ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Investigation on the Luminescence Properties of InMO4 (M = V5+, Nb5+, Ta5+) Crystals Doped with Tb3+ or Yb3+ Rare Earth Ions

  • Pablo Botella*
    Pablo Botella
    Department of Engineering Sciences and Mathematics, Luleå University of Technology, SE-97187 Luleå, Sweden
    *E-mail: [email protected]
  • Francesco Enrichi
    Francesco Enrichi
    Department of Engineering Sciences and Mathematics, Luleå University of Technology, SE-97187 Luleå, Sweden
    Department of Molecular Sciences and Nanosystems, Ca’ Foscari University of Venice, via Torino 155, 30172 Venezia, Italy
  • Alberto Vomiero
    Alberto Vomiero
    Department of Engineering Sciences and Mathematics, Luleå University of Technology, SE-97187 Luleå, Sweden
    Department of Molecular Sciences and Nanosystems, Ca’ Foscari University of Venice, via Torino 155, 30172 Venezia, Italy
  • Juan E. Muñoz-Santiuste
    Juan E. Muñoz-Santiuste
    Departamento de Física, MALTA Consolider Team, Escuela Politécnica Superior, Universidad Carlos III de Madrid, Avenida de la Universidad 30, E-28913 Leganés, Spain
  • Alka B. Garg
    Alka B. Garg
    High Pressure and Synchrotron Radiation Physics Division, Bhabha Atomic Research Centre, Mumbai 400085, India
    More by Alka B. Garg
  • Ananthanarayanan Arvind
    Ananthanarayanan Arvind
    Process Development Division, Bhabha Atomic Research Centre, Mumbai 400085, India
  • Francisco J. Manjón
    Francisco J. Manjón
    Instituto de Diseño para la Fabricación y Producción Automatizada, MALTA Consolider Team, Universitat Politècnica de València, Camí de Vera s/n, 46022 València, Spain
  • Alfredo Segura
    Alfredo Segura
    Departamento de Física Aplicada-ICMUV, Universidad de Valencia, MALTA Consolider Team, Edificio de Investigación, C. Dr. Moliner 50, 46100 Burjassot, Spain
  • , and 
  • Daniel Errandonea
    Daniel Errandonea
    Departamento de Física Aplicada-ICMUV, Universidad de Valencia, MALTA Consolider Team, Edificio de Investigación, C. Dr. Moliner 50, 46100 Burjassot, Spain
Cite this: ACS Omega 2020, 5, 5, 2148–2158
Publication Date (Web):January 30, 2020
https://doi.org/10.1021/acsomega.9b02862

Copyright © 2022 American Chemical Society. This publication is licensed under CC-BY.

  • Open Access

Article Views

1785

Altmetric

-

Citations

LEARN ABOUT THESE METRICS
PDF (3 MB)
Supporting Info (2)»

Abstract

We explore the potential of Tb- and Yb-doped InVO4, InTaO4, and InNbO4 for applications as phosphors for light-emitting sources. Doping below 0.2% barely change the crystal structure and Raman spectrum but provide optical excitation and emission properties in the visible and near-infrared (NIR) spectral regions. From optical measurements, the energy of the first/second direct band gaps was determined to be 3.7/4.1 eV in InVO4, 4.7/5.3 in InNbO4, and 5.6/6.1 eV in InTaO4. In the last two cases, these band gaps are larger than the fundamental band gap (being indirect gap materials), while for InVO4, a direct band gap semiconductor, the fundamental band gap is at 3.7 eV. As a consequence, this material shows a strong self-activated photoluminescence centered at 2.2 eV. The other two materials have a weak self-activated signal at 2.2 and 2.9 eV. We provide an explanation for the origin of these signals taking into account the analysis of the polyhedral coordination around the pentavalent cations (V, Nb, and Ta). Finally, the characteristic green (5D47FJ) and NIR (2F5/22F7/2) emissions of Tb3+ and Yb3+ have been analyzed and explained.

Introduction

ARTICLE SECTIONS
Jump To

Light-emitting diodes (LEDs) have attracted much attention in recent decades because of their properties of high brightness. (1) Lanthanide-doped oxides are suitable materials for these applications, particularly because of its robustness and flexibility for hosting different dopants. (1) Indium metal oxides with InMO4 (M = V5+, Nb5+, Ta5+) stoichiometry are included among them. These compounds are of interest not only because of its potential use as phosphors for LEDs (2,3) but also because of their ability to act as photocatalytic materials (4−6) and gas sensors. (7,8) All these applications are intimately related to the electronic band structure of the material. By modifying the electronic band structure, the optical and electronic properties of a given material can be tailored for specific applications. Several methods can be chosen to tune materials properties, such as high-pressure techniques, (9−14) ion irradiation, (15,16) and doping. (2,3,17) In particular, mechanical techniques modify material properties by deforming the lattice of the crystal. In contrast, chemical techniques, such as doping, modify the crystal structure very slightly for a doping below 1%, thus remaining the structure almost identical to the undoped sample. (2,3,17) However, the dopant, even in very small proportions, introduces localized electronic levels that have a significant impact on the electronic and optical properties. In this context, it has been demonstrated that doping InMO4 (M = V5+, Nb5+, Ta5+) materials improves their performance as photocatalysts. (18−22) Moreover, they have also shown to be good host materials for rare earth (RE) ions, being the luminescence properties useful for LEDs. (2,3,23,24)
The crystal structures of indium niobate (InNbO4) and indium tantalate (InTaO4) are isomorphic and belong to the monoclinic space group P2/c of the wolframite structure (see Figure 1, left). (9−11,25) The structure has two formulae per unit cell (Z = 2); Nb(Ta)5+ occupies the 2e sites, while In3+ occupies the 2f sites. Both Nb(Ta) and In cations feature a 6-fold coordination. In fact, both InO6 and Nb(Ta)O6 octahedral units are the building blocks of the structure because the crystal structure is constructed by edge- and corner-sharing of InO6 and Nb(Ta)O6 zigzag chains parallel to the c direction and layered in the a direction.

Figure 1

Figure 1. Crystal structure of the wolframite-type InNb(Ta)O4 host and coordination environments for Nb5+(Ta5+) and In3+ (left) and orthorhombic InVO4 host and coordination environments for V5+ and In3+ (right) (for the color code of the structure, the reader is referred to the digital version).

Indium vanadate (InVO4) crystallizes in the orthorhombic space group Cmcm (Z = 4) with In3+ and V5+ atoms occupying 4a and 4c sites, respectively. (26,27) The structure is composed of InO6 octahedral units and VO4 tetrahedral units as building blocks. InO6 octahedra are edge-sharing along the c axis, forming chains that are connected through VO4 tetrahedral units (see Figure 1, right). The octahedral units are more regular than the ones in the wolframite structures of the other compounds, and VO4 units are not linked between them.
It has been recently shown that InVO4, InNbO4, and InTaO4 are wide band gap semiconductors. (10) These studies clarified the discrepancies reported in the literature about their band gap energy (Eg) and about their band gap nature. Many of the controversies were due to a wrong assignment of the fundamental absorption edge to the absorption of light by defects. (9,10) InVO4 was shown to be a direct band gap semiconductor along the Y → Y direction with an Eg of 3.62(5) eV, whereas InNbO4 and InTaO4 are indirect semiconductors along the Y → Γ-B direction with Eg values of 3.63(5) and 3.79(5) eV, respectively. In all the compounds, states at the bottom of the conduction band (CB) are dominated by V 3d, Nb 4d, or Ta 5d, and O 2p states dominate the upper part of the valence band (VB).
Extensive experimental and theoretical works on doping InVO4, InNbO4, and InTaO4 compounds by using nonmetal and metal elements have been reported. (17−22) However, regarding RE elements, mainly three of them (Eu3+, Tm3+, and Dy3+) have been used for doping such compounds, (2,3,23,24,28−31) and only one work has been reported on Tb-doped InTaO4. (31) Besides, it has been shown that InVO4, InNbO4, and InTaO4 can be self-activated phosphors depending on the synthesis process, which can lead to modification of the morphology, pH, and M/In molar ratio and consequently of the luminescence properties. (28,31−33)
Here, we report a detailed study on the luminescence and optical properties of InMO4 (M = V5+, Nb5+, Ta5+) compounds by comparing undoped materials and materials doped with Tb3+ or Yb3+ RE ions. The self-activated luminescence of the undoped samples and the influence of the host lattice in the characteristic green and near-infrared (NIR) region emission lines of Tb3+ and Yb3+, respectively, are also studied.
Table 1. Atomic % of the Dopants in InVO4, InNbO4, and InTaO4 Compounds
dopant sampleTb3+Yb3+
InVO40.190.17
InNbO40.090.08
InTaO40.070.15

Results and Discussion

ARTICLE SECTIONS
Jump To

Structural and Vibrational Analysis

X-ray diffraction (XRD) patterns of all doped samples were collected in order to determine the crystal structure and purity and were compared to the XRD patterns of the undoped samples previously reported. (9,11,12)Table 2 and Figure 2 show the Rietveld refinement results and the XRD patterns (bars/columns under data represent the simulated XRD of undoped orthorhombic InVO4 and monoclinic InNbO4 and InTaO4), respectively. All diffraction peaks correspond to the orthorhombic InVO4 and wolframite InNbO4 and InTaO4 crystal structures. The host lattice was barely affected by the small concentration of the dopants used. Similar results have been also observed using other doping elements. (2,3,17) In these structures, In3+ has 6-fold octahedral coordination in all the samples, with an ionic radius of 0.8 Å. Considering the same valence and coordination, Tb3+ and Yb3+ have ionic radii of 0.923 and 0.868 Å, respectively. Therefore, RE ions can be assumed to occupy the In3+ sites in the InVO4, InNbO4, and InTaO4 host lattices. Results from XRD and RS are consistent with this hypothesis. Only some residual material from precursors have been observed in the case of InTaO4 doped with Yb3+ (see Table 2). Even though RE ions possess bigger ionic radii than In3+, it is observed that there is a small diminution of the lattice parameters leading to a reduction of the unit cell volume less than 1%, contrary to what would be expected (see Table 2). These variations of the lattice parameters could be ascribed to a distortion of the octahedral units when a foreign RE element is introduced into the crystal structure, which reduces the unit cell volume to accommodate the RE ions. As we will see later, these modifications will be reflected in the self-activated PL properties of the materials.

Figure 2

Figure 2. XRD patterns of InVO4, InNbO4, and InTaO4 doped samples with Tb3+ or Yb3+. Bars/columns data represent the standard ICSD charts of the undoped orthorhombic InVO4 (ICSD-237482) and the undoped monoclinic InNbO4 (ICSD-257869) and InTaO4 (ICSD-72569), respectively. The height of the bars is proportional to the theoretical intensity of the peaks. Tables indicating the index, positions, and intensities of all reflections are included in the Supporting Information.

Table 2. Unit Cell Parameters and Goodness of the Rietveld Refinement for InVO4, InNbO4, and InTaO4 Compounds and the Corresponding Doped Samples with Tb3+ or Yb3+ from our XRD Experiments and the Contribution of the Residual Precursor Materials Found on the InTaO4:Yb Sample
 lattice parameters goodness of the fit
samplea (Å)b (Å)c (Å)β (°)ΔV (%)RpRwpRexp
InVO45.7588.5306.587  3.828.845.93
InVO4:Tb5.7478.5066.563 –0.87.0312.885.55
InVO4:Yb5.7448.5016.565 –0.96.6611.625.51
InNbO44.8365.7715.14491.13 5.89.545.8
InNbO4:Tb4.8305.7585.12991.19–0.64.938.14.32
InNbO4:Yb4.8325.7605.12991.17–0.64.86.475.86
InTaO44.8265.7755.15591.37 2.925.932.89
InTaO4:Tb4.8215.7675.14891.37–0.44.015.825.18
InTaO4:Yb4.8235.7675.14991.35–0.34.756.264.68
InTaO4:YbIn2O3Ta2O5InTaO4
contribution (%)7.46.286.4
Raman measurements also support XRD observations. As can be seen in Figure 3, the Raman signal of the doped samples is similar to that of the previously reported undoped samples. (9,11,12) Although the dopants slightly modify the unit cell, no appreciable shifts or broadenings of the peaks were observed on the results. This is due to a small local disorder introduced in the crystalline network. Notice that if the dopants were located at interstitial sites, that is, not substituting indium, more important changes (likely with the appearance of new Raman modes) could have been found in the Raman spectrum. Therefore, Raman measurements support the claim that the RE atoms substitute In. Regarding the small changes in Raman frequencies, this can be related to the small unit cell volume change associated with doping. The reduction of the unit cell volume due to dopants can be seen as the effect of an external applied pressure to the material equivalent to 0.64, 1.14, and 0.7 GPa for InVO4, InNbO4 and InTaO4, respectively. (9−12) These pressures would shift the Raman modes about 3–5 cm–1; however, we have observed all the shifts to be lower than 2 cm–1, that is, within the instrumental resolution.

