Next Article in Journal
A Critical Review on Recent Advancements in Aluminium-Based Metal Matrix Composites
Previous Article in Journal
Hydrofluoric Acid-Free Synthesis of MIL-101(Cr)-SO3H
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Defect Structures of Rare Earth-Doped Lutetium Oxide and Impacts of Li Co-Dopant

Kazuo Inamori School of Engineering, New York State College of Ceramics, Alfred University, Alfred, NY 14802, USA
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(5), 413; https://doi.org/10.3390/cryst14050413
Submission received: 26 March 2024 / Revised: 18 April 2024 / Accepted: 25 April 2024 / Published: 28 April 2024
(This article belongs to the Section Polycrystalline Ceramics)

Abstract

:
Defect complexes consisting of point defects induced by the doping of rare earth elements (Nd, Er) into lutetium oxide (Lu2O3) host were investigated with respect to defect formation energies and defect configurations using atomistic simulations with General Utility Lattice Program (GULP). The site preferences of the substitutional point defects of the dopants and the occupation between the two available cationic sites, the 8b and 24d sites, were analyzed. Additionally, the impacts of Li on the doping of rare earth elements into Lu2O3 were revealed from the viewpoints of energy and structure. Dopant pairs in the nearest neighbor configurations (8b + 8b), (8b + 24d), and (24d + 24d) were considered. The results contribute to the understanding of structures of defects in rare earth-doped Lu2O3.

1. Introduction

Oxides with cubic crystal structures have been of great research interest in the last two decades. Lutetium oxide (Lu2O3), with a C-type rare earth crystal structure (the bixbyite structure), is one of them. Properties such as high melting point, wide band gap, high density, and good thermal conductivity make it a good candidate as a solid-state laser host [1,2,3]. Nd (e.g., emission at around 1.06 μ m ) and Er (e.g., emission at around 1.5 and 3 μ m ) are recognized as ideal active laser ions for applications in the near and mid-infrared ranges. However, the main emission wavelengths may differ depending on the structure of the host. The optimal concentration of rare earth dopant in the popular laser hosts, such as Nd in YAG, is usually below 1.0 at%. One of the strong research interests in the bixbyite host is its potential capability of hosting high concentrations of rare earth elements, for example, 10 at % of Yb [4]. Experimental investigations have been continuously developing the processing, including powder synthesis and sintering, and optical properties, such as the analysis of optical spectra, to achieve high transparency. Despite the progress in fabrication routes and the well-established knowledge of the electronic structures of free rare earth ions, our understanding of the structure–property relationships in rare earth-doped bixbyite structure systems, especially with respect to optical properties, is still limited. For example, optical quenching occurs with high concentrations of doping, which limits the doping concentration to a certain level for an optimal emission intensity. According to the investigations of Nd-doped Lu2O3 by Goget [5], the emission cross-section (emission probability) of 4F3/2     4/I11/2 decreased as the concentration increased in the range of 1–5 at%. It was indicated that the dopant occupations of the two Lu sites influence the concentration measurement. Brunn [6] investigated Nd-doped single-crystal Lu2O3; the emission peak with the largest cross-section was at 1076 nm with 0.6% (1.71 × 1023 cm−3) doping concentration. Strong luminescence quenching was observed in samples with higher concentrations (1%, 3%). From the viewpoint of the crystal structure, higher concentrations of dopants will induce more structural distortions. These findings reflected a strong correlation between optical properties and defect structures in Nd3+: Lu2O3. Luminescence quenching in Nd3+: YAG has been attributed to several possible mechanisms of energy transfer and energy dissipation between nearby Nd dopants; for example, the cross-relaxation via the 4I15/2 manifold [7,8]. The splitting of energy states due to an electric field acting upon the atom is called Stark splitting [7]. So, more specifically, the luminescence quenching is related to the electronic transitions between Stark levels of the manifolds. The crystal fields are different in different hosts, making the Stark splitting of rare earth dopants in YAG and Lu2O3 different; thus, their quenching mechanisms may be different. Defects induce perturbations to the crystal field and thus to the energy states of the dopants. Therefore, investigating the local structures of defects around the dopants is essential for achieving a better understanding of the optical quenching mechanisms. Zhou [9,10] fabricated a 3 at% Nd3+: Lu2O3 ceramic using pressureless sintering with a reduced H2 atmosphere. A high transparency of 75.5% (theoretical transparency is 81.7%) at 1080 nm was achieved. Luminescence peaks were found at 1079 nm. EXAFS (extended X-ray absorption fine structure) results indicated greater disorder in the sample with the higher luminescence intensity. However, the detailed local disordered structure was unclear due to the practical limitations of that characterization technique.
The synthesis of 1 mol% Er: Lu2O3 hollow microspheres was performed via a urea-based homogeneous precipitation method by Jia [11]. Three distinct up-conversion luminescence emissions were identified under 980 nm excitation. They were 540 nm, 566 nm, and 660 nm, corresponding to the electronic transitions of 2H11/2     4I15/2, 4S3/2     4I15/2, and 4F9/2     4I15/2, respectively. However, the site specification and partial contributions to the spectra from cationic sites remained unknown. Information about site preference is important for the analysis of spectra from the bixbyite structure, which offers more than one substitutional cation site for the dopants. Merkle [12] investigated the spectra of 0.22 at% Er3+-doped Lu2O3; emissions at 1576 nm and 1601 nm were found to be promising for laser operation at cryogenic temperatures. Calculations of the cross-sections and radiative lifetime were performed based on two assumptions, which are difficult to verify experimentally. These were: (1) that Er3+ enters the 8b site without distortions; and (2) that it is random for Er3+ to enter 8b and 24d sites, and the concentration of 24d sites, which mainly contribute to the absorption spectra, is three-fourths of the total Er dopant concentration. So, information about the site occupations, local distortions, and respective spectra contributions is vital because they are the basis for evaluations of optical properties. The dependence of emission cross-section on the doping concentration was also observed in another work [13]. The sample with 11 at% Er doping concentration was found to possess the best laser performance. It was suggested that a homogeneous distribution of Er contributed to low quenching. Two mechanisms were proposed by Wang [14] for the up-conversion emissions (green band of 4S3/2     4I15/2 centers at 565 nm, red band of 4F9/2     4I15/2 centers at 660 nm) under 980 nm excitation. The main point of up-conversion is the excitation of electrons in excited levels to higher excited levels, either in a single Er3+ or between multiple Er3+. In the latter case, energy transfer from one excited Er3+ to another is required. However, the question remains as to which Er3+ pair is involved in the energy transfer, 8b or 24d sites. Further investigations [15] found that the two main up-conversion emissions responded differently to increasing Er3+ doping concentrations. The green emission reached its maximum intensity at 3 at% (increased doping from 2 at% to 10 at%). Further increases in doping concentration, beyond 3 at%, decreased its emission intensity. Cross-relaxation was proposed to be responsible for the observed quenching phenomenon. Kränkel [3] discussed three mid-infrared laser emissions in Er-doped Lu2O3. The emission at 3 μ m is related to two manifolds, 4I11/2 and 4I13/2. A relatively high doping concentration (>5%) was essential for laser gain from this electronic transition due to the influence of concentration on the lifetime of two terminal levels. It was mentioned that the disordered structure of the host would lead to changes in many aspects, such as absorption or emission line width, acceptance bandwidth, and even a decrease in thermal conductivity.
Co-doping with Li has been found to enhance the luminescence intensity of rare earth-doped oxides such as MgO. But the impact of co-doping Li in rare earth-doped bixbyite Lu2O3 has received much less attention so far. Modulation of up-conversion luminescence by Li in multi-doping Nd/Yb/Er: Lu2O3 nano-powders was investigated by Liu et al. [16]. The doped samples were prepared using a chemical co-precipitation method. The energy transfers between dopants were proposed to be influenced by Li co-doping via the Foxter–Dexter relation [17] (the energy transfer probability between adjacent atoms being inversely proportional to the sixth power of interatomic distance). The substitution of Li on the Lu site was proposed to be more favorable for the intensity increase than Li interstitials by reducing the symmetry of the local crystal field. Co-doping with Li was also found to impact the optical properties of rare earth-doped Lu2O3 by Li et al. [18]. Detailed descriptions of the defect structures are necessary to better understand the role played by the presence of Li.
Above all, the local defect structures of optically active rare earth ions are closely related to the optical properties of the doped systems. However, characterizing the defects at the atomistic scale experimentally presents significant challenges. Therefore, the main objective of this paper is to investigate the point defect structures and the configurations of defect pairs and defect complexes in rare earth-doped (Nd, Er) Lu2O3 with and without co-doped Li using atomistic simulations.

