Next Article in Journal
Assessment of Musculoskeletal Pain and Physical Demands Using a Wearable Smartwatch Heart Monitor among Precast Concrete Construction Workers: A Field Case Study
Next Article in Special Issue
Triangular Silver Nanoparticles Synthesis: Investigating Potential Application in Materials and Biosensing
Previous Article in Journal
Visible—Light Driven Systems: Effect of the Parameters Affecting Hydrogen Production through Photoreforming of Organics in Presence of Cu2O/TiO2 Nanocomposite Photocatalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of the Structural Perfection of a LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) Double-Doped Single Crystal Using the Raman Spectra Excited by Laser Lines in the Visible (532 nm) and Near-IR (785 nm) Regions

by
Nikolay Sidorov
1,
Mikhail Palatnikov
1,
Alexander Pyatyshev
2,* and
Alexander Skrabatun
2,3
1
Tananaev Institute of Chemistry—Subdivision of the Federal Research Centre “Kola Science Centre of the Russian Academy of Sciences”, “Academic Town”, 26a, Murmansk Region, 184209 Apatity, Russia
2
P.N. Lebedev Physical Institute of the Russian Academy of Sciences, Leninskiy Prospekt 53, 119991 Moscow, Russia
3
Physics Department, Faculty of Fundamental Sciences, Bauman Moscow State Technical University, 2nd Baumanskaya St. 5/1, 105005 Moscow, Russia
*
Author to whom correspondence should be addressed.
Appl. Sci. 2023, 13(4), 2348; https://doi.org/10.3390/app13042348
Submission received: 19 December 2022 / Revised: 1 February 2023 / Accepted: 9 February 2023 / Published: 11 February 2023
(This article belongs to the Special Issue Nanotechnology and Functional Nanomaterials)

Abstract

:
A compositionally homogeneous nonlinear optical single crystal of double-doped LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) was obtained. Fine features of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal structure were studied from the Raman spectra of the first and second orders upon excitation by laser lines in the visible (532 nm) and near-IR (785 nm) regions. When the Raman spectrum was excited by a 785 nm laser line in the frequency range of 1000–2000 cm−1 for the first time, a number of low-intensity lines in the range of 900–2000 cm−1, corresponding to the second-order Raman spectrum, were discovered. The same lines also appear in the spectrum upon excitation by a laser line with a wavelength of 532 nm, but their intensities are significantly (by an order of magnitude or more) lower. It is shown that in the structure of the double-doped LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%), the crystal oxygen-octahedral clusters MeO6 (Me–Li, Nb, Gd, Mg) are slightly distorted, and in addition, the value R = [Li]/[Nb] ≈ 1 is close to that for a nominally pure stoichiometric crystal.

1. Introduction

An urgent task of modern physical materials science is to obtain highly advanced optical materials with a low photorefraction effect (optical damage) based on a nonlinear optical single crystal of lithium niobate (LiNbO3). Lithium niobate is an oxygen-octahedral phase of variable composition with a wide homogeneity region on the phase diagram [1,2]. The set of optical properties of this material—the magnitudes of electro-optical and nonlinear optical coefficients, the sensitivity to the holographic recording of information, and the possibility of obtaining laser generation with frequency self-doubling—make it universal for optical applications, such as quartz in acoustics. One of the reasons for the universality of lithium niobate as a phase of variable composition is the possibility of controlling its properties over a wide range by varying the composition through doping and changing the stoichiometry, which is especially attractive for the development of integrated optical devices.
Among the properties that strongly depend on the composition is the effect of a photoinduced change in the refractive indices (optical damage), which leads to the distortion of the wave front of a laser beam passing through the crystal [1]. The existence of the photorefraction effect causes two important alternative practical problems: finding ways to suppress its (i.e., obtaining non-photorefractive, optical-damage resistant”) compositions for traditional applications, and the optimization of photorefractive properties; for example, increasing the sensitivity and speed of information recording. For practice, the first task is much more important, since the use of a lithium niobate crystal in nonlinear optics as an active nonlinear laser medium, as well as for converting and modulating laser radiation, remains dominant. This task is especially relevant in the current intensive development of the method of optical frequency conversion in the regime of phase quasi-phase matching on regular domain structures (PPLN—“periodically-poled LiNbO3”) [3].
LiNbO3:Mg crystals are currently the most promising for the creation of highly advanced materials for radiation conversion [3,4,5,6,7,8]. In the case of double doping, when one of the alloying elements is magnesium, it can be used to create optical materials of increased compositional homogeneity with a minimum photorefractive response time and a high resistance to optical damage [9,10,11,12,13,14,15,16]. In highly perfect double-doped LiNbO3 crystals, the fundamental absorption edge shifts to shorter wavelengths, and a noticeable increase in nonlinear optical coefficients is observed vs. one-time doped crystals [11,12,14,16,17]. Co-doping simultaneously with two “non-photorefractive” cations (Mg2+ and Gd3+) makes it possible to control the ordering of structural units of the cationic sublattice along the polar axis and the polarizability of MeO6 clusters (Me–Li, Nb, vacancy, and impurity metal) more finely than single doping. It also allows one to control the type and concentration of point and complex defects with localized electrons, which determine the magnitude of the photorefraction effect (optical damage) [1,11]. This is due to the fact that both the Mg2+ and Gd3+ cations occupy mainly lithium positions in LiNbO3 crystals [2]. Therefore, in co-doped LiNbO3:Gd:Mg crystals, there is competition for the lithium positions between the Mg2+ and Gd3+ cations, which leads to a change in the number and type of defects, which are shallow and deep electron traps that increase the photorefraction effect. There is no such competition in LiNbO3:Mg and LiNbO3:Gd [1,2].
Thus, it is of considerable interest to establish the possibility of controlling the optical and electrical properties of a LiNbO3 crystal by doping LiNbO3 crystals with various cations or groups of cations.
Raman spectroscopy (RS) is a well-known method for studying various organic and inorganic compounds. This method has a high sensitivity, which makes it possible to study objects under difficult conditions. In particular, art objects of various ages [18,19,20], food [21,22], viruses and bacteria [23,24,25], recycled polyethylene terephthalate [26], and so on, have previously been studied. The high sensitivity of RS to the slightest changes in the interaction between structural units is an important factor for studying the crystalline perfection of LiNbO3. An important feature of the polar ferroelectric LiNbO3 is that the transverse (TO) and longitudinal (LO) vibrations of the crystal lattice are characterized by a strong interaction with electromagnetic radiation, which excites the Raman spectra [2,27]. As mentioned above, lithium niobate is a photorefractive crystal. In this regard, an additional volumetrically ordered sublattice of nano- and microstructures appears in the illuminated region of the crystal. In this region, there are changes in the refractive index, permittivity, conductivity, and other parameters due to the photoreaction effect. As a result, the Raman scattering lines are shifted in frequency, their intensity changes, and they broaden. For this reason, the Raman spectra obtained upon excitation by lasers with different wavelengths can noticeably differ from each other [28]. This may be due to the different sensitivities of micro- and macrostructures, MeO6 clusters, point defects in the form of irregularly located main and impurity atoms, and the effects of structural disorder and anharmonicity that exist in the LiNbO3 crystal as a photorefractive phase of variable composition [1,2,27], and the effects of radiation over a wide spectral range.
In this work, in the frequency range of 50–4000 cm−1, with the excitation of the spectra in the visible (532 nm) and near-IR (785 nm) regions, the full Raman spectrum of a double-doped LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) is measured. Gd3+ and Mg2+ cations, which have different valences and ionic radii, are the non-photorefractive additives that differently affect the state of the crystal defect sublattice, spontaneous polarization along the polar axis, the geometry of NbO6 oxygen-octahedral clusters, and the effect of photorefraction (optical damage). Previously, the Raman spectra of LiNbO3:Gd:Mg crystals of different compositions were studied in the literature only in the range of 50–1000 cm−1 [27,29]. According to the first-order Raman spectra in the frequency range of 50–1000 cm−1 in single-doped LiNbO3:Gd3+ and LiNbO3:Mg2+ crystals at low concentrations of Gd3+ (0.002 wt.%) and Mg2+ (0.030–0.078 wt.%), there is an increase in the ordering of structural units of the cationic sublattice along the polar axis, and a noticeable decrease in the photorefraction effect [25,27]. The same effects are observed for heavily doped Mg2+ (0.40–0.65 wt.%) crystals of double-doped LiNbO3:Gd3+:Mg2+ at low concentrations of the second doping component Gd3+ (0.001–0.23 wt.%) [27,29]. At high levels of magnesium doping (>3.0 wt.%), additional lines appear in the Raman spectra that conform to pseudoscalar fundamental lattice vibrations of the A2 symmetry type, which are not allowed in the Raman spectra for the R3c symmetry space group [30]. It was also found that a magnification of the Mg amount in single and co-doped LiNbO3 leads to an increase in the Raman intensity on vibrations of the E modes without any spectral shift [9,31,32].

2. Materials and Methods

A single crystal of LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) was grown in air from a melt using the Czochralski method on a Kristall 2 (ZavodKristall Ltd., Likino village, Russia) induction growth setup with automatic control of the crystal diameter (Figure 1) in a platinum crucible 75 mm in diameter with a small (2.5 deg/cm) axial gradient in the direction of the polar axis (z-cut, (001)). The single-crystal boule rotation speed was ~16–18 rpm, and the crystal pulling speed was ~0.7 mm/h. In this case, the crystal growth rate was ~1.02–1.04 mm/h. The technological parameters of growth corresponded to the condition of a flat crystallization front. The direct alloying of the congruent melt ([Li]/[Nb] = 0.946) with MgO and Gd2O3 was carried out. To relieve thermoelastic stresses, the grown crystal was subjected to heat treatment in an air atmosphere at 1500 K for 15 h in a high-temperature furnace, Lantan (Voroshilovgradsky zavod electronnogo mashinostroeniya, Voroshilovgrad, USSR).
Growing optically and compositionally uniform doubly doped lithium niobate crystals is a non-trivial technological problem. Doping additives, as a rule, have impurity distribution coefficients that differ (in our case, strongly differ) in magnitude: KD(Mg) ≈ 1.1, KD(Gd) ≈ 0.25 [33]. Consequently, the melt composition near the solidification front during crystal growth can be simultaneously enriched in one alloying component and depleted in another. Thus, the composition of a doped crystal during growth can change significantly from the cone to the end of the crystal, which usually leads to a decrease in its compositional and optical homogeneity. In this case, the characteristics of the crystal can significantly vary within its different parts. To minimize such effects, it is necessary to apply changes in parameters that are natural for the process of crystal growth. These are the speed of rotation and movement of the crystal, temperature gradients in the melt and growth zone, and different conjunctions of these parameters. In our case, such technological methods, including the use of a special design of the thermal unit, which creates small temperature gradients at the crystallization front, the use of low crystallization rates, particular melt preparations before crystal growth, long postgrowth annealing, and suitable conditions for the electrothermal treatment of the crystal, are effective.
Figure 1. Kristall 2 growth setup for growing lithium niobate crystals using the Czochralski method—(a); crystal LiNbO3:Gd3+(0.003):Mg2+ (0.65 wt.%) (b). Figure is reproduced with the permission of Elsevier from the paper [34].
Figure 1. Kristall 2 growth setup for growing lithium niobate crystals using the Czochralski method—(a); crystal LiNbO3:Gd3+(0.003):Mg2+ (0.65 wt.%) (b). Figure is reproduced with the permission of Elsevier from the paper [34].
Applsci 13 02348 g001
When growing a single crystal of double-doped LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%), we used the design of a thermal unit with double “warming”, which allows for the creation of an isothermal zone in the volume of the platinum screen for the postgrowth annealing of the crystal, and the growth of the crystal under the conditions of a small temperature gradient at the crystallization front. The design of the thermal unit was similar to that used in [34]. Figure 1b shows the grown LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal.
The crystal growth process was finished when the LiNbO3:Gd:Mg crystal’s mass of less than 275 g was reached. In this case, about ~25% of the total mass of the melt crystallized. The parameters of the LiNbO3:Gd:Mg crystal growth process (pulling speed, rod rotation speed, and temperature gradient at the crystallization front) were selected experimentally, based on the need to obtain a flat crystallization front, which should ensure a sufficiently high structural perfection of the crystal. The estimated mass during the growth of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal was determined from the readings of the weight sensor of the growth setup; the accurate mass of the crystal was determined by weighing the grown crystal after separating the seed from the crystal boule. The grown crystals had a flat crystallization front and geometric dimensions: diameter (Ø) ≈ 38 mm, length of the cylindrical part Lc ≈ 40 mm. Impurities were introduced into the charge in the form of high-purity oxides of MgO and Gd2O3 with a concentration of foreign admixtures of no greater than 3 × 10−4 wt.%, followed by thorough mixing. The melt before the start of crystal growth was kept for 8–11 h under superheating conditions by 180–200 °C relative to the melting temperature (Tmelt = 1263 °C) of lithium niobate, to homogenize the impurities in the melt. After growth, the Mg crystal was annealed at 1200 °C in a growth setup for 10 h, and then cooled at a rate of ~50 deg/h. Long-term post-growth annealing in the isothermal zone (same as in [34]) under a platinum screen is required to homogenize the composition of the doped crystal, and to remove the thermal and mechanical stresses.
The monodomainization of the LiNbO3:Gd:Mg crystal was carried out via high-temperature electrodiffusion annealing, namely by applying a constant electric voltage to the polar cuts of the crystal during its cooling, at a rate of 20 deg/h in the temperature range ~1230–870 °C.
For the synthesis of the lithium niobate charge, Nb2O5 grade A was used, and produced using Technical Specifications No. 1763-025-00545484-2000 at Solikamsk magnezium works (Solikamsk, Russia), and with Li2CO3 of high purity, with an amount of foreign admixtures of no greater than 3 × 10−4 wt.%. From these initial components, a granular charge of a congruent composition ([Li2O] = 48.6 mol.%) with a high bulk density (~3.4 g/cm3) was obtained, using the synthesis–granulation method. The preparation of the charge is described in detail in [35].
To estimate the impurity content in the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal, and the distribution features of the change in their concentration along the length of the crystal, samples were studied from the upper (cone) and lower cylindrical parts of the boule. Inductively coupled plasma atomic emission spectroscopy (ICP-AES) on an Optima 8300 ICP-OES (Mg) (PerkinElmer, Waltham, MA, USA) spectrometer, and atomic absorption spectrometry (AAS) on a Kvant-FA instrument (Zn) (GRANAT, Saint Petersburg, Russia) were used to determine the concentration of dopants in milled samples. Table 1 shows the admixture concentrations of the granular mixture and the studied LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) single crystal.
The sample for studying the Raman spectra was a parallelepiped with dimensions of 4.93 × 5.98 × 8.41 mm3, and edges coinciding with the direction of the principal crystallographic axes. The faces of the parallelepiped were thoroughly burnished.
To record the Raman spectra in the visible region, we used a BWS465-532S i-Raman Plus spectrometer (B&W Tek, Plainsboro Township, NJ, USA), which has a wavelength of 532 nm and allows for the recording of spectra in the range of 50–4000 cm−1. The power of laser radiation during the registration of spectra was 30 mW. The numerical aperture was ≈0.22. To record the Raman spectra in the near-IR region, a BWS465-785H i-Raman Plus spectrometer (B&W Tek, Plainsboro Township, NJ, USA) was used, with an excitation wavelength of 785 nm and with the allowance of recording spectra in the range of 50–2850 cm−1. The power of laser radiation during the registration of spectra was 340 mW. The numerical aperture was ≈0.22. The laser spot size at the focus was 85 μm. All spectra were recorded at room temperature using backscattering geometry. In order to minimize the local influence of the exciting laser radiation, in each experiment, we selected the optimal regimes of radiation, focusing on the crystals under study and the accumulation time of the useful signal.

3. Results and Discussion

The composition of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal on the phase diagram of the Nb2O5-Li2O system is within the homogeneity region (solid solution region). Therefore, the unit cell of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal, such as the unit cell of the ferroelectric phase of the nominally pure LiNbO3 crystal, is characterized by the space symmetry group C 3 V 6 (R3c) and contains two formula units (10 atoms) [2]. In this case, the doping of the Gd3+ and Mg2+ cations in the structure of the LiNbO3 crystal violate the order of alternation of the Li+ and Nb5+ cations, and vacancies (V) along the polar axis, and it distorts the geometry of the oxygen-octahedral MeO6 clusters, which is characteristic of a nominally pure crystal of congruent composition. The main Li+ and Nb5+ ions, as well as the doping of the Mg2+ and Gd3+ ions, occupy the C3 position, while the O2− ions occupy the C1 position. The phonon dispersion curve thus has 30 vibrational branches, of which 27 are optical and 3 are acoustic. The optical vibrational representation of a LiNbO3 crystal has the following form:
Γ = 5A1(z) + 5A2 + 10E(x, y).
In Raman scattering and IR absorption at k = 0 (at the center of the Brillouin zone), active 4A1(z) + 9E(x, y) dipole-active fundamental vibrations occur, respectively, along and perpendicular to the polar axis. Due to the polar nature of all optical vibrations in a crystal, they are split into longitudinal (LO) and transverse (TO). Thus, in the Raman spectra, under the condition of the propagation of phonons along the main crystallographic axes, taking into account the LO-TO splitting, 26 lines corresponding to fundamental phonons should appear [2]. There are also A1(z) + E(x, y) acoustic and 5A2 optically inactive fundamental vibrations that should not be detected in the Raman and IR absorption spectra. It follows from the form of the Raman tensors [16] that only nondegenerate phonons of A1(z) symmetry appear in the polarization (zz); in polarizations (xy), (xz), (yx), (yz), (zx), (zy)—only doubly degenerate phonons of E(x, y)—symmetry types. In polarizations (xx) and (yy), phonons A1(z) and E(x, y) of symmetry types must be present simultaneously.
Figure 2 and Figure 3 show the Raman spectra of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal in some backscattering geometries, recorded upon excitation by laser lines with a wavelength of 532 and 785 nm, respectively. The experimentally observed frequencies and their assignments are presented in Table 2. The assignment of experimentally recorded frequencies was carried out on the basis of widely cited articles [36,37], in which crystals of various stoichiometry, powders, and solid solutions were studied using RS and IR spectroscopy at various temperatures. For some registered Raman lines, there is an alternative assignment, shown in parentheses. As this table shows, we did not assign all registered Raman lines. Thus, upon the excitation of Raman radiation via laser radiation in the near-IR range (785 nm), a line with a frequency of 615 cm−1 is observed. This line is close to the calculated value (610 cm−1) for the 4A1(z)TO mode [37]. In one of the first works on Raman spectroscopy in lithium niobate [38], a line with a frequency of 603 cm−1 was observed, which the authors attributed to an A1(z)TO + E(x, y)TO vibration. Other theoretical calculations [39,40,41] refer to the 603 cm−1 line for the fully symmetrical fundamental A1(z)TO-mode. The difference in the frequency of this line can be attributed to the choice of the calculation model, as well as the presence in our work of a lithium niobate crystal co-doped with magnesium and gadolinium. When excited by laser radiation in the visible range (532 nm), a line with a frequency of 732 cm−1 was registered in the Raman spectrum. Taking into account the double doping of the studied crystal and the error of the used spectrometer, it can be attributed to second-order Raman scattering [42].
According to Figure 2, the registered Raman spectra contain intense lines with frequencies: 250, 629–632 cm−1, corresponding to the A1(TO) fundamental vibrations of the symmetry type along the polar z axis; lines with frequencies 864–872 cm−1, corresponding to fundamental vibrations A1(LO) of the symmetry type, and six lines with frequencies of 152, 262, 321–327, 359–362, 429–432, and 578–581 cm−1, corresponding to fundamental doubly degenerate vibrations E(TO)-symmetries perpendicular to the polar z axis. It is important to note that in the spectra of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal in the range 1000–2000 cm−1, upon excitation by a laser line in the visible range (532 nm), we found three broad bands in all the studied scattering geometries. Moreover, one of them has a close spectral position for different geometries.
It is important to note that in the Raman spectra of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) in the range of 900–2000 cm−1, upon excitation by a laser line with a wavelength of 532 nm, we observed a number of lines of very low intensity in the involved scattering geometries (Figure 2). At the same time, when the spectrum is excited by the 785 nm laser line, the intensities of the lines in the range of 900–2000 cm−1 are higher by an order of magnitude or more (Figure 3). Sufficiently intense lines in the range of 900–2000 cm−1, corresponding to the second-order Raman spectrum, were also observed in LiNbO3:Tb(2.24 wt.%), LiTaO3:Cr(0.2): Nd(0.45 wt.%), and LiNbxTa1−xO3 [43,44,45], characterized by a disordered cationic sublattice and distorted (compared to clusters of a congruent crystal) MeO6 clusters, in the region above 1000 cm−1. It should be noted that there are no lines in the recorded Raman spectra whose frequencies are greater than the precise value of the doubled frequency of the 4A1(z)LO mode (872 × 2 = 1744 cm−1).
For the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) double doped crystal, we did not observe a low-intensity line with a frequency of 120 cm−1 corresponding to two-particle states of acoustic phonons with a total wave vector close to zero, which is confidently observed in the scattering geometries involved x ( z z , z y ) x ¯ and y ( z z , z x ) y ¯ in the spectra of all nonstoichiometric LiNbO3 crystals [2,27,29]. The absence of a 120 cm−1 Raman line indicates that the R = [Li]/[Nb] value approaches unity, which is characteristic of a nominally net LiNbO3 crystal of stoichiometric composition [2,27,46]. The appearance of this low-intensity line in the Raman spectrum indicates a strong anharmonic interaction of the lowest frequency fundamental A1(TO)-mode (quasi-soft mode) with the acoustic continuum [46]. In the Raman spectrum of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%), we also did not find any lines corresponding to pseudoscalar vibrations of the A2 symmetry type with frequencies of 209, 230, and 880 cm−1, previously recorded in crystals of LiNbO3 heavily doped with magnesium [30]. The appearance in the Raman spectrum of lines corresponding to pseudoscalar vibrations of the A2 symmetry type is possible due to a decrease in the local point group of the symmetry of the crystal from C3V to C3 due to the distortion of oxygen-octahedral MeO6 clusters by dopant cations [2,47]. The optical representation for the point symmetry group C3 is 9A(z) + 9E(x, y) [2]. Thus, pseudoscalar vibrations of the A2 symmetry type, forbidden in the Raman spectrum for C3V by the selection rules, pass into vibrations of the A1(z) symmetry type permitted by the selection rules for C3. The fact that there are no lines in the Raman spectrum corresponding to A2 vibrations of the symmetry type indicates an insignificant effect of the doping Gd3+ and Mg2+ cations on the geometry of oxygen-octahedral MeO6 clusters in the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal.
Figure 3 shows the Raman spectra of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal obtained in the same backscattering geometries upon excitation by a laser line with a wavelength of 785 nm. From Figure 3, we can see that the Raman spectrum upon excitation by a laser line with a wavelength of 785 nm contains many more lines than are allowed by the selection rules, taking into account the LO-TO splitting for the space group R3c ( C 3 V 6 ), with two formula units in the unit cell [2]. The recorded spectra contain lines corresponding to the fundamental vibrations of the crystal lattice (<900 cm−1), as well as second-order Raman lines (in the range 900–2000 cm−1). Namely, bands with maxima near the frequencies 1036–1046, 1122–1131, 1202, 1279, 1291, 1351, 1398, 1411, 1522–1525, 1595, 1709, 1720, 1853–1862, and 1963 cm−1 correspond to the second-order Raman spectra, the frequencies of which are significantly higher than the frequencies corresponding to the fundamental modes of the crystal lattice, located in the range of 150–900 cm−1. These lines correspond to overtone processes (bound states of optical phonons).
Thus, we have found significant differences in the Raman spectra of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal obtained upon excitation by laser radiation at wavelengths of 532 and 785 nm. We managed to register the Raman spectra of the first and second orders of the crystal LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%). When excited by a near-IR laser line (785 nm), more lines are observed in the recorded second-order Raman spectrum. Differences in the Raman spectra of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) obtained upon excitation by laser lines with wavelengths of 532 and 785 nm can be explained via a different mechanism of interaction, with visible and near-IR radiation of microstructural features and structural defects of the crystal, which determine the phonon–phonon interaction (bound states of phonons) and the anharmonicity of the vibrations of the crystal lattice.
Methods for calculating the bound states of phonons in crystals were proposed earlier in the theoretical works [48,49,50]. In these papers, many-particle states of phonons are associated with the anharmonicity of vibrations of the crystal lattice. Using Green’s function method, one can calculate the density of two-phonon states ρ2(ω) as follows. In the simplest approximation, known as the quasi-Newtonian, the density of single-particle states ρ1(ω) can be written in the following form:
ρ 1 ( ω ) = a ω 0 ω
Here a = ω 0 V 2 ω 0 2 π 2 s 3 , V—unit of cell volume, ω0 is the frequency near the center of the Brillouin zone, and s—speed of sound.
The one-particle Green’s function is written as:
D 1 ( k , ω ) = ω ( k ) 2 [ 1 ω ω ( k ) + 1 2 i Γ 1 ω + ω ( k ) 1 2 i Γ ]
Then, the two-particle Green’s function is written as follows:
D 2 ( k , ω ) = 2 F ( ω ) 1 1 2 g 4 F ( ω )
where g4 is the anharmonicity constant and F(ω) is defined as
F ( ω ) = i ( 2 π ) 4 d 3 k D 1 ( k , ω ω ) D 1 ( k , ω ) d ω
Integrating (4) once, we obtain:
F ( ω ) = 1 4 ω 0 2 a 0 Δ ω ω 2 ( ω 0 ω ) + i Γ d ω
where Δ is a small part of the dispersion curve of optical phonons. Finally, we can obtain the final expression for ρ2(ω):
ρ 2 ( k , ω ) 2 π ω 0 2 Im F ( ω ) [ 1 1 2 g 4 Re F ( ω ) ] 2 + [ 1 2 g 4 Im F ( ω ) ] 2
As can be seen from Figure 3, there are two second-order Raman lines (1853 and 1963 cm−1) of LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%), the frequencies of which are noticeably higher than the exact value overtone frequency of the mode 4A1(z)LO (1752 cm−1). Let us apply the methods for calculating the bound states of the optical phonons to describe these lines. Figure 4 illustrates a comparison between the experimentally observed intensity of Raman scattering, and a theoretical calculation.
As shown in this picture, for the band at 1853 cm−1 (curve 2 in Figure 4), a satisfactory agreement for the bound state of the 4A1(LO) polar mode with ν0 = 876 cm−1 is achieved at s = 1000, Δ = 0.1 ω0, g4 = 4.61·10–18, and Γ = 7.53 1012. For the 1963 cm−1 band (curve 3 in Figure 4), a good agreement is obtained at the values s = 1000, Δ = 0.1 ω0, g4 = 7.54·10–18, and Γ = 7.53·1012. Near the center of the Brillouin zone, the dispersion curve of the 4A1(z)-mode occupies a small area, which explains the choice of the parameter Δ in Equation (5).

4. Conclusions

In this work, the full Raman spectra of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal, which is a promising material for laser radiation conversion, are recorded in the backscattering geometries x ( z z , z y ) x ¯ , y ( z z , z x ) y ¯ , and z ( x x , y y , x y ) z ¯ . We have given an interpretation of the registered Raman lines, including lines with ambiguous assignments. The Raman spectra show that, in the structure of the LiNbO3:Gd3+(0.003):Mg2+ (0.65 wt.%) crystal, oxygen-octahedral clusters of MeO6 (Me–Li, Nb, Gd, and Mg) are slightly distorted, and in addition, the value of R = [Li]/[Nb] is increased compared to that for a congruent crystal. That is, during doping, an increase in the stoichiometry of the crystal occurs. The fact of the increase in stoichiometry is confirmed by the fact that the Raman spectrum of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal lacks a low-intensity line with a frequency of 120 cm−1, corresponding to two-particle states of acoustic phonons with a total wave vector that is close to zero. Its absence in the spectrum also indicates a small anharmonic interaction of the lowest-frequency fundamental vibration A1(TO)-symmetry type (quasi-soft mode) with the acoustic continuum. A low-intensity line with a frequency of 120 cm−1 is confidently observed in the spectra of non-stoichiometric LiNbO3 crystals, nominally pure and doped, and this is absent in the Raman spectra of a stoichiometric crystal [2]. The results obtained indicate a high degree of structural perfection of the crystal and allow us to state that the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal is close in some of its properties to the stoichiometric LiNbO3 crystal. One of the properties of stoichiometric and magnesium-doped LiNbO3 crystals that is important for creating materials for laser radiation conversion on periodically polarized submicron-sized domains with flat boundaries [4,5,6] is a low value of the coercive field (≈2.3 kV/cm). In a congruent LiNbO3 crystal, the coercive field is much higher, ≈23.0 kV/cm.
When the Raman spectrum of a LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal is excited by a laser line with a wavelength of 532 nm, the intensities of the lines corresponding to the second-order spectrum are significantly (by an order of magnitude or more) less than the intensities of the second-order lines observed at excitation with a 785 nm laser line. At the same time, in the Raman spectrum in the region below 900 cm−1, no lines were found to correspond to the second-order spectrum, as well as lines with frequencies of 209, 230, 298, and 880 cm−1 crystals of LiNbO3:Mg [30]. A comparison of the second-order Raman spectra shows that the near-IR laser line makes it possible to detect more second-order Raman lines.

Author Contributions

Conceptualization, N.S. and A.S.; methodology, N.S., M.P. and A.S.; software, A.S.; formal analysis, A.S.; investigation, A.P.; resources, N.S. and M.P.; data curation, A.P.; writing—original draft preparation, A.S., A.P. and N.S.; writing—review and editing, M.P.; visualization, A.P.; supervision, N.S. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by Ministry of Science and Higher Education Russian Federation scientific topic № 0186-2022-0002 (FMEZ-2022-0016) and the RFBR and BRFBR (grant No. 20-52-04001 Bel_mol_a).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data on this research will be available from the corresponding author, A.P., upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Volk, T.; Wöhlecke, M. Lithium niobate. In Defects, Photorefraction and Ferroelectric Switching; Springer: Berlin/Heidelberg, Germany, 2008. [Google Scholar]
  2. Sidorov, N.V.; Volk, T.R.; Mavrin, B.N.; Kalinnikov, V.T. Lithium Niobate: Defects, Photorefraction, Vibrational Spectra; Polaritons: Nauka, Moscow, 2003. [Google Scholar]
  3. Yu, Y.; Chen, X.; Cheng, L.; Dong, Y.; Wu, C.; Li, S.; Fu, Y.; Jin, G. High repetition rate multiple optical parametric oscillator by an aperiodically poled lithium niobate around 1.57 and 3.84 μm. Opt. Laser Technol. 2017, 97, 187–190. [Google Scholar] [CrossRef]
  4. 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] [PubMed]
  5. 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]
  6. 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]
  7. Wang, Z.; Li, B.; Wang, Y.; Han, R.; Yang, Y.; Yu, Y.; Jin, G. Study on Mid-Infrared Energy Conversion of a Doubly Resonant Optical Parametric Oscillator Using Aperiodically Poled Lithium Niobate. Appl. Sci. 2022, 12, 1739. [Google Scholar] [CrossRef]
  8. Zhang, Z.; Liu, H.; Wang, Y.; Wang, X.; Zhao, Y.; Yu, Y.; Jin, G. Theoretical and experimental study on gain competition adjustment of intracavity pumped dual-wavelength optical parametric oscillator using an aperiodically poled lithium niobate at approximately 3.30 and 3.84 μm. Infrared Phys. Technol. 2022, 123, 104167. [Google Scholar] [CrossRef]
  9. Quispe-Siccha, R.; Mejía-Uriarte, E.V.; Villagrán-Muniz, M.; Jaque, D.; Solé, J.G.; Jaque, F.; Sato-Berrú, R.Y.; Camarillo, E.; Hernándes, J.A.; Murrieta, H.S. The effect of Nd and Mg doping on the micro-Raman spectra of LiNbO3 single-crystals. J. Phys. Condens. Matter 2009, 21, 145401. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, J.; Liu, A.; Chen, Y.; Tu, X.; Zheng, Y. Growth and optical properties of Pr-Mg co-doped LiNbO3 crystal using Bridgman method. Phys. B Condens. Matter 2021, 624, 413419. [Google Scholar] [CrossRef]
  11. Dai, L.; Wang, L.; Han, X.; Shao, Y.; Liu, C.; Xu, Y. Defect structure and optical damage resistance of Mg:Ru:Fe:LiNbO3 crystals with various [Li]/[Nb] ratios. J. Alloy. Compd. 2018, 778, 827–832. [Google Scholar] [CrossRef]
  12. Yang, C.; Tu, X.; Wang, S.; Xiong, K.; Chen, Y.; Zheng, Y.; Shi, E. Growth and properties of Pr3+ doped LiNbO3 crystal with Mg2+ incorporation: A potential material for quasi-parametric chirped pulse amplification. Opt. Mater. 2020, 105, 109893. [Google Scholar] [CrossRef]
  13. Long, S.W.; Yang, M.M.; Ma, D.C.; Zhu, Y.Z.; Lin, S.P.; Wang, B. Enhanced red emissions and higher quenching temperature based on the intervalence charge transfer in Pr3+ doped LiNbO3 with Mg2+ incorporation. Opt. Mater. Express 2019, 9, 1062–1071. [Google Scholar] [CrossRef]
  14. Zhao, Y.; Zheng, H.; Yu, Y.; Wang, Y.; Liu, H.; Jin, G. 2.1 μm self frequency conversion optical parameter oscillator based on Nd3+ doped MgO:PPLN. Opt. Laser Technol. 2021, 143, 107348. [Google Scholar] [CrossRef]
  15. Zhang, P.; Yin, J.; Zhang, L.; Liu, Y.; Hong, J.; Ning, K.; Chen, Z.; Wang, X.; Shi, C.; Hang, Y. Efficient enhanced 1.54 μm emission in Er/Yb: LiNbO3 crystal codoped with Mg2+ ions. Opt. Mater. 2014, 36, 1986–1990. [Google Scholar] [CrossRef]
  16. Zhang, P.; Hang, Y.; Yin, J.; Zhao, C.; Gong, J.; He, M.; Zhang, L. Growth and properties of LiNbO3 co-doped with Yb3+/Er3+/Mg2+. J. Cryst. Growth 2012, 363, 118–121. [Google Scholar] [CrossRef]
  17. Sidorov, N.V.; Teplyakova, N.A.; Bobreva, L.A.; Palatnikov, M.N. Optical Properties and Defects of Double Doped Crystals LiNbO3:Mg(5.05):Fe(0.009) and LiNbO3:Zn(4.34):Fe(0.02) (mol%). J. Struct. Chem. 2019, 60, 1765–1772. [Google Scholar] [CrossRef]
  18. Vermeersch, E.; Pincé, P.; Jehlička, J.; Culka, A.; Rousaki, A.; Vandenabeele, P. Micro-Raman spectroscopy on pigments of painted pre-Islamic ceramics from the Kur River Basin (Fars Province, Iran): The case of manganese oxides identification. J. Raman Spectrosc. 2022, 53, 1402–1414. [Google Scholar] [CrossRef]
  19. Rousaki, A.; Costa, M.; Saelens, D.; Lycke, S.; Sánchez, A.; Tuñón, J.; Ceprián, B.; Amate, P.; Montejo, M.; Mirão, J.; et al. A comparative mobile Raman study for the on field analysis of the Mosaico de los Amores of the Cástulo Archaeological Site (Linares, Spain). J. Raman Spectrosc. 2019, 51, 1913–1923. [Google Scholar] [CrossRef]
  20. Conti, C.; Botteon, A.; Bertasa, M.; Colombo, C.; Realini, M.; Sali, D. Portable Sequentially Shifted Excitation Raman spectroscopy as an innovative tool for in situ chemical interrogation of painted surfaces. Analyst 2016, 141, 4599–4607. [Google Scholar] [CrossRef] [PubMed]
  21. Ullah, R.; Khan, S.; Ali, H.; Bilal, M.; Saleem, M.; Mahmood, A.; Ahmed, M. Raman-spectroscopy-based differentiation between cow and buffalo milk. J. Raman Spectrosc. 2017, 48, 692–696. [Google Scholar] [CrossRef]
  22. pour, S.O.; Afshari, R.; Landry, J.; Pillidge, C.; Gill, H.; Blanch, E. Spatially offset Raman spectroscopy: A convenient and rapid tool to distinguish cheese made with milks from different animal species. J. Raman Spectrosc. 2021, 52, 1705–1711. [Google Scholar] [CrossRef]
  23. Pezzotti, G.; Boschetto, F.; Ohgitani, E.; Fujita, Y.; Shin-Ya, M.; Adachi, T.; Yamamoto, T.; Kanamura, N.; Marin, E.; Zhu, W.; et al. Raman Molecular Fingerprints of SARS-CoV-2 British Variant and the Concept of Raman Barcode. Adv. Sci. 2021, 9, 2103287. [Google Scholar] [CrossRef] [PubMed]
  24. Hernández-Arteaga, A.; Ojeda-Galván, H.; Rodríguez-Aranda, M.; Toro-Vázquez, J.; Sánchez, J.; José-Yacamán, M.; Navarro-Contreras, H. Determination of the denaturation temperature of the Spike protein S1 of SARS-CoV-2 (2019 nCoV) by Raman spectroscopy. Spectrochim. Acta A 2022, 264, 120269. [Google Scholar] [CrossRef]
  25. Saleem, M.; Ali, S.; Khan, M.B.; Amin, A.; Bilal, M.; Nawaz, H.; Hassan, M. Optical diagnosis of hepatitis B virus infection in blood plasma using Raman spectroscopy and chemometric techniques. J. Raman Spectrosc. 2020, 51, 1067–1077. [Google Scholar] [CrossRef]
  26. Peñalver, R.; Zapata, F.; Arroyo-Manzanares, N.; López-García, I.; Viñas, P. Raman spectroscopic strategy for the discrimination of recycled polyethylene terephthalate in water bottles. J. Raman Spectrosc. 2022, 54, 107–112. [Google Scholar] [CrossRef]
  27. Sidorov, N.; Palatnikov, M.; Kadetova, A. Raman Scattering in Non-Stoichiometric Lithium Niobate Crystals with a Low Photorefractive Effect. Crystals 2019, 9, 535. [Google Scholar] [CrossRef]
  28. Kruk, A.A.; Sidorov, N.V.; Yanichev, A.A.; Palatnikov, M.N. Raman Spectra of Copper-Doped Lithium Niobate Crystals as a Function of Excitation Wavelength. J. Appl. Spectrosc. 2014, 81, 1–6. [Google Scholar] [CrossRef]
  29. Sidorov, N.; Serebryakov, Y. Investigation of structural peculiarities of impure lithium niobate crystals by Raman spectroscopy. Vib. Spectrosc. 1994, 6, 215–223. [Google Scholar] [CrossRef]
  30. Sidorov, N.V.; Palatnikov, M.N. Raman spectra of lithium niobate crystals heavily doped with zinc and magnesium. Opt. Spectrosc. 2016, 121, 842–850. [Google Scholar] [CrossRef]
  31. Rahman, M.K.R.; Riscob, B.; Bhatt, R.; Bhaumik, I.; Ganesamoorthy, S.; Vijayan, N.; Bhagavannarayana, G.; Karnal, A.K.; Nair, L. Investigations on Crystalline Perfection, Raman Spectra and Optical Characteristics of Transition Metal (Ru) Co-Doped Mg:LiNbO3 Single Crystals. ACS Omega 2021, 6, 10807–10815. [Google Scholar] [CrossRef] [PubMed]
  32. Tejerina, M.; da Silva, K.P.; Goñi, A.; Torchia, G. Hydrostatic-pressure dependence of Raman-active optical phonons in Nd:Mg:LiNbO3. Opt. Mater. 2013, 36, 581–583. [Google Scholar] [CrossRef]
  33. Palatnikov, M.; Biryukova, I.; Sidorov, N.; Denisov, A.; Kalinnikov, V.; Smith, P.; Shur, V. Growth and concentration dependencies of rare-earth doped lithium niobate single crystals. J. Cryst. Growth 2006, 291, 390–397. [Google Scholar] [CrossRef]
  34. Palatnikov, M.; Sidorov, N.; Kadetova, A.; Teplyakova, N.; Makarova, O.; Manukovskaya, D. Concentration threshold in optically nonlinear LiNbO3:Tb crystals. Opt. Laser Technol. 2020, 137, 106821. [Google Scholar] [CrossRef]
  35. Palatnikov, M.N.; Sidorov, N.V.; Biryukova, I.V.; Shcherbina, O.B.; Kalinnikov, V.T. Granulated charge for growth of lithium niobate single crystals. Perspect. Mater. 2011, 2, 93–97. [Google Scholar] [CrossRef]
  36. Margueron, S.; Bartasyte, A.; Glazer, A.M.; Simon, E.; Hlinka, J.; Gregora, I.; Gleize, J. Resolved E-symmetry zone-centre phonons in LiTaO3 and LiNbO3. J. Appl. Phys. 2012, 111, 104105. [Google Scholar] [CrossRef]
  37. Sanna, S.; Neufeld, S.; Rüsing, M.; Berth, G.; Zrenner, A.; Schmidt, W.G. Raman scattering efficiency in LiTaO3 and LiNbO3 crystals. Phys. Rev. B 2015, 91, 224302. [Google Scholar] [CrossRef]
  38. Barker, J.A.S.; Loudon, R. Dielectric Properties and Optical Phonons in LiNbO3. Phys. Rev. 1967, 158, 433–445. [Google Scholar] [CrossRef]
  39. Veithen, M.; Ghosez, P. First-principles study of the dielectric and dynamical properties of lithium niobate. Phys. Rev. B 2002, 65, 214302. [Google Scholar] [CrossRef]
  40. Schmidt, W.G.; Albrecht, M.; Wippermann, S.; Blankenburg, S.; Rauls, E.; Fuchs, F.; Rödl, C.; Furthmüller, J.; Hermann, A. LiNbO3 ground- and excited-state properties from first-principles calculations. Phys. Rev. B 2008, 77, 035106. [Google Scholar] [CrossRef]
  41. Friedrich, M.; Riefer, A.; Sanna, S.; Schmidt, W.G.; Schindlmayr, A. Phonon dispersion and zero-point renormalization of LiNbO3 from density-functional perturbation theory. J. Phys. Condens. Matter 2015, 27, 385402. [Google Scholar] [CrossRef]
  42. Yang, X.; Lan, G.; Li, B.; Wang, H. Raman Spectra and Directional Dispersion in LiNbO3 and LiTaO3. Phys. Status Solidi 1987, 142, 287–300. [Google Scholar] [CrossRef]
  43. Gorelik, V.S.; Abdurakhmonov, S.D. Overtone Raman Scattering in Lithium Niobate Single Crystals Doped with Terbium. Crystallogr. Rep. 2022, 67, 252–255. [Google Scholar] [CrossRef]
  44. Sidorov, N.; Palatnikov, M.; Pyatyshev, A. Raman Scattering in a Double-Doped Single Crystal LiTaO3:Cr(0.2):Nd(0.45 wt%). Photonics 2022, 9, 712. [Google Scholar] [CrossRef]
  45. Sidorov, N.; Palatnikov, M.; Pyatyshev, A.; Sverbil, P. Second-Order Raman Scattering in Ferroelectric Ceramic Solid Solutions LiNbxTa1−xO3. Crystals 2022, 12, 456. [Google Scholar] [CrossRef]
  46. Anikiev, A.A.; Umarov, M.F.; Scott, J.F. Processing and characterization of improved congruent lithium niobate. AIP Adv. 2018, 8, 115016. [Google Scholar] [CrossRef]
  47. Fontana, M.D.; Bourson, P. Microstructure and defects probed by Raman spectroscopy in lithium niobate crystals and devices. Appl. Phys. Rev. 2015, 2, 040602. [Google Scholar] [CrossRef]
  48. Ruvalds, J.; Zawadowski, A. Two-Phonon Resonances and Hybridization of the Resonance with Single-Phonon States. Phys. Rev. B 1970, 2, 1172–1175. [Google Scholar] [CrossRef]
  49. Zawadowski, A.; Ruvalds, J. Indirect Coupling and Antiresonance of Two Optic Phonons. Phys. Rev. Lett. 1970, 24, 1111–1114. [Google Scholar] [CrossRef]
  50. Ruvalds, J.; Zawadowski, A. Resonances of two phonons from different dispersion branches. Solid State Commun. 1971, 9, 129–132. [Google Scholar] [CrossRef]
Figure 2. Raman spectra in the band of 50–2000 cm−1 for the backscattering geometry of a LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal obtained upon excitation by a laser line with a wavelength of 532 nm.
Figure 2. Raman spectra in the band of 50–2000 cm−1 for the backscattering geometry of a LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal obtained upon excitation by a laser line with a wavelength of 532 nm.
Applsci 13 02348 g002aApplsci 13 02348 g002b
Figure 3. Raman spectra in the band of 50–2000 cm−1 for the backscattering geometry of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal obtained upon excitation by a laser line with a wavelength of 785 nm.
Figure 3. Raman spectra in the band of 50–2000 cm−1 for the backscattering geometry of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal obtained upon excitation by a laser line with a wavelength of 785 nm.
Applsci 13 02348 g003aApplsci 13 02348 g003b
Figure 4. Raman scattering spectra of the crystal LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) in the region of two-phonon states: 1—experimental curve, 2 and 3—calculated dependences.
Figure 4. Raman scattering spectra of the crystal LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) in the region of two-phonon states: 1—experimental curve, 2 and 3—calculated dependences.
Applsci 13 02348 g004
Table 1. Concentrations of admixtures (C, wt.%) in the granular charge, as well as in the cone and end parts of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal.
Table 1. Concentrations of admixtures (C, wt.%) in the granular charge, as well as in the cone and end parts of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal.
AdmixtureConcentration C·10−3, wt.%
In ChargeCone of the CrystalEnd of the Crystal
Mn<0.2<0.2<0.2
Ni<0.3<0.3<0.3
Al<0.3<0.31
Fe<0.30.320.38
Cr, Cu, V0.30.30.3
Pb, Sn<0.5<0.5<0.5
Bi0.50.50.5
Mg0.50.530.58
Si, Ti, Mo, Ca, Co111
Sb2.11.72
Zr<101010
Table 2. The frequencies of transverse (TO) and longitudinal (LO) fundamental polar vibrations observed in the Raman spectrum of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal obtained in this work, as well as their assignment using [36,37]. Below are the line frequencies corresponding to the second-order spectrum.
Table 2. The frequencies of transverse (TO) and longitudinal (LO) fundamental polar vibrations observed in the Raman spectrum of the LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) crystal obtained in this work, as well as their assignment using [36,37]. Below are the line frequencies corresponding to the second-order spectrum.
Assignmentλ0 = 532 nmλ0 = 785 nm
ν, cm−1
x ( z z , z y ) x ¯ y ( z z , z x ) y ¯ z ( x x , y y , x y ) z ¯ x ( z z , z y ) x ¯ y ( z z , z x ) y ¯ z ( x x , y y , x y ) z ¯
1E(TO)152152152153153153
1E(LO) 186182188
2E(LO) 238237237237
1A1(TO)250250 253256
3E(TO) 262 262
3E(LO) 296
4E(TO)321321327322322328
5E(TO)
(5E(LO))
359362 367365363
6E(LO)
(7E(TO))
429432432436433429
8E(TO)581581578578
4A1(TO)629632 629629
4A1(LO)
(9E(LO))
875864872876876875
104610361041
1122 1131
1202
1281 129112791291
1319 13511398
1411
1475
154815121539152315251522
1595
172517341739170917201720
1853 1862
1963
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

Sidorov, N.; Palatnikov, M.; Pyatyshev, A.; Skrabatun, A. Investigation of the Structural Perfection of a LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) Double-Doped Single Crystal Using the Raman Spectra Excited by Laser Lines in the Visible (532 nm) and Near-IR (785 nm) Regions. Appl. Sci. 2023, 13, 2348. https://doi.org/10.3390/app13042348

AMA Style

Sidorov N, Palatnikov M, Pyatyshev A, Skrabatun A. Investigation of the Structural Perfection of a LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) Double-Doped Single Crystal Using the Raman Spectra Excited by Laser Lines in the Visible (532 nm) and Near-IR (785 nm) Regions. Applied Sciences. 2023; 13(4):2348. https://doi.org/10.3390/app13042348

Chicago/Turabian Style

Sidorov, Nikolay, Mikhail Palatnikov, Alexander Pyatyshev, and Alexander Skrabatun. 2023. "Investigation of the Structural Perfection of a LiNbO3:Gd3+(0.003):Mg2+(0.65 wt.%) Double-Doped Single Crystal Using the Raman Spectra Excited by Laser Lines in the Visible (532 nm) and Near-IR (785 nm) Regions" Applied Sciences 13, no. 4: 2348. https://doi.org/10.3390/app13042348

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