Next Article in Journal
A Rapid and Simple UHPLC-MS/MS Method for Quantification of Plasma Globotriaosylsphingosine (lyso-Gb3)
Previous Article in Journal
The Anthelmintic Quassinoids Ailanthone and Bruceine a Induce Infertility in the Model Organism Caenorhabditis elegans by an Apoptosis-like Mechanism Induced in Gonadal and Spermathecal Tissues
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural and Spectroscopic Effects of Li+ Substitution for Na+ in LixNa1-xCaGd0.5Ho0.05Yb0.45(MoO4)3 Scheelite-Type Upconversion Phosphors

by
Chang-Sung Lim
1,*,
Aleksandr Aleksandrovsky
2,3,
Maxim Molokeev
4,5,6,
Aleksandr Oreshonkov
7,8 and
Victor Atuchin
9,10,11,*
1
Department of Aerospace Advanced Materials and Chemical Engineering, Hanseo University, Seosan 31962, Korea
2
Laboratory of Coherent Optics, Kirensky Institute of Physics Federal Research Center KSC SB RAS, 660036 Krasnoyarsk, Russia
3
Institute of Nanotechnology, Spectroscopy and Quantum Chemistry, Siberian Federal University, 660041 Krasnoyarsk, Russia
4
Laboratory of Crystal Physics, Kirensky Institute of Physics, Federal Research Center KSC SB RAS, 660036 Krasnoyarsk, Russia
5
Institute of Engineering Physics and Radioelectronics, Siberian Federal University, 660041 Krasnoyarsk, Russia
6
Department of Physics, Far Eastern State Transport University, 680021 Khabarovsk, Russia
7
Laboratory of Molecular Spectroscopy, Kirensky Institute of Physics, Federal Research Center KSC SB RAS, 660036 Krasnoyarsk, Russia
8
School of Engineering and Construction, Siberian Federal University, 660041 Krasnoyarsk, Russia
9
Laboratory of Optical Materials and Structures, Institute of Semiconductor Physics, SB RAS, 630090 Novosibirsk, Russia
10
Research and Development Department, Kemerovo State University, 650000 Kemerovo, Russia
11
Department of Industrial Machinery Design, Novosibirsk State Technical University, 630073 Novosibirsk, Russia
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(23), 7357; https://doi.org/10.3390/molecules26237357
Submission received: 27 October 2021 / Revised: 24 November 2021 / Accepted: 30 November 2021 / Published: 3 December 2021
(This article belongs to the Section Inorganic Chemistry)

Abstract

:
A set of new triple molybdates, LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45, was successfully manufactured by the microwave-accompanied sol–gel-based process (MAS). Yellow molybdate phosphors LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 with variation of the LixNa1-x (x = 0, 0.05, 0.1, 0.2, 0.3) ratio under constant doping amounts of Ho3+ = 0.05 and Yb3+ = 0.45 were obtained, and the effect of Li+ on their spectroscopic features was investigated. The crystal structures of LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 (x = 0, 0.05, 0.1, 0.2, 0.3) at room temperature were determined in space group I41/a by Rietveld analysis. Pure NaCaGd0.5Ho0.05Yb0.45(MoO4)3 has a scheelite-type structure with cell parameters a = 5.2077 (2) and c = 11.3657 (5) Å, V = 308.24 (3) Å3, Z = 4. In Li-doped samples, big cation sites are occupied by a mixture of (Li,Na,Gd,Ho,Yb) ions, and this provides a linear cell volume decrease with increasing Li doping level. The evaluated upconversion (UC) behavior and Raman spectroscopic results of the phosphors are discussed in detail. Under excitation at 980 nm, the phosphors provide yellow color emission based on the 5S2/5F45I8 green emission and the 5F55I8 red emission. The incorporated Li+ ions gave rise to local symmetry distortion (LSD) around the cations in the substituted crystalline structure by the Ho3+ and Yb3+ ions, and they further affected the UC transition probabilities in triple molybdates LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45. The complex UC intensity dependence on the Li content is explained by the specificity of unit cell distortion in a disordered large ion system within the scheelite crystal structure. The Raman spectra of LixNa1-xCaGd0.5(MoO4)3 doped with Ho3+ and Yb3+ ions were totally superimposed with the luminescence signal of Ho3+ ions in the range of Mo–O stretching vibrations, and increasing the Li+ content resulted in a change in the Ho3+ multiplet intensity. The individual chromaticity points (ICP) for the LiNaCaGd(MoO4)3:Ho3+,Yb3+ phosphors correspond to the equal-energy point in the standard CIE (Commission Internationale de L’Eclairage) coordinates.

Graphical Abstract

1. Introduction

Complex molybdate crystals have become a subject of extensive investigation due to their diverse crystal chemistry, high chemical stability, and specific physical properties, which are promising for applications in such fields as electronics, laser systems, electrochemistry, and electro-photonics [1,2,3,4,5,6,7,8,9]. These lanthanide-activated double molybdates provide stable chemical and physical characteristics with excellent optical properties for favorable lanthanide admittance, relatively low phonon energy, and versatile applications in such fields as solar cells, lasers, biomedical and optoelectronic devices, optical sensors, etc. [10,11,12,13,14,15,16]. Many molybdate crystals are appropriate for the incorporation of rare earth (Ln) ions in their structure, and the materials are considered as potential hosts for the creation of phosphors to be used in photonic structures [2,3,4,7,17,18,19,20,21,22,23]. Among such crystals, scheelite-type (ST) molybdates are particularly interesting in terms of the search for new structures, including structure-modulation effects, and promising spectroscopic characteristics [2,21,24,25,26,27,28,29,30]. Scheelites of general composition ABO4 (A = A2+ cation, B = Mo, W) crystallize in space group I41/a. Generally, an ST structure is very stable and highly tolerant to mixed cation accommodation at the A position, which opens up the possibility for the creation of solid solutions with wide-range doping by Ln ions. In accordance with this, new binary and ternary molybdates with the ST structure and cation disorder at the A position were synthesized and their basic properties were evaluated [21,23,24,25,26,27,28,29,30,31]. Commonly, in combination with Ln3+ ions, such lower valence ions as Li+, Na+, Ag+, Ca2+, Sr2+, and Pb2+ are used to reach the average charge balance at the A position because these ions have appropriate effective radii [32].
In ST solid solutions, different combinations of Ln3+ ions can be reached without a loss of structural quality; this opens the way for the investigation of energy transfer (ET) effects in phosphor systems with binary and ternary doping [33,34,35,36,37,38,39]. In particular, the incorporation of appropriate Ln3+ pairs provides enhanced frequency upconversion (UC) properties under near-infrared laser excitation [21,29,30,31,33,34,35,36,38,39]. For Ho3+/Yb3+-doped UC phosphors, yellow emissions can be derived via a co-doping system based on the red emission bands from 5F55I8 transitions and green emission bands from 5S2/ 5F45I8 transitions [40,41,42,43]. Extremely wide color tunability was reported for Ho3+/Yb3+ UC phosphor upon the addition of Er3+ ions [44]. Laser active Ho3+ ions play the role of activator, and Yb3+ ions, as an efficient sensitizer with a high absorption cross section at 980 nm, could enhance the UC efficiency through the ET process between the activator and sensitizer induced by a unique energy level configuration.
The preparation of complex molybdates can be carried out via several specific processes [45,46,47,48,49,50,51,52,53,54,55]. Among the different methods, microwave accompanied sol–gel-based synthesis (MAS) can provide high-quality crystalline materials in a short processing period. The powder products fabricated by the MAS route are commonly characterized by a homogeneous microstructure and high chemical reproducibility, in reference to the nominal starting reagent ratio that is governed by the sol preparation step with high atom intermixing [56,57]. Compared with the usual methods, microwave synthesis has the advantages of shortening the reaction time and resulting in products with a small particle size, narrow particle size distribution, and high purity [58,59,60,61,62,63]. At the final step, the high-temperature microwave treatment efficiently stimulates crystal lattice formation [64,65].
Previously, a new division of triple ST molybdates NaTLn(MoO4)3 (T = Ca, Pb; Ln = La, Gd) was designed and prepared by the MAS process, and their properties, as promising phosphor hosts, were reported [21,31,64,65,66]. Also, the LixNa1-xCaLa0.5Ho0.05Yb0.45(MoO4)3 solid solution was evaluated to determine the effect of Li+ substitution for Na+ on the frequency upconversion, and the ST structure was observed in the range of x = 0.05−0.4 [67]. To extend the capabilities of ST crystal engineering, it is interesting to reveal ions and ion combinations appropriate for incorporation at the mixed A position in complex ST crystals. In relation to this, the triple ST solid solutions (Li,Na)MLn(MoO4)3 (M = Ca, Sr, Pb) are of particular interest because the presence of Li+ ions can affect the photoluminescence properties of Ln3+ activators [67]. Thus, the present contribution is aimed at the preparation of LixNa1-xCaLa(MoO4)3;Yb3+,Ho3+ compounds by the MAS-based method and evaluation of their spectroscopic properties. The influence of incorporated Li+ ions was investigated—to enhance the UC transition probabilities by the Ho3+ and Yb3+ ions in the ST molybdate LixNa1-xCaGd0.5(MoO4)3—and is discussed in terms of the local symmetry distortion (LSD) around the cations in the substituted crystalline structure, The crystal structures and morphologies of the synthesized particles were evaluated by X-ray diffraction (XRD) and scanning electron microscopy (SEM), respectively. The spectroscopic characteristics were investigated under the consideration of efficient UC emissions, individual chromaticity points (ICP) according to Commission Internationale de L’Eclairage (CIE), and Raman scattering.