2. Simulation Methodology

The program employed for the current simulations was the General Utility Lattice Program (GULP 6.0), which is based on classical force field methods. The static lattice method [19] was the methodology applied and it is outlined below.
The interactions between atoms are modelled by a series of functions with unique parameters depending on the nature of interactions between the different atoms. For materials that are considered mainly ionic, the lattice energy calculations include both long-range electrostatic interactions (the Coulomb term) and short-range interactions, namely the London interaction, which is known as the Van Der Waals interaction (or dispersion interaction, for historical reasons), and a repulsion term considering the Pauli exclusion rules. The lattice energy expressed by the three interaction terms is:
E Lattice = φ Coulomb + φ London + φ Repulsion
The repulsion term in the Born–Mayer exponential form can be combined with the dispersive term, which is induced by interactions between the instantaneous dipole moments and their induced instantaneous dipole moments. The combined term is known as the Buckingham potential, which was used in the current calculations. The two-body Buckingham potential is expressed as:
φ ij Buckingham r ij = Aexp r ij ρ C 6 r ij 6
A, ρ , and C6 are the Buckingham parameters. It is computationally impractical to include interactions between all atoms in the solid as the number of atoms are on a scale of 1023 and Buckingham interations rapidly diminish with interatomic distance. Distance cut-offs were introduced to maintain a balance between efficiency and accuracy. Typically, a value of 5–10 Å is used. A cut-off distance of 10 Å and 12 Å were set for cation-oxygen and oxygen-oxygen potentials, respectively, in current simulations. The Ewald summation method was employed in the calculations of electrostatic interactions due to the slow convergence of 1/r summations. The arrangement of ions in the system was then determined by energy minimization processes with respect to all relevant structure factors, unit cell parameters, and atomic coordinates [19].
Atoms are treated as point ions, with the shell model proposed by Dick and Overhauser [13] being used to account for the polarizability of the ion. The shell model mimics the polarizability by defining an atom or ion as an entity consisting of a shell and a core that interact through the spring constant K cs . If the charges distributed on the shell of the ion are Q s , the polarizability of the free ion can be expressed as:
α = Q s 2 K cs
The sum of the core and shell charges is equal to the formal oxidation state of the ion. If all charges are on the core, the atom is considered to be unpolarizable. In the current simulations, the cationic cores possessed formal charges, and the charge distributions of O2− were 0.869 e and −2.869 e on the core and shell, respectively. The K cs of O2− was 74.9.
The Mott–Littleton (ML) method, or the so-called two-region strategy, was used for calculations of defect formation energies. Atoms in the regions with different distances to the assigned defect center are treated differently. The ML method divides the whole system studied into two regions. Atoms in Region 1, closest to the defect, are treated explictly atomistically. Region 2 is divided into two. Region 2b is treated as a continuous dielectric medium. The energy was evaluated using classical continuum theory, and atomic displacements were determined by bulk polarization. Region 2a is an interface region that is treated both atomistically and as a dielectric continuum to provide consistency between Region 1 and Region 2b [19,20,21]. The radii of Region 1 and Region 2a were set to be 20 Å and 35 Å.

3. Results and Discussion

3.1. Intrinsic Point Defects in Pristine Lutetium Oxide

Pristine crystalline Lu2O3 under ambient condition possesses a stable C-type rare earth crystal structure with space group Ia 3 ¯ (No. 206), isomorphous with Y2O3, the bixbyite structure. The bixbyite crystal structure offers two types of cationic sites with different site symmetries: C3i (Wyckoff position 8b) and C2 (Wyckoff position 24d) [22]. The bixbyite structure can be seen as a pseudo-fluorite structure with ordered arrays of oxygen vacancies [23]. From the viewpoint of crystallography, the sixfold coordination of both types of Lu sites can be derived by removing two oxygen atoms from the eightfold coordination of cations in the fluorite structure [24], that is, removing two oxygen atoms from eight Lu (8b) centered cubes. These two coordination environments are differently distorted instead of being identically perfect. It causes the whole structure to be relatively more complex compared with simple cubic structures such as fluorite or rock salt. This way of observing the bixbyite structure provides convenience for locating potential interstitial sites within Lu2O3.
The Buckingham potentials used in calculations are listed in Table 1 [25,26]. The structural parameters of the current simulations are listed in Table 2. After bulk optimization, the relaxed lattice parameter is 10.285 Å (0.08 Å or 0.8%; smaller than the experimental result [22]), and the lattice energy is −138.273 eV. The relaxed crystal structure of the (a) unit cell and (b) octahedral coordination of 8b and 24d sites are shown in Figure 1.
Intrinsic point defects in Lu2O3 include, using Kröger–Vink notation, vacancies V Lu / / / , V O and interstitials Lu i , O i / / . In order to calculate the defect energies of defect complexes involving interstitials systematically, the coordinates of the available interstitial sites in the Lu2O3 unit cell are required. But these coordinates were not explicitly listed in the literature. So, a re-evaluation of potentially available interstitial sites was conducted based on the experimentally obtained crystal structure of Lu2O3 [22]. The crystal structure of Lu2O3 was observed in the following two parts from which potential interstitial sites were identified. On the one hand, as shown in Figure 2a and as mentioned above, each Lu 8b site can be seen as being at the body center of a distorted cube, and the distribution of six coordinated oxygen atoms can be seen as removing two oxygen atoms from the eight body corners. All 48 oxygen atoms in the Lu2O3 unit cell can be seen to be in the eight distorted cubes. So, in total, there are 16 oxygen vacancies, i.e., potential interstitial sites, viewed from these Lu 8b-centered cubes. On the other hand, there is a fluorite type, a face-centered cube at the central part of the Lu2O3 unit cell, with eight Lu 8b sites at body corners and six Lu 24d sites at face centers. According to this center structure, another 13 potential interstitial points can be identified: they are 12 middle sites between adjacent Lu 8b sites and 1 at the body center, as shown in Figure 2b. These interstitial sites belong to the 16c, 24d, and 8a Wyckoff positions, respectively [27]. The fractional coordinates of the identified interstitial sites in the Lu2O3 unit cell are listed in Appendix A. An orthogonal view of those identified interstitial sites is shown in Figure 2c.
Results of the calculated defect formation energies of the four intrinsic point defects in pristine Lu2O3 are listed in Table 3.
Using the lowest formation energy of each point defect, the Schottky and Frenkel defect energies (per defect) were obtained:
Schottky :   nil 2 V Lu / / / + 3 V O + Lu 2 O 3 Δ H = 4.910   eV Frenkel   Lu :   nil V Lu / / / + Lu i Δ H = 7.840   eV Frenkel   O :   nil V O + O i / / Δ H = 3.544   eV
Comparing the Schottky and Frenkel defect energies, the O Frenkel is the most favorable, consisting of one oxygen vacancy point defect and one oxygen interstitial point defect. This indicates that in undoped Lu2O3, intrinsically, there is a greater probability of finding O vacancies compensated by O interstitials compared with a Lu vacancy compensated by a Lu interstitial. At this point in the analysis, the point defects are isolated.
When point defects approach close to each other, they may associate with each other. The effective charges of defects may promote clustering, forming defect pairs or defect complexes. The association energy in the current context is defined as: E association = E complex E isolated . So, the higher the association, the lower (more negative) the association energy, the lower the defect formation energy of the complex and the more favorable it is. Although Schottky disorder is less favorable than O Frenkel disorder, the association energies of the complexes were calculated systematically. Potential defect complexes include: ( V Lu / / / + V O ) pair , ( 2 V Lu / / / + V O ) complex , ( 2 V O + V Lu / / / ) complex , ( 3 V O + V Lu / / / ) complex , ( 2 V O + 2 V Lu / / / ) complex , and ( 2 V Lu / / / + 3 V O ) complex . All five defect complexes were considered, with 300 potential configurations being examined. The lowest association energies of each defect complex are shown in Table 4.
With the number of point defects increasing, a lower association energy is obtained. These results reveal a preference for clustering of the Schottky defects. The relaxed local structure of Lu2O3 that contains one ( 2 V Lu / / / + 3 V O ) complex is shown in Figure 3. Due to the two oxygen vacancies of the cluster present, the seven neighboring Lu ions’ coordination numbers were reduced by one. As a result, the coordination polyhedron changes from an octahedron to a distorted cuboid. The mean Lu-O distances of those cuboids are: 2.145 Å, 2.139 Å, 2.132 Å, 2.135 Å, 2.144 Å, 2.147 Å, and 2.146 Å. Ample void space emerges due to the clustering of these point defects, i.e., the vacancies. The dimensions of this void space concern the distance between these seven neighboring Lu ions. The minimum and maximum distances are 3.709 Å and 7.587 Å. This void space may be important for accommodating dopants of considerable size. If Schottky defect complexes are of high concentration, they will exert vital influences on physical properties, such as densities, and mechanical properties such as hardness and strength. It is noteworthy for studies related to dopants with a considerable size, as it is undoubtedly available space for interstitial point defects or even clusters.

3.2. Defect Structures of Rare Earth-Doped Lutetium Oxide

Neodymium (Nd) and erbium (Er) are the rare earth elements under current investigation. A primary problem of studying the doping mechanisms is to determine the preferential form of existence and preferential occupation site of the dopant in the host. As mentioned, there are plenty of potential interstitial sites in Lu2O3. Trivalent Nd and Er are isovalent with the Lu in the host, which means there is zero effective charge upon the substitution for Lu, and they possess similar ionic sizes. It is reasonable to assume a certain ease for them to be doped into the Lu2O3 host. Three sites were tested for interstitial Nd and Er: (0.625, 0.125, 0.375), (0.25, 0.50, 0.75), and (0.50, 0.50, 0.50). The lattice energies of Nd2O3, Er2O3, and Li2O are −129.01 eV, −134.97 eV, and −29.67 eV after bulk optimizations when using the crystal structures from experimental results [28,29,30,31,32,33].
Table 5 lists the calculated defect formation energies of Nd i , Nd Lu × , Er i , and Er Lu × . It first shows that 16c is the preferable site for interstitial Nd and Er. Furthermore, both Nd and Er prefer 8b as the substitution site. This agrees with other simulation results, which reveal a relationship between preferential sites and ionic size [34]. However, the differences between the defect formation energies of substitutional point defects concerning the two Wyckoff positions are not significantly large. They are Δ E 1 = 0.108   eV in Nd: Lu2O3 and Δ E 2 = 0.041   eV in Er: Lu2O3. Assuming that the energy differences are independent of temperature and constant entropy, statistically, the ratio of site occupation or concentration of 8b and 24d sites can be expressed by the Boltzmann distribution, where C is concentration: [35]
C 8 b C 24 d + C 8 b = exp Δ E kT exp Δ E kT + 3
As a function of absolute temperature, from (5), the concentration ratio between the 8b and 24d sites can be straightforwardly derived as plotted in Figure 4. The temperature range shown in the plot is from 200 to 2150 K, that is, from room temperature to up to 1800 °C, a high temperature that might be reached during fabrication processes (e.g., sintering) of bulk polycrystalline ceramic materials.
Regarding the bixbyite structure, the quantity ratio between 8b and 24d sites in Lu2O3 is approximately 0.33. However, will dopants occupy those sites strictly according to the quantity ratio of sites available? Our results show that the answer is negative. As shown, neither of the two dopants achieves the maximum ratio of 0.33. Within the temperature range considered, C8b/C24d reaches its maximum of 0.18 in Nd: Lu2O3 and 0.26 in Er: Lu2O3 at around 1800 °C. If the average sintering temperature is 1500 °C, then in Nd: Lu2O3, among all substituted sites, around 14% of them are 8b sites and 86% are 24d sites. In Er: Lu2O3, the concentration percentages are around 20% and 80% for 8b and 24d sites, respectively. Due to the smaller energy difference, at a specific temperature that provides sufficient energy for doping, more 8b sites are expected to be substituted in Er: Lu2O3 than Nd: Lu2O3, starting from an identical total doping quantity. From the viewpoint of energy, at 1500 °C, to reach a ratio of 0.33, Δ E is expected to be as small as 0.016 eV. Apparently, for Nd and Er, Δ E 1 and Δ E 2 are not small enough.
Putting the calculated point defect formation energies from Table 3 and Table 5 into quasi-chemical reactions (in the Appendix A), the ones with the lowest enthalpy (per rare earth dopant) of Nd and Er dopants in substitutional and interstitial sites are:
Nd 2 O 3 Lu 2 O 3 2 Nd Lu × + Lu 2 O 3 Δ H = 0.161   eV
Nd 2 O 3 Lu 2 O 3 2 Nd i + 3 O i / / Δ H = 13.887   eV
Er 2 O 3 Lu 2 O 3 2 Er Lu × + Lu 2 O 3 Δ H = 0.179   eV
Er 2 O 3 Lu 2 O 3 2 Er i + 3 O i / / Δ H = 14.209   eV
The negative enthalpy in (6) indicates that the reaction is exothermal. Compared with the interstitial site, substitution is (much) more favorable. If interstitials with positive effective charges appear, O i / / was found to be more favorable than V Lu / / / as the charge compensator.
To prove the preference for substitution from another aspect, the defect formation energies of Nd i + V Lu / / / pair and Er i + V Lu / / / pair were calculated. V Lu / / / was placed at (0.25, 0.25, 0.25), the 8b site, and all identified interstitial sites were tested for Nd and Er interstitials.
Figure 5 shows the results of the defect energies of defect pairs (a rare earth interstitial and a Lu vacancy) versus relaxed distance between the two point defects. At 0 Å, the interstitial rare earth dopant fills the Lu vacancy, becoming a substitutional defect. From 0 Å to around 4.5 Å, no data points are shown in the graph because all interstitial dopants initially placed in this range were relaxed to Lu sites. Nd interstitials were found to be directly relaxed to the (0.25, 0.25, 0.25) Lu site. For the Er interstitial, the surrounding Lu can be relaxed to the initial Lu vacancy. Then, the dopant interstitial fills the newly generated Lu vacancy, forming Er Lu × and leaving no vacancy at last. An example of this process is shown in Figure 6.
Initially, the Lu vacancy was placed at the 8b site (0.25, 0.25, 0.25) and the Er interstitial was placed at (0.75, 0.5, 0.25). After the energy minimization, the (0.25, 0.25, 0.25) position was filled by its neighboring Lu atom (Lu* as indicated), which was at a 24d site (0.53, 0.50, 0.25) initially. Alternatively, one could consider that the Lu vacancy moved to a neighboring 24d site. Then, the Er interstitial was found to occupy the initial position of Lu*, finally forming an Er Lu × point defect at the 24d site (0.53, 0.50, 0.25).
Another aspect worth investigating is the favorability of the configuration of dopant pairs. Experimental studies have shown that energy transfer between Eu dopants plays a significant role in the optical properties of Eu-doped bixbyite oxides [36]. For Nd-doped YAG, luminescence quenching phenomena were closely related to energy transfer and up/down conversions among adjacent Nd dopants. The mechanisms were discussed by Danielmeyer [8]. Calculations of the defect formation energies of dopant pairs will reveal which pairing configuration is more favorable for energy transfer, considering that the energy transfer probability is a function of interatomic distance [17]. A shorter distance indicates a greater probability of energy transferring between dopants. Three types of pairs were considered: (8b + 8b), (24d + 24d), and (8b + 24d). A total of 54 configurations were examined for each dopant.
As shown in Figure 7, the defect formation energy depends on the sites (8b and 24d) occupied by the pair and the distance between two substitutional point defects. The dashed square highlights the lowest energy in each type of pair. The (8b + 8b) pair has the lowest defect formation energy, and the order of energetic favorability is (8b + 8b) > (8b + 24d) > (24d + 24d). This order is the same for Nd- and Er-doped systems. In an (8b + 8b) pair, the most favorable interatomic distance is 8.907 Å, which is the third nearest neighbor (3rd NN) separation. Similarly, the third nearest neighbor configuration is also the most favored by the (8b + 24d) pair, with interatomic distances of 6.191 Å and 6.186 Å for Nd and Er dopants. Only (24d + 24d) most favors the nearest neighbor (NN) configuration, with interatomic distances of 3.524 Å and 3.478 Å for Nd and Er, respectively. According to the Foxter–Dexter relation [17], the energy transfer probability between adjacent atoms is inversely proportional to the sixth power of interatomic distance. The order of energy-transferring probability between two rare earth dopants among the three dopant pairs is then (24d + 24d) > (8b + 24d) > (8b + 8b). Table 6 lists the defect energies and inter-dopant distances of the NN and third NN configurations for the three types of pairs for both Nd and Er. When positioning two dopants in close proximity, such as in the nearest neighbor (NN) configuration of (8b + 24d) and (24d + 24d), Nd: Lu2O3 exhibits a higher degree of structural relaxation compared with Er: Lu2O3. The relaxed inter-dopant distances for the (8b + 24d) and (24d + 24d) NN configurations in Nd: Lu2O3 are 3.513 Å and 3.524 Å, respectively, which are greater than those in Er: Lu2O3, 3.466 Å and 3.478 Å. This discrepancy can be attributed to the larger ionic radius of Nd relative to Er. Note that the differences in the defect energies between the NN and the third NN configurations of (8b + 8b) and (8b + 24d) are on a scale of 0.001 eV in both Nd and Er doping systems, indicating that both configurations will appear in experimental practice. As for (24d + 24d), the configurations that are energetically close to the NN are the sixth or seventh NN. The favorability of NN configuration is relatively more significant in Nd- than Er-doped Lu2O3. Considering that there are twice as many Lu 24d sites than Lu 8b sites in the structure and considering the relative favorability of NN configuration, the energy transfer probability between two rare earth dopants in the (24d + 24d) pair may be the highest.

3.3. Defect Complexes Induced by Li Co-Doping

As noted in the Introduction, co-doping with Li has been experimentally shown to have an impact on optical properties. From the last section, it was shown that doping of Nd and Er into Lu2O3 requires relatively little energy. The substitution on the Lu site maintains the electroneutrality. No extra charge compensator is required. As for the co-doping of Li, the topics of interest are: (1) Does Li make it further easier for the doping of Nd and Er into the Lu2O3 host? (2) How will the inter-dopants distances change by co-doping Li? For the first question, comparing the enthalpy of the quasi-chemical reactions with and without Li will provide an answer. The second question requires calculations of the defect formation energies of the defect complexes containing the rare earth dopants and Li to determine the most favorable structures of the various defect complexes. With the incorporation of Li, two additional charged point defects Li i and Li Lu / / may be introduced. Li favors the 8b and 24d sites as the substitutional and interstitial point defects; the defect formation energies of Li i (8a) and Li Lu / / (24d) are −4.374 eV and 40.161 eV, respectively, as listed in Table 7.
Possible quasi-chemical reactions in Li-doped Lu2O3 (per Li dopant) are shown in (7).
3 2 Li 2 O Lu 2 O 3 2 Li i + Li Lu / / + 1 2 Lu 2 O 3 Δ H = 1.868   eV
Li 2 O Lu 2 O 3 2 Li i + O i / / Δ H = 2.997   eV
3 2 Li 2 O Lu 2 O 3 3 Li i + V Lu / / / + 1 2 Lu 2 O 3 Δ H = 3.545   eV
1 2 Li 2 O Lu 2 O 3 Li Lu / / + V O + 1 2 Lu 2 O 3 Δ H = 6.699   eV
3 2 Li 2 O Lu 2 O 3 3 Li Lu / / + 2 Lu i + 1 2 Lu 2 O 3 Δ H = 8.968   eV
The most favorable way to accommodate Li in Lu2O3 is Li i and Li Lu / / coexisting, as shown in reaction (7a). The two types of point defect of opposite effective charge can compensate each other. Consider Li i and Li Lu / / separately; it can be found that Li i is the more favorable point defect by comparing reactions (7b) and (7d), which are the reactions with the lowest enthalpies for Li i and Li Lu / / , respectively.
After introducing Li into Nd- and Er-doped Lu2O3, the lowest enthalpy reactions are:
Nd 2 O 3 + 3 2 Li 2 O Lu 2 O 3 2 Nd Lu × + 2 Li i + Li Lu / / + 3 2 Lu 2 O 3 Δ H = 2.641   eV  
Er 2 O 3 + 3 2 Li 2 O Lu 2 O 3 2 Er Lu × + 2 Li i + Li Lu / / + 3 2 Lu 2 O 3 Δ H = 2.983 eV
By comparing (8a) and (8b) with (6a) and (6c), substitutional rare earth dopants do not benefit energetically from co-doping with Li, which only brings extra charged point defects that need to be compensated. Remember in reactions (6b) and (6d) that the favorable charge compensator for Nd i and Er i point defects is O i / / . After introducing Li, Li Lu / / replaces O i / / as their favorable charge compensator:
Nd 2 O 3 + 3 2 Li 2 O Lu 2 O 3 2 Nd i + 3 Li Lu / / + 3 2 Lu 2 O 3 Δ H = 13.303   eV .
Er 2 O 3 + 3 2 Li 2 O Lu 2 O 3 2 Er i + 3 Li Lu / / + 3 2 Lu 2 O 3 Δ H = 13.629   eV .
By comparing reactions (9a) and (9b) with reactions (8a) and (8b), the advantage of substitutional sites to interstitial sites of rare earth dopants are not challenged by Li co-doping, as the energy required for interstitial point defects is still about four times higher than for substitutional point defects.
Considering the reactions (8), to determine which has the lowest enthalpy in Li-co-doped RE: Lu2O3 (RE = Nd, Er), it is necessary to consider the defect complex 2 RE Lu × + 2 Li i + Li Lu / / complex . For 2 RE Lu × , the nearest neighbor (NN) configurations were used. All three types of pairs were considered, (8b + 8b), (8b + 24d), and (24d + 24d). In searching for the most favorable structure of the defect complex, the point defects were placed and tested in the following way: The center of the dopant pair was fixed by placing two dopants first. Then, Li i , Li Lu / / were tested in a range from dmin to dmax. dmax is the distance between the pair center and the second nearest interstitial sites for Li i and the second nearest Lu sites for Li Lu / / . For each type of pair, the Wyckoff positions of the nearby sites tested for Li i and Li Lu / / point defects are listed in Table 8. A total of 67 configurations of 2 RE Lu × + 2 Li i + Li Lu / / complex were tested for each doping system.
Table 9 lists the inter-dopants (the rare earth dopants) distances of the most favorable configurations of 2 Nd Lu × + 2 Li i + Li Lu / / complex and 2 Er Lu × + 2 Li i + Li Lu / / complex , as well as the Wyckoff positions of the Li i and Li Lu / / point defects in the defect complexes after relaxation.
Investigating the effects of Li on the distances between the nearest neighbor dopants is important, because for each of the three types of pairs—(8b + 8b), (8b + 24d), and (24d + 24d)—the energy transfer probability is the highest in the nearest neighbor configurations. Comparing the inter-dopant distances of 2 RE Lu × from Table 6 and the 2 RE Lu × + 2 Li i + Li Lu / / complex from Table 9, one finds that the inter-dopant distances change on a scale of 0.01–0.1 Å due to the incorporation of Li point defects near the rare earth dopant pairs. For 2 Nd Lu × , nearby Li point defects cause the inter-dopant distance of NN (8b + 8b) to decrease. This effect is the same for the NN (8b + 8b) of 2 Er Lu × . However, for NN (8b + 24d) and NN (24d + 24d) pairs, the effects of Li are opposite in 2 Nd Lu × and 2 Er Lu × . The inter-dopant distances of NN (8b + 24d) and NN (24d + 24d) are increased in 2 Nd Lu × but decreased in 2 Er Lu × when co-doped with Li. From above, in Nd: Lu2O3, the inter-dopant energy transfer is encouraged only in (8b + 8b) among the three types of NN configurations by Li co-doping. In Er: Lu2O3, inter-dopant energy transfer in all three types of NN configurations is encouraged by Li co-doping. So, the impact of Li on inter-dopant distance depends on the type of dopant and dopant pairs. But in both Nd- and Er-doped systems, Li has the least impact on the inter-dopant distance of the (8b + 8b) pair (a decrease of 1.0% and 0.6% in Nd- and Er-doped systems, respectively) and the most significant impact on the (24d + 24d) pair (an increase of 3.0% in the Nd-doped system; a decrease of 2.5% in the Er-doped system). These results are meaningful for the spectroscopic analysis of rare earth elements in Lu2O3 and provide a theoretical basis for co-doping with Li to control the concentration quenching of the rare earth elements in Lu2O3.

4. Conclusions

This paper investigated pristine and rare earth (Nd, Er)-doped Lu2O3 with respect to point defects and structures of defect complexes. The influence of Li as a co-dopant was considered. The defect formation energies were calculated, and the most energetically favorable defect structures were determined. In pristine Lu2O3, associations between the five individual point defects in the Schottky defect result in clustering, forming a local concentration of vacancies. According to the calculation results, the rare earth dopants, Nd and Er, energetically prefer the substitutional Lu sites compared with interstitial sites in the Lu2O3 host. This result supports the experimental characterization results, where the X-ray diffraction patterns of Nd: Lu2O3 [9] and Er: Lu2O3 [11,15] ceramic samples showed no impurity phase, indicating substitutions of Lu by Nd and Er dopants. The 8b site is more favorable than the 24d site for the substitutional Nd and Er point defects. Nevertheless, the energy difference is minor, on a scale of 0.01 eV–0.1 eV. The difference is smaller in Er: Lu2O3 than in Nd: Lu2O3. Attention was drawn to dopant pairs, considering the importance of the interatomic distance between dopants for energy transfer. Among the three types of dopant pairs, (8b + 8b), (8b + 24d), and (24d + 24d), the (8b + 8b) pair in the third nearest neighbor configuration possesses the lowest defect formation energy. Only the (24d + 24d) pair was found to most favor the nearest neighbor configuration. Therefore, energy transfer is predicted to occur with more probability in the (24d + 24d) configuration. Li Lu / / and Li i are introduced by co-doping Li, and the energy required for the formation of charge compensators for these two charged point defects increases the enthalpy of doping Nd and Er into the Lu2O3 host. So, it is not energetically favorable to co-dope Li. Structurally, the effects of co-doping Li on the inter-dopant distances in the nearest neighbor configurations were found to be different in Nd and Er doping systems. Li has the least impact on the inter-dopant distances in the (8b + 8b) pair and a more significant impact on the (24d + 24d) pair.
The results can assist in the experimental analysis of the optical properties of rare earth-doped Lu2O3 and provide a basis to further investigate the concentration-induced luminescence quenching phenomena.

Author Contributions

Conceptualization, A.N.C. and Y.W.; investigation, Y.Z., A.N.C. and Y.W.; methodology, A.N.C.; supervision, A.N.C. and Y.W.; writing—original draft, Y.Z.; writing—review and editing, A.N.C. and Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding authors.

Acknowledgments

The formulation of the research topic in this work was crafted by Yiquan Wu and Alastair N. Cormack. We thank the Inamori School of Engineering at NYS College of Ceramics, Alfred University, for providing the computational resources for this work. Y.Z. would like to thank Alfred University for studentships.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Table A1. Fractional coordinates of interstitial sites in Lu2O3 unit cell.
Table A1. Fractional coordinates of interstitial sites in Lu2O3 unit cell.
From 8 cation centered cubes
Wyckoff position: 16c
1. (0.625, 0.125, 0.375)2. (0.875, 0.375, 0.125)
3. (0.125, 0.125, 0.125)4. (0.375, 0.375, 0.375)
5. (0.875, 0.625, 0.375)6. (0.625, 0.875, 0.125)
7. (0.375, 0.625, 0.125)8. (0.125, 0.875, 0.375)
9. (0.375, 0.875, 0.625)10. (0.125, 0.625, 0.875)
11. (0.625, 0.625, 0.625) 12. (0.875, 0.875, 0.875)
13. (0.375, 0.125, 0.875)14. (0.125, 0.375, 0.625)
15. (0.625, 0.375, 0.875)16. (0.875, 0.125, 0.625)
From central fluorite-type cube
Wyckoff position: 24d
17. (0.75, 0.5, 0.25)18. (0.5, 0.75, 0.25)
19. (0.25, 0.5, 0.25)20. (0.5, 0.25, 0.25)
21. (0.75, 0.25, 0.5)22. (0.25, 0.25, 0.5)
23. (0.75, 0.75, 0.5)24. (0.25, 0.75, 0.5)
25. (0.5, 0.75, 0.75)26. (0.75, 0.5, 0.75)
27. (0.5, 0.25, 0.75)28. (0.25, 0.5, 0.75)
Wyckoff position: 8a
29. (0.5, 0.5, 0.5)
Table A2. Quasi-chemical reactions of doping Nd into Lu2O3.
Table A2. Quasi-chemical reactions of doping Nd into Lu2O3.
Reaction Δ H Per Nd Dopant (eV)
1 2 Nd 2 O 3 Lu 2 O 3 Nd i + 3 V O + 3 V Lu / / / + 3 2 Lu 2 O 3 40.083
1 2 Nd 2 O 3 Lu 2 O 3 Nd i + V O + O i / / + V Lu / / / + 1 2 Lu 2 O 3 22.619
Nd 2 O 3 Lu 2 O 3 Nd Lu × + Nd i + 3 V O + 3 V Lu / / / + 2 Lu 2 O 3 19.961
Nd 2 O 3 Lu 2 O 3 2 Nd i + V O + 4 O i / / 17.430
Nd 2 O 3 Lu 2 O 3 2 Nd Lu × + 4 V O + O i / / + 2 V Lu / / / + 2 Lu 2 O 3 15.659
Nd 2 O 3 Lu 2 O 3 2 Nd i + 2 V Lu / / / + Lu 2 O 3 15.531
3 2 Nd 2 O 3 Lu 2 O 3 3 Nd i + 3 O i / / + V Lu / / / + 1 2 Lu 2 O 3 14.435
Nd 2 O 3 Lu 2 O 3 2 Nd i + 3 O i / / 13.887
Nd 2 O 3 Lu 2 O 3 2 Nd Lu × + 3 V O + 2 V Lu / / / + 2 Lu 2 O 3 12.116
3 2 Nd 2 O 3 Lu 2 O 3 Nd Lu × + 2 Nd i + V O + 4 O i / / + 1 2 Lu 2 O 3 11.567
Nd 2 O 3 Lu 2 O 3 Nd Lu × + Nd i + V O + O i / / + V Lu / / / + Lu 2 O 3 11.229
2 Nd 2 O 3 Lu 2 O 3 Nd Lu × + 3 Nd i + 3 O i / / + V Lu / / / + Lu 2 O 3 10.786
3 2 Nd 2 O 3 Lu 2 O 3 Nd Lu × + 2 Nd i + 3 O i / / + 1 2 Lu 2 O 3 9.204
Nd 2 O 3 Lu 2 O 3 Nd Lu × + Nd i + V Lu / / / + Lu 2 O 3 7.685
Nd 2 O 3 Lu 2 O 3 2 Nd Lu × + V O + O i / / + Lu 2 O 3 3.380
Nd 2 O 3 Lu 2 O 3 2 Nd Lu × + Lu 2 O 3 −0.161
Table A3. Quasi-chemical reactions of doping Er into Lu2O3.
Table A3. Quasi-chemical reactions of doping Er into Lu2O3.
Reaction Δ H Per Er Dopant (eV)
1 2 Er 2 O 3 Lu 2 O 3 Er i + 3 V O + 3 V Lu / / / + 3 2 Lu 2 O 3 40.405
1 2 Er 2 O 3 Lu 2 O 3 Er i + V O + O i / / + V Lu / / / + 1 2 Lu 2 O 3 22.941
Er 2 O 3 Lu 2 O 3 Er Lu × + Er i + 3 V O + 3 V Lu / / / + 2 Lu 2 O 3 20.292
Er 2 O 3 Lu 2 O 3 2 Er i + V O + 4 O i / / 17.753
Er 2 O 3 Lu 2 O 3 2 Er Lu × + 4 V O + O i / / + 2 V Lu / / / + 2 Lu 2 O 3 15.998
Er 2 O 3 Lu 2 O 3 2 Er i + 2 V Lu / / / + Lu 2 O 3 15.854
3 2 Er 2 O 3 Lu 2 O 3 3 Er i + 3 O i / / + V Lu / / / + 1 2 Lu 2 O 3 14.757
Er 2 O 3 Lu 2 O 3 2 Er i + 3 O i / / 14.209
Er 2 O 3 Lu 2 O 3 2 Er Lu × + 3 V O + 2 V Lu / / / + 2 Lu 2 O 3 12.454
3 2 Er 2 O 3 Lu 2 O 3 Er Lu × + 2 Er i + V O + 4 O i / / + 1 2 Lu 2 O 3 11.895
Er 2 O 3 Lu 2 O 3 Er Lu × + Er i + V O + O i / / + V Lu / / / + Lu 2 O 3 11.560
2 Er 2 O 3 Lu 2 O 3 Er Lu × + 3 Er i + 3 O i / / + V Lu / / / + Lu 2 O 3 11.113
3 2 Er 2 O 3 Lu 2 O 3 Er Lu × + 2 Er i + 3 O i / / + 1 2 Lu 2 O 3 9.532
Er 2 O 3 Lu 2 O 3 Er Lu × + Er i + V Lu / / / + Lu 2 O 3 8.016
Er 2 O 3 Lu 2 O 3 2 Er Lu × + V O + O i / / + Lu 2 O 3 3.722
Er 2 O 3 Lu 2 O 3 2 Er Lu × + Lu 2 O 3 0.179

References

  1. Uehara, H.; Yasuhara, R.; Tokita, S.; Kawanaka, J.; Murakami, M.; Shimizu, S. Efficient continuous wave and quasi-continuous wave operation of a 28 μm Er: Lu2O3 ceramic laser. Opt. Express 2017, 25, 18677–18684. [Google Scholar] [CrossRef] [PubMed]
  2. Lu, J.; Takaichi, K.; Uematsu, T.; Shirakawa, A.; Musha, M.; Ueda, K.; Yagi, H.; Yanagitani, T.; Kaminskii, A.A. Promising ceramic laser material: Highly transparent Nd3+: Lu2O3 ceramic. Appl. Phys. Lett. 2002, 81, 4324–4326. [Google Scholar] [CrossRef]
  3. Krankel, C. Rare-earth-doped sesquioxides for diode-pumped high-power lasers in the 1-, 2-, and 3-μm spectral range. IEEE J. Sel. Top. Quantum Electron. 2015, 21, 250–262. [Google Scholar] [CrossRef]
  4. Liu, Z.; Ikesue, A.; Li, J. Research progress and prospects of rare-earth doped sesquioxide laser ceramics. J. Eur. Ceram. Soc. 2021, 41, 3895–3910. [Google Scholar] [CrossRef]
  5. Goget, G.A.; Guyot, Y.; Guzik, M.; Boulon, G.; Ito, A.; Goto, T.; Yoshikawa, A.; Kikuchi, M. Nd3+-doped Lu2O3 transparent sesquioxide ceramics elaborated by the Spark Plasma Sintering (SPS) method. Part 1: Structural, thermal conductivity and spectroscopic characterization. Opt. Mater. 2015, 41, 3–11. [Google Scholar] [CrossRef]
  6. Von Brunn, P.; Heuer, A.M.; Fornasiero, L.; Huber, G.; Kränkel, C. Efficient laser operation of Nd3+: Lu2O3 at various wavelengths between 917 nm and 1463 nm. Laser Phys. 2016, 26, 084003. [Google Scholar] [CrossRef]
  7. Ikesue, A.; Aung, Y.L.; Lupei, V. Ceramic Lasers; Cambridge University Press: New York, NY, USA, 2013. [Google Scholar]
  8. Danielmeyer, H.G.; Blätte, M.; Balmer, P. Fluorescence Quenching in Nd: YAG. Appl. Phys. 1973, 1, 269–274. [Google Scholar] [CrossRef]
  9. Zhou, D.; Shi, Y.; Xie, J.; Ren, Y.; Yun, P. Fabrication and luminescent properties of Nd3+-Doped Lu2O3 transparent ceramics by pressureless sintering. J. Am. Ceram. Soc. 2009, 92, 2182–2187. [Google Scholar] [CrossRef]
  10. Zhou, D.; Ren, Y.; Xu, J.; Shi, Y.; Jiang, G.; Zhao, Z. Fine grained Nd3+: Lu2O3 transparent ceramic with enhanced photoluminescence. J. Eur. Ceram. Soc. 2014, 34, 2035–2039. [Google Scholar] [CrossRef]
  11. Jia, G.; You, H.; Zheng, Y.; Liu, K.; Guo, N.; Zhang, H. Synthesis and characterization of highly uniform Lu2O3: Ln3+ (Ln = Eu, Er, Yb) luminescent hollow microspheres. CrystEngComm 2010, 12, 2943–2948. [Google Scholar] [CrossRef]
  12. Merkle, L.D.; Gabirielyan, N.T.; Kacik, N.J.; Sanamyan, T.; Zhang, H.; Yu, H.; Wang, J.; Dubinskii, M. Er: Lu2O3-Laser-related spectroscopy. Opt. Mater. Express 2013, 3, 1992. [Google Scholar] [CrossRef]
  13. Uehara, H.; Tokita, S.; Kawanaka, J.; Konishi, D.; Murakami, M.; Shimizu, S.; Yasuhara, R. Optimization of laser emission at 28 μm by Er: Lu2O3 ceramics. Opt. Express 2018, 26, 3497. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, N.L.; Zhang, X.Y.; Wang, P.H. Fabrication and spectroscopic characterization of Er3+: Lu2O3 transparent ceramics. Mater. Lett. 2013, 94, 5–7. [Google Scholar] [CrossRef]
  15. Wang, N.; Zhang, X.; Wang, P. Synthesis of Er3+: Lu2O3 nanopowders by carbonate co-precipitation process and fabrication of transparent ceramics. J. Alloys Compd. 2015, 652, 281–286. [Google Scholar] [CrossRef]
  16. Liu, X.; Li, T.; Hu, W.; Zhao, P. Simultaneous size manipulation and up-conversion luminescence modulation of Lu2O3: Nd3+/Yb3+/Er3+ nanospheres by Li+ ion doping. Mater. Res. Bull. 2019, 113, 161–168. [Google Scholar] [CrossRef]
  17. Dexter, D.L.; Schulman, J.H. Theory of concentration quenching in inorganic phosphors. J. Chem. Phys. 1954, 22, 1063–1070. [Google Scholar] [CrossRef]
  18. Li, L.; Wei, X.; Cao, X.; Deng, K.; Chen, Q.; Chen, Y.; Guo, C.; Yin, M. Upconversion luminescence enhancement in Yb3+/Tm3+-codoped Lu2O3 nanocrystals induced by doping of Li+ ions. J. Nanosci. Nanotechnol. 2011, 11, 9892–9898. [Google Scholar] [CrossRef]
  19. Catlow, C.R.A.; Bell, R.G.; Gale, J.D. Computer modelling as a technique in materials chemistry. J. Mater. Chem. 1994, 4, 781. [Google Scholar] [CrossRef]
  20. Du, J.; Cormack, A.N. (Eds.) Atomistic Simulations of Glasses; Wiley: Hoboken, NJ, USA, 2022. [Google Scholar]
  21. Tilley, R.J.D. Defects in Solids; John Wiley & Sons: Hoboken, NJ, USA, 2008. [Google Scholar]
  22. Guzik, M.; Pejchal, J.; Yoshikawa, A.; Ito, A.; Goto, T.; Siczek, M.; Lis, T.; Boulon, G. Structural investigations of Lu2O3 as single crystal and polycrystalline transparent ceramic. Cryst. Growth Des. 2014, 14, 3327–3334. [Google Scholar] [CrossRef]
  23. Lee, D.; Gao, X.; Sun, L.; Jee, Y.; Poplawsky, J.; Farmer, T.O.; Fan, L.; Guo, E.; Lu, Q.; Heller, W.; et al. Colossal oxygen vacancy formation at a fluorite-bixbyite interface. Nat. Commun. 2020, 11, 1371. [Google Scholar] [CrossRef]
  24. Hanic, F.; Hartmanova, M.; Aida, G.G.; Urusovskaya, A.A.; Bagdasarov, K.S. Real Structure of Undoped Y2O3 Single Crystals. Acta Cryst. B 1936, 40, 76–82. [Google Scholar] [CrossRef]
  25. Lewis, G.V.; Catlow, C.R.A. Potential models for ionic oxides. J. Phys. C Solid State Phys. 1985, 18, 1149. [Google Scholar] [CrossRef]
  26. Zhao, Y.; Cormack, A.N.; Wu, Y. Atomistic simulations of defect structures in rare earth doped magnesium oxide. Crystals 2024, 14, 384. [Google Scholar] [CrossRef]
  27. Aroyo, M.; Perez, J.M.; Capillas, C.; Kroumova, E.; Ivantchev, S.; Madariaga, G.; Kirov, A.; Wondratschek, H. Bilbao Crystallographic Server: I. Databases and crystallographic computing programs. Zeitschrift Kristallographie 2006, 221, 15–27. [Google Scholar] [CrossRef]
  28. Faucher, M.; Pannetier, J.; Charreire, Y.; Caro, P. X-ray structure analysis and molecular conformation of tert-butyloxycarbonyI-L-prolylproline (Boo-Pro-Pro): Errata. Refinement of the Nd2O3and Nd2O2S structures at 4 K. Acta Cryst. B 1982, 38, 344. [Google Scholar] [CrossRef]
  29. Islam, M.M.; Bredow, T.; Minot, C. Theoretical analysis of structural, energetic, electronic, and defect properties of Li2O. J. Phys. Chem. B 2006, 110, 9413–9420. [Google Scholar] [CrossRef] [PubMed]
  30. Goel, P.; Choudhury, N.; Chaplot, S.L. Superionic behavior of lithium oxide Li2O: A lattice dynamics and molecular dynamics study. Phys. Rev. B 2004, 70, 174307. [Google Scholar] [CrossRef]
  31. Ohno, H.; Konishi, S.; Noda, K.; Takeshita, H.; Yoshida, H.; Watanabe, H.; Matsuo, T. Conductivities of a sintered pellet and a single crystal of Li2O. J. Nucl. Mater. 1983, 118, 242–247. [Google Scholar] [CrossRef]
  32. Guo, Q.; Zhao, Y.; Jiang, C.; Mao, W.L.; Wang, Z.; Zhang, J.; Wang, Y. Pressure-induced cubic to monoclinic phase transformation in erbium sesquioxide Er2O3. Inorg. Chem. 2007, 46, 6164–6169. [Google Scholar] [CrossRef]
  33. Berard, M.F.; Wilder, D.R. Cation Self-Diffusion in Polycrystalline Y2O3 and Er2O3. J. Am. Ceram. Soc. 1969, 52, 85–88. [Google Scholar] [CrossRef]
  34. Levy, M.R.; Stanek, C.R.; Chroneos, A.; Grimes, R.W. Defect chemistry of doped bixbyite oxides. Solid State Sci. 2007, 9, 588–593. [Google Scholar] [CrossRef]
  35. Stanek, C.R.; McClellan, K.J.; Uberuaga, B.P.; Sickafus, K.E.; Levy, M.R.; Grimes, R.W. Determining the site preference of trivalent dopants in bixbyite sesquioxides by atomic-scale simulations. Phys. Rev. B Condens. Matter. Mater. Phys. 2007, 75, 134101. [Google Scholar] [CrossRef]
  36. Buijs, M.; Meyerink, A.; Blasse, G. Energy transfer between Eu3+ ions in a lattice with two different crystallographic sites: Y2O3: Eu3+, Gd2O3: Eu3+ and Eu2O3. J. Lumin. 1987, 37, 9–20. [Google Scholar] [CrossRef]
Figure 1. Relaxed crystal structure of Lu2O3. Dark grey: Lu 8b site. Light grey: Lu 24d site. White: O. (a) Lu2O3 unit cell and (b) coordination of Lu sites.
Figure 1. Relaxed crystal structure of Lu2O3. Dark grey: Lu 8b site. Light grey: Lu 24d site. White: O. (a) Lu2O3 unit cell and (b) coordination of Lu sites.
Crystals 14 00413 g001
Figure 2. Identified interstitial sites in Lu2O3: (a) 16 ordered oxygen vacancies in 8 distorted cubes centered around Lu 8b sites; (b) 13 interstitial sites in the central fluorite-type cube; (c) interstitial sites viewed from <100>. Dark grey: Lu 8b site. Light grey: Lu 24d site. White: O. Dashed: potential interstitial site.
Figure 2. Identified interstitial sites in Lu2O3: (a) 16 ordered oxygen vacancies in 8 distorted cubes centered around Lu 8b sites; (b) 13 interstitial sites in the central fluorite-type cube; (c) interstitial sites viewed from <100>. Dark grey: Lu 8b site. Light grey: Lu 24d site. White: O. Dashed: potential interstitial site.
Crystals 14 00413 g002aCrystals 14 00413 g002b
Figure 3. Configuration of ( 2 V Lu / / / + 3 V O ) complex . Dark grey and striped: Lu atom. White: O atom.
Figure 3. Configuration of ( 2 V Lu / / / + 3 V O ) complex . Dark grey and striped: Lu atom. White: O atom.
Crystals 14 00413 g003
Figure 4. Concentration ratio of 8b and 24d sites in Nd: Lu2O3 and Er: Lu2O3.
Figure 4. Concentration ratio of 8b and 24d sites in Nd: Lu2O3 and Er: Lu2O3.
Crystals 14 00413 g004
Figure 5. Defect energy between V Lu / / / and interstitial isovalent dopants.
Figure 5. Defect energy between V Lu / / / and interstitial isovalent dopants.
Crystals 14 00413 g005
Figure 6. Illustration of an initially distant Er interstitial occupying a Lu vacancy. Fuzzy white: O atom; fuzzy light grey: Lu atom.
Figure 6. Illustration of an initially distant Er interstitial occupying a Lu vacancy. Fuzzy white: O atom; fuzzy light grey: Lu atom.
Crystals 14 00413 g006
Figure 7. Variations of energy with distance between substitutional Nd or Er point defects: (a) Nd-doped; (b) Er-doped.
Figure 7. Variations of energy with distance between substitutional Nd or Er point defects: (a) Nd-doped; (b) Er-doped.
Crystals 14 00413 g007aCrystals 14 00413 g007b
Table 1. Buckingham potentials used in calculations.
Table 1. Buckingham potentials used in calculations.
Pairwise InteractionA (eV)ρ (Å)C6 (eV•Å6)
Lu-O1347.10.34300
O-O22,7640.149027.879
Nd-O1379.90.36010
Li-O235.10.35440
Er-O1381.50.34920
Table 2. Optimized atomic positions of bixbyite Lu2O3.
Table 2. Optimized atomic positions of bixbyite Lu2O3.
AtomFractional CoordinateWyckoff PositionCN
Lu1(0.250, 0.250, 0.250)8b6
Lu2(0.471, 0.000, 0.250)24d6
O(0.390, 0.152, 0.381)48e4
Table 3. Defect formation energies in pristine Lu2O3.
Table 3. Defect formation energies in pristine Lu2O3.
Intrinsic Point DefectDefect Formation Energy (eV)
V Lu / / / 8b50.014
24d50.709
V O 48e20.933
Lu i 8a−33.402
24d−32.669
16c−34.334
O i / / 8a−12.390
24d−11.016
16c−13.845
Table 4. Association energies of Schottky defect complexes in Lu2O3.
Table 4. Association energies of Schottky defect complexes in Lu2O3.
DefectEffective Charge (e)Association Energy (eV)
( V Lu / / / + V O ) pair −1−4.265
( 2 V Lu / / / + V O ) complex −4−5.402
( 2 V O + V Lu / / / ) complex +1−7.509
( 3 V O + V Lu / / / ) complex +3−8.638
( 2 V O + 2 V Lu / / / ) complex −2−10.835
( 2 V Lu / / / + 3 V O ) complex 0−14.917
Table 5. Point defects in Nd (Er): Lu2O3.
Table 5. Point defects in Nd (Er): Lu2O3.
Point DefectWyckoff PositionDefect Formation Energy (eV)
Nd Lu × 8b4.471
24d4.579
Er Lu × 8b1.832
24d1.873
Nd i 8a−27.073
24d−26.183
16c−29.851
Er i 8a−30.756
24d−29.947
16c−32.507
Table 6. Relaxed structures of 2 Nd Lu × .
Table 6. Relaxed structures of 2 Nd Lu × .
2 N d L u × ConfigurationInter-Dopant Distance (Å)Defect Energy (eV)
(8b + 8b)NN5.1508.948
3rd NN *8.9078.942
(8b + 24d)NN3.5139.035
3rd NN *6.1919.031
(24d + 24d)NN *3.5249.144
2 E r L u × ConfigurationInter-Dopant Distance (Å)Defect Energy (eV)
(8b + 8b)NN5.1453.665
3rd NN *8.9073.664
(8b + 24d)NN3.4663.703
3rd NN *6.1863.702
(24d + 24d)NN *3.4783.744
* denotes the configuration of the lowest defect formation energy in each type of pair.
Table 7. Defect formation energies of point defects of Li in Lu2O3.
Table 7. Defect formation energies of point defects of Li in Lu2O3.
Point DefectWyckoff PositionDefect Formation Energy (eV)
Li i 8a−4.374
24d−4.916
16c−4.375
Li Lu / / 8b40.068
24d40.161
Table 8. Wyckoff positions of the sites tested for Li i and Li Lu / / .
Table 8. Wyckoff positions of the sites tested for Li i and Li Lu / / .
2 R E L u ×
in the Complex
Nearby   Interstitial   Sites   Tested   for   L i i
Nearby   Substitutional   Sites   for   L i L u / /
NN (8b + 8b)8a, 16c, 24d24d
NN (8b + 24d)16c, 24d8b, 24d
NN (24d + 24d)8a, 24d24d
Table 9. Relaxed structures of the 2 RE Lu × + 2 Li i + Li Lu / / complex .
Table 9. Relaxed structures of the 2 RE Lu × + 2 Li i + Li Lu / / complex .
2 N d L u ×  in the ComplexRelaxed Wyckoff Positions of  L i i Relaxed Wyckoff Positions of  L i L u / / Relaxed Inter-Dopant Distance (Å)
NN (8b + 8b)48e48e5.101
NN (8b + 24d)48e24d3.560
NN (24d + 24d)8a, 48e48e3.629
2 E r L u ×  in the ComplexRelaxed Wyckoff Positions of  L i i Relaxed Wyckoff Positions of  L i L u / / Relaxed Inter-Dopant Distance (Å)
NN (8b + 8b)48e48e5.112
NN (8b + 24d)48e48e3.411
NN (24d + 24d)16c, 48e16c3.390
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, Y.; Cormack, A.N.; Wu, Y. Defect Structures of Rare Earth-Doped Lutetium Oxide and Impacts of Li Co-Dopant. Crystals 2024, 14, 413. https://doi.org/10.3390/cryst14050413

AMA Style

Zhao Y, Cormack AN, Wu Y. Defect Structures of Rare Earth-Doped Lutetium Oxide and Impacts of Li Co-Dopant. Crystals. 2024; 14(5):413. https://doi.org/10.3390/cryst14050413

Chicago/Turabian Style

Zhao, Yanfeng, Alastair N. Cormack, and Yiquan Wu. 2024. "Defect Structures of Rare Earth-Doped Lutetium Oxide and Impacts of Li Co-Dopant" Crystals 14, no. 5: 413. https://doi.org/10.3390/cryst14050413

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

Article Metrics

Back to TopTop