Figure 3

Figure 3. RS spectra of InVO4, InNbO4, and InTaO4 compounds and the corresponding doped samples with Tb3+ or Yb3+.

Thus, doping the sample with Tb3+ or Yb3+ does not modify appreciably the phonon frequencies of the compounds studied here. This also supports that RE are substituting the In3+ atoms, what excludes the possibility that they are placed in an interstitial position, which would, in principle, could give rise to local vibrational modes observable as RS peaks, such as those found in ZnO. (41)

Optical and Photoluminescence Properties

Optical reflectance and photoluminescence measurements of InVO4, InVO4:Tb, and InVO4:Yb samples are shown in Figure 4. The optical reflectance measurements for all samples are similar (see Figure 4, top), showing a broad asymmetric band feature from 2.8 to 4.6 eV with a maximum around 4.1 eV. For comparison, the calculated reflectance (R) has been also included. These data have been estimated using the calculated refractive index (n) by Mondal et al. (42) and using the Fresnel equation in the special case of normal incidence when the sample is immersed in air:
(1)

Figure 4

Figure 4. Optical reflectance (top) and PLE/PL spectra (bottom) of InVO4, InVO4:Tb, and InVO4:Yb (short dashed-dotted line corresponds to PLE data). The data were normalized for a better comparison of the emitted signals.

The trend in the calculated reflectance is in good agreement with the observed spectra. However, due to the use of two different functionals, there is displacement of band gap energy from 3.13 eV with density functional theory calculations to 4.02 eV with tight binding calculations.
The broad asymmetric reflectance band can be well described by two Gaussian functions peaking at 3.7 and 4.1 eV. According to Mondal et al., (42) this energy region is due to the direct allowed interband transitions between the valence band and conduction band states caused by the charge transfer (CT) from O(p) to V(d) atoms inside the tetrahedral units. Thus, following previous results, we attribute these maximums to the two first direct optical allowed transition of InVO4 at the Y → Y and Γ → Γ points of the Brillouin zone (BZ). (10,27,42) The first direct transition value is in very good agreement with the energy gap value determined from our optical absorption measurements (Eg = 3.6 eV) (10) and consistent with the direct band gap nature. A reflectance maximum requires a strong optical absorption onset, which is the signature of an allowed direct transition. On the contrary, in indirect transitions, the optical transitions are very weak due to the need of phonon participation to conserve the momentum. Therefore, in direct gap semiconductors, the reflectance maximum and the fundamental absorption edge are expected at a similar energy. However, in indirect gap semiconductors, the first maximum of the reflectance corresponds to the first direct allowed transition, which is at a much higher energy than the indirect transition as it will be seen for the other two materials studied here.
The photoluminescence excitation (PLE) results are similar for all samples, showing a broad band starting at 3.4 eV and peaking around 4.1 eV (see Figure 4, bottom, short dashed-dotted line). For the PL signal, a very broad band similar for all the samples is seen peaking at 2.2 eV in the visible region (see Figure 4, bottom, solid line). InVO4 is a self-activated phosphor material due to the CT inside the vanadate group VO43–. The vanadate oxoanion in a distorted tetrahedral coordination different from the ideal Td symmetry where the transitions are spin-forbidden, presenting a self-activated luminescence due to the spin–orbit interaction that makes partly allowed the transitions. (43−45) In our case, the tetrahedral vanadate presents two different bond distances to the oxygen atoms, which makes the Td symmetry degraded to the subgroup C2v, giving rise to luminescence.
The PLE spectra show that the self-activated PL is excited by band to band transitions. These transitions were attributed to the direct transition from the ground state (1A1) due to the oxygen 2p localized states to the excited Teltow (T) levels of the 3d vanadate states (see Figure 5, top). Due to the 3d metal character, the first excited level gives rise to four states, which, following the same nomenclature as in literature, are called 1T1, 1T2, 3T1, and 3T2, with a proposed level ordering as 3T13T2 < 1T1 < 1T2. (44)

Figure 5

Figure 5. Energy level diagram of InVO4 doped with Yb (top) and InNbO4 doped with Yb or Tb (ET stands for electron transfer. CR stands for cross-relaxation. Dashed arrows represent nonradiative processes, and solid arrows correspond to PL emission and excitation).

Following previous analysis, (42−45) the PLE spectra can be deconvoluted using two Gaussian functions as it can be seen in Figure 4, bottom. The Gaussian functions are peaking at 3.7 and 4.1 eV, which agree well to the direct transitions observed in the reflectance measurements. These energies are attributed to the direct transition from 1A1 to 1T2 (Ex1) and 1T1 (Ex2), respectively (see Figure 5, top). The PL spectra were also deconvoluted using two Gaussian functions peaking at 2.1 and 2.4 eV. These energy levels correspond to the radiative decay of 3T21A1 (Em1) and 3T11A1 (Em2) inside the VO43– group.
In ref (46), a modest absorption in the visible light region above 2.5 eV was observed due to the existence of oxygen vacancies and defects in the InVO4 compound. However, they interpreted the observed luminescence as a consequence of these defects, forming a donor–acceptor pair that involves a deep donor state located at ∼0.7 eV below of the conduction band and an acceptor state that is located at ∼0.3 eV above the valence band. As it is discussed above and in ref (47), the luminescence is due to the CT inside the VO4 group and not to lattice defects or impurities as color centers. Instead, these defects or impurities play a role in the luminescence efficiency and exciton lifetime because they act as trap centers. Thus, the strong luminescence, a consequence of the distorted tetrahedral vanadate, could be affected by the presence of defects such as oxygen vacancies.
A small shift (<0.1 eV) can be observed in the PLE measurements, which could be attributed to the effect of the concentration and the distortion grade of the VO43– groups, which affects the lattice parameters and influences the energy level positions of the excited states 1T1 and 1T2. (44)
Only the self-activated PL emission from the host material was observed in the visible region, without any signal from the characteristic green emission lines of Tb3+, which lay in the same spectral region (see Figure 5, bottom). However, in a previous work, (8) it has been observed that there are characteristic Eu3+ lines even when InVO4 was doped with concentrations below to 2%. Different reasons have been suggested to the suppression of the emission lines of Tb3+ such as an inefficient energy transfer from the host material to the RE ions, a significant back-transfer rate, or a loss mechanism due to the Tb–V interaction via intervalence absorption. (48,49)
For the InVO4:Yb sample, the PL signal was detected in the NIR (see Figure 4, bottom). We can attribute the PL signal in the NIR to the transitions from 2F5/2 to 2F7/2 energy levels of the Yb3+ atoms as it can be seen in the energy level diagram (see Figure 5, top) calculated in the way previously indicated using the structural data. Although the energy level scheme of Yb3+ is very simple and contains two multiplets, the 2F7/2 ground state and the 2F5/2 excited state, the electronic energy level scheme resulting from the simulation cannot be assumed as very accurate. This fact is related with the strong interaction of Yb3+ions with the lattice vibration that usually gives rise to strong vibronic sidebands. In this case, the Yb3+ NIR luminescence consists of a broad band with small shoulders being difficult to identify the electronic transitions. Despite this uncertainty and to maintain the hypothesis that Yb3+ ions are actually substituting In3+ ions, we can attribute the PL signal in the NIR to the transitions from the 2F5/2 to 2F7/2 energy level of the Yb3+ atoms as it can be seen in the energy level diagram (see Figure 5, top). The proposed PL mechanism is as follows (see Figure 5, top): the UV photons are absorbed by the VO43– groups in the host matrix (fundamental absorption from O 2p in the valence band to the V 3d levels in the conduction band generating an electron–hole pair), which transfer part of the energy to the Yb3+ ions by a nonradiative mechanism (the remaining energy is self-emitted), in which the electron–hole is captured by Yb3+ ions. The excited Yb3+ ions come back to the ground state through a radiative transition (due to the thermal motion, all An (n = 1, 2, 3) levels are populated, and consequently, emissions from all of them are expected). (50)
Figure 6 shows the reflectance (top) and photoluminescence (bottom) spectra of InNbO4, InNbO4:Tb, and InNbO4:Yb samples. The calculated reflectance is also included, as estimated from the dielectric function calculated by Li et al. (51) and using eq 1 once the refractive index was calculated using the approximation , where ε1 is the real part of the dielectric function. Despite the displacement due to the different functional used, the calculated reflectance describes well the experimental trend.

Figure 6

Figure 6. Optical reflectance (top) and PLE/PL spectra (bottom) of InNbO4, InNbO4:Tb, and InNbO4:Yb (short dasheddotted line corresponds to PLE data). The data were normalized for a better comparison of the emitted signals.

The optical reflectance spectra are similar for all the samples, showing two features around 4.7 and 5.3 eV. We identify these energies with the direct transitions from the VB to the CB at Γ → Γ and Z → Z points in the BZ, respectively. (10) These transitions yield different contributions to the dielectric function in different directions (anisotropic material), specifically for the xx component of the dielectric tensor, which show a shift to higher energies of the maximum with respect to yy and zz components. (51)
The PLE spectrum of undoped InNbO4, in the measured range, shows a PLE peak with a maximum centered at 4.6 eV. This energy matches the direct transition observed by optical reflectance, and it is due to the charge transfer from the filled oxygen p states to the empty niobate d states inside of the octahedral unit NbO43–. (23,28) However, due to the limitations of the PLE setup, we cannot observe the second maximum at a higher energy. From our previous work, (10) we know that InNbO4 is an indirect wide band gap semiconductor with the valence band maximum (VBM) at the Y point and the conduction band minimum (CBM) in a point between the Γ and B direction of the BZ. Our optical absorption measurements yield an indirect energy gap value of 3.6 eV, which is smaller than the values found by PLE and optical reflectance. This is due to the indirect gap nature of the material as discussed previously.
For undoped InNbO4, no fluorescence was found in the visible region (see Figure 6, bottom). Blasse et al. (28) reported similar results for undoped InNbO4 under UV radiation and only very weak blue emission at liquid nitrogen temperature. On the contrary, Feng et al. (52) found, in InNbO4 nanofibers and nanoparticles, a significant PL signal centered at 2.9 eV. This value well corresponds with the features found in the PLE results of the doped samples.
The PLE measurements of the doped samples exhibit several features apart from the fundamental absorption at 4.6 eV. In the case of InNbO4:Tb, the fundamental absorption also lays at the same level as the 7DJ of the Tb3+ ions where most probably overlap both absorption bands (see Figure 5, bottom). In general, the excitation of Tb3+ in the UV spectral region may have different origins. (53) It can be attributed to the charge transfer (CT) from the orbitals 2p of O2– to the 4f of Tb3+ (54,55) or to the spin-allowed transition between the 7F6 ground state and 7DJ (J = 1, 2, 3, 4, 5) multiplets (low spin 4f75d excited states of Tb3+). Similarly, the feature around 4 eV may be related to the energy level 9DJ of the Tb3+ ion. This band is due to the spin-forbidden transition between the 7F6 ground state and 9DJ (J = 3, 4, 5, 6) multiplets of the Tb3+ ion (high spin 4f75d excited states of Tb3+). (56,57) The feature at 2.9 eV, which also has been observed for the InNbO4:Yb sample, could be related to impurity states (i.e., oxygen vacancies) introduced by the doping process or to a transition in the NbO67– octahedral having an Oh symmetry that give rise to self-activated luminescence (i.e., the intrinsic luminescence of the material, which does not originate from doping) in that region. (52,58) These distorted octahedral units, with lower symmetry than that of the ideal octahedron, could be the responsible for the decrease in the lattice parameter observed by XRD measurements.
In the case of InNbO4:Yb, an additional band in the PLE spectrum is observed around 3.4 eV, which we called (E1) in the energy level diagram (see Figure 5, bottom). This band only was observed by monitoring the emission in the NIR region; thus, this level transfers all the energy to the Yb atom that in turn radiatively decays emitting the characteristic NIR lines of the Yb3+ ions. This E1 level is related to the CT band of the 2p orbital of oxygen to 4f orbital of Yb3+ ions, which would directly populate the excited 2F5/2 level (see Figure 5, bottom). (59,60)
The PL of the doped samples, besides the abovementioned feature around 2.9 eV, shows the characteristic green line emissions of Tb3+ atoms due to the 5D4 to 7FJ (J = 0, 1, 2, 3, 4, 5, 6) transitions (see Figure 5, bottom). In this case, the results obtained from the energy level simulation in highly satisfactory. Using the calculated 5D4 and 7FJ energy level positions, we can reproduce the position and width (due to the overlapping of the different transitions) of the observed emission bands. This fact strongly supports the hypothesis that Tb3+ are in substitutional configuration in the InNbO4 host matrix.
The 5D47F5 transition always has the largest probability. This fact comes from the largest values of the reduced matrix elements both the electric dipole and the magnetic dipole ones for this transition. (61) We do not observe the 5D37FJ luminescence, which is expected in the same spectral region around 2.9 eV. As commented previously, this feature is intrinsic of the material and not from the dopants. The 5D37FJ luminescence is expected to be obtained when Tb excitation takes place at higher energies because the large 5D35D4 energy separation (more than six lattice phonons) that makes the nonradiative 5D35D4 relaxation highly unlikely. (62) The absence of this luminescence indicates that the ET channel (Figure 5, bottom) from the host to the Tb ions does not take place at energies above 3.0 eV. Additionally, cross-relaxation mechanisms are often argued as the main 5D35D4 depopulation channel in some other oxides with a similar Tb concentration. (63,64)
For the sample doped with Yb3+, the characteristic NIR lines were observed at the same energy as in the InVO4 matrix but showing sharper emission peaks, suggesting a different interaction with the lattice and allowing a successful simulation from our crystal field calculation. Henderson and Imbusch showed that the electron–phonon coupling modifies the 4f electron wave-function description by introducing opposite-parity ones. (62) As a general result, the electron–phonon coupling takes part in many phenomena besides the vibronic sidebands, including shapes and widths of spectral lines and modification of the relaxation rates. (64) On this way, the different spectral shape obtained for the InNbO4 (and for InTaO4, see further in the text) samples must be strongly related with the different (richest) phonon structures obtained in the Raman spectra for these samples.
Similar results are expected in the case of InTaO4 as it has the same crystal structure as InNbO4. The reflectance (top) and photoluminescence (bottom) measurements of the InTaO4 undoped matrix and the doped samples are presented in Figure 7. As in InNbO4, the reflectance spectra exhibit two features around 5.6 and 6.1 eV, which are associated with the direct transitions at the same points in the BZ as in InNbO4, given the close similarity in the electronic band structure of both compounds. (9,10) The calculated reflectance was estimated in the same way as in the InNbO4, showing that these maximums are an intrinsic feature of the dielectric function.

