried out to look at the optical characteristics of Sn-doped CdWO4 below room temperature. The 2D mapping of the temperature-dependent PL spectra with intensity is shown in Figure 5d–f. There are emission peaks at low temperatures. Internal quantum efficiency (IQE), which is measured as the ratio of integrated PL intensity between 15 and 300 K and assumes an internal efficiency of 100% at a low temperature of 15 K, was found to be 64.7%, 55.9%, and 48.9% for 0, 1%, and 5% Sn, respectively. This shows a relative decrease in the IQE of 5% Sn-doped CdWO4 NRDs of about 49% compared to pure CdWO4 NRDs. The PL intensity, however, monotonically decreases in Sn CdWO4 NRDs. Sn doping in CdWO4 causes surface recombination of electrons and holes, which causes the PL intensity to decrease as Sn content rises. The recombination and transfer of photogenerated electron–hole pairs also can be caused by a decrease in PL intensity as Sn content increases. It is generally acknowledged that, from an application standpoint, Sn-doped CdWO4 can be stronger than the photocatalytic performance [32]. The CIE 1931 chromaticity coordinates (X, Y) were next acquired to assess the performance of the Sn-doped CdWO4 NRDs on color luminescent emission. Figure 6a depicts the resulting chromaticity diagram. It demonstrates that increased Sn content caused a change in the emission color from sky blue to light green, which was attributed to the increased defect density.
The temperature-dependent photoluminescence excitation (PLE) spectra of undoped and Sn-doped CdWO4 for the most intense peak was at the direct bandgap at 300 and 15 K temperatures, as shown in Figure 6b–d [55,56]. For CdWO4, the maximum intensity of the PLE band is located at 4.65 eV. A similar profile was obtained for the 5% Sn-doped CdWO4 sample with a band located at 4.69 eV. The temperature-dependent photoluminescence emission spectra of undoped and Sn-doped CdWO4 for the emission peak are located at 2.56 eV (484 nm), as shown in Figure S8 (Supporting Information). Electron-phonon coupling could explain the temperature dependence of emission peaks [57]. Due to the thermal quenching caused by nonradiative recombination, the emission intensity drops as the temperature rises. Figure 6e reveals the photoluminescence mechanism of Cd1−xSnxWO4 NRDs with various Sn concentrations. At low temperatures and room temperature, the primary emission peak for both doped and undoped samples was located at approximately 4.65 eV for CdWO4, which corresponds to the direct energy gap. After doping the 5% Sn to the CdWO4, the energy gap increases to 4.69 eV. The band gap obtained from the pristine CdWO4 NRDs matched what was reported elsewhere [57,58]. It was found that increased Sn doping caused an increase in energy gap values, which indicates that the doping caused the absorption edge to shift toward higher energies. This increase can be explained by the Bureshtain–Moss Shift phenomenon, in which the levels close to the conduction beam are dense with electrons which requires more energy for the electrons to travel, giving the impression that the energy gap is expanding. Similar behavior was also observed in the reported Sn-doped cadmium oxide [59].
The temperature-dependent integrated PL intensity has been studied at temperatures ranging from 15 to 300 K to learn more about the origin of PL emission. Figure 7a–c depicts the variance in Integrated PL intensity as a function of inverse temperature. The temperature dependence of PL intensity has been significantly changed by Sn-doping. A similar response was observed in Cu-doped BaWO4, NiWO4, and CdWO4, allocated to dopant-induced intermediate energy levels [27,28,32]. For further brief analysis, the integrated PL intensity as a function of inverse temperature for CdWO4 (0% Sn) has been fitted using the Arrhenius equation, as shown below [60]:
I ( T ) = I 0 [ 1 + A exp ( E 1 / k B T ) ]
where I 0 is the integrated PL intensity at temperature 15 K, I(T) is the integrated PL intensity at temperature T, and k B is Boltzmann’s constant. We propose that the quenching mechanism of PL intensities for CdWO4 NRDs is connected to two non-radiative recombination processes based on fitting data. So, the experimental data (Figure 7b,c) is fit by using a modified Arrhenius equation, as shown below:
I ( T ) = I 0 [ 1 + A exp ( E 1 / k B T ) + B exp ( E 2 / k B T ) ]
where E 1 and E 2 implies the activation energies of the non radiative recombination for thermal quenching of PL intensities at a higher and lower temperature. A and B are the constants that represent the ratio of τ R (radiative lifetime of carriers) and τ 0 (non-radiative lifetime of carriers). Table S4 provides an overview of the best-fit parameters (Supporting Information). From the best fitting of the data using the Arrhenius equation, activation energy E 1 for PL quenching is found to be 145.72 meV for CdWO4. E 1 emerged as a potential barrier between the localized potential minima and the non-radiative centers in the numerous quantum wells in this case [32]. As observed in the best-fitted modified Arrhenius equation, a decrease in the fitted value of E 1 from 192.73 to 108.9 meV with Sn concentration increases, ranging from 1 to 5%, suggesting that carriers can easily escape from confined states at higher temperatures. The constant A increases from 154 to 1321, with an increase in Sn from 0 to 5% due to a decrease in τ 0 . The density of non-radiative recombination centers decreases as Sn content increases, which causes a decrease in A , consistent with previous earlier reports [32,60]. The low E 2 and B values obtained for the 1 and 5% Sn-doped CdWO4 NRDs consist of quenching at low temperatures.
Figure 8a–c shows the photoluminescence time decay curves of Cd1−xSnxWO4 NRDs (x = 0, 1, and 5%). All the curves fit well with double-order exponential decay, as shown below [61]:
I ( t ) = I 0 + A 1 exp ( t τ 1 ) + A 2 exp ( t τ 2 )
where I represents the photoluminescence intensity, I0 is the photoluminescence intensity at 0, and A1 and A2 are constants. τ1 and τ2 represent the decay times for the exponential components. Average decay lifetimes can be calculated as follows [61]:
I ( t ) = ( A 1 τ 1 2 + A 2 τ 2 2 ) / ( A 1 τ 1 + A 2 τ 2 )
Using Equation (6), it is possible to calculate the decay times of Cd1−xSnxWO4 NRDs (x = 0, 1, and 5%); they were found to be 1.11, 0.93, and 1.16 ns, respectively. This result shows lifetimes varied with increasing Sn2+ concentration. It can strongly prove the energy transfer process existing from Sn2+ ions [61].