2. Experimental Section

In this experiment, molybdate solid solution LixNa1-xCaGd0.5(MoO4)3: Ho3+0.05/Yb3+0.45 was designed to achieve efficient UC luminescent characteristics by variation of the Li/Na ratio (x= 0, 0.05, 0.1, 0.2, and 0.3) under fixed rare earth element contents of Gd3+ = 0.5, Ho3+ = 0.05, and Yb3+ = 0.45. Na2MoO4∙2H2O, Ca(NO3)2∙4H2O, Gd(NO3)3∙6H2O, (NH4)6Mo7O24∙4H2O, LiNO3, and Ho(NO3)3∙5H2O with 99.0% purity were purchased from Sigma-Aldrich, USA. Yb(NO3)3∙5H2O, with 99.9% purity, was purchased from Sigma-Aldrich, USA. Besides these, citric acid (CA) at 99.5% purity was obtained from Daejung Chemicals, Korea. Distilled water (DW), ethylene glycol (EG, A.R.), and NH4OH (A.R.) were used to bring about the transparent sol formation.
The sample notation introduced according to the nominal compositions is given in Table 1. Initially, for the sol preparation of (a) NCGM:HY, Ca(NO3)2∙4H2O at 0.4 mol%, Na2MoO4∙2H2O at 0.2 mol%, and (NH4)6Mo7O24∙4H2O at 0.171 mol% were dissolved in 80 mL 8M NH4OH with 20 mL EG. To make the sol of (b) LiNCGM:HY-0.05 for Li0.05Na0.95, Ca(NO3)2∙4H2O at 0.4 mol%, Na2MoO4∙2H2O at 0.19 mol%, LiNO3 at 0.01 mol %, and (NH4)6Mo7O24∙4H2O at 0.171 mol% were used. For the compositions of (c) LiNCGM:HY-0.1 for Li0.1Na0.9, (d) LiNCGM:HY-0.2 for Li0.2Na0.8, and (e) LiNCGM:HY-0.3 for Li0.3Na0.7, the reagent sets (c) Na2MoO4∙2H2O at 0.18 mol% and LiNO3 at 0.02 mol %, (d) Na2MoO4∙2H2O at 0.16 mol% and LiNO3 at 0.04 mol %, and (e) Na2MoO4∙2H2O at 0.14 mol% and LiNO3 at 0.06 mol % were applied. Subsequently, Gd(NO3)3∙6H2O at 0.2 mol%, Yb(NO3)3∙5H2O at 0.18 mol%, and Ho(NO3)3∙5H2O at 0.02 mol% were carefully weighed and dissolved very slowly in 100 mL of DW under slight heat treatment. Then, these two kinds of prepared solutions were slowly co-mixed and vigorously stirred. The CA/CM (molar ratio of CA to the total cation metal (CM) ions) was adjusted to 2:1. The intermixed solutions were 180–200 mL in volume, and they were heated slowly to ~80–100 °C in a 450 mL Pyrex glass. At this stage, the solutions were in a highly transparent state. Then, the solutions were subjected to the MAS-derived treatment. The typical procedure applied at this stage can be found elsewhere [21,31,64,65,66,67]. Then, the obtained dried black gels were ground and annealed at 800 °C for 16 h in air, with intervals of 100 °C between 600 and 800 °C. After the annealing process, pink-colored particles were obtained for the samples.
The structural properties of synthesized samples were evaluated by XRD analysis. The powder XRD patterns of the LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 particles for Rietveld analysis were precisely recorded over the angle range of 2θ = 5−90° at room temperature using a D/MAX 2200 (Rigaku in Japan) diffractometer with Cu Kα radiation and θ-2θ geometry. The 2θ size step was 0.02°, and the counting time was 5 s per step. The TOPAS 4.2 package was applied for the Rietveld analysis [68]. The typical microstructure and surface morphology of the obtained particles were observed using SEM (JSM-5600, JEOL in Japan) methods. The PL spectra were recorded at room temperature using a spectrophotometer (Perkin Elmer LS55 in UK). The Raman spectral measurements were performed using a LabRam Aramis (Horiba Jobin-Yvon in France) device with spectral resolution of 2 cm−1. The 514.5 nm line of an Ar ion laser was used as an excitation source and, to avoid sample decomposition, the power on the samples was kept at the 0.5 mW level.

3. Results and Discussion

The XRD patterns measured for the samples listed in Table 1 are shown in Figure 1 and Figures S1–S3 (Supporting Information). In general, the patterns are similar. In Figure 1, for comparison, the XRD patterns of NCGM:HY (x = 0) and LiNCGM:HY-0.3 (x = 0.3) are presented. As can be seen, there is no significant difference even for the highest Li content. All peaks in all patterns were successfully indexed by the tetragonal cell (I41/a) with cell parameters close to those of CaMoO4 [69]. Therefore, the crystal structure of CaMoO4 was taken as a starting model for Rietveld refinement. The Ca2+ ion site was considered as the one occupied by a mixture of Li+, Ca2+, Na+, Gd3+, Ho3+, and Yb3+ ions (Figure 2) with fixed partial occupations according to the nominal sample composition. The refinements were stable and gave low R-factors (Table 1, Figure 1 and Figures S1–S3). The atom coordinates and the main bond lengths are summarized in Tables S1 and S2, respectively.
As is known, the effective radius of the Li+ ion is noticeably lower than that of the Na+ ion; therefore, substitution of Li+ for Na+ in the LiNCGM:HY compounds should induce a decrease in the average ion radius IR(Li/Na/Ca/Gd/Ho/Yb) of the A position and a related unit cell volume decrease in the ST structures. The dependencies of cell parameters and unit cell volume on IR(Li/Na/Ca/Gd/Ho/Yb) in the LiNCGM:HY compounds are shown in Figure 3. The IR values were calculated on the basis of the nominal compositions and the known system of ion radii [32]. For comparison, the same dependencies of two previously studied Gd-containing systems, NaCaGd(MoO4)3:Er,Yb and NaCaGd(MoO4)3:Ho,Yb [64,65], are also given in Figure 3. It should be pointed out that only the ST structure type was obtained in Refs. [50,51], and the structural parameters of NaCaGd(MoO4)3:Er,Yb and NaCaGd(MoO4)3:Ho,Yb remain unknown. For this reason, in the present work, the Rietveld refinement of the structures was implemented on the basis of the previously recorded XRD data [64,65]. The refinements were stable and gave low R-factors (Tables S3 and S4). From the observation of Figure 3, it is evident that, in the LiNCGM:HY compounds, the cell parameters and cell volume continuously decrease with decreasing IR(Li/Na/Ca/Gd/Ho/Yb) value or increasing Li content. This clearly proves the suggested chemical formulas of the LiNCGM:HY solid solutions. As to NaCaGd(MoO4)3:Er,Yb and NaCaGd(MoO4)3:Ho,Yb, with increasing doping level, the cell parameters and cell volume decrease proportionally to the IR value, according to the general trend observed in ST molybdates [66]. However, the decrease rate due to Li+ incorporation, instead of Na+, in LiNCGM:HY is strongly lower than that in NaCaGd(MoO4)3:Er,Yb and NaCaGd(MoO4)3:Ho,Yb due to the rare earth ion incorporation instead of Gd3+. Moreover, with increasing Li content, the points related to the LiNCGM:HY system move away from the point of CdMoO4, which is an ST molybdate with the lowest unit cell volume [70]. To verify this effect, it would be valuable to see the behavior in other related systems, and this is possible for LixNa1-xCaLa0.5 (MoO4)3:Ho0.05Yb0.45 (LiNCLM:HY) [67]. To determine the structural parameters of the LiNCLM:HY solid solutions, Rietveld refinement was carried out for the LiNCLM:HY compounds on the basis of the previously reported XRD data [67]. The refinements were stable and gave low R-factors (Table S5). The obtained results are shown in Figure 4 in association with other known La-containing systems NaTLa(MoO4)3:Er,Yb (T = Ca, Sr, Pb) [21,31,66]. As can be seen, the variation in the unit cell volume with increasing Li+ content in the LiNCLM:HY crystals is nearly the same as that in LiNCGM:HY. Comparatively, NaTLa(MoO4)3:Er,Yb (T = Ca, Sr, Pb) solutions strongly follow the general trend of ST molybdates [66]. Thus, the substitution of Li+ for Na+ in LiNCGM:HY and LiNCLM:HY crystals generates a specific effect of unit cell compression that is not governed by the general trend of ST structures. It can be supposed that similar effects could be observed in other LixNa1-xTLn(MoO4)3 solid solutions. Further details of the crystal structures of the compounds listed in Table 2 and Tables S3–S5 may be obtained from Fachinformationszentrum Karlsruhe, 76344 Eggenstein-Leopoldshafen, Germany (fax: (+49)7247-808-666; E-mail: [email protected]; http://www.fiz-karlsruhe.de/request_for_deposited_data.html) by quoting the deposition number: 2117686-2117704.
The SEM images obtained for the representative compositions (a) NCGM:HY, (b) LiNCM:HY-0.05, (c) LiNCGM:HY-0.1, (d) LiNCGM:HY-0.2, and (e) LiNCGM:HY-0.3 are shown in Figure 5. As can be seen, the particle morphology is similar for all five samples. The samples contain uniform partly coalescent particles of 3–10 μm in size. However, the Li-containing samples presented in Figure 5b–e show a slightly smaller characteristic particle size of 3–5 μm, compared to that of the NCGM:HY sample (5–10 μm). Faceted forms were not detected, and this may be due to the comparatively short time of high-temperature treatment used in the MAS processing. In particular, agglomerated grains can be observed, and these could be induced by the material interdiffusion between the grains. Previously, similar grain agglomeration was observed in many other oxide materials subjected to high-temperature annealing [71,72,73]. Thus, MAS-derived synthesis, when applied to LiNCGM doped with Yb3+/Ho3+, provides powder products with an uniform micrograin morphology.
The Raman spectra recorded for LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 (x = 0, 0.05, 0.1, 0.2, 0.3) are shown in Figure 6a. As can be seen in Table S1, all of the investigated compounds contain only one crystallographically independent MoO4 tetrahedron, and these units occupy the sites with S4 point symmetry. It follows from the factor group analysis that Raman-active [MoO4]2− ion vibrations can be listed as: Ag, symmetric stretching; Bg + Eg, antisymmetric stretching; Ag + Bg, symmetric bending; and Bg + Eg, antisymmetric bending [28,74]. The strongest spectral band in Figure 6a at 884 cm−1 is related to the ν1 Ag symmetric stretching vibration, and the ν3 antisymmetric stretching vibrations are located in the range of 720–860 cm−1. Part of the spectra between 820 and 1350 cm−1 is covered by the luminescence of Ho3+ ions, and the determination of exact positions of the peaks related to antisymmetric stretching vibrations is therefore impossible [3]. The Raman spectrum decomposition in the range of free MoO4 rotation and bending vibrations for Li0.05Na0.95CaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 is presented in Figure 6b. The spectral contour in the range of 300–360 cm−1 consists of two lines related to symmetric stretching, and two lines were found in the range of 380–425 cm−1 (antisymmetric bending). The spectral peak at 208 cm−1 is shown with a single line and corresponds to the free rotation of MoO4 tetrahedra [75]. Thus, the number of observed peaks in this region is in accordance with the group theoretical analysis, confirming the high crystallinity of the samples.
It is interesting to note that the Raman profiles of LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 (x= 0, 0.05, 0.1, 0.2, 0.3) in the range of free rotation and bending modes of MoO4 tetrahedra are sensitive to the Li content, as seen in Figure 7. The range of MoO4 stretching vibrations is beyond the scope of this discussion because of the overlapping of the Raman signal with the Ho3+ luminescence. As was mentioned earlier, the substitution of Li+ for Na+ should induce a decrease in the average ion radius IR(Li/Na/Ca/Gd/Ho/Yb) of the A position and a related unit cell volume decrease. In turn, this should lead to minor changes in the geometry of molybdenum–oxygen tetrahedra, namely, changes in bond lengths (see Table S2) and angles that should affect the Raman shift of vibration modes. The graphical representation of MoO4 bending vibrations is presented in Figure 8, and we can suppose that the angle variation in MoO4 units should affect the symmetric bending, while the bond length variation should lead to changes in the Raman shift of antisymmetric bending vibrations. Thus, we can summarize that the Li content variation in LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 leads to minor changes in the MoO4 bond length that are consistent with the changes in the Raman profiles in the range of antisymmetric bending vibrations. For the free MoO4 rotation vibration, the Raman line position variation is less than 1 cm−1.
The UC emission spectra of Na1-xLixCaGd(MoO4)3:Ho,Yb at x = 0, 0.05, 0.1, 0.2, and 0.3 excited at 980 nm at room temperature are shown in Figure 9. Under the excitation at 980 nm, the samples exhibited yellow emission composed of red and green emission bands of the Ho3+ ion, namely, the 5S2/5F45I8 band is in green and the 5F55I8 band is in red [42]. The peak intensity variation of the green and red bands, as well as the integral intensity of UC luminescence over the whole spectrum, shows its complex behavior, as seen in Figure S4, maximizing at two different Li contents of x = 0.1 and 0.3. The incorporation of Li+ ions into the lattice of hosts instead of ions with a larger radius is known to be efficient for controlling the luminescence of doping ions via the crystal field variation affecting these luminescing ions [76]. Li+ ions commonly do not destroy or alter the local symmetry of the rare earth ions in the lattice, but they can provide more appropriate UC intensities [76,77]. This is the case for the crystal structure under study. All Na, Li Ca, Gd, Ho, and Yb ions occupy the same site in the scheelite crystal structure. The introduction of Li instead of Na leads to a decrease in the tetragonal unit cell parameters a and c. However, it is a specific feature of the scheelite-type crystal structure under study that this decrease is rather tiny for x = 0.05 − 0.2, and at x = 0.3, a much stronger decrease in the unit cell size along the c axis is observed, indicating a more prominent distortion of the cell and of local structural elements within it. Specifically, all large ions mentioned above in the scheelite structure are coordinated by eight oxygen ions forming polyhedra with the local symmetry S4. The distances between a large cation and oxygen are not all equal, but take two unequal values, However, for the crystal under study with Li content x = 0, the difference between these values (2.484 Å) is below the accuracy of measurement by XRD. At x = 0.3, the Me–O distance values are noticeably different, being 2.44 and 2.53 Å, indicating a distortion of the large cation polyhedron. An additional factor that may be at work in the case of the scheelite structure is the noticeable electric field due to the stochastic cation distribution over equivalent sites. This factor may lead to variability of the cation distribution characteristics upon the introduction of Li ions. The existence of several possible mechanisms working as the Li content is varied produces, in our opinion, the complex dependence of the UC luminescence in Figure 9. The similarity of this dependency for both green and red bands means, as we suggest, that the introduction of Li influences the two-stage UC excitation channel, while the de-excitation channel probability variations play a less important role, resulting in minor changes in the relative intensities of red and green bands.
It is interesting to compare the behavior of LiNCGM:HY studied in this paper and that of LiNCLM:HY [67] at a different Li content. In the case of LiNCLM:HY, the green upconversion band experiences two maxima, while the red upconversion band experiences a single maximum at the Li content x = 0.2. The explanation for the different behavior of the two hosts arises from the different ionic radii of La and Gd. Ho and Yb ions experience, in general, a stronger crystal field when they occupy the Gd position than when they occupy the La position in the crystal structure of the host. As a result, the LiNCGM:HY behavior is more complex in the case of varied Li content.
The CIE diagram and individual chromaticity points (ICP) for (x, y) of LiNCGM:HY phosphors are shown in Figure 10. The ICP of the CIE for samples (a), (b), (c), and (d) are exhibited by the legend in Figure 10A. The calculated values for chromaticity coordinates are x = 0.374 and y = 0.409 for (a), x = 0.486 and y = 0.368 for (b), x = 0.431 and y = 0.522 for (c), and x = 0.386, and y = 0.428 for (d), corresponding to the equal-energy point in the standard CIE diagram. As can be seen, LiNCGM:HY phosphors provide emission in the yellowish region.

4. Conclusions

New MAS-derived triple molybdate LiNCGM:HY phosphors under the variation of LixNa1-x (x = 0, 0.05, 0.1, 0.2, 0.3, 0.4) were successfully manufactured. The resultant particles after annealing at 800 °C for 16 h provided well-crystallized ST tetragonal phases with particles of size 3–10 μm. The crystal structures of LiNCGM:HY phosphors at room temperature were determined in space group I41/a by Rietveld analysis. NCGM:HY has a scheelite-type structure with cell parameters a = 5.24782 (11) and c = 11.5107 (3) Å, V = 317.002 (17) Å3, Z = 4. In doped samples, the sites are occupied by a mixture of (Li,Na,Gd,Ho,Yb) ions, and this provides a linear cell volume decrease with increasing doping level. Under the excitation derived from 980 nm, the final particles led to the formation of yellow emissions based on the 5S2/ 5F45I8 green emission and the 5F55I8 red emission. The incorporated Li+ ions gave rise to local symmetry distortion around the cations in the substituted crystal structure by the Ho3+ and Yb3+ ions, and further affected the UC transition probabilities in the quadruple molybdate of LiNCGM:HY. The Raman spectra of LiNCGM doped with Ho3+ and Yb3+ ions were totally covered by the luminescence signal of Ho3+ ions, and increasing the Li content resulted in a difference in the multiplet Ho3+ intensity. Variation of Raman line positions was observed in the range of MoO4 bending vibrations depending on the Li content.

Supplementary Materials

The following are available online, Figure S1: Difference Rietveld plots of LiNCGM:HY-0.05, Figure S2: Difference Rietveld plots of LiNCGM:HY-0.1, Figure S3: Difference Rietveld plots of LiNCGM:HY-0.2, Figure S4: Variations of integral intensity and intensity of specific UC bands upon Li content, Table S1: Fractional atomic coordinates and isotropic displacement parameters (Å2) of LiNCGM:HY samples, Table S2: Main bond lengths (Å) of LiNCGM:HY samples, Table S3: Main parameters of processing and refinement of the NaCaGd(MoO4)3:xEr,yYb samples [64], Table S4: Main parameters of processing and refinement of the NaCaGd(MoO4)3:xHo,yYb samples [65], Table S5: Main parameters of processing and refinement of the LixNa1-xCaLa(MoO4)3:0.05Ho,0.45Yb samples [66], related cif files.

Author Contributions

Conceptualization, C.-S.L. and V.A.; methodology, C.-S.L.; software, M.M. and A.O.; formal analysis, A.A., M.M. and A.O.; data curation, C.-S.L.; writing—original draft preparation, C.-S.L., A.A., M.M. and A.O.; writing—review and editing, V.A. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Research Program through the Campus Research Foundation funded by Hanseo University in 2021 (211Yunghap06).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, D.; Randall, C.A.; Pang, L.X.; Wang, H.; Wu, X.G.; Guo, J.; Zhang, G.Q.; Shui, L.; Yao, X. Microwave dielectric properties of Li2(M2+)2Mo3O12 and Li3(M3+)Mo3O12 (M = Zn, Ca, Al, and In) lyonsite-related-type ceramics with ultra-low sintering temperatures. J. Am. Ceram. Soc. 2011, 94, 802–805. [Google Scholar] [CrossRef]
  2. Meert, K.W.; Morozov, V.A.; Abakumov, A.M.; Hadermann, J.; Poelman, P.; Smet, P.F. Energy transfer in Eu3+ doped scheelites: Use as thermographic phosphor. Opt. Exp. 2014, 22, A961–A972. [Google Scholar] [CrossRef]
  3. Lim, C.S.; Aleksandrovsky, A.; Molokeev, M.; Oreshonkov, A.; Atuchin, V. The modulated structure and frequency upconversion properties of CaLa2(MoO4)4:Ho3+/Yb3+ phosphors prepared by microwave synthesis. Phys. Chem. Chem. Phys. 2015, 17, 19278–19287. [Google Scholar] [CrossRef] [PubMed]
  4. Kurilchik, S.; Loiko, P.; Yasukevich, A.; Trifonov, V.; Volokitina, A.; Vilejshikova, E.; Kisel, V.; Mateos, X.; Baranov, A.; Goriev, O.; et al. Orthorombic Yb:Li2Zn2(MoO4)3—A novel potential crystal for broadly tunable lasers. Laser Phys. Lett. 2017, 14, 085804. [Google Scholar] [CrossRef]
  5. Savina, A.A.; Morozov, V.A.; Buzlukov, A.L.; Arapova, I.Y.; Stefanovich, S.Y.; Baklanova, Y.V.; Denisova, T.A.; Medvedeva, N.I.; Bardet, M.; Hadermann, J.; et al. New solid electrolyte Na9Al(MoO4)6: Structure and Na+ ion conductivity. Chem. Mater. 2017, 29, 8901–8913. [Google Scholar] [CrossRef]
  6. Solodovnikov, S.F.; Atuchin, V.V.; Solodovnikova, Z.A.; Khyzhun, O.Y.; Danylenko, M.I.; Pishchur, D.P.; Plyusnin, P.E.; Pugachev, A.M.; Gavrilova, T.A.; Yelisseyev, A.P.; et al. Synthesis, structural, thermal, and electronic properties of palmierite-related double molybdate α-Cs2Pb(MoO4). Inorg. Chem. 2017, 56, 3276–3286. [Google Scholar] [CrossRef]
  7. Sinha, S.; Kumar, K. Studies ion up/down-conversion emission of Yb3+ sensitized Er3+ doped MLa2(MoO4)4 (M = Ba, Sr and Ca) phosphors for thermometry and optical heating. Opt. Mater. 2018, 75, 770–780. [Google Scholar] [CrossRef]
  8. Goulas, A.; Chi-Tangyie, G.; Wang, D.; Zhang, S.; Ketharam, A.; Vaidhyanathan, B.; Reaney, I.M.; Cadman, D.A.; Whittow, W.G.; Vardaxoglou, J.Y.; et al. Microstructure and microwave dielectric properties of 3D printed low loss Bi2Mo2O9 ceramics for LTCC applications. Appl. Mater. Today 2020, 21, 100862. [Google Scholar] [CrossRef]
  9. Chimitova, O.D.; Bazarov, B.G.; Bazarova, J.G.; Atuchin, V.V.; Azmi, R.; Sarapulova, A.E.; Mikhailova, D.; Balachandran, G.; Fiedler, A.; Geckle, U.; et al. The crystal growth and properties of novel magnetic double molybdate RbFe5(MoO4)7 with mixed Fe3+/Fe2+ states and 1D negative thermal expansion. CrystEngComm 2021, 23, 3297–3307. [Google Scholar] [CrossRef]
  10. Li, A.; Xu, D.; Tang, Y.; Wang, Z.; Li, Z.; Zhang, Y. Solid solution Na(Gd/La)(MoO4)2:Yb3+/Er3+ upconversion nanocrystals with simultaneously enhanced photothermal conversion efficiency and luminescence intensity. J. Lumin. 2021, 239, 118356. [Google Scholar] [CrossRef]
  11. Li, K.; Deun, R.V. Mutual energy transfer luminescent properties in novel CsGd(MoO4)2:Yb3+, Er3+/Ho3+ phosphors for solid-state lighting and solar cells. Phys. Chem. Chem. Phys. 2019, 21, 4746–4754. [Google Scholar] [CrossRef]
  12. Li, A.; Xu, D.; Zhang, Y.; Lin, H.; Yang, S.; Chen, Z.; Shao, Y. Upconversion luminescence and energy-transfer mechanism of NaGd(MoO4)2:Yb3+/Er3+ microcrystals. J. Am. Ceram. Soc. 2016, 99, 1657–1663. [Google Scholar] [CrossRef]
  13. Hua, Y.; Hussain, S.K.; Yu, J.S. Samarium (iii) and terbium (iii) ion-doped NaLa(MoO4)2 phosphors for versatile applications. New J. Chem. 2019, 43, 10645–10657. [Google Scholar] [CrossRef]
  14. Atuchin, V.V.; Aleksandrovsy, A.S.; Bazarov, B.G.; Bazarova, J.G.; Chimitova, O.D.; Denisenko, Y.G.; Gavrilova, T.A.; Krylov, A.S.; Maximovskiy, E.A.; Molokeev, M.S.; et al. Exploration of structural, vibrational and spectroscopic properties of self-activated orthorhombic double molybdate RbEu(MoO4)2 with isolated MoO4 units. J. Alloy. Compd. 2019, 785, 692–697. [Google Scholar] [CrossRef] [Green Version]
  15. Zhang, Z.; Wu, Y.; Suo, H.; Zhao, X.; Guo, C. Optical thermometric properties of square-like KLa(MoO4)2:Yb3+/Er3+ up-converting phosphors. Mater. Res. Bull. 2021, 133, 111079. [Google Scholar] [CrossRef]
  16. Hu, W.; Hu, F.; Li, X.; Fang, H.; Zhao, L.; Chen, Y.; Duan, C.K.; Yin, M. Optical thermometry of a Tm3+/Yb3+ co-doped LiLa(MoO4)2 up-conversion phosphor with a high sensitivity. RSC Adv. 2016, 6, 84610–84615. [Google Scholar] [CrossRef]
  17. Haque, M.; Kim, D.-K. Luminescent properties of Eu3+ activated MLa2(MoO4)4 based (M = Ba, Sr and Ca) novel red-emitting phosphors. Mater. Lett. 2009, 63, 793–796. [Google Scholar] [CrossRef]
  18. Atuchin, V.V.; Chimitova, O.D.; Gavrilova, T.A.; Molokeev, M.S.; Kim, S.J.; Surovtsev, N.V.; Bazarov, B.G. Synthesis, structural and vibrational properties of microcrystalline RbNd(MoO4)2. J. Cryst. Growth 2011, 318, 683–686. [Google Scholar] [CrossRef]
  19. Wang, Y.; Lin, C.; Zheng, H.; Sun, D.; Li, L.; Chen, B. Fluorescent and chromatic properties of visible-emitting phosphor KLa(MoO4)2:Sm3+. J. Alloy. Compd. 2013, 559, 123–128. [Google Scholar] [CrossRef]
  20. Shi, P.; Xia, Z.; Molokeev, M.S.; Atuchin, V.V. Crystal chemistry and luminescence properties of red-emitting CsGd1−xEux(MoO4)2 solid-solution phosphors. Dalton Trans. 2014, 43, 9669–9676. [Google Scholar] [CrossRef]
  21. Lim, C.S.; Aleksandrovsky, A.S.; Molokeev, M.S.; Oreshonkov, A.S.; Atuchin, V.V. Microwave synthesis and spectroscopic properties of ternary scheelite-type molybdate phosphors NaSrLa(MoO4)3:Er3+,Yb3+. J. Alloy. Compd. 2017, 713, 156–163. [Google Scholar] [CrossRef]
  22. Li, K.; Van Deun, R. Photoluminescence and energy transfer properties of a novel molybdate KBaY(MoO4)3:Ln3+ (Ln3+ = Tb3+, Eu3+, Sm3+, Tb3+/Eu3+, Tb3+/Sm3+) as a multi-color emitting phosphor for UV w-LEDs. Dalton Trans. 2018, 47, 6995–7004. [Google Scholar] [CrossRef]
  23. Sha, X.; Chen, B.; Zhang, X.; Zhang, J.; Xu, S.; Li, X.; Sun, J.; Zhang, Y.; Wang, X.; Zhang, Y.; et al. Pre-assessments of optical transition, gain performance and temperature sensing of Er3+ in NaLn(MoO4)2 (Ln = Y, La, Gd and Lu) single crystals by using their powder-formed samples derived from traditional solid state reaction. Opt. Laser Technol. 2021, 140, 107012. [Google Scholar] [CrossRef]
  24. Abakumov, A.M.; Morozov, V.A.; Tsirlin, A.A.; Verbeeck, J.; Hadermann, J. Cation ordering and flexibility of the BO42– tetrahedra in incommensurately modulated CaEu2(BO4)4 (B = Mo, W) Scheelites. Inorg. Chem. 2014, 53, 9407–9415. [Google Scholar] [CrossRef] [PubMed]
  25. Krishnan, R.; Thirumalai, J.; Mahalingam, V.; Mantha, S.; Lavanya, M. Synthesis, luminescence and photometric characteristics of Ca0.5La(MoO4)2:Ln3+ (Ln = Eu, Tb, Dy) phosphors. Mater. Chem. Phys. 2015, 162, 41–49. [Google Scholar] [CrossRef]
  26. Moura, J.V.; Pinheiro, G.S.; Silveira, J.V.; Freire, P.T.; Viana, B.C.; Luz-Lima, C. NaCe(MoO4)2 microcrystals: Hydrothermal synthesis, characterization and photocatalytic performance. J. Phys. Chem. Solids 2017, 111, 258–265. [Google Scholar] [CrossRef]
  27. Wang, H.; Yang, T.; Feng, L.; Ning, Z.; Liu, M.; Lai, X.; Gao, D.; Bi, J. Energy Transfer and Multicolor Tunable Luminescence Properties of NaGd0.5Tb0.5− xEux(MoO4)2 Phosphors for UV-LED. J. Electron. Mater. 2018, 47, 6494–6506. [Google Scholar] [CrossRef]
  28. Tokarev, M.G.; Potanina, E.A.; Orlova, A.I.; Khainakov, S.A.; Boldin, M.S.; Lantsev, E.A.; Sakharov, N.V.; Murashov, A.A.; Garcia-Granda, S.; Nokhrin, A.V.; et al. Thermal expansion of scheelite-like molybdate powders and ceramics. Inorg. Mater. 2019, 55, 730–736. [Google Scholar] [CrossRef]
  29. Lim, C.S.; Aleksandrovsky, A.; Atuchin, V.; Molokeev, M.; Oreshonkov, A. Microwave-Employed Sol–Gel Synthesis of Scheelite-Type Microcrystalline AgGd(MoO4)2:Yb3+/Ho3+ Upconversion Yellow Phosphors and Their Spectroscopic Properties. Crystals 2020, 10, 1000. [Google Scholar]
  30. Xu, Z.; Li, S.; Meng, Q.; Wang, X.; Zhu, Q.; Li, J.G. Generalized synthesis of NaLn(MoO4)2 nano/microcrystals (Ln = La–Lu and Y): The effects of lanthanide contraction, structure, and down-/up-conversion luminescence. J. Alloy. Compd. 2020, 830, 154676. [Google Scholar] [CrossRef]
  31. Lim, C.S.; Aleksandrovsky, A.S.; Molokeev, M.S.; Oreshonkov, A.S.; Ikonnikov, D.A.; Atuchin, V.V. Triple molybdate scheelite-type upconversion phosphor NaCaLa(MoO4)3:Er3+/Yb3+: Structural and spectroscopic properties. Dalton Trans. 2015, 45, 15541–15551. [Google Scholar] [CrossRef] [Green Version]
  32. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  33. Zhai, Y.; Li, J.; Li, X.; Dong, Y.; Wang, Y.; Song, S. Synthesis and luminescent properties of NaLa(MoO4)2:Eu3+,Tb3+ phosphors by microwave-assisted sol–gel method. J. Sol-Gel Sci. Technol. 2015, 74, 544–549. [Google Scholar] [CrossRef]
  34. Lim, C.S.; Atuchin, V.V.; Aleksandrovsky, A.S.; Molokeev, M.S. Preparation of NaSrLa(WO4)3:Ho3+/Yb3+ ternary tungstates and their upconversion photoluminescence properties. Mater. Lett. 2016, 181, 38–41. [Google Scholar] [CrossRef]
  35. Wu, J.; Du, S.; Wang, Y. Photosensitizer coated upconversion nanoparticles for triggering reactive oxygen species under 980 nm near-infrared excitation. J. Mater. Chem. B 2019, 7, 7306–7313. [Google Scholar] [CrossRef] [PubMed]
  36. Lim, C.S.; Atuchin, V.V.; Aleksandrovsky, A.S.; Denisenko, Y.G.; Molokeev, M.S.; Oreshonkov, A.S. Fabrication of microcrystalline NaPbLa(WO4)3:Yb3+/Ho3+ phosphors and their upconversion photoluminescent characteristics. Korean J. Mater. Res. 2019, 29, 741–746. [Google Scholar] [CrossRef]
  37. Xie, H.; Chen, C.; Li, J.; He, Y.; Wang, N. Sol-gel synthesis and luminescent performance of Eu3+, Lu3+ co-doped Ca0.3Sr0.7Mo1-xWxO4 red-emitting phosphor. Inorg. Nano-Metal Chem. 2020, 51, 1297–1305. [Google Scholar] [CrossRef]
  38. Krut’ko, V.A.; Komova, M.G.; Pominova, D.V. Synthesis and luminescence of Gd2-x-yYbxEr(Ho)yGeMoO8 germanate-molibdates with scheelite structure. J. Solid State Chem. 2020, 292, 121704. [Google Scholar] [CrossRef]
  39. Xie, J.; Cheng, L.; Tang, H.; Wang, Z.; Sun, H.; Lu, L.; Mi, X.; Liu, Q.; Zhang, X. Wide range color tunability and efficient energy transfer of novel NaCaGd(WO4)3:Tb3+,Eu3+ phosphors with excellent thermal stability for pc-WLEDs. Inorg. Chem. Front. 2021, 8, 4517–4527. [Google Scholar] [CrossRef]
  40. Kenyon, A. Recent developments in rare-earth doped materials for optoelectronics. Prog. Quantum Electron. 2002, 26, 225–284. [Google Scholar] [CrossRef]
  41. Zhou, J.; Liu, Q.; Feng, W.; Sun, Y.; Li, F. Upconversion Luminescent Materials: Advances and Applications. Chem. Rev. 2015, 115, 395–465. [Google Scholar] [CrossRef] [PubMed]
  42. Lim, C.S.; Atuchin, V.V.; Aleksandrovsky, A.S.; Molokeev, M.S.; Oreshonkov, A.S. Incommensurately modulated structure and spectroscopic properties of CaGd2(MoO4)4:Ho3+/Yb3+ phosphors for up-conversion applications. J. Alloy. Compd. 2017, 695, 737–746. [Google Scholar] [CrossRef] [Green Version]
  43. Denisenko, Y.G.; Atuchin, V.V.; Molokeev, M.S.; Wang, N.; Jiang, X.; Aleksandrovsky, A.S.; Krylov, A.S.; Oreshonkov, A.S.; Sedykh, A.E.; Volkova, S.S.; et al. Negative thermal expansion in one-dimension of a new double sulfate AgHo(SO4)2 with isolated SO4 tetrahedra. J. Mater. Sci. Technol. 2021, 76, 111–121. [Google Scholar] [CrossRef]
  44. Das, S.; Reddy, A.A.; Babu, S.S.; Prakash, G.V. Tunable visible upconversion emission in Er3+/Yb3+-codoped KCaBO3 phosphors by introducing Ho3+ ions. Mater. Lett. 2014, 120, 232–235. [Google Scholar] [CrossRef]
  45. Ryu, J.H.; Yoon, J.-W.; Lim, C.S.; Shim, K.B. Microwave-assisted synthesis of barium molybdate by a citrate complex method and oriented aggregation. Mater. Res. Bull. 2005, 40, 1468–1476. [Google Scholar] [CrossRef]
  46. Tang, J.; Cheng, C.; Chen, Y.; Huang, Y. Yellow–green upconversion photoluminescence in Yb3+, Ho3+ co-doped NaLa(MoO4)2 phosphor. J. Alloy. Compd. 2014, 609, 268–273. [Google Scholar] [CrossRef]
  47. Chance, W.M.; Loye, H.-C.Z. Synthesis, structure, and optical properties of a series of quaternary oxides, K2Ba(MO4)2 (M = Cr, Mo, W). Solid State Sci. 2014, 28, 90–94. [Google Scholar] [CrossRef]
  48. Savina, A.; Atuchin, V.; Solodovnikov, S.; Solodovnikova, Z.; Krylov, A.; Maximovskiy, E.; Molokeev, M.; Oreshonkov, A.; Pugachev, A.; Khaikina, E. Synthesis, structural and spectroscopic properties of acentric triple molybdate Cs2NaBi(MoO4)3. J. Solid State Chem. 2015, 225, 53–58. [Google Scholar] [CrossRef]
  49. Lim, C.S. Highly modulated structure and upconversion photoluminescence properties of PbGd2(MoO4)4:Er3+/Yb3+ phosphors. Mater. Res. Bull. 2016, 75, 211–216. [Google Scholar] [CrossRef]
  50. Vidya, S.; Solomon, S.; Thomas, J.K. Single step combustion synthesis of nanocrystalline scheelite Ba0.5Sr0.5MoO4 for optical and LTCC applications: Its structural, optical and dielectric properties. J. Electroceramics 2016, 36, 142–149. [Google Scholar] [CrossRef]
  51. Mhiri, M.; Badri, A.; Ben Amara, M. Synthesis and crystal structure of NaMgFe(MoO4)3. Acta Crystallogr. Sect. E Crystallogr. Commun. 2016, 72, 864–867. [Google Scholar] [CrossRef] [PubMed]
  52. Atuchin, V.; Aleksandrovsky, A.; Molokeev, M.; Krylov, A.; Oreshonkov, A.; Zhou, D. Structural and spectroscopic properties of self-activated monoclinic molybdate BaSm2(MoO4)4. J. Alloy. Compd. 2017, 729, 843–849. [Google Scholar] [CrossRef] [Green Version]
  53. Gül, G.Ç.; Kurtuluş, F. Cr and Co doped Cs5Bi(MoO4)4: Microwave assisted synthesis, characterization and specification of optical properties. Optik 2017, 132, 153–163. [Google Scholar]
  54. Shi, T.; Zhang, F.; Li, Y.; Gao, L.; Yang, Z.; Pan, S. Structural Diversity of Molybdate Iodate and Fluoromolybdate: Syntheses, Structures, and Calculations on Na3(MoO4)(IO3) and Na3Cs(MoO2F4)2. Inorg. Chem. 2020, 59, 3034–3041. [Google Scholar] [CrossRef]
  55. Jendoubi, I.; Ptak, M.; Zid, M.F. Optical properties and visible-light-driven photocatalytic activity of Na2Mn2(MoO4)3. J. Solid State Chem. 2021, 300, 122245. [Google Scholar] [CrossRef]
  56. Rybakov, K.I.; Olevsky, E.A.; Krikun, E.V. Microwave Sintering: Fundamentals and Modeling. J. Am. Ceram. Soc. 2013, 96, 1003–1020. [Google Scholar] [CrossRef]
  57. Kitchen, H.J.; Vallance, S.R.; Kennedy, J.L.; Tapia-Ruiz, N.; Carassiti, L.; Harrison, A.; Whittaker, A.G.; Drysdale, T.D.; Kingman, S.W.; Gregory, D.H. Modern microwave methods in solid-state inorganic materials chemistry: From fundamentals to manufacturing. Chem. Rev. 2014, 114, 1170–1206. [Google Scholar] [CrossRef] [PubMed]
  58. Ryu, J.H.; Yoon, J.W.; Lim, C.S.; Oh, W.C.; Shim, K.B. Microwave-assisted synthesis of nanocrystalline MWO4 (M: Ca, Ni) via water-based citrate complex precursor. Ceram. Int. 2005, 31, 883–888. [Google Scholar] [CrossRef]
  59. Lim, C.S. Upconversion photoluminescence properties of SrY2(MoO4)4:Er3+/Yb3+ phosphors synthesized by a cyclic microwave-modified sol-gel method. Inf. Phys. Technol. 2014, 67, 371–376. [Google Scholar] [CrossRef]
  60. Ryu, J.H.; Koo, S.-M.; Chang, D.S.; Yoon, J.-W.; Lim, C.S.; Shim, K.B. Microwave-assisted synthesis of PbWO4 nano-powders via a citrate complex precursor and its photoluminescence. Ceram. Int. 2006, 32, 647–652. [Google Scholar] [CrossRef]
  61. Das, S.; Som, S.; Yang, C.Y.; Lu, C.H. Optical temperature sensing properties of SnO2:Eu3+ microspheres prepared via the microwave assisted solvothermal process. Mater. Res. Bull. 2018, 97, 101–108. [Google Scholar] [CrossRef]
  62. Shakhova, I.; Rozova, M.G.; Burova, D.; Filimonov, D.S.; Drozhzhin, O.A.; Abakumov, A.M. Microwave-assisted hydrothermal synthesis, structure and electrochemical properties of the Na3V2-yFeyO2x(PO4)2F3-2x electrode materials for Na-ion batteries. J. Solid State Chem. 2020, 281, 121010. [Google Scholar] [CrossRef]
  63. Abraham, M.; Kunti, A.K.; Thejas, K.K.; Amador-Mendez, N.; Gogneau, N.; Nishanth, K.G.; Tchernycheva, M.; Das, S. The elevated colour rendering of white-LEDs by microwave-synthesized red-emitting (Li, Mg)3RbGe8O18:Mn4+ nanophosphors. Dalton Trans. 2021, 50, 3044–3059. [Google Scholar] [CrossRef]
  64. Lim, C.S. Microwave sol–gel derived NaCaGd(MoO4)3:Er3+/Yb3+ phosphors and their upconversion photoluminescence properties. Infrared Phys. Technol. 2016, 76, 353–359. [Google Scholar] [CrossRef]
  65. Lim, C.S. Microwave sol-gel derived Ho3+/Yb3+ co-doped NaCaGd(MoO4)3 phosphors and their upconversion photoluminescence. J. Korean Ceram. Soc. 2016, 53, 456–462. [Google Scholar] [CrossRef] [Green Version]
  66. Lim, C.S.; Aleksandrovsky, A.S.; Atuchin, V.V.; Molokeev, M.S.; Oreshonkov, A.S. Microwave sol-gel synthesis, microstructural and spectroscopic properties of scheelite-type ternary molybdate upconversion phosphor NaPbLa(MoO4)3:Er3+/Yb3+. J. Alloy. Compd. 2020, 826, 152095. [Google Scholar] [CrossRef] [Green Version]
  67. Lim, C.S. Synthesis of Microcrystalline LiNaCaLa(MoO4)3:Yb3+/Ho3+ Upconversion Phosphors and Effect of Li+ on Their Spectroscopic Properties. Trans. Electr. Electron. Mater. 2019, 20, 60–66. [Google Scholar] [CrossRef]
  68. Bruker AXS. TOPAS V4: General Profile and Structure Analysis Software for Powder Diffraction Data—User’s Manual; Bruker AXS: Karlsruhe, Germany, 2008. [Google Scholar]
  69. Hazen, R.M.; Finger, L.W.; Mariathasan, J.W. High-pressure crystal chemistry of scheelite-type tungstates and molybdates. J. Phys. Chem. Solids 1985, 46, 253–263. [Google Scholar] [CrossRef]
  70. Daturi, M.; Busca, G.; Borel, M.M.; Leclaire, A.; Piaggio, P. Vibrational and XRD Study of the System CdWO4−CdMoO4. Solid State Chem. 1997, 101, 4358–4369. [Google Scholar] [CrossRef]
  71. Atuchin, V.; Gavrilova, T.; Grivel, J.-C.; Kesler, V. Electronic structure of layered titanate Nd2Ti2O7. Surf. Sci. 2008, 602, 3095–3099. [Google Scholar] [CrossRef]
  72. Atuchin, V.; Bazarov, B.; Gavrilova, T.; Grossman, V.; Molokeev, M.; Bazarova, Z. Preparation and structural properties of nonlinear optical borates K2(1−x)Rb2xAl2B2O7, 0 < x < 0.75. J. Alloy. Compd. 2012, 515, 119–122. [Google Scholar]
  73. Galashov, E.N.; Atuchin, V.V.; Gavrilova, T.A.; Korolkov, I.V.; Mandrik, Y.M.; Yelisseyev, A.P.; Xia, Z.G. Synthesis of Y3Al5O12:Ce3+ phosphor in the Y2O3-Al metal-CeO2 ternary system. J. Mater. Sci. 2017, 52, 13033–13039. [Google Scholar] [CrossRef]
  74. Kroumova, E.; Aroyo, M.I.; Mato, J.M.P.; Kirov, A.; Capillas, C.; Ivantchev, S.; Wondratschek, H. Bilbao Crystallographic Server: Useful Databases and Tools for Phase-Transition Studies. Phase Transit. 2003, 76, 155–170. [Google Scholar] [CrossRef]
  75. Thongtem, T.; Kungwankunakorn, S.; Kuntalue, B.; Phuruangrat, A.; Thongtem, S. Luminescence and absorbance of highly crystalline CaMoO4, SrMoO4, CaWO4 and SrWO4 nanoparticles synthesized by co-precipitation method at room temperature. J. Alloy. Compd. 2010, 506, 475–481. [Google Scholar] [CrossRef]
  76. Cheng, Q.; Sui, J.; Cai, W. Enhanced upconversion emission in Yb3+ and Er3+ codoped NaGdF4 nanocrystals by introducing Li+ ions. Nanoscale 2012, 4, 779–784. [Google Scholar] [CrossRef] [PubMed]
  77. Bai, Y.; Wang, Y.; Yang, K.; Zhang, X.; Song, Y.; Wang, C. Enhanced upconverted photoluminescence in Er3+ and Yb3+ codoped ZnO nanocrystals with and without Li+ ions. Opt. Commun. 2008, 281, 5448–5452. [Google Scholar] [CrossRef]
Figure 1. Difference Rietveld plots of (a) NCGM:HY and (b) LiNCGM:HY-0.3.
Figure 1. Difference Rietveld plots of (a) NCGM:HY and (b) LiNCGM:HY-0.3.
Molecules 26 07357 g001aMolecules 26 07357 g001b
Figure 2. The crystal structure of LiNCGM:HY crystals. The unit cell is outlined. Lone atoms are omitted for clarity.
Figure 2. The crystal structure of LiNCGM:HY crystals. The unit cell is outlined. Lone atoms are omitted for clarity.
Molecules 26 07357 g002
Figure 3. The dependence of (a) cell parameter a, (b) cell parameter c, and (c) cell volume V on the averaged ion radius IR(Li,Na/Ca/Gd/Ho/Er/Yb) in LiNCGM:HY and related scheelite-type molybdates.
Figure 3. The dependence of (a) cell parameter a, (b) cell parameter c, and (c) cell volume V on the averaged ion radius IR(Li,Na/Ca/Gd/Ho/Er/Yb) in LiNCGM:HY and related scheelite-type molybdates.
Molecules 26 07357 g003aMolecules 26 07357 g003b
Figure 4. The dependence of unit cell volume V on the averaged ion radius IR(Li,Na/T/Ln) in LixNa1-xTLn(MoO4)3 scheelite-type molybdates.
Figure 4. The dependence of unit cell volume V on the averaged ion radius IR(Li,Na/T/Ln) in LixNa1-xTLn(MoO4)3 scheelite-type molybdates.
Molecules 26 07357 g004
Figure 5. Scanning electron microscopy images of the synthesized (a) NCGM, (b) LiNCM:HY-0.05, (c) LiNCGM:HY-0.1, (d) LiNCGM:HY-0.2, and (e) LiNCGM:HY-0.3 particles.
Figure 5. Scanning electron microscopy images of the synthesized (a) NCGM, (b) LiNCM:HY-0.05, (c) LiNCGM:HY-0.1, (d) LiNCGM:HY-0.2, and (e) LiNCGM:HY-0.3 particles.
Molecules 26 07357 g005
Figure 6. (a) Raman spectra of LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 (x = 0, 0.05, 0.1, 0.2, 0.3); (b) Raman spectrum decomposition in the range of MoO4 tetrahedra rotation and bending.
Figure 6. (a) Raman spectra of LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 (x = 0, 0.05, 0.1, 0.2, 0.3); (b) Raman spectrum decomposition in the range of MoO4 tetrahedra rotation and bending.
Molecules 26 07357 g006
Figure 7. Raman shift of bands related to antisymmetric (a) and symmetric (b) bending and free rotation (c) of MoO4 tetrahedra in LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 (x = 0, 0.05, 0.1, 0.2, 0.3).
Figure 7. Raman shift of bands related to antisymmetric (a) and symmetric (b) bending and free rotation (c) of MoO4 tetrahedra in LixNa1-xCaGd0.5(MoO4)3:Ho3+0.05/Yb3+0.45 (x = 0, 0.05, 0.1, 0.2, 0.3).
Molecules 26 07357 g007aMolecules 26 07357 g007b
Figure 8. Antisymmetric (left) and symmetric (right) bending modes of MoO4 tetrahedra.
Figure 8. Antisymmetric (left) and symmetric (right) bending modes of MoO4 tetrahedra.
Molecules 26 07357 g008
Figure 9. The UC photoluminescence spectra of (a) NCGM:HY, (b) LiNCGM:HY-0.05, (c) LiNCGM:HY-0.1, (d) LiNCGM:HY-0.2, and (e) LiNCGM:HY-0.3 particles excited under 980 nm at room temperature.
Figure 9. The UC photoluminescence spectra of (a) NCGM:HY, (b) LiNCGM:HY-0.05, (c) LiNCGM:HY-0.1, (d) LiNCGM:HY-0.2, and (e) LiNCGM:HY-0.3 particles excited under 980 nm at room temperature.
Molecules 26 07357 g009
Figure 10. (A) CIE chromaticity diagram for the LiNCGM:HY phosphors and (B) calculated chromaticity coordinate (x, y) values. The emission points for the samples are shown with the legend.
Figure 10. (A) CIE chromaticity diagram for the LiNCGM:HY phosphors and (B) calculated chromaticity coordinate (x, y) values. The emission points for the samples are shown with the legend.
Molecules 26 07357 g010
Table 1. Abbreviations used as sample notation for LixNa1-xCaGd(MoO4)3:Ho3+0.05/Yb3+0.45.
Table 1. Abbreviations used as sample notation for LixNa1-xCaGd(MoO4)3:Ho3+0.05/Yb3+0.45.
Scheme 0.Chemical Composition
NCGM:HYNaCaGd0.5Ho0.05Yb0.45(MoO4)3
LiNCGM:HY-0.05Li0.05Na0.95CaGd0.5Ho0.05Yb0.45(MoO4)3
LiNCGM:HY-0.1Li0.1Na0.9CaGd0.5Ho0.05Yb0.45(MoO4)3
LiNCGM:HY-0.2Li0.2Na0.8CaGd0.5Ho0.05Yb0.45(MoO4)3
LiNCGM:HY-0.3Li0.3Na0.7CaGd0.5Ho0.05Yb0.45(MoO4)3
Table 2. Main parameters of processing and refinement of the LixNa1-xCaGd0.5Ho0.05Yb0.45(MoO4)3 samples.
Table 2. Main parameters of processing and refinement of the LixNa1-xCaGd0.5Ho0.05Yb0.45(MoO4)3 samples.
x.0.0.05.0.1.0.2.0.3.
Sp.Gr.I41/aI41/aI41/aI41/aI41/a
a, Å5.2077 (2)5.2074 (2)5.20679 (11)5.20342 (11)5.19940 (8)
c, Å11.3657 (5)11.3615 (6)11.3597 (3)11.3508 (3)11.3043 (2)
V, Å3308.24 (3)308.09 (3)307.968 (15)307.329 (15)306.572 (12)
Z44444
interval, º10–9010–9010–9010–9010–90
No. of reflections6363636363
No. of refined parameters77777
Rwp, %19.7918.3915.4815.8215.44
Rp, %14.5813.0010.5010.9810.59
Rexp, %14.0314.3713.9313.3813.63
χ21.411.281.111.181.13
RB, %7.593.662.192.821.41
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lim, C.-S.; Aleksandrovsky, A.; Molokeev, M.; Oreshonkov, A.; Atuchin, V. Structural and Spectroscopic Effects of Li+ Substitution for Na+ in LixNa1-xCaGd0.5Ho0.05Yb0.45(MoO4)3 Scheelite-Type Upconversion Phosphors. Molecules 2021, 26, 7357. https://doi.org/10.3390/molecules26237357

AMA Style

Lim C-S, Aleksandrovsky A, Molokeev M, Oreshonkov A, Atuchin V. Structural and Spectroscopic Effects of Li+ Substitution for Na+ in LixNa1-xCaGd0.5Ho0.05Yb0.45(MoO4)3 Scheelite-Type Upconversion Phosphors. Molecules. 2021; 26(23):7357. https://doi.org/10.3390/molecules26237357

Chicago/Turabian Style

Lim, Chang-Sung, Aleksandr Aleksandrovsky, Maxim Molokeev, Aleksandr Oreshonkov, and Victor Atuchin. 2021. "Structural and Spectroscopic Effects of Li+ Substitution for Na+ in LixNa1-xCaGd0.5Ho0.05Yb0.45(MoO4)3 Scheelite-Type Upconversion Phosphors" Molecules 26, no. 23: 7357. https://doi.org/10.3390/molecules26237357

Article Metrics

Back to TopTop