Figure 7

Figure 7. Optical reflectance (top) and PLE/PL spectra (bottom) of InTaO4, InTaO4:Tb, and InTaO4:Yb (short dasheddotted line corresponds to PLE data). The data were normalized for a better comparison of the emitted signals.

In the PLE measurement of the undoped sample, no maximum was observed due to the spectral range limitation. InTaO4 is a wider indirect band gap semiconductor (3.75 eV) by comparison to InNbO4. (9,10) Thus, the direct transitions are expected to be at a higher energy than those in InNbO4. This is because the distances of Ta–O in the TaO6 octahedra are smaller than those in the NbO6 octahedra, which makes the band gap wider.
In the PLE results of doped samples, basically, the same two features as in InNbO4:Tb can be seen for InTaO4:Tb. These features are due to Tb+3 ion absorption bands, and no contribution of the host in this case is possible due to the position of the fundamental absorption band at higher energies. However, an additional band can be seen around 3.9 eV, which gives rise to a broad band luminescence in the visible region. This band can be clearly seen for InTaO4:Yb (see Figures 6 and 7 for comparison), and most probably, in the case of InTaO4:Tb, this band overlaps with the self-absorption band of the Tb+3 ion at 3.9 eV. This signal was only observed for doped samples and again can be attributed to the defects introduced by the doping process or probably the formation of an additional distorted TaO67– octahedral site with symmetry C3v as suggested by Chukova et al. (58)
Brixner et al. (31) reports that a properly prepared InTaO4 exhibits self-activated PL around 3 eV, and Zeng et al. (33) observed the PL signal for InTaO4 nanofibers and nanoparticles with a broad peak centered at 2.7 eV. In our measurements, the weak PL signal around 2.9 eV for InTaO4 (see Figure 7, bottom) is due to the same process as discussed for InNbO4. A PL broad band from 1.7 to 2.6 eV is seen for InTaO4:Yb in the visible region. This can be due to a self-activated PL of the host material associated with defects and vacancies, created by doping, as it is also present in the PL signal from InTaO4:Tb. However, this emission is not observed in InNbO4 samples with Oh symmetry. Thus, most probably in the tantalate structure, there are octahedra with C3v symmetry that generate this additional excited and emission level in the matrix. (58) Similar as in InNbO4, the characteristic green emissions of Tb3+ atoms in the visible region are also observed in InTaO4. About the NIR emissions of Yb3+, it can be noticed that the emissions are stronger in InNbO4 than that in InTaO4 when normalized to the predominant emission at 2.24 eV.
Concerning time-resolved measurements, the decay curves and lifetimes of the characteristic emission lines of Tb3+ and Yb3+ and the self-activated InVO4 PL signal are shown in Figure 8 and Table 3, respectively. All the RE3+ emissions show double exponential decay, which is frequently observed when the excitation energy is transferred from the donor MO43– (M = V, Nb, Ta) to the activator ion Tb3+ or Yb3+ in this case. (65) In nonheavily doped samples, the characteristic behavior of the luminescence decays shows a typical initial fast decay (related with the transfer mechanism and faster as higher the concentration is) followed by a long living tail (mainly associated with the pure radiative decay).

Figure 8

Figure 8. Time-resolved PL decay in the visible and NIR spectral regions for each matrix. The excitation wavelength was 310 nm, and emission wavelengths were 550 nm for visible and 995 nm for NIR.

Table 3. Lifetime Values of the Characteristic Emission Lines of Tb3+, Yb3+, and Self-Activated InVO4 Band
sampleemission (nm)B1τ1 (μs)B2τ2 (μs)τaverage (μs)
InVO4550    69
InVO4:Tb550    71
InVO4:Yb550    66
InVO4:Yb (NIR)995250025239823341997
InNbO4:Tb55079312344232210
InNbO4:Yb (NIR)9972439621659340249
InTaO4:Tb550226225755263202
InTaO4:Yb (NIR)99926321071325461241
The decay time of the self-activated PL of InVO4 shows similar values even when it is doped by RE3+ ions, indicating that the decay time of the self-activated emission of (VO4)3– is not affected by doping. This behavior has been previously observed for Sr3La(VO4)3 when changing the doping concentration of Eu3+ (no variation of the decay time was observed), and the values agree with the values reported in the literature for the self-activated emission of the VO4 tetrahedral units embedded in different structures. (43,66) This implies that the electron transfer (ET) between (VO4)3– and RE3+ ions is not owing to the cross-relaxation between 3T1, 3T21A1 of (VO4)3– and 5D47FJ of Tb3+, and the 2F5/22F7/2 of Yb3+. (43)
The characteristic lifetime emission of the Yb3+ in the NIR region (997 nm) is, in general, long-lived (around 1 ms), (67) which is consistent with our measured decay time for Yb3+ ion in the InVO4 host. However, when Yb3+ is hosted in InNbO4 or InTaO4, it shows similar values (0.2 ms) in both compounds due to the same environment of the RE ions in both isostructural compounds but with an unusual short decay time. Such a short lifetime could be attributed to the specific local environment or to a fast recombination such as defect states in these nanostructured materials. The short decay time could be a good feature for white LED application because it avoids saturation at a high excitation. (68) The lifetime of the characteristic emissions of Tb3+ in the visible region (550 nm) shows similar values in both InNbO4 and InTaO4 and is consistent with other compounds (69) due to the same environment of the RE as we commented previously for the case of Yb3+ ions. An interesting fact to explore in the future is the influence of the difference of maximum phonon energies in lifetime differences. (9,11,12)
Among the possible applications, these materials may be used as phosphors both in the visible and NIR spectral regions. In the visible region, an important parameter for the quality of the emitted color is given by the CIE chromatic coordinates. The calculated CIE (x, y) coordinates of InVO4, InNbO4:Tb, and InTaO4:Tb are shown in Figure 9, while Table 4 lists the CIE coordinates values, correlation color temperature (CCT), color rendering indices (CRI), and the color emitted from each sample. In InVO4, as it is a self-activated phosphor, they are only estimated for the undoped sample. InVO4, InNbO4:Tb, and InTaO4:Tb present yellowish-orange, green-yellow, and greenish-blue color code coordinates, respectively. Among them, self-activated InVO4 presents a good CRI of 71, which is already interesting for lighting applications. Furthermore, proper combination of the three phosphors may further improve the quality of the light emission; thus, they can be considered promising candidates for white LEDs or NIR emitting sources in the case of Yb-doped samples. The capability of converting UV photons in visible or NIR photons has also potential applications, improving the efficiency of silicon-based solar cells.

Figure 9

Figure 9. CIE diagram of InVO4, InNbO4:Tb, and InTaO4:Tb.

Table 4. Chromaticity Coordinates (CIE), Correlated Color Temperature (CCT), Color Rendering Indices (CRI), and Color Emitted for InVO4, InNbO4:Tb, and InTaO4:Tb Samples
samplexyCCT (K)CRI (%)color
InVO40.430.47356171between warm white and neutral white
InNbO4:Tb0.260.33989537overcast sky, slightly blue-green
InTaO4:Tb0.340.58537927between daylight and sunlight

Conclusions

ARTICLE SECTIONS
Jump To

Doping InVO4, InNbO4, and InTaO4 with Tb3+ or Yb3+ up to 0.2% at. concentration does not change the crystal structure and phonon frequencies of the host materials but provides peculiar optical excitation and emission properties in the visible and NIR spectral regions. The energy of the two first direct transitions was estimated in InVO4 at 3.7/4.1 eV in the Y → Y and Γ → Γ points in the BZ and in InNbO4 and InTaO4 at 4.7/5.3 and 5.6/6.1 eV in the Γ → Γ and Z → Z points in the BZ, respectively. InVO4, being a direct band gap semiconductor, showed a strong self-activated photoluminescence centered at 2.2 eV, in comparison with the indirect InNbO4 and InTaO4 semiconductors that showed weak self-activated signals at 2.2 and 2.9 eV. These signals were related to the irregular tetrahedral VO43– in InVO4 and the octahedra Nb(Ta)O43– in InNbO4 and InTaO4, respectively. The characteristic green (5D47FJ) and NIR (2F5/22F7/2) emission of Tb3+- and Yb3+-doped materials were analyzed, demonstrating to be potential candidates for applications as phosphors for white LED lighting and NIR emitting sources and improving the efficiency of silicon-based solar cells.