3. Methods and Materials

The Cd1−xSnxWO4 (x = 0, 1, 3, and 5 %) NRDs were prepared using in situ hydrothermal methods [26]. Scheme 1 shows the schematic representation of the synthesis process of Cd1−xSnxWO4 (x = 0, 1, 3, and 5%) samples using in situ hydrothermal methods. In this method, the stoichiometric quantity of tungstic acid sodium salt dihydrate (Na2WO4·2H2O), tin chloride dihydrate (SnCl2·2H2O), and cadmium chloride hydrate (CdCl2·H2O) were dissolved separately in ddH2O (double-distilled water). The tungstic acid sodium salt dihydrate solution was then mixed with cadmium chloride hydrate, followed by the dropwise addition of desired tin chloride dihydrate while constantly stirring. The resulting solution, which had different concentrations of Sn, was stirred continuously for an hour before being transferred individually to a 100 mL Teflon-lined stainless-steel autoclave, which made up around 80% of the total capacity. Without shaking or stirring during the heating period, all of the autoclaves with various Sn content solutions were kept at 180 °C for 12 h at once, after which they were allowed to cool naturally to room temperature. The precipitates were collected, filtered, and washed five times with ddH2O before being dried at 80 °C overnight in a hot air oven. At first look, the samples’ peculiar color suggested that an increase in the Sn content may have impacted the structural and optical characteristics of CdWO4 [62]. Keep in mind that both before and after the reaction, the pH level is near 7. Samples were analyzed morphologically using field-emission scanning electron microscopy (FESEM, JEOL JSM-7000F). TEM characterization was carried out through field-emission transmission electron microscopy (FE-TEM, JEM-F200). The samples were structurally characterized using PXRD at the National Synchrotron Radiation Research Center (NSRRC) Taiwan light source (TLS) 01C2 beamline in Hsinchu, Taiwan (λ = 0.77491 Å). Raman spectra of all the samples were studied using a micro-Raman spectrometer coupled with a microscope (Leica, Wetzlar, Germany) and a 532 nm wavelength laser. The TLS 03A1 beamline of the NSRRC in Taiwan was used to perform the temperature-dependent PL observations. Two-photon fluorescence lifetime imaging (TP-FLIM) has been used to visualize Sn-doped CdWO4 NRDs. This measurement was performed with the objective UPlanFLN 40/0.75 (Olympus, Shinjuku, Tokyo, Japan). TP-FLIM measurements and the lifetime imaging of Sn-doped CdWO4 NRDs were performed using the excitation of a Ti-sapphire laser (Chameleon Ultra-II; pulse duration 140 fs; repetition rate 80 MHz). The wavelength of excitation used was 760 nm. The excitation light power was controlled by a neutral density filter and was measured before the microscope objective from 15 to 35 mW. The emission was collected in the range of 450 to 650 nm with a single photon counting system Pico Harp 300 (Pico Quant, Germany) and a thermoelectrically cooled PMT conjugated with microscope Olympus IX 71. Fluorescence lifetime images of Sn-doped CdWO4 NRDs were obtained and analyzed using the SymPho Time software package, Version 5.2.4.0 (Pico Quant, Germany).

