Next Article in Journal
AgSn[Bi1−xSbx]Se3: Synthesis, Structural Characterization, and Electrical Behavior
Previous Article in Journal
Microstructure Evaluation Study of Al5083 Alloy Using EBSD Technique after Processing with Different ECAP Processes and Temperatures
Previous Article in Special Issue
Investigation of Elastic Properties of the Single-Crystal Nickel-Base Superalloy CMSX-4 in the Temperature Interval between Room Temperature and 1300 °C
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Extreme Values of Young’s Modulus and the Negative Poisson’s Ratios of Rhombic Crystals

by
Valentin A. Gorodtsov
and
Dmitry S. Lisovenko
*
Ishlinsky Institute for Problems in Mechanics RAS, Prosp. Vernadskogo 101-1, 119526 Moscow, Russia
*
Author to whom correspondence should be addressed.
Crystals 2021, 11(8), 863; https://doi.org/10.3390/cryst11080863
Submission received: 4 May 2021 / Revised: 14 July 2021 / Accepted: 22 July 2021 / Published: 25 July 2021
(This article belongs to the Special Issue Elasticity of Crystalline Materials)

Abstract

:
The extreme values of Young’s modulus for rhombic (orthorhombic) crystals using the necessary and sufficient conditions for the extremum of the function of two variables are analyzed herein. Seven stationary expressions of Young’s modulus are obtained. For three stationary values of Young’s modulus, simple analytical dependences included in the sufficient conditions for the extremum of the function of two variables are revealed. The numerical values of the stationary and extreme values of Young’s modulus for all rhombic crystals with experimental data on elastic constants from the well-known Landolt-Börnstein reference book are calculated. For three stationary values of Young’s modulus of rhombic crystals, a classification scheme based on two dimensionless parameters is presented. Rhombic crystals ((CH 3 ) 3 NCH 2 COO·(CH) 2 (COOH) 2 , I, SC(NH 2 ) 2 , (CH 3 ) 3 NCH 2 COO·H 3 BO 3 , Cu-14 wt%Al, 3.0wt%Ni, NH 4 B 5 O 8 · 4H 2 O, NH 4 HC 2 O 4 · 1/2H 2 O, C 6 N 2 O 3 H 6 and CaSO 4 ) having a large difference between maximum and minimum Young’s modulus values were revealed. The highest Young’s modulus among the rhombic crystals was found to be 478 GPa for a BeAl 2 O 4 crystal. More rigid materials were revealed among tetragonal (PdPb 2 ; maximum Young’s modulus, 684 GPa), hexagonal (graphite; maximum Young’s modulus, 1020 GPa) and cubic (diamond; maximum Young’s modulus, 1207 GPa) crystals. The analytical stationary values of Young’s modulus for tetragonal, hexagonal and cubic crystals are presented as special cases of stationary values for rhombic crystals. It was found that rhombic, tetragonal and cubic crystals that have large differences between their maximum and minimum values of Young’s modulus often have negative minimum values of Poisson’s ratio (auxetics). We use the abbreviated term auxetics instead of partial auxetics, since only the latter were found. No similar relationship between a negative Poisson’s ratio and a large difference between the maximum and minimum values of Young’s modulus was found for hexagonal crystals.

1. Introduction

Anisotropic materials occupy an important place in modern technical applications. While the description of the linear elastic properties of isotropic media requires only two independent elastic constants, the number of important elastic constants increases with decreasing symmetry of materials. The deformation of anisotropic crystalline bodies depends not only on the locations of external forces in relation to the body, but also on the orientation of the crystallographic axes inside it. In addition, restrictions on such important elastic engineering characteristics (combinations of elastic constants), such as Young’s moduli, Poisson’s ratios and shear moduli, are reduced. In particular, if Poisson’s ratios in isotropic media have restrictions of 1 below and 0.5 above, then for crystals of all seven symmetry systems, including the most symmetric cubic system, there are no general restrictions on the values and signs of Poisson’s ratios [1].
An analysis of the variability of Poisson’s ratios and Young’s moduli of a large number of real crystals of all seven crystal systems (cubic, hexagonal, rhombohedral, tetragonal, rhombic, monoclinic and triclinic) was carried out in [2,3], based on extensive information on experimental elastic constants in the Landolt–Börnstein reference book [4]. In [2], the extrema of Poisson’s ratios and correlations of the extrema with the values of elastic anisotropy indices, generalizing the classical Zener exponent, were found. The extrema of Poisson’s ratios, together with the extrema of Young’s moduli, were also established for real crystals of all crystalline systems in [3], limited to a one-parameter set of orientations. General analytical results for the extrema of the basic engineering moduli of materials of any crystal symmetry were obtained in [5,6,7]. In [6], the stationary values and extrema of Young’s modulus and shear modulus were analyzed. In [7], conditions for the stationary values, maxima and minima of the three engineering moduli of anisotropic elastic materials were derived.
Several studies have been devoted to the analysis of the extreme values of Young’s modulus and Poisson’s ratio for crystals of some particular symmetry systems and examples of real crystals. In [8], a variational Lagrangian analysis of the extrema of Young’s modulus for cubic and hexagonal crystals was supplemented with examples of classifications and results for some crystals. In [9], general expressions for the extrema of Young’s moduli of six constant tetragonal crystals were established, and results were given for many materials. In [10,11], analytical expressions for the extreme values of Poisson’s ratio for cubic crystals were obtained. These analytical relationships were used to calculate the extreme values of known crystals. It has been demonstrated that high absolute values of the extrema of Poisson’s ratio can be observed for specific orientations of some crystals. In [10], indium–thallium alloys were such crystals. In [11], most of their attention was paid to metastable cubic metal alloys and analyzing of the role of the elastic anisotropy coefficient, which vanishes in the limit of an isotropic medium. Stationary and extreme values of Young’s moduli and Poisson’s ratios for hexagonal crystals were established in [12] based on an analytical analysis of the angular orientations of crystals and several dimensionless anisotropy characteristics that disappear in the isotropic limit. Numerical results were obtained on the basis of 147 hexagonal crystals. The anisotropy coefficients made it possible to construct classification schemes for the distributions of the extrema of Young’s modulus and Poisson’s ratio of real crystals.
The history of materials with negative Poisson’s ratio dates back to the publication on crystalline pyrite in the well-known monograph by A.E.H. Love [13]. Experimental research and qualitative analysis by R.S. Lakes of negative Poisson’s ratios for metal and polymer foams [14,15] had a great influence on further studies of various materials and designs. The proposal by K.E. Evans of replacing the longer phrase “negative-Poisson’s-ratio materials” with the term auxetics [16,17] has become generally accepted.
The first theoretical studies of auxetics by K.W. Wojciechowsci [18,19] dealt with a 2D isotropic lattice built from 2D anisotropic molecules. In [20], Tretiakov K.V. and Wojciechowski K.W. studied the features of the formation of auxetics in the isotropic 2D solid phase depending on the 2D molecular geometry. In [21], the same authors analyzed the formation of auxetics, partially auxetics and nonauxetics among 2D crystals of five crystal systems with anisotropic 2D molecules in the form of rigid cyclic tetramers. In the article by K.W. Wojciechowski and A.C. Branka [22], approximations of free volume and Monte Carlo simulation revealed the decisive role of the hexagonal shape of the molecule in the 2D isotropic lattice model, which leads to auxeticity due to mirror symmetry breaking (chirality). However, in [23], K.W. Wojciechowski demonstrated that a 2D isotropic model with 2D anisotropic molecules such as cyclic trimers can form a nonchiral phase with a negative Poisson’s ratio.
A new series of studies by K.W. Wojciechowski, K.V. Tretiakov, J.W. Narojczyk, P.M. Piglowski and their collaborators has concerned the auxeticity of 3D model materials with 2D thin layers and 1D narrow channels (“nanolayers” and “nanochannels”) [24,25,26,27,28,29,30,31,32]. The auxetic properties of the composites of spherical particles in some main matrix and nanochannels [24,25,28,29,30] or nanolayers [26,27,31] depended on their orientations, relative particle diameters and filling densities. It was shown in [32] that the orthogonal combination of nanochannels and nanolayers can lead, with a sufficiently large size of spherical inclusions, to the absence of auxetic properties of the composite.
Another line of research into the mechanism of auxeticity was undertaken by J.N. Grima, K.E. Evans and A. Alderson et al. [33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58]. The concepts of the mechanism of auxeticity of materials in these articles were based on rotations of simplified 2D geometric structures from triangular, square, rectangular and rhombic forms (etc.). Following A. Alderson and K.E. Evans, the auxetic nature of the deformation of a number of crystals (zeolites, silicates, α -crystobalite and β -crystobalite in particular) was associated with rotation and dilation of 3D tetrahedral and rotating 3D cuboidal microstructures [59,60,61,62,63,64,65,66,67,68]. It was shown in [69] that the auxeticity and negative linear compressibility of Boron Arsenate arise mainly due to deformations of framework tetrahedra. In [70], the manifestation of auxeticity and negative linear compressibility was discussed in the case of the formation of a 3D microstructure of a metamaterial due to stretching in out-of-plane direction of the original 2D “rotating squares”. In [71], the possibility of auxeticity for a broad range of loading directions and negative linear compressibility for a small number of such directions was discussed for a 3D metamaterial composed of arrowhead-like structural units. In [72], the role of the rearrangement of the 3D microstructure of boron arsenanite under shear deformation in auxeticity and negative linear compressibility was discussed. An important feature of the shearing deformation of tetrahedra on the projection planes is the distortion of the rotating squares.
Auxetic materials are often found among natural anisotropic materials. There are particularly many of them (about three hundred) among highly symmetric cubic crystals [73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103]. Since the negativity of Poisson’s ratio usually corresponds to the selected directions of crystal orientation [7,11], in this case we actually focus on partial auxetics. Fewer auxetics are found among crystals of lower symmetry.
In this article, we consider the problem of stationary and extreme values of Young’s modulus, and the question of the relationship between the extrema of Young’s modulus and the value and sign of Poisson’s ratio. Section 2 begins with a presentation of Young’s modulus versus crystal orientation angles. Then six anisotropy coefficients are introduced as linear combinations of the compliance triples that disappear in the isotropic limit. Anisotropy coefficients for 18 crystals are shown in Table 1. A more complete list is provided in the Supplementary Material. In Section 3, the analysis of the second derivatives made it possible to find the extrema of Young’s modulus for 140 rhombic crystals, shown partially in Table 2 and completely in the Supplementary Material. The dependence of three stationary values of Young’s modulus on the anisotropy coefficients is presented in the form of a classification scheme. An analysis of the extrema of Poisson’s ratios showed that more than 50 rhombic crystals are auxetic; about 30 of them correspond to the ratio E max / E min > 3 (Table 3 and Supplementary Material). In Section 4, the stationary values of Young’s modulus for cubic, hexagonal and tetragonal crystals are discussed briefly as special cases of the rhombic system. In Section 5, conclusions are given.

2. Young’s Modulus

This expression for the reciprocal of Young’s modulus is obtained as the ratio of the tensile force uniformly distributed over the transverse surface to the relative elongation using Hooke’s law for an anisotropic material. Young’s modulus E n for anisotropic materials depends on the tensor compliance coefficients s i j k l and direction of the axis of extension [104]:
1 E n = s i j k l n i n j n k n l .
Here n i are the components of the unit vector n , which is directed along the axis of extension. Rhombic crystals are characterized by nine independent matrix compliance coefficients s 11 , s 22 , s 33 , s 44 , s 55 , s 66 , s 12 , s 13 and s 23 [105]. The matrix of compliance coefficients is represented as follows
s 11 s 12 s 13 0 0 0 s 12 s 22 s 23 0 0 0 s 13 s 23 s 33 0 0 0 0 0 0 s 44 0 0 0 0 0 0 s 55 0 0 0 0 0 0 s 66 .
Using the matrix compliance coefficients, the expression for Young’s modulus of rhombic crystals can be written as
E 1 ( n ) = s 11 n 1 4 + s 22 n 2 4 + s 33 n 3 4 + ( 2 s 23 + s 44 ) n 2 2 n 3 2 + ( 2 s 13 + s 55 ) n 1 2 n 3 2 + ( 2 s 12 + s 66 ) n 1 2 n 2 2 .
If the orientation of the crystalline rod in the crystallographic coordinate system is described with three Euler’s angles φ , θ , ψ , then using the relationship between the unit vector n and Euler’s angles φ , θ ,
n = sin φ sin θ cos φ sin θ cos θ ,
the expression of Young’s modulus E for rhombic crystals can be rewritten as follows.
E 1 ( φ , θ ) = s 11 sin 4 φ + s 22 cos 4 φ + 0.25 ( 2 s 12 + s 66 ) sin 2 2 φ sin 4 θ + + s 33 cos 4 θ + 0.25 ( 2 s 23 + s 44 ) cos 2 φ + ( 2 s 13 + s 55 ) sin 2 φ sin 2 2 θ
The dependence of Young’s modulus for rhombic crystals is a periodic function φ and θ with periods T φ = π and T θ = π .
It is convenient to introduce six anisotropy coefficients of rhombic crystals for analyzing the variability of Young’s modulus:
Δ 1 s 11 s 12 0.5 s 66 , Δ 2 s 22 s 12 0.5 s 66 , Δ 3 s 11 s 13 0.5 s 55 , Δ 4 s 33 s 13 0.5 s 55 , Δ 5 s 22 s 23 0.5 s 44 , Δ 6 s 33 s 23 0.5 s 44 ,
which disappear in the limit of an isotropic medium. The number of anisotropy coefficients of rhombic crystals is greater than those of the cubic, hexagonal and tetragonal crystals. The last crystals have one [105,106], two [8,12] and three [9] anisotropy coefficients, respectively. The values of the anisotropy coefficients for some rhombic crystals are given in Table 1, and in Table S1 from the Supplementary Material the values for all rhombic crystals from the reference book [4] are presented.

3. Stationary and Extreme Values of Young’s Modulus

The necessary conditions for extremum of Young’s modulus are the stationarity conditions
E ( φ , θ ) φ = 0 , E ( φ , θ ) θ = 0 .
These conditions using (2) lead to a system of equations:
Δ 1 + Δ 2 sin 2 φ Δ 2 sin 2 θ + Δ 0 cos 2 θ sin 2 θ sin 2 φ = 0 s 11 sin 4 φ + s 22 cos 4 φ + 0.25 2 s 12 + s 66 sin 2 2 φ s 13 + 0.5 s 55 sin 2 φ s 23 + 0.5 s 44 cos 2 φ sin 2 θ + + s 13 + 0.5 s 55 sin 2 φ + s 23 + 0.5 s 44 cos 2 φ s 33 cos 2 θ sin 2 θ = 0
Here Δ 0 = Δ 6 Δ 4 = Δ 5 Δ 3 = s 13 s 23 + 0.5 s 55 0.5 s 44 . Four solutions to the first equation are θ = 0 ; φ = 0 ; φ = π / 2 ; tan 2 θ = Δ 0 / Δ 1 + Δ 2 sin 2 φ Δ 2 . By substituting them into the second equation of system (5), one can find seven stationary values of Young’s modulus.
At φ = π / 2 and θ = π / 2 stationary value
E 1 = E [ 100 ] = 1 s 11
is achieved. It corresponds to stretching in the [100] direction.
The second stationary value of Young’s modulus
E 2 = E [ 010 ] = 1 s 22 = 1 s 11 + Δ 2 Δ 1
is achieved at φ = 0 , θ = π / 2 and φ = π , θ = π / 2 . It corresponds to stretching in the [ 0 1 ¯ 0 ] and [010] directions.
The third value also has a simple form,
E 3 = E [ 001 ] = 1 s 33 = 1 s 11 + Δ 4 Δ 3 ,
and is achieved at θ = 0 and an arbitrary angle φ . This stationary value corresponds to stretching in the [001] direction.
At φ = 0 the fourth stationary value of Young’s modulus has the form
E 4 = E 2 E 3 Δ 5 + Δ 6 2 E 2 Δ 5 2 + E 3 Δ 6 2 + 2 E 3 1 Δ 5 E 2 Δ 5 Δ 6
at the limitations
tan 2 θ = Δ 6 Δ 5 0 .
This value corresponds to stretching in the (100) plane. Young’s moduli E 2 and E 3 also lie in the (100) plane.
At φ = π / 2 the fifth stationary value of Young’s modulus has the form
E 5 = E 1 E 3 Δ 3 + Δ 4 2 E 1 Δ 3 2 + E 3 Δ 4 2 + 2 E 3 1 Δ 3 E 1 Δ 3 Δ 4
with the limitations
tan 2 θ = Δ 4 Δ 3 0 .
This value corresponds to stretching in the (010) plane. Young’s moduli E 1 and E 3 also lie in the (010) plane.
At θ = π / 2 the sixth stationary value of Young’s modulus has the form
E 6 = E 1 E 2 Δ 1 + Δ 2 2 E 1 Δ 1 2 + E 2 Δ 2 2 + 2 E 2 1 Δ 1 E 1 Δ 1 Δ 2
with the limitation
tan 2 φ = Δ 2 Δ 1 0 .
This value corresponds to stretching in the (001) plane. Young’s moduli E 1 and E 2 also lie in the (001) plane.
The seventh stationary value of Young’s modulus has the form (2) with the constraints
tan 2 θ = Δ 0 Δ 1 + Δ 2 sin 2 φ Δ 2 0 ,
0 sin 2 φ = Δ 0 Δ 5 Δ 2 Δ 6 Δ 0 2 + Δ 0 Δ 2 Δ 6 Δ 1 + Δ 2 1 .
We further investigate these stationary points using the sufficient condition for the extremum of the function of two variables. If at the indicated stationary points from the second derivatives of Young’s modulus,
A = 2 E φ 2 , B = 2 E φ θ , C = 2 E θ 2 ,
a combination is formed
D = A C B 2 ,
then at D > 0 extremes of Young’s modulus are achieved at the corresponding stationary point (maximum at A < 0 and C < 0 or minimum at A > 0 and C > 0 ). In the case D < 0 , extrema are absent at the stationary point, and at D = 0 additional analysis is required [107].
In the case of a stationary point φ = π / 2 , θ = π / 2 , we have E = E 1 and
D = 16 Δ 1 Δ 2 s 11 4 , A = 4 Δ 1 s 11 2 , C = 4 Δ 2 s 11 2 , B = 0 .
Then, according to the sufficient condition for the extremum of the function, the value of Young’s modulus E 1 will be extremal if Δ 1 Δ 3 > 0 . The value E 1 will be the maximum at Δ 1 < 0 or Δ 3 < 0 and the minimum at Δ 1 > 0 or Δ 3 > 0 .
In the case of a stationary points φ = 0 , θ = π / 2 and φ = π , θ = π / 2 , we have E = E 2 and
D = 16 Δ 2 Δ 5 s 22 4 , A = 4 Δ 2 s 22 2 , C = 4 Δ 5 s 22 2 , B = 0 .
The value of Young’s modulus E 2 will be extreme if Δ 2 Δ 5 > 0 . The value E 2 is the maximum at Δ 2 < 0 or Δ 5 < 0 and minimum at Δ 2 > 0 or Δ 5 > 0 .
At θ = 0 and an arbitrary angle φ we have E = E 3 and combination of coefficients D vanishes. As a result, additional analysis is required for each specific crystal. The value E 3 will be the extremum of 44 (from 142) rhombic crystals. For example, such crystals are Ga ( E 3 = E max ), NH 4 ClO 4 ( E 3 = E max ), Al 2 SiO 5 ( E 3 = E max ), BaSO 4 ( E 3 = E max ), Cs 2 SO 4 ( E 3 = E min ) and LiCsSO 4 ( E 3 = E min ), AgTlSe ( E 3 = E min ), TbF 3 ( E 3 = E min ) (see Table 2 and Table S2 in the Supplementary Material).
For the stationary values of Young’s modulus— E 4 , E 5 , E 6 and E 7 —the second derivatives A, B, C and D have a cumbersome analytical form. Therefore, only numerical analysis of them for 142 rhombic crystals was carried out. The results of this analysis are presented in Table 2 and Table S2 in the Supplementary Material. In Table S2 in the Supplementary Material for the values of Young’s modulus E 4 , E 5 , E 6 and E 7 , the values of the angles at which they are achieved are also given. In these tables, the global maximum and minimum values of Young’s modulus are shown in bold. An analysis of the variability of Young’s modulus showed that the value E 7 is the inflection point for all rhombic crystals from [4].
The largest differences between the maximum and minimum values of Young’s modulus were found in (CH 3 ) 3 NCH 2 COO·(CH) 2 (COOH) 2 ( E max / E min = 12.7 ), I ( E max / E min = 10.1 ) and SC(NH 2 ) 2 ( E max / E min = 9.96 ; for the second set of elastic constants E max / E min = 11.8 ), (CH 3 ) 3 NCH 2 COO·H 3 BO 3 ( E max / E min = 9.92 ), Cu-14 wt% Al, 3.0 wt% Ni ( E max / E min = 7.51 ), NH 4 B 5 O 8 · 4H 2 O ( E max / E min = 7.39 ), NH 4 HC 2 O 4 · 1/2H 2 O ( E max / E min = 5.83 ), C 6 N 2 O 3 H 6 ( E max / E min = 5.58 ), CaSO 4 ( E max / E min = 5.4 ). The maximum Young’s modulus was revealed in BeAl 2 O 4 ( E max = 478 GPa). Thus, among rhombic crystals, no materials were found for which E max > 500 GPa, in contrast to materials with tetragonal, hexagonal and cubic anisotropy (see Section 4).
Among rhombic crystals, more than 50 auxetics (materials with negative Poisson’s ratios) were detected. The smallest values of Poisson’s ratio were for C 6 N 2 O 3 H 6 ( ν min = 0.91 ), Cu-14wt %Al, 3.0wt %Ni ( ν min = 0.70 ), I ( ν min = 0.48 ), (CH 3 NHCH 2 COOH) 3 · CaCl 2 ( ν min = 0.48 ), AgTlSe ( ν min = 0.42 ), (CH 3 ) 3 NCH 2 COO·H 3 BO 3 ( ν min = 0.39 ), Sr(COOH) 2 · 2H 2 O ( ν min = 0.39 ), SC(NH 2 ) 2 ( ν min = 0.37 ). As can be seen, most materials that have a maximum value of ratio E max / E min have the smallest values of Poisson’s ratio. The data from Table 3 and Table S3 from the Supplementary Material confirm this. Thirty-three rhombic crystals with E max / E min > 3 are shown in Table 3. Twenty-four crystals of them have negative Poisson’s ratios. The remaining ten crystals from this table have small positive values for the minimum Poisson’s ratio. These values range from 0 to 0.1. The values E max / E min and extremum values of Poisson’s ratio for all rhombic crystals from [4] are given in Table S3 from the Supplementary Material.
In Figure 1, the classification scheme for three stationary values of Young’s modulus E 1 , E 2 and E 3 depending on two dimensionless parameters is presented, α = ( Δ 2 Δ 1 ) / s 11 and β = ( Δ 4 Δ 3 ) / s 11 . The points indicate the values of dimensionless parameters α and β for 142 rhombic crystals from [4]. Most crystals fall into the area 1 < α < 1 and 1 < β < 1 . There are six zones on the classification scheme, in which various inequalities between the stationary values of Young’s modulus E 1 , E 2 , E 3 are satisfied. For each of these zones, the surface of Young’s moduli for some rhombic crystals are shown in Figure 2.

4. Young’s Moduli of Tetragonal, Hexagonal and Cubic Crystals

Above, the stationary values of Young’s modulus for rhombic crystals were shown. Below we present the stationary values of Young’s modulus for tetragonal, hexagonal and cubic crystals as special cases of rhombic crystals. Rhombic crystals are characterized by nine independent compliance coefficients s 11 , s 22 , s 33 , s 44 , s 55 , s 66 , s 12 , s 13 and s 23 , and six anisotropy coefficients (see Formulas (3)).

4.1. Tetragonal Crystals

Tetragonal crystals have six independent compliance coefficients, which are obtained under three conditions, s 11 = s 22 , s 44 = s 55 and s 13 = s 23 , for nine compliance coefficients that were given previously.
The expression of Young’s modulus for six-constant tetragonal crystals takes the form
E 1 ( φ , θ ) = s 11 0.5 Δ 1 sin 2 2 φ sin 4 θ + s 33 cos 4 θ + 0.25 ( 2 s 13 + s 44 ) sin 2 2 θ .
The dependence of Young’s modulus for six-constant tetragonal crystals is a periodic function φ , θ with periods T φ = π / 2 and T θ = π . Such crystals will already have three anisotropy coefficients,
Δ 1 s 11 s 12 0.5 s 66 , Δ 2 s 11 s 12 0.5 s 44 , Δ 3 s 33 s 13 0.5 s 44 ,
and five stationary values of Young’s modulus.
1. At φ = π / 2 , θ = π / 2 , φ = 0 , θ = π / 2 the first stationary value has the form
E 1 = E [ 100 ] = E [ 010 ] = 1 s 11
and is achieved by stretching in the [100] and [010] directions.
2. At θ = 0 and an arbitrary angle φ second stationary value
E 2 = E [ 001 ] = 1 s 33
takes place when stretched in the [001] direction.
3. At φ = π / 4 , θ = π / 2 the stationary value
E 3 = 1 s 11 0.5 Δ 1
is achieved by stretching in the (001) plane.
4. At φ = 0 , φ = π / 2 and limitation
tan 2 θ = Δ 3 Δ 2 0
the fourth stationary value has the form
E 4 = E 1 E 2 Δ 2 + Δ 3 2 E 1 Δ 2 2 + E 2 Δ 3 2 + 2 E 1 1 Δ 3 E 2 Δ 2 Δ 3 .
This value corresponds to stretching in the (100) (at φ = 0 ) and (010) (at φ = π / 2 ) planes. Young’s moduli E 1 and E 2 also lie in the (100) and (010) planes.
5. In this case, the system of Equations (5) is greatly simplified, and it is possible to obtain a simple form for the fifth stationary value:
E 5 = E 2 E 3 2 Δ 2 + 2 Δ 3 Δ 1 2 4 E 2 Δ 3 2 + E 3 2 Δ 2 Δ 1 2 + 4 Δ 3 E 3 1 Δ 3 E 2 2 Δ 2 Δ 1 ,
which is achieved at φ = π / 4 , φ = 3 π / 4 and limitation
tan 2 θ = 2 Δ 3 2 Δ 2 Δ 1 0 .
Young’s moduli E 2 , E 3 and E 5 lie in the same plane.
A detailed analysis of the extreme values of Young’s modulus for six-constant and seven-constant tetragonal crystals was carried out in [108].
The largest differences between the maximum and minimum values of Young’s modulus were found in Hg 2 I 2 ( E max / E min = 34.6 ), Hg 2 Br 2 ( E max / E min = 29.8 ), Hg 2 Cl 2 ( E max / E min = 24.0 ), TeO 2 ( E max / E min = 12.6 ; for the second set of elastic constants E max / E min = 14.2 ) and (NH 2 ) 2 CO ( E max / E min = 11.6 ; for the second set of elastic constants E max / E min = 24.1 ). The maximum Young’s modulus with E max > 500 GPa was revealed in PdPb 2 ( E max = 684 GPa), stishovite ( E max = 654 GPa). Among tetragonal crystals, 50 auxetics were found. Crystals with minimum Poisson’s ratios of less than 0.5 are Hg 2 Br 2 ( ν min = 1.02 ), Hg 2 I 2 ( ν min = 0.96 ), Hg 2 Cl 2 ( ν min = 0.91 ) and (NH 2 ) 2 CO ( ν min = 0.8 ; for the second set of elastic constants ν min = 0.98 ), TeO 2 ( ν min = 0.80 ; for the second set of elastic constants ν min = 0.85 ) and FeGe 2 ( ν min = 0.77 ) [2,109]. Thus, tetragonal crystals with lowest Poisson’s ratio have the greatest ratio E max / E min . Note that the minimum value of Poisson’s ratio for Hg 2 Br 2 is less than 1 (less than the lower boundary for isotropic materials).

4.2. Hexagonal Crystals

Hexagonal crystals have five independent compliance coefficients, which are obtained under four conditions s 11 = s 22 , s 44 = s 55 , s 13 = s 23 , s 66 = 2 ( s 11 s 12 ) for nine compliance coefficients of rhombic crystals previously given. The expression of Young’s modulus for hexagonal crystals will take the form
E 1 ( θ ) = s 11 sin 4 θ + s 33 cos 4 θ + 0.25 ( 2 s 13 + s 44 ) sin 2 2 θ .
Young’s modulus of hexagonal crystals depends on only one Euler’s angle θ . The dependence of Young’s modulus is a periodic function θ with a period T θ = π . Hexagonal crystals already have two anisotropy coefficients:
Δ 1 s 11 s 12 0.5 s 44 , Δ 2 s 33 s 13 0.5 s 44
and three stationary values of Young’s modulus.
1. At θ = π / 2 the first stationary value has the form
E 1 = E ( 0001 ) = 1 s 11
and achieved by stretching in the (0001) plane.
2. At θ = 0 second stationary value
E 2 = E [ 0001 ] = 1 s 33
takes place in tension in the [0001] direction.
3. When limiting
tan 2 θ = Δ 2 Δ 1 0
the third stationary value has the form
E 3 = E 1 E 2 Δ 1 + Δ 2 2 E 1 Δ 1 2 + E 2 Δ 2 2 + 2 E 1 1 Δ 2 E 2 Δ 1 Δ 2 .
Young’s moduli E 1 , E 2 and E 3 lie in the same plane.
A detailed analysis of the extreme values of Young’s modulus and Poisson‘s ratio for hexagonal crystals was carried out in [12]. In this article, a classification scheme for the extreme values of Young’s modulus E 1 , E 2 and E 3 , depending on two dimensionless parameters, is also given. The largest differences between the maximum and minimum values of Young’s modulus were found in graphite ( E max / E min = 71.8 ), which has the greatest ratio among rhombic, tetragonal, hexagonal and cubic crystals. A large difference ( E max / E min > 5 ) was also revealed in RbNiCl 3 ( E max / E min = 5.52 ) and CsNiF 3 ( E max / E min = 5.72 for one experimental set of compliance coefficients and 10.6 for the second set of compliance coefficients) [12]. Maximum Young’s modulus with E max > 500 GPa were detected in graphite ( E max = 1020 GPa), WC ( E max = 827 GPa), SiC ( E max = 556 GPa), Re ( E max = 588 GPa) and Ru ( E max = 550 GPa). Graphite with hexagonal anisotropy and diamond with cubic anisotropy have the highest Young’s moduli ( E max > 1 TPa) among the rhombic, tetragonal, hexagonal and cubic crystals from [4].
Among hexagonal crystals, six auxetics have been detected [12]. These crystals are MoS 2 ( ν min = 0.28 ), C 7 H 12 ( ν min = 0.15 ), Zn ( ν min = 0.07 ), MnAs ( ν min = 0.04 ), Be-Cu at 2.4% ( ν min = 0.04 ), Be ( ν min = 0.005 ) and Be-Cu at 1.1% Cu ( ν min = 0.005 ). This number of crystalline auxetics is the smallest among rhombic, tetragonal, hexagonal and cubic crystals. For hexagonal crystals, no relationship between the ratio E max / E min and negativity of Poisson’s ratio was found, unlike rhombic, tetragonal and cubic crystals.

4.3. Cubic Crystals

Cubic crystals have only three independent compliance coefficients, s 11 = s 22 = s 33 , s 44 = s 55 = s 66 and s 12 = s 13 = s 23 . The expression of Young’s modulus for cubic crystals has the form
E 1 ( φ , θ ) = s 11 0.5 Δ ( sin 2 2 θ + sin 4 θ sin 2 2 φ ) .
The dependence of Young’s modulus is a periodic function φ , θ with periods T φ = π / 2 and T θ = π . Cubic crystals are characterized by one anisotropy coefficient
Δ s 11 s 12 0.5 s 44
and have three stationary values of Young’s modulus.
1. At θ = 0 and an arbitrary angle φ φ = π / 2 , θ = π / 2 ; φ = 0 , θ = π / 2 —the first stationary value has the form
E 1 = E [ 100 ] = E [ 010 ] = E [ 001 ] = 1 s 11
and is achieved by stretching in the [100], [010] and [001] directions.
2. At φ = 0 , θ = π / 4 ; φ = π / 2 , θ = π / 4 ; φ = π / 4 , θ = π / 2 the second stationary value
E 2 = E [ 110 ] = 1 s 11 0.5 Δ = E 1 1 Δ E 1 / 2
is achieved by stretching in the [110] direction.
3. At φ = π / 4 , tan θ = 2 the third stationary has the form
E 3 = E [ 111 ] = 1 s 11 2 Δ / 3 = E 1 1 2 Δ E 1 / 3 = E 2 1 Δ E 2 / 6
and corresponds to stretching in the [111] direction. This value is conveniently obtained from the fifth stationary value for tetragonal crystals.
Whether the magnitude of Young’s modulus is the maximum or minimum depends on the sign and value of the anisotropy coefficient Δ . For a subclass of cubic crystals with Δ > 0 from (18)–(20), inequalities follow:
E [ 111 ] > E [ 110 ] > E [ 100 ] .
For example, Li, Na, K, Rb, Cs, Ca, Fe, Ni, Cu, Ag, Au, Al, C, Si and Ge have positive anisotropy coefficients ( Δ ).
For a subclass of cubic crystals with Δ < 0 from (18)–(20), opposite inequalities follow:
E [ 100 ] > E [ 110 ] > E [ 111 ] .
For example, V, Cr, Mo and Nb have negative anisotropy coefficients ( Δ ).
The maximum Young’s moduli with E max > 500 GPa were detected in diamond ( E max = 1207 GPa), Ir ( E max = 649 GPa; for the second set of elastic constants E max = 620 GPa), ReO 3 ( E max = 571 GPa; for the second set of elastic constants E max = 478 GPa), NbC 0.865 ( E max = 526 GPa), SiC ( E max = 511 GPa; for the second set of elastic constants E max = 547 GPa) and CeB 6 ( E max = 508 GPa; for the second set of elastic constants E max = 472 GPa). The largest differences between the maximum and minimum values of Young’s modulus were found in InTl (25at%Tl) ( E max / E min = 32.5 ), InTl (28.13at%Tl) ( E max / E min = 26.6 ), InTl (27at%Tl) ( E max / E min = 25.0 ), InTl (30.16at%Tl) ( E max / E min = 21.0 ), NiCr 2 O 4 ( E max / E min = 20.8 ), CuAuZn 2 ( E max / E min = 15.8 ), Au 23 Cu 30 Zn 47 ( E max / E min = 10.8 ), InTl (39.06at%Tl) ( E max / E min = 10.8 ) and CuSi (4.17at%Si) ( E max / E min = 10.2 ). InTl alloys are shape memory materials. Additionally, a minimum value of Poisson’s ratio of less than 1 was detected in some InTl alloys [11,86,102]: ν min = 1.17 for InTl (25at%Tl) and ν min = 1.02 for InTl (27at%Tl). Some other crystals also have large negative Poisson’s ratio values: ν min = 0.81 for InTl (28.13at%Tl), ν min = 0.77 for InTl (30.16at%Tl), ν min = 0.59 for InTl (39.06at%Tl), ν min = 0.77 for NiCr 2 O 4 , ν min = 0.72 for CuAuZn 2 , ν min = 0.62 for Au 23 Cu 30 Zn 47 and ν min = 0.16 for CuSi (4.17at%Si). In the case of cubic crystals, a relationship between the maximum ratio E max / E min and the negativity of Poisson’s ratio can also be observed. All these crystals with negative Poisson’s ratios have positive anisotropy ratios ( Δ ).

5. Conclusions

In the article, the variability of Young’s moduli of rhombic crystals was analyzed. Analytical expressions of seven stationary values were obtained. Three stationary values always exist. Four other values occur when the additional conditions are met. In the case of rhombic crystals, the six stationary values of Young’s modulus were revealed upon tension in the (100), (010) and (001) planes. Three of these values have a simple form and correspond to stretching in the [100], [010] and [001] directions. In addition, these six stationary values of Young’s modulus can be extremes under certain conditions. The seventh stationary value is the inflection point for all 142 rhombic crystals indicated in [4].
Analytical stationary values of Young’s modulus for tetragonal, hexagonal and cubic crystals were written out as special cases of rhombic crystals. Tetragonal crystals already have five stationary values of Young’s modulus, whereas hexagonal and cubic crystals have three. In the case of tetragonal and hexagonal crystals, all stationary values can be global extrema under certain conditions. For cubic crystals, only two stationary values are global extrema ( E [ 100 ] or E [ 111 ] ).
In the article, a numerical analysis of the stationary and extreme values of Young’s modulus of rhombic crystals was also carried out, and the angles at which these values were revealed were determined. For three stationary values of Young’s moduli of rhombic crystals corresponding to tension in the [100], [010] and [001] directions, a classification scheme based on two dimensionless parameters was presented. Rhombic crystals with strong anisotropy ( E max / E min ) were detected.
More than 50 auxetics have been identified among rhombic crystals. The largest differences between the maximum and minimum values of Young’s modulus of rhombic crystals were found in (CH 3 ) 3 NCH 2 COO·(CH) 2 (COOH) 2 ( E max / E min = 12.7 ), I ( E max / E min = 10.1 ), SC(NH 2 ) 2 ( E max / E min = 9.96 ; for the second set of elastic constants E max / E min = 11.8 ), (CH 3 ) 3 NCH 2 COO·H 3 BO 3 ( E max / E min = 9.92 ), Cu-14 wt%Al, 3.0wt%Ni ( E max / E min = 7.51 ), NH 4 B 5 O 8 · 4H 2 O ( E max / E min = 7.39 ), NH 4 HC 2 O 4 · 1/2H 2 O ( E max / E min = 5.83 ), C 6 N 2 O 3 H 6 ( E max / E min = 5.58 ) and CaSO 4 ( E max / E min = 5.4 ). Most of these crystals have negative minimum values of Poisson’s ratio: (CH 3 ) 3 NCH 2 COO·(CH) 2 (COOH) 2 ( ν min = 0.05 ), I ( ν min = 0.48 ), SC(NH 2 ) 2 ( ν min = 0.37 ), (CH 3 ) 3 NCH 2 COO·H 3 BO 3 ( ν min = 0.39 ), Cu-14wt%Al, 3.0wt%Ni ( ν min = 0.70 ), NH 4 B 5 O 8 · 4H 2 O ( ν min = 0.10 ), C 6 N 2 O 3 H 6 ( ν min = 0.91 ) and CaSO 4 ( ν min = 0.05 ). Twenty-four of the thirty-three rhombic crystals with E max / E min > 3 have negative Poisson’s ratios.
The same relationship between these factors was revealed for crystals with tetragonal and cubic anisotropy. The largest differences between the maximum and minimum values of Young’s modulus of tetragonal crystals were found in Hg 2 I 2 ( E max / E min = 34.6 ), Hg 2 Br 2 ( E max / E min = 29.8 ), Hg 2 Cl 2 ( E max / E min = 24.0 ), TeO 2 ( E max / E min = 12.6 ; for the second set of elastic constants E max / E min = 14.2 ) and (NH 2 ) 2 CO ( E max / E min = 11.6 ; for the second set of elastic constants E max / E min = 24.1 ). All these crystals have negative minimum values of Poisson’s ratio: Hg 2 I 2 ( ν min = 0.96 ), Hg 2 Br 2 ( ν min = 1.02 ), Hg 2 Cl 2 ( ν min = 0.91 ), TeO 2 ( ν min = 0.80 ; for the second set of elastic constants ν min = 0.85 ) and (NH 2 ) 2 CO ( ν min = 0.8 ; for the second set of elastic constants ν min = 0.98 ). In the case of cubic crystals, the largest differences between the maximum and minimum values of Young’s modulus were found in InTl (25at%Tl) ( E max / E min = 32.5 ), InTl (28.13at%Tl) ( E max / E min = 26.6 ), InTl (27at%Tl) ( E max / E min = 25.0 ), InTl (30.16at%Tl) ( E max / E min = 21.0 ), NiCr 2 O 4 ( E max / E min = 20.8 ), CuAuZn 2 ( E max / E min = 15.8 ), Au 23 Cu 30 Zn 47 ( E max / E min = 10.8 ), InTl (39.06at%Tl) ( E max / E min = 10.8 ) and CuSi (4.17at%Si) ( E max / E min = 10.2 ). All these crystals have negative minimum values of Poisson’s ratio: InTl (25at%Tl) ( ν min = 1.17 ), InTl (28.13at%Tl) ( ν min = 0.81 ), InTl (27at%Tl) ( ν min = 1.02 ), InTl (30.16at%Tl) ( ν min = 0.77 ), NiCr 2 O 4 ( ν min = 0.77 ), CuAuZn 2 ( ν min = 0.72 ), Au 23 Cu 30 Zn 47 ( ν min = 0.62 ), InTl (39.06at%Tl) ( ν min = 0.59 ) and CuSi (4.17at%Si) ( ν min = 0.16 ). For hexagonal crystals, the relationship between the largest ratio E max / E min and the minimum value of Poisson’s ratio was not revealed.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/cryst11080863/s1. Table S1: Values of anisotropy coefficients Δ 1 , Δ 2 , Δ 3 , Δ 4 , Δ 5 and Δ 6 of rhombic crystals. Table S2: Values of Young’s modulus, E 1 , E 2 , E 3 , E 4 , E 5 , E 6 and E 7 , for rhombic crystals. Global maximum and minimum values are shown in bold. Angle values are given in degrees. Table S3: The values of the minimum and maximum Young’s moduli E min and E max and their ratios E max / E min ; and the values of the minimum and maximum Poisson’s ratios ν min and ν max .

Author Contributions

Conceptualization, methodology, V.A.G. and D.S.L.; software, formal analysis, D.S.L.; writing—original draft preparation, D.S.L. All authors have read and agreed to the published version of the manuscript.

Funding

The reported study was funded by the Government program of IPMech RAS AAAA-A20-120011690136-2 and by RFBR, project number 20-31-70035.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ting, T.C.T.; Chen, T. Poisson’s ratio for anisotropic elastic materials can have no bounds. Quart. J. Mech. Appl. Math. 2005, 58, 73–82. [Google Scholar] [CrossRef]
  2. Lethbridge, Z.A.D.; Walton, R.I.; Marmier, A.S.H.; Smith, C.W.; Evans, K.E. Elastic anisotropy and extreme Poisson’s ratios in single crystals. Acta Mater. 2010, 58, 6444–6451. [Google Scholar] [CrossRef] [Green Version]
  3. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S. Auxetic mechanics of crystalline materials. Mech. Solids 2010, 45, 529–545. [Google Scholar] [CrossRef]
  4. Nelson, D.F. (Ed.) Second and Higher Order Elastic Constants. In Landolt-Börnstein—Group III Condensed Matter; Springer: Berlin/Heidelberg, Germany, 1992; Volume 29a. [Google Scholar] [CrossRef]
  5. Ting, T.C.T. The stationary values of Young’s modulus for monoclinic and triclinic materials. J. Mech. 2005, 21, 249–253. [Google Scholar] [CrossRef]
  6. Ting, T.C.T. Explicit expression of the stationary values of Young’s modulus and the shear modulus for anisotropic elastic materials. J. Mech. 2005, 21, 255–266. [Google Scholar] [CrossRef]
  7. Norris, A.N. Extreme values of Poisson’s ratio and other engineering moduli in anisotropic materials. J. Mech. Mater. Struc. 2006, 1, 793–812. [Google Scholar] [CrossRef] [Green Version]
  8. Cazzani, A.; Rovati, M. Extrema of Young’s modulus for cubic and transversely isotropic solids. Int. J. Solids Struct. 2003, 40, 1713–1744. [Google Scholar] [CrossRef]
  9. Cazzani, A.; Rovati, M. Extrema of Young’s modulus for elastic solids with tetragonal symmetry. Int. J. Solids Struct. 2005, 42, 5057–5096. [Google Scholar] [CrossRef]
  10. Norris, A.N. Poisson’s ratio in cubic materials. Proc. Roy. Soc. A 2006, 462, 3385–3405. [Google Scholar] [CrossRef]
  11. Epishin, A.I.; Lisovenko, D.S. Extreme values of the Poisson’s ratio of cubic crystals. Tech. Phys. 2016, 61, 1516–1524. [Google Scholar] [CrossRef]
  12. Gorodtsov, V.A.; Lisovenko, D.S. Extreme values of Young’s modulus and Poisson’s ratio of hexagonal crystals. Mech. Mater. 2019, 134, 1–8. [Google Scholar] [CrossRef]
  13. Love, A.E.H. A Treatise on the Mathematical Theory of Elasticity; Cambridge University Press: Cambridge, UK, 1892; Volume 1. [Google Scholar]
  14. Lakes, R.S. Foam structures with a negative Poisson’s ratio. Science 1987, 235, 1038–1040. [Google Scholar] [CrossRef]
  15. Friis, E.A.; Lakes, R.S.; Park, J.B. Negative Poisson’s ratio polymeric and metallic foams. J. Mater. Sci. 1988, 23, 4406–4414. [Google Scholar] [CrossRef]
  16. Evans, K.E.; Nkansah, M.A.; Hutchinson, I.J.; Rogers, S.C. Molecular network design. Nature 1991, 353, 124. [Google Scholar] [CrossRef]
  17. Evans, K.E. Auxetic polymers: A new range of materials. Endeavour 1991, 15, 170–174. [Google Scholar] [CrossRef]
  18. Wojciechowski, K.W. Constant thermodynamic tension Monte Carlo studies of elastic properties of a two-dimensional system of hard cyclic hexamers. Mol. Phys. 1987, 61, 1247–1258. [Google Scholar] [CrossRef]
  19. Wojciechowski, K.W. Two-dimensional isotropic system with a negative Poisson ratio. Phys. Lett. A 1989, 137, 60–64. [Google Scholar] [CrossRef]
  20. Tretiakov, K.V.; Wojciechowski, K.W. Poisson’s ratio of simple planar ‘isotropic’ solids in two dimensions. Phys. Status Solidi B 2007, 244, 1038–1046. [Google Scholar] [CrossRef]
  21. Tretiakov, K.V.; Wojciechowski, K.W. Auxetic, partially auxetic, and nonauxetic behaviour in 2D crystals of hard cyclic tetramers. Phys. Status Solidi RRL 2020, 14, 2000198. [Google Scholar] [CrossRef]
  22. Wojciechowski, K.W.; Brańka, A.C. Negative Poisson ratio in a two-dimensional “isotropic” solid. Phys. Rev. A 1989, 40, 7222–7225. [Google Scholar] [CrossRef]
  23. Wojciechowski, K.W. Non-chiral, molecular model of negative Poisson ratio in two dimensions. J. Phys. A 2003, 36, 11765–11778. [Google Scholar] [CrossRef]
  24. Narojczyk, J.W.; Kowalik, M.; Wojciechowski, K.W. Influence of nanochannels on Poisson’s ratio of degenerate crystal of hard dimers. Phys. Status Solidi B 2016, 253, 1324–1330. [Google Scholar] [CrossRef]
  25. Tretiakov, K.V.; Piglowski, P.M.; Hyzorek, K.; Wojciechowski, K.W. Enhanced auxeticity in Yukawa systems due to introduction of nanochannels in [001]-direction. Smart Mater. Struct. 2016, 25, 054007. [Google Scholar] [CrossRef]
  26. Piglowski, P.M.; Wojciechowski, K.W.; Tretiakov, K.V. Partial auxeticity induced by nanoslits in the Yukawa crystal. Phys. Status Solidi RRL 2016, 10, 566–569. [Google Scholar] [CrossRef]
  27. Pigłowski, P.; Narojczyk, J.; Poźniak, A.; Wojciechowski, K.; Tretiakov, K. Auxeticity of Yukawa systems with nanolayers in the (111) crystallographic plane. Materials 2017, 10, 1338. [Google Scholar] [CrossRef] [Green Version]
  28. Tretiakov, K.; Pigłowski, P.; Narojczyk, J.; Bilski, M.; Wojciechowski, K. High partial auxeticity Induced by nanochannels in [111]-direction in a simple model with Yukawa interactions. Materials 2018, 11, 2550. [Google Scholar] [CrossRef] [Green Version]
  29. Tretiakov, K.V.; Pigłowski, P.M.; Narojczyk, J.W.; Wojciechowski, K.W. Selective enhancement of auxeticity through changing a diameter of nanochannels in Yukawa systems. Smart Mater. Struct. 2018, 27, 115021. [Google Scholar] [CrossRef]
  30. Narojczyk, J.W.; Wojciechowski, K.W.; Tretiakov, K.V.; Smardzewski, J.; Scarpa, F.; Piglowski, P.M.; Kowalik, M.; Imre, A.R.; Bilski, M. Auxetic properties of a f.c.c. crystal of hard spheres with an array of [001]-nanochannels filled by hard spheres of another diameter. Phys. Status Solidi B 2019, 256, 1800611. [Google Scholar] [CrossRef] [Green Version]
  31. Narojczyk, J.; Wojciechowski, K. Poisson’s ratio of the f.c.c. hard sphere crystals with periodically stacked (001)-nanolayers of hard spheres of another diameter. Materials 2019, 12, 700. [Google Scholar] [CrossRef] [Green Version]
  32. Narojczyk, J.W.; Wojciechowski, K.W.; Smardzewski, J.; Imre, A.R.; Grima, J.N.; Bilski, M. Cancellation of auxetic properties in f.c.c. hard sphere crystals by hybrid layer-channel nanoinclusions filled by hard spheres of another diameter. Materials 2021, 14, 3008. [Google Scholar] [CrossRef]
  33. Grima, J.N.; Evans, K.E. Auxetic behavior from rotating squares. J. Mater. Sci. Lett. 2000, 19, 1563–1565. [Google Scholar] [CrossRef]
  34. Grima, J.N.; Jackson, R.; Alderson, A.; Evans, K.E. Do zeolites have negative Poisson’s ratios? Adv. Mater. 2000, 12, 1912–1918. [Google Scholar] [CrossRef]
  35. Grima, J.N.; Evans, K.E. Self expanding molecular networks. Chem. Commun. 2000, 1531–1532. [Google Scholar] [CrossRef]
  36. Grima, J.N.; Alderson, A.; Evans, K.E. Negative Poisson’s ratios from rotating rectangles. Comput. Methods Sci. Technol. 2004, 10, 137–145. [Google Scholar] [CrossRef] [Green Version]
  37. Grima, J.N.; Alderson, A.; Evans, K.E. Auxetic behaviour from rotating rigid units. Phys. Status Solidi B 2005, 242, 561–575. [Google Scholar] [CrossRef]
  38. Grima, J.N.; Gatt, R.; Alderson, A.; Evans, K.E. On the potential of connected stars as auxetic systems. Molec. Simul. 2005, 31, 925–935. [Google Scholar] [CrossRef] [Green Version]
  39. Grima, J.N.; Gatt, R.; Alderson, A.; Evans, K.E. On the origin of auxetic behaviour in the silicate α-cristobalite. J. Mater. Chem. 2005, 15, 4003–4005. [Google Scholar] [CrossRef]
  40. Grima, J.N.; Evans, K.E. Auxetic behavior from rotating triangles. J. Mater. Sci. 2006, 41, 3193–3196. [Google Scholar] [CrossRef]
  41. Grima, J.N.; Gatt, R.; Alderson, A.E.; Evans, K. An alternative explanation for the negative Poisson’s ratios in α-cristobalite. Mater. Sci. Eng. A 2006, 423, 219–224. [Google Scholar] [CrossRef]
  42. Grima, J.N.; Gatt, R.; Ravirala, N.; Alderson, A.; Evans, K.E. Negative Poisson’s ratios in cellular foam materials. Mater. Sci. Eng. A 2006, 423, 214–218. [Google Scholar] [CrossRef]
  43. Grima, J.N.; Zammit, V.; Gatt, R.; Alderson, A.; Evans, K.E. Auxetic behaviour from rotating semi-rigid units. Phys. Status Solidi B 2007, 244, 866–882. [Google Scholar] [CrossRef]
  44. Grima, J.N.; Gatt, R.; Zammit, V.; Williams, J.J.; Evans, K.E.; Alderson, A.; Walton, R.I. Natrolite: A zeolite with negative Poisson’s ratios. J. Appl. Phys. 2007, 101, 086102. [Google Scholar] [CrossRef]
  45. Grima, J.N.; Farrugia, P.S.; Caruana, C.; Gatt, R.; Attard, D. Auxetic behaviour from stretching connected squares. J. Mater. Sci. 2008, 43, 5962–5971. [Google Scholar] [CrossRef]
  46. Attard, D.; Grima, J.N. Auxetic behaviour from rotating rhombi. Phys. Status Solidi B 2008, 245, 2395–2404. [Google Scholar] [CrossRef]
  47. Grima, J.N.; Farrugia, P.S.; Gatt, R.; Attard, D. On the auxetic properties of rotating rhombi and parallelograms: A preliminary investigation. Phys. Status Solidi B 2008, 245, 521–529. [Google Scholar] [CrossRef]
  48. Attard, D.; Manicaro, E.; Grima, J.N. On rotating rigid parallelograms and their potential for exhibiting auxetic behaviour. Phys. Status Solidi B 2009, 246, 2033–2044. [Google Scholar] [CrossRef]
  49. Grima, J.N.; Cassar, R.N.; Gatt, R. On the effect of hydrostatic pressure on the auxetic character of NAT-type silicates. J. Non-Cryst. Solids 2009, 355, 1307–1312. [Google Scholar] [CrossRef]
  50. Attard, D.; Manicaro, E.; Gatt, R.; Grima, J.N. On the properties of auxetic rotating stretching squares. Phys. Status Solidi B 2009, 246, 2045–2054. [Google Scholar] [CrossRef]
  51. Grima, J.N.; Gatt, R.; Ellul, B.; Chetcuti, E. Auxetic behaviour in non-crystalline materials having star or triangular shaped perforations. J. Non-Cryst. Solids 2010, 356, 1980–1987. [Google Scholar] [CrossRef]
  52. Grima, J.N.; Gatt, R. Perforated sheets exhibiting negative Poisson’s ratios. Adv. Eng. Mater. 2010, 12, 460–464. [Google Scholar] [CrossRef]
  53. Grima, J.N.; Manicaro, E.; Attard, D. Auxetic behaviour from connected different-sized squares and rectangles. Proc. Roy. Soc. A 2011, 467, 439–458. [Google Scholar] [CrossRef] [Green Version]
  54. Grima, J.N.; Chetcuti, E.; Manicaro, E.; Attard, D.; Camilleri, M.; Gatt, R.; Evans, K.E. On the auxetic properties of generic rotating rigid triangles. Proc. Roy. Soc. A 2012, 468, 810–830. [Google Scholar] [CrossRef]
  55. Gatt, R.; Mizzi, L.; Azzopardi, K.M.; Grima, J.N. A force-field based analysis of the deformation mechanism in α-cristobalite. Phys. Status Solidi B 2015, 252, 1479–1485. [Google Scholar] [CrossRef]
  56. Gatt, R.; Mizzi, L.; Azzopardi, J.I.; Azzopardi, K.M.; Attard, D.; Casha, A.; Briffa, J.; Grima, J.N. Hierarchical auxetic mechanical metamaterials. Sci. Rep. 2015, 5, 8395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Dudek, K.K.; Gatt, R.; Mizzi, L.; Dudek, M.R.; Attard, D.; Evans, K.E.; Grima, J.N. On the dynamics and control of mechanical properties of hierarchical rotating rigid unit auxetics. Sci. Rep. 2017, 7, 46529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Attard, D.; Casha, A.; Grima, J. Filtration properties of auxetics with rotating rigid units. Materials 2018, 11, 725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Evans, K.E.; Alderson, A. Rotation and dilation deformation mechanisms for auxetic behaviour in the α-cristobalite tetrahedral framework structure. Phys. Chem. Min. 2001, 28, 711–718. [Google Scholar] [CrossRef]
  60. Alderson, A.; Evans, K.E. Molecular origin of auxetic behavior in tetrahedral framework silicates. Phys. Rev. Lett. 2002, 89, 225503. [Google Scholar] [CrossRef]
  61. Alderson, A.; Alderson, K.L.; Evans, K.E.; Grima, J.N.; Williams, M.R.; Davies, P.J. Molecular modeling of the deformation mechanisms acting in auxetic silica. Comput. Methods Sci. Technol. 2004, 10, 117–126. [Google Scholar] [CrossRef]
  62. Alderson, A.; Alderson, K.L.; Evans, K.E.; Grima, J.N.; Williams, M.R.; Davies, P.J. Modelling the deformation mechanisms, structure-property relationships and applications of auxetic nanomaterials. Phys. Status Solidi B 2005, 242, 499–508. [Google Scholar] [CrossRef]
  63. Alderson, A.; Alderson, K.L.; Evans, K.E.; Grima, J.N.; Williams, M.S. Modelling of negative Poisson’s ratio nanomaterials: Deformation mechanisms, structure-property relationships and applications. J. Metastab. Nanocryst. Mater. 2005, 23, 55–58. [Google Scholar] [CrossRef]
  64. Alderson, A.; Evans, K.E. Deformation mechanisms leading to auxetic behaviour in the α-cristobalite and α-quartz structures of both silica and germania. J. Phys. Condens. Matter 2008, 21, 025401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Grima, J.N.; Zammit, V.; Gatt, R.; Attard, D.; Caruana, C.; Chircop Bray, T.G. On the role of rotating tetrahedra for generating auxetic behavior in NAT and related systems. J. Non-Cryst. Solids 2008, 354, 4214–4220. [Google Scholar] [CrossRef]
  66. Attard, D.; Grima, J.N. A three-dimensional rotating rigid units network exhibiting negative Poisson’s ratios. Phys. Status Solidi B 2012, 249, 1330–1338. [Google Scholar] [CrossRef]
  67. Azzopardi, K.M.; Brincat, J.P.; Grima, J.N.; Gatt, R. Anomalous elastic properties in stishovite. RSC Adv. 2015, 5, 8974–8980. [Google Scholar] [CrossRef]
  68. Nazaré, F.; Alderson, A. Models for the prediction of Poisson’s ratio in the ‘α-cristobalite’ tetrahedral framework. Phys. Status Solidi B 2015, 252, 1465–1478. [Google Scholar] [CrossRef]
  69. Grima-Cornish, J.N.; Vella-Żarb, L.; Grima, J.N. Negative linear compressibility and auxeticity in boron arsenate. Annal. Physik 2020, 532, 1900550. [Google Scholar] [CrossRef]
  70. Grima-Cornish, J.N.; Grima, J.N.; Attard, D. A novel mechanical metamaterial exhibiting auxetic behavior and negative compressibility. Materials 2020, 13, 79. [Google Scholar] [CrossRef] [Green Version]
  71. Dudek, K.K.; Attard, D.; Gatt, R.; Grima-Cornish, J.N.; Grima, J.N. The multidirectional auxeticity and negative linear compressibility of a 3D mechanical metamaterial. Materials 2020, 13, 2193. [Google Scholar] [CrossRef]
  72. Grima-Cornish, J.N.; Vella-Żarb, L.; Wojciechowski, K.W.; Grima, J.N. Shearing deformations of β-cristobalite-like boron arsenate. Symmetry 2021, 13, 977. [Google Scholar] [CrossRef]
  73. Milstein, F.; Huang, K. Existence of a negative Poisson ratio in fcc crystals. Phys. Rev. B 1979, 19, 2030–2033. [Google Scholar] [CrossRef]
  74. Baughman, R.H.; Shacklette, J.M.; Zakhidov, A.A.; Stafström, S. Negative Poisson’s ratios as a common feature of cubic metals. Nature 1998, 392, 362–365. [Google Scholar] [CrossRef]
  75. Paszkiewicz, T.; Pruchnik, M.; Wolski, S. Slowness surfaces and energy focusing patterns of auxetic cubic media. Comput. Methods Sci. Technol. 2004, 10, 183–195. [Google Scholar] [CrossRef] [Green Version]
  76. Paszkiewicz, T.; Wolski, S. Anisotropic properties of mechanical characteristics and auxeticity of cubic crystalline media. Phys. Status Solidi B 2007, 244, 966–977. [Google Scholar] [CrossRef] [Green Version]
  77. 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]
  78. Jasiukiewicz, C.; Paszkiewicz, T.; Wolski, S. Auxetic properties and anisotropy of elastic material constants of 2D crystalline media. Phys. Status Solidi B 2008, 245, 562–569. [Google Scholar] [CrossRef] [Green Version]
  79. Brańka, A.C.; Wojciechowski, K.W. Auxeticity of cubic materials: The role of repulsive core interaction. J. Non-Cryst. Solids 2008, 354, 4143–4145. [Google Scholar] [CrossRef]
  80. Branka, A.C.; Heyes, D.M.; Wojciechowski, K.W. Auxeticity of cubic materials. Phys. Status Solidi B 2009, 246, 2063–2071. [Google Scholar] [CrossRef]
  81. Branka, A.C.; Heyes, D.M.; Wojciechowski, K.W. Auxeticity of cubic materials under pressure. Phys. Status Solidi B 2011, 248, 96–104. [Google Scholar] [CrossRef]
  82. Branka, A.C.; Heyes, D.M.; Maćkowiak, S.; Pieprzyk, S.; Wojciechowski, K.W. Cubic materials in different auxetic regions: Linking microscopic to macroscopic formulations. Phys. Status Solidi B 2012, 249, 1373–1378. [Google Scholar] [CrossRef]
  83. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S. Cubic auxetics. Dokl. Phys. 2011, 56, 399–402. [Google Scholar] [CrossRef]
  84. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S. Relation of Poisson’s ratio on average with Young’s modulus. Auxetics on average. Dokl. Phys. 2012, 57, 174–178. [Google Scholar] [CrossRef]
  85. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S. Classification of cubic auxetics. Phys. Status Solidi B 2013, 250, 2038–2043. [Google Scholar] [CrossRef]
  86. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S.; Volkov, M.A. Negative Poisson’s ratio for cubic crystals and nano/microtubes. Phys. Mesomech. 2014, 17, 97–115. [Google Scholar] [CrossRef]
  87. Krasavin, V.V.; Krasavin, A.V. Auxetic properties of cubic metal single crystals. Phys. Status Solidi B 2014, 251, 2314–2320. [Google Scholar] [CrossRef]
  88. Ho, D.T.; Park, S.D.; Kwon, S.Y.; Park, K.; Kim, S.Y. Negative Poisson’s ratios in metal nanoplates. Nat. Commun. 2014, 5, 3255. [Google Scholar] [CrossRef]
  89. Ho, D.T.; Kim, H.; Kwon, S.Y.; Kim, S.Y. Auxeticity of face-centered cubic metal (001) nanoplates. Phys. Status Solidi B 2015, 252, 1492–1501. [Google Scholar] [CrossRef]
  90. Ho, D.T.; Park, S.D.; Kwon, S.Y.; Han, T.S.; Kim, S.Y. Negative Poisson’s ratio in cubic materials along principal directions. Phys. Status Solidi B 2016, 253, 1288–1294. [Google Scholar] [CrossRef]
  91. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S.; Volkov, M.A. Two-layer tubes from cubic crystals. Dokl. Phys. 2016, 61, 604–610. [Google Scholar] [CrossRef]
  92. Lisovenko, D.S.; Baimova, J.A.; Rysaeva, L.K.; Gorodtsov, V.A.; Rudskoy, A.I.; Dmitriev, S.V. Equilibrium diamond-like carbon nanostructures with cubic anisotropy: Elastic properties. Phys. Status Solidi B 2016, 253, 1295–1302. [Google Scholar] [CrossRef]
  93. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S.; Volkov, M.A. Two-layered tubes from cubic crystals: Auxetic tubes. Phys. Status Solidi B 2017, 254, 1600815. [Google Scholar] [CrossRef]
  94. Lisovenko, D.S.; Baimova, Y.A.; Rysaeva, L.K.; Gorodtsov, V.A.; Dmitriev, S.V. Equilibrium structures of carbon diamond-like clusters and their elastic properties. Phys. Solid State 2017, 59, 820–828. [Google Scholar] [CrossRef]
  95. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S. Longitudinal elastic tension of two-layered plates from isotropic auxetics-nonauxetics and cubic crystals. Eur. J. Mech. A Solids 2017, 63, 122–127. [Google Scholar] [CrossRef]
  96. Gorodtsov, V.A.; Lisovenko, D.S.; Lim, T.C. Three-layered plate exhibiting auxeticity based on stretching and bending modes. Compos. Struct. 2018, 194, 643–651. [Google Scholar] [CrossRef]
  97. Rysaeva, L.K.; Baimova, J.A.; Lisovenko, D.S.; Gorodtsov, V.A.; Dmitriev, S.V. Elastic properties of fullerites and diamond-like phases. Phys. Status Solidi B 2019, 256, 1800049. [Google Scholar] [CrossRef] [Green Version]
  98. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S.; Volkov, M.A. Thin homogeneous two-layered plates of cubic crystals with different layer orientation. Phys. Mesomech. 2019, 22, 261–268. [Google Scholar] [CrossRef]
  99. Bryukhanov, I.A.; Gorodtsov, V.A.; Lisovenko, D.S. Chiral Fe nanotubes with both negative Poisson’s ratio and Poynting’s effect. Atomistic simulation. J. Phys. Cond. Matt. 2019, 31, 475304. [Google Scholar] [CrossRef] [PubMed]
  100. Bryukhanov, I.A.; Gorodtsov, V.A.; Lisovenko, D.S. Modeling of the mechanical properties of chiral metallic nanotubes. Phys. Mesomech. 2020, 23, 477–486. [Google Scholar] [CrossRef]
  101. Bielejewska, N.; Brańka, A.C.; Pieprzyk, S.; Yevchenko, T. Another look at auxeticity of 2D square media. Phys. Status Solidi B 2020, 257, 2000485. [Google Scholar] [CrossRef]
  102. Gorodtsov, V.A.; Lisovenko, D.S. Auxetics among materials with cubic anisotropy. Mech. Solids 2020, 55, 461–474. [Google Scholar] [CrossRef]
  103. Volkov, M.A.; Gorodtsov, V.A.; Fadeev, E.P.; Lisovenko, D.S. Stretching of chiral tubes obtained by rolling-up plates of cubic crystals with various orientations. J. Mech. Mater. Struct. 2021, 16, 139–157. [Google Scholar] [CrossRef]
  104. Sirotin, Y.I.; Shaskolskaya, M.P. Fundamentals of Crystal Physics; Mir: Moscow, Russia, 1982. [Google Scholar]
  105. Nye, J.F. Physical Properties of Crystals; Clarendon Press: Oxford, UK, 1957; 329p. [Google Scholar]
  106. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S. Young’s modulus of cubic auxetics. Lett. Mater. 2011, 1, 127–132. [Google Scholar] [CrossRef] [Green Version]
  107. Smirnov, V.I. A Course of Higher Mathematics, Vol. I: Elementary Calculus; Pergamon Press: Oxford, UK, 1964; 546p. [Google Scholar]
  108. Gorodtsov, V.A.; Tkachenko, V.G.; Lisovenko, D.S. Extreme values of Young’s modulus of tetragonal crystals. Mech. Mater. 2021, 154, 103724. [Google Scholar] [CrossRef]
  109. Goldstein, R.V.; Gorodtsov, V.A.; Lisovenko, D.S.; Volkov, M.A. Auxetics among 6-constant tetragonal crystals. Lett. Mater. 2015, 5, 409–413. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Classification scheme for stationary values of Young’s modulus ( E 1 , E 2 and E 3 ) for rhombic crystals. The points indicate the values of dimensionless parameters α and β for 142 rhombic crystals from [4].
Figure 1. Classification scheme for stationary values of Young’s modulus ( E 1 , E 2 and E 3 ) for rhombic crystals. The points indicate the values of dimensionless parameters α and β for 142 rhombic crystals from [4].
Crystals 11 00863 g001
Figure 2. Young’s modulus surfaces for rhombic crystals related to one of six zones: AgTlSe ( E 1 > E 2 > E 3 ) (a), CaCO 3 ( E 1 > E 3 > E 2 ) (b), Ga ( E 3 > E 1 > E 2 ) (c), Ni 3 B ( E 3 > E 2 > E 1 ) (d), ZnSb ( E 2 > E 3 > E 1 ) (e) or TbF 3 ( E 2 > E 1 > E 3 ) (f).
Figure 2. Young’s modulus surfaces for rhombic crystals related to one of six zones: AgTlSe ( E 1 > E 2 > E 3 ) (a), CaCO 3 ( E 1 > E 3 > E 2 ) (b), Ga ( E 3 > E 1 > E 2 ) (c), Ni 3 B ( E 3 > E 2 > E 1 ) (d), ZnSb ( E 2 > E 3 > E 1 ) (e) or TbF 3 ( E 2 > E 1 > E 3 ) (f).
Crystals 11 00863 g002aCrystals 11 00863 g002b
Table 1. Values of anisotropy coefficients Δ 1 , Δ 2 , Δ 3 , Δ 4 , Δ 5 and Δ 6 of some rhombic crystals.
Table 1. Values of anisotropy coefficients Δ 1 , Δ 2 , Δ 3 , Δ 4 , Δ 5 and Δ 6 of some rhombic crystals.
Crystals Δ 1 , Δ 2 , Δ 3 , Δ 4 , Δ 5 , Δ 6 ,
TPa 1 TPa 1 TPa 1 TPa 1 TPa 1 TPa 1
CaSO 4 42.2 47.5 6.57 8.02 8.16 4.33
CaCO 3 1.71 4.54 13.0 7.73 3.482.48
BaMnF 4 , s E 9.2 1.117.433.912.919.1
Cs 2 SO 4 2.64.42.88.65.49.4
Ga4.26.01.95 1.76 2.1 3.41
In 4 Se 3 15.8 9.15 43.028.0 7.24 2.76
PbBr 2 101 111 51.655.911.125.6
LiGaO 2 , s E 0.552.35 1.85 1.15 2.351.25
MgBaF 4 3.95 3.65 11.38.81.7 1.1
Co 2 SiO 4 2.0 1.09 2.46 0.73 0.95 2.31
Rb 2 SO 4 2.85 3.55 2.353.951.653.95
AgTlSe 4.85 11.087.6143104143
NaNO 2 54.5 74.2 2.5 23.8 19.9 21.5
α -S40.052.026.5 14.5 18.0 71.0
26.8 9.6 35.3 34.5 96.360.7
TbF 3 1.55 2.25 8.6510.5 3.72 1.9
Ni 3 B4.373.732.771.63.633.1
α -U 0.6 1.22 1.97 2.09 5.323.38
ZnSb2.35 0.05 7.657.25 9.5 7.5
Table 2. Values of Young’s moduli E 1 , E 2 , E 3 , E 4 , E 5 , E 6 and E 7 for some rhombic crystals. Global maximum and minimum values are shown in bold.
Table 2. Values of Young’s moduli E 1 , E 2 , E 3 , E 4 , E 5 , E 6 and E 7 for some rhombic crystals. Global maximum and minimum values are shown in bold.
Crystals E 1 E 2 E 3 E 4 E 5 E 6 E 7
GPa GPa GPa GPa GPa GPa GPa
Al 2 SiO 5 188Min251310Max247Min259Max
CaSO 4 90.9Max175Max10590.571.632.4Min
CaCO 3 144Max75.8Min82.089.5Max66.3Min85.8
BaMnF 4 , s E 58.836.6Min29.9Min45.290.1Max
BaSO 4 58.1Min53.292.6Max36.5Min73.2Max
Cs 2 SO 4 32.7Min30.9Min27.5Min32.933.433.7Max
BeAl 2 O4478Max386Max417372Min374379
MgSiO 3 190148Min192182Min202Max
Mg 2 SiO 4 297Max171203171Min199
Ga82.0Min71.4Min118Max95.5Max
In 4 Se 3 23.8Min58.8Max37.062.7Max
I3.05Min9.717.586.24Min30.8Max13.5Max
La 2 CuO 4 117Min116Min159161161202Max
PbBr 2 19.724.718.227.0Max38.7Max10.1Min
LiGaO 2 , s E 137110Min125132Max118Min139Max130
MgBaF 4 70.969.486.2129Max61.9Min72.8
C 6 N 2 O 3 H 6 7.04.91Min3.51Min19.6Max13.7
Co 2 SiO 4 240Max138170133Min165
Mg 2 GeO 4 282Max161187153Min173
Ni 2 SiO 4 270Max175Min189192
KNO 3 26.4Max20.115.414.4Min19.2
K 2 SeO 4 40.3Max39.5Max30.624.7Min37.6
K 2 SO 4 42.445.944.247.4Max46.7Max38.4Min
K 2 ZnCl 4 15.6Max15.7Max20.8Max15.214.813.3Min
RbHSO 4 22.930.132.6Max14.3Min15.7Min31.5Max
Rb 2 SO 4 38.839.836.5Min40.6Max40.1Max36.9Min
Rb 2 ZnBr 4 12.212.9Max16.9Max12.510.0Min
Al 2 SiO 5 242Max153Min279325Max196Min225
AgNO 3 11.3Min13.5Min29.2Max25.6Max
AgTlSe18.714.4Min9.17Min38.849.6Max
NaBF 4 39.4Max29.951.7Max13.9Min29.6
Na 2 GeO 3 , s E 66.083.8Max71.494.7Max54.5Min
NaNO 2 25.1Max49.8Max54.1Max33.725.015.9Min
Na 2 SO 4 65.4Max93.5Max58.842.6Min51.163.0
SrSO 4 45.5Min45.987.7Max39.0Min74.6Max
α -S14.1Min12.033.3Max11.5Min18.7Max
13.4Max9.01Min13.319.3Max10.8Min11.6
Mn 2 SiO 4 198Max116Min146143
TbF 3 101Min163Max85.1Min166Max
Tl 2 SO 4 21.7Min21.0Min27.728.2Max27.922.2
Ni 3 B150Min166Min182246Max203232
α -U204Max149Min209288Max170Min189
ZnSb72.5Min87.7Max74.659.8Min101Max81.7
Table 3. The values of the minimum and maximum Young’s moduli E min , E max and their ratios, E max / E min ; and the values of the minimum and maximum Poisson’s ratios, ν min and ν max .
Table 3. The values of the minimum and maximum Young’s moduli E min , E max and their ratios, E max / E min ; and the values of the minimum and maximum Poisson’s ratios, ν min and ν max .
Crystals E min E max E max / E min ν min ν max
(CH 3 ) 3 NCH 2 COO·(CH) 2 (COOH) 2 2.9237.012.7 0.05 0.91
I3.0530.810.1−0.481.31
SC(NH 2 ) 2 2.3923.89.96 0.28 1.00
2.0924.611.8 0.37 1.07
(CH 3 ) 3 NCH 2 COO·H 3 BO 3 1.8518.39.89 0.39 1.22
Cu-14 wt% Al 3.0 wt% Ni22.31677.49 0.70 1.43
NH 4 B 5 O 8 · 4H 2 O6.8550.77.40 0.10 0.85
NH 4 HC 2 O 4 · 1/2H 2 O10.561.05.810.050.82
C 6 N 2 O 3 H 6 3.5119.65.58 0.91 1.05
(Fe,Mg) 2 (Al,Fe + 3 ) 9 O 6 SiO 4 (O,OH) 2 57.83125.40 0.20 0.95
CaSO 4 32.41755.40 0.05 0.76
AgTlSe9.1749.55.40 0.42 1.07
CH 3 COOLi·2H 2 O11.653.84.640.040.68
CaPb(CN) 4 · 5H 2 O9.7943.54.440.070.71
(CD) 4 N 2 4.4619.24.300.000.71
KB 5 O 8 · 4H 2 O10.243.14.220.060.82
Cd(COOH) 2 8.0633.34.13 0.09 0.98
C 24 H 18 3.2213.14.07 0.06 0.77
Ca(COOH) 2 11.847.84.05 0.23 0.81
C 14 H 12 N 2 2.8411.03.87 0.02 0.76
PbBr 2 10.138.73.83 0.19 0.90
NaBF 4 13.951.73.72 0.05 0.71
(CH 3 NHCH 2 COOH) 3 · CaCl 2 12.644.93.56 0.48 0.76
Na 2 C 4 H 4 O 6 · 2H 2 O11.037.93.45−0.050.88
C 6 H 4 (NO 2 ) 2 5.6819.43.42 0.01 0.60
NIPC2.398.153.41−0.160.84
NaNO 2 15.954.13.400.090.64
CsSCN6.3821.13.310.010.78
(CH 3 ) 3 NCH 2 COO·CaCl 2 · 2H 2 O7.6323.63.09 0.08 0.79
C 6 H 8 O 7 H 2 O8.6126.43.070.000.74
C 5 H 10 ClNO 4 9.9030.33.060.050.61
ZnSO 47 H 2 O16.148.83.03 0.15 0.71
15.630.01.92−0.040.66
[CN 3 H 6 ] 2 C 8 H 4 O 4 3.6110.93.02 0.28 0.94
BaMnF 4 , s E 29.990.13.01 0.05 0.87
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gorodtsov, V.A.; Lisovenko, D.S. The Extreme Values of Young’s Modulus and the Negative Poisson’s Ratios of Rhombic Crystals. Crystals 2021, 11, 863. https://doi.org/10.3390/cryst11080863

AMA Style

Gorodtsov VA, Lisovenko DS. The Extreme Values of Young’s Modulus and the Negative Poisson’s Ratios of Rhombic Crystals. Crystals. 2021; 11(8):863. https://doi.org/10.3390/cryst11080863

Chicago/Turabian Style

Gorodtsov, Valentin A., and Dmitry S. Lisovenko. 2021. "The Extreme Values of Young’s Modulus and the Negative Poisson’s Ratios of Rhombic Crystals" Crystals 11, no. 8: 863. https://doi.org/10.3390/cryst11080863

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