Next Article in Journal
Innovative Powder Pre-Treatment Strategies for Enhancing Maraging Steel Performance
Previous Article in Journal
Influence of Process Parameter and Build Rate Variations on Defect Formation in Laser Powder Bed Fusion SS316L
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Features of the Defect Structure of LiNbO3:Mg:B Crystals of Different Composition and Genesis

by
Roman A. Titov
1,
Alexandra V. Kadetova
1,2,
Diana V. Manukovskaya
1,*,
Maxim V. Smirnov
1,
Olga V. Tokko
2,
Nikolay V. Sidorov
1,
Irina V. Biryukova
1,
Sofja M. Masloboeva
1 and
Mikhail N. Palatnikov
1
1
Tananaev Institute of Chemistry–Subdivision of the Federal Research Centre “Kola Science Centre of the Russian Academy of Sciences” (ICT KSC RAS), Apatity 184209, Murmansk Region, Russia
2
Solid State Physics Department, Petrozavodsk State University (PetrSU), Petrozavodsk 185910, Republic of Karelia, Russia
*
Author to whom correspondence should be addressed.
Materials 2025, 18(2), 436; https://doi.org/10.3390/ma18020436
Submission received: 9 December 2024 / Revised: 6 January 2025 / Accepted: 15 January 2025 / Published: 18 January 2025
(This article belongs to the Topic Advances in Computational Materials Sciences)

Abstract

:
We proposed and investigated a refinement of technology for obtaining Mg-doped LiNbO3 (LN) crystals by co-doping it with B. LN:Mg (5.0 mol%) is now the most widely used material based on bulk lithium niobate. It is suitable for light modulation and transformation. We found that non-metal boron decreases threshold concentrations of the target dopant in many ways. In addition, we earlier determined that the method of boron introduction into the LN charge strongly affects the LN:B crystal structure. So we investigated the point structural defects of two series of LN:Mg:B crystals obtained by different doping methods, in which the stage of dopant introduction was different. We investigated the features of boron cation localization in LN:Mg:B single crystals. We conducted the study using XRD (X-ray diffraction) analysis. We have confirmed that the homogeneous doping method introduces an additional defect (MgV) into the structure of LN:Mg:B single crystals. Vacancies in niobium positions (VNb) are formed as a compensator for the excess positive charge of point structural defects. According to model calculations, boron is localized in most cases in the tetrahedron face common with the vacant niobium octahedron from the first layer (VNbIO6). The energy of the Coulomb interaction is minimal in the LN:Mg:B crystal (2.57 mol% MgO and 0.42 × 10−4 wt% B in the crystal); it was obtained using the solid-phase doping technology. The solid-phase doping technology is better suited for obtaining boron-containing crystals with properties characteristic of double-doped crystals (LN:Mg:B).

1. Introduction

Lithium niobate single crystal (LiNbO3, LN) is a ferroelectric nonlinear optical material that has attracted the attention of researchers for many decades due to its unique combination of properties and possible applications [1,2,3,4,5,6,7]. Recent studies have shown that LN single crystals are promising for a large number of applications. They can be used in the development of materials for holographic recording of information and coherent optical information processing systems [8,9,10,11], for generating terahertz radiation [12], for biomedicine [13], for nanoplasmonics [14,15], for creating compact crystalline accelerators that implement the generation of electron beams and soft X-ray radiation [16], for optical manipulation of micro- and nanoparticles [17,18] and water droplets [19,20,21]. Since LN is a phase of variable composition, the unique characteristics of the crystal are determined by the state of imperfection of its oxygen-octahedral structure. The Mg-doped and co-doped crystals are of the greatest interest, and materials based on LN:Mg find the widest application among all bulk LN [1,2,3,4,9,10,11]. However, practically important devices are made of Mg-doped bulk LN:Mg (5.0 mol%). This strongly suppresses optical damage and makes it suitable for laser radiation transformation as a part of periodically poled LN (PPLN) devices [10,11]. An amount of 5.0 mol% Mg in LN is considered an optimal concentration that balances useful properties and amount of defects. Meanwhile, 5.0 mol% is very close to the concentration threshold of 5.5 mol% MgO. Crystals with the dopant concentration above threshold change their properties and physical characteristics very sharply, which includes appearance of a great number of defects [2,9]. The defects prevent the use of above-threshold LN crystals in optics. Moreover, the closer the Mg concentration is to the threshold, the greater is the number of such defects, and 5.0 mol% still generates plenty of defects. Thus, it is of great relevance to find such a technology, that would provide LN:Mg crystals with the given properties but with the least possible defectivity. And that is the aim of this paper.
The crystal anionic structure contains big oxygen octahedra and small-volume vacant tetrahedral voids, Figure 1. Metallic cations are usually introduced into the O6 octahedral voids of LN. The influence of small-volume tetrahedral voids O4 of the LiNbO3 crystal on the formation of the defect structure remains unknown. Such features of the defect structure can seriously affect its physical characteristics. And metal cations cannot be introduced into the tetrahedral voids of O4 due to their large ionic radius.
The phase diagram of the Li2O-Nb2O5 system has a number of features [22]. Obtaining LN crystals of stoichiometric composition is complicated. However, technologies have now been developed that make it possible to grow LN crystals of stoichiometric (SLN) and near-stoichiometric (NSLN) composition. One approach is to grow crystals from a melt with an excess of an alkaline component (58.6 mol% Li2O) [1]. The disadvantages of this method are the low growth rate and large non-uniformity of the refractive index along the crystal growth axis. The VTE (Vapor Transport Equilibrium) method involves the diffusion of gaseous lithium into the structure of an LN sample [9,23]. This method increases the stoichiometry of only thin plates; therefore, it is unsuitable for large crystals. Another approach is the HTTSSG (High-Temperature Top Seeded Solution Growth) technology. It consists of growing crystals from a charge of congruent composition (R = [Li]/[Nb] = 0.946) with the addition of alkali metal fluxes, in particular K2O [24,25]. The main disadvantage of the method is the high concentration of potassium (1–2 × 10−2 wt%) in the grown crystals.
An alternative technological approach for obtaining NSLN is the addition of boron-containing doping components (B2O3, H3BO3) to a charge of congruent composition [26,27]. Boron-containing crystals of lithium niobate LiNbO3:B (hereinafter referred to as LN:B) have high structural and compositional uniformity and a low photorefractive effect. The non-metallic element boron is included in the crystal structure at the level of trace amounts (~4 × 10−4 mol%) and is localized in the faces of vacant tetrahedral voids [26]. The chemically active element boron structures the melt, binds excess niobium, regulates the cations of impurity metals inevitably present in the charge and prevents their transition into the structure of the growing crystal [26,27]. Such NSLN crystals should have a coercive field value as low as SLN, which is also crucial for the conversion of laser radiation in PPLN.
Currently, PPLN devices are made from LN crystals heavily doped with magnesium (≈5.0 mol% MgO) [28,29,30]. The compositional homogeneity of such crystals can decrease [31] due to the formation of a large number of structural defects [32].
Let us consider possible point defects and the models of their distribution. According to the Li-vacancy compensation model, in the crystal lattice of a congruent LN (CLN), there are ~1 mol. % NbLi point defects and ~4 mol % VLi point defects [33,34]. NbLi defects disrupt the ideal alternation of structural units of the cation sublattice in LiNbO3 crystals; the defects cause the photorefraction effect. To reduce the concentration of such defects, optical-damage-resistant ions (ODRIs) are deliberately introduced into the LN charge. This is a group of metal cations, for example, Zn2+, Mg2+, In3+, etc. [2,35,36,37,38]. However, such doping can not only reduce the concentration of NbLi defects but also form other defects (NbV, MeLi, MeV).
Thus, the classical ODRI dopant (Mg) reduces the photorefraction effect and the content of NbLi defects, and a non-metal (B) affects the “crystal-melt” system, the content of NbLi defects and the optical properties of such crystals. The mutual influence of these two dopants would be a fruitful approach to obtaining LN materials suitable for PPLN application. However, it should be handled with care, as an unexpected synergetic effect can arise. This means that approaches to the co-doping technology, appearing defects and their combinations should be investigated thoroughly [39,40].
The influence of boron on the structural features of LN crystals can be traced using Raman spectra. For example, Raman spectra revealed a distortion of the anion sublattice of LN:B crystals in work [41]. This was not observed in CLN and SLN crystals. A more informative method may be XRD. It provides reliable information on the coordinates of atoms and their spatial arrangement in the crystal lattice. This helps to determine the features of the localization of point defects in the crystal structure. The results of such studies, the main objects of which were LN:B crystals and some LiNbO3:Mg:B crystals, are given in [26,27,39]. However, XRD can be used to determine the position of atoms only heavier than boron.
Calculations of the Coulomb interaction can be both an independent method and part of a larger modeling; it is an important and indispensable tool in modern materials science. For example, the Coulomb interaction was applied and studied in various fields of materials science [42,43,44,45,46,47,48,49,50,51]. It is especially interesting that computer simulation including Coulomb interaction calculations were applied to study Mg-doped LN [52,53,54].
The work [26] determined the features of localization of boron in the structure of LN:B crystals obtained using homogeneous and direct doping technologies. The approach consisted of calculating the energy of the Coulomb interaction of point charges of fragments (clusters) of the crystal structure constructed on the basis of XRD data [27], with a boron cation. Boron was placed in different faces of vacant tetrahedral voids. Having analyzed 10 calculated clusters of different configurations, we established that in LN:B crystals, boron is localized preferentially in the faces of tetrahedra common with vacant niobium octahedra (VNb) [26]. This approach can provide important information about the features of boron localization in the structure of LN:B crystals, in particular, double-doped LN:Mg:B crystals containing point structural defects. The calculations in this work were also compared with the data on monodoped crystals LN:B(1-HG) and LN:B(2-SP) obtained in [26]. LN:B(1-HG) crystal was obtained by homogeneous doping, and LN:B(2-SP) by solid phase.
Thus, the aim of this work is to study the structural features of LN:Mg:B crystals of different genesis. We grew crystals from a charge obtained using homogeneous (HG) and solid-phase (SP) doping technologies in this work. The crystals were grown by stepwise dilution of the initial melt. The study was carried out by the Rietveld refinement of XRD patterns and using model calculations. Calculations revealed the features of localization of B3+ ions in the faces of vacant tetrahedral voids of such crystals, taking into account the detected point defects. An important task of the study is also the justification of the choice of the optimal doping technology for obtaining LN:Mg:B crystals with the least amount of structural defects.