4. Conclusions

In this work, the monoclinic structure of Cd1−xSnxWO4 (x = 0, 1, 3, and 5%) NRDs has been synthesized using in situ hydrothermal methods. The FE-SEM and TEM images of pure CdWO4 and Sn-doped CdWO4 clearly show the nanorod-like morphology. An increase in Sn content from 0 to 5% causes an increase in aggregation, a decrease in diameter from 65 to 34 nm, and a decrease in length from 317 to 115 nm, resulting in an astonishing ~31% reduction in aspect ratio. According to the EDS analysis, there are no impurities in the synthesized samples. SAED patterns confirm the formation of a monoclinic structure. The Rietveld refined XRD patterns demonstrate the single phase without showing any impurities. Increasing Sn concentration may broaden the diffraction peak, associated with a reduction in crystallite size and increased strain. A Raman shift was observed with increased Sn2+ concentration, indicating lattice distortion in the Sn2+-doped CdWO4 NRDs. The emission peak profile did not change significantly as Sn concentrations changed; however, Sn concentrations greatly impacted photoluminescence intensity. The chromaticity diagram shows that an increase in the Sn content caused a change in the emission color from sky blue to light green, which was attributed to the increased defect density. The intensity of the emission peak decreases with increasing temperature due to thermal quenching by nonradiative recombination. The average lifetime of Sn2+-doped CdWO4 NRDs was obtained from the fitted photoluminescence time decay curves and confirmed that the average lifetimes varied with increasing Sn2+ concentration. It can strongly prove the energy transfer process existing between Sn2+ and Cd2+ ions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232315123/s1.

Author Contributions

K.M., A.C.G. and S.Y.W. wrote, conceived, and designed the experiments. M.-K.H., T.-E.H., H.-H.C. and T.-Y.L. contributed to the data analysis. B.V.K. and P.M.R. synthesized the samples. Y.-H.W., B.-H.L., T.S.C., A.K. and C.-L.C. contributed to the PL and Raman measurements. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministry of Science and Technology (MOST) grant number MOST-111-2112-M-259-013.

Informed Consent Statement

Not applicable.