Experimental Section

ARTICLE SECTIONS
Jump To

Undoped InMO4 (M = V5+, Nb5+, Ta5+) powders were prepared by a solid-state reaction following the method reported in previous works. (9−12) The composition and purity of the samples were confirmed by energy-dispersive X-ray spectroscopy analysis (EDAX) using a transmission electron microscope operated at 200 kV. The crystal structure was verified by powder (XRD) measurements and Raman spectroscopy (RS) measurements at room conditions, and the results were previously published. (9,11,12)
Polycrystalline InMO4:RE (M = V5+, Nb5+, Ta5+; RE = Tb3+, Yb3+) doped samples having Tb3+ or Yb3+ concentrations below 0.2% (see Table 1) were synthesized with the solid-state reaction method using predried powders of In2O3, Nb2O5, Ta2O5, V2O5, Yb2O3, and Tb4O7 (purity of >99.9%). The low doping concentration was selected following a previous work on Eu-doped InVO4 with the aim of obtaining good luminescence properties without affecting the crystal structure of the host material. For the doped tantalate and niobate, respective binary oxides were weighed in a stoichiometric ratio, thoroughly ground in a pestle and mortar, compacted by cold pressing into cylinders of 12.5 mm in diameter and 5 mm in height, and fired at 1100 °C for 24 h in a box-type resistive furnace followed by another heat treatment at 1200 °C for 24 h. For the vanadate, the first heating was carried out at 700 °C for 24 h followed by second heat treatment at 850 °C. All the samples are in a powder form.
Compositions were confirmed by EDS. The EDS measurements were performed using an Oxford Instruments X Max 80 EDS system attached to a Philips XL30ESEM. During EDS measurements, an accelerating voltage of 30 kV was employed. Since the samples are nonconducting, the ESEM was operated in environmental mode, where surface charge buildup on the sample was neutralized using water vapor. The accuracy of the EDS measurements was ensured by measurements of standard samples of known compositions in environmental mode. Each determination is the average of 32 runs, and at least three places were analyzed on each sample.
The crystal structures, as well as the possible structural modification introduced by the dopants, have been studied by XRD and RS techniques. For XRD measurements, a laboratory-based powder XRD using a rotating-anode generator (RAG) with a Mo (λ = 0.7107 Å) anode and a MAR345 area detector was used. RS measurements were excited with the 632.8 nm line of a He-Ne gas laser using a power of 2 mW. The scattered light was collected through a 50×/0.35 objective and sent to a Horiba Jobin Yvon LabRAM HR spectrometer with an edge filter cutting Raman signals below ∼50 cm–1. The signal was dispersed by a grating of 1200 grooves/mm and detected by a thermoelectrically cooled multichannel charge-coupled device detector enabling a spectral resolution below 2 cm–1.
For optical reflectance measurements in the UV–Vis–NIR at normal incidence, an optical setup consisting of a deuterium lamp, fused silica lenses, reflecting optics objectives, and a UV–Vis spectrometer was used. (34)
Photoluminescence excitation (PLE) and emission (PL) spectra were recorded by an Edinburgh Instruments FLS980 photoluminescence spectrometer. A continuous-wave xenon lamp was used as an excitation source for steady-state measurements, coupled to a double-grating monochromator for wavelength selection. The light emitted from the sample was collected by a double-grating monochromator and recorded by a photon counting R928P photomultiplier tube cooled at −20 °C in the visible spectral region and a R5509-73 photomultiplier tube cooled at −80 °C in the NIR spectral region.
The PLE signal was measured by following the emission signal at 550 nm (2.25 eV) and in the case of samples doped with Yb3+ also at 995 nm (1.25 eV). PL emission was measured in the visible region and in the case of samples doped with Yb3+ and also in the NIR region by using 310 nm (4 eV) excitation.
Time-resolved PL emission was obtained in multichannel scaling (MCS) mode, exciting the sample by a microsecond xenon flash lamp with a pulse duration of 1–2 μs and a repetition frequency of 10 Hz by using the same PMT detectors described above. The decay time was measured for the same signals as in PLE measurements. All abovementioned measurements were carried out at room temperature.

Computational Methods

The energy level scheme of Tb3+ and Yb3+ ions inside the crystals was modeled using a parametrized one-electron Hamiltonian in the 4fn ground configuration. The usual description includes both the several superimposed atomic interactions, which generate the 2S+1LJ multiplets, and the effect of the crystal field felt by the shielded 4f shell electron, when a rare earth ion incorporates into a solid host. The crystal field reflects the local symmetry of the RE location and is responsible for the breaking down of the 2S+1LJ degeneracy giving rise to the Stark levels. The total Hamiltonian can be expressed as (35)
(2)
Most of the parameters in the atomic Hamiltonian (EAVE, F(k), ζ4f, α, β, γ, M(j), P(k), and T(r)) were fixed to previously reported values in the present calculation. (35−37) Only EAVE, the Slater F2 parameter (for Tb3+), and the spin–orbit ζ4f parameter are slightly varied to properly estimate the position and separation of the multiplets involved in the observed transitions.
The number of nonvanishing parameters in the crystal field Hamiltonian depends on the point symmetry of the rare earth site in the host. For the studied structures, and due to the similarity of their ionic radius, we can assume that the RE3+ ion replaces the In3+ ions in sites with D2 local symmetry for InMO4 (M = Ta, Nb) compounds or C2h local symmetry for InVO4 crystals. In these symmetries, the degeneracies of the 2S+1LJ multiplets are completely lifted. By appropriated selection of the crystal field quantization axis, and a subsequent suitable rotation around the z axis to get B22 = 0, both symmetries can be described by a crystal field Hamiltonian having only 14 nonvanishing parameters. The crystal field Hamiltonian takes the form
(3)
The CF parameters were calculated using a modified version of the simple overlap model (SOM) (38) that correlates the bond distance and the bond valence as in the usual bond valence model (details can be obtained from refs (14) and (39)). The crystallographic positions of the In3+ ion and its oxygen ligands obtained by ab initio calculations for every host matrix were used neglecting the small distortion of the In3+ site when occupied by a RE3+ ion. Following standard convention in the description of the crystal field interaction, the rotationally invariant crystal field strength parameter defined as (40)
(4)
has been also calculated to simplify the comparison of the crystal-field interaction in the different structures.
The obtained parameters and the energy level positions for each host and dopant ion are included in the Supporting Information. The energy level schemes obtained from these data has been used to analyze the optical spectra.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.9b02862.

  • Crystal field parameters and crystal field strength values, calculated energy level positions for Tb3+ and Yb3+ ion in InNbO4, InTaO4, and InVO4, oxygen position and relative charge for oxygen ligands in In site used to obtain the crystal field parameters, graphical representations of the In local environment, and Miller indices, d-space, dispersion angle, and simulated X-ray diffraction intensity of InVO4, InNbO4, and InTaO4 (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
  • Authors
    • Francesco Enrichi - Department of Engineering Sciences and Mathematics, Luleå University of Technology, SE-97187 Luleå, SwedenDepartment of Molecular Sciences and Nanosystems, Ca’ Foscari University of Venice, via Torino 155, 30172 Venezia, Italy
    • Alberto Vomiero - Department of Engineering Sciences and Mathematics, Luleå University of Technology, SE-97187 Luleå, SwedenDepartment of Molecular Sciences and Nanosystems, Ca’ Foscari University of Venice, via Torino 155, 30172 Venezia, ItalyOrcidhttp://orcid.org/0000-0003-2935-1165
    • Juan E. Muñoz-Santiuste - Departamento de Física, MALTA Consolider Team, Escuela Politécnica Superior, Universidad Carlos III de Madrid, Avenida de la Universidad 30, E-28913 Leganés, Spain
    • Alka B. Garg - High Pressure and Synchrotron Radiation Physics Division and , Bhabha Atomic Research Centre, Mumbai 400085, IndiaOrcidhttp://orcid.org/0000-0003-4050-8469
    • Ananthanarayanan Arvind - Process Development Division and , Bhabha Atomic Research Centre, Mumbai 400085, India
    • Francisco J. Manjón - Instituto de Diseño para la Fabricación y Producción Automatizada, MALTA Consolider Team, Universitat Politècnica de València, Camí de Vera s/n, 46022 València, SpainOrcidhttp://orcid.org/0000-0002-3926-1705
    • Alfredo Segura - Departamento de Física Aplicada-ICMUV, Universidad de Valencia, MALTA Consolider Team, Edificio de Investigación, C. Dr. Moliner 50, 46100 Burjassot, SpainOrcidhttp://orcid.org/0000-0002-9979-1302
    • Daniel Errandonea - Departamento de Física Aplicada-ICMUV, Universidad de Valencia, MALTA Consolider Team, Edificio de Investigación, C. Dr. Moliner 50, 46100 Burjassot, SpainOrcidhttp://orcid.org/0000-0003-0189-4221
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

The authors thank the financial support from the Spanish Ministerio de Ciencia, Innovación y Universidades, Spanish Research Agency (AEI), Generalitat Valenciana, and European Fund for Regional Development (ERDF, FEDER) under grants no. MAT2016-75586-C4-1/2-P, RTI2018-101020-B-I00, RED2018-102612-T (MALTA Consolier Team), and Prometeo/2018/123 (EFIMAT). P.B. acknowledges financial support from the Kempe Foundation and the Knut & Alice Wallenberg Foundation via a doctoral studentship. A.B.G. thanks the support provided by Universitat de Valencia to perform a research stay (Atracció de Talent, VLC-CAMPUS).

References

ARTICLE SECTIONS
Jump To

This article references 69 other publications.

  1. 1
    Wang, Z. L. From nanogenerators to piezotronics-A decade-long study of ZnO nanostructures. MRS Bull. 2012, 37, 814,  DOI: 10.1557/mrs.2012.186
  2. 2
    Li, Y.; Xu, S. The contribution of Eu3+ doping concentration on the modulation of morphology and luminescence properties of InVO4:Eu3+. RSC Adv. 2018, 8, 31905,  DOI: 10.1039/C8RA02716A
  3. 3
    Tang, A.; Ma, T.; Gu, L.; Zhao, Y.; Zhang, J.; Zhang, H.; Shao, F.; Zhang, H. Luminescence properties of novel red-emitting phosphor InNb1-xPxO4:Eu3+ for white light emitting-diodes. Mater. Sci.-Pol. 2015, 33, 331334,  DOI: 10.1515/msp-2015-0050
  4. 4
    Ye, J.; Zou, Z.; Arakawa, H.; Oshikiri, M.; Shimoda, M.; Matsushita, A.; Shishido, T. Correlation of crystal and electronic structures with photophysical properties of water splitting photocatalysts InMO4 (M= V5+, Nb5+, Ta5+). J. Photochem. Photobiol., A 2002, 148, 79,  DOI: 10.1016/S1010-6030(02)00074-6
  5. 5
    Oshikiri, M.; Boero, M.; Ye, J.; Zou, Z.; Kido, G. Electronic structures of promising photocatalysts InMO4 (M = V, Nb, Ta) and BiVO4 for water decomposition in the visible wavelength region. J. Chem. Phys. 2002, 117, 7313,  DOI: 10.1063/1.1507101
  6. 6
    Zou, Z.; Ye, J.; Arakawa, H. Structural properties of InNbO4 and InTaO4: correlation with photocatalytic and photophysical properties. Chem. Phys. Lett. 2000, 332, 271,  DOI: 10.1016/S0009-2614(00)01265-3
  7. 7
    Balamurugan, C.; Vijayakumar, E.; Subramania, A. Synthesis and characterization of InNbO4 nanopowder for gas sensors. Talanta 2012, 88, 115120,  DOI: 10.1016/j.talanta.2011.10.017
  8. 8
    Chen, L.; Liu, Y.; Lu, Z.; Zeng, D. Shape-controlled synthesis and characterization of InVO4 particles. J. Colloid Interface Sci. 2006, 295, 440444,  DOI: 10.1016/j.jcis.2005.09.051
  9. 9
    Errandonea, D.; Popescu, C.; Garg, A. B.; Botella, P.; Martinez-García, D.; Pellicer-Porres, J.; Rodríguez-Hernández, P.; Muñoz, A.; Cuenca-Gotor, V.; Sans, J. A. Pressure-induced phase transition and band-gap collapse in the wide-band-gap semiconductor InTaO4. Phys. Rev. B 2016, 93, 035204  DOI: 10.1103/PhysRevB.93.035204
  10. 10
    Botella, P.; Errandonea, D.; Garg, A. B.; Rodríguez-Hernandez, P.; Muñoz, A.; Achary, S. N.; Vomiero, A. High-pressure characterization of the optical and electronic properties of InVO4, InNbO4, and InTaO4. SN Appl. Sci. 2019, 1, 389,  DOI: 10.1007/s42452-019-0406-7
  11. 11
    Garg, A. B.; Errandonea, D.; Popescu, C.; Martinez-García, D.; Pellicer-Porres, J.; Rodríguez-Hernández, P.; Muñoz, A.; Botella, P.; Cuenca-Gotor, V. P.; Sans, J. A. Pressure-Driven Isostructural Phase Transition in InNbO4: In Situ Experimental and Theoretical Investigations. Inorg. Chem. 2017, 56, 5420,  DOI: 10.1021/acs.inorgchem.7b00437
  12. 12
    Errandonea, D.; Gomis, O.; García-Domene, B.; Pellicer-Porres, J.; Katari, V.; Achary, S. N.; Tyagi, A. K.; Popescu, C. New Polymorph of InVO4: A High-Pressure Structure with Six-Coordinated Vanadium. Inorg. Chem. 2013, 52, 12790,  DOI: 10.1021/ic402043x
  13. 13
    Errandonea, D.; Tu, C.; Jia, G.; Martín, I. R.; Rodríguez-Mendoza, U. R.; Lahoz, F.; Torres, M. E.; Lavín, V. Effect of pressure on the luminescence properties of Nd3+ doped SrWO4 laser crystal. J. Alloys Compd. 2008, 451, 212214,  DOI: 10.1016/j.jallcom.2007.04.180
  14. 14
    Muñoz-Santiuste, J. E.; Lavín, V.; Rodríguez-Mendoza, U. R.; Ferrer-Roca, C.; Errandonea, D.; Martínez-García, D.; Rodríguez-Hernández, P.; Muñoz, A.; Bettinelli, M. Experimental and theoretical study on the optical properties of LaVO4 crystals under pressure. Phys. Chem. Chem. Phys. 2018, 20, 27314,  DOI: 10.1039/C8CP04701D
  15. 15
    Yoo, S. H.; Kum, J. M.; Cho, S. O. Tuning the electronic band structure of PCBM by electron irradiation. Nanoscale Res. Lett. 2011, 6, 545,  DOI: 10.1186/1556-276X-6-545
  16. 16
    Tapasztó, L.; Dobrik, G.; Nemes-Incze, P.; Vertesy, G.; Lambin, P.; Biró, L. P. Tuning the electronic structure of graphene by ion irradiation. Phys. Rev. B 2008, 78, 233407,  DOI: 10.1103/PhysRevB.78.233407
  17. 17
    Shih, H.-R.; Liu, K.-T.; Teoh, L.-G.; Wei, L.-K.; Chang, Y.-S. Synthesis and photoluminescence properties of (La,Pr) co-doped InVO4 phosphor. Microelectron. Eng. 2015, 148, 1013,  DOI: 10.1016/j.mee.2015.07.007
  18. 18
    Lu, M.; Li, Q.; Zhou, C.; Zhang, C.; Shi, H. Effects of nonmetal elements doping on the electronic structures of InNbO4: first-principles calculations. Mater. Res. Express 2018, 5, 075505  DOI: 10.1088/2053-1591/aace04
  19. 19
    Song, Y.; Sun, Z.; Wu, Y.; Chai, Z.; Wang, X. Investigation of the Preferential Doping Site and Regulating on the Visible Light Response and Redox Performance for Fe- and/or La Doped InNbO4. Inorg. Chem. 2018, 57, 85588567,  DOI: 10.1021/acs.inorgchem.8b01287
  20. 20
    Rakesh, K.; Khaire, S.; Bhange, D.; Dhanasekaran, P.; Deshpande, S. S.; Awate, S. V.; Gupta, N. M. Role of doping-induced photochemical and microstructural properties in the photocatalytic activity of InVO4 for splitting of water. J. Mater. Sci. 2011, 46, 54665476,  DOI: 10.1007/s10853-011-5489-5
  21. 21
    Wetchakun, N.; Wanwaen, P.; Phanichphant, S.; Wetchakun, K. Influence of Cu doping on the visible-light-induced photocatalytic activity of InVO4. RSC Adv. 2017, 7, 13911,  DOI: 10.1039/C6RA27138C
  22. 22
    Malingowski, A. C.; Stephens, P. W.; Huq, A.; Huang, Q.; Khalid, S.; Khalifah, P. G. Substitutional Mechanism of Ni into the Wide-Band-Gap Semiconductor InTaO4 and Its Implications for Water Splitting Activity in the Wolframite Structure Type. Inorg. Chem. 2012, 51, 60966103,  DOI: 10.1021/ic202715c
  23. 23
    Su, L.; Fan, X.; Cai, G.; Jin, Z. Tunable luminescence properties and energy transfer of Tm3+, Dy3+, and Eu3+ co-activated InNbO4 phosphors for warm-white-lighting. Ceram. Int. 2016, 42, 1599416006,  DOI: 10.1016/j.ceramint.2016.07.105
  24. 24
    Tang, A.; Zhang, D. F.; Yang, L. Synthesis and luminescence properties of novel red emitting phosphor InNbO4:Eu3+ for white light emitting diodes. Russ. Chem. Bull. 2012, 61, 2172,  DOI: 10.1007/s11172-012-0304-2
  25. 25
    Errandonea, D.; Ruiz-Fuertes, J. A brief review of the effects of pressure on wolframite-type oxides. Crystals 2018, 8, 71,  DOI: 10.3390/cryst8020071
  26. 26
    Baran, E. J. Materials belonging to the CrVO4 structure type: preparation, crystal chemistry and physicochemical properties. J. Mater. Sci. 1998, 33, 2479,  DOI: 10.1023/A:1004380530309
  27. 27
    López-Moreno, S.; Rodríguez-Hernández, P.; Muñoz, A.; Errandonea, D. First-Principles Study of InVO4 under Pressure: Phase Transitions from CrVO4 to AgMnO4-Type Structure. Inorg. Chem. 2017, 56, 26972711,  DOI: 10.1021/acs.inorgchem.6b02867
  28. 28
    Blasse, G.; Bril, A. Luminescence of Phosphors Based on Host Lattices ABO4 (A is Sc, In; B is P, V, Nb). J. Chem. Phys. 1969, 50, 2974,  DOI: 10.1063/1.1671493
  29. 29
    Tang, A.; Gu, L.; Shao, F.; Liu, X.; Zhao, Y.; Chen, H.; Zhang, H. Influence of Bi3+ content on photoluminescence of InNbO4:Eu3+, Bi3+ for white light-emitting diodes. Mater. Sci.-Pol. 2017, 35, 435439,  DOI: 10.1515/msp-2017-0053
  30. 30
    Shia, Z.-R.; Chen, H.-L.; Tsai, Y.-Y.; Wu, S.; Chang, Y.-S. Synthesis and Photoluminescence Properties of InVO4:Eu3+ Phosphors Prepared using Sol–Gel Method. ECS Trans. 2010, 28, 145154,  DOI: 10.1149/1.3367220
  31. 31
    Brixner, L. H.; Chen, H.-Y. On the structural and luminescent properties of the InTa1–xNbxO4 system. Mater. Res. Bull. 1980, 15, 607612,  DOI: 10.1016/0025-5408(80)90140-3
  32. 32
    Shen, J.; Yang, H.; Shen, Q.; You, Z. Synthesis and Characterization of InVO4 Nano-materials and their Photoluminescence Properties. Procedia Eng. 2014, 94, 6470,  DOI: 10.1016/j.proeng.2013.11.043
  33. 33
    Zeng, G.-S.; Yu, J.; Zhu, H.-Y.; Liu, H.-L.; Xing, Q.-J.; Bao, S.-K.; He, S.; Zou, J.-P.; Au, C.-T. Controllable synthesis of InTaO4 catalysts of different morphologies using a versatile sol precursor for photocatalytic evolution of H2. RSC Adv. 2015, 5, 37603,  DOI: 10.1039/C5RA03638K
  34. 34
    Lacomba-Perales, R.; Ruiz-Fuertes, J.; Errandonea, D.; Martínez-García, D.; Segura, A. Optical absorption of divalent metal tungstates: Correlation between the band-gap energy and the cation ionic radius. EPL (Eur. Lett.) 2008, 83, 37002,  DOI: 10.1209/0295-5075/83/37002
  35. 35
    Carnall, W. T.; Goodman, G. L.; Rajnak, K.; Rana, R. S. A systematic analysis of the spectra of the lanthanides doped into single crystal LaF3. J. Chem. Phys. 1989, 90, 3443,  DOI: 10.1063/1.455853
  36. 36
    Gruber, J. B.; Nash, K. L.; Yow, R. M.; Sardar, D. K.; Valiev, U. V.; Uzokov, A. A.; Burdick, G. W. Spectroscopic and magnetic susceptibility analyses of the 7FJ and 5D4 energy levels of Tb3+(4f8) in TbAlO3. J. Lumin. 2008, 128, 12711284,  DOI: 10.1016/j.jlumin.2007.12.041
  37. 37
    Cascales, C.; Zaldo, C. Spectroscopic Characterization and Systematic Crystal-Field Modeling of Optically Active Rare Earth RE3+ Ions in the Bismuth Germanate BiY1-xRxGeO5 Host. Chem. Mater. 2006, 18, 37423753,  DOI: 10.1021/cm060785t
  38. 38
    Porcher, P.; Couto Dos Santos, M.; Malta, O. Relationship between phenomenological crystal field parameters and the crystal structure: The simple overlap model. Phys. Chem. Chem. Phys. 1999, 1, 397405,  DOI: 10.1039/a803807d
  39. 39
    Hernández-Rodríguez, M. A.; Muñoz-Santiuste, J. E.; Lavín, V.; Lozano-Gorrín, A. D.; Rodríguez-Hernández, P.; Muñoz, A.; Venkatramu, V.; Martín, I. R.; Rodríguez-Mendoza, U. R. High pressure luminescence of Nd3+ in YAlO3 perovskite nanocrystals. A crystal-field analysis. J. Chem. Phys. 2018, 148, 044201  DOI: 10.1063/1.5010150
  40. 40
    Chang, N. C.; Gruber, J. B.; Leavitt, R. P.; Morrison, C. A. Optical spectra, energy levels, and crystal-field analysis of tripositive rare-earth ions in Y2O3. II. Kramers ions in C2 sites. J. Chem. Phys. 1982, 76, 3877,  DOI: 10.1063/1.443530
  41. 41
    Gluba, M. A.; Nickel, N. H.; Karpensky, N. Interstitial zinc clusters in zinc oxide. Phys. Rev. B 2013, 88, 245201,  DOI: 10.1103/PhysRevB.88.245201
  42. 42
    Mondal, S.; Appalakondaiah, S.; Vaitheeswaran, G. High pressure structural, electronic, and optical properties of polymorphic InVO4 phases. J. Appl. Phys. 2016, 119, 085702  DOI: 10.1063/1.4942182
  43. 43
    Zhou, J.; Huang, F.; Xu, J.; Chen, H.; Wang, Y. Luminescence study of a self-activated and rare earth activated Sr3La(VO4). J. Mater. Chem. C 2015, 3, 30233028,  DOI: 10.1039/C4TC02783C
  44. 44
    Ronde, H.; Blasse, G. The nature of the electronic transitions of the vanadate group. J. Inorg. Nucl. Chem. 1978, 40, 215219,  DOI: 10.1016/0022-1902(78)80113-4
  45. 45
    Nakajima, T.; Isobe, M.; Tsuchiya, T.; Ueda, Y.; Manabe, T. Correlation between Luminescence Quantum Efficiency and Structural Properties of Vanadate Phosphors with Chained, Dimerized, and Isolated VO4 Tetrahedra. J. Phys. Chem. C 2010, 114, 51605167,  DOI: 10.1021/jp910884c
  46. 46
    Van de Krol, R.; Ségalini, J.; Enache, C. S. Influence of point defects on the performance of InVO4 photoanodes. J. Photonics Energy 2011, 1, 016001  DOI: 10.1117/1.3564926
  47. 47
    Nakajima, T.; Isobe, M.; Tsuchiya, T.; Ueda, Y.; Manabe, T. Photoluminescence property of vanadates M2V2O7 (M: Ba, Sr and Ca). Opt. Mater. 2010, 32, 16181621,  DOI: 10.1016/j.optmat.2010.05.021
  48. 48
    Delosh, R. G.; Tien, T. Y.; Gibbons, E. F.; Zacmanidis, P. J.; Stadler, H. L. Strong Quenching of Tb3+ Emission by Tb–V Interaction in YPO4–YVO4. J. Chem. Phys. 1970, 53, 681,  DOI: 10.1063/1.1674044
  49. 49
    Blasse, G.; Bril, A. Investigations of Tb3+- activated phosphors. Philips Res. Rep. 1967, 22, 481504
  50. 50
    Huignard, A.; Gacoin, T.; Boilot, J.-P. Synthesis and Luminescence Properties of Colloidal YVO4:Eu Phosphors. Chem. Mater. 2000, 12, 10901094,  DOI: 10.1021/cm990722t .
    Wei, X.; Huang, S.; Chen, Y.; Guo, C.; Yin, M.; Xu, W. Energy transfer mechanisms in Yb3+ doped YVO4 near-infrared downconversion phosphor. J. Appl. Phys. 2010, 107, 103107,  DOI: 10.1063/1.3425794
  51. 51
    Li, G. L.; Yin, Z. Theoretical insight into the electronic, optical and photocatalytic properties of InMO4 (M = V, Nb, Ta) photocatalysts. Phys. Chem. Chem. Phys. 2011, 13, 28242833,  DOI: 10.1039/B921143H
  52. 52
    Feng, H.; Hou, D.; Huang, Y.; Hu, X. Facile synthesis of porous InNbO4 nanofibers by electrospinning and their enhanced visible-light-driven photocatalytic properties. J. Alloys Compd. 2014, 592, 301305,  DOI: 10.1016/j.jallcom.2013.12.261
  53. 53
    Back, M.; Massari, A.; Boffelli, M.; Gonella, F.; Riello, P.; Cristofori, D.; Riccò, R.; Enrichi, F. Optical investigation of Tb3+-doped Y2O3 nanocrystals prepared by Pechini-type sol–gel process. J. Nanopart. Res. 2012, 14, 792,  DOI: 10.1007/s11051-012-0792-x
  54. 54
    Muenchausen, R. E.; Jacobsohn, L. G.; Bennett, B. L.; McKigney, E. A.; Smith, J. F.; Valdez, J. A.; Cooke, D. W. Effects of Tb doping on the photoluminescence of Y2O3:Tb nanophosphors. J. Lumin. 2007, 126, 838842,  DOI: 10.1016/j.jlumin.2006.12.004
  55. 55
    Flores-Gonzalez, M. A.; Ledoux, G.; Roux, S.; Lebbou, K.; Perriat, P.; Tillement, O. Preparing nanometer scaled Tb-doped Y2O3 luminescent powders by the polyol method. J. Solid State Chem. 2005, 178, 989997,  DOI: 10.1016/j.jssc.2004.10.029
  56. 56
    Meng, Q.; Chen, B.; Xu, W.; Yang, Y.; Zhao, X.; Di, W.; Lu, S.; Wang, X.; Sun, J.; Cheng, L.; Yu, T.; Peng, Y. Size-dependent excitation spectra and energy transfer in Tb3+-doped Y2O3 nanocrystalline. J. Appl. Phys. 2007, 102, 093505093501,  DOI: 10.1063/1.2803502
  57. 57
    Pavitra, E.; Raju, G. S. R.; Ko, Y. H.; Yu, J. S. A novel strategy for controllable emissions from Eu3+ or Sm3+ ions co-doped SrY2O4:Tb3+ phosphors. Phys. Chem. Chem. Phys. 2012, 14, 1129611307,  DOI: 10.1039/c2cp41722g
  58. 58
    Chukova, O.; Nedilko, S.; Titov, Y.; Sheludko, V. Crystallographic Features and Nature of Luminescence Centres of the Niobate and Tantalate Compounds with Layered Perovskite-Like Structure. Open Mater. Sci. J. 2018, 12, 213,  DOI: 10.2174/1874088X01812010002
  59. 59
    Van Pieterson, L.; Heeroma, M.; de Heer, E.; Meijerink, A. Charge transfer luminescence of Yb3+. J. Lumin. 2000, 91, 177193,  DOI: 10.1016/S0022-2313(00)00214-3
  60. 60
    Boulon, G. Why so deep research on Yb3+-doped optical inorganic materials?. J. Alloys Compd. 2008, 451, 111,  DOI: 10.1016/j.jallcom.2007.04.148
  61. 61
    Kuboniwa, S.; Hoshina, T. Luminescent Properties of Tb3+ in Oxygen-Dominated Compounds. J. Phys. Soc. Jpn. 1972, 32, 10591068,  DOI: 10.1143/JPSJ.32.1059
  62. 62
    Henderson, B.; Imbusch, G. F. Optical spectroscopy of inorganic solids; 1st ed., Academic press: Oxford: New York, 1989; pp 388403.
  63. 63
    Anisimov, V. A.; Dmitryuk, A. V.; Karapetyan, G. O. Kinetic study of Tb3+(5D3) luminescence in phosphate glasses. J. Appl. Spectrosc. 1985, 43, 747,  DOI: 10.1007/BF00660581
  64. 64
    May, P. S.; Sommer, K. D. Tb3+ Luminescence in Tb-Doped and Tb/Gd-Doped CsCdBr3 Crystals: 5D45D3 Cross-Relaxation Rates in Tb3+ Pair. J. Phys. Chem. A 1997, 101, 9571,  DOI: 10.1021/jp972199g
  65. 65
    Sharma, K. G.; Singh, N. Re-dispersible CaWO4:Tb3+ nanoparticles: Synthesis, characterization and photoluminescence studies. J. Lumin. 2013, 139, 98103,  DOI: 10.1016/j.jlumin.2013.02.006
  66. 66
    Sayer, M. Luminescence in the Alkali Metavanadates. Phys. Status Solidi (a) 1970, 1, 269,  DOI: 10.1002/pssa.19700010209
  67. 67
    Laversenne, L.; Guyot, Y.; Goutaudier, C.; Cohen-Adad, M. T.; Boulon, G. Optimization of spectroscopic properties of Yb3+-doped refractory sesquioxides: cubic Y2O3, Lu2O3 and monoclinic Gd2O3. Opt. Mater. 2001, 16, 475483,  DOI: 10.1016/S0925-3467(00)00095-1
  68. 68
    Bharat, L. K.; Jeon, S. K.; Krishna, K. G.; Yu, J. S. Rare-earth free self-luminescent Ca2KZn2(VO4)3 phosphors for intense white light-emitting diodes. Sci. Rep. 2017, 7, 42348,  DOI: 10.1038/srep42348
  69. 69
    Boutinaud, P.; Cavalli, E.; Bettinelli, M. Emission quenching induced by intervalence charge transfer in Pr3+- or Tb3+-doped YNbO4 and CaNb2O6. J. Phys.: Condens. Matter 2007, 19, 386230,  DOI: 10.1088/0953-8984/19/38/386230

Cited By

This article is cited by 20 publications.

  1. Alka B. Garg, David Vie, Placida Rodriguez-Hernandez, Alfonso Muñoz, Alfredo Segura, Daniel Errandonea. Accurate Determination of the Bandgap Energy of the Rare-Earth Niobate Series. The Journal of Physical Chemistry Letters 2023, 14 (7) , 1762-1768. https://doi.org/10.1021/acs.jpclett.3c00020
  2. Tarik Ouahrani, Alka B. Garg, Rekha Rao, Placida Rodríguez-Hernández, Alfonso Muñoz, Michael Badawi, Daniel Errandonea. High-Pressure Properties of Wolframite-Type ScNbO4. The Journal of Physical Chemistry C 2022, 126 (9) , 4664-4676. https://doi.org/10.1021/acs.jpcc.1c10483
  3. Sibang Liu, Gaoliang Wang, Liuyang Xu, Huanxia Jia, Xianke Sun, Honglei Yuan. Synthesis and luminescence characteristics of an efficient broadband NIR phosphor: Cr3+ activated GaTa0.5Nb0.5O4. Ceramics International 2023, 49 (20) , 33401-33406. https://doi.org/10.1016/j.ceramint.2023.04.010
  4. A. Lysak, E. Przeździecka, A. Wierzbicka, R. Jakiela, Z. Khosravizadeh, M. Szot, A. Adhikari, A. Kozanecki. Temperature dependence of the bandgap of Eu doped {ZnCdO/ZnO}30 multilayer structures. Thin Solid Films 2023, 781 , 139982. https://doi.org/10.1016/j.tsf.2023.139982
  5. Latifa Bettadj, Reda M. Boufatah, Tarik Ouahrani, Mohammed Benaissa. Stability of doped and undoped ScNbO4 compound: Insight from first principle calculations. Materials Science in Semiconductor Processing 2023, 163 , 107545. https://doi.org/10.1016/j.mssp.2023.107545
  6. A. Dwivedi, A. Roy, S. B. Rai. Photoluminescence behavior of rare earth doped self-activated phosphors ( i.e. niobate and vanadate) and their applications. RSC Advances 2023, 13 (24) , 16260-16271. https://doi.org/10.1039/D3RA00629H
  7. Amol Nande, Swati Raut, S. J. Dhoble. Charge Transfer in Rare-Earth-Doped Inorganic Materials. 2023, 31-58. https://doi.org/10.1007/978-981-99-4145-2_2
  8. Reda M. Boufatah, Tarik Ouahrani, Mohammed Benaissa. Electronic and optical properties of wolframite-type ScNbO$$_4$$: the effect of the rare-earth doping. The European Physical Journal B 2022, 95 (10) https://doi.org/10.1140/epjb/s10051-022-00427-5
  9. Ehab M. Ahmed, Amaal Mohamed, Mortady I. Youssif, Nasser A. El‐Ghamaz, May M. El‐shabaan. Effect of Y 3+ on the structural and photoluminescence properties of yttrium‐doped sodium borate glass. Luminescence 2022, 37 (9) , 1455-1464. https://doi.org/10.1002/bio.4317
  10. Qianqian Zhang, Dongjie Liu, Peipei Dang, Hongzhou Lian, Guogang Li, Jun Lin. Two Selective Sites Control of Cr 3+ ‐Doped ABO 4 Phosphors for Tuning Ultra‐Broadband Near‐Infrared Photoluminescence and Multi‐Applications. Laser & Photonics Reviews 2022, 16 (2) https://doi.org/10.1002/lpor.202100459
  11. L.L. Alves, J.S. Souza, A.F. Lima, M.V. Lalic. Electronic, optical, and photocatalytic properties of the wolframite InNbO4 and InTaO4 compounds. Optical Materials 2022, 123 , 111781. https://doi.org/10.1016/j.optmat.2021.111781
  12. Jie Zhao, Xiao Guo, Qiang He, Fei Wu, Binghua Yao. Construction of N-CQDs/InNbO4 composites for the removal of ipronidazole: Performance and degradation mechanism. Journal of Solid State Chemistry 2021, 304 , 122567. https://doi.org/10.1016/j.jssc.2021.122567
  13. Jiayu Liao, Qiudi Chen, Xiaochen Niu, Peixiong Zhang, Huiyu Tan, Fengkai Ma, Zhen Li, Siqi Zhu, Yin Hang, Qiguo Yang, Zhenqiang Chen. Energy Transfer and Cross-Relaxation Induced Efficient 2.78 μm Emission in Er3+/Tm3+: PbF2 mid-Infrared Laser Crystal. Crystals 2021, 11 (9) , 1024. https://doi.org/10.3390/cryst11091024
  14. Qi Sun, Wei Liu, Xue Xiao, Yanhua Song, Xiangting Zhang, Dan Zhang, Haifeng Zou. Ionic liquid-assisted two-phase synthesis of Lu7O6F9:Yb3+, Er3+ phosphors and their morphological control, color-tunable up-conversion luminescence and temperature sensing behavior. Ceramics International 2021, 47 (15) , 21147-21160. https://doi.org/10.1016/j.ceramint.2021.04.118
  15. Yongbin Hua, Jung‐Uk Kim, Jae Su Yu. Charge transfer band excitation of La 3 NbO 7 :Sm 3+ phosphors induced abnormal thermal quenching toward high‐sensitivity thermometers. Journal of the American Ceramic Society 2021, 104 (8) , 4065-4074. https://doi.org/10.1111/jace.17805
  16. D.L. Shruthi, G.N. Anil Kumar, A. Jagannatha Reddy. Solid solution of novel LixByGdEu (WO4)2 (B=Na, K) red phosphors: Influence of Na/K substitution on microstructures, Judd-Ofelt and luminescence properties for WLED applications. Ceramics International 2021, 47 (11) , 16342-16357. https://doi.org/10.1016/j.ceramint.2021.02.214
  17. P Botella, S López-Moreno, D Errandonea, F J Manjón, J A Sans, D Vie, A Vomiero. High-pressure characterization of multifunctional CrVO 4. Journal of Physics: Condensed Matter 2020, 32 (38) , 385403. https://doi.org/10.1088/1361-648X/ab9408
  18. Afiq Radzwan, Abdullahi Lawal, Amiruddin Shaari, Idris Muhammad Chiromawa, Summanuwa Timothy Ahams, Rashid Ahmed. First-principles calculations of structural, electronic, and optical properties for Ni-doped Sb2S3. Computational Condensed Matter 2020, 24 , e00477. https://doi.org/10.1016/j.cocom.2020.e00477
  19. Vassily A. Medvedev, Daria V. Mamonova, Ilya E. Kolesnikov, Anastasiya R. Khokhlova, Mikhail D. Mikhailov, Alina A. Manshina. Synthesis and luminescence properties of YVO4: Nd3+, Er3+ and Tm3+ nanoparticles. Inorganic Chemistry Communications 2020, 118 , 107990. https://doi.org/10.1016/j.inoche.2020.107990
  20. Daniel Errandonea. High pressure crystal structures of orthovanadates and their properties. Journal of Applied Physics 2020, 128 (4) https://doi.org/10.1063/5.0016323
  • Abstract

    Figure 1

    Figure 1. Crystal structure of the wolframite-type InNb(Ta)O4 host and coordination environments for Nb5+(Ta5+) and In3+ (left) and orthorhombic InVO4 host and coordination environments for V5+ and In3+ (right) (for the color code of the structure, the reader is referred to the digital version).

    Figure 2

    Figure 2. XRD patterns of InVO4, InNbO4, and InTaO4 doped samples with Tb3+ or Yb3+. Bars/columns data represent the standard ICSD charts of the undoped orthorhombic InVO4 (ICSD-237482) and the undoped monoclinic InNbO4 (ICSD-257869) and InTaO4 (ICSD-72569), respectively. The height of the bars is proportional to the theoretical intensity of the peaks. Tables indicating the index, positions, and intensities of all reflections are included in the Supporting Information.

    Figure 3

    Figure 3. RS spectra of InVO4, InNbO4, and InTaO4 compounds and the corresponding doped samples with Tb3+ or Yb3+.

    Figure 4

    Figure 4. Optical reflectance (top) and PLE/PL spectra (bottom) of InVO4, InVO4:Tb, and InVO4:Yb (short dashed-dotted line corresponds to PLE data). The data were normalized for a better comparison of the emitted signals.

    Figure 5

    Figure 5. Energy level diagram of InVO4 doped with Yb (top) and InNbO4 doped with Yb or Tb (ET stands for electron transfer. CR stands for cross-relaxation. Dashed arrows represent nonradiative processes, and solid arrows correspond to PL emission and excitation).

    Figure 6

    Figure 6. Optical reflectance (top) and PLE/PL spectra (bottom) of InNbO4, InNbO4:Tb, and InNbO4:Yb (short dasheddotted line corresponds to PLE data). The data were normalized for a better comparison of the emitted signals.

    Figure 7

    Figure 7. Optical reflectance (top) and PLE/PL spectra (bottom) of InTaO4, InTaO4:Tb, and InTaO4:Yb (short dasheddotted line corresponds to PLE data). The data were normalized for a better comparison of the emitted signals.

    Figure 8

    Figure 8. Time-resolved PL decay in the visible and NIR spectral regions for each matrix. The excitation wavelength was 310 nm, and emission wavelengths were 550 nm for visible and 995 nm for NIR.

    Figure 9

    Figure 9. CIE diagram of InVO4, InNbO4:Tb, and InTaO4:Tb.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 69 other publications.

    1. 1
      Wang, Z. L. From nanogenerators to piezotronics-A decade-long study of ZnO nanostructures. MRS Bull. 2012, 37, 814,  DOI: 10.1557/mrs.2012.186
    2. 2
      Li, Y.; Xu, S. The contribution of Eu3+ doping concentration on the modulation of morphology and luminescence properties of InVO4:Eu3+. RSC Adv. 2018, 8, 31905,  DOI: 10.1039/C8RA02716A
    3. 3
      Tang, A.; Ma, T.; Gu, L.; Zhao, Y.; Zhang, J.; Zhang, H.; Shao, F.; Zhang, H. Luminescence properties of novel red-emitting phosphor InNb1-xPxO4:Eu3+ for white light emitting-diodes. Mater. Sci.-Pol. 2015, 33, 331334,  DOI: 10.1515/msp-2015-0050
    4. 4
      Ye, J.; Zou, Z.; Arakawa, H.; Oshikiri, M.; Shimoda, M.; Matsushita, A.; Shishido, T. Correlation of crystal and electronic structures with photophysical properties of water splitting photocatalysts InMO4 (M= V5+, Nb5+, Ta5+). J. Photochem. Photobiol., A 2002, 148, 79,  DOI: 10.1016/S1010-6030(02)00074-6
    5. 5
      Oshikiri, M.; Boero, M.; Ye, J.; Zou, Z.; Kido, G. Electronic structures of promising photocatalysts InMO4 (M = V, Nb, Ta) and BiVO4 for water decomposition in the visible wavelength region. J. Chem. Phys. 2002, 117, 7313,  DOI: 10.1063/1.1507101
    6. 6
      Zou, Z.; Ye, J.; Arakawa, H. Structural properties of InNbO4 and InTaO4: correlation with photocatalytic and photophysical properties. Chem. Phys. Lett. 2000, 332, 271,  DOI: 10.1016/S0009-2614(00)01265-3
    7. 7
      Balamurugan, C.; Vijayakumar, E.; Subramania, A. Synthesis and characterization of InNbO4 nanopowder for gas sensors. Talanta 2012, 88, 115120,  DOI: 10.1016/j.talanta.2011.10.017
    8. 8
      Chen, L.; Liu, Y.; Lu, Z.; Zeng, D. Shape-controlled synthesis and characterization of InVO4 particles. J. Colloid Interface Sci. 2006, 295, 440444,  DOI: 10.1016/j.jcis.2005.09.051
    9. 9
      Errandonea, D.; Popescu, C.; Garg, A. B.; Botella, P.; Martinez-García, D.; Pellicer-Porres, J.; Rodríguez-Hernández, P.; Muñoz, A.; Cuenca-Gotor, V.; Sans, J. A. Pressure-induced phase transition and band-gap collapse in the wide-band-gap semiconductor InTaO4. Phys. Rev. B 2016, 93, 035204  DOI: 10.1103/PhysRevB.93.035204
    10. 10
      Botella, P.; Errandonea, D.; Garg, A. B.; Rodríguez-Hernandez, P.; Muñoz, A.; Achary, S. N.; Vomiero, A. High-pressure characterization of the optical and electronic properties of InVO4, InNbO4, and InTaO4. SN Appl. Sci. 2019, 1, 389,  DOI: 10.1007/s42452-019-0406-7
    11. 11
      Garg, A. B.; Errandonea, D.; Popescu, C.; Martinez-García, D.; Pellicer-Porres, J.; Rodríguez-Hernández, P.; Muñoz, A.; Botella, P.; Cuenca-Gotor, V. P.; Sans, J. A. Pressure-Driven Isostructural Phase Transition in InNbO4: In Situ Experimental and Theoretical Investigations. Inorg. Chem. 2017, 56, 5420,  DOI: 10.1021/acs.inorgchem.7b00437
    12. 12
      Errandonea, D.; Gomis, O.; García-Domene, B.; Pellicer-Porres, J.; Katari, V.; Achary, S. N.; Tyagi, A. K.; Popescu, C. New Polymorph of InVO4: A High-Pressure Structure with Six-Coordinated Vanadium. Inorg. Chem. 2013, 52, 12790,  DOI: 10.1021/ic402043x
    13. 13
      Errandonea, D.; Tu, C.; Jia, G.; Martín, I. R.; Rodríguez-Mendoza, U. R.; Lahoz, F.; Torres, M. E.; Lavín, V. Effect of pressure on the luminescence properties of Nd3+ doped SrWO4 laser crystal. J. Alloys Compd. 2008, 451, 212214,  DOI: 10.1016/j.jallcom.2007.04.180
    14. 14
      Muñoz-Santiuste, J. E.; Lavín, V.; Rodríguez-Mendoza, U. R.; Ferrer-Roca, C.; Errandonea, D.; Martínez-García, D.; Rodríguez-Hernández, P.; Muñoz, A.; Bettinelli, M. Experimental and theoretical study on the optical properties of LaVO4 crystals under pressure. Phys. Chem. Chem. Phys. 2018, 20, 27314,  DOI: 10.1039/C8CP04701D
    15. 15
      Yoo, S. H.; Kum, J. M.; Cho, S. O. Tuning the electronic band structure of PCBM by electron irradiation. Nanoscale Res. Lett. 2011, 6, 545,  DOI: 10.1186/1556-276X-6-545
    16. 16
      Tapasztó, L.; Dobrik, G.; Nemes-Incze, P.; Vertesy, G.; Lambin, P.; Biró, L. P. Tuning the electronic structure of graphene by ion irradiation. Phys. Rev. B 2008, 78, 233407,  DOI: 10.1103/PhysRevB.78.233407
    17. 17
      Shih, H.-R.; Liu, K.-T.; Teoh, L.-G.; Wei, L.-K.; Chang, Y.-S. Synthesis and photoluminescence properties of (La,Pr) co-doped InVO4 phosphor. Microelectron. Eng. 2015, 148, 1013,  DOI: 10.1016/j.mee.2015.07.007
    18. 18
      Lu, M.; Li, Q.; Zhou, C.; Zhang, C.; Shi, H. Effects of nonmetal elements doping on the electronic structures of InNbO4: first-principles calculations. Mater. Res. Express 2018, 5, 075505  DOI: 10.1088/2053-1591/aace04
    19. 19
      Song, Y.; Sun, Z.; Wu, Y.; Chai, Z.; Wang, X. Investigation of the Preferential Doping Site and Regulating on the Visible Light Response and Redox Performance for Fe- and/or La Doped InNbO4. Inorg. Chem. 2018, 57, 85588567,  DOI: 10.1021/acs.inorgchem.8b01287
    20. 20
      Rakesh, K.; Khaire, S.; Bhange, D.; Dhanasekaran, P.; Deshpande, S. S.; Awate, S. V.; Gupta, N. M. Role of doping-induced photochemical and microstructural properties in the photocatalytic activity of InVO4 for splitting of water. J. Mater. Sci. 2011, 46, 54665476,  DOI: 10.1007/s10853-011-5489-5
    21. 21
      Wetchakun, N.; Wanwaen, P.; Phanichphant, S.; Wetchakun, K. Influence of Cu doping on the visible-light-induced photocatalytic activity of InVO4. RSC Adv. 2017, 7, 13911,  DOI: 10.1039/C6RA27138C
    22. 22
      Malingowski, A. C.; Stephens, P. W.; Huq, A.; Huang, Q.; Khalid, S.; Khalifah, P. G. Substitutional Mechanism of Ni into the Wide-Band-Gap Semiconductor InTaO4 and Its Implications for Water Splitting Activity in the Wolframite Structure Type. Inorg. Chem. 2012, 51, 60966103,  DOI: 10.1021/ic202715c
    23. 23
      Su, L.; Fan, X.; Cai, G.; Jin, Z. Tunable luminescence properties and energy transfer of Tm3+, Dy3+, and Eu3+ co-activated InNbO4 phosphors for warm-white-lighting. Ceram. Int. 2016, 42, 1599416006,  DOI: 10.1016/j.ceramint.2016.07.105
    24. 24
      Tang, A.; Zhang, D. F.; Yang, L. Synthesis and luminescence properties of novel red emitting phosphor InNbO4:Eu3+ for white light emitting diodes. Russ. Chem. Bull. 2012, 61, 2172,  DOI: 10.1007/s11172-012-0304-2
    25. 25
      Errandonea, D.; Ruiz-Fuertes, J. A brief review of the effects of pressure on wolframite-type oxides. Crystals 2018, 8, 71,  DOI: 10.3390/cryst8020071
    26. 26
      Baran, E. J. Materials belonging to the CrVO4 structure type: preparation, crystal chemistry and physicochemical properties. J. Mater. Sci. 1998, 33, 2479,  DOI: 10.1023/A:1004380530309
    27. 27
      López-Moreno, S.; Rodríguez-Hernández, P.; Muñoz, A.; Errandonea, D. First-Principles Study of InVO4 under Pressure: Phase Transitions from CrVO4 to AgMnO4-Type Structure. Inorg. Chem. 2017, 56, 26972711,  DOI: 10.1021/acs.inorgchem.6b02867
    28. 28
      Blasse, G.; Bril, A. Luminescence of Phosphors Based on Host Lattices ABO4 (A is Sc, In; B is P, V, Nb). J. Chem. Phys. 1969, 50, 2974,  DOI: 10.1063/1.1671493
    29. 29
      Tang, A.; Gu, L.; Shao, F.; Liu, X.; Zhao, Y.; Chen, H.; Zhang, H. Influence of Bi3+ content on photoluminescence of InNbO4:Eu3+, Bi3+ for white light-emitting diodes. Mater. Sci.-Pol. 2017, 35, 435439,  DOI: 10.1515/msp-2017-0053
    30. 30
      Shia, Z.-R.; Chen, H.-L.; Tsai, Y.-Y.; Wu, S.; Chang, Y.-S. Synthesis and Photoluminescence Properties of InVO4:Eu3+ Phosphors Prepared using Sol–Gel Method. ECS Trans. 2010, 28, 145154,  DOI: 10.1149/1.3367220
    31. 31
      Brixner, L. H.; Chen, H.-Y. On the structural and luminescent properties of the InTa1–xNbxO4 system. Mater. Res. Bull. 1980, 15, 607612,  DOI: 10.1016/0025-5408(80)90140-3
    32. 32
      Shen, J.; Yang, H.; Shen, Q.; You, Z. Synthesis and Characterization of InVO4 Nano-materials and their Photoluminescence Properties. Procedia Eng. 2014, 94, 6470,  DOI: 10.1016/j.proeng.2013.11.043
    33. 33
      Zeng, G.-S.; Yu, J.; Zhu, H.-Y.; Liu, H.-L.; Xing, Q.-J.; Bao, S.-K.; He, S.; Zou, J.-P.; Au, C.-T. Controllable synthesis of InTaO4 catalysts of different morphologies using a versatile sol precursor for photocatalytic evolution of H2. RSC Adv. 2015, 5, 37603,  DOI: 10.1039/C5RA03638K
    34. 34
      Lacomba-Perales, R.; Ruiz-Fuertes, J.; Errandonea, D.; Martínez-García, D.; Segura, A. Optical absorption of divalent metal tungstates: Correlation between the band-gap energy and the cation ionic radius. EPL (Eur. Lett.) 2008, 83, 37002,  DOI: 10.1209/0295-5075/83/37002
    35. 35
      Carnall, W. T.; Goodman, G. L.; Rajnak, K.; Rana, R. S. A systematic analysis of the spectra of the lanthanides doped into single crystal LaF3. J. Chem. Phys. 1989, 90, 3443,  DOI: 10.1063/1.455853
    36. 36
      Gruber, J. B.; Nash, K. L.; Yow, R. M.; Sardar, D. K.; Valiev, U. V.; Uzokov, A. A.; Burdick, G. W. Spectroscopic and magnetic susceptibility analyses of the 7FJ and 5D4 energy levels of Tb3+(4f8) in TbAlO3. J. Lumin. 2008, 128, 12711284,  DOI: 10.1016/j.jlumin.2007.12.041
    37. 37
      Cascales, C.; Zaldo, C. Spectroscopic Characterization and Systematic Crystal-Field Modeling of Optically Active Rare Earth RE3+ Ions in the Bismuth Germanate BiY1-xRxGeO5 Host. Chem. Mater. 2006, 18, 37423753,  DOI: 10.1021/cm060785t
    38. 38
      Porcher, P.; Couto Dos Santos, M.; Malta, O. Relationship between phenomenological crystal field parameters and the crystal structure: The simple overlap model. Phys. Chem. Chem. Phys. 1999, 1, 397405,  DOI: 10.1039/a803807d
    39. 39
      Hernández-Rodríguez, M. A.; Muñoz-Santiuste, J. E.; Lavín, V.; Lozano-Gorrín, A. D.; Rodríguez-Hernández, P.; Muñoz, A.; Venkatramu, V.; Martín, I. R.; Rodríguez-Mendoza, U. R. High pressure luminescence of Nd3+ in YAlO3 perovskite nanocrystals. A crystal-field analysis. J. Chem. Phys. 2018, 148, 044201  DOI: 10.1063/1.5010150
    40. 40
      Chang, N. C.; Gruber, J. B.; Leavitt, R. P.; Morrison, C. A. Optical spectra, energy levels, and crystal-field analysis of tripositive rare-earth ions in Y2O3. II. Kramers ions in C2 sites. J. Chem. Phys. 1982, 76, 3877,  DOI: 10.1063/1.443530
    41. 41
      Gluba, M. A.; Nickel, N. H.; Karpensky, N. Interstitial zinc clusters in zinc oxide. Phys. Rev. B 2013, 88, 245201,  DOI: 10.1103/PhysRevB.88.245201
    42. 42
      Mondal, S.; Appalakondaiah, S.; Vaitheeswaran, G. High pressure structural, electronic, and optical properties of polymorphic InVO4 phases. J. Appl. Phys. 2016, 119, 085702  DOI: 10.1063/1.4942182
    43. 43
      Zhou, J.; Huang, F.; Xu, J.; Chen, H.; Wang, Y. Luminescence study of a self-activated and rare earth activated Sr3La(VO4). J. Mater. Chem. C 2015, 3, 30233028,  DOI: 10.1039/C4TC02783C
    44. 44
      Ronde, H.; Blasse, G. The nature of the electronic transitions of the vanadate group. J. Inorg. Nucl. Chem. 1978, 40, 215219,  DOI: 10.1016/0022-1902(78)80113-4
    45. 45
      Nakajima, T.; Isobe, M.; Tsuchiya, T.; Ueda, Y.; Manabe, T. Correlation between Luminescence Quantum Efficiency and Structural Properties of Vanadate Phosphors with Chained, Dimerized, and Isolated VO4 Tetrahedra. J. Phys. Chem. C 2010, 114, 51605167,  DOI: 10.1021/jp910884c
    46. 46
      Van de Krol, R.; Ségalini, J.; Enache, C. S. Influence of point defects on the performance of InVO4 photoanodes. J. Photonics Energy 2011, 1, 016001  DOI: 10.1117/1.3564926
    47. 47
      Nakajima, T.; Isobe, M.; Tsuchiya, T.; Ueda, Y.; Manabe, T. Photoluminescence property of vanadates M2V2O7 (M: Ba, Sr and Ca). Opt. Mater. 2010, 32, 16181621,  DOI: 10.1016/j.optmat.2010.05.021
    48. 48
      Delosh, R. G.; Tien, T. Y.; Gibbons, E. F.; Zacmanidis, P. J.; Stadler, H. L. Strong Quenching of Tb3+ Emission by Tb–V Interaction in YPO4–YVO4. J. Chem. Phys. 1970, 53, 681,  DOI: 10.1063/1.1674044
    49. 49
      Blasse, G.; Bril, A. Investigations of Tb3+- activated phosphors. Philips Res. Rep. 1967, 22, 481504
    50. 50
      Huignard, A.; Gacoin, T.; Boilot, J.-P. Synthesis and Luminescence Properties of Colloidal YVO4:Eu Phosphors. Chem. Mater. 2000, 12, 10901094,  DOI: 10.1021/cm990722t .
      Wei, X.; Huang, S.; Chen, Y.; Guo, C.; Yin, M.; Xu, W. Energy transfer mechanisms in Yb3+ doped YVO4 near-infrared downconversion phosphor. J. Appl. Phys. 2010, 107, 103107,  DOI: 10.1063/1.3425794
    51. 51
      Li, G. L.; Yin, Z. Theoretical insight into the electronic, optical and photocatalytic properties of InMO4 (M = V, Nb, Ta) photocatalysts. Phys. Chem. Chem. Phys. 2011, 13, 28242833,  DOI: 10.1039/B921143H
    52. 52
      Feng, H.; Hou, D.; Huang, Y.; Hu, X. Facile synthesis of porous InNbO4 nanofibers by electrospinning and their enhanced visible-light-driven photocatalytic properties. J. Alloys Compd. 2014, 592, 301305,  DOI: 10.1016/j.jallcom.2013.12.261
    53. 53
      Back, M.; Massari, A.; Boffelli, M.; Gonella, F.; Riello, P.; Cristofori, D.; Riccò, R.; Enrichi, F. Optical investigation of Tb3+-doped Y2O3 nanocrystals prepared by Pechini-type sol–gel process. J. Nanopart. Res. 2012, 14, 792,  DOI: 10.1007/s11051-012-0792-x
    54. 54
      Muenchausen, R. E.; Jacobsohn, L. G.; Bennett, B. L.; McKigney, E. A.; Smith, J. F.; Valdez, J. A.; Cooke, D. W. Effects of Tb doping on the photoluminescence of Y2O3:Tb nanophosphors. J. Lumin. 2007, 126, 838842,  DOI: 10.1016/j.jlumin.2006.12.004
    55. 55
      Flores-Gonzalez, M. A.; Ledoux, G.; Roux, S.; Lebbou, K.; Perriat, P.; Tillement, O. Preparing nanometer scaled Tb-doped Y2O3 luminescent powders by the polyol method. J. Solid State Chem. 2005, 178, 989997,  DOI: 10.1016/j.jssc.2004.10.029
    56. 56
      Meng, Q.; Chen, B.; Xu, W.; Yang, Y.; Zhao, X.; Di, W.; Lu, S.; Wang, X.; Sun, J.; Cheng, L.; Yu, T.; Peng, Y. Size-dependent excitation spectra and energy transfer in Tb3+-doped Y2O3 nanocrystalline. J. Appl. Phys. 2007, 102, 093505093501,  DOI: 10.1063/1.2803502
    57. 57
      Pavitra, E.; Raju, G. S. R.; Ko, Y. H.; Yu, J. S. A novel strategy for controllable emissions from Eu3+ or Sm3+ ions co-doped SrY2O4:Tb3+ phosphors. Phys. Chem. Chem. Phys. 2012, 14, 1129611307,  DOI: 10.1039/c2cp41722g
    58. 58
      Chukova, O.; Nedilko, S.; Titov, Y.; Sheludko, V. Crystallographic Features and Nature of Luminescence Centres of the Niobate and Tantalate Compounds with Layered Perovskite-Like Structure. Open Mater. Sci. J. 2018, 12, 213,  DOI: 10.2174/1874088X01812010002
    59. 59
      Van Pieterson, L.; Heeroma, M.; de Heer, E.; Meijerink, A. Charge transfer luminescence of Yb3+. J. Lumin. 2000, 91, 177193,  DOI: 10.1016/S0022-2313(00)00214-3
    60. 60
      Boulon, G. Why so deep research on Yb3+-doped optical inorganic materials?. J. Alloys Compd. 2008, 451, 111,  DOI: 10.1016/j.jallcom.2007.04.148
    61. 61
      Kuboniwa, S.; Hoshina, T. Luminescent Properties of Tb3+ in Oxygen-Dominated Compounds. J. Phys. Soc. Jpn. 1972, 32, 10591068,  DOI: 10.1143/JPSJ.32.1059
    62. 62
      Henderson, B.; Imbusch, G. F. Optical spectroscopy of inorganic solids; 1st ed., Academic press: Oxford: New York, 1989; pp 388403.
    63. 63
      Anisimov, V. A.; Dmitryuk, A. V.; Karapetyan, G. O. Kinetic study of Tb3+(5D3) luminescence in phosphate glasses. J. Appl. Spectrosc. 1985, 43, 747,  DOI: 10.1007/BF00660581
    64. 64
      May, P. S.; Sommer, K. D. Tb3+ Luminescence in Tb-Doped and Tb/Gd-Doped CsCdBr3 Crystals: 5D45D3 Cross-Relaxation Rates in Tb3+ Pair. J. Phys. Chem. A 1997, 101, 9571,  DOI: 10.1021/jp972199g
    65. 65
      Sharma, K. G.; Singh, N. Re-dispersible CaWO4:Tb3+ nanoparticles: Synthesis, characterization and photoluminescence studies. J. Lumin. 2013, 139, 98103,  DOI: 10.1016/j.jlumin.2013.02.006
    66. 66
      Sayer, M. Luminescence in the Alkali Metavanadates. Phys. Status Solidi (a) 1970, 1, 269,  DOI: 10.1002/pssa.19700010209
    67. 67
      Laversenne, L.; Guyot, Y.; Goutaudier, C.; Cohen-Adad, M. T.; Boulon, G. Optimization of spectroscopic properties of Yb3+-doped refractory sesquioxides: cubic Y2O3, Lu2O3 and monoclinic Gd2O3. Opt. Mater. 2001, 16, 475483,  DOI: 10.1016/S0925-3467(00)00095-1
    68. 68
      Bharat, L. K.; Jeon, S. K.; Krishna, K. G.; Yu, J. S. Rare-earth free self-luminescent Ca2KZn2(VO4)3 phosphors for intense white light-emitting diodes. Sci. Rep. 2017, 7, 42348,  DOI: 10.1038/srep42348
    69. 69
      Boutinaud, P.; Cavalli, E.; Bettinelli, M. Emission quenching induced by intervalence charge transfer in Pr3+- or Tb3+-doped YNbO4 and CaNb2O6. J. Phys.: Condens. Matter 2007, 19, 386230,  DOI: 10.1088/0953-8984/19/38/386230
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.9b02862.

    • Crystal field parameters and crystal field strength values, calculated energy level positions for Tb3+ and Yb3+ ion in InNbO4, InTaO4, and InVO4, oxygen position and relative charge for oxygen ligands in In site used to obtain the crystal field parameters, graphical representations of the In local environment, and Miller indices, d-space, dispersion angle, and simulated X-ray diffraction intensity of InVO4, InNbO4, and InTaO4 (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect