Next Article in Journal
Thickness-Dependent Sign Change of the Magnetoresistance in VTe2 Thin Films
Previous Article in Journal
The Diminishing Role of the Nucleation Rate as Crystallization Develops in Avrami-Type Models
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanical Characterization of Anhydrous Microporous Aluminophosphate Materials: Tridimensional Incompressibility, Ductility, Isotropy and Negative Linear Compressibility

by
Francisco Colmenero
1,2,*,
Álvaro Lobato
1 and
Vicente Timón
2
1
Department of Chemical-Physics, Facultad de Ciencias Químicas, Universidad Complutense de Madrid, 28040 Madrid, Spain
2
Instituto de Estructura de la Materia, Consejo Superior de Investigaciones Científicas, 28006 Madrid, Spain
*
Author to whom correspondence should be addressed.
Solids 2022, 3(3), 457-499; https://doi.org/10.3390/solids3030032
Submission received: 21 June 2022 / Revised: 26 July 2022 / Accepted: 4 August 2022 / Published: 16 August 2022

Abstract

:
Here, a detailed mechanical characterization of five important anhydrous microporous aluminophosphate materials (VPI-5, ALPO-8, ALPO-5, ALPO-18, and ALPO-31) is performed using first principles methods based on periodic density functional theory. These materials are characterized by the presence of large empty structural channels expanding along several different crystallographic directions. The elasticity tensors, mechanical properties, and compressibility functions of these materials are determined and analyzed. All of these materials have a common elastic behavior and share many mechanical properties. They are largely incompressible at zero pressure, the compressibilities along the three crystallographic directions being frequently smaller than 5 TPa 1 . Notably, the compressibilities of ALPO-5 and ALPO-31 along the three principal directions are smaller than this threshold. Likewise, the compressibilities of ALPO-18 along two directions are smaller than 5 TPa 1 . All of the considered materials are shear resistant and ductile due to the large bulk to shear moduli ratio. Furthermore, all of these materials have very small mechanical anisotropies. ALPO-18 exhibits the negative linear compressibility (NLC) phenomenon for external pressures in the range P = 1.21 to P = 2.70 GPa. The minimum value of the compressibility along the [1 0 0] direction, k a = 30.9 TPa 1 , is encountered for P = 2.04 GPa. The NLC effect in this material can be rationalized using the empty channel structural mechanism. The effect of water molecule adsorption in the channels of ALPO-18 is assessed by studying the hydrated ALPO-18 material (ALPO-18W). ALPO-18W is much more compressible and less ductile than ALPO-18 and does not present NLC effects. Finally, the effect of aging and pressure polymorphism in the mechanical properties of VPI-5 and ALPO-5 is studied. As hydration, aging leads to significant variations in the elastic properties of VPI-5 and increases substantially its compressibility. For ALPO-5, pressure polymorphism has a small impact in its elasticity at zero pressure but a large influence at high pressure.

Graphical Abstract

1. Introduction

Aluminophosphate (ALPO) compounds are important synthetic microporous materials whose structure is characterized by the presence of large structural channels expanding along several different crystallographic directions [1]. Due to their high surface area and pore volume, ALPO materials have been employed in a wide range of important applications [1]. However, the full tensorial elasticity of these compounds, determining their behavior under stress and their mechanical performance in the applications, have surprisingly not been studied. A detailed mechanical characterization of a representative set of ALPO materials is performed in the present work. The set of materials considered includes VPI-5 [2,3,4,5], ALPO-8 [6,7], ALPO-5, [8,9] ALPO-18 [10,11], and ALPO-31 [12]. Specifically, the ALPO materials studied in this paper have been extensively used for compound adsorption [13,14,15,16,17,18,19,20,21,22,23], encapsulation [13,18,24,25,26] and separation [27,28,29], synthesis and catalysis [30,31,32,33,34,35,36,37,38,39,40,41,42,43,44], sensor fabrication [45], and energy storage and heat transformation applications [46,47,48,49]. The most widely used ALPO material is ALPO-5, a highly versatile material [14,18,19,23,24,25,26,30,31,32,33,34,35,36,40,45] which is very well-known for its applications in the encapsulation of laser dyes [18,24] and as a host for the formation of single-walled carbon nanotubes [30,31,32,33,34,35,36]. The range of applicability of aluminophosphate microporous solids is being greatly extended from the well-known fields of absorption, catalysis, and ion exchange. Important applications investigate their potential use as catalysts in the manufacture of chemicals, for increasingly selective acid and oxidation catalysis, in biomedical applications as drug delivery, in devices that require special electronic or optical properties, and as advanced functional materials [1].
The mechanical properties of natural and synthetic materials exhibiting high porosity are extremely interesting from the point of view of applications, three important examples being those of zeolites [50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67], metal organic frameworks (MOF) [68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103], and 3D carbon (3DC) materials [104,105,106,107,108,109,110,111,112,113,114,115,116]. Since their mechanical properties vary substantially from one compound to other, they have a large spectrum of applications. Interesting enhanced mechanical characteristics found for these materials include increased flexibility [21,29,32,34,37,57,58,59,60,61,62] and responsible behavior upon guest molecule adsorption and under temperature and pressure perturbations [56,57,67,69,76,98,99,110,111,117,118,119,120,121,122,123,124,125,126,127,128]. These properties make these materials exceptionally appropriate for mechanical damping and mechanical energy storage [55,84,85,86,91,114,115,120], compound adsorption [70,98,99,110,111,117,118,119,120,121,122], separation and storage [113,123,124], drug-delivery [125], and sensing applications [45,116,120]. Furthermore, negative [64,71,73,74,77,78,79,92,93,94,95,96,97,98,99,100,101] or zero [55] linear compressibility and negative [51,54,65,66,74,83,100] or zero [100] Poisson’s ratio phenomena have been encountered for many of these materials. The negative-linear compressibility (NLC) [129,130,131], zero linear compressibility (ZLC) [132,133,134,135] and negative Poisson’s ratio phenomena (NPR) [136,137,138,139] have multiple potential applications [129,131,133,134,135,139,140,141,142,143,144,145,146,147,148], the most well-known being the development of ultrasensitive pressure sensors and actuators [129,139,140]. The relevance of the research on the behavior of highly porous materials under the effect of pressure in materials science has further increased since the application of high pressures to this type of materials has allowed for the design of new advanced functional materials, thus expanding the limits of conventional synthetic chemistry. Interesting amorphous materials, glasses, and crystalline compounds have been obtained from pressure induced phase transitions and structural rearrangements [149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164].
The enormous amount of research devoted to the synthesis and characterization of the microporous aluminophosphate materials by X-ray, neutron and synchrotron diffraction, NMR, and spectroscopic techniques and to the study of their thermal stability, thermodynamic, and general physical and chemical properties is strongly in contrast with the poor current mechanical characterization of these materials [165,166,167,168,169,170,171,172]. As far as we know, only the compressibility and phase transitions induced by the application of isotropic pressures in ALPO-5 [167], VPI-5 [168,169,170] and ALPO-17 [171,172] have been studied. Therefore, the number of ALPO materials whose elasticity has been investigated is very small and the only mechanical property considered is the compressibility. Although these studies provided interesting results, the understanding of the mechanics of these materials is strongly deficient. The importance of the knowledge of tensorial elastic properties of porous materials is well-known in Earth sciences, where the effect of porosity on the elasticity of rocks and minerals is an important topic of direct interest for applications in petroleum exploration and production [173,174,175,176,177,178], clay swelling and radioactive nuclear waste storage [179,180,181,182,183,184,185,186,187,188], and for characterizing mechanical instabilities as shear failures [183,184,185,186,187,188]. Knowledge of the full elastic tensor is fundamental for assessing the mechanical stability of a given material or structure [189,190,191,192]. In materials science and technology, mechanical engineering, the pharmaceutical industry, and many other branches of science [193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240], the knowledge of the anisotropic mechanical properties, including the Young and shear moduli, Poisson’s ratio, and elastic anisotropy measures, is fundamental for the assessment of many important properties of materials, composites and products as their hardness [193,194,195,196,197] ductility [198,199,200,201,202,203,204], anisotropy [205,206,207,208,209,210,211,212] shear strength [74,82,215,216,217,225,226,227] compaction, tableting and milling performance [88,217,218,221,222,227,234,235], durability and degradability [207,208,238,239], and for the material screening and selection [217,218,219,220,221,222,223,228,239,240]. As shown in this paper, the mechanical properties of different polymorphic forms of the same ALPO material vary significantly. The knowledge of the elastic properties of the different polymorphic forms of a given material is important for selecting the polymorph with the best performance for a given application [217,218,219,220,221,222,223,226,227,228,229,230,231,234,236,237]. For microporous materials, the importance of the analysis of the full tensorial elasticity has been highlighted by the study of the mechanical stability and mechanical properties of selected zeolites [51,54,60,61,64,65,66,241,242,243] and MOFs [73,74,80,81,83,90,100,101]. A wide range of experimental methods, complementary to the use of diamond anvil cells (DAC) under hydrostatic pressures, such as elastic wave propagation measurements [244,245,246,247], the inelastic X-ray scattering [242,243,248,249,250,251], Brillouin scattering [53,54,252,253], nanoindentation [68,80,81,83,220,224,229,230,231,254,255,256,257,258,259] and ellipsometry [81,260] techniques, impedance spectroscopy [261,262,263,264], real contact area measurements [265], or methods based on DAC with non-hydrostatic pressures [168,264,265,266,267,268,269,270] could be used to provide a more complete set of mechanical properties of ALPO materials. These techniques have already been used in some cases for zeolites and MOFs [53,54,68,80,81,241,254]. Alternatively, force field [51,52,65,90,243,271,272,273] or accurate first principles methods [60,61,64,65,66,68,71,87,100,101,264,274,275,276] could be employed. First-principles calculations based on density functional theory represent a versatile, efficient, and accurate method for calculating the mechanical properties of materials [60,61,100,101,264,270,277].
In the present work, first principles solid-state methodology is employed to determine the mechanical properties and compressibility functions of ALPO materials. From a physical point of view, the increase in porosity in a given material should reduce its bulk and shear moduli and isotropy, making the material weaker, more compressible, and less shear resistant. However, these expectations are largely unsatisfied in many cases and the mechanics of porous materials is highly variable. In the present work, anhydrous ALPO compounds are surprisingly found to be very incompressible, ductile, and isotropic materials. The ductility of these materials results from the high bulk to shear moduli ratios [198,200], even although the shear moduli are quite significant. The incompressibility, ductility, and isotropy of the anhydrous ALPO materials could make them advantageous in many practical applications with respect to other more compressible, brittle, or anisotropic materials due to their larger stability and controllability under operation. However, these characteristics are shown to be highly dependent on the pressure and humidity conditions and material aging. In recent works [100,101], the negative linear compressibility phenomenon was found in some microporous metal organic frameworks due to the presence of empty structural channels in their crystal structures. In these materials, the widening of the channels along the direction of minimum compressibility under the effect of pressure leads to an increase of the dimension of the crystal along this direction. This effect, however, could or could not be found in multi-channel materials due to compensation effects associated with the distinct compressional behavior of the different channels along the several crystallographic directions. However, the present results show that the NLC phenomenon is indeed observed in ALPO-18 material in which the compressional behavior of its largest channels is dominant and leads to a substantial NLC effect along the [1 0 0] crystallographic direction.

2. Methods

The mechanical properties and compressibility functions of the ALPO materials studied [2,3,4,5,6,7,8,9,10,11,12], were determined using first principles solid-state methods based on periodic density functional theory, employing plane wave basis sets and pseudopotential functions for the description of the inner atomic electrons [278]. All of the computational works were carried out utilizing the Cambridge Serial Total Energy Program (CASTEP) computer code [279] interfaced with the Materials Studio program suite [280]. The main energy-density functional used in this work was the specialized version of Perdew-Becke-Ernzerhof (PBE) functional [281] for solid materials, PBEsol [282]. The PBE functional supplemented with D2 Grimme empirical correction [283] was also employed to assess the importance of the dispersion interactions in the structures and elastic properties of the selected ALPOs and for the description of ALPO-18 hydrated material (ALPO-18W) [11]. The good performance of PBEsol functional for anhydrous materials is well documented [284,285,286,287,288,289,290]. The pseudopotential functions utilized in this work were norm-conserving pseudopotentials [291] provided by the CASTEP program. Further computational details and material data of the compounds studied are given in Table S1 of the Supplementary Materials (SM). The Broyden–Fletcher–Goldfarb–Shanno (BFGS) method [292] was employed to optimize completely all of the atomic positions and unit cell parameters of the studied materials. The crystal structure optimizations were carried out, using the experimental structures as starting point, with stringent convergence criteria. The thresholds in the variation of the total energy, maximum atomic force, maximum atomic displacement and maximum stress are 2.5 × 10 6 eV/atom, 0.005 eV/Å, 2.5 × 10 4 Å, and 0.0025 GPa, respectively. The software REFLEX included in the Materials Studio program package [279] was used to derive the X-ray powder diffraction patterns [293] of the selected ALPOs from the experimental and computed crystal structures using CuK α radiation (λ = 1.540598 Å).
The elastic constants, the matrix elements of the stiffness tensor [294], needed to calculate the mechanical properties of the materials considered and to study the mechanical stability of their crystal structures, were determined from stress-strain relationships using the finite deformation method (FDM) [295]. The theory of elasticity in solid state physics, is a mathematically well-defined theory which relies on the quantum mechanical definition of the stress tensor [296]. In the FDM, the individual elastic constants are determined from the stress tensors resulting from the response of the material to finite programmed symmetry-adapted strains [295]. The FDM has been satisfactorily utilized to describe the elastic response of many solid materials, including uranyl-containing compounds [286,288,289,297,298,299,300,301,302,303,304,305,306], organic crystals [307,308,309,310,311], and metal-organic compounds [100,101,290,312,313,314]. The reliability of this method has been confirmed recently by the experimental verification of the negative area compressibility effect in silver oxalate [314] which was predicted using the first principles methodology [313]. The full set of elastic constants can also be computed using other methods [315,316,317,318,319,320,321,322,323] such as density functional perturbation theory (DFPT) [315,316,317,318] or the well-known strain or stress fluctuation formalisms from both Monte Carlo and molecular dynamics simulations [319,320,321,322,323]. The ElAM computer program [324] was employed to generate the tridimensional representations of the mechanical properties as a function of the orientation of the applied strain.

3. Results and Discussion

3.1. Crystal Structures

The computed crystal structures of VPI-5, ALPO-8 and ALPO-31 are shown in Figure 1 and the structures of ALPO-18 and ALPO-18W are displayed in Figure 2. The structure of VPI-5 [2,3,4,5], consists of alternating AlO 4 and PO 4 tetrahedra [325] which share vertices to form four-and six-membered rings (4-MR and 6-MR). The four-and six-membered rings are linked together in the (0 0 1) plane to form exceptionally large hexagonal 18-membered rings (18-MR) with horizontal and vertical dimensions ω = 13.84 Å and h = 15.84 Å measured as the distances between opposite oxygen atoms (see Figure 1A) These rings delimitate empty structural channels expanding along [0 0 1], the 18-MR channels having hexagonal cross-sections. The corresponding horizontal and vertical pore apertures for the 18-MR channels can obtained by discounting two times the oxygen ionic radii (R(O) = 1.33 Å), p ω = 11.2 Å, and p h = 13.2 Å.
The crystal structure of ALPO-8 [6,7], is formed by alternating corner sharing AlO 4 and PO 4 tetrahedra forming 4-, 6- and 14-membered rings as shown in Figure 1B. The 14-membered rings have a distorted rectangular shape and have horizontal and vertical dimensions of ω = 11.40 Å and h = 10.55 Å (pore apertures of 8.7 and 7.9 Å ). VPI-5 and ALPO-8 belong to a group of molecular sieves with framework structures containing very large channels. The term “extra-large pore” material was coined by Davis et al. [326] to specify microporous materials having channels with apertures larger than those of the classical 12-MR rings. ALPO-31 [12] contains one-dimensional channels circumscribed by 4-, 6-and 12-tetrahedrally coordinated atoms, that is, 4-, 6-and 12-membered rings in [0 0 1] plane (Figure 1C), the latter being nearly circular and having a diameter of d = 7.9 Å (pore aperture p d = 5.2 Å ). ALPO-5,8-9 as ALPO-31, contains 4-, 6-and 12-MR channels expanding along [0 0 1], the latter having a nearly circular cross section with a diameter of d = 10.1 Å (pore aperture p d = 7.4 Å ).
As shown in Figure 2A, the structure of ALPO-18 [10,11] exhibits 4-and 8-membered rings in (0 0 1) plane. The 8-membered nearly circular rings, have a diameter of d = 6.5 Å . A perspective view of the 8-MR channels is plotted in Figure 2B. The ALPO materials have commonly also channels expanding along several directions. While these channels have generally smaller apertures, the case of ALPO-18 is noticeable since also large channels are observed when the material is viewed from other directions. As can be observed in the second and third subfigures of Figure 2A, large 8-membered rings are also observed when the structure of ALPO-18 is seen from [1 0 0] and [1 1 0] directions. Figure 2C shows the crystal structure of hydrated ALPO-18 (ALPO-18W). Two water molecules per formula unit are adsorbed within the channels of ALPO-18. The structure of ALPO-18 changes significantly upon hydration. The space symmetry of this material changes from monoclinic ( C 2 / c ) to triclinic ( P 1 ). In fact, some of the aluminum atoms change their coordination environment from tetrahedral to octahedral to account for the presence of additional water molecules [11].
The computed lattice parameters for VPI-5, ALPO-8, ALPO-5, ALPO-18, ALPO-18W, and ALPO-31 along with the experimental parameters are reported in Table 1. The average difference of the computed and experimental unit cell volumes is quite good, 2.5 and 2.3%, for the PBE and PBEsol functionals. The impact of introduction of dispersion corrections in these materials is relatively small and the average difference of the computed and experimental unit cell volumes is reduced by only 0.3%. Since the improvement due to the inclusion of dispersion corrections was small, the PBEsol functional was used for all anhydrous materials to retain the ab initio character of the computations. Additional details about the impact of including dispersion interactions in the calculations will be given in Section 3.6.3. Dispersion corrections were only included for ALPO-18W since, as it is well-known [100,288,289,290], they significantly improve the hydrogen bond geometries in the structures of hydrated materials. The X-ray diffraction patterns of VPI-5, ALPO-8, ALPO-5, ALPO-18, and ALPO-31, generated from the computed and experimental structures [4,6,9,10,12], are compared in Figure 3. The agreement is excellent. A detailed comparison of the positions of the main reflections in the X-ray diffraction patterns for these ALPO materials is provided in Tables S2–S6 of the SM. Similarly, Tables S7–S11 of the SM provide a comparison of the computed and experimental interatomic distances in the crystal structures of these materials. The computed PO and AlO average distances are 1.52 and 1.72 Å , respectively, which are in good agreement with the experimental values of 1.51 and 1.70 Å .

3.2. Stiffness Tensors and Mechanical Stability

The computed stiffness tensors of VPI-5, ALPO-8, ALPO-5, ALPO-18 and ALPO-31 are provided in Table 2. The number of non-vanishing and non-equivalent elements in the matrix representation of the symmetric stiffness tensor depends on the space symmetry of the corresponding crystal structure [192,325]. The P 6 3 c m and P 6 c c structures of VPI-5 and ALPO-5 are hexagonal and have nine non-vanishing elastic constants in their stiffness matrices, five of which are non-equivalent ( C 11 , C 33 , C 44 , C 12 ,   C 13 ). ALPO-8 is orthorhombic ( C m c 2 1 ) and, therefore, its stiffness tensor has nine non-vanishing elements all of which are non-equivalent. The number of nonvanishing elastic constants for the monoclinic structure of ALPO-18 ( C 2 / c ), increases to thirteen due to its lower symmetry. For ALPO-31 (trigonal, R 3 ¯ h ) there are fifteen non-vanishing elastic constants, seven of which are non-equivalent ( C 11 , C 33 , C 44 , C 12 ,   C 13 , C 14 , C 15 ). A crystal structure is mechanically stable, if an only if, the Born mechanical stability conditions are fulfilled [189,190,191,192]. The generic Born mechanical stability condition can be written in mathematical form as an algebraic condition on the eigenvalues of the matrix representation of the stiffness tensor: the elastic matrix must be positive definite, that is, all its eigenvalues must be greater than zero [192]. A numerical diagonalization of the stiffness tensors of all of the ALPO materials was carried out. Since all of the elastic matrix eigenvalues for all materials were positive, they are mechanically stable.

3.3. Mechanical Properties

The computed stiffness tensors were employed to determine the mechanical properties of polycrystalline aggregates of the considered ALPO materials using the Voigt [327], Reuss [328] and Hill [329] schemes. The corresponding formulas for these approximations may be found in several sources (for example, Weck et al. [330]). The results obtained using the three approaches were quite similar for all materials. As was found in many previous papers [100,101,286,288,290,298,299,300,301,302,303,304,305,306,307,308,309,310,311,312,313,331], the values of the calculated bulk moduli in the Reuss approximation gave the best approximation to the single crystal bulk moduli of these materials. The mechanical properties for VPI-5, ALPO-8, ALPO-5, ALPO-18 and ALPO-31 in the Reuss approximation are given in Table 3. As can be observed, these five materials are characterized by very small elastic anisotropies, since the universal anisotropy indices ( A U ) [212] are 0.66, 0.83, 0.49, 0.46, and 0.36, respectively. The similarity of the computed mechanical properties in the Voigt, Reuss, and Hill approximations is a direct consequence of the low mechanical anisotropy. For crystalline systems with strong anisotropy, large differences should be expected [286,330,331]. The single-crystal bulk moduli, B s c , reported in Table 3, were not determined from fits of calculated pressure-volume data to high-order Birch-Murnaghan equations of state [332], as was customary in previous works [100,101,286,288,290,298,299,300,301,302,303,304,305,306,307,308,309,310,311,312,313,331], since the fitting parameters were rather dependent on the range of pressure employed. Similar difficulties were found in the study of the compressibility of several minerals [333]. The computed bulk moduli were derived from the compressibility functions reported in Section 3.5 and Section 3.6, which were determined from accurate six order polynomial fits to pressure-volume data in the pressure range from 0 to 5 GPa.
The computed values of the bulk modulus, B , the inverse of compressibility, are significant: 60.49, 68.27, 88.22, 73.56, and 80.17 GPa for VPI-5, ALPO-8, ALPO-5, ALPO-18 and ALPO-31, respectively. This means that these materials are very incompressible under the effect of external isotropic pressures. ALPOs are also resistant with respect to the application to uniaxial pressures since the values of the Young’s moduli, 60.46, 68.18, 81.43, 53.35, 71.39 GPa, respectively, are substantial. The shear modulus, G , represents the resistance to plastic deformation. The calculated values of G , 22.67, 25.56, 30.24, 19.34, and 26.41 GPa, respectively, are quite large is comparison with the values found for other microporous materials [53,54,64,66,72,73,74,75,80,81,82,100]. Therefore, ALPOs are resistant with respect to the application of external uniform and uniaxial pressures and shear stresses.
The Cauchy pressure term, defined in terms of the elastic constants as C P = ( C 11 C 44 ) , was proposed by Pettifor [199] as an indicator of the angular character of atomic bonding. The values of C P are positive and large, 22.49, 29.07, 41.56, 54.63, and 32.32 GPa, for VPI-5, ALPO-8, ALPO-5, ALPO-18, and ALPO-31, respectively, reflecting a largely angular bonding in these materials. The value of the Cauchy pressure term is particularly large for ALPO-18. C P is also related with the brittle/ductile character of crystals [198,200,201,202,203,204]. Large values of C P are associated with highly ductile materials. The ductility index, D = B / G , was proposed by Pugh [198], as a standard measure of the ductility of a material. A value of D = 1.75, separates the brittle and ductile materials [201,208]. All of the computed values of D , 2.66, 2.67, 2.92, 3.80, and 3.04, for VPI-5, ALPO-8, ALPO-5, ALPO-18, and ALPO-31, respectively, are much larger than 1.75. Therefore, the five materials are ductile. An improved ductility criterium has been provided recently by Niu et al. [200]. In their work, these authors noticed that the intrinsic ductility index, defined as the ratio of the Cauchi pressure term to the Young’s modulus, D I = ( C 11 C 44 ) / E , is strongly correlated hyperbolically with the Pugh’s ratio. As shown in Table 3, the values obtained for the intrinsic ductility index for the ALPO materials considered ranges from 1.02 to 0.18 and are in the same range as that for common metals [200]. The value of D I for ALPO-18, 1.02, is close to that of Pt (0.98 ± 0.01) or Nb (1.00 ± 0.01). For ALPO-5, D I = 0.51, coincides with that of K (0.51 ± 0.01). The intrinsic ductility indices of ALPO-31 and ALPO-8, 0.45, and 0.43, respectively, are close to the ductility for Al (0.44 ± 0.04). Finally, the value of D I for VPI-5, 0.37, is near to that of Cu (0.38 ± 0.04).
The Vickers hardness ( H ) measures the resistance of a given material to indentation. A series of representative values of H for interesting materials may be obtained from several published papers [193,194,195,196,197]. As a reference, talc and halite ( H = 0.26 and 0.30) are very weak, calcite and sphalerite ( H = 1.5 and 1.8) are weak, fluorite and apatite ( H = 3.0 and 5.1) have medium hardness, quartz and zirconia ( H = 11.1 and 13.0) are hard and corundum and diamond ( H = 21.5 and 96.0) are very hard. The computed values of H for VPI-5, ALPO-8, ALPO-5, ALPO-18, and ALPO-31 are 0.94, 1.22, 1.13, 0.63, and 0.70, respectively, which correspond to relatively weak materials.

3.4. Mechanical Properties as a Function of the Orientation of the Applied Strain

In the previous subsection, a general view of the elasticity of the ALPO materials was achieved and average values of the elastic moduli, Poisson’s ratios, and ductility, hardness, and elastic anisotropy indices were reported. A more detailed understanding of the elasticity of these materials is provided by the analysis of the variation of the mechanical properties with the strain orientation. Three dimensional representations of the dependence of the elastic moduli and Poisson’s ratios for VPI-5, ALPO-8, ALPO-5, ALPO-18 and ALPO-31 with respect to the direction of the applied strain are displayed in Figure 4, Figure 5, Figure 6, Figure 7 and Figure 8, respectively. These figures explain the low elastic anisotropy of these materials since all elastic moduli have a smooth variation with respect to the direction of the applied strain. The elastic properties of VPI-5 and ALPO-5 (with hexagonal space symmetries) show a nice orientational dependence which is axially symmetric around z axis.
Figure 4, Figure 5, Figure 6, Figure 7 and Figure 8 ( k and E ) show that there are not preferred directions along which the compressibility and Young’s modulus are negative, near to zero or particularly small. Therefore, ALPOs are very incompressible in all directions and there are not clear directions for material fracture when isotropic or uniaxial pressures are applied. Likewise, these figures show a smooth directional dependence of the shear modulus ( E ), without special directions associated with small values of this property (see the projections of the surfaces of minimum shear modulus). Therefore, there are not crystallographic planes along which shear failure can be predicted. The presence of shear slippages imposed serious limitations to the mechanical properties of some microporous materials including MOFs [72,75,80,81,82] and carbon nanotube composites [215,216]. No signs of auxeticity (negative Poisson’s ratios [334]) were found for any of the materials investigated since the Poisson’s ratios are always positive for all strain directions. This is in contrast with the elasticity of other microporous materials (for example for zeolites), for which negative or zero Poisson’s ratios were frequently encountered [51,52,53,54,64,65,66,67].

3.5. Compressibility Functions

The crystal structures of VPI-5, ALPO-8, ALPO-5, ALPO-18, and ALPO-31 were fully optimized under different external isotropic pressures in the pressure range from −0.5 to 5.0 GPa. The computed unit cell volumes and lattice parameters for VPI-5, ALPO-8, ALPO-5, ALPO-31 and ALPO-18 at different pressures are plotted in Figure 9, Figure 10, Figure 11, Figure 12 and Figure 13, respectively. The calculated volumetric compressibilities, k V = 1 / V · ( V / P ) P , and the linear compressibilities, k l = 1 / l · ( l / P ) P ( l = a ,   b ,   c ) along the three crystallographic directions between P = 0.0 and P = 4.0 GPa are also displayed in these figures. The values of the calculated unit cell volumes, lattice parameters and compressibilities are given in Tables S12–S21. The structure of ALPO-18 was also optimized under the effect of different uniaxial pressures (see Figure 14 and Tables S22 and S23). The calculated compressibilities of these materials at zero pressure are collected in Table 4. From this table, it follows that the volumetric compressibilities at zero pressure are very small, the most compressible material being VPI-5 ( k V = 16.46 TPa 1 ) and the less compressible one being ALPO-31 ( k V = 11.06 TPa 1 ).
The linear compressibilities along the different directions are frequently smaller than 5 TPa 1 . For ALPO-5 and ALPO-31, the three linear compressibilities are smaller than this threshold. The same occurs for two linear compressibilities of ALPO-18, although the compressibility along b direction is also very near to this limit ( k b = 5.17 TPa 1 ). For VPI-5 and ALPO-8, only the compressibilities along c direction satisfy k c < 5 TPa 1 . However, the value of k c for ALPO-8 is the lowest linear compressibility found for all of the ALPO materials considered, ( k c = 1.81 TPa 1 ). The criterium usually used for zero linear compressibility (ZLC) [102,103,132,133,134,135] is that the absolute value of the linear compressibility along a certain direction is smaller than 1.0 TPa 1 , |   k l   | 1.0 TPa 1 [133]. While this criterium is not met for ALPOs, the presence of three simultaneously small linear compressibilities is very infrequent. The term near zero tridimensional linear compressibility (NZTLC) is proposed for materials satisfying, 1 |   k l   | 5.0 for l = a ,   b ,   c . ALPO-5, ALPO-31 and, in practical terms, ALPO-18 are NZTLC materials at zero pressure.
As shown in Figure 9, the compressibilities of VPI-5 along the three crystallographic directions are lower than 10 TPa 1 for the full range of pressure considered. The compressibility along the c direction is smaller than 5 TPa 1 from 0 to 4 GPa except for applied pressures near 4.0 GPa. Therefore, as shown in Figure S1, the isotropic compression of VPI-5, leads to very small changes of its structure. For ALPO-8 (Figure 10), although the linear compressibility along a direction remains small from 0 to 4 GPa and attains a minimum near P = 2.75 GPa ( k a = 3.4 TPa 1 ), k b and k c increase rapidly and reach maxima near P = 2.5 GPa. Consequently, the volumetric compressibility increases from 0 to 2.5 GPa and then decreases up to 4 GPa. Since the compressibility along c direction is very small at zero pressure, and it is a strongly decreasing function as the pressure diminishes, the presence of negative values of k a under tension (negative pressure) is highly probable. For ALPO-5 (Figure 11), as for VPI-5, the linear compressibilities remain small in the range from 0 to 4.0 GPa. However, for ALPO-31 (Figure 12), as for ALPO-8, the compressibilities increase largely as the pressure increases and the volumetric compressibility reach a maximum near P = 2.0 GPa. The behavior of ALPO-18 under pressure is extremely anomalous and is studied in the next Subsection.

3.6. Negative Linear Compressibility (NLC) in ALPO-18

3.6.1. Isotropic Negative Linear Compressibility (INLC)

The three lattice parameters of VPI-5, ALPO-8, ALPO-5, and ALPO-31 decrease invariably under isotropic compression. However, as can be observed in Figure 13B, the a lattice parameter of ALPO-18 increases sharply from P = 1.21 to P = 2.70 GPa. Therefore ALPO-18 exhibits the isotropic negative linear compressibility (INLC) phenomenon [129,130,131] in this pressure range. The minimum value of the compressibility along the a direction is encountered at P = 2.04 GPa, k a = 30.9 TPa 1 .
The INLC effect in ALPO-18 can be rationalized in terms of the empty channel structural mechanism [100,335]. The deformation of the crystal structure of ALPO-18 induced by the application of increasing isotropic pressures is illustrated in Figure 15. In this figure, the optimized crystal structures at five different pressures, P = 1.00, 1.75, 2.00 2.25, and 2.50 GPa, are displayed.
As can be observed, the width and height of main 8-MR channels expanding along [0 0 1] increase and decrease substantially under increasing pressure. The widening of the channels along [1 0 0], which coincides with the direction of minimum compressibility in ALPO-18 (see Figure 7), leads to an increase of the a lattice parameter and to the INLC effect in this material. The impact of the deformation of the channels expanding along [1 0 0] and [1 1 0] (Figure 2) in the dimensions of the crystal is much smaller, as shown in Figure S2. Therefore, the dominance of the deformation of the 8-MR channels expanding along [0 0 1] makes observable the NLC effect based in the empty structural mechanism in the multichannel ALPO-18 material.

3.6.2. Anisotropic Negative Volumetric Compressibility (ANVC) Effect

In previous works, the INLC effect due to the empty channel structural mechanism was observed to be accompanied by the anisotropic volumetric NLC effect (ANLC) [100,336], i.e., the increase of the volume of a material when an external anisotropic pressure is applied to it. This effect was discovered in 2015 by Baughman and Fonseca [336] in porous materials and, independently, by Colmenero in 2019 for non-porous materials as the cyclic oxocarbon acids [307], oxalic acid [308] and uranyl squarate monohydrate [311]. The unit cell volumes, lattice parameters and compressibilities of ALPO-18 under the effect of increasing uniaxial pressures along the direction of minimum compressibility, [1 0 0], are shown in Figure 14 and provided in Tables S22 and S23 of the SM. As can be appreciated, the unit cell volume increases under tension from P = −1.0 up to P = −0.20 GPa. Therefore, ALPO-18 exhibits the ANVC effect in this pressure range. The minimum value of the compressibility is found at P = −0.76 GPa, k V = −6.0 TPa 1 .

3.6.3. Effect of Dispersion Interactions in the NLC Effect of ALPO-18

In Section 3.1, the influence of dispersion interactions in the crystal structures of the considered ALPO materials was shown to be small. However, due to the relevance of the NLC phenomenon, the crystal structure of ALPO-18 was also completely optimized under the effect of different isotropic pressures using the PBE functional supplemented with Grimme’s dispersion corrections [283] The computed values of the a lattice parameter are compared with those obtained using the PBEsol functional in Figure 16. The results are quite simitar, thus confirming the NLC effect in ALPO-18 and the good performance of the PBEsol functional for the description of anhydrous materials [191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279,280,281,282,283,284]. Using the dispersion corrected treatment, ALPO-18 displays an even larger isotropic NLC effect from P = 0.61 GPa to P = 2.50 GPa. The minimum value of the compressibility along the [1 0 0] direction is k a = −38.7 TPa 1 at P = 1.74 GPa.

3.7. Effect of Hydration in the Mechanical Properties of ALPO-18

In this Section, the effect of the presence of water molecules adsorbed in the structural channels of ALPO-18 on the mechanical properties of this material is studied. This is relevant from the point of view of applications since if one desires to take advantage of the mechanical properties of ALPOs, such as their large incompressibility and ductility, the influence of water adsorption should be considered. If the impact in the elastic properties is large, hydration should be avoided as much as possible. The calculated lattice parameters of ALPO-18W are in given in Table 1. The computed unit cell volume differs from the experimental value [11] by only 0.5%. The computed stiffness tensor and mechanical properties of ALPO-18W are reported in Table 5 and Table 6, respectively, and the dependence of its mechanical properties on the orientation of the applied strain is shown in Figure S3. The unit cell volumes, lattice parameters and compressibilities of ALPO-18W under different isotropic pressures are shown in Figure S4 and given in Tables S24 and S25.
Since ALPO-18 is triclinic, all of the elements of the matrix representation of its elastic tensor are non-vanishing and non-equivalent. As with ALPO-18, ALPO-18W is characterized by large bulk, Young’s, and shear moduli. However, due to the adsorption of water molecules, the bulk and shear moduli of ALPO-18W become much smaller and larger, respectively, than those of ALPO-18. Consequently, although ALPO-18W is also ductile, the ductility index is smaller. Since the intrinsic ductility index ( D I ) is strongly correlated with Pugh’s ratio [190], its value is reduced from 1.02 to 0.18. Therefore, hydration makes this material more compressible and less ductile. The universal anisotropy index is very small, A U = 0.24, as with the other ALPO materials. As expected from the small elastic anisotropy index, the dependence of the mechanical properties on the direction of the applied strain is smooth. No preferred directions for fracture or shear failures nor negative Poisson’s ratios are observed.
In a recent work [100], a strong reduction of the NLC effect in titanium oxalate dihydrate was also found as a result of water molecule adsorption, although the NLC effect in this microporous metal organic framework does not disappear completely. In contrast, for some microporous zeolites and MOFs [56,57,98,99], water intrusion leads to a strong increase of the unit-cell volume and NLC effects. The strong influence of the presence of guest molecules in the structural channels of porous materials in their mechanical properties has been found by several research groups in previous works [87,88,337]. In fact, Terracina et al. [88], showed that the main source of structural instability in the MOF HKUST-1 during compaction was the presence water molecules adsorbed by the powdered samples and a new tableting method preserving the crystal structure and porosity of the pristine powders was reported. The influence of guest molecule adsorption in the elastic properties of microporous materials is highly dependent on the material under study and the type of interaction between the molecules with the walls of the channels and between the molecules themselves. The presence of water in contact with the material may increase the internal tensions, lead to phase transformations or even be the origin of crack propagation and fracture [238]. The important NLC effect induced by water or guest molecule intrusion should be distinguished from the conventional NLC phenomenon, encountered for ALPO-18 in this paper, due to the need of specifying the origin of pressure and the requirement of the description of the interaction of a variable number of guess molecules with the material for his theoretical study. It is a common belief that the absorption of water molecules in the channels of a microporous material should reduce its compressibility due to the stiffening of the structure due to increased density [338]. However, the opposite its true in ALPO-18. The counterintuitive softening upon adsorption of guest molecules in microporous materials was first observed by Mouhat et al. [339] and Canepa et al. [340]. It is difficult to find an explanation for the softening in ALPO-18 based on the changes in the chemical bonding due to the drastic geometric rearrangement occurring upon water absorption. This requires a further study which is out of the scope of the present work.
The great influence of water adsorption in the structure of ALPO-18, underlines the need of using non-hydrous pressure transmitting media (PTM) to measure their compressibilities or experimental methods which are not based in the use of DAC in order to study its full tensorial elasticity. The compressibility of hydrated ALPO-5 [165,166,167] material was measured experimentally using DACs and different pressure transmitting media. The measured compressibilities were highly dependent of the PTM used. Furthermore, the compressibilities measured using the same PTM vary significantly from one study to another. For example, the volumetric compressibilities of ALPO-5W using a 16:3:1 methanol-ethanol-water (MEW) mixture as PTM measured by Kim et al. [166] and Lotti et al. [167] were 19.8 and 45.0 TPa 1   (corresponding to measured bulk moduli of 50.5(7) and 22.2(9) GPa), respectively. For VPI-5W, Alabarse et al. [168], using silicone oil as PTM, obtained of volumetric compressibility of 41.2 Tpa 1 ( B = 24.3(5) Gpa). The present results for anhydrous VPI-5 and ALPO-5 and the experimental results for VPI-5W and ALPO-5W show that, as for ALPO-18, the influence of the presence of water in the channels of these materials in the elastic properties are substantial. Again, the use of complementary experimental methods not based in the use of DAC with a PTM for the study of the full tensorial elasticity of these materials is suggested.

3.8. Effect of Aging in the Elastic Properties of VPI-5

De Oñate Martinez et al. [5], noted that the space symmetry of VPI-5 depends on the method of preparation of this material and that aging also leads to space symmetry variations. Clearly, the origin of this effect is the small differences in the relative thermodynamic stabilities of the different crystal structures of ALPO materials [341,342]. Therefore, to assess the influence of aging, the monoclinic C 1 m 1 crystal structure reported by De Oñate Martinez et al. [5] obtained from an aged sample of anhydrous VPI-5 was employed in order to compute its mechanical properties. The computed lattice parameters are given in Table S26 of the SM and the calculated stiffness tensor and mechanical properties are given in Table 5 and Table 6, respectively. The dependence of its mechanical properties on the direction of the applied strain is shown in Figure S5. The bulk modulus for the monoclinic structure of VPI-5 diminishes significantly with respect to that for hexagonal VPI-5 (from 60.5 to 37.1 GPa) and, consequently, the ductility index is largely reduced (from 2.66 to 1.77). The elastic anisotropy in monoclinic VPI-5 ( A U = 0.24) is smaller than in hexagonal VPI-5 ( A U = 0.66). Although the directional dependence of the elastic properties for the monoclinic structure are significantly modified, the axial symmetry around z axis is conserved (Figure S5B) and no preferred directions for fracture and shear failure are observed. Therefore, aging in VPI-5, as hydration in ALPO-18, reduces its incompressibility and ductility. The computed unit-cell volume, lattice parameters and compressibilities of monoclinic VPI-5 as a function of the external isotropic pressure are shown in Figure S6 and given in Tables S27 and S28 of the SM. While the linear compressibilities of hexagonal VPI-5 remains small as pressure increases, the same is not true for the monoclinic structure.

3.9. Effect of Pressure Polymorphism

3.9.1. VPI-5

Since in this paper we are interested in studying the behavior of ALPO materials under the effect of pressure, the relative thermodynamic stability of the P 6 3 c m 4 and C 1 m 1 [5] structures of VPI-5 under pressure was investigated. Both crystal structures were fully optimized under the effect of different external isotropic pressures and the corresponding enthalpies were determined. As shown in Figure S7 of the SM, the X-ray diffraction patterns of both structures at zero pressure are remarkably similar. The positions of the main reflections in the X-ray diffraction patterns of both structures are reported in Tables S2 and S29, respectively. The computed unit cell volumes and enthalpies are compared in Figure 17. In this figure the volumes and enthalpies of the P 6 3 c m structure have been doubled since the unit-cell of the C 1 m 1 structure is twice as large as the hexagonal unit cell. As can be seen, while the enthalpies of both structures are very close at zero pressure, the monoclinic structure is increasingly more stable as pressure increases. The difference of the enthalpies of both polymorphs, 0.3 kJ per formula unit at zero pressure, becomes 28.6 kJ at P = 5.0 GPa. Therefore, the VPI-5 monoclinic polymorph appears to be significantly more stable than the hexagonal one at high pressure conditions. The transition pressure between these structures is conditioned by thermodynamic and kinetic considerations [343]. In the initial studies concerning the structures of ALPO materials [344], great effort was devoted to the identification of the symmetry of their structures. Present results show that the difficulty in the identification, is further complicated by pressure polymorphism. At the same time, the results point to a form of obtaining monoclinic VPI-5 by submitting hexagonal VPI-5 to high isotropic pressures. As was shown in Section 3.8, the hexagonal-monoclinic polymorphic transformation reduces the incompressibility and ductility of VPI-5 substantially. It should be noted that an additional monoclinic C 1 m 1 structure for VPI-5 under pressure was also recently obtained by Fabbiani et al. [345] with a volume four times that of the hexagonal structure and two times that of the monoclinic structure reported by De Oñate Martinez et al. [5]. As shown in Table S30, the X-ray diffraction pattern derived from the structure of De Oñate Martinez et al. [5] is nearly the same as that derived from this structure, the difference in the positions of the main reflections being lower than 0.1 ° . This structure was obtained under the effect of isotropic pressure, in agreement with the present results favoring the monoclinic structures under pressure.

3.9.2. ALPO-5

For ALPO-5, the relative thermodynamic stability of the P cc 2 [8] and P 6 c c [9] structures under pressure was investigated. Figure S8 of the SM shows the great similarity of X-ray diffraction patterns of both structures at zero pressure. The positions of the main reflections in the X-ray diffraction patterns of both structures are given in Tables S4 and S31, respectively. The unit cell volumes and enthalpies associated with both structures under the effect different external pressure are compared in Figure 18. The volumes and enthalpies of the P 6 c c structure were doubled in Figure 18 (the orthorhombic unit cell is twice as large as the hexagonal one). Both structures are nearly degenerate at zero pressure. However, the orthorhombic structure is increasingly more stable as the pressure increases. The enthalpy difference becomes 12.1 kJ per formula unit ( AlPO 4 ) at P = 5.0 GPa. The hexagonal-orthorhombic polymorphism in the anhydrous and hydrated forms of ALPO-5 is a long-standing problem [8,9,346,347,348,349,350,351,352,353,354,355,356]. The presence of one or another polymorph is not only dependent on the temperature but also on the method used for the synthesis of this compound [356]. The results obtained here show that the orthorhombic structure is the high-pressure polymorph. The orthorhombic structure should be obtainable submitting the hexagonal polymorph to high pressure. Similarly, the synthesis of ALPO-5 at sufficiently high pressure should favor the production of the orthorhombic form, independently of the synthetic method employed.
The computed lattice parameters of the P cc 2 structure [8] at zero pressure are given in Table S26 of the SM. The calculated stiffness tensor and mechanical properties are reported in Table 5 and Table 6, respectively. The dependence of its mechanical properties on the direction of the applied strain is shown in Figure S9. In contrast with the result obtained for VPI-5 in previous Section, the variation of the elastic properties for the orthorhombic structure with respect to those of the hexagonal one is exceedingly small. For example, the bulk modulus for the hexagonal polymorph, B = 88.2 GPa, becomes 88.6 GPa. The intrinsic ductility index, D I = 0.51, is unchanged. The directional dependence of the elastic properties for the two polymorphs is quite similar. The axial symmetry around the z axis for the hexagonal structure is slightly distorted in the orthorhombic polymorph (Figure S9B). Consequently, the impact of pressure polymorphism in the mechanical properties of ALPO-5 at zero pressure is very small. The computed unit-cell volumes, lattice parameters, and compressibilities of orthorhombic ALPO-5 as a function of the applied isotropic pressure are displayed in Figure S10 and given in Tables S32 and S33 of the SM. Although the influence of the pressure polymorphism in the elastic properties of ALPO-5 is small, the linear compressibilities along a and c directions and the volumetric compressibility of the orthorhombic polymorph are strongly affected by the increase of pressure. Only the linear compressibility b direction remains nearly constant under pressure with a value close to k b = 5 TPa 1 .

3.10. Comparison with Experimental Data

There are very few data to compare the results of the present paper with experimental data [165,166,167,168,169,170,171,172]. They are mostly for hydrated ALPO materials and limited to compressibility values measured using the DAC technique using a given PTM. There appears to exist a large difference between the experimental values of the compressibility measured using this technique and the theoretical results for empty porous structures. The same is true for the experimental results obtained with and without a PTM [150], using two different PTMs or from two different studies using the same PTM [166,167,168]. For different PTMs, involving different molecules, the collisions of molecules with the surfaces of the material considered produce different effects. In many cases, pressure induced transitions and pressure induced amorphizations (PIA) appear at very different pressures for different PTMs [166,167,168]. For instance, for ALPO-5W [166], a PIA was observed at 15.9 GPa using liquid nitrogen as PTM and at 8.5 GPa using silicone oil (non-pore-penetrating PTM). For dense crystal structures accurate values of the compressibilities are generally obtained [314]. The measurements performed for some ALPO materials indicate large compressibilities. For example, for a crystalline sample of dehydrated VPI-5, Alabarse et al. [169] using a DAC with silicone oil observed a pressure induced phase transition to ALPO-8 beginning at 0.8 GPa, which do not appear from the theoretical calculations, obtained a compressibility of k = 80.6 TPa 1 (bulk modulus B = 12.4(2) GPa) for VPI-5 from a fit to a to a second order Birch−Murnaghan (2-BM) equation of state (EOS) from pressure−volume data below 1.6 GPa and the same value for ALPO-8 from data below 3.4 GPa. For dehydrated ALPO-17, Alabarse et al. [171] using a DAC with silicone oil observed a pressure induced amorphization beginning at 1 GPa and obtained a compressibility of k = 32 TPa 1 (bulk modulus B = 31.2(5) GPa) from a fit to a to a 3-BM EOS from pressure−volume data below 3.1 GPa. The results obtained in this work showed that the compressibilities obtained from fits to a BM EOS are highly dependent on the pressure range used in the fits and that many pressure-volume data points should be used to obtain reliable compressibilities. Furthermore, the emergence of pressure induced amorphizations should influence substantially in the measured compressibilities due to the reduction of the volume involved in the pore collapse. NLC phenomena were also observed for ALPO materials in previous works. However, the NLC effects observed were much less significant than that found for ALPO-18 in this work. For ALPO-17, Alabarse et al. [171], found a small increase of the a lattice parameter near the PIA associated to the pore collapse. Likewise, for ALPO-5W, Kim et al. [166] found that a lattice expands at small pressures before it starts to contract using a MEW mixture as a PTM (probably due to molecule pore intrusion).
The results obtained using theoretical techniques are unique. The underlying reason for the different theoretical and experimental results based in DAC must be the different origin of pressure in these experimental techniques (collisional mechanism) and in the theoretical calculations. In the theoretical calculations, the pressure is defined in terms of the stress tensor and the elastic properties are determined from the stress tensor resulting from an applied strain and the action of molecules over the material is not invoked. The data presented in this paper are rigorous quantum mechanical results and were fully revised and carried out twice to check their reproducibility. Furthermore, the compressibilities obtained from the computed elastic tensors and fits of computed pressure-volume data are in excellent agreement. Thus, it is concluded that the data obtained from the theoretical calculations and experimental measurements using DAC correspond to different physical quantities for highly porous materials. To obtain a better agreement between theoretical and experimental results, either experimental measurements not based in DAC or quantum molecular dynamic calculations with specific solvents simulating the effect of the different PTMs should be carried out. The effect of temperature should also be investigated since the present theoretical results correspond to zero temperature.

4. Conclusions

In this work, the crystal structures, mechanical properties, and compressibility functions of five important anhydrous microporous aluminophosphate materials have been determined using first principles methods based on density functional theory. The calculated crystal structures and associated X-ray diffraction patterns are in good agreement with their experimental counterparts. The elastic tensors of all of these materials have been reported and the mechanical stability of their structures has been confirmed. A detailed mechanical characterization is performed, and a rich set of mechanical properties was derived. This set includes the bulk, shear and Young’s moduli, as well as ductility, hardness, and elastic anisotropy indices. The elastic behavior of the five materials shares many common mechanical properties such as high incompressibility, ductility, and low elastic anisotropy. Their intrinsic ductility indices are in the same range as that for common metals. The analysis of the dependence of the mechanical properties of ALPOs in the orientation of the applied strain, show that they are resistant with respect to the application of external isotropic and uniaxial pressures and shear stresses. A smooth directional dependence is found in all of the cases and no special directions for material fracture or shear instability are encountered. The only previous study from which some clue about the incompressibility of ALPO materials was found is the work by Polisi et al. [210] where the dehydration mechanism of hydrated ALPO-5 was studied. The small volume change of this material under the effect of temperature led these authors to state that ALPO-5 was one of the most rigid zeolite frameworks. While this finding concerns only the thermal behavior of hydrated ALPO-5, the present results show that high incompressibility is a general property of anhydrous ALPO materials under the effect of pressure.
The crystal structures of all of the materials were completely optimized under the effect of different isotropic pressures and the linear and volumetric compressibilities were determined. At zero pressure, the ALPO materials have small linear compressibilities along the three crystallographic directions. The tridimensional incompressibility of ALPO-5, ALPO-18 and ALPO-31 is notable since the compressibilities along the three principal directions are lower or close to 5 TPa 1 . The incompressibility ALPO-8 and ALPO-31 materials is lost at high pressures. ALPO-18 displays an extremely anomalous mechanical behavior at relatively low external pressures. It exhibits a large negative linear compressibility effect between P = 1.21 and P = 2.70 GPa. The minimum value of the compressibility along [1 0 0] direction, k a = 30.9 TPa 1 , is encountered at P = 2.04 GPa. The NLC phenomenon in this material can be rationalized using the empty channel structural mechanism. The width and height of main 8-MR channels expanding along [0 0 1] increase and decrease substantially under increasing pressure. The widening of the channels along [1 0 0], coinciding with the direction of minimum compressibility in ALPO-18, leads to the increase of the a lattice parameter and to the NLC effect in this material. Furthermore, ALPO-18 exhibits the anisotropic negative volumetric compressibility effect for uniaxial pressures applied along the [1 0 0] direction.
The effect of water molecule adsorption in the channels of ALPO-18 in its elastic properties is assessed by studying the mechanical behavior of the hydrated ALPO-18 material (ALPO-18W). ALPO-18W is much more compressible and less ductile than ALPO-18 and does not present the NLC effect found in ALPO-18. The effect of aging and pressure polymorphism in the mechanical properties of VPI-5 and ALPO-5 is also studied. As hydration, aging and pressure polymorphism leads to significant variations in the elastic properties of VPI-5 and reduces its incompressibility and ductility. For ALPO-5, pressure polymorphism has only a small impact in its elasticity at zero pressure but a large influence at high pressure.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/solids3030032/s1, The Supplementary Information associated with this article contains: (a) material data and calculation parameters; (b) interatomic distances and angles; (c) positions of the main reflections in the X-ray diffraction patterns; (d) unit-cell volumes, lattice parameters and compressibilities under isotropic and anisotropic pressures.

Author Contributions

Conceptualization: F.C.; investigation: F.C., Á.L. and V.T.; writing—original draft: F.C.; writing—review and editing: F.C., Á.L. and V.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Ministry of Science, Innovation and Universities (project PGC2018-094814-B-C21). One of the authors (VT) was supported by the Ministry of Science, Innovation and Universities within the Project FIS2016-77726-C3-1-P.

Data Availability Statement

Not available.

Acknowledgments

Supercomputer time by the CTI-CSIC center is gratefully acknowledged. We want to thank A.M. Fernández for reading the manuscript and many helpful comments.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationship that could have appeared to influence the work reported in this paper.

References

  1. Wright, P.A. Microporous Framework Solids; RSC Publishing: Cambridge, UK, 2008. [Google Scholar]
  2. Davis, M.E.; Garces, J.; Saldarriaga, C.; Crowder, C. A molecular sieve with eighteen-membered rings. Nature 1988, 331, 698–699. [Google Scholar] [CrossRef]
  3. McCusker, L.B.; Baerlocher, C.; Jahn, E.; Bülow, M. The Triple Helix inside the Large-Pore Aluminophosphate Molecular Sieve VPI-5. Zeolites 1991, 11, 308–313. [Google Scholar] [CrossRef]
  4. Poojary, D.M.; Perez, D.M.; Clearfield, J.O. Crystal Structures of Dehydrated VPI-5 and H1 Aluminum Phosphates from X-ray Powder Data. J. Phys. Chem. C 1992, 96, 7709–7714. [Google Scholar] [CrossRef]
  5. De Oñate Martinez, J.; McCusker, L.B.; Baerlocher, C. Characterization and Structural Analysis of Differently Pre-pared Samples of Dehydrated VPI-5. Microporous Mesoporous Mater. 2000, 34, 99–113. [Google Scholar] [CrossRef]
  6. Dessau, R.M.; Schlenker, J.L.; Higgins, J.B. Structure Determination and Rietveld Refinement of Aluminophosphate Molecular Sieve AlPO4-8. Zeolites 1990, 10, 522–524. [Google Scholar] [CrossRef]
  7. Richardson, J.W.; Vogt, E.T.C. Structure determination and rietveld refinement of aluminophosphate molecular sieve AIPO4-8. Zeolites 1992, 12, 13–19. [Google Scholar] [CrossRef]
  8. Mora, A.J.; Fitch, A.N.; Cole, M.; Goyal, R.; Jones, R.H.; Jobic, H.C.; Carr, S.W. The Structure of the Calcined Aluminophosphate AlPO4-5 Determined by High Resolution X-Ray and Neutron Powder Diffraction. J. Mater. Chem. 1996, 6, 1831–1835. [Google Scholar] [CrossRef]
  9. Klap, G.J.; van Koningsveld, H.; Graafsma, H.; Schreurs, A.M.M. Absolute Configuration and Domain Structure of AlPO4-5 Studied by Single Crystal X-ray Diffraction. Microporous Mesoporous Mater. 2000, 38, 403–412. [Google Scholar] [CrossRef]
  10. Simmen, A.; McCusker, L.B.; Baerlocher, C.W.; Meier, W. The Structure Determination and Rietveld Refinement of the Aluminophosphate AlPO4-18. Zeolites 1991, 11, 654–661. [Google Scholar] [CrossRef]
  11. Poulet, G.; Tuelm, A.; Sautet, P.A. Combined Experimental and Theoretical Evaluation of the Structure of Hydrated Microporous Aluminophosphate AlPO4-18. J. Phys. Chem. B 2005, 109, 22939–22946. [Google Scholar] [CrossRef]
  12. Bennett, J.M.; Kirchner, R.M. The Structure of Calcined AlPO4-31: A New Framework Topology Containing One-Dimensional 12-Ring Pores. Zeolites 1992, 12, 338–342. [Google Scholar] [CrossRef]
  13. Endregard, M.; Nicholson, D.G.; Stocker, M.; Beagle, B. Cobalticenium Ions Adsorbed on Large-pore Aluminophosphate VPI-5 Studied by X-Ray Absorption, 13CSolid-state NMR and FTlR Spectroscopy. J. Mater. Chem. 1995, 5, 485–491. [Google Scholar] [CrossRef]
  14. Endregard, M.; Nicholson, D.G.; Stocker, M.; Lamble, G.J. Adsorption and Thermal Decomposition of Cobalticenium Ions on AlPO4-5 Studied by X-Ray Absorption Spectroscopy, 13C Solid-State NMR and FTlR. J. Mater. Chem. 1995, 5, 785–791. [Google Scholar] [CrossRef]
  15. Parton, R.F.; Thibault-Starzyk, F.; Reynders, R.A.; Grobet, P.J.; Jacobs, P.A.; Bezoukhanova, C.P.; Sun, W.; Wu, Y. Stacked Phthalocyanines in VPI-5 Pores as Evidenced by CPDOR 1H27Al NMR. Mol. Catal. A 1995, 97, 183–186. [Google Scholar] [CrossRef]
  16. Kärger, J.; Keller, W.; Pfeifer, H.; Ernst, S.; Weitkamp, J. Unexpectedly Low Translational Mobility of Methane and Tetrafluoromethane in the Large-Pore Molecular Sieve VPI-5. Microporous Mater. 1995, 3, 401–408. [Google Scholar] [CrossRef]
  17. Jin, Y.M.; Chon, H. A Novel Method for Encapsulation of Dyes into AlPO4-8 Molecular Sieve. Chem. Commun. 1996, 1996, 135–136. [Google Scholar] [CrossRef]
  18. Ganschow, M.; Schulz-Ekloff, G.; Wark, M.; Wendschuh-Josties, M.; Wöhrle, D. Microwave-Assisted Preparation of Uniform Pure and Dye-Loaded AlPO4-5 Crystals with Different Morphologies for Use as Microlaser Systems. J. Mater. Chem. 2001, 11, 1823–1827. [Google Scholar] [CrossRef]
  19. Gonzalez-Platas, J.; Breton, J.; Girardet, C. Physisorption in a Molecular Helicoidal Cavity: Application to AlPO4-5. Langmuir 1995, 11, 197–203. [Google Scholar] [CrossRef]
  20. Hartmann, M.; Kevan, L. Generation of Ion-Exchange Capacity by Silicon Incorporation into the Aluminophosphate VPI-5/AlPO4-5 Molecular Sieve System. J. Chem. Soc. Faraday Trans. 1996, 92, 3661–3667. [Google Scholar] [CrossRef]
  21. Garcia-Carmona, J.; Fanovich, M.A.; Llibre, J.; Rodriguez-Clemente, R.; Domingo, C. Processing of Microporous VPI-5 Molecular Sieve by Using Supercritical CO2: Stability and Adsorption Properties. Microporous Mesoporous Mater. 2000, 54, 127–137. [Google Scholar] [CrossRef]
  22. Van Heyden, H.; Mintova, S.; Bein, T. AlPO4-18 Nanocrystals Synthesized Under Microwave Irradiation. J. Mater. Chem. 2006, 16, 514–518. [Google Scholar] [CrossRef]
  23. Shutilov, A.; Grenev, I.V.; Kikhtyanin, O.V.; Gavrilov, V.Y. Adsorption of Molecular Hydrogen on Aluminophosphate Zeolites at 77 K. Kinet. Catal. 2012, 53, 137–144. [Google Scholar] [CrossRef]
  24. Weiß, O.; Loerke, J.; Wüstefeld, U.; Marlow, F.; Schüth, F.J. Host-Guest Interactions and Laser Activity in AlPO4-5/Laser Dye Composites. J. Solid State Chem. 2002, 67, 302–309. [Google Scholar] [CrossRef]
  25. Yao, M.; Wang, T.; Yao, Z.; Duan, D.; Chen, S.; Liu, Z.; Liu, R.; Lu, S.; Yuan, Y.; Zou, B.; et al. Pressure-Driven Topological Transformations of Iodine Confined in One-Dimensional Channels. J. Phys. Chem. C 2013, 117, 25052–25058. [Google Scholar] [CrossRef]
  26. Guo, J.; Wang, C.; Xu, J.; Deng, J.; Yan, R.P.; Sharma, R.; Xu, R. Encapsulation of Bulky Solvent Molecules into the Channels of Aluminophosphate Molecular Sieve and its Negative Influence on the Thermal Stability of Open-Framework. Inorg. Chem. Commun. 2018, 91, 67–71. [Google Scholar] [CrossRef]
  27. Carreon, M.L.; Li, S.; Carreon, M.A. AlPO4-18 Membranes for CO2/CH4 Separation. Chem. Commun. 2012, 48, 2310–2312. [Google Scholar] [CrossRef]
  28. Wu, T.; Lucero, J.; Zong, Z.; Elsaidi, S.K.; Thallapally, P.K.; Carreon, M.A. Microporous Crystalline Membranes for Kr/Xe Separation: Comparison Between ALPO-18, SAPO-34, and ZIF-8. ACS Appl. Nano Mater. 2018, 1, 463–470. [Google Scholar] [CrossRef]
  29. Wang, B.; Gao, F.; Zhang, F.; Xing, W.; Zhou, R. Highly Permeable and Oriented AlPO4-18 Membranes Prepared Using Directly Synthesized Nanosheets for CO2/CH4 Separation. J. Mater. Chem. A 2019, 7, 13164–13172. [Google Scholar] [CrossRef]
  30. Tang, Z.K.; Sun, H.D.; Wang, J.; Li, J.C. Silver oxalate: Mechanical properties and extreme negative mechanical phenomena. Appl. Phys. Lett. 1998, 73, 2287–2289. [Google Scholar] [CrossRef]
  31. Launois, P.; Moret, R.; Le Bolloc’h, D.; Albouya, P.A.; Tang, Z.K.; Li, G. Carbon Nanotubes Synthesised in Channels of AlPO4-5 Single Crystals: First X-ray Scattering Investigations. Solid State Commun. 2000, 116, 99–103. [Google Scholar] [CrossRef]
  32. Li, Z.M.; Tang, Z.K.; Liu, H.J.; Wang, N.; Chan, C.T.; Saito, R.; Okada, S.; Li, G.D.; Chen, J.S.; Nagasawa, N.; et al. Polarized Absorption Spectra of Single-Walled 4 Å Carbon Nanotubes Aligned in Channels of an AlPO4-5 Single Crystal. Phys. Rev. Lett. 2001, 87, 127401. [Google Scholar] [CrossRef] [PubMed]
  33. Li, G.D.; Tang, Z.K.; Wang, N.; Chen, J.S. Structural Study of the 0.4-nm Single-Walled Carbon Nanotubes Aligned in Channels of AlPO4-5 crystal. Carbon 2002, 40, 917–921. [Google Scholar]
  34. Zhai, J.P.; Tang, Z.K.; Li, Z.M.; Li, I.L.; Jiang, F.Y.; Shen, P.; Hu, X. Carbonization Mechanism of Tetrapropylammonium-hydroxide in Channels of AlPO4-5 Single Crystals. Chem. Mater. 2006, 18, 1505–1511. [Google Scholar] [CrossRef]
  35. Zhai, J.P.; Li, Z.M.; Liu, H.J.; Li, I.L.; Sheng, P.; Hu, H.J.; Tang, Z.K. Catalytic Effect of Metal Cations on the Formation of Carbon Nanotubes inside the Channels of AlPO4-5 crystal. Carbon 2006, 44, 1151–1157. [Google Scholar] [CrossRef]
  36. Yang, W.; Sun, W.; Zhao, S.; Yin, X. Single-Walled Carbon Nanotubes Prepared in Small AlPO4-5 and CoAPO-5 Molecular Sieves by Low-Temperature Hydrocracking. Microporous Mesoporous Mater. 2016, 219, 87–92. [Google Scholar] [CrossRef]
  37. Concepcion, P.; Lopez Nieto, J.M. Novel Synthesis of Vanadium Cobalt Aluminophosphate Molecular Sieve of AEI Structure (VCoAPO-18) and its Catalytic Behavior for Ethane Oxidation. Catal. Commun. 2001, 2, 363–367. [Google Scholar] [CrossRef]
  38. Concepcion, P.; Blasco, T.; Lopez Nieto, J.M.; Vidal-Moya, A.; Martinez-Arias, A. Preparation, Characterization and Reactivity of V- and/or Co-Containing AlPO4-18 Materials (VCoAPO-18) in the Oxidative Dehydrogenation of Ethane. Microporous Mesoporous Mater. 2004, 67, 215–227. [Google Scholar] [CrossRef]
  39. Dai, W.; Li, N.; Guan, N.; Hunger, M. Unexpected Methanol-to-Olefin Conversion Activity of Low-Silica Aluminophosphate Molecular Sieves. Catal. Commun. 2011, 16, 124–127. [Google Scholar] [CrossRef]
  40. Liu, D.; Zhang, B.; Liu, X.; Li, J. Cyclohexane Oxidation over AFI Molecular Sieves: Effects of Cr, Co Incorporation and Crystal Size. Catal. Sci. Technol. 2015, 5, 3394–3402. [Google Scholar] [CrossRef]
  41. Preeth, M.E.; Umasankari, A.; Rekha, C.H.; Palanichamy, P.; Sivakumar, T.; Pandurangan, A. Selective Oxidation of Cyclohexane to KA Oil Over Ce-AlPO-18 Molecular Sieves. Int. J. Eng. Technol. 2018, 7, 352–354. [Google Scholar] [CrossRef]
  42. Chen, J.; Thomas, J.M.; Wright, P.A. Silicoaluminophosphate Number Eighteen (SAPO-18): A New Microporous Solid Acid Catalyst. Catal. Lett. 1994, 28, 241–248. [Google Scholar] [CrossRef]
  43. Thomas, J.M.; Greaves, G.N.; Sanka, G.; Wright, P.A.; Chen, J.; Dent, A.J.; Marchese, L. On the Nature of the Active Site in a CoAPO-18 Solid Acid Catalyst. Angew. Chem. Int. Ed. 1994, 33, 1871–1873. [Google Scholar] [CrossRef]
  44. Zubowa, H.L.; Richter, M.; Roost, U.; Parlitz, B.; Fricke, R. Synthesis and Catalytic Properties of Substituted AlPO4-31 Molecular Sieves. Catal. Lett. 1993, 19, 67–79. [Google Scholar] [CrossRef]
  45. Chen, Y.; Zhang, D.; Li, F.; Gao, F.; Feng, C.; Wen, S.; Ruan, S. Humidity Sensor Based on AlPO4-5 Zeolite with High Responsivity and its Sensing Mechanism. Sens. Actuators B 2015, 212, 242–247. [Google Scholar] [CrossRef]
  46. Ristic, A.; Logar, N.Z.; Henninger, S.K.; Kaucic, V. The Performance of Small-Pore Microporous Aluminophosphates in Low-Temperature Solar Energy Storage: The Structure–Property Relationship. Adv. Funct. Mater. 2012, 22, 1952–1957. [Google Scholar] [CrossRef]
  47. Henninger, S.K.; Schmidt, F.P.; Henning, H.M. Water Adsorption Characteristics of Novel Materials for Heat Transformation Applications. Applied Thermal Eng. 2010, 30, 1692–1702. [Google Scholar] [CrossRef]
  48. Henninger, S.K.; Schmidt, F.P.; Henning, H.M. Characterisation and Improvement of Sorption Materials with Molecular Modeling for the Use in Heat Transformation Applications. Adsorption 2011, 17, 833–843. [Google Scholar] [CrossRef]
  49. Henninger, S.K.; Jeremias, F.; Kummer, H.; Schossig, P.; Henning, H.M. Novel Sorption Materials for Solar Heating and Cooling. Energy Procedia 2012, 30, 279–288. [Google Scholar] [CrossRef]
  50. Khosrovani, N.; Sleight, A.W. Flexibility of Network Structures. J. Solid State Chem. 1996, 121, 2–11. [Google Scholar] [CrossRef]
  51. 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]
  52. 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]
  53. Sanchez-Valle, C.; Lethbridge, Z.A.D.; Sinogeikin, S.V.; Williams, J.J.; Walton, R.I.; Evans, K.E.; Bass, J.D. Negative Poisson’s Ratios in Siliceous Zeolite MFI-Silicalite. J. Chem. Phys. 2008, 128, 184503. [Google Scholar] [CrossRef] [PubMed]
  54. Sanchez-Valle, C.; Sinogeikin, V.; Lethbridge, Z.A.D.; Walton, R.I.; Smith, C.W.; Evans, K.E.; Bass, J.D. Brillouin Scattering Study on the Single Crystal of Natrolite and Analcime Zeolites. J. Appl. Phys. 2005, 98, 53508. [Google Scholar] [CrossRef]
  55. Eroshenko, V.; Regis, R.C.; Soulard, M.; Patarin, J. Energetics: A New Field of Applications for Hydrophobic Zeolites. J. Am. Chem. Soc. 2001, 123, 8129–8130. [Google Scholar] [CrossRef] [PubMed]
  56. Lee, Y.; Hriljac, J.A.; Vogt, T.; Parise, J.B.; Artioli, G. Phase Transition of Zeolite Rho at High-Pressure. J. Am. Chem. Soc. 2001, 123, 12732–12733. [Google Scholar] [CrossRef]
  57. Lee, Y.; Vogt, T.; Hriljac, J.A.; Parise, J.B.; Artioli, G. Pressure-Induced Volume Expansion of Zeolites in the Natrolite Family. J. Am. Chem. Soc. 2002, 124, 5466–5475. [Google Scholar] [CrossRef]
  58. Gatta, G.D.; Lee, Y. Zeolites at high pressure: A review. Mineral. Mag. 2014, 78, 267–291. [Google Scholar] [CrossRef]
  59. Arletti, R.; Ferro, O.; Quartieri, S.; Sani, A.; Tabacchi, G.; Vezzalini, G. Structural Deformation Mechanisms of Zeolites under Pressure. Am. Mineral. 2003, 88, 1416–1422. [Google Scholar] [CrossRef]
  60. Astala, R.; Auerbach, S.M.; Monson, P.A. Density Functional Theory Study of Silica Zeolite Structures: Stabilities and Mechanical Properties of SOD, LTA, CHA, MOR, and MFI. J. Phys. Chem. B 2004, 108, 9208–9215. [Google Scholar] [CrossRef]
  61. Astala, R.; Auerbach, S.M.; Monson, P.A. Normal mode approach for predicting the mechanical properties of solids from first principles: Application to compressibility and thermal expansion of zeolites. Phys. Rev. B 2005, 71, 014112. [Google Scholar] [CrossRef]
  62. Sartbaeva, A.; Wells, S.A.; Tracy, M.M.J.; Thorpe, M.F. The flexibility window in zeolites. Nat. Mater. 2006, 5, 962–965. [Google Scholar] [CrossRef] [PubMed]
  63. Fletcher, R.E.; Wells, S.A.; Leung, K.M.; Edwards, P.P.; Sartbaeva, A. Intrinsic flexibility of porous materials; theory, modelling and the flexibility window of the EMT zeolite framework. Acta Crystallogr. B 2015, 71, 641–647. [Google Scholar] [CrossRef] [PubMed]
  64. Coudert, F.X. Systematic investigation of the mechanical properties of pure silicazeolites: Stiffness, anisotropy, and negative linear compressibility. Phys. Chem. Chem. Phys. 2013, 15, 16012–16018. [Google Scholar] [CrossRef]
  65. Evans, J.D.; Coudert, F.X. Predicting the Mechanical Properties of Zeolite Frameworks by Machine Learning. Chem. Mater. 2017, 29, 7833–7839. [Google Scholar] [CrossRef]
  66. Gaillac, R.; Chibani, S.; Coudert, F.X. Speeding up discovery of auxetic zeolite frameworks by machine learning. Chem. Mater. 2020, 32, 2653–2663. [Google Scholar] [CrossRef]
  67. Santoro, M.; Veremeienko, V.; Polisi, M.; Fantini, R.; Alabarse, F.; Arletti, R.; Quatieri, S.; Svitlyk, V.; Van der Lee, A.; Rouquette, J.; et al. Insertion and Confinement of H2O in Hydrophobic Siliceous Zeolites at High Pressure. J. Phys. Chem. C 2019, 123, 17432–17439. [Google Scholar] [CrossRef]
  68. Bahr, D.F.; Reid, J.A.; Mook, W.M.; Bauer, C.A.; Stumpf, R.; Skulan, A.J.; Moody, N.R.; Simmons, B.A.; Shindel, M.M.; Allendorf, M.D. Mechanical properties of cubic zinc carboxylate IRMOF-1 metal-organic framework crystals. Phys. Rev. B 2007, 76, 184106. [Google Scholar] [CrossRef]
  69. Chapman, K.W.; Halder, G.J.; Chupas, P.J. Pressure-Induced Amorphization and Porosity Modification in a Metal−Organic Framework. J. Am. Chem. Soc. 2009, 131, 17546–17547. [Google Scholar] [CrossRef]
  70. Fairen-Jimenez, D.; Moggach, S.A.; Wharmby, T.; Wright, P.A.; Parsons, S.; Düren, S. Opening the Gate: Framework Flexibility in ZIF-8 Explored by Experiments and Simulations. J. Am. Chem. Soc. 2011, 133, 8900–8902. [Google Scholar] [CrossRef]
  71. Li, W.; Probert, M.R.; Kosa, M.; Bennett, T.D.; Thirumurugan, A.; Burwood, R.P.; Parinello, M.; Howard, J.A.; Cheetham, A.K. Negative Linear Compressibility of a Metal−organic Framework. J. Am. Chem. Soc. 2012, 134, 11940–11943. [Google Scholar] [CrossRef]
  72. Wu, H.; Yildirim, T.; Zhou, W.J. Exceptional Mechanical Stability of Highly Porous Zirconium Metal-Organic Framework UiO-66 and Its Important Implications. J. Phys. Chem. Lett. 2013, 4, 925–930. [Google Scholar] [CrossRef]
  73. Ortiz, A.U.; Boutin, A.; Fuchs, A.H.; Coudert, F.X. Anisotropic Elastic Properties of Flexible Metal-Organic Frameworks: How Soft are Soft Porous Crystals? Phys. Rev. Lett. 2012, 109, 195502. [Google Scholar] [CrossRef] [PubMed]
  74. Ortiz, A.U.; Boutin, A.; Fuchs, A.H.; Coudert, F.X. Organic Frameworks with Wine-Rack Motif: What Determines their Flexibility and Elastic Properties? J. Chem. Phys. 2013, 138, 174703. [Google Scholar] [CrossRef]
  75. Ortiz, A.U.; Boutin, A.; Fuchs, A.H.; Coudert, F.X. Investigating the Pressure-Induced Amorphization of Zeolitic Imidazolate Framework ZIF-8: Mechanical Instability Due to Shear Mode Softening. J. Phys. Chem. Lett. 2013, 4, 1861–1865. [Google Scholar] [CrossRef] [PubMed]
  76. Coudert, F.X. Responsive Metal–Organic Frameworks and Framework Materials: Under Pressure, Taking the Heat, in the Spotlight, with Friends. Chem. Mater. 2015, 27, 1905–1916. [Google Scholar] [CrossRef]
  77. Cai, W.; Katrusiak, A. Giant negative linear compression positively coupled to massive thermal expansion in a metal–organic framework. Chem. Commun. 2014, 5, 4337–4342. [Google Scholar] [CrossRef]
  78. Cai, W.; Gładysiak, A.; Anioła, M.; Smith, V.J.; Barbour, L.J.; Katrusiak, A. Giant Negative Area Compressibility Tunable in a Soft Porous Framework Material. J. Am. Chem. Soc. 2015, 137, 9296–9301. [Google Scholar] [CrossRef]
  79. Serra-Crespo, P.; Dikhtiarenko, A.; Stavitski, E.; Juan-Alcañiz, J.; Kapteijn, F.; Coudert, F.X.; Gascon, J. Experimental evidence of negative linear compressibility in the MIL-53 metal–organic framework family. CrystEngComm 2015, 17, 276–280. [Google Scholar] [CrossRef]
  80. Tan, J.C.; Bennett, T.D.; Cheetham, A.K. Chemical structure, network topology, and porosity effects on the mechanical properties of Zeolitic Imidazolate Frameworks. Proc. Natl. Acad. Sci. USA 2010, 107, 9938–9943. [Google Scholar] [CrossRef]
  81. Tan, J.C.; Cheetham, A.K. Mechanical properties of hybrid inorganic–organic framework materials: Establishing fundamental structure–property relationships. Chem. Soc. Rev. 2011, 40, 1059–1080. [Google Scholar] [CrossRef]
  82. Tan, J.C.; Civalleri, B.; Lin, C.C.; Valenzano, L.; Galvelis, R.; Chen, P.F.; Bennett, T.D.; Mellot-Draznieks, C.; Zicovich-Wilson, C.M.; Cheetham, A.K. Exceptionally Low Shear Modulus in a Prototypical Imidazole-Based Metal-Organic Framework. Phys. Rev. Lett. 2012, 108, 095502. [Google Scholar] [CrossRef] [PubMed]
  83. Ryder, M.R.; Tan, J.C. Explaining the mechanical mechanisms of zeolitic metal–organic frameworks: Revealing auxeticity and anomalous elasticity. Dalton Trans. 2016, 45, 4154–4161. [Google Scholar] [CrossRef] [PubMed]
  84. Yot, P.G.; Boudene, Z.; Macia, J.; Granier, D.; Vanduyfhuys, L.; Verstraelen, T.; Speybroeck, V.V.; Devic, T.; Serre, C.; Ferey, G.; et al. Metal–organic frameworks as potential shock absorbers: The case of the highly flexible MIL-53(Al). Chem. Commun. 2014, 50, 9462–9464. [Google Scholar] [CrossRef] [PubMed]
  85. Banlusan, K.; Antillon, E.; Strachan, A. Mechanisms of Plastic Deformation of Metal−Organic Framework-5. J. Phys. Chem. C 2015, 119, 25845–25852. [Google Scholar] [CrossRef]
  86. Banlusan, K.; Amornkitbamrung, V. J Effects of Free Volume on Shock-Wave Energy Absorption in A Metal–Organic Framework: A Molecular Dynamics Investigation. Phys. Chem. C 2020, 124, 17027–17038. [Google Scholar] [CrossRef]
  87. Fraux, G.; Coudert, F.X.; Boutin, A.; Fuchs, A.H. Forced intrusion of water and aqueous solutions in microporous materials: From fundamental thermodynamics to energy storage devices. Chem. Soc. Rev. 2017, 46, 7421–7437. [Google Scholar] [CrossRef]
  88. Terracina, A.; Todaro, M.; Mazaj, M.; Agnello, S.; Gelardi, F.M.; Buscarino, G. Unveiled the Source of the Structural Instability of HKUST-1 Powders upon Mechanical Compaction: Definition of a Fully Preserving Tableting Method. J. Phys. Chem. C 2019, 123, 1730–1741. [Google Scholar] [CrossRef]
  89. Redfern, L.R.; Robison, L.; Wasson, M.C.; Goswami, S.; Lyu, J.; Islamoglu, T.; Chapman, K.W.; Farha, O.K. Isolating the Role of the Node-Linker Bond in the Compression of UiO-66 Metal−Organic Frameworks. J. Am. Chem. Soc. 2019, 141, 4365–4371. [Google Scholar] [CrossRef]
  90. Moghadam, P.Z.; Rogge, S.M.; Li, A.; Chow, C.M.; Wieme, J.; Moharrami, N.; Aragones-Anglada, M.; Conduit, G.; Gomez-Gualdron, D.A.; Van Speybroeck, V.; et al. Structure-Mechanical Stability Relations of Metal-Organic Frameworks via Machine Learning. Matter 2019, 19, 219–234. [Google Scholar] [CrossRef]
  91. Zhou, X.; Miao, Y.; Suslick, K.S.; Dlott, D.D. Mechanochemistry of Metal–Organic Frameworks under Pres-sure and Shock. Acc. Chem. Res. 2020, 53, 2806–2815. [Google Scholar] [CrossRef]
  92. Zeng, Q.; Wang, K.; Qiao, Y.; Li, X.; Zou, B. Negative Linear Compressibility Due to Layer Sliding in a Layered Metal−Organic Framework. J. Phys. Chem. Lett. 2017, 8, 1436–1441. [Google Scholar] [CrossRef] [PubMed]
  93. Zeng, Q.; Wang, K.; Zou, B. Large Negative Linear Compressibility in InH(BDC)2 from Framework Hinging. J. Am. Chem. Soc. 2017, 139, 15648–15651. [Google Scholar] [CrossRef]
  94. Zeng, Q.; Wang, K.; Zou, B. Negative Linear Compressibility Response to Pressure in Multitype Wine-Rack Metal−Organic Frameworks. ACS Mater. Lett. 2020, 2, 291–295. [Google Scholar] [CrossRef]
  95. Yan, Y.; O’Connor, A.E.; Kanthasamy, G.; Atkinson, G.; Allan, D.R.; Blake, A.J.; Schroder, M. Unusual and Tunable Negative Linear Compressibility in the Metal-Organic Framework MFM-133(M) (M = Zr, Hf). J. Am. Chem. Soc. 2018, 140, 3952–3958. [Google Scholar] [CrossRef] [PubMed]
  96. Feng, G.; Zhang, W.; Dong, L.; Li, W.; Cai, W.; Wei, W.; Ji, L.; Lin, Z.; Lu, P. Negative area compressibility of a hydrogen bonded two-dimensional material. Chem. Sci. 2019, 10, 1309–1315. [Google Scholar] [CrossRef] [PubMed]
  97. Chen, Z.; Xu, B.; Li, Q.; Meng, Y.; Quan, Z.; Zou, B. Selected Negative Linear Compressibilities in the Metal−Organic Framework of [Cu(4,4′-bpy)2 (H2O)2]·SiF6. Inorg. Chem. 2020, 59, 1715–1722. [Google Scholar] [CrossRef]
  98. Zajdel, P.; Chorążewski, M.; Leão, J.B.; Jensen, G.V.; Bleuel, M.; Zhang, H.F.; Feng, T.; Luo, D.; Li, M.; Lowe, A.R.; et al. Inflation Negative Compressibility during Intrusion−Extrusion of a Non-Wetting Liquid into a Flexible Nanoporous Framework. Phys. Chem. Lett. 2021, 12, 4951–4957. [Google Scholar] [CrossRef]
  99. Tortora, M.; Zajdel, P.; Lowe, A.R.; Chorążewski, M.; Leão, J.B.; Jensen, G.V.; Bleuel, M.; Giacomello, M.; Casciola, C.M.; Meloni, S.; et al. Giant Negative Compressibility by Liquid Intrusion into Superhydrophobic Flexible Nanoporous Frameworks. Nano Lett. 2021, 21, 2848–2853. [Google Scholar] [CrossRef]
  100. Colmenero, F. Negative Linear Compressibility in Nanoporous Metal–Organic Frameworks Rationalized by the Empty Channel Structural Mechanism. Phys. Chem. Chem. Phys. 2021, 23, 8508–8524. [Google Scholar] [CrossRef]
  101. Colmenero, F.; Lobato, A.; Timón, V. Compressing the Channels in the Crystal Structure of Copper Squarate Metal-Organic Framework. Solids 2022, 3, 374–384. [Google Scholar] [CrossRef]
  102. Zeng, Q.; Wang, K.; Zou, B. Near Zero Area Compressibility in a Perovskite-Like Metal−Organic Frameworks [C(NH2 )3][Cd(HCOO)3]. ACS Appl. Mater. Interfaces 2018, 10, 23481–23484. [Google Scholar] [CrossRef] [PubMed]
  103. Yu, Y.; Zeng, Q.; Chen, Y.; Jiang, L.; Wang, K.; Zou, B. Extraordinarily Persistent Zero Linear Compressibility in Metal-Organic Framework MIL-122(In). ACS Mater. Lett 2020, 2, 519–523. [Google Scholar] [CrossRef]
  104. Zhang, X.; Sui, Z.; Xu, B.; Yue, S.; Luo, Y.; Zhan, W.; Liu, B.A. Mechanically strong and highly conductive graphene aerogel and its use as electrodes for electrochemical power source. J. Mater. Chem. 2011, 21, 6494–6497. [Google Scholar] [CrossRef]
  105. Mecklenburg, M.; Schuchardt, A.; Mishra, Y.K.; Kaps, S.; Adelung, R.; Lotnyk, A.; Kienle, L.; Schulte, K. Aerographite: Ultra Lightweight, Flexible Nanowall, Carbon Microtube Material with Outstanding Mechanical Performance. Adv. Mater. 2012, 24, 3486–3490. [Google Scholar] [CrossRef]
  106. Worsley, M.A.; Kucheyev, S.O.; Mason, H.E.; Merrill, M.D.; Mayer, B.P.; Lewicki, J.; Valdez, C.A.; Suss, M.E.; Stadermann, M.; Pauzauskie, M.J.; et al. Mechanically robust 3D graphene macroassembly with high surface area. Chem. Commun. 2012, 48, 8428–8430. [Google Scholar] [CrossRef] [PubMed]
  107. Hu, H.; Zhao, Z.; Wan, W.; Gogotsi, Y.; Qiu, J. Ultralight and Highly Compressible Graphene Aerogels. Adv. Mater. 2013, 25, 2219–2223. [Google Scholar] [CrossRef]
  108. Zhu, C.; Han, T.Y.; Duoss, E.B.; Golobic, A.M.; Kuntz, J.; Spadaccini, C.M.; Worsley, M.A. Highly compressible 3D periodic graphene aerogel microlattices. Nat. Commun. 2015, 6, 6962. [Google Scholar] [CrossRef]
  109. Lei, J.; Liu, Z. The structural and mechanical properties of graphene aerogels based on Schwarz-surface-like graphene models. Carbon 2018, 130, 741–748. [Google Scholar] [CrossRef]
  110. Wu, Y.; Yi, N.; Huang, L.; Zhang, T.; Fang, S.; Chang, H.; Li, N.; Oh, J.; Lee, J.A.; Kozlov, M. Three-dimensionally bonded spongy graphene material with super compressive elasticity and near-zero Poisson’s ratio. Nat. Commun. 2015, 6, 6141. [Google Scholar] [CrossRef]
  111. Chen, B.; Ma, Q.; Tan, C.; Lim, T.T.; Huang, L.; Zhang, H. Carbon-Based Sorbents with Three-Dimensional Architectures for Water Remediation. Small 2015, 11, 3319–3336. [Google Scholar] [CrossRef]
  112. Xu, X.; Zhang, Q.; Yu, Y.; Chen, W.; Hu, H.; Li, H. Naturally Dried Graphene Aerogels with Superelasticity and Tunable Poisson’s Ratio. Adv. Mater. 2016, 28, 9223–9230. [Google Scholar] [CrossRef] [PubMed]
  113. Robertson, M.C.; Mokaya, R. Microporous activated carbon aerogels via a simple subcritical drying route for CO2 capture and hydrogen storage. Microporous Mesoporous Mater. 2013, 179, 151–156. [Google Scholar] [CrossRef]
  114. Patil, S.P.; Shendye, P.; Markert, B. Molecular investigation of mechanical properties and fracture behavior of graphene aerogel. J. Phys. Chem. B 2020, 124, 6132–6139. [Google Scholar] [CrossRef] [PubMed]
  115. Patil, S.P.; Kulkarni, A.; Markert, B. Shockwave response of graphene aerogels: An all-atom simulation study. Comput. Mater. Sci. 2021, 189, 110252. [Google Scholar] [CrossRef]
  116. Cho, H.J.; Kim, I.D.; Jung, S.M. Multifunctional Inorganic Nanomaterial Aerogel Assembled into fSWNT Hydrogel Platform for Ultraselective NO2 Sensing. ACS Appl. Mater. Interfaces 2020, 12, 10637–10647. [Google Scholar] [CrossRef]
  117. Barthelet, K.; Marrot, J.; Riou, D.; Férey, G. A Breathing Hybrid Organic-Inorganic Solid with Very Large Pores and High Magnetic Characteristics. Angew. Chem. Int. Ed. 2022, 41, 281–284. [Google Scholar] [CrossRef]
  118. Bradshaw, D.; Claridge, J.B.; Cussen, E.J.; Prior, T.J.; Rosseinsky, M. Design, Chirality, and Flexibility in Nanoporous Molecule-Based Materials. J. Acc. Chem. Res. 2005, 38, 273–282. [Google Scholar] [CrossRef]
  119. Serre, C.; Mellot-Draznieks, C.; Surble, C.; Audebrand, N.; Filinchuk, Y.; Ferey, G. Role of Solvent-Host Interactions That Lead to Very Large Swelling of Hybrid Frameworks. Science 2007, 315, 1828–1831. [Google Scholar] [CrossRef]
  120. Férey, G.; Serre, C. Large breathing effects in three-dimensional porous hybrid matter: Facts, analyses, rules and consequences. Chem. Soc. Rev. 2009, 38, 1380–1399. [Google Scholar] [CrossRef]
  121. Sato, H.; Kosaka, W.; Matsuda, R.; Hori, A.; Hijikata, Y.; Belosludov, R.V.; Sakaki, S.; Takata, M.; Kitagawa, S. Self-Accelerating CO Sorption in a Soft Nanoporous Crystal. Science 2014, 343, 167–170. [Google Scholar] [CrossRef]
  122. Sanchez-Gonzalez, E.; Mileo, P.G.M.; Sagastuy-Breña, M.; Raziel-Alvarez, J.; Reynolds, J.E.; Villarreal, A.; Gutierrez-Alejandre, A.; Ramirez, J.; Balmaseda, J.; González-Zamora, E.; et al. Highly reversible sorption of H2S and CO2 by an environmentally friendly Mg-based MOF. J. Mater. Chem. A 2018, 6, 16900–16909. [Google Scholar] [CrossRef]
  123. Remy, T.; Baron, G.V.; Denayer, J.F.M. Modeling the Effect of Structural Changes during Dynamic Separation Processes on MOFs. Langmuir 2011, 27, 13064–13071. [Google Scholar] [CrossRef] [PubMed]
  124. Remy, T.; Ma, L.; Maes, M.; De Vos, D.E.; Baron, G.V.; Denayer, J.F.M. Vapor-Phase Adsorption and Separation of Ethylbenzene and Styrene on the Metal−Organic Frameworks MIL-47 and MIL-53(Al). Ind. Eng. Chem. Res. 2012, 51, 14824–14833. [Google Scholar] [CrossRef]
  125. Horcajada, P.; Serre, C.; Maurin, G.; Ramsahye, N.A.; Balas, F.; Vallet-Regi, M.; Sebban, M.; Taulelle, F.; Férey, G. Flexible Porous Metal-Organic Frameworks for a Controlled Drug Delivery. J. Am. Chem. Soc. 2008, 130, 6774–6780. [Google Scholar] [CrossRef]
  126. Millange, F.; Serre, C.; Guillou, N.; Férey, G.; Walton, R.I. Structural Effects of Solvents on the Breathing of Metal–Organic Frameworks: An In Situ Diffraction Study. Angew. Chem. Int. Ed. 2008, 47, 4100–4105. [Google Scholar] [CrossRef] [PubMed]
  127. Allendorf, M.D.; Houk, R.J.T.; Andruszkiewicz, L.; Talin, A.; Pikarsky, J.; Choudhury, A.; Gall, K.A.; Hesketh, P.J. Stress-induced chemical detection using flexible metal-organic frameworks. J. Am. Chem. Soc. 2008, 130, 14404–14405. [Google Scholar] [CrossRef] [PubMed]
  128. Yot, P.G.; Vanduyfhuys, L.; Alvarez, E.; Rodriguez, J.; Itíe, J.P.; Fabry, P.; Guillou, N.; Devic, T.; Beurroies, I.; Llewellyn, P.L.; et al. Metal−Organic Frameworks as Potential Shock Absorbers: The Case of The Highly Flexible MIL-53(Al). Chem. Sci. 2016, 7, 446–450. [Google Scholar] [CrossRef]
  129. Baughman, R.H.; Stafstrom, S.; Cui, C.; Dantas, S.O. Compressibilities in One or More Dimensions. Science 1998, 279, 1522–1524. [Google Scholar] [CrossRef]
  130. Lakes, R.S.; Wojciechowski, K.W. Negative Compressibility, Negative Poisson’s Ratio and Stability. Phys. Stat. Sol. B 2008, 245, 545–551. [Google Scholar] [CrossRef]
  131. Cairns, A.B.; Goodwin, A.L. Negative Linear Compressibility. Phys. Chem. Chem. Phys. 2015, 17, 20449–20465. [Google Scholar] [CrossRef]
  132. Xie, Y.M.; Yang, X.; Shen, J.; Yan, X.; Ghaedizadeh, A.; Rong, J.; Huang, X.; Zhou, S. Designing orthotropic materials for negative or zero compressibility. Int. J. Solids Struct. 2014, 51, 4038–4051. [Google Scholar] [CrossRef]
  133. Jiang, X.; Yang, Y.; Molokeev, M.S.; Gong, L.; Liang, F.; Wang, S.; Liu, L.; Wu, X.; Li, X.; Wu, S.; et al. Zero Linear Compressibility in Nondense Borates with a “Lu-Ban Stool“-Like Structure. Adv. Mater. 2018, 30, 1801313. [Google Scholar] [CrossRef]
  134. Jiang, X.; Molokeev, M.S.; Dong, L.; Dong, Z.; Wang, N.; Kang, L.; Li, X.; Li, Y.; Tian, C.; Peng, S.; et al. Anomalous mechanical materials squeezing three-dimensional volume compressibility into one dimension. Nat. Commun. 2020, 11, 5593. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, L.; Dove, M.T.; Shi, J.; Sun, B.; Hu, D.; Wang, D. Adjustable uniaxial zero thermal expansion and zero linear compressibility in unique hybrid semiconductors: The role of the organic chain. Dalton Trans. 2020, 49, 719–728. [Google Scholar] [CrossRef] [PubMed]
  136. Lakes, R.S. Foam Structures with a Negative Poisson’s Ratio. Science 1987, 235, 1038–1040. [Google Scholar] [CrossRef] [PubMed]
  137. 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]
  138. Lakes, R.S. Negative-Poisson’s-Ratio Materials: Auxetic Solids. Annu. Rev. Mater. Res. 2017, 47, 63–81. [Google Scholar] [CrossRef]
  139. Evans, K.E.; Alderson, A. Auxetic Materials: Functional Materials and Structures from Lateral Thinking! Adv. Mater. 2000, 12, 617–628. [Google Scholar] [CrossRef]
  140. Spinks, G.M.; Wallace, G.G.; Fifield, L.S.; Dalton, L.R.; Mazzoldi, A.; De Rossi, D.; Khayrullin, I.I.; Baughman, R.H. Pneumatic Carbon Nanotube Actuators. Adv. Mater. 2002, 14, 1728–1732. [Google Scholar] [CrossRef]
  141. Nicolaou, Z.G.; Motter, A.E. Mechanical Metamaterials with Negative Compressibility Transitions. Nat. Mater. 2012, 11, 608–613. [Google Scholar] [CrossRef]
  142. Cairns, A.B.; Catafesta, J.; Levelut, C.; Rouquette, J.; Lee, A.; Peters, V.D.; Thompson, A.L.; Dmitriev, V.; Haines, J.; Goodwin, A.L. Giant Negative Linear Compressibility in Zinc Dicyanoaurate. Nat. Mater. 2013, 12, 212–216. [Google Scholar] [CrossRef] [PubMed]
  143. Fang, N.; Xi, D.; Xu, J.; Ambati, M.; Srituravanich, W.; Sun, C.; Zhang, X. Ultrasonic metamaterials with negative modulus. Nat. Mater. 2006, 5, 452–456. [Google Scholar] [CrossRef] [PubMed]
  144. Aliev, A.V.; Oh, J.; Kozlov, M.E.; Kuznetsov, A.A.; Fang, S.; Fonseca, A.F.; Ovalle, R.; Lima, M.D.; Haque, M.H.; Gartstein, Y.N.; et al. Giant-Stroke, Superelastic Carbon Nanotube Aerogel Muscles. Science 2009, 323, 1575–1578. [Google Scholar] [CrossRef]
  145. Uhoya, W.; Tsoi, G.; Vohra, Y.K.; McGuire, M.A.; Sefat, A.S.; Sales, B.C.; Mandrus, D.; Weir, S.T. nomalous Compressibility Effects and Superconductivity of EuFe2As2 under High Pressures. J. Phys. Cond. Matter. 2010, 22, 292202. [Google Scholar] [CrossRef] [PubMed]
  146. Alderson, A.; Rasburn, J.; Ameer-Beg, S.; Mullarkey, P.G.; Perrie, W.; Evans, K.E. An auxetic filter: A tuneable filter displaying enhanced size selectivity or defouling properties. Ind. Eng. Chem. Res. 2000, 39, 654–665. [Google Scholar] [CrossRef]
  147. Alderson, A.; Rasburn, J.; Evans, K.E.; Grima, J.N. Modelling of the mechanical and mass transport properties of auxetic molecular sieves: An idealised organic (polymeric honeycomb) host–guest system. Membr. Technol. 2001, 137, 6–8. [Google Scholar] [CrossRef]
  148. Rasburn, J.; Alderson, A.; Ameer-Beg, S.; Mullarkey, P.G.; Perrie, W.; Evans, K.E.; Perrie, W.; Evans, K.E. Auxetic structures for variable permeability systems. AIChE J. 2001, 47, 2623–2626. [Google Scholar] [CrossRef]
  149. Greaves, G.N.; Meneau, F.; Sapelkin, A.; Colyer, L.M.; Gwynn, I.; Wade, S.; Sankar, G. The Rheology of Collapsing Zeolites Amorphized by Temperature and Pressure. Nat. Mater. 2003, 2, 622–628. [Google Scholar] [CrossRef]
  150. Haines, J.; Levelut, C.; Isambert, A.; Hebert, P.; Kohara, S.; Keen, D.A.; Hamouda, T.; Andraul, D. Topologically Ordered Amorphous Silica Obtained from the Collapsed Siliceous Zeolite, Silicalite-1-F, A Step toward Perfect Glasses. J. Am. Chem. Soc. 2009, 131, 12333–12338. [Google Scholar] [CrossRef]
  151. Wang, L.; Wang, W.; Chen, L.; Shen, Z. Formation of a unique glass by spark plasma sintering of a zeolite. J. Mater. Res. 2009, 24, 3241–3245. [Google Scholar] [CrossRef]
  152. Moggach, S.A.; Bennett, T.D.; Cheetham, A.K. The Effect of Pressure on ZIF-8: Increasing Pore Size with Pressure and the Formation of a High-Pressure Phase at 1.47 GPa. Angew. Chem. 2009, 121, 7221–7223. [Google Scholar] [CrossRef]
  153. Hwang, G.C.; Shin, T.J.; Blom, D.A.; Vogt, T.; Lee, Y. Pressure-Induced Amorphization of Small Pore Zeolites—The Role of Cation-H2O Topology and Antiglass Formation. Sci. Rep. 2015, 5, 15056. [Google Scholar] [CrossRef] [PubMed]
  154. Santoro, M.; Gorelli, F.; Haines, J.; Cambon, O.; Levelut, C.; Garbarino, G. Silicon carbonate phase formed from carbon dioxide and silica under pressure. Proc. Natl. Acad. Sci. USA 2011, 108, 7689–7692. [Google Scholar] [CrossRef] [PubMed]
  155. Santoro, M.; Gorelli, F.A.; Bini, R. Carbon enters silica forming a cristobalite-type CO2-SiO2 solid solution. Nat. Commun. 2013, 4, 1557. [Google Scholar] [CrossRef] [PubMed]
  156. Santoro, M.; Gorelli, F.; Bini, R.; Salamat, A.; Garbarino, G.; Levelut, C.; Cambon, O.; Haines, J. High-Pressure Synthesis of a Polyethylene/Zeolite Nano-Composite Material. Nat. Commun. 2014, 5, 3761. [Google Scholar] [CrossRef]
  157. Jorda, J.F.; Rey, F.; Sastre, G.; Valencia, S.; Palomino, M.; Corma, A.; Segura, A.; Errandonea, D.; Lacomba, R.; Manjon, F.J.; et al. Synthesis of a Novel Zeolite through a Pressure-Induced Reconstructive Phase Transition Process. Angew. Chem. Int. Ed. 2013, 125, 10652–10656. [Google Scholar] [CrossRef]
  158. Zhou, R.; Qu, B.; Dai, J.; Zeng, X.C. Unravelling the crystal structure of the high-pressure phase of silicon carbonate. Phys. Rev. X 2014, 4, 011030. [Google Scholar]
  159. Marques, M.; Morales-Garcia, A.; Menendez, J.M.; Baonza, V.G.; Recio, J.M. A novel crystalline SiCO compound. Phys. Chem. Chem. Phys. 2015, 17, 25055–25060. [Google Scholar] [CrossRef]
  160. Santamaria-Perez, D.; Marqueño, T.; MacLeod, S.; Ruiz-Fuertes, J.; Daisenberger, D.; Chuliá-Jordan, R.; Errandonea, D.; Jordá, J.L.; Rey, F.; McGuire, C.; et al. Structural Evolution of CO2-Filled Pure Silica LTA Zeolite under High- Pressure High-Temperature Conditions. Chem. Mater. 2017, 29, 4502–4510. [Google Scholar] [CrossRef]
  161. Marqueño, T.; Santamaria-Perez, D.; Ruiz-Fuertes, J.; Chuliá-Jordán, R.; Jordá, J.L.; Rey, F.; McGuire, C.; Kavner, A.; MacLeod, A.S.; Daisenberger, D.; et al. An Ultrahigh CO2-Loaded Silicalite-1 Zeolite: Structural Stability and Physical Properties at High Pressures and Temperatures. Inorg. Chem. 2018, 57, 6447–6455. [Google Scholar] [CrossRef]
  162. Thibaud, J.M.; Rouquette, J.; Hermet, P.; Dziubek, K.; Gorelli, M.; Santoro, M.; Garbarino, G.; Alabarse, F.G.; Cambon, O.; Di Renzo, F.; et al. Saturation of the Siliceous Zeolite TON with Neon at High Pressure. J. Phys. Chem. C 2017, 121, 4283–4292. [Google Scholar] [CrossRef]
  163. Tan, C.; Liu, Z.; Yonezawa, Y.; Sukenaga, S.; Ando, M.; Shibata, H.; Sasaki, Y.; Okubo, T.; Wakihara, T. Unique Crystallization Behavior in Zeolite Synthesis under External High Pressures. Chem. Commun. 2020, 56, 2811–2814. [Google Scholar] [CrossRef]
  164. Deneyer, A.; Ke, J.; Devos, M.; Dusselier, M. Zeolite Synthesis under Nonconventional Conditions: Reagents, Reactors, and Modi Operandi. Chem. Mater. 2020, 32, 4884–4919. [Google Scholar] [CrossRef]
  165. Lv, H.; Yao, M.; Li, Q.; Liu, R.; Liu, B.; Lu, S.; Jiang, L.; Cui, W.; Liu, Z.; Liu, J.; et al. The structural stability of AlPO4-5 zeolite under pressure: Effect of the pressure transmission medium. J. Appl. Phys. 2012, 111, 112615. [Google Scholar] [CrossRef]
  166. Kim, T.; Lee, Y.; Jang, Y.N.; Shin, J.; Hong, S.B. Contrasting high-pressure compression behaviors of AlPO4-5 and SSZ-24 with the same AFI framework topology. Microporous Mesoporous Mater. 2013, 169, 42–46. [Google Scholar] [CrossRef]
  167. Lotti, P.; Gatta, G.D.; Comboni, D.; Merlini, M.; Pastero, L.; Hanfland, M. AlPO4-5 zeolite at high pressure: Crystal-fluid interaction and elastic behavior. Microporous Mesoporous Mater. 2016, 228, 158–167. [Google Scholar] [CrossRef]
  168. Alabarse, F.G.; Silly, G.; Haidoux, A.; Levelut, C.; Bourgogne, D.; Flank, A.M.; Lagarde, P.; Pereira, A.S.; Bantignies, J.L.; Cambon, O.; et al. Effect of H2O on the Pressure-Induced Amorphization of AlPO4-54·xH2O. Phys. Chem. C 2014, 118, 3651–3663. [Google Scholar] [CrossRef]
  169. Alabarse, F.G.; Brubach, J.; Roy, P.; Haidoux, A.; Levelut, C.; Bantignies, J.L.; Cambon, O.; Haines, J. AlPO4-54 − AlPO4-8 Structural Phase Transition and Amorphization under High Pressure. Mechanism of H2O Insertion and Chemical Bond Formation in AlPO4-54·xH2O at High Pressure. J. Phys. Chem. C 2015, 119, 7771–7779. [Google Scholar] [CrossRef]
  170. Alabarse, F.G.; Rouquette, J.; Coasne, B.; Haidoux, A.; Paulmann, C.; Cambon, O.; Haines, J. Mechanism of H2O insertion and chemical bond formation in AlPO(4)-54·xH2O at high pressure. J. Am. Chem. Soc. 2015, 137, 584–587. [Google Scholar] [CrossRef]
  171. Alabarse, F.G.; Silly, G.; Brubach, J.B.; Roy, P.; Haidoux, A.; Levelut, C.; Bantignies, J.B.; Kohara, S.; Floch, S.L.; Cambon, O.; et al. Anomalous Compressibility and Amorphization in AlPO4-17, the Oxide with the Highest Negative Thermal Expansion. Phys. Chem. C 2017, 121, 6852–6863. [Google Scholar] [CrossRef]
  172. Alabarse, F.G.; Joseph, B.; Lausi, A.; Haines, J. Effect of H2O on the Pressure-Induced Amorphization of Hydrated AlPO4-17. Molecules 2019, 24, 2864. [Google Scholar] [CrossRef] [PubMed]
  173. Domenico, S.N. Elastic properties of unconsolidated porous sand reservoirs. Geophysics 1977, 42, 1339–1368. [Google Scholar] [CrossRef]
  174. Eberhart-Phillips, D.; Han, D.; Zoback, M. Empirical relationships among seismic velocity, effective pressure, porosity, and clay content in sandstone. Geophysics 1989, 54, 82–89. [Google Scholar] [CrossRef]
  175. Thomsen, L. Elastic Anisotropy Due to Aligned Cracks in Porous Rock. Geophys. Prospect. 1995, 3, 805–829. [Google Scholar] [CrossRef]
  176. Dvorkin, J.; Nur, A. Elasticity of high-porosity sandstones: Theory for two North Sea data sets. Geophysics 1996, 61, 1363–1370. [Google Scholar] [CrossRef]
  177. Nur, A.; Mavko, G.; Dvorkin, J.; Galmudi, D. Diagnosing high-porosity sandstones: Strength and permeability from porosity and velocity. Lead. Edge 1998, 17, 357–362. [Google Scholar] [CrossRef]
  178. Schön, J.H. Chapter 6 Elastic Properties. In Physical Properties of Rocks, A Workbook. Handbook of Petroleum Exploration and Production; Schön, J.H., Ed.; Elsevier: Amsterdam, The Netherlands, 2011; Volume, 8, pp. 149–243. [Google Scholar]
  179. Alonso, E.E.; Vaunat, J.; Gens, A. Modelling the mechanical behaviour of expansive clays. Eng. Geol. 1999, 54, 173–183. [Google Scholar] [CrossRef]
  180. Åkesson, A.; Kristensson, O. Mechanical modeling of MX-80–Development of constitutive laws. Phys. Chem. Earth 2008, 33, S504–S507. [Google Scholar] [CrossRef]
  181. Tisato, N.; Marelli, S. Laboratory measurements of the longitudinal and transverse wave velocities of compacted bentonite as a function of water content, temperature, and confining pressure. J. Geophys. Res. Solid Earth 2013, 118, 3380–3393. [Google Scholar] [CrossRef]
  182. Keller, L.M. Porosity anisotropy of Opalinus Clay: Implications for the poroelastic behaviour. Geophys. J. Int. 2017, 208, 1443–1448. [Google Scholar] [CrossRef]
  183. Kenigsberg, A.R.; Rivière, J.; Marone, C.; Saffer, D.M. Evolution of Elastic and Mechanical Properties during Fault Shear: The Roles of Clay Content, Fabric Development, and Porosity. J. Geophys. Res. Solid Earth 2019, 124, 10968–10982. [Google Scholar] [CrossRef]
  184. Kenigsberg, A.R.; Rivière, J.; Marone, C.; Saffer, D.M. The Effects of Shear Strain, Fabric, and Porosity Evolution on Elastic and Mechanical Properties of Clay-Rich Fault Gouge. J. Geophys. Res. Solid Earth 2020, 125, e2019JB018612. [Google Scholar] [CrossRef]
  185. Liu, K.; Sheng, J.; Zhang, Z. A simulation study of the effect of clay swelling on fracture generation and porosity change in shales under stress anisotropy. Eng. Geol. 2020, 278, 105829. [Google Scholar] [CrossRef]
  186. Sveinsson, H.A.; Ning, F.; Cao, P.; Fang, B.; Malthe-Sørenssen, A. Grain-Size-Governed Shear Failure Mechanism of Polycrystalline Methane Hydrates. J. Phys. Chem. C 2021, 125, 10034–10042. [Google Scholar] [CrossRef]
  187. Yu, C.; Ji, S.; Li, Q. Effects of porosity on seismic velocities, elastic moduli and Poisson’s ratios of solid materials and rocks. J. Rock Mech. Geotech. Eng. 2016, 8, 35–49. [Google Scholar] [CrossRef]
  188. Brantut, N.; Stefanou, I.; Sulem, J. Dehydration-induced instabilities at intermediate depths in subduction zones. J. Geophys. Res. Solid Earth 2017, 122, 6087–6107. [Google Scholar] [CrossRef]
  189. Born, M. On the Stability of Crystal Lattices. Math. Proc. Camb. Phil. Soc. 1940, 36, 160–172. [Google Scholar] [CrossRef]
  190. Milstein, F. Theoretical elastic behaviour of crystals at large strains. J. Mater. Sci. 1980, 15, 1071–1084. [Google Scholar] [CrossRef]
  191. Karki, B.B.; Ackland, G.J.; Crain, J. Elastic instabilities in crystals from ab initio stress–strain relations. J. Phys. Condens. Matter 1997, 9, 8579–8589. [Google Scholar] [CrossRef]
  192. Mouhat, F.; Coudert, F.X. Necessary and Sufficient Elastic Stability Conditions in Various Crystal Systems. Phys. Rev. B 2014, 90, 224104. [Google Scholar] [CrossRef]
  193. Gao, F. Hardness estimation of complex oxide materials. Phys. Rev. B 2004, 69, 094113. [Google Scholar] [CrossRef]
  194. Šimůnek, A.; Vackář, J. Hardness of Covalent and Ionic Crystals: First-Principle Calculations. Phys. Rev. Lett. 2006, 96, 085501. [Google Scholar] [CrossRef] [PubMed]
  195. Niu, H.; Wei, P.; Sun, Y.; Chen, C.X.; Franchini, C.; Li, D.; Li, Y. Electronic, optical, and mechanical properties of superhard cold-compressed phases of carbon. Appl. Phys. Lett. 2011, 99, 031901. [Google Scholar] [CrossRef]
  196. Chen, X.Q.; Niu, H.; Li, D.; Li, Y. Modeling Hardness of Polycrystalline Materials and Bulk Metallic Glasses. Intermetallics 2011, 19, 1275–1281. [Google Scholar] [CrossRef]
  197. Liu, X.; Wang, H.; Wang, W.; Fu, Z. Simple Method for the Hardness Estimation of Inorganic Crystals by the Bond Valence Model. Inorg. Chem. 2016, 55, 11089–11095. [Google Scholar] [CrossRef]
  198. Pugh, S.F. Relations between the Elastic Moduli and the Plastic Properties of Polycrystalline Pure Metals. Philos. Mag. 1954, 45, 823–843. [Google Scholar] [CrossRef]
  199. Pettifor, D.G. Theoretical predictions of structure and related properties of intermetallics. Mater. Sci. Technol. 1992, 8, 345–349. [Google Scholar] [CrossRef]
  200. Niu, H.; Chen, X.Q.; Liu, P.; Xing, W.; Cheng, X.; Li, D.; Li, Y. Extra-electron induced covalent strengthening and generalization of intrinsic ductile-to-brittle criterion. Sci. Rep. 2012, 2, 718. [Google Scholar] [CrossRef]
  201. Bouhadda, Y.; Djella, S.; Bououdina, M.; Fenineche, N.; Boudouma, Y. Structural and Elastic Properties of LiBH4 for Hydrogen Storage Applications. J. Alloys Compd. 2012, 534, 20–24. [Google Scholar] [CrossRef]
  202. Gschneidner, K.; Russell, A.; Pecharsky, A.; Morris, J.; Zhang, Z.; Lograsso, T.; Hsu, D.; Lo, H.C.; Ye, Y.; Slager, A.; et al. A family of ductile intermetallic compounds. Nat. Mater. 2003, 2, 587–591. [Google Scholar] [CrossRef]
  203. Ritchie, R.O. The conflicts between strength and toughness. Nat. Mater. 2011, 10, 817–822. [Google Scholar] [CrossRef] [PubMed]
  204. Hwang, E.; Cuddy, E.; Lin, J.; Kaufman, J.L.; Shaw, A.; Conway, P.L.J.; Pribram-Jones, A.; Laws, K.J.; Bassman, L. Predicting ductility in quaternary -like alloys. Phys. Rev. Mater. 2021, 5, 033604. [Google Scholar] [CrossRef]
  205. Zener, C. Elasticity and Anelasticity of Metals; University of Chicago Press: Chicago, IL, USA, 1948. [Google Scholar]
  206. Chung, D.H.; Buessem, W.R. The Elastic Anisotropy of Crystals. J. Appl. Phys. 1967, 38, 2010–2012. [Google Scholar] [CrossRef]
  207. Tvergaard, V.; Hutchinson, J.H. Microcracking in ceramics induced by thermal expansion or elastic anisotropy. J. Am. Ceram. Soc. 1988, 71, 157–166. [Google Scholar] [CrossRef]
  208. Ravindran, P.; Fast, L.; Korzhavyi, P.A.; Johansson, B.; Wills, J.; Eriksson, O. Density Functional Theory for Calculation of Elastic Properties of Orthorhombic Crystals: Application to TiSi2. J. Appl. Phys. 1997, 84, 4891–4904. [Google Scholar] [CrossRef]
  209. Ledbetter, H.; Migliori, A.A. A general elastic-anisotropy measure. J. Appl. Phys. 2006, 100, 063516. [Google Scholar] [CrossRef]
  210. Lloveras, P.; Castan, T.; Porta, M.; Planes, A.; Saxena, A. Influence of Elastic Anisotropy on Structural Nanoscale Textures. Phys. Rev. Lett. 2008, 100, 165707. [Google Scholar] [CrossRef]
  211. Kube, C.M. Elastic anisotropy of crystals. AIP Adv. 2016, 6, 095209. [Google Scholar] [CrossRef]
  212. Ranganathan, S.I.; Ostoja-Starzewski, M. Universal Elastic Anisotropy Index. Phys. Rev. Lett. 2008, 101, 055504. [Google Scholar] [CrossRef]
  213. Curtin, W.A. Theory of Mechanical Properties of Ceramic-Matrix Composites. J. Am. Ceram. Soc. 1991, 74, 2837–2845. [Google Scholar] [CrossRef]
  214. Coleman, J.N.; Khan, U.; Blau, W.J.; Gun’ko, Y.K. Small but strong: A review of the mechanical properties of carbon nanotube–polymer composites. Carbon 2006, 44, 1624–1652. [Google Scholar] [CrossRef]
  215. Salvetat, J.P.; Briggs, G.A.D.; Bonard, J.M.; Bacsa, R.R.; Kulik, A.J.; Stöckli, T.; Burnham, N.A.; Forró, L. lastic and Shear Moduli of Single-Walled Carbon Nanotube Ropes. Phys. Rev. Lett. 1999, 82, 944–947. [Google Scholar] [CrossRef]
  216. Kis, A.; Csányi, G.; Salvetat, J.P.; Lee, T.N.; Couteau, E.; Kulik, A.J.; Benoit, W.; Brugger, J.; Forró, L. Reinforcement of single-walled carbon nanotube bundles by intertube bridging. Nat. Mater. 2004, 3, 153–157. [Google Scholar] [CrossRef] [PubMed]
  217. Beyer, T.; Day, G.M.; Price, S.L. The prediction, morphology, and mechanical properties of the polymorphs of paracetamol. J. Am. Chem. Soc. 2001, 123, 5086–5094. [Google Scholar] [PubMed]
  218. Reddy, C.M.; Basavoju, S.; Desiraju, G.R. Structural basis for bending of organic crystals. Chem. Commun. 2005, 2005, 2439–2441. [Google Scholar] [CrossRef] [PubMed]
  219. Reddy, C.M.; Krishna, G.R.; Ghosh, S. Mechanical properties of molecular crystals—Applications to crystal engineering. CrystEngComm 2010, 12, 2296–2314. [Google Scholar] [CrossRef]
  220. Raju, K.B.; Ranjan, S.; Vishnu, V.S.; Bhattacharya, M.; Bhattacharya, B.; Mukhopadhyay, A.K.; Reddy, C.M. Rationalizing Distinct Mechanical Properties of Three Polymorphs of a Drug Adduct by Nanoindentation and Energy Frameworks Analysis: Role of Slip Layer Topology and Weak Interactions. Cryst. Growth Des. 2018, 8, 3927–3937. [Google Scholar] [CrossRef]
  221. Fabbiani, F.P.A.; Pulham, C.R. High-pressure studies of pharmaceutical compounds and energetic materials. Chem. Soc. Rev. 2006, 35, 932–942. [Google Scholar] [CrossRef]
  222. Fabbiani, F.P.A.; Allan, D.R.; David, W.I.F.; Davidson, A.J.; Lennie, A.R.; Parsons, S.; Pulham, C.R.; Warren, J.E. High-pressure studies of pharmaceuticals: An exploration of the behavior of piracetam. Cryst. Growth Des. 2007, 7, 1115–1124. [Google Scholar] [CrossRef]
  223. Neumann, M.A.; Van de Streek, J.; Fabbiani, F.P.A.; Hidber, P.; Grassmann, O. Combined crystal structure prediction and high-pressure crystallization in rational pharmaceutical polymorph screening. Nat. Commun. 2015, 6, 7793. [Google Scholar] [CrossRef]
  224. Meier, M.; John, E.; Wieckhusen, D.; Wirth, W.; Peukert, W. Influence of mechanical properties on impact fracture: Prediction of the milling behaviour of pharmaceutical powders by nanoindentation. Powder Technol. 2009, 188, 301–313. [Google Scholar] [CrossRef]
  225. Karki, S.; Friščić, T.; Fábián, L.; Laity, P.R.; Day, G.M.; Jones, W. Improving Mechanical Properties of Crystalline Solids by Cocrystal Formation: New Compressible Forms of Paracetamol. Adv. Mater. 2009, 21, 3905–3909. [Google Scholar] [CrossRef]
  226. Varughese, S.; Kiran, M.S.R.N.; Solanko, K.A.; Bond, A.D.; Ramamurty, U.; Desiraju, G.R. Interaction anisotropy and shear instability of aspirin polymorphs established by nanoindentation. Chem. Sci. 2011, 2, 2236–2242. [Google Scholar] [CrossRef]
  227. SeethaLekshmi, S.; Kiran, M.S.R.N.; Ramamurty, U.; Varughese, S. Molecular basis for the mechanical response of sulfa drug crystals. Chem. Eur. J. 2019, 25, 526–537. [Google Scholar] [CrossRef] [PubMed]
  228. Egar, M.; Janković, B.; Lah, N.; Ilić, I.; Srčič, S. Nanomechanical properties of selected single pharmaceutical crystals as a predictor of their bulk behaviour. Pharm. Res. 2015, 3, 469–481. [Google Scholar] [CrossRef] [PubMed]
  229. Azuri, I.; Meirzadeh, E.; Here, D.; Cohen, S.R.; Rappe, A.; Lahav, M.; Lubomirsky, I.; Kronik, L. Unusually Large Young’s Moduli of Amino Acid Molecular Crystals. Angew. Chem. Int. Ed. 2015, 54, 13566–13570. [Google Scholar] [CrossRef]
  230. Rupasinghe, T.M.; Hutchins, K.M.; Bandaranayake, B.S.; Ghorai, S.; Karunatilake, C.; Bučar, D.K.; Swenson, D.C.; Arnold, M.A.; MacGillivray, L.R.; Tivanski, A.V. Mechanical Properties of a Series of Macro- and Nanodimensional Organic Cocrystals Correlate with Atomic Polarizability. J. Am. Chem. Soc. 2015, 137, 12768–12771. [Google Scholar] [CrossRef]
  231. Mohamed, M.R.; Mishra, M.K.; Al Harbi, L.M.; Al Ghamdic, M.S.; Ramamurty, U. Anisotropy in the mechanical properties of organic crystals: Temperature dependence. RSC Adv. 2015, 5, 64156–64162. [Google Scholar] [CrossRef]
  232. Mishra, M.K.; Mishra, K.; Narayan, A.; Reddy, C.M.; Vangala, V.R. Structural Basis for Mechanical Anisotropy in Polymorphs of a Caffeine–Glutaric Acid Cocrystal. Cryst. Growth Des. 2020, 20, 6306–6315. [Google Scholar] [CrossRef]
  233. Mishra, M.K.; Sun, C.C. Conformation Directed Interaction Anisotropy Leading to Distinct Bending Behaviors of Two ROY Polymorphs. Cryst. Growth Des. 2020, 20, 4764–4769. [Google Scholar] [CrossRef]
  234. Jain, A.; Shah, H.S.; Johnson, P.R.; Narang, A.S.; Morris, K.R.; Haware, R.V. Crystal anisotropy explains structure-mechanics impact on tableting performance of flufenamic acid polymorphs. Eur. J. Pharm. Biopharm. 2018, 132, 83–92. [Google Scholar] [CrossRef] [PubMed]
  235. Singaraju, A.B.; Bahl, D.; Wang, C.; Swenson, D.C.; Sun, C.C.; Stevens, L.L. Molecular Interpretation of the Compaction Performance and Mechanical Properties of Caffeine Cocrystals: A Polymorphic Study. Mol. Pharmaceutics 2020, 17, 21–31. [Google Scholar] [CrossRef] [PubMed]
  236. Zhang, K.; Sun, C.C.; Liu, Y.; Wang, C.; Shi, P.; Xu, J.; Wu, S.; Gong, J. Fabrication of a microcapsule extinguishing agent with a core–shell structure for lithium-ion battery fire safety. Chem. Mater. 2021, 33, 1053–1060. [Google Scholar] [CrossRef]
  237. Vaksler, Y.; Idrissi, A.; Urzhuntseva, V.V.; Shishkina, S.V. Quantum Chemical Modeling of Mechanical Properties of Aspirin Polymorphic Modifications. Cryst. Growth Des. 2021, 21, 2176–2186. [Google Scholar] [CrossRef]
  238. Wang, L.; Wang, C.; Wu, S.; Fan, Y.; Li, X. Influence of the mechanical properties of biomaterials on degradability, cell behaviors and signaling pathways: Current progress and challenges. Biomater. Sci. 2020, 8, 2714–2733. [Google Scholar] [CrossRef]
  239. Mitragotri, M.; Llahann, J. Physical approaches to biomaterial design. Nat. Mater. 2009, 8, 15–23. [Google Scholar] [CrossRef]
  240. Jahan, A.; Ismail, M.Y.; Sapuan, S.; Mustapha, F. Material screening and choosing methods—A review. Mater. Des. 2010, 31, 696–705. [Google Scholar] [CrossRef]
  241. Li, Z.; Nevitt, M.N.; Ghose, S. Elastic constants of sodalite Na4Al3Si3O12Cl. Appl. Phys. Lett. 1989, 55, 1730–1731. [Google Scholar] [CrossRef]
  242. Lethbridge, Z.A.D.; Walton, R.I.; Bosak, A.; Krisch, M. Single-crystal elastic constants of the zeolite analcime measured by inelastic X-ray scattering. Chem. Phys. Lett. 2009, 471, 286–289. [Google Scholar] [CrossRef]
  243. Williams, J.J.; Evans, K.E.; Walton, R.I. On the elastic constants of the zeolite chlorosodalite. Appl. Phys. Lett. 2006, 88, 021914. [Google Scholar] [CrossRef]
  244. Birch, F. The Velocity of Compressional Waves in Rocks to 10 Kilobars, Part 1. J. Geophys Res. 1960, 65, 1083–1102. [Google Scholar] [CrossRef]
  245. Neighbours, J.R.; Schacher, G.E. Determination of Elastic Constants from Sound-Velocity Measurements in Crystals of General Symmetry. J. Appl. Phys. 1967, 38, 5366–5375. [Google Scholar] [CrossRef]
  246. Christensen, N.I. Poisson′s ratio and crustal seismology. J. Geophys. Res. 