Acknowledgments

We would like to thank the Ministry of Science and Technology (MOST) of the Republic of China for the financial support through grant numbers MOST-107-2112-M-259-005-MY3, MOST-109-2811-M-259-506, MOST-110-2112-M-259-006, MOST-110-2112-M-259-009, MOST-111-2112-M-259-014, MOST-111-2811-M-259-009, and MOST-111-2112-M-259-013.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yue, D.; Li, Q.F.; Lu, W.; Wang, Q.; Wang, M.N.; Li, C.Y.; Jin, L.; Shi, Y.R.; Wang, Z.L.; Hao, J.H. Multi-color luminescence of uniform CdWO4 nanorods through Eu3+ ion doping. J. Mater. Chem. C 2015, 3, 2865–2871. [Google Scholar] [CrossRef]
  2. Li, D.; Bai, X.J.; Xu, J.; Ma, X.G.; Zhu, Y.F. Synthesis of CdWO4 nanorods and investigation of the photocatalytic activity. Phys. Chem. Chem. Phys. 2014, 16, 212–218. [Google Scholar] [CrossRef]
  3. Shad, N.A.; Sajid, M.M.; Amin, N.; Javed, Y.; Akhtar, K.; Ahmad, G.; Hassan, S.; Ikram, M. Photocatalytic degradation performance of cadmium tungstate (CdWO4) nanosheets-assembly and their hydrogen storage features. Ceram. Int. 2019, 45, 19015–19021. [Google Scholar] [CrossRef]
  4. Mikhailik, V.B.; Kraus, H.; Miller, G.; Mykhaylyk, M.S.; Wahl, D. Luminescence of CaWO4, CaMoO4, and ZnWO4 scintillating crystals under different excitations. J. Appl. Phys. 2005, 97, 083523. [Google Scholar] [CrossRef] [Green Version]
  5. Dai, Q.L.; Song, H.W.; Pan, G.H.; Bai, X.; Zhang, H.; Qin, R.F.; Hu, L.Y.; Zhao, H.F.; Lu, S.Z.; Ren, X.G. Surface defects and their influence on structural and photoluminescence properties of CdWO4: Eu3+ nanocrystals. J. Appl. Phys. 2007, 102, 054311. [Google Scholar] [CrossRef]
  6. Zhai, Y.; Wang, M.; Zhao, Q.; Yu, J.; Li, X. Fabrication and Luminescent properties of ZnWO4:Eu3+, Dy3+ white light-emitting phosphors. J. Lumin. 2016, 172, 161. [Google Scholar] [CrossRef]
  7. Hao-Yang, H.; Hai-Ping, X.; Jian-Xu, H.; Yue-Pin, Z.; Hao-Chuan, J.; Hong-Bing, C. Fluorescence properties of divalent and trivalent europium ions in CdWO4 single crystals grown by the Bridgman method. Chin. Phys. B 2013, 22, 027804. [Google Scholar]
  8. Hojamberdiev, M.; Kanakala, R.; Ruzimuradov, O.; Yan, Y.; Zhu, G.; Xu, Y. Besom-like CdWO4 structures composed of single-crystalline nanorods grown under a simple hydrothermal process in ultra-wide pH range. Opt. Mater. 2012, 34, 1954. [Google Scholar] [CrossRef]
  9. Zhang, J.; Zhao, T.; Wang, B.; Li, L.; Zou, L.; Gan, S. PEG-assisted hydrothermal synthesis and photoluminescence of CdMoO4:Tb3+ green phosphor. J. Phys. Chem. Solids 2015, 79, 14. [Google Scholar] [CrossRef]
  10. Lotem, H.; Burshtein, Z. Method for complete determination of a refractive-index tensor by bireflectance: Application to CdWO4. Opt. Lett. 1987, 12, 561–563. [Google Scholar] [CrossRef]
  11. Wang, Y.; Guan, X.; Li, L.; Lin, H.; Wang, X.; Li, G. Solvent-driven polymorphic control of CdWO4 nanocrystal for photocatalytic performances. New J. Chem. 2012, 36, 1852–1858. [Google Scholar] [CrossRef]
  12. Saito, N.; Sonoyama, N.; Sakata, T. Analysis of the Excitation and Emission Spectra of Tungstates and Molybdate. Bull. Chem. Soc. Jpn. 1996, 69, 2191. [Google Scholar] [CrossRef]
  13. Jia, R.; Zhang, G.; Wu, Q.; Yaping, D. Preparation, structures and photoluminescent enhancement of CdWO4-TiO2 composite nanofilms. Appl. Surf. Sci. 2006, 253, 2038–2042. [Google Scholar] [CrossRef]
  14. Lou, Z.D.; Hao, J.H.; Cocivera, M. Luminescence of ZnWO4 and CdWO4 thin films prepared by spray pyrolysis. J. Lumin. 2002, 99, 349. [Google Scholar] [CrossRef]
  15. Greskovich, C.D.; Cusano, D.; Hoffman, D.; Riedner, R.J. Ceramic scintillators for advanced, medical X-ray detectors. Am. Ceram. Soc. Bull. 1992, 71, 1120. [Google Scholar]
  16. Nagornaya, L.; Burachas, S.; Vostretrou, Y.; Martynov, V.; Ryzhikov, V.J. Studies of ways to reduce defects in CdWO4 single crystals. Cryst. Growth 1998, 198, 877. [Google Scholar] [CrossRef]
  17. Oishi, S.; Hirao, M. Effect of starting compositions on the growth of calcium tungstate crystals from sodium tungstate flux. Bull. Chem. Soc. Jpn. 1990, 63, 984. [Google Scholar] [CrossRef] [Green Version]
  18. Liao, H.-W.; Wang, Y.-F.; Liu, X.-M.; Li, Y.-D.; Qian, Y.-T. Hydrothermal Preparation and Characterization of Luminescent CdWO4 Nanorods. Chem. Mater. 2000, 12, 2819–2821. [Google Scholar] [CrossRef]
  19. Tanaka, K.; Sonobe, D. Synthesis of CdWO4 thin films showing photoluminescence at room temperature by pulsed laser ablation using photoluminescenceless CdO/WO3 mixed oxide targets: Effect of plume forming condition and target-substrate distance. Appl. Surf. Sci. 1999, 140, 138. [Google Scholar] [CrossRef]
  20. Blasse, G.; Brixner, L.H. Ultraviolet emission from ABO4-type niobates, tantalates and tungstates. Chem. Phys. Lett. 1990, 173, 409. [Google Scholar] [CrossRef]
  21. Cho, W.; Yashima, M.; Kakihana, M.; Kuso, A.; Sakata, T.; Yoshimura, M. Room-temperature preparation of the highly crystallized luminescent CaWO4 film by an electrochemical method. Appl. Phys. Lett. 1995, 66, 1027. [Google Scholar] [CrossRef]
  22. Srinivas, M.; Modi, D.; Patel, N.; Verma, V.; Murthy, K.V.R. Photoluminescence Studies and Core–Shell Model Approach for Rare Earth doped CdWO4 Nano Phosphor. J. Inorg. Organomet. Polym. 2014, 24, 988. [Google Scholar] [CrossRef]
  23. Premjit Singh, N.; Premananda Singh, N.; Rajmuhon Singh, N.; Mohondas Singh, N. Photoluminescence studies of CdWO4:Sm3+ phosphor: Concentration and annealing effect. Optik 2017, 144, 490. [Google Scholar] [CrossRef]
  24. Ye, D.; Li, D.Z.; Chen, W.; Shao, Y.; Xiao, G.C.; Sun, M.; Fu, X.Z. Characterization and properties of Eu3+-doped CdWO4 prepared by a hydrothermal method. Res. Chem. Intermed. 2009, 35, 675–683. [Google Scholar] [CrossRef]
  25. Narsimha, K.; Anuradha, N.; Sudarshan, K.; Gandhi, A.C.; Raju, A.K.; Reddy, P.M.; Mone, R.; Upender, G.; Kumar, B.V. One-pot hydrothermal preparation and defect-enhanced photocatalytic activity of Bi-doped CdWO4 nanostructures. Phys. Chem. Chem. Phys. 2022, 24, 8775–8786. [Google Scholar] [CrossRef]
  26. Narsimha, K.; Babu, M.S.; Anuradha, N.; Guda, S.; Kumar, B.K.; Mallesh, D.; Upender, G.; Reddy, P.M.; Kumar, B.V. Preparation and characterization of CdWO4:Cu nanorods with enhanced photocatalytic performance under sunlight irradiation. New J. Chem. 2020, 44, 2380–2388. [Google Scholar] [CrossRef]
  27. Stephen, S.K.; Varghese, T. Structural modifications and extended spectral response of nanocrystalline Ba1−xCuxWO4 samples. Mater. Chem. Phys. 2021, 258, 123901. [Google Scholar] [CrossRef]
  28. Tri, N.L.M.; Duc, D.S.; Van Thuan, D.; Tahtamouni, T.A.; Pham, T.-D.; Tran, D.T.; Thi Phuong Le Chi, N.; Nguyen, V.N. Superior photocatalytic activity of Cu doped NiWO4 for efficient degradation of benzene in air even under visible radiation. Chem. Phys. 2019, 525, 110411. [Google Scholar] [CrossRef]
  29. Ney, V.; Venkataraman, V.; Wilhelm, F.; Rogalev, A.; Ney, A. Structural and magnetic properties of Cu-doped ZnO epitaxial films at the coalescence limit—A superparamagnetic CuO-ZnO nanocomposite. J. Appl. Phys. 2019, 126, 143904. [Google Scholar] [CrossRef]
  30. Kumar, C.M.N.; Xiao, Y.; Lunkenheimer, P.; Loidl, A.; Ohl, M. Crystal structure, incommensurate magnetic order, and ferroelectricity in Mn1−xCuxWO4 (0 ≤ x ≤ 0.19). Phys. Rev. B 2015, 91, 235149. [Google Scholar] [CrossRef] [Green Version]
  31. Choudhury, B.; Choudhury, A.; Borah, D. Interplay of dopants and defects in making Cu doped TiO2 nanoparticle a ferromagnetic semiconductor. J. Alloys Compd. 2015, 646, 692–698. [Google Scholar] [CrossRef]
  32. Gandhi, A.C.; Chiu, H.H.; Ho, M.K.; Hsu, T.E.; Li, T.Y.; Wu, Y.H.; Vijaya Kumar, B.; Muralidhar Reddy, P.; Lin, B.H.; Cheng, C.L.; et al. Modulation of Magnetic and Luminescence Properties via Control Cu-Doped in CdWO4 Nanorods for Photocatalytic Applications. ACS Appl. Nano Mater. 2022, 5, 14811–14823. [Google Scholar] [CrossRef]
  33. Zhang, Y.; Zhu, X.; Zhao, Y.; Zhang, Q.; Dai, Q.; Lu, L.; Zhang, L. CdWO4:Eu3+ Nanostructures for Luminescent Applications. ACS Appl. Nano Mater. 2019, 2, 7095–7102. [Google Scholar] [CrossRef]