2. Materials and Methods

Single crystals of LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) with a diameter of 35–38 mm and a cylindrical part length of 35–40 mm were grown in the (001) direction in an air atmosphere from platinum crucibles with a diameter of 85 mm by the Czochralski method on a growth setup Kristall-2 (Voroshilovgradsky zavod electronnogo mashinostroeniya, Voroshilovgrad, USSR). The setup is equipped with a thyristor generator and an automated control system. The thermal conditions and technological modes of growing single crystals are given in [39]. Numbers in designations of crystals LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) mean number of a crystal. Further explanations and details are given below.
Two types of charges were used to grow LN:Mg:B(1-8) single crystals. We used a homogeneously doped charge for growing LN:Mg:B(1-4-HG) crystals [55]. Boron and magnesium were introduced at the stage of niobium hydroxide precipitation from high-purity niobium-containing fluoride solutions. After that, we carried out high-temperature synthesis granulation of the Nb2O5:B:Mg-Li2CO3 mixture.
We used a solid-phase doped charge for growing LN:Mg:B(5-8-SP) crystals [56]. Boron and magnesium were introduced at the stage of preparation of the Li2CO3-Nb2O5-H3BO3-MgO mixture. This mixture is made in order to carry out the synthesis granulation of the charge. Boron was added in the form of boric acid (H3BO3), magnesium was added in the form of MgO (high purity, concentration of foreign impurities at the level of <5 × 10−4 wt%) [39]. In both cases, the following substances were used for the synthesis of the charge: niobium pentoxide Nb2O5 grade A produced using Technical Specifications 1763-025-00545484-2000 at Solikamsk magnesium works (Solikamsk, Russian Federation); high-purity lithium carbonate Li2CO3 with a concentration of impurities at a level of <3 × 10−4 wt% [39]. To synthesize the HG charge, HF (99.99, Vekton Ltd., Saint Petersburg, Russia) and NH4OH (25% solution, Komponent-reaktiv Ltd., Moscow, Russia) were used.
When calculating the required amount of Li2CO3 to obtain a LiNbO3 charge of congruent composition (R = 0.946), attention was paid to the magnesium content in the mixture. The concentration of boron can be compared with the concentration of trace amounts of impurities [26,41]. Therefore, it was not taken into account when determining the amount of Li2CO3. Synthesis granulation of mixtures (Nb2O5:B:Mg-Li2CO3 and Li2CO3-Nb2O5-H3BO3-MgO) was carried out at a temperature of ~1235–1245 °C for 5 h. The heating rate of the mixtures was ~200 °C/h [39].
Two series of crystals LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) were grown from the HG- and SP-doped charges by stepwise dilution of the initial melt. After growing each LN:Mg:B single crystal, the nominally pure lithium niobate charge of congruent composition was added to the melt remaining in the crucible [39]. Table 1 shows the content of doping elements in the grown crystals of LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP). The content of dopants in the melt is given in [39,40]. The crystals were grown in the following order: LN:Mg:B(4-HG) → LN:Mg:B(3-HG) → LN:Mg:B(2-HG) → LN:Mg:B(1-HG); LN:Mg:B(8-SP) → LN:Mg:B(7-SP) → LN:Mg:B(6-SP) → LN:Mg:B(5-SP).
Table 1 shows that in LN:Mg:B(3-HG) and LN:Mg:B(8-SP) crystals, the magnesium concentrations are very close to each other.
The concentration of magnesium in the charge and crystals was determined by AES (ICPE-9000, Shimadzu, Japan, Kyoto, 2011) with an accuracy of 4 × 10−3%. The boron content was determined by inductively coupled plasma mass spectrometry (ELAN 9000 DRC-e, PerkinElmer, Hopkinton, MA, USA) with an accuracy of 1 × 10−6%.
To carry out Rietveld refinement, XDR patterns of the studied crystals were recorded on a DRON-6 diffractometer (NPP Burevestnik, Saint Petersburg, Russian Federation) in monochromatic CuKα radiation (λ = 1.54178 Å) in the range of scattering angles 2θ from 5 to 145°. The Rietveld method was used to determine the unit cell parameters, the position of the doping element (Mg) in the structure, the type of intrinsic defects and their concentration. The MRIA and FULL PROF programs were used.
During the Rietveld refinement of powder XRD patterns of the LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) samples studied in the work, all possible models of the arrangement of intrinsic [57,58,59] and doping defects in the structure of the lithium niobate crystal were refined. Magnesium doping atoms were located in the positions of lithium (MgLi), niobium (MgNb) and the empty octahedron (Mgv). The position of boron in the structure was not specified due to its low concentration (~10−3 mol%) in the studied samples. A total of 30 probabilistic models of defect structure were refined for each sample. The criteria for selecting the final model for describing the defect structure were the minimum values of the agreement factors (Rw, Rwp), the stability of the refined parameters during Rietveld refinement, and the values of electroneutrality. The latter is calculated using the formula A + 5B + 2C = 6, where A is the number of lithium atoms, B is the number of niobium atoms in the main position, and in the defect node, C is the number of magnesium atoms.
The calculation of the total energy of the Coulomb interaction of point charges (EC, eV) of the fragment of the structure of LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) with the considered cation B3+ in the sp2-hybrid state was carried out using the Coulomb potential [26]:
E C = k i = 1 n q i q B r i B
where qi and qB are the values of the charge of the interacting particles in fractions of an electron; riB is the distance between the centers of the interacting charges [Å]; and k is a constant (eV⋅Å), expressed by the formula [26]:
k = e2/(4 · π · ε0 ·10−10) = 14.41971
where e is the electron charge, and ε0 is the dielectric constant.
When calculating the energy of Coulomb interaction, we used data obtained by X-ray structural analysis. The number of significant figures after the decimal point is 3. This determines the accuracy of the calculations.
The system (cluster) under consideration includes six oxygen octahedra in two layers: lower (I) and upper (II). For example, the notation for the placement of intrinsic cations in a cluster without defects and with ideal order would be as follows: LiIO6, NbIO6, VIO6, LiIIO6, NbIIO6 and VIIO6. Filling the constructed O6 octahedra with Li and Nb cations corresponds to filling the real LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals with these cations. The composition of such a cluster takes into account two tetrahedral voids formed by the selected oxygen octahedra. It should be noted that the system under consideration is not electrically neutral. In this work, we study the change in the energy of the Coulomb interaction of the B3+ cation with the surrounding fragment of the real crystal structure of LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals. We included in the consideration the defects NbLi, NbV, MgLi, MgV and VNb. The position of B3+ was taken to be equidistant from the oxygen atoms in each face of the vacant tetrahedral voids. The cluster framework is formed by 20 oxygen anions O2–. The number and localization of the main metal cations (Li+ and Nb5+) in the clusters differ and depend on the coordinates of the atoms, the type and number of defects introduced into the cluster. Structural defects NbLi, NbV, MgLi and MgV were detected in LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals during XRD patterns analysis. These defects determined the choice of models for describing the localization of intrinsic point defects in the studied crystals: the niobium vacancy model (M2 [35,59]) and the model of filling empty octahedra (M3 [1,59]).
In this work, we considered the formation of the following defects of different types: NbLi4• (niobium cation localized in the lithium octahedron), NbV5• (niobium cation localized in the vacant octahedron), MgLi (magnesium cation localized in the lithium octahedron), MgV2• (magnesium cation localized in the vacant octahedron), VNb5′ (vacancy in the niobium octahedron). The vacancy in the niobium octahedron is formed within the framework of charge compensation of the defects NbLi5+, NbV5+, MgLi2+ and MgV2+. Since the model calculations were performed using a limited fragment of the crystal structure (6 octahedra), we took into account the point charge of the defects introduced into the cluster, i.e., NbLi5+, NbV5+, MgLi2+ and MgV2+. The point charge may differ from the defect charge relative to the lattice.
The methodology for evaluating the influence of doping technology and the set of defective clusters on the features of the localization of the boron cation in the faces of vacant tetrahedral voids are given in [26].
For convenient description of the considered clusters, we introduce the following designations: «A.B.C», Table 2. In this designation, «A» is the crystal number (according to Table 1), «B» is the type of the considered cluster (0—without defects, 1—contains one NbLi defect, 2—contains one NbV defect, 3—contains one MgLi defect, 4—contains one MgV defect), «C» is the position of the considered defect (NbLi, NbV, MgLi or MgV) in the cluster (0—the defect is absent, 1—the defect is located in the 1st layer of the cluster, 2—in the 2nd layer of the cluster). For example, the entry «3.0.0» means that the defect-free cluster of crystal 3 is considered; the entry «4.2.1» means that the cluster of crystal 4 is considered, in which the NbV defect is located in the 1st layer.
As a result, 64 clusters were constructed and calculated based on XRD analysis data:
-
8 defect-free clusters (1.0.0–8.0.0);
-
16 clusters containing an NbLi defect (1.1.1–8.1.1, 1.1.2–8.1.2);
-
16 clusters containing an NbV defect (1.2.1–8.2.1, 1.2.2–8.2.2);
-
16 clusters containing an MgLi defect (1.3.1–8.3.1, 1.3.2–8.3.2);
-
8 clusters containing an MgV defect (1.4.1–4.4.2).
Several possible cationic compositions of model clusters of LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals are considered in this work. The periods of the unit cells (Table 3, Figure 1, work [40]) and the type and coordinates of the defects under consideration (NbLi, NbV, MgLi and MgV: Table 3 and work [39]) were taken into account when forming the clusters.

3. Results and Discussion

3.1. XRD Studies

Table 3 shows the atomic coordinates (x/a, y/b, z/c), site occupancies (G), and R-factors obtained during Rietveld refinement. Table 3 shows the data for the samples LN:Mg:B(2, 3-HG) and LN:Mg:B(5, 6-SP); for the other samples, the data are published in [39].
Analysis of the obtained models of the location of doping defects in the studied crystals showed that the mechanism of magnesium incorporation into the structure depends on the doping technology of the samples. Magnesium occupies only lithium positions in samples grown from the SP charge, and most of the magnesium atoms are additionally located in the empty octahedron in samples grown from the HG charge (Table 3, [39]). And the doping atoms are not located in the regular position of lithium; they are shifted either to the center of the octahedron or to the oxygen plane.
All samples contain intrinsic defects: niobium in the lithium site (NbLi), and niobium in the empty oxygen octahedron (Nbv). Charge compensation, when defects (NbLi, Nbv, MgLi, MgV) arise, occurs due to the formation of niobium vacancies (VNb). The number of lithium vacancies is small or equal to zero.
The concentration of MeV defects in samples grown from the HG charge is higher than in samples grown from the SP charge (Table 3, [39]). The presence of defects of this type disorders the cation sublattice along the polar axis and increases the defectiveness of the structure.
The periods and volume of the unit cell for all the studied double-doped crystals exceed the corresponding values for the NSLN (HTTSSG) crystal (a = 5.1428 Å, c = 13.8443 Å, V = 317.10 Å3 [39]). In crystals grown from HG charge, a non-monotonic dependence of the values of the unit cell periods on the dopant concentration in the crystal is observed, Figure 2. In crystals grown from the SP charge, when the magnesium concentration increases, the unit cell volume first decreases (LN:Mg:B(6-SP) crystal), and then the volume increases (LN:Mg:B(7 and 8-SP) crystals), Figure 2. The highest period value is achieved in the LN:Mg:B(8-SP) crystal at the maximum magnesium concentration in this series. The opposite trend is observed in crystals grown from the HG charge: the highest period value is recorded in the LN:Mg:B(1-HG) crystal which contains the minimum magnesium concentration in this series.

3.2. Coulomb Energy Calculation Studies

Figure 3 shows the EC value of clusters 1.0.0–8.0.0. These clusters correspond to fragments of LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals. They contain a boron cation in one of the seven faces of vacant tetrahedral voids considered in this work. The ideal alternation of the main metals (Li, Nb) and vacancies (V) along the polar axis of the crystal is considered for lithium niobate in these clusters. The alternation excludes the presence of point structural defects. Figure 3 additionally demonstrates the dependence of EC of the comparison Cluster 1. It is constructed on the basis of XRD structural data [1] for the CLN crystal. We obtained the data in an earlier study [26]. The EC for all possible positions of the boron cation in the faces of vacant tetrahedral voids of the clusters 1.0.0–8.0.0 is greater than for the control cluster, Figure 3. This can be explained by a more multicomponent “crystal-melt” system during the production of LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals, in contrast to the melt of a congruent composition which is simpler in composition. To synthesize the doped charge, which was used to grow LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals, in addition to the main components Li2CO3 and Nb2O5, MgO and H3BO3 were added to the reaction mixture. Such additives influence the type and concentration of electrochemical complexes in the melt. They determine the features of both the primary and secondary structure of the grown crystals.
The EC values of boron cations for clusters 1.0.0–8.0.0 have fairly close values (Δmax EC = 6 eV) in all positions considered, Figure 3. No significant differences in EC that could be explained by the technology of obtaining crystals and different concentrations of doping elements in them could be detected in the case of Clusters 1.0.0–8.0.0. The minor difference in EC found is due to the individual characteristics of each crystal (different periods of the unit cell, different distances between cations and anions of the clusters). More significant differences in EC will be shown further when considering point structural defects in clusters.
Figure 4a,b shows the results of EC calculations in Clusters 1.1.1–8.1.1 (abscissa axis in the upper part of the graph) and 1.1.2–8.1.2 (abscissa axis in the lower part of the graph). Recall that in Clusters 1.1.1–8.1.1, the NbLi defect is located in layer I, and VNb is located in layer II; in Clusters 1.1.2–8.1.2, the NbLi defect is located in layer II, and VNb is located in layer I. Common to all the clusters considered is the fact that the most energetically favorable position of boron is its localization on the face of the tetrahedron adjacent to the vacant niobium octahedron (VNbIO6), Figure 4. The EC of Clusters 1.1.2–8.1.2 for the VNbIO6 position is in the range from −593 to −604 eV, respectively. The EC of Clusters 1.1.1–8.1.1 for the VNbIIO6 position is in the range from −564 to −574 eV, respectively, Figure 4.
The EC of most of the possible boron positions calculated for HG crystal clusters has a large difference in values when compared with the EC calculated for SP crystal clusters (Figure 4).
For example, the EC for the VIO6 position in Clusters 1.1.2–4.1.2 of HG crystals varies from −429 to −449 eV, respectively (Figure 4). And the EC for SP crystals (Clusters 5.1.2–8.1.2) for the VIO6 position varies in a wider range from −443 to −451 eV, respectively. The EC for the OI4-OII4 position of HG crystals also varies greatly, Figure 4a: 1.1.2 (−435 eV), 2.1.2 (−410 eV), 3.1.2 (−414 eV), 4.1.2 (−416 eV). This is not observed for the OI4-OII4 position of SP crystals, Figure 4b: EC for Clusters 5.1.2–8.1.2 varies within the range from −426 to −435 eV. Such differences in EC for similar positions of the boron cation in clusters of the same configuration, but for crystals of different genesis, can be due to the following reasons.
Firstly, the homogeneous doping technology involves the introduction of a doping element at the stage of Nb2O5 precursor preparation. On the one hand, this increases the distribution coefficient of the dopant [39], and on the other hand, it can lead to a more complex composition of the ionic complexes. They participate in the crystallization process.
Secondly, the LN:Mg:B(1-4-HG) crystal series contains a higher concentration of magnesium cations than the LN:Mg:B(5-8-SP) crystal series (3.6–4.2 versus 2.57–3.82 mol%, respectively), Table 1.
Thirdly, the EC calculation method takes into account the relative positions of the oxygen framework, main cations, and point defects (as well as their charge component). In this case, the coordinates of both the main cations and defects may differ, see Table 3 and work in [39]. And the defects are determined by the technology of production, the type and concentration of dopants and the technological parameters of crystal growth. Such differences also appear when calculating EC.
NbV is the next defect. It was found in LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals (Table 3) and work [39]. The results of model calculations taking into account the presence of this defect in Clusters 1.2.1–8.2.2 are shown in Figure 5a,b. As in the case of Clusters 1.1.1–8.1.2, the localization of boron is most likely in the face of the tetrahedron which is common with the vacant niobium octahedron, Figure 5. The EC for the VNbIIO6 position in Clusters 1.2.1–4.2.1 and 5.2.1–8.2.1 varies within the range of −554 to −561 eV and −562 to −566 eV, respectively. In turn, the EC for the VNbIO6 position in Clusters 1.2.2–4.2.2 and 5.2.2–8.2.2 varies within the range of −551 to −574 eV and −544 to −549 eV, respectively. In the case of the formation of the NbV defect, the localization of the boron cation will be energetically favorable in the tetrahedron faces adjacent to VNb from both the first and second layers. In this case, the absolute minimum of the EC values corresponds to the position of VNbIO6 for Clusters 2.2.2 and 4.2.2 (−573 and −574 eV, respectively).
The behavior of the dependence of EC on the localization of the boron cation in seven different positions of Clusters 1.2.1–8.2.1 is similar for different crystals. Significant differences are observed for the NbVIO6, OI4–OII4, and LiIIO6 positions of Cluster 3.2.1 compared to Clusters 1.2.1, 2.2.1, 4.2.1, and 5.2.1–8.2.1, Figure 5. The decrease in EC for the indicated boron positions in Cluster 3.2.1 is due to the low location of the niobium cation in the NbVIO6 octahedron of the LN:Mg:B(3-HG) crystal, Table 3. As a result, the distance between the NbVIO6 defect from the first layer and the boron cations in the NbVIO6, OI4-OII4 and LiIIO6 positions increases for this crystal. This leads to a decrease in EC for these boron positions when Cluster 3.2.1 is considered.
The MgLi point defect was detected in all the crystals under study. As in the case of the formation of NbLi (Figure 4) and NbV (Figure 5) defects, the MgLi defect in Clusters 1.3.1–8.3.2 has a minimum EC value in cases where boron is localized near VNb, Figure 6. The absolute minimum EC is characteristic of the VNbIO6 position when boron is introduced into the magnesium-substituted lithium octahedron of the second layer (MgLiIIO6), Figure 6.
The EC remains almost unchanged when considering Clusters 1.3.1–8.3.2 with the exception of Clusters 1.3.2–4.3.2 (VIO6 and OI4–OII4 positions), Figure 6a. The MgLi defect in LN:Mg:B(2, 4-HG) crystals is localized lower than in LN:Mg:B(1, 3-HG) crystals (Table 3 and work [39]). This explains such different values of EC for the VIO6 and OI4-OII4 positions of Clusters 1.3.2–8.3.2, Figure 6a.
For Cluster 2.4.1, the EC in the positions of MgVIO6, OI4-OII4 and LiIIO6 is 7–15 eV less than the EC for the corresponding positions of boron in Clusters 1.4.1, 3.4.1 and 4.4.1, Figure 7. These differences, as in the previous cases, are explained by the peculiarity of the localization of the defect in question in the oxygen octahedron. The MgV defect in the oxygen octahedron of the LN:Mg:B(2-HG) crystal is localized lower than in the LN:Mg:B(1, 3, 4-HG) crystals (Table 3 and work [39]).
Thus, at this stage, we can conclude the following:
  • We considered Clusters 1.1.1–8.3.2 and 1.4.1–4.4.2. They contain NbLi, NbV, MgLi and MgV defects compensated by VNb. In them, the most energetically favorable position for the boron cation and the system as a whole will be its localization in the oxygen plane adjacent to the vacant niobium octahedron. The exception is Clusters 1.2.1–8.2.2. They contain the NbV defect. For these clusters, the presence of a boron cation in the faces of vacant tetrahedral voids common with the vacant niobium octahedron from the second layer (VNbIIO6) will be more energetically favorable. The exception is Clusters 2.2.2 and 4.2.2, for which the minimum EC is still characteristic of the VNbIO6 position, Figure 5.
  • The mutual arrangement of the structural units of the clusters (oxygen framework, main cations and structural defects) of real magnesium–boron co-doped LN crystals depends on several factors and affects the localization of boron cations in the structure. The factors are the doping technology, the type and concentration of dopants.
As mentioned above, there are several vacancy models for describing the defect structure of the LN crystal. The results of XRD analysis for boron-containing LN crystals were obtained: for LN:B(1-HG) and LN:B(2-SP) in [26]; LN:Mg:B(1, 4-HG) and LN:Mg:B(7, 8-SP) in [39]; LN:Mg:B(1, 4-HG) and LN:Mg:B(5-8-SP) in [40]; LN:Mg:B(2, 3-HG) in the present work. The results show that in the studied series of crystals, positively charged defect centers are compensated by vacancies in the niobium position. We believe that this is not a coincidence. Boron as an active complexing agent performs several actions: it binds excess niobium cations in a melt of congruent composition, complexes regulated cations of impurity metals, and structures the melt [27]. From a melt of congruent composition (with the addition of magnesium and boron or only boron), initially excess in niobium, a crystal grows which is close in stoichiometry to 1. The stoichiometry values of co-doped crystals LN:Mg:B(1-4-HG) and LN:Mg:B(4-8-SP) are 0.997, 0.971, 1.014, 0.985, 0.982, 0.973, 0.965 and 0.996, respectively, (Table 3 and work [39]). The stoichiometry values of single-doped crystals LN:B(1-HG) and LN:B(2-SP) are 0.985 and 1.033, respectively [26]. Consequently, part of the excess niobium is incorporated into lithium and vacant octahedra. To a greater extent, this mechanism is realized in LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals (Table 3 and work [39]); to a lesser extent, it is realized in LN:B(1-HG) and LN:B(2-SP) crystals [26]). In the case of LN:Mg:B(1-4 HG) and LN:Mg:B(5-8-SP) crystals, magnesium cations additionally act as a non-photorefractive additive (ODRI) and are incorporated into lithium octahedra. Magnesium is incorporated into vacant octahedra only in LN:Mg:B(1-4-HG) crystals. As a result, the proportion of occupied niobium octahedra decreases, and VNb is formed. Thus, when boron-containing dopants appear in the reaction mixture, vacancies in the niobium position (VNb5−) begin to compensate for the excess positive charge of point structural defects, instead of lithium vacancies (VLi-). That is, the presence of boron in the reaction mixture determines the type of vacancy model. It is worth noting that niobium vacancies are also observed in crystals doped with Er and Tb [60,61].
The lithium vacancy model cannot explain the NSLN composition of LN:Mg:B and LN:B crystals. To compensate for the positive defects, many more lithium vacancies are required than niobium vacancies. With such a number of lithium vacancies, the proportion of occupied lithium octahedra will decrease sharply which will lead to a decrease in the stoichiometry of such crystals.
In favor of the stated hypothesis, one can consider Figure 8. It shows the dependence of the change in the lengths of B-O bonds in the faces of vacant tetrahedral voids of the considered clusters on the concentration of magnesium in LN:Mg:B(1-4 HG) and LN:Mg:B(5-8-SP) crystals. It is well known that niobium octahedra are smaller than lithium and vacant ones because five covalent bonds and one ionic bond are formed between niobium and oxygens in NbO6 octahedra [1]. Boron has a significantly smaller ionic radius (for trivalent B3+—0.15 Å [62]) than oxygen (1.26 Å [62]). Therefore, it is difficult for boron to stay in the large-area faces of the tetrahedron. These are the faces common with the octahedra LiIO6, LiIIO6, VIO6, VIIO6 and the face OI4-OII4. The area of the faces common with NbIO6 and NbIIO6 is much smaller (Figure 8). Boron stays in them better.
Thus, the localization of boron near vacant niobium octahedra can be explained from several sides at once. This localization is explained by the change in the stoichiometry of the doped charge of congruent composition, the difference in the charges of B3+ and VI/IINb5− and the steric factor.
We have considered particular cases of real crystals in our paper. These cases include a specific defect. It was detected by the X-ray structural analysis method. The defect is compensated by VNb. However, the proportion of such “defective” clusters in the structure of a real crystal is very small. Therefore, in order to evaluate the influence of the set and number of the considered defect-containing clusters on the total EC in each specific position of the boron cation (seven positions), it is necessary to take into account the population factors G of the positions of the intrinsic metal cations (Li, Nb), defects (NbLi, NbV, MgLi, MgV) and the proportion of vacant (VV) and vacant niobium octahedra (VNb) (Table 3 and work [39]). This differs from the case when the ideal filling of the cluster with basic cations and vacancies is considered without taking into account defects (Figure 3). Therefore, we approach the next stage of adaptation of real X-ray structural data.
The formula for the calculation of the real contribution of energy of each boron position (EGC) to the whole crystal 1 is as follows (Cluster 1.0.0 is given as an example):
E G C ( 1.0 . 0 / L i I O 6 ) = E C ( 1.0 . 0 / L i I O 6 ) ·   K 1.0 . 0
where E G C ( 1.0 . 0 / L i I O 6 ) is the Coulomb interaction energy of a boron cation located in the LiIO6 face of a tetrahedron of Cluster 1.0.0, E C ( 1.0 . 0 / L i I O 6 ) is a corresponding value from Figure 3, and K1.0.0 is a coefficient taking into account the site population factors of the main metal cations and defects in the structure of the considered Cluster 1.0.0 (see [39]).
The coefficient is calculated using the formula:
K1.0.0 = G(LiLi) · G(NbNb) · G(VV)
where G(LiLi) is the population factor of Li atoms in lithium sites, G(NbNb) is the population factor of Nb atoms in niobium sites, and G(VV) is the amount of strictly vacant octahedra in vacant sites calculated using the Formula (5):
G1–4(VV) = 1 − G(NbV) − G(MgV)
where G(NbV) is the population factor of Nb atoms occupying vacant octahedra, and G(MgV) is the population factor of Mg atoms occupying vacant octahedra (only for crystals 1–4).
Strictly speaking, we cannot discuss vacancy ‘population’, since these are octahedra free of any cations; they are not populated. However, LN structure vacancies are crucial; we must take their amount into consideration.
Calculations of EGC(i) and Ki for all seven possible boron positions (in LiIO6, NbIO6, VIO6, OI 4-OII4, LiIIO6, NbIIO6, and VIIO6 faces of tetrahedra), made for Cluster 1.0.0 according to Formulas (3)–(5), will be similar for Clusters 2.0.0–4.0.0. For Clusters 5.0.0–8.0.0 in (5) G(VV) calculation considers only G(NbV):
G5–8(VV) = 1 − G(NbV)
The calculations for Clusters 1.1.1–8.1.1, 1.1.2-8.1.2 were similar, but they included data for the corresponding defect (NbLi) and compensated for the niobium vacancy (VNb). The formula for these clusters is given as an example, and was calculated correspondingly (Cluster 1.1.1 is given as an example):
E G C ( 1.1.1 / N b L i I O 6 ) = E C ( 1.1.1 / N b L i I O 6 ) ·   K 1.1.1 1.1.2
where E G C ( 1.1 . 1 / N b L i I O 6 ) is the Coulomb interaction energy of a boron cation located in the N b L i I O 6 face of a tetrahedron of Cluster 1.1.1, E C ( 1.1 . 1 / N b L i I O 6 ) is a corresponding value from Figure 4, and K1.1.1–1.1.2 is a coefficient calculated using the formula:
K1.1.1–1.1.2 = G(LiLi) · G(NbNb) · G(NbLi) · G1–4(VV) · G(VNb)
where G(NbLi) is the population factor of Nb atoms occupying lithium octahedra, and G(VNb) is the amount of vacant niobium octahedra calculated as:
G(VNb) = 1 − G(NbNb)
The coefficient K1.1.1–1.1.2 is used because the structure of Clusters 1.1.1 and 1.1.2 is different only in the mutual location of defects along the Z axis; all other data (the amount of defects in the cluster, the site population factor for them) are the same. The coefficients are identical for Clusters 2.1.1 and 2.1.2 (K2.1.1–2.1.2), 3.1.1 and 3.1.2 (K3.1.1–3.1.2), 4.1.1 and 4.1.2 (K4.1.1–4.1.2), etc.
For crystals 5–8, Formula (8) has a slightly different way of G(Vv) calculation (at an example of crystal 5):
K5.1.1–5.1.2 = G(LiLi) · G(NbNb) · G(NbLi) · G5–8(VV) · G(VNb)
For the remaining types of defects, the resulting Coulomb energy of the corresponding cluster was calculated according to the same algorithm as in Equations (7)–(10), but with the corresponding coefficients. We omitted them as their form is very similar and quite obvious. G values are taken from [39] or Table 3 and EC values from Figure 4, Figure 5, Figure 6 and Figure 7.
We obtained EGC(i) and summed them for crystals 1–4 in each position (crystal 1 is given as an example):
E s u m C ( L N : M g : B ( 1 H G ) / L i I O 6 ) = E G C ( 1.0.0 / L i I O 6 ) + E G C ( 1.1.1 / N b L i I O 6 ) + E G C ( 1.1.2 / L i I O 6 ) + E G C ( 1.2.1 / L i I O 6 ) + E G C ( 1.2.2 / L i I O 6 ) + E G C ( 1.3.1 / M g L i I O 6 ) + E G C ( 1.3.2 / L i I O 6 ) + E G C ( 1.4.1 / L i I O 6 ) + E G C ( 1.4.2 / L i I O 6 )
Since in crystals 5–8, the defective structure is slightly different, the least members of Equation (11) are absent (crystal 5 is given as an example):
E s u m C ( L N : M g : B ( 5 S P ) / L i I O 6 ) = E G C ( 5.0.0 / L i I O 6 ) + E G C ( 5.1.1 / N b L i I O 6 ) + E G C ( 5.1.2 / L i I O 6 ) + E G C ( 5.2.1 / L i I O 6 ) + E G C ( 5.2.2 / L i I O 6 ) + E G C ( 5.3 . 1 / M g L i I O 6 ) + E G C ( 5.3.2 / L i I O 6 )
The fundamental calculation method is given in [26]. The calculation results are shown in Figure 9. For comparison, Figure 9 also shows the EC values for LN:B(1-HG) and LN:B(2-SP) crystals studied in [26].
Figure 9 shows that the total EC dependences of LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals are 21–45 eV higher than EC of defect-free Clusters 1.0.0–8.0.0 (Figure 3). Such an increase in EC is due to all defects and their population coefficients.
The minimum value of EC for each individual dependence in Figure 9 is observed for the positions LiIO6, VIO6, OI4-OII4, LiIIO6 and VIIO6. This agrees well with our previously obtained results [26] and with the results in Figure 3. And the position of the dependences of EC on the localization of boron in different faces of the tetrahedrons has some features. For example, the minimum values of EC are characteristic of the LN:Mg:B(5-SP) crystal (Figure 9). The crystal has the minimum concentration of magnesium and boron of the two studied series of crystals, Table 1. Then, when the concentrations of magnesium and boron in the LN:Mg:B(6, 7, 8-SP) crystals increase, EC increases and takes on close values, Figure 9. If we move on to the LN:Mg:B(1-4-HG) crystals, EC behaves in the same way as in the case of the LN:Mg:B(5-8-SP) crystals. EC is minimal for the LN:Mg:B(1-HG) crystal with the minimal concentration of magnesium and boron in the series. The EC value of the LN:Mg:B(1-HG) crystal is approximately at the same level as the EC value of the LN:Mg:B(6, 7, 8-SP) crystals, Figure 9.
The minimum EC in each series of crystals is found in samples LN:Mg:B(1-HG) and LN:Mg:B(5-SP) (Figure 9) with the minimum concentration of magnesium and boron, Table 1. Two series were grown by dilution. That is, a nominally pure charge of congruent (R = 0.946) composition was added to the crucible after growing a crystal with the maximum concentration of magnesium and boron. This technique probably reduces the concentration of NbLi, NbV, MgLi and MgV defects in LN:Mg:B(4 → 1-HG) crystals and the MgLi defect in LN:Mg:B(8 → 5) crystals (Table 3 and work [39]). This is confirmed by the total occupancy of defect positions in crystals: 0.078 LN:Mg:B(4-HG) → 0.075 LN:Mg:B(3-HG) → 0.073 LN:Mg:B(2-HG) → 0.052 LN:Mg:B(1-HG); 0.053 LN:Mg:B(8-SP) → 0.065 LN:Mg:B(7-SP) → 0.060 LN:Mg:B(6-SP) → 0.051 LN:Mg:B(5-SP). If the decrease in magnesium concentration in the crystal series is explained by the dilution of the melt with a nominally pure charge, then for boron, this is also associated with its evaporation during the process of growing LN:Mg:B crystals.
In this case, the distribution coefficient close to 1 is favorable for the use of LN crystals. This is typical for LN:Mg:B(5-8-SP) crystals. For LN:Mg:B(1-4-HG) crystals, the distribution coefficient is significantly greater than 1, and it decreases rapidly over a fairly short concentration range [39]. Taking into account the data in Figure 9, we can say the following:
  • Growing LN:Mg:B crystals from a higher concentration to a lower one reduces the overall defectiveness of the crystals.
  • Despite the overall decrease in defectiveness, the LN:Mg:B(5-SP) crystal has the most stable structure. This is probably due to the absence of the MgV defect in it.
Co-doping with magnesium and boron promotes the evolution of the defect structure of LN:Mg:B crystals compared to single-doped LN:B. And the defectiveness of HG-doped crystals is higher. Indeed, NbLi and NbV defects are present in LN:B(1-HG) and LN:B(2-SP) crystals. The total population of these defects (0.025 and 0.028, respectively) is less than the total population of defects for LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals (Table 3 and work [39]).
The minimum EC of 10 crystals LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP), LN:B(1-HG) and LN:B(2-SP) is possessed by the crystal LN:B(1-HG), Figure 9. EC is slightly higher for the crystal LN:Mg:B(5-SP). EC of this crystal in almost all boron positions is lower than for the crystal LN:B(2-SP). The doping technologies for LN:B(2-SP) and LN:Mg:B(5-SP) crystals are the same, Figure 9. When monodoping is used, boron should be introduced into the LN crystal using a homogeneous method. And if the task is to obtain double-doped crystals (LN:Mg:B), then the technology of direct solid-phase doping is more advanced. In such crystals, the sequence of the main and doping cations along the growth axis will be preserved, and magnesium will not be introduced into vacant octahedra.
The contents of Mg and B in LN:Mg:B(3-HG) and LN:Mg:B(8-SP) crystals are close to each other. However, these crystals were obtained using different doping technologies. We compared the final calculation result of these crystals in Figure 9. The EC values for LN:Mg:B(8-SP) are much lower than those for LN:Mg:B(3-HG) for all boron positions. Moreover, the energy difference for each of the positions ΔEC = EC(LN:Mg:B(3-HG)) − EC(LN:Mg:B(8-SP)) is almost the same as the difference between the EC of LN:B(1-HG) and LN:B(2-SP). This finding confirms the importance of describing the complete doping and charge preparation technology and the growth parameters of doped LN single crystals when obtaining and interpreting any data, because crystals with the same dopant content obtained from charges of different genesis have different properties, from structure to distribution coefficient. Our conclusion confirms similar data for LN:Zn [63] obtained earlier. The conclusion about different properties of crystals with the same dopant concentration obtained by different technologies can partially explain the difference in the interpretation of many data concerning doped LN single crystals.

4. Conclusions

A study of two series of crystals LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) grown by the melt dilution method was carried out. Rietveld refinement for LN:Mg:B(2-HG) and LN:Mg:B(3-HG) was investigated for the first time. Models of boron cation localization in the faces of tetrahedrons of the LN structure were calculated. Analysis of the obtained models showed that the mechanism of magnesium incorporation into the crystal structure depends on the doping technology of the samples. It is shown that the HG method ensures the formation of a point defect MgV in the structure of LN:Mg:B(1-4-HG) crystals. It was not detected in LN:Mg:B(5-8-SP) crystals. All studied crystals were grown from a doped charge of congruent composition (R = 0.946) but have a stoichiometry value close to unity (0.965 < R < 1.014).
The presence of boron in the reaction mixture determines the type of realized vacancy model in the boron-containing crystal. This does not depend on the method of its introduction and the presence/absence of the second doping element. In such crystals, vacancies in the niobium position (VNb5−) compensate for the excess positive charge of point structural defects.
In this work, based on the data of real crystals, 64 clusters of various configurations were constructed and calculated. It was experimentally established that the mutual arrangement of the main (Li, Nb) and impurity (NbLi, NbV, MgLi, MgV) structural units of the clusters depends on the doping technology, type, concentration of dopants and crystal growth method. The mutual arrangement of the main (Li, Nb) and doping (NbLi, NbV, MgLi, MgV) structural units of the clusters affects the localization of boron cations in the structure of LN:Mg:B crystals. In Clusters 1.1.1–8.3.2 and 1.4.1–4.4.2, it is most advantageous for the boron cation to be localized in the tetrahedron face common with VNb. In this case, the absolute minimum of EC is achieved when boron is localized in the tetrahedron face common with VNbIO6. In this case, point defects NbLi, MgLi or MgV are located in the second layer of clusters. For Clusters 1.2.1–8.2.2, the opposite trend is observed: the absolute minimum of EC is recorded for the case when the NbV defect is localized in layer I, and the vacancy in the niobium position is localized in layer II. However, for Clusters 2.2.2 and 4.2.2, the trend characteristic of clusters with NbLi, MgLi or MgV defects is preserved. And the localization of boron near VNb is due to the increase in the stoichiometry of boron-containing crystals (≈1), grown from a charge of congruent composition (0.946), with a charge difference (B3+ and VI/IINb5−) and a steric factor.
Taking into account all the considered defect-containing and defect-free crystal clusters showed that the minimum EC in each series of grown crystals is possessed by the LN:Mg:B(1-HG) and LN:Mg:B(5-SP) samples, and the absolute minimum values are characteristic of the second crystal. This fact is very important for practice. This result is due to the fact that the melt was diluted when the crystals were grown, from the maximum concentration of dopants to the minimum. And this reduces the development of the defective structure of the crystals (reduction in the content of NbLi, NbV, MgLi and MgV in LN:Mg:B(4 → 1-HG) crystals; reduction in the content of the MgLi defect in LN:Mg:B(8 → 5-SP) crystals). The total population of defect positions in LN:Mg:B crystals decreases consistently (with the exception of the LN:Mg:B(8-SP) crystal). Thus, growing LN:Mg:B crystals from higher to lower concentrations reduces the total defectiveness of the crystals.
It has been established that double doping with magnesium and boron promotes the evolution of the defect structure of LN:Mg:B crystals compared to LN:B crystals. Its main contribution is reduced to a decrease in the concentration of NbLi defects.
Taking into account all the considered defect-containing and defect-free clusters of LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP), LN:B(1-HG) and LN:B(2-SP) crystals helps to make an important conclusion. When monodoping is used, boron should be introduced into the LN crystal by the homogeneous doping method, but when there is a need to obtain co-doped crystals, the direct solid-phase doping technology is more advanced.
Studies have shown that LN:Mg:B crystals with a small number of structural defects can be obtained at relatively low dopant concentrations (~2.50 mol% MgO) by the SP method. The absence of MgV-type defects in LN:Mg:B-SP crystals also supports this conclusion.
Our calculations also showed that LN:Mg:B(3-HG) and LN:Mg:B(8-SP) crystals with close dopant concentrations have different properties. Therefore, when comparing the properties of doped LN single crystals, it is critically important to describe all the details of the charge production, the technological parameters of growth for a complete understanding of the patterns of formation of these properties.

Author Contributions

Conceptualization, A.V.K. (ICT), R.A.T., D.V.M. and M.V.S.; methodology, R.A.T., A.V.K. (ICT) and S.M.M.; software, R.A.T.; validation, I.V.B.; formal analysis, O.V.T. and M.V.S.; investigation, R.A.T., A.V.K. (PetrSU), O.V.T., D.V.M. and I.V.B.; resources, M.N.P., I.V.B. and S.M.M.; data curation, R.A.T. and A.V.K. (PetrSU); writing—original draft preparation, R.A.T., A.V.K. (PetrSU), N.V.S. and D.V.M.; writing—review and editing, M.N.P., R.A.T., N.V.S. and D.V.M.; visualization, R.A.T.; supervision, R.A.T.; project administration, M.N.P.; funding acquisition, M.N.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Russian Science Foundation grant 24-13-20004. Reference to the grant page can be found here: https://rscf.ru/en/project/24-13-20004/ (accessed on 13 May 2024).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data required to reproduce these findings are available from authors A.V.K., R.A.T. and D.V.M. on reasonable request due to privacy.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Prokhorov, A.M.; Kuz’minov, Y.S. Physics and Chemistry of Crystalline Lithium Niobate; Adam Hilger: New York, NY, USA, 1990; 337p. [Google Scholar]
  2. Volk, T.; Wohlecke, M. Lithium niobate. In Defects, Photorefraction and Ferroelectric Switching; Springer: Berlin/Heidelberg, Germany, 2008; 250p. [Google Scholar]
  3. Fontana, M.D.; Bourson, P. Microstructure and defects probed by Raman spectroscopy in lithium niobate crystals and devices. J. Appl. Phys. Rev. 2015, 2, 040602. [Google Scholar] [CrossRef]
  4. Hong, N.; Cui, J.R.; Kim, H.J.; Shaffer, R.G.; Vinh, N.Q. New insights into refractive indices and birefringence of undoped and MgO-doped lithium niobate crystals at high temperatures. Opt. Mater. 2023, 145, 114365. [Google Scholar] [CrossRef]
  5. Schmidt, F.; Kozub, A.L.; Biktagirov, T.; Eigner, C.; Silberhorn, C.; Schindlmayr, A.; Schmidt, W.G.; Gerstmann, U. Free and defect-bound (bi)polarons in LiNbO3: Atomic structure and spectroscopic signatures from ab initio calculations. Phys. Rev. Res. 2020, 2, 043002. [Google Scholar] [CrossRef]
  6. Du, Y.; Pang, Z.; Zou, Y.; Zhu, B.; Liu, L.; Zhang, X.; Wang, C. Proton exchange-enhanced surface activated bonding for facile fabrication of monolithic lithium niobate microfluidic chips. Chem. Eng. J. 2024, 496, 154046. [Google Scholar] [CrossRef]
  7. Yu, S.; Fang, Z.; Zhou, Y.; Zhu, Y.; Huang, Q.; Ma, Y.; Liu, J.; Zhang, H.; Wang, M.; Cheng, Y. A high-power narrow-linewidth microlaser based on active-passive lithium niobate photonic integration. Opt. Laser Techn. 2024, 176, 110927. [Google Scholar] [CrossRef]
  8. Guo, Y.; Liu, L.; Liu, D.; Deng, S.; Zhi, Y. Absorption characteristic and nonvolatile holographic recording in LiNbO3:Cr:Cu crystals. Appl. Opt. 2005, 44, 7106–7111. [Google Scholar] [CrossRef]
  9. Sidorov, N.V.; Volk, T.R.; Mavrin, B.N.; Kalinnikov, V.T. Lithium Niobate: Defects, Photorefraction, Vibrational Spectrum, Polaritons; Nauka: Moscow, Russia, 2003; p. 255. (In Russian) [Google Scholar]
  10. Arizmendi, L. Photonic applications of lithium niobate crystals. Phys. Status Solidi 2004, 201, 253–283. [Google Scholar] [CrossRef]
  11. Gunter, P.; Huignard, J.-P. Photorefractive Materials and Their Applications; N.Y. Springer Science + Busines Media. LLC: New York, NY, USA, 2007; 647P. [Google Scholar] [CrossRef]
  12. Wang, Y.; Wang, R.; Yuan, J.; Wang, Y. Terahertz generation from Cu ion implantation into lithium niobate. J. Luminesc. 2014, 147, 242–244. [Google Scholar] [CrossRef]
  13. Blázquez-Castro, A.; García-Cabañes, A.; Carrascosa, M. Biological applications of ferroelectric materials. Appl. Phys. Rev. 2018, 5, 041101. [Google Scholar] [CrossRef]
  14. Wang, C.; Su, H.; Zhang, J.; Zhao, H. Surface metallization from ab initio theory and subwavelength coupling via surface plasmon polaritons in Cu-doped lithium niobate/tantalate owing to charge accumulation. Appl. Surf. Sci. 2021, 551, 149294. [Google Scholar] [CrossRef]
  15. Ma, X.; Xiong, Z.; Huo, D.; Wang, Y.; Su, H.; Wang, C.; Zhao, H. Long-ranged surface plasmon polaritons and their coupling with upconversion emissions in indium-tin-oxide-coated erbium and iron codoped LiNbO3. J. Opt. Soc. Am. B 2021, 38, 2984–2992. [Google Scholar] [CrossRef]
  16. Kukhtarev, N.V.; Kukhtareva, T.V.; Stargell, G.; Wang, J.C. Pyroelectric and photogalvanic crystal accelerators. J. Appl. Phys. 2009, 106, 014111. [Google Scholar] [CrossRef]
  17. Carrascosa, M.; García-Cabañes, A.; Jubera, M.; Ramiro, J.B.; Agulló-López, F. LiNbO3: A photovoltaic substrate for massive parallel manipulation and patterning of nano-objects. Appl. Phys. Rev. 2015, 2, 040605. [Google Scholar] [CrossRef]
  18. Mambetova, K.M.; Shandarov, S.M.; Tatyannikov, A.I.; Smirnov, S.V. Aggregation of dielectric nanoparticles on the x-cut of LiNbO3:Cu crystal by electric fields of photorefractive holograms. Russ. Phys. J. 2019, 62, 658–663. [Google Scholar] [CrossRef]
  19. Puerto, A.; Méndez, A.; Arizmendi, L.; García-Cabañes, A.; Carrascosa, M. Optoelectronic manipulation, trapping, splitting, and merging of water droplets and aqueous biodroplets based on the bulk photovoltaic effect. Phys. Rev. Appl. 2020, 14, 024046. [Google Scholar] [CrossRef]
  20. Zaltron, A.; Ferraro, D.; Meggiolaro, A.; Cremaschini, S.; Carneri, M.; Chiarello, E.; Sartori, P.; Pierno, M.; Sada, C.; Mistura, G. Optofluidic platform for the manipulation of water droplets on engineered LiNbO3 surfaces. Adv. Mater. Interf. 2022, 9, 2200345. [Google Scholar] [CrossRef]
  21. Zhang, X.; Mugisha, E.R.; Mi, Y.; Liu, X.; Wang, M.; Gao, Z.; Gao, K.; Shi, L.; Chen, H.; Yan, W. Photovoltaic cycling to-and-fro actuation of a water-microdroplet for automatic repeatable solute acquisition on oil-infused hydrophobic LN:Fe surface. ACS Photonics 2021, 8, 639–647. [Google Scholar] [CrossRef]
  22. Azuma, Y.; Uda, S. Electric current induced compositional variation in LiNbO3 fiber crystal grown by a micro-pulling down method. J. Cryst. Growth 2007, 306, 217–224. [Google Scholar] [CrossRef]
  23. Palatnikov, M.N.; Shcherbina, O.B.; Sandler, V.A.; Sidorov, N.V. Study of a stoichiometric lithium tantalate crystal obtained by VTE (Vapor transport equilibration) processing. Perspect. Mater. 2011, 2, 659–664. (In Russian) [Google Scholar]
  24. Lengyel, K.; Peter, A.; Kovacs, L.; Corradi, G.; Palfalvi, L.; Hebling, J.; Unferdorben, M.; Dravecz, G.; Hajdara, I.; Szaller, Z.; et al. Growth, defect structure, and THz application of stoichiometric lithium niobate. Appl. Phys. Rev. 2015, 2, 040601. [Google Scholar] [CrossRef]
  25. Polgar, K.; Peter, A.; Kovacs, L.; Corradi, G.; Szaller, Z. Growth of stoichiometric LiNbO3 single crystals by top seeded solution growth method. J. Cryst. Growth 1997, 177, 211–216. [Google Scholar] [CrossRef]
  26. Titov, R.; Kadetova, A.; Tokko, O.; Sidorov, N.; Palatnikov, M.; Teplyakova, N.; Masloboeva, S.; Biryukova, I.; Efremov, I.; Manukovskaya, D. Influence of Doping Technology on the Stoichiometry and Features of the Localization of B3+ Cations in LiNbO3:B Single Crystals. Crystals 2023, 13, 1245. [Google Scholar] [CrossRef]
  27. Palatnikov, M.N.; Sidorov, N.V.; Kadetova, A.V.; Titov, R.A.; Biryukova, I.V.; Makarova, O.V.; Manukovskaya, D.V.; Teplyakova, N.A.; Efremov, I.N. Growing, structure and optical properties of LiNbO3:B crystals, a material for laser radiation transformation. Materials 2023, 16, 732. [Google Scholar] [CrossRef] [PubMed]
  28. Shur, V.Y.; Akhmatkhanov, A.R.; Baturin, I.S. Micro- and nano-domain engineering in lithium niobate. Appl. Phys. Rev. 2015, 2, 040604. [Google Scholar] [CrossRef]
  29. Kemlin, V.; Jegouso, D.; Debray, J.; Boursier, E.; Segonds, P.; Boulanger, B.; Ishizuki, H.; Taira, T.; Mennerat, G.; Melkonian, J.-M.; et al. Dual-wavelength source from 5% MgO:PPLN cylinders for the characterization of nonlinear infrared crystals. Opt. Express 2013, 21, 28886–28891. [Google Scholar] [CrossRef]
  30. Murray, R.T.; Runcorn, T.H.; Guha, S.; Taylor, J.R. High average power parametric wavelength conversion at 3.31–3.48 μm in MgO:PPLN. Opt. Express 2017, 25, 6421–6430. [Google Scholar] [CrossRef]
  31. Sidorov, N.V.; Teplyakova, N.A.; Palatnikov, M.N. Influence of the method of doping on uniformity and optical properties of LiNbO3:Mg crystals. Phys. Chem. Asp. Study Clust. Nanostr. Nanomater. 2021, 13, 383–391. (In Russian) [Google Scholar]
  32. Wang, F.; Sun, D.; Liu, Q.; Song, Y.; Zhang, F.; Zhou, W.; Sang, Y.; Wang, D.; Liu, H. Growth of large size near-stoichiometric lithium niobate single crystals with low coercive field for manufacturing high quality periodically poled lithium niobate. Opt. Mater. 2022, 125, 112058. [Google Scholar] [CrossRef]
  33. Iyi, N.; Kitamura, K.; Izumi, F.; Yamamoto, J.K.; Hayashi, T.; Asano, H.; Kimura, S. Comparative study of defect structures in lithium niobate with different compositions. J. Solid State Chem. 1992, 101, 340–352. [Google Scholar] [CrossRef]
  34. Blumel, J.; Born, E.; Metzger, T. Solid state NMR study supporting the lithium vacancy defect model in congruent lithium niobate. J. Phys. Chem. Solids 1994, 55, 589–593. [Google Scholar] [CrossRef]
  35. Sánchez-Dena, O.; Villalobos-Mendoza, S.D.; Farías, R.; Fierro-Ruiz, C.D. Lithium Niobate Single Crystals and Powders Reviewed—Part II. Crystals 2020, 10, 990. [Google Scholar] [CrossRef]
  36. Kovács, L.; Kocsor, L.; Tichy-Rács, É.; Lengyel, K.; Béncs, L.; Corradi, G. Hydroxyl ion probing transition metal dopants occupying Nb sites in stoichiometric LiNbO3. Opt. Mater. Express 2019, 9, 4506–4516. [Google Scholar] [CrossRef]
  37. Kong, Y.; Bo, F.; Wang, W.; Zheng, D.; Liu, H.; Zhang, G.; Rupp, R.; Xu, J. Recent Progress in Lithium Niobate: Optical Damage, Defect Simulation, and On-Chip Devices. Adv. Mater. 2019, 32, 1806452. [Google Scholar] [CrossRef] [PubMed]
  38. Sulyanov, S.; Volk, T. Lattice Parameter of Optical Damage Resistant In-Doped LiNbO3 Crystals. Crystals 2018, 8, 210. [Google Scholar] [CrossRef]
  39. Palatnikov, M.; Makarova, O.; Kadetova, A.; Sidorov, N.; Teplyakova, N.; Biryukova, I.; Tokko, O. Structure, Optical Properties and Physicochemical Features of LiNbO3:Mg,B Crystals Grown in a Single Technological Cycle: An Optical Material for Converting Laser Radiation. Materials 2023, 16, 4541. [Google Scholar] [CrossRef]
  40. Kadetova, A.V.; Tokko, O.V.; Prusskii, A.I.; Smirnov, M.V.; Palatnikov, M.N.; Biryukova, I.V. The role of doping technology in the formation of nonlinear optical properties of LiNbO3:Mg:B crystals. Opt. Mater. 2024, 156, 115921. [Google Scholar] [CrossRef]
  41. Sidorov, N.V.; Teplyakova, N.A.; Makarova, O.V.; Palatnikov, M.N.; Titov, R.A.; Manukovskaya, D.V.; Birukova, I.V. Boron influence on defect structure and properties of lithium niobate crystals. Crystals 2021, 11, 458. [Google Scholar] [CrossRef]
  42. Vafek, O.; Kang, J. Renormalization Group Study of Hidden Symmetry in Twisted Bilayer Graphene with Coulomb Interactions. Phys. Rev. Lett. 2020, 125, 257602. [Google Scholar] [CrossRef]
  43. Cea, T.; Guinea, F. Coulomb interaction, phonons, and superconductivity in twisted bilayer graphene. Appl. Phys. Sci. 2021, 118, e2107874118. [Google Scholar] [CrossRef]
  44. Shi, J.; Zhao, N.; Yan, D.; Song, J.; Fu, W.; Li, Z. Design of a mechanically strong and highly stretchable thermoplastic silicone elastomer based on coulombic interactions. J. Mater. Chem. A 2020, 8, 5943–5951. [Google Scholar] [CrossRef]
  45. Chiou, D.-S.; Yu, H.J.; Hung, T.-H.; Lyu, Q.; Chang, C.-K.; Lee, J.S.; Lin, L.-C.; Kang, D.-Y. Highly CO2 Selective Metal–Organic Framework Membranes with Favorable Coulombic Effect. Adv. Funct. Mater. 2021, 31, 2006924. [Google Scholar] [CrossRef]
  46. Erkensten, D.; Brem, S.; Malic, E. Exciton-exciton interaction in transition metal dichalcogenide monolayers and van der Waals heterostructures. Phys. Rev. B 2021, 103, 045426. [Google Scholar] [CrossRef]
  47. Rømer, A.T.; Hirschfeld, P.J.; Andersen, B.M. Superconducting state of Sr2RuO4 in the presence of longer-range Coulomb interactions. Phys. Rev. B 2021, 104, 064507. [Google Scholar] [CrossRef]
  48. Miao, S.; Wang, T.; Huang, X.; Chen, D.; Lian, Z.; Wang, C.; Blei, M.; Taniguchi, T.; Watanabe, K.; Tongay, S.; et al. Strong interaction between interlayer excitons and correlated electrons in WSe2/WS2 moiré superlattice. Nat. Commun. 2021, 12, 3608. [Google Scholar] [CrossRef] [PubMed]
  49. Kiczynski, M.; Gorman, S.K.; Geng, H.; Donnelly, M.B.; Chung, Y.; He, Y.; Keizer, J.G.; Simmons, M.Y. Engineering topological states in atom-based semiconductor quantum dots. Nature 2022, 606, 694–699. [Google Scholar] [CrossRef] [PubMed]
  50. Li, Q.; Ning, D.; Wong, D.; An, K.; Tang, Y.; Zhou, D.; Schuck, G.; Chen, Z.; Zhang, N.; Liu, X. Improving the oxygen redox reversibility of Li-rich battery cathode materials via Coulombic repulsive interactions strategy. Nat. Commun. 2022, 13, 112. [Google Scholar] [CrossRef] [PubMed]
  51. Huang, J.; Wang, Y.; Wang, Y.; Wei, S.; Huang, Q.; Bai, S.; Wang, F.; Liu, Y.; Li, Z.; Chen, G.; et al. Altered S–S Coulomb Interactions in Polysulfides Associated with VS2/S Nanocomposite Electrodes for Na–S Batteries. ACS Appl. Nano Mater. 2024, 7, 7674–7683. [Google Scholar] [CrossRef]
  52. Donnerberg, H.J.; Tomlinson, S.M.; Catlows, C.R.A. Defects in LiNbO3. II. Computer simulation. J. Phys. Chem. Solids 1991, 52, 201–210. [Google Scholar] [CrossRef]
  53. Messerschmidt, S.; Krampf, A.; Vittadello, L.; Imlau, M.; Nörenberg, T.; Eng, L.M.; Emin, D. Small-polaron hopping and low-temperature (45–225 K) photo-induced transient absorption in magnesium-doped lithium niobate. Crystals 2020, 10, 809. [Google Scholar] [CrossRef]
  54. Kalimullah, N.M.M.; Shukla, K.; Shelke, A.; Habib, A. Stiffness tensor estimation of anisotropic crystal using point contact method and unscented Kalman filter. Ultrasonics 2023, 131, 106939. [Google Scholar] [CrossRef]
  55. Masloboeva, S.M.; Efremov, I.N.; Biryukova, I.V.; Palatnikov, M.N. Preparation and Characterization of Lithium Niobate Single Crystals Activated with Magnesium and Boron. Inorg. Mater. 2021, 57, 1271–1278. [Google Scholar] [CrossRef]
  56. Biryukova, I.V.; Efremov, I.N.; Masloboeva, S.M.; Palatnikov, M.N. Effect of the Doped Growth Charge Preparation Process on the Growth Conditions and Properties of LiNbO3:B:Mg Single Crystals. Inorg. Mater. 2022, 58, 948–955. [Google Scholar] [CrossRef]
  57. Lerner, P.; Legras, C.; Dumas, J.P. Stoechiométrie des monocristaux de métaniobate de lithium. J. Cryst. Growth 1968, 3, 231–235. [Google Scholar] [CrossRef]
  58. Peterson, G.E.; Carnevale, A. 93Nb NMR Linewidths in Nonstoichiometric Lithium Niobate. J. Chem. Phys. 1972, 6, 4848–4851. [Google Scholar] [CrossRef]
  59. Zotov, N.; Boysen, H.; Frey, F.; Metzger, T.; Born, E. Cation substitution models of congruent LiNbO3 investigated by X-ray and neutron powder diffraction. J. Phys. Chem. Solids 1994, 55, 145–152. [Google Scholar] [CrossRef]
  60. Palatnikov, M.N.; Kadetova, A.V.; Aleshina, L.A.; Sidorova, O.V.; Sidorov, N.V.; Biryukova, I.V.; Makarova, O.V. Growth, structure, physical and chemical characteristics in a series of LiNbO3:Er crystals of different composition grown in one technological cycle. Opt. Laser Technol. 2022, 147, 107671. [Google Scholar] [CrossRef]
  61. Palatnikov, M.N.; Sidorov, N.V.; Kadetova, A.V.; Teplyakova, N.A.; Makarova, O.V.; Manukovskaya, D.V. Concentration threshold in optically nonlinear LiNbO3:Tb crystals. Opt. Laser Technol. 2021, 137, 106821. [Google Scholar] [CrossRef]
  62. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Cryst. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  63. Palatnikov, M.N.; Birukova, I.V.; Masloboeva, S.M.; Makarova, O.V.; Manukovskaya, D.V.; Sidorov, N.V. The search of homogeneity of LiNbO3 crystals grown of charge with different genesis. J. Cryst. Growth 2014, 386, 113–118. [Google Scholar] [CrossRef]
Figure 1. A fragment of the LiNbO3 crystal structure which shows the relative positions of the oxygen octahedra O6 (vacant and containing Li or Nb cations) and the tetrahedra O4.
Figure 1. A fragment of the LiNbO3 crystal structure which shows the relative positions of the oxygen octahedra O6 (vacant and containing Li or Nb cations) and the tetrahedra O4.
Materials 18 00436 g001
Figure 2. Dependence of the unit cell periods (a and c, Å) on the magnesium concentration in LN:Mg:B(1-4-HG) (bulk dots) and LN:Mg:B(5-8-SP) (empty dots) crystals. Data for LN:Mg:B(1,4-HG) and LN:Mg:B(5-8-SP) crystals are taken from [40]).
Figure 2. Dependence of the unit cell periods (a and c, Å) on the magnesium concentration in LN:Mg:B(1-4-HG) (bulk dots) and LN:Mg:B(5-8-SP) (empty dots) crystals. Data for LN:Mg:B(1,4-HG) and LN:Mg:B(5-8-SP) crystals are taken from [40]).
Materials 18 00436 g002
Figure 3. Dependence of EC on the position of the B3+ cation in the faces of the tetrahedrons of clusters 1.0.0–8.0.0, and in control Cluster 1 [26]. The positions of boron in the clusters under consideration are indicated on the abscissa axis. A slanted arrow indicates which part of the graph is enlarged at an inset.
Figure 3. Dependence of EC on the position of the B3+ cation in the faces of the tetrahedrons of clusters 1.0.0–8.0.0, and in control Cluster 1 [26]. The positions of boron in the clusters under consideration are indicated on the abscissa axis. A slanted arrow indicates which part of the graph is enlarged at an inset.
Materials 18 00436 g003
Figure 4. Dependence of EC on the position of the B3+ cation in the tetrahedron face in clusters containing the NbLi defect: (a)—1.1.1–4.1.2; (b)—5.1.1–8.1.2. The abscissa axis in the upper part of the figures shows the positions of boron in Clusters 1.1.1–8.1.1 (bulk dots), and the lower part of the figures shows the positions of boron in Clusters 1.1.2–8.1.2 (empty dots). A slanted arrow indicates which part of the graph is enlarged at an inset. Straight arrows indicate which abscissa axis (upper or lower) should be taken into account when analyzing the graph.
Figure 4. Dependence of EC on the position of the B3+ cation in the tetrahedron face in clusters containing the NbLi defect: (a)—1.1.1–4.1.2; (b)—5.1.1–8.1.2. The abscissa axis in the upper part of the figures shows the positions of boron in Clusters 1.1.1–8.1.1 (bulk dots), and the lower part of the figures shows the positions of boron in Clusters 1.1.2–8.1.2 (empty dots). A slanted arrow indicates which part of the graph is enlarged at an inset. Straight arrows indicate which abscissa axis (upper or lower) should be taken into account when analyzing the graph.
Materials 18 00436 g004
Figure 5. Dependence of EC on the position of the B3+ cation in the tetrahedron faces in clusters containing the NbV defect: (a)—1.2.1–4.2.2.; (b)—5.2.1–8.2.2. The positions of boron are indicated on the abscissa axis in the upper part of the figures in clusters 1.2.1–8.2.1 (bulk dots), in the lower part of the figures—in clusters 1.2.2–8.2.2 (empty dots). A slanted arrow indicates which part of the graph is enlarged at an inset. Straight arrows indicate which abscissa axis (upper or lower) should be taken into account when analyzing the graph.
Figure 5. Dependence of EC on the position of the B3+ cation in the tetrahedron faces in clusters containing the NbV defect: (a)—1.2.1–4.2.2.; (b)—5.2.1–8.2.2. The positions of boron are indicated on the abscissa axis in the upper part of the figures in clusters 1.2.1–8.2.1 (bulk dots), in the lower part of the figures—in clusters 1.2.2–8.2.2 (empty dots). A slanted arrow indicates which part of the graph is enlarged at an inset. Straight arrows indicate which abscissa axis (upper or lower) should be taken into account when analyzing the graph.
Materials 18 00436 g005
Figure 6. Dependence of EC on the position of the B3+ cation in the tetrahedron faces in clusters containing the MgLi defect: (a)—1.3.1–4.3.2; (b)—5.3.1–8.3.2. The positions of boron are indicated on the abscissa axis in the upper part of the figures in Clusters 1.3.1–8.3.1 (bulk dots), in the lower part of the figures—in Clusters 1.3.2–8.3.2 (empty dots). A slanted arrow indicates which part of the graph is enlarged at an inset. Straight arrows indicate which abscissa axis (upper or lower) should be taken into account when analyzing the graph.
Figure 6. Dependence of EC on the position of the B3+ cation in the tetrahedron faces in clusters containing the MgLi defect: (a)—1.3.1–4.3.2; (b)—5.3.1–8.3.2. The positions of boron are indicated on the abscissa axis in the upper part of the figures in Clusters 1.3.1–8.3.1 (bulk dots), in the lower part of the figures—in Clusters 1.3.2–8.3.2 (empty dots). A slanted arrow indicates which part of the graph is enlarged at an inset. Straight arrows indicate which abscissa axis (upper or lower) should be taken into account when analyzing the graph.
Materials 18 00436 g006
Figure 7. Dependence of EC on the position of the B3+ cation in the tetrahedron faces in Clusters 1.4.1–4.4.2 containing the MgV defect. The abscissa axis in the upper part of the figure shows the positions of boron in the clusters. A slanted arrow indicates which part of the graph is enlarged at an inset. Straight arrows indicate which abscissa axis (upper or lower) should be taken into account when analyzing the graph.
Figure 7. Dependence of EC on the position of the B3+ cation in the tetrahedron faces in Clusters 1.4.1–4.4.2 containing the MgV defect. The abscissa axis in the upper part of the figure shows the positions of boron in the clusters. A slanted arrow indicates which part of the graph is enlarged at an inset. Straight arrows indicate which abscissa axis (upper or lower) should be taken into account when analyzing the graph.
Materials 18 00436 g007
Figure 8. Dependence of the change in the lengths of B-O bonds (Å) in the faces of the tetrahedra of the considered clusters on the concentration of magnesium (mol%) in LN:Mg:B(1-4 HG) and LN:Mg:B(5-8-SP) crystals.
Figure 8. Dependence of the change in the lengths of B-O bonds (Å) in the faces of the tetrahedra of the considered clusters on the concentration of magnesium (mol%) in LN:Mg:B(1-4 HG) and LN:Mg:B(5-8-SP) crystals.
Materials 18 00436 g008
Figure 9. Dependence of EC on the position of the boron cation (seven positions), taking into account the population factors of sites in LN:Mg:B(1-4-HG); LN:Mg:B(5-8-SP); LN:B(1-HG); LN:B(2-SP) crystals [26]. A slanted arrow indicates which part of the graph is enlarged at an inset.
Figure 9. Dependence of EC on the position of the boron cation (seven positions), taking into account the population factors of sites in LN:Mg:B(1-4-HG); LN:Mg:B(5-8-SP); LN:B(1-HG); LN:B(2-SP) crystals [26]. A slanted arrow indicates which part of the graph is enlarged at an inset.
Materials 18 00436 g009
Table 1. Concentration of magnesium and boron in LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals [39].
Table 1. Concentration of magnesium and boron in LN:Mg:B(1-4-HG) and LN:Mg:B(5-8-SP) crystals [39].
Doping TechnologyHGSP
Sample number12345678
[Mg] in crystal, mol%3.63.73.94.22.572.733.253.82
[B] in crystal cone, 10−4 wt%0.591.01.79.00.420.761.132.07
Table 2. Possible designations of letters in the cluster number “A.B.C.”.
Table 2. Possible designations of letters in the cluster number “A.B.C.”.
ABC
10 No defects
1 NbLi
2 NbV
3 MgLi
4 MgV
0 No defects
1 Defect in layer I
2 Defect in layer II
2
3
4
5
6
7
8
Table 3. The atomic coordinates (x/a, y/b, z/c), site occupancies (G, site occupancy), and R-factors obtained during Rietveld refinement for the samples LN:Mg:B(2, 3-HG) and LN:Mg:B(5, 6-SP).
Table 3. The atomic coordinates (x/a, y/b, z/c), site occupancies (G, site occupancy), and R-factors obtained during Rietveld refinement for the samples LN:Mg:B(2, 3-HG) and LN:Mg:B(5, 6-SP).
Gx/ay/bz/c Gx/ay/bz/c
LN:Mg:B(2-HG): C(Mg) = 3.7 mol%
Rwp(%) = 6.76, Rp(%) = 9.62
LN:Mg:B(3-HG): C(Mg) = 3.9 mol %
Rwp(%) = 6.59 Rp(%) = 8.44
Nb0.956000Nb0.925000
O1.00.05510.32490.0686O1.00.05470.32550.0634
Li0.958000.2950Li0.969000.2783
NbLi0.014000.2732NbLi0.013000.2670
NbV0.017000.1550NbV0.018000.1109
MgLi0.017000.2690MgLi0.018000.2900
MgV0.025000.1212MgV0.026000.1470
LN:Mg:B(5-SP): C(Mg) = 2.57 mol%;
Rwp(%) = 6.78, Rp(%) = 6.39
LN:Mg:B(6-SP): C(Mg) = 2.73 mol %
Rwp(%) = 8.12, Rp(%) = 6.36
Nb0.957000Nb0.955000
O1.00.05440.34730.0637O1.000.06420.33870.0643
Li0.963000.2807Li0.957000.2821
NbLi0.015000.2803NbLi0.012000.2900
NbV0.009000.1100Nbv0.017000.1100
MgLi0.027000.2756MgLi0.031000.2877
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

Titov, R.A.; Kadetova, A.V.; Manukovskaya, D.V.; Smirnov, M.V.; Tokko, O.V.; Sidorov, N.V.; Biryukova, I.V.; Masloboeva, S.M.; Palatnikov, M.N. Features of the Defect Structure of LiNbO3:Mg:B Crystals of Different Composition and Genesis. Materials 2025, 18, 436. https://doi.org/10.3390/ma18020436

AMA Style

Titov RA, Kadetova AV, Manukovskaya DV, Smirnov MV, Tokko OV, Sidorov NV, Biryukova IV, Masloboeva SM, Palatnikov MN. Features of the Defect Structure of LiNbO3:Mg:B Crystals of Different Composition and Genesis. Materials. 2025; 18(2):436. https://doi.org/10.3390/ma18020436

Chicago/Turabian Style

Titov, Roman A., Alexandra V. Kadetova, Diana V. Manukovskaya, Maxim V. Smirnov, Olga V. Tokko, Nikolay V. Sidorov, Irina V. Biryukova, Sofja M. Masloboeva, and Mikhail N. Palatnikov. 2025. "Features of the Defect Structure of LiNbO3:Mg:B Crystals of Different Composition and Genesis" Materials 18, no. 2: 436. https://doi.org/10.3390/ma18020436

APA Style

Titov, R. A., Kadetova, A. V., Manukovskaya, D. V., Smirnov, M. V., Tokko, O. V., Sidorov, N. V., Biryukova, I. V., Masloboeva, S. M., & Palatnikov, M. N. (2025). Features of the Defect Structure of LiNbO3:Mg:B Crystals of Different Composition and Genesis. Materials, 18(2), 436. https://doi.org/10.3390/ma18020436

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