Next Article in Journal
Evaluation of Smeared Constitutive Laws for Tensile Concrete to Predict the Cracking of RC Beams under Torsion with Smeared Truss Model
Next Article in Special Issue
Optimization of Residual Stress Measurement Conditions for a 2D Method Using X-ray Diffraction and Its Application for Stainless Steel Treated by Laser Cavitation Peening
Previous Article in Journal
Effect of Mg on the Formation of Periodic Layered Structure during Double Batch Hot Dip Process in Zn-Al Bath
Previous Article in Special Issue
The Number of Subgrain Boundaries in the Airfoils of Heat-Treated Single-Crystalline Turbine Blades
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Relationship between Young’s Modulus and Planar Density of Unit Cell, Super Cells (2 × 2 × 2), Symmetry Cells of Perovskite (CaTiO3) Lattice

1
Faculty of Mechanical Engineering and Design, Kaunas University of Technology, LT-51424 Kaunas, Lithuania
2
Department of Materials Science and Engineering, Sharif University of Technology, Tehran 11365-9466, Iran
*
Authors to whom correspondence should be addressed.
Submission received: 5 January 2021 / Revised: 1 March 2021 / Accepted: 2 March 2021 / Published: 6 March 2021
(This article belongs to the Special Issue X-ray Diffraction of Functional Materials)

Abstract

:
Calcium titanate-CaTiO3 (perovskite) has been used in various industrial applications due to its dopant/doping mechanisms. Manipulation of defective grain boundaries in the structure of perovskite is essential to maximize mechanical properties and stability; therefore, the structure of perovskite has attracted attention, because without fully understanding the perovskite structure and diffracted planes, dopant/doping mechanisms cannot be understood. In this study, the areas and locations of atoms and diffracted planes were designed and investigated. In this research, the relationship between Young’s modulus and planar density of unit cell, super cells (2 × 2 × 2) and symmetry cells of nano CaTiO3 is investigated. Elastic constant, elastic compliance and Young’s modulus value were recorded with the ultrasonic pulse-echo technique. The results were C11 = 330.89 GPa, C12 = 93.03 GPa, C44 = 94.91 GPa and E = 153.87 GPa respectively. Young’s modulus values of CaTiO3 extracted by planar density were calculated 162.62 GPa, 151.71 GPa and 152.21 GPa for unit cell, super cells (2 × 2 × 2) and symmetry cells, respectively. Young’s modulus value extracted by planar density of symmetry cells was in good agreement with Young’s modulus value measured via ultrasonic pulse-echo.

Graphical Abstract

1. Introduction

Perovskites have a general formula of ABO3. In these structures, the A site cation is a typical lanthanide, alkaline or alkaline-earth metal with 12-fold oxygen coordination, and the B-site is any one of a variety of transition metal cations [1]. Calcium titanate (CaTiO3) was established in 1839 by a Russian mineralogist Perovski, and materials with the same type of CaTiO3 were introduced as the perovskite structure. CaTiO3 has ionic bonds, as well as the ionic radii of Ca2+, O2− and Ti4+ are 1   Å , 1.40   Å and 0.6   Å , respectively [2]. In recent years, researchers have focused on developing perovskites and their mechanical properties in order to obtain a high yield. Furthermore, CaTiO3 is a well-known component in ferroelectric perovskite category, which has been considerably utilized as a dopant/doping in electronic materials due to its dielectric manner and flexibility in structural transformations [3,4]. The modulus of elasticity (E) or Young’s modulus is defined as the proportion of the stress to the strain, created by the stress on the body when the body is in the elastic region [5]. The elastic constants are specified from the lattice crystal deformation against force. These elastic moduli are: Young’s modulus, shear modulus and volumetric modulus. These modules are registered via inherent elastic properties of materials and their resistance to deformation due to loading. Elastic behavior of materials is described by models such as Cauchy elastic, hypo-elastic and hyper-elastic. A hyper-elastic is a constitutive model for ideally elastic material that responds against stress gain from a strain energy density function, while for hypo-elastic material, their governing equation is independent of finite strain quantity except in the linearized state [6]. The elastic properties are intimately connected to the crystal structure, the intrinsic character of bonding between the atoms and the anisotropic nature of materials [7,8]; therefore, elastic constants can be derived from crystal lattice calculations [9]. There are several studies on the relationship between elastic constants and planes/directions in a lattice structure, for example, in [10,11,12]. One of the most accurate methods to measure the elastic stiffness constants and Young’s modulus is to determine the velocity of long-wavelength acoustic waves through the ultrasonic pulse-echo technique [13]. In a crystal structure, points, directions and planes are described with an indexing scheme, and planar density is obtained as the number of atoms per unit area, which are centered on a specific crystallographic plane with a defined index [14]. Since the discovery of X-rays at the end of the 19th century, this method has been often used for material characterization [15]. It is used to identify the atomic-scale structure of different materials in a variety of states [16]. X-ray diffraction is the only method that provides the specification of both the mechanical and microstructural character of each diffracted plane. These planes are used as a strain to quantify Young’s modulus in one or more planes/directions of the diffraction vector [17]. In forming, designing and manufacturing equipment industries, the use of non-destructive, accurate and convenient methods to determine the mechanical properties of materials is particularly important. Mechanical tests, such as tensile, strike and collision tests, are destructive. Ultrasonic methods are very time-consuming and require operator expertise in this area, and theoretical methods require time-consuming density functional theory (DFT) calculation and may need verification with experimental tests. Our proposed method only needs the XRD analysis, which is a routine test and calculation of planar density; therefore, it can be very significant in terms of industrial application. In this study, the effects of cell size on the accuracy of Young’s modulus calculation were considered. Locations of atoms and diffracted planes of unit cell, super cells (2 × 2 × 2) and symmetry cells of CaTiO3 are designed and investigated. The super cell is a cell that describes the same crystal but has a larger volume than a unit cell. By extension of a unit cell proportional to the lattice vectors, the super cells are generated. In super cells (2 × 2 × 2), the extension is twice of unit cell length in each direction; likewise, for super cells (8 × 8 × 8), the extension is 8 times. The result extracted by symmetry cells was in good agreement with results recorded via ultrasonic technique. Therefore, this new approach of exploration of reliable Young’s modulus quantity based on XRD is proposed for either single crystal or polycrystalline of CaTiO3.

2. Experimental

2.1. Materials

In this study, for synthesis CaTiO3, titanium (IV) butoxide, calcium chloride dehydrate, sodium hydroxide and ethanol reagents were purchased from Sigma Aldrich (Taufkirchen, Germany) and deionized water as the solvent for dispersions was prepared.

2.2. Instrumentation

In this research, a Bruker D8 Advance X-ray diffractometer (Kaunas, Lithuania) with Cu K α radiation was used. The powder X-ray diffraction was taken at 40 kV and 40 mA, and it was registered at a scanning rate of 2.5 degrees/minute and a step size of 0.02 degrees. The XRD peaks were interpreted by High Score X’Pert software (4.9.0) analysis to get the output ASC type files. The pulse-echo technique was applied for the determination of sound velocity for both transverse and longitudinal ultrasonic signals. For ultrasonic measurement, the model of pulser receiver and oscilloscope were Panametrics Co. (waltham, MA, USA) and Iwatsu (Tokyo, Japan) (100 MHz), respectively. For powder pressing, the model of mechanical machines was CD04-Z and CIP (CP 360). Additionally, the specific surface area of the sample was investigated by desorption isotherms of nitrogen (N2) gas via using a Brunauer-Emmett-Teller (BET) apparatus Gemini V analyzer, micrometrics GmbH (Tehran, Iran). Moreover, transmission electron microscopy (TEM) CM 10-Philips (Tehran, Iran) with acceleration voltage from 50 to 80 KV was utilized.

2.3. Methods

2.3.1. Synthesis of Nano-Powder CaTiO3

Calcium titanate (CaTiO3) was synthesized by solvothermal method. A simple procedure, namely the solvothermal method, was performed for the synthesis of CaTiO3 (Figure S1). In the first step, (1) calcium chloride dehydrate was stirred with ethanol and deionized water. (2) Titanium (IV) butoxide and ethanol were added to the system drop by drop, under stirring for around 10 min (750 rpm). The molar ratio of ingredients was achieved to calcium chloride dehydrate = 1, ethanol = 5, Titanium (IV) butoxide = 1 and deionized water = 100 respectively. (3) To create pH = 14, sodium hydroxide solution was utilized. (4) The produced solution was placed into the autoclave and the temperature was ~250 °C for 5 h. (5) Afterward, the product was under the drying conditions involved at 110 °C and 0.76 bar, respectively. (6) After a day, the mixture was washed, (7) filtered and dried (110 °C for 4 h), respectively. This method was used in previous studies [18,19].

2.3.2. X-ray Diffraction of CaTiO3 and Planar Density Calculations

Combining X-ray diffraction of crystalline CaTiO3 and calculation of planar density values of each diffracted plane was performed. In our study, the atomic density of each plane was considered as the planar density, which was determined as the area of atoms with the center positioned at the plane divided by the total area of the plane, and it is a determinant factor for mechanical properties of each plane. Planar density is a unitless parameter, and its value is less than 1 in each cell. Furthermore, the values of planar density are related to the positions and situations of atoms in the planes. For determination of atomic area, the Crystal Maker, Version 10.2.2 software was performed. First of all, the three-dimensional (3D) geometry of crystal structures was designed, and then, from the intersection area of each diffracted plane with atoms located at the plane, the atomic area was calculated. When an atom with diameter D was involved completely, the atomic area will be   A = π ( D 2 ) 2 ; otherwise, it will be a percentage of this amount.

2.3.3. Ultrasonic Pulse-Echo Technique of CaTiO3

An ultrasonic wave is a type of elastic wave spread in the medium with high frequency to obtain the Young’s modulus value of samples. Mastering the ultrasonic parameters can be used to acquire more accurate values of mechanical properties [20]. Recently, different studies on mechanical properties have been done by ultrasonic techniques. Basically, the crossing of longitudinal and transverse waves in nano- or microstructures is performed at different velocities. Each returned velocity is considered as the represented properties.
For ultrasonic measurements based on the Christoffel procedure, the first cubic specimen of CaTiO3 was prepared by cold isostatic press. The schematics of ultrasonic measurement are depicted in Figure 1a. The main part of the ultrasonic system is the pulser-receiver, which creates an electric pulse and stimulates the probe. Furthermore, the produced pulses enter the specimen, and after a sweep, they can be received via a probe. In this measurement, some drops of water were utilized to prevent the depreciation of waves in the air, and the effect of hand pressure on the probe was decreased [21].
At any position in the sample, a local coordinate is adjusted, such as X1, the radial coordinate; X2, the circumferential coordinate; and X3, the axial coordinate. Vi/j denotes the velocity of an ultrasound wave propagating in the Xi direction with particle displacements in the Xj direction. Vi/j with the same i and j is longitudinal, and with i ≠ j is related to the transverse waves. For the measurement of quasi-longitudinal or quasi-transverse velocity (Vij/ij), the specimen should be cut (bezel) on the edges of the surfaces perpendicular to the X directions. A sketch of the sample is represented in Figure 1b.

3. Results

3.1. X-ray Diffraction of CaTiO3 and Planar Density Calculations

The XRD pattern of CaTiO3 is presented in Figure 2. The characteristic peaks of CaTiO3 correspond to the report in Ref [22]. The crystal structure of CaTiO3 is cubic, the atomic positions of Ti are at (000), Ca at ( 1 2 , 1 2 , 1 2 ) and O at ( 1 2 , 0 , 0 ), ( 0 , 1 2 , 0 ), ( 0 , 0 , 1 2 ). According to X-ray powder diffraction results, the lattice parameter is 3.79 ± 0.02 Å, which is in good corresponds with the amount recorded in the Ref [23]. In addition, crystallographic parameters (Table S1) of CaTiO3 and analyzed data by X’Pert [24] nasiri are recorded as the cell volume = 54.44 ±   0.01   Å 3 and crystal density = 4.14 ± 0.01 g/cm3, and the space group is Pm-3m. In addition, the crystal size of CaTiO3 was calculated by the Monshi–Scherrer equation (Figure S2) [25] and BET analysis. The crystal size values were registered at ~59.10 and 63.02 nm, respectively. The Monshi–Scherrer method is described in Section 2 of the supporting information. Furthermore, a TEM image of CaTiO3 is shown in Figure S3. According to the images shown in Figure S3, the size of CaTiO3 particles basically corresponds to the crystallite size, and it is clear that particles of powder have nanoscale and size can be reported almost ±50 nm.
For the evaluation of cells as the results, the comprehensive calculations of the planar density of diffracted planes in the unit cell, super cells (2 × 2 × 2) and super cells (8 × 8 × 8) of CaTiO3 lattice are presented in Figures S4–S6 respectively. In addition, the locations of atoms, geometry of planes and calculations of planar density of (211) super cell (4 × 4 × 4), (211) super cell (8 × 8 × 8), (221) super cell (4 × 4 × 4), (221) super cell (8 × 8 × 8), (311) super cell (3 × 3 × 3), (311) super cell (4 × 4 × 4), (311) super cell (8 × 8 × 8), (222) super cell (3 × 3 × 3) and (222) super cell (8 × 8 × 8) are depicted briefly in Figure 3, Figure 4, Figure 5 and Figure 6 respectively. Furthermore, the completed calculations with their figures are shown in Figures S7–S10.

3.2. Investigation of Results Obtained from Ultrasonic Pulse-Echo Technique of CaTiO3

Taking into account the Christoffel equation, the connection between ultrasonic phase velocity and the stiffness matrix is given as follows:
( C ijkl l j l l ρ V 2 δ ik ) α k   =   0
where V is the ultrasonic phase velocity, C ijkl is the general stiffness matrix, ρ is the material density, l is the orientation of propagation, α k is the polarization direction and δ ik is the Kronecker delta (note that i, j, k, I = 1 to 3). For the extraction and calculation of elastic constants from ultrasonic measurements based on the Christoffel equation, with the propagation in X1, X2 and X3 directions, all of the diagonal elements of the stiffness matrix are obtained. For the determination of whole constants, we cut the specimen on the edges of the surfaces perpendicular to principal directions (bezel) and the velocity was measured from the propagation of ultrasound wave normal to these planes.
Based on Equations (1)–(5) [26,27] and the measured velocity according to the Table 1, stiffness constants values were obtained. C 11 is in the agreement with longitudinal distortion and longitudinal compression/tension, so C 11 can be described as the hardness. Moreover, the transverse distortion is connected to the C 12 , and C 12 is obtained from the transverse expansion correlated to the Poisson’s ratio. C 44 is based on the shear modulus, as well as C 44 is in the settlement with C 11 and C 12   [26].
C 11   =   ρ V 1 1 2
C 22   =   ρ V 2 2 2
C 66   =   ρ V 1 2 2 =   ρ V 2 1 2
C 12   =   ( C 11 + C 66 2 ρ V 12 12 2 ) ( C 22 + C 66 2 ρ V 12 12 2 )     C 66
C 44   =   ρ V 2 3 2 =   ρ V 3 2 2
After substitution and calculation, C11, C12 and C44 were registered at 330.89, 93.03 and 94.91 GPa respectively. These values of CaTiO3 were in good agreement with the values submitted in the [28,29,30]. Moreover, with the ultrasonic technique, longitudinal and transverse waves can be utilized for determining Young’s modulus quantity [31,32]. The longitudinal and transverse waves of CaTiO3 sample are shown in Figure 7. In this method, by measuring the waves velocity and density of specimen, the determination of Young’s modulus quantity was carried out (Equation (6)).
E   =   4 ρ ( L t s ) 2 ( 3 t s 2 4 t l 2 ) t s 2 t l 2
where, t s and t l are differences between two echo in longitudinal and transverse waves, respectively [33,34]. According to the results shown in Figure 7, t s and t l values are calculated as 5.75 and 3.01 μs, respectively. In addition, the density of the specimen is recorded as 3857.30 Kg m 3 , and the length of the specimen after powder pressing reached 11.21 mm. After calculation, Young’s modulus value of CaTiO3 was 153.87 GPa. This value corresponds with the value reported by Ramajo et al. [35].

3.3. Calculations: Relationship between Elastic Stiffness-Compliance Constants, Young’s Modulus and Planar Density Extracted through the Unit Cell, Super Cells (2 × 2 × 2) and Symmetry Cells of CaTiO3 Lattice

Three elastic constants of CaTiO3 were calculated via the ultrasonic technique. For the cubic CaTiO3 system, the relationship between stiffness ( C ij ) and compliance constant ( S ij ) are provided in Equations (7)–(9) [27,36]. The values resulted via Equations (7)–(9) are 0.0034, −0.0007 and 0.0105 GPa for S 11 , S 12 and S 44 , respectively. Furthermore, Young’s modulus of each diffracted plane of CaTiO3 can be written as Equation (10) [37].
S 11 =   C 11 + C 12 ( C 11 C 12 ) ( C 11 + 2 C 12 )
S 12 =   C 12 ( C 11 C 12 ) ( C 11 + 2 C 12 )
S 44 =   1 C 44
1 E h k l   =   S 11   2 [ ( S 11 S 12 ) 1 2 S 44 ] [ h 2 k 2 +   k 2 l 2 +   l 2 h 2 ( h 2 +   k 2 +   l 2 ) ]
The planar density and Young’s modulus values related to the each diffracted plane of the unit, super (2 × 2 × 2), symmetry and super (8 × 8 × 8) cells of CaTiO3 lattice are tabulated in Table 2.

4. Discussion

According to Table 2 and Figure 3, Figure 4, Figure 5 and Figure 6, the expanded cells have an optimum matrix, and in this case, achieving the optimum matrix is introduced as the symmetry cells. An optimum matrix is the minimum extension for a specific plane of the unit cell to a super cell from which the density plane of that plane does not change. For example, symmetry cell (optimum matrix) of (311) plane is (3 × 3 × 3), which means that after extending to a greater matrix such as (4 × 4 × 4) or (8 × 8 × 8), planar density values will be similar (Figure 5a–c). Real planar density values of each plane are obtained from the symmetry cell, because once the symmetry of each plane is reached, with the extension of that plane to infinity (real plane), the planar density does not change. In addition, to recognize the symmetry cell, knowing some parameters such as crystal lattice, locations of atoms in the planes and index of planes is essential. Therefore, to determine Young’s modulus values based on the planar density of CaTiO3, the symmetry cells should be found. It is very interesting that symmetrical or real values are related to the super cells of the (8 × 8 × 8) matrix, because in matrix (8 × 8 × 8), lattice correspondence in all directions is available; therefore, real planar density values should be calculated for the super cell of (8 × 8 × 8) matrix. To confirm this, calculations of real planar density and geometry of atoms and planes of (211), (221), (311) and (222) in super cells (8 × 8 × 8) are presented in Figure 3b, Figure 4b, Figure 5c and Figure 6b, respectively. It is clear that finding the exact situation of planes and geometries is sophisticated, but with when they are known, the results obtained from Young’s modulus values will have fewer errors. The basic supposition is that when the planar density is raised, the motion of atoms with the mechanism of dislocation movement needs high forces. Dislocations are regions in the lattice where an additional plane of atoms have been included abstracted from an ideal crystal (without imperfections). Dislocations are caused by losing acoustic energy, and this matter will affect the values of wavelength and time of ultrasonic waves [38].
The force (W), which is needed for the movement of atoms in each plane, is obtained from Equation (11) [39].
W   =   E 2 ( 1 + ν )   b 2 l
In Equation (11), E is Young’s modulus, b is Burgers vector, l is dislocation length and ν is Poisson’s ratio. The higher value of force is in accordance with the modulus of elasticity (Young’s modulus), which would be higher.
To compare Young’s modulus values of CaTiO3 in a unit cell, super cells (2 × 2 × 2) and symmetry cells, the fitting of Young’s modulus values extracted by each diffracted plane versus planar density values is presented in Figure 8. According to the results (shown in the Figure 8) and the straight fitting line, Young’s modulus values of unit cell, super cells (2 × 2 × 2) and symmetry cells were calculated as 162.62 ± 0.4 GPa, 151.71 ± 0.4 GPa and 152.21 ± 0.4 GPa, respectively. As expected, the Young’s modulus value of symmetry cells of CaTiO3 (152.21 ± 0.4 GPa) is in good agreement with experimental Young’s modulus value extracted via ultrasonic-echo technique (153.87 ± 0.2 GPa). Moreover, Young’s modulus value of unit cell (162.62 ± 0.4 GPa) has a greater difference with experimental Young’s modulus value, and as a result, the unit cell of CaTiO3 cannot be represented as whole cells. This is because in a unit cell of CaTiO3, crystalline defects are not considered and is especially controlling of deformation, and displacement of atoms in the planes is related to the dislocation networks [40]. Further, a unit cell of CaTiO3 is not involved in imperfections (such as dislocations, Frenkel defect and Schottky defect) with respect to the super cell [41]; therefore, the slope line value of the unit cell is reported (37.23) to be less than the slope line value of super cells (2 × 2 × 2) (63.67) and symmetry cells (62.41). Consequently, the effect of imperfections in expanded cells (super cells) is very impressive, so the unit cell of CaTiO3 is considered as the ideal lattice, while symmetry cells such as (8 × 8 × 8) of CaTiO3 are real lattices [42]; this is consistent with the experimental Young’s modulus. It is clear that each imperfection will be caused by a decreasing Young’s modulus [43], and in Figure 8, this matter is confirmed when the Young’s modulus value (intercept) in the unit cell of CaTiO3 is higher than in super cells (2 × 2 × 2) and symmetry cells. Apparently, a unit cell of CaTiO3 is represented by the volume of a real crystal, so the unit cell is useful to acquire theoretical density. Nevertheless, calculations of planar density based on the unit cell were obtained, but with errors.

5. Conclusions

  • CaTiO3 as a category of perovskite is successfully synthesized via solvothermal method.
  • Crystal size values of CaTiO3 are calculated as ~59.10 and 63.02 through the Monshi-Scherrer method and BET analysis, and the crystal size values were confirmed by TEM image.
  • Planar density is responsible for modulus of elasticity of that plane; therefore, for the first time, comprehensive calculations of geometry, location and planar density values of CaTiO3 were shown.
  • Elastic stiffness constants and Young’s modulus values of CaTiO3 were obtained by ultrasonic-echo method (C11 = 330.89, C12 = 93.03, C44 = 94.91 GPa and E = 153.87 ± 0.2 GPa).
  • Young’s modulus values of CaTiO3 extracted by planar density and least square method were calculated as 162.62 ± 0.4, 151.71 ± 0.4 and 152.21 ± 0.4 GPa for unit cell, super cells (2 × 2 × 2) and symmetry cells, respectively.
  • The Young’s modulus value of CaTiO3 reported by symmetry cells is in good agreement with Young’s modulus value reported by ultrasonic-echo technique and the literatre.
  • A unit cell of CaTiO3 is not representative of the distribution of atoms on the planes; therefore, to obtain the real value of planar density and find the symmetry of distribution of atoms on the planes, expanded cells and utilizing symmetry cells are suggested.
  • Obtaining the planar density values based on unit cell or each super cells except for (8 × 8 × 8) is an estimation.
  • The real value of Young’s modulus of CaTiO3 should be extracted by symmetry cells or super cells (8 × 8 × 8).
  • The value of Young’s modulus of CaTiO3 extracted with this method can be applied for industrial applications.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/1996-1944/14/5/1258/s1, Figure S1: Synthesis route of CaTiO3, Table S1: Crystallographic parameters of each individual XRD pattern related to the CaTiO3, Figure S2: Linear plot of modified Scherrer equation related to the CaTiO3, Figure S3: TEM image of CaTiO3 powder, Figure S4: Geometry of planes and calculations of planar density of (a) (100), (b) (110), (c) (111), (d) (200), (e) (210), (f) (211), (g) (220), (h) (221), (i) (310), (j) (311) and (k) (222) related to the unit cell of CaTiO3, Figure S5: Geometry of planes and calculations of planar density of (a) (100), (b) (110), (c) (111), (d) (200), (e) (210), (f) (211), (g) (220), (h) (221), (i) (310), (j) (311) and (k) (222) related to the super cells (2 × 2 × 2) of CaTiO3, Figure S6: Geometry of planes and calculations of planar density of (a) (100), (b) (110), (c) (111), (d) (200), (e) (210), (f) (220), (g) (310) (4 × 4 × 4) and (h) (310) (8 × 8 × 8) related to the super cells (8 × 8 × 8) of CaTiO3, Figure S7: Geometry of planes and calculations of planar density of (a) (211) super cell (4 × 4 × 4) and (b) (211) super cell (8 × 8 × 8), Figure S8: Geometry of planes and calculations of planar density of (a) (221) super cell (4 × 4 × 4) and (b) (221) super cell (8 × 8 × 8), Figure S9: Geometry of planes and calculations of planar density of (a) (311) super cell (3 × 3 × 3), (b) (311) super cell (4 × 4 × 4) and (c) (311) super cell (8 × 8 × 8), Figure S10: Geometry of planes and calculations of planar density of (a) (222) super cell (3 × 3 × 3), (b) (222) super cell (8 × 8 × 8).

Author Contributions

Conceptualization, M.R. and A.P.; methodology, M.R. and S.N.; investigation, A.V. and A.D.; data curation, A.D.; writing—original draft, M.R.; writing—review and editing, S.N. and G.J.; resources, G.J.; supervision and validation A.P. and G.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a grant No. S-MIP-19-43 from the Research Council of Lithuania.

Data Availability Statement

Data supporting the findings of this study are available from the corresponding author upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Atta, N.F.; Galal, A.; El-Ads, E.H. Perovskite Nanomaterials-Synthesis, Characterization, and Applications; Intechopen: London, UK, 2016. [Google Scholar]
  2. Reshmi Varma, P.C. Low-dimensional perovskites. In Perovskite Photovoltaics: Basic to Advanced Concepts and Implementation; Elsevier: Amsterdam, The Netherlands, 2018; pp. 197–229. ISBN 9780128129159. [Google Scholar]
  3. Nakamura, T.; Sun, P.H.; Shan, Y.J.; Inaguma, Y.; Itoh, M.; Kim, I.N.S.; Sohn, J.H.O.; Ikeda, M.; Kitamura, T.; Konagaya, H. On the perovskite-related materials of high dielectric permittivity with small temperature dependence and low dielectric loss. Ferroelectrics 1997, 196, 205–209. [Google Scholar] [CrossRef]
  4. Wang, X.; Xu, C.N.; Yamada, H.; Nishikubo, K.; Zheng, X.G. Electro-mechano-optical conversions in Pr3+-doped BaTiO3-CaTiO3 ceramics. Adv. Mater. 2005, 17, 1254–1258. [Google Scholar] [CrossRef]
  5. Ma, Y.Z.; Sobernheim, D.; Garzon, J.R. Glossary for Unconventional Oil and Gas Resource Evaluation and Development. In Unconventional Oil and Gas Resources Handbook: Evaluation and Development; Elsevier Inc.: Amsterdam, The Netherlands, 2016; pp. 513–526. ISBN 9780128025369. [Google Scholar]
  6. Slaughter, W.S. The Linearized Theory of Elasticity; Birkhäuser Boston: New York, NY, USA, 2002. [Google Scholar]
  7. Kuma, S.; Woldemariam, M.M. Structural, Electronic, Lattice Dynamic, and Elastic Properties of SnTiO3 and PbTiO3 Using Density Functional Theory. Adv. Condens. Matter Phys. 2019, 2019. [Google Scholar] [CrossRef] [Green Version]
  8. Nakamura, M. Elastic constants of some transition- metal- disilicide single crystals. Metall. Mater. Trans. A 1994, 25, 331–340. [Google Scholar] [CrossRef]
  9. Paszkiewicz, T.; Wolski, S. Elastic properties of cubic crystals: Every’s versus Blackman’s diagram. J. Phys. Conf. Ser. 2008, 104, 012038. [Google Scholar] [CrossRef]
  10. Ching, W.Y.; Rulis, P.; Misra, A. Ab initio elastic properties and tensile strength of crystalline hydroxyapatite. Acta Biomater. 2009, 5, 3067–3075. [Google Scholar] [CrossRef] [Green Version]
  11. Holec, D.; Friák, M.; Neugebauer, J.; Mayrhofer, P.H. Trends in the elastic response of binary early transition metal nitrides. Phys. Rev. B-Condens. Matter Mater. Phys. 2012, 85, 64101. [Google Scholar] [CrossRef] [Green Version]
  12. Knowles, K.M.; Howie, P.R. The Directional Dependence of Elastic Stiffness and Compliance Shear Coefficients and Shear Moduli in Cubic Materials. J. Elast. 2015, 120, 87–108. [Google Scholar] [CrossRef] [Green Version]
  13. Stair, K.; Liu, J.; Asta, M. Undefined Ultra-sonic measurement and computation of elastic constants. In Proceedings of the ASEE Annual Conference and Exposition, Kansas City, MO, USA, 13–15 September 2006. [Google Scholar]
  14. Kelly, A.; Knowles, K.M. Crystallography and Crystal Defects; John Wiley & Sons, Ltd.: Chichester, UK, 2012; ISBN 9781119961468. [Google Scholar]
  15. Epp, J. X-ray Diffraction (XRD) Techniques for Materials Characterization. In Materials Characterization Using Nondestructive Evaluation (NDE) Methods; Elsevier Inc.: Amsterdam, The Netherlands, 2016; pp. 81–124. ISBN 9780081000571. [Google Scholar]
  16. Badawi, F.; Villain, P. Stress and elastic-constant analysis by X-ray diffraction in thin films. J. Appl. Crystallogr. 2003, 36, 869–879. [Google Scholar] [CrossRef]
  17. Waseda, Y.; Matsubara, E.; Shinoda, K. X-ray Diffraction Crystallography X-ray Diffraction Crystallography Introduction, Examples and Solved Problems; Springer: Berlin/Heidelberg, Germany, 2011. [Google Scholar]
  18. Dong, W.; Song, B.; Meng, W.; Zhao, G.; Han, G. A simple solvothermal process to synthesize CaTiO3 microspheres and its photocatalytic properties. Appl. Surf. Sci. 2015, 349, 272–278. [Google Scholar] [CrossRef]
  19. Wang, Y.; Niu, C.G.; Wang, L.; Wang, Y.; Zhang, X.G.; Zeng, G.M. Synthesis of fern-like Ag/AgCl/CaTiO3 plasmonic photocatalysts and their enhanced visible-light photocatalytic properties. RSC Adv. 2016, 6, 47873–47882. [Google Scholar] [CrossRef]
  20. ASTM E797—95 Standard Practice for Measuring Thickness by Manual Ultrasonic Pulse-Echo Contact Method; ASTM International: West Conshohocken, PA, USA, 2001.
  21. ASTM E214—05 Standard Practice for Immersed Ultrasonic Testing by the Reflection Method Using Pulsed Longitudinal Waves; ASTM International: West Conshohocken, PA, USA, 2007.
  22. Sahoo, S.; Parashar, S.K.S.; Ali, S.M. CaTiO3 nano ceramic for NTCR thermistor based sensor application. J. Adv. Ceram. 2014, 3, 117–124. [Google Scholar] [CrossRef] [Green Version]
  23. Cockayne, E.; Burton, B.P. Phonons and static dielectric constant in CaTiO3 from first principles. Phys. Rev. B-Condens. Matter Mater. Phys. 2000, 62, 3735–3743. [Google Scholar] [CrossRef] [Green Version]
  24. Janusas, T.; Urbaite, S.; Palevicius, A.; Nasiri, S.; Janusas, G. Biologically Compatible Lead-Free Piezoelectric Composite for Acoustophoresis Based Particle Manipulation Techniques. Sensors 2021, 21, 483. [Google Scholar] [CrossRef] [PubMed]
  25. Rabiei, M.; Palevicius, A.; Monshi, A.; Nasiri, S.; Vilkauskas, A.; Janusas, G. Comparing Methods for Calculating Nano Crystal Size of Natural Hydroxyapatite Using X-ray Diffraction. Nanomaterials 2020, 10, 1627. [Google Scholar] [CrossRef]
  26. Li, Y.; Thompson, R.B. Relations between elastic constants Cij and texture parameters for hexagonal materials. J. Appl. Phys. 1990, 67, 2663–2665. [Google Scholar] [CrossRef] [Green Version]
  27. Huntington, H.B. The Elastic Constants of Crystals. Solid State Phys.-Adv. Res. Appl. 1958, 7, 213–351. [Google Scholar]
  28. De Jong, M.; Chen, W.; Angsten, T.; Jain, A.; Notestine, R.; Gamst, A.; Sluiter, M.; Ande, C.K.; Van Der Zwaag, S.; Plata, J.J.; et al. Charting the complete elastic properties of inorganic crystalline compounds. Sci. Data 2015, 2, 150009. [Google Scholar] [CrossRef] [Green Version]
  29. Tariq, S.; Ahmed, A.; Saad, S.; Tariq, S. Structural, electronic and elastic properties of the cubic CaTiO3 under pressure: A DFT study. AIP Adv. 2015, 5, 077111. [Google Scholar] [CrossRef]
  30. Sakhya, A.P.; Maibam, J.; Saha, S.; Chanda, S.; Dutta, A.; Sharma, B.I.; Thapa, R.; Sinha, T. Electronic structure and elastic properties of ATiO3 (A = Ba, Sr, Ca) perovskites: A first principles study. IJPAP 2015, 53, 102–109. [Google Scholar]
  31. Wang, H.; Prendiville, P.L.; McDonnell, P.J.; Chang, W.V. An ultrasonic technique for the measurement of the elastic moduli of human cornea. J. Biomech. 1996, 29, 1633–1636. [Google Scholar] [CrossRef]
  32. Bray, D.E.; Stanley, R.K. Nondestructive Evaluation: A Tool in Design, Manufacturing and Service; CRC press: Boca Raton, FL, USA, 1996. [Google Scholar]
  33. Figliola, R.S.; Beasley, D.E. Theory and Design for Mechanical Measurements, 7th ed.; John Wiley & Sons: Hoboken, NJ, USA, 2019. [Google Scholar]
  34. Rabiei, M.; Palevicius, A.; Dashti, A.; Nasiri, S.; Monshi, A.; Vilkauskas, A.; Janusas, G. Measurement Modulus of Elasticity Related to the Atomic Density of Planes in Unit Cell of Crystal Lattices. Materials 2020, 13, 4380. [Google Scholar] [CrossRef]
  35. Ramírez, M.A.; Parra, R.; Reboredo, M.M.; Varela, J.A.; Castro, M.S.; Ramajo, L. Elastic modulus and hardness of CaTiO3, CaCu3Ti4O12 and CaTiO3/CaCu3Ti4O12 mixture. Mater. Lett. 2010, 64, 1226–1228. [Google Scholar] [CrossRef]
  36. Papaconstantopoulos, D.A.; Mehl, M.J. Tight-Binding Method in Electronic Structure. In Encyclopedia of Condensed Matter Physics; Elsevier Inc.: Amsterdam, The Netherlands, 2005; pp. 194–206. ISBN 9780123694010. [Google Scholar]
  37. Zhang, J.M.; Zhang, Y.; Xu, K.W.; Ji, V. Young’s modulus surface and Poisson’s ratio curve for cubic metals. J. Phys. Chem. Solids 2007, 68, 503–510. [Google Scholar] [CrossRef]
  38. Mason, W.P. Effect of Dislocations on Ultrasonic Wave Attenuation in Metals. Bell Syst. Tech. J. 1955, 34, 903–942. [Google Scholar] [CrossRef]
  39. Reed-Hill, R.E.; Abbaschian, R. PHYSICAL METALLURGY PRINCIPLES; Boston PWS Publishing Company: Boston, MA, USA, 2009; ISBN 978-0-495-08254-5. [Google Scholar]
  40. Berdichevsky, V. Energy of dislocation networks. Int. J. Eng. Sci. 2016, 103, 35–44. [Google Scholar] [CrossRef]
  41. Hull, D.; Bacon, D.J. Introduction to Dislocations, 5th ed.; Butterworth-Heinemann: Burlington, MA, USA, 2011; ISBN 9780080966724. [Google Scholar]
  42. Englert, U. Symmetry Relationships between Crystal Structures. Applications of Crystallographic Group Theory in Crystal Chemistry. By Ulrich Müller. Angew. Chem. Int. Ed. 2013, 52, 11973. [Google Scholar] [CrossRef]
  43. Duparc, O.H.; Polytechnique, É.; Duparc, O.B.M.H. A review of some elements in the history of grain boundaries, centered on Georges Friedel, the coincident “site” lattice and the twin index Crystallography History View project ANR FluTi View project A review of some elements in the history of grain boundaries, centered on Georges Friedel, the coincident “site” lattice and the twin index. Artic. J. Mater. Sci. 2011, 46, 4116–4134. [Google Scholar]
Figure 1. Schematic of (a) ultrasonic pulse instrument and (b) a sketch of prepared CaTiO3 sample.
Figure 1. Schematic of (a) ultrasonic pulse instrument and (b) a sketch of prepared CaTiO3 sample.
Materials 14 01258 g001
Figure 2. X-ray diffraction of CaTiO3 (powder sample).
Figure 2. X-ray diffraction of CaTiO3 (powder sample).
Materials 14 01258 g002
Figure 3. Geometry of planes and calculations of planar density of (a) (211) super cell (4 × 4 × 4) and (b) (211) super cell (8 × 8 × 8) (which shows and emphasizes the symmetry of (8 × 8 × 8) super cells).
Figure 3. Geometry of planes and calculations of planar density of (a) (211) super cell (4 × 4 × 4) and (b) (211) super cell (8 × 8 × 8) (which shows and emphasizes the symmetry of (8 × 8 × 8) super cells).
Materials 14 01258 g003aMaterials 14 01258 g003b
Figure 4. Geometry of planes and calculations of planar density of (a) (221) super cell (4 × 4 × 4) and (b) (221) super cell (8 × 8 × 8) (which shows and emphasizes the symmetry of (8 × 8 × 8) super cells).
Figure 4. Geometry of planes and calculations of planar density of (a) (221) super cell (4 × 4 × 4) and (b) (221) super cell (8 × 8 × 8) (which shows and emphasizes the symmetry of (8 × 8 × 8) super cells).
Materials 14 01258 g004aMaterials 14 01258 g004b
Figure 5. The concept of a symmetry cell; geometry of planes and calculations of planar density of (a) (311) super cell (3 × 3 × 3), (b) (311) super cell (4 × 4 × 4) and (c) (311) super cell (8 × 8 × 8).
Figure 5. The concept of a symmetry cell; geometry of planes and calculations of planar density of (a) (311) super cell (3 × 3 × 3), (b) (311) super cell (4 × 4 × 4) and (c) (311) super cell (8 × 8 × 8).
Materials 14 01258 g005aMaterials 14 01258 g005b
Figure 6. Geometry of planes and calculations of planar density of (a) (222) super cell (3 × 3 × 3) and (b) (222) super cell (8 × 8 × 8).
Figure 6. Geometry of planes and calculations of planar density of (a) (222) super cell (3 × 3 × 3) and (b) (222) super cell (8 × 8 × 8).
Materials 14 01258 g006
Figure 7. Recorded signals extracted via (a) longitudinal waves and (b) transverse waves of CaTiO3 specimen.
Figure 7. Recorded signals extracted via (a) longitudinal waves and (b) transverse waves of CaTiO3 specimen.
Materials 14 01258 g007
Figure 8. Young’s modulus versus planar density values of each diffracted plane related to the (a) symmetry cells, (b) super cells (2 × 2 × 2) and (c) unit cell of CaTiO3.
Figure 8. Young’s modulus versus planar density values of each diffracted plane related to the (a) symmetry cells, (b) super cells (2 × 2 × 2) and (c) unit cell of CaTiO3.
Materials 14 01258 g008
Table 1. The values of longitudinal and transverse velocity of the sample.
Table 1. The values of longitudinal and transverse velocity of the sample.
Longitudinal Velocity
(m/s)
Transverse Velocity
(m/s)
Quasi Longitudinal or Quasi Transverse (m/s)
V1/1 = 9261.85V2/3 = 4960.5V12/12 = 4976.63
V2/2 = 8013.51V1/2 = 4283.65
Table 2. Planar density and Young’s modulus values of the unit cell, super cells (2 × 2 × 2) and symmetry cells of CaTiO3.
Table 2. Planar density and Young’s modulus values of the unit cell, super cells (2 × 2 × 2) and symmetry cells of CaTiO3.
IndexPlanar Density
of Unit Cell
Planar Density
of Super Cell (2 × 2 × 2)
Planar Density
of Symmetry Cells
Planar Density
of Super Cell (8 × 8 × 8)
Young’s Modulus
(GPa)
(100)0.930.930.93 in (2 × 2 × 2)0.93290.059
(110)0.510.510.51 in (2 × 2 × 2)0.51221.652
(111)0.040.040.04 in (2 × 2 × 2)0.04179.354
(200)0.640.640.64 in (2 × 2 × 2)0.64290.059
(210)0.410.410.41 in (2 × 2 × 2)0.41194.176
(211)0.160.250.25 in (2 × 2 × 2)0.25150.612
(220)0.60.60.6 in (2 × 2 × 2)0.6129.810
(221)0.460.290.31 in (4 × 4 × 4)0.31109.622
(310)0.240.240.23 in (4 × 4 × 4)0.23186.471
(311)0.040.030.02 in (3 × 3 × 3)0.02140.386
(222)0.990.880.88 in (3 × 3 × 3)0.8883.615
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rabiei, M.; Palevicius, A.; Nasiri, S.; Dashti, A.; Vilkauskas, A.; Janusas, G. Relationship between Young’s Modulus and Planar Density of Unit Cell, Super Cells (2 × 2 × 2), Symmetry Cells of Perovskite (CaTiO3) Lattice. Materials 2021, 14, 1258. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14051258

AMA Style

Rabiei M, Palevicius A, Nasiri S, Dashti A, Vilkauskas A, Janusas G. Relationship between Young’s Modulus and Planar Density of Unit Cell, Super Cells (2 × 2 × 2), Symmetry Cells of Perovskite (CaTiO3) Lattice. Materials. 2021; 14(5):1258. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14051258

Chicago/Turabian Style

Rabiei, Marzieh, Arvydas Palevicius, Sohrab Nasiri, Amir Dashti, Andrius Vilkauskas, and Giedrius Janusas. 2021. "Relationship between Young’s Modulus and Planar Density of Unit Cell, Super Cells (2 × 2 × 2), Symmetry Cells of Perovskite (CaTiO3) Lattice" Materials 14, no. 5: 1258. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14051258

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