  34. Nobre, F.X.; Nogueira, I.C.; Souza, G.D.S.; Matos, J.M.E.D.; Couceiro, P.R.D.C.; Brito, W.R.; de la Cruz, J.P.; Leyet Ruiz, Y. Structural and Optical Properties of Ca0.99Cu0.01WO4 Solid Solution Synthesized by Sonochemistry Method at Room Temperature. Inorg. Chem. 2020, 59, 6039–6046. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, X.F.; Su, W.; Hu, X.H.; Liu, H.H.; Sheng, M.Q.; Zhou, Q.Y. Synthesis and characterization of Fe-doped CdWO4 nanoparticles with enhanced photocatalytic activity. Mater. Res. Express 2019, 6, 035507. [Google Scholar] [CrossRef]
  36. Hao, C.T.; Zhou, Y.Z.; Dang, Y.; Chai, S.N.; Han, G.P.; Li, Z.L.; Zhang, H.Z.; Zhang, Y.C. The partial substitution of Cd by La ions in CdWO4 nanocrystal for the efficiently enhanced electrochemical sensing of BPA. Electrochim. Acta 2019, 318, 581–589. [Google Scholar] [CrossRef]
  37. Wang, F.; Han, Y.; Lim, C.S.; Lu, Y.; Wang, J.; Xu, J.; Chen, H.; Zhang, C.; Hong, M.; Liu, X. Simultaneous phase and size control of upconversion nanocrystals through lanthanide doping. Nature 2010, 463, 1061–1065. [Google Scholar] [CrossRef]
  38. Mingjiang, Y.; Xun, J.; Zhangn, Z.; Zhou, Y. Eu3+-doped CdWO4 phosphor forred-lightemission: Hydrothermal preparation and blue light excitation. Ceram. Int. 2014, 40, 16189–16194. [Google Scholar]
  39. Singh, N.P.; Singh, N.R.; Singh, N.R.; Singh, N.M. Effect of doping ion concentration on the photoluminescence behavior of CdWO4:Tb3+ phosphor synthesized via co-precipitation method. Indian J. Phys. 2018, 92, 1461–1466. [Google Scholar] [CrossRef]
  40. Bish, D.L.; Howard, S.A. Quantitative phase analysis using the Rietveld method. J. Appl. Cryst. 1988, 21, 86–91. [Google Scholar] [CrossRef]
  41. Wang, Y.; Wu, C.; Geng, L.; Chen, S. Unexpected formation of scheelite-structured Ca1−xCdxWO4 (0 ≤ x ≤ 1) continuous solid solutions with tunable photoluminescent and electronic properties. Phys. Chem. Chem. Phys. 2017, 19, 23204–23212. [Google Scholar] [CrossRef] [PubMed]
  42. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. A 2016, 32, 751. [Google Scholar] [CrossRef]
  43. Pundareekam Goud, J.; Mahamoud, S.A.; James Raju, K.C. Structural, dielectric and impedance study of Bi and Li co-substituted Ba0.50Sr0.50TiO3 ceramics for tunable microwave devices applications. J. Mater. Sci. Mater. 2018, 29, 3611–3620. [Google Scholar] [CrossRef]
  44. Moghadam, M.T.T.; Babamoradi, M.; Azimirad, R. Effect of Hydrothermal Reaction Temperature on the Photocatalytic Properties of CdWO4-RGO Nanocomposites. J. Nanostruct. 2019, 9, 600–609. [Google Scholar]
  45. Liu, Y.; Li, L.; Wang, R.; Li, J.; Huang, J.; Zhang, W. Multi-mode photocatalytic performances of CdS QDs modified CdIn2S4/CdWO4 nanocomposites with high electron transfer ability. J. Nanopart. Res. 2018, 20, 319. [Google Scholar] [CrossRef]
  46. Mai, M.; Feldmann, C. Microemulsion-base synthesis and luminescence of nanoparticulate CaWO4, ZnWO4, CaWO4: Tb, and CaWO4: Eu. J. Mater. Sci. 2012, 47, 1427–1435. [Google Scholar] [CrossRef]
  47. Harshan, H.; Priyanka, K.P.; Sreedevi, A.; Jose, A.; Varghese, T. Structural, optical and magnetic properties of nanophase NiWO4 for potential applications. Eur. Phys. J. B 2018, 91, 287. [Google Scholar] [CrossRef]
  48. Ross-Medgaarden, E.I.; Wachs, I.E. Structural determination of bulk and surface tungsten oxides with UV-vis diffuse reflectance spectroscopy and Raman spectroscopy. J. Phys. Chem. C 2007, 111, 15089–15099. [Google Scholar] [CrossRef]
  49. Wang, Y.; Wu, W.; Fu, X.; Liu, M.; Cao, J.; Shao, C.; Chen, S. Metastable scheelite CdWO4:Eu3+ nanophosphors: Solvothermal synthesis, phase transitions and their polymorph-dependent luminescence properties. Dye. Pigm. 2017, 147, 283–290. [Google Scholar] [CrossRef]
  50. Naiara, A.L.; Lorena, D.S.A.; Siu-Li, M.; Carlos, A.C.F.; Alexandre, M.; Jean-Claude, M.; Maria, I.B.B. NiWO4 powders prepared via polymeric precursor method for application as ceramic luminescent pigments. J. Adv. Ceram. 2020, 9, 55–63. [Google Scholar]
  51. Dabre, K.V.; Dhoble, S.J. Synthesis and Photoluminescence properties of Eu3+, Sm3+ and Pr3+ doped CaZnWO6 phosphor for phosphor converted LED. J. Lumin. 2014, 150, 55–58. [Google Scholar] [CrossRef]
  52. Ruiz-Fuertes, J.; Friedrich, A.; Pellicer-Porres, J.; Errandonea, D.; Segura, A.; Morgenroth, W.; Haussühl, E.; Tu, C.Y.; Polian, A. Structure Solution of the High-Pressure Phase of CuWO4 and Evolution of the Jahn–Teller Distortion. Chem. Mater. 2011, 23, 4220–4226. [Google Scholar] [CrossRef]
  53. Dutta, D.P.; Ningthoujam, R.S.; Tyagi, A.K. Luminescence properties of Sm3+ doped YPO4: Effect of solvent, heat treatment, Ca2+/W6+co-doping and its hyperthermia application. AIP Adv. 2012, 2, 042184. [Google Scholar] [CrossRef]
  54. Singh, N.S.; Ningthoujam, R.S.; Phaomei, G.; Singh, S.D.; Vinud, A.; Vatsa, R.K. Redispersion and film formation of GdVO4:Ln3+ (Ln3+ = Dy3+, Eu3+, Sm3+, Tm3+) nanoparticles: Particle size and luminescence studies. Dalton Trans. 2012, 41, 4404–4412. [Google Scholar] [CrossRef] [Green Version]
  55. Klein, P.B.; Nwagwu, U.; Edgar, J.H.; Freitas, J.A. Photoluminescence investigation of the indirect band gap and shallow impurities in icosahedral B12As2. J. Appl. Phys. 2012, 112, 013508. [Google Scholar] [CrossRef]
  56. Gregory, M.K.; Astrid, M.M.; Nathan, S.L.; Harry, A.A. Photoluminescence-based measurements of the energy gap and diffusion length of Zn3P2. Appl. Phys. Lett. 2019, 95, 112103. [Google Scholar]
  57. Zhang, C.L.; Guo, D.L.; Xu, W.N.; Hu, C.G.; Chen, Y.X. Radiative/Nonradiative Recombination Affected by Defects and Electron-Phone Coupling in CdWO4 Nanorods. J. Phys. Chem. C 2016, 120, 12218–12225. [Google Scholar] [CrossRef]
  58. Laasner, R. G0W0 band structure of CdWO4. J. Phys. Condens. Matter 2014, 26, 125503. [Google Scholar] [CrossRef]
  59. Kadhim, A.A.; Baida, M.A.; Madyan, A.K. Influence of Sn doping ratio on the structural and optical properties of CdO films prepared by laser induced plasma. Iraq. J. Phys. 2020, 18, 1–8. [Google Scholar]
  60. Fang, Y.; Wang, L.; Sun, Q.; Lu, T.; Deng, Z.; Ma, Z.; Jiang, Y.; Jia, H.; Wang, W.; Zhou, J.; et al. Investigation of temperature-dependent photoluminescence in multi-quantum wells. Sci. Rep. 2015, 5, 12718. [Google Scholar] [CrossRef] [Green Version]
  61. Wang, L.; Moon, B.K.; Park, S.H.; Kim, J.H.; Shi, J.; Kim, K.H.; Jeong, J.H. Synthesis and photoluminescence of Bi3+, Eu3+ doped CdWO4 phosphors: Application of energy level rules of Bi3+ ions. New J. Chem. 2016, 40, 3552. [Google Scholar] [CrossRef]
  62. Šroubek, Z.; Ždánský, K. Electron Spin Resonance of Cu2+ Ion in CdWO4, ZnWO4, and MgWO4 Single Crystals. J. Chem. Phys. 1966, 44, 3078–3083. [Google Scholar] [CrossRef]
Figure 1. (ad) FE-SEM images of CdWO4 and 1, 3, 5% Sn-doped CdWO4 NRDs. (eh) The histogram distribution in the diameter of Cd1−xSnxWO4 (x = 0, 1, 3, and 5%) NRDs.
Figure 1. (ad) FE-SEM images of CdWO4 and 1, 3, 5% Sn-doped CdWO4 NRDs. (eh) The histogram distribution in the diameter of Cd1−xSnxWO4 (x = 0, 1, 3, and 5%) NRDs.
Ijms 23 15123 g001
Figure 2. (ac) TEM images, (df) high-resolution TEM images, and (gi) SAED pattern of 1, 3, and 5% Sn-doped CdWO4 NRDs (top to bottom).
Figure 2. (ac) TEM images, (df) high-resolution TEM images, and (gi) SAED pattern of 1, 3, and 5% Sn-doped CdWO4 NRDs (top to bottom).
Ijms 23 15123 g002
Figure 3. (ad) The refined PXRD patterns of Cd1−xSnxO4 with x = 0 (CdWO4), 1, 3, and 5%. (e) Variation of lattice constants with Sn concentrations. (f) Diffraction peak (−1 1 1) from all the samples where the vertical dashed line marks the peak center of the CdWO4 nanorod. (g) Variation of crystallite size and microstrain with Sn concentrations.
Figure 3. (ad) The refined PXRD patterns of Cd1−xSnxO4 with x = 0 (CdWO4), 1, 3, and 5%. (e) Variation of lattice constants with Sn concentrations. (f) Diffraction peak (−1 1 1) from all the samples where the vertical dashed line marks the peak center of the CdWO4 nanorod. (g) Variation of crystallite size and microstrain with Sn concentrations.
Ijms 23 15123 g003
Figure 4. (a) Raman spectra of Cd1−xSnxWO4 (x = 0, 1, 3, and 5%) NRDs (The asterisks (*) indicate the internal stretching modes). (b) A plot of the vibrational band observed at 898 cm−1 from 0 to 5% Sn-doped CdWO4 NRDs. (c) The integrated intensity vs. Sn doping concentrations.
Figure 4. (a) Raman spectra of Cd1−xSnxWO4 (x = 0, 1, 3, and 5%) NRDs (The asterisks (*) indicate the internal stretching modes). (b) A plot of the vibrational band observed at 898 cm−1 from 0 to 5% Sn-doped CdWO4 NRDs. (c) The integrated intensity vs. Sn doping concentrations.
Ijms 23 15123 g004
Figure 5. (ac) The excitation and emission spectrum obtained at 15 K (black) and 300 K (red) (df), a 2D plot of the emission spectrum, where the vertical axis represents temperature varying from 15 to 300 K.
Figure 5. (ac) The excitation and emission spectrum obtained at 15 K (black) and 300 K (red) (df), a 2D plot of the emission spectrum, where the vertical axis represents temperature varying from 15 to 300 K.
Ijms 23 15123 g005
Figure 6. (a) CIE 1931 color space chromaticity diagram in the (X, Y) coordinate system shows a shift in the emission color from light blue to sky blue of the PL emission with increased Sn-content. The plot on the right-hand side shows the magnified rectangular area. (bd) Temperature-dependent photoluminescence spectra of undoped and Sn-doped CdWO4 for the most intense peak at the direct bandgap at room temperature and lower temperatures. (e) Schematic diagram of photoluminescence mechanism of Cd1−xSnxWO4 (x = 0, 1 and 5%) NRDs.
Figure 6. (a) CIE 1931 color space chromaticity diagram in the (X, Y) coordinate system shows a shift in the emission color from light blue to sky blue of the PL emission with increased Sn-content. The plot on the right-hand side shows the magnified rectangular area. (bd) Temperature-dependent photoluminescence spectra of undoped and Sn-doped CdWO4 for the most intense peak at the direct bandgap at room temperature and lower temperatures. (e) Schematic diagram of photoluminescence mechanism of Cd1−xSnxWO4 (x = 0, 1 and 5%) NRDs.
Ijms 23 15123 g006
Figure 7. (ac) The temperature-dependent integrated PL intensity of Cd1−xSnxWO4 (x = 0, 1, and 5%) NRDs.
Figure 7. (ac) The temperature-dependent integrated PL intensity of Cd1−xSnxWO4 (x = 0, 1, and 5%) NRDs.
Ijms 23 15123 g007
Figure 8. (ac) Photoluminescence decay curves of Cd1−xSnxWO4 (x = 0, 1, and 5%) NRDs.
Figure 8. (ac) Photoluminescence decay curves of Cd1−xSnxWO4 (x = 0, 1, and 5%) NRDs.
Ijms 23 15123 g008
Scheme 1. Schematic representation of solution preparation, filtering and drying of Cd1−xSnxWO4 (x = 0, 1, 3, and 5%) samples using in situ hydrothermal methods.
Scheme 1. Schematic representation of solution preparation, filtering and drying of Cd1−xSnxWO4 (x = 0, 1, 3, and 5%) samples using in situ hydrothermal methods.
Ijms 23 15123 sch001
Table 1. Summary of the crystallite sizes, lattice parameters, microstrains, and volume obtained from PXRD of Sn-doped CdWO4 NRDs.
Table 1. Summary of the crystallite sizes, lattice parameters, microstrains, and volume obtained from PXRD of Sn-doped CdWO4 NRDs.
Sn-DopingCrystallite Size (D) nmLattice Parameters (Å)η (%)Volume (Å3)
abc
0625.02985.86105.07530.01619149.570
1%405.02795.86445.07780.02495149.673
3%395.02705.86835.07940.02506149.786
5%385.02495.86965.07920.02624149.812
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Manjunatha, K.; Ho, M.-K.; Hsu, T.-E.; Chiu, H.-H.; Li, T.-Y.; Kumar, B.V.; Reddy, P.M.; Chan, T.S.; Wu, Y.-H.; Lin, B.-H.; et al. Precise Sn-Doping Modulation for Optimizing CdWO4 Nanorod Photoluminescence. Int. J. Mol. Sci. 2022, 23, 15123. https://doi.org/10.3390/ijms232315123

AMA Style

Manjunatha K, Ho M-K, Hsu T-E, Chiu H-H, Li T-Y, Kumar BV, Reddy PM, Chan TS, Wu Y-H, Lin B-H, et al. Precise Sn-Doping Modulation for Optimizing CdWO4 Nanorod Photoluminescence. International Journal of Molecular Sciences. 2022; 23(23):15123. https://doi.org/10.3390/ijms232315123

Chicago/Turabian Style

Manjunatha, K., Ming-Kang Ho, Tsu-En Hsu, Hsin-Hao Chiu, Tai-Yue Li, B. Vijaya Kumar, P. Muralidhar Reddy, Ting San Chan, Yu-Hao Wu, Bi-Hsuan Lin, and et al. 2022. "Precise Sn-Doping Modulation for Optimizing CdWO4 Nanorod Photoluminescence" International Journal of Molecular Sciences 23, no. 23: 15123. https://doi.org/10.3390/ijms232315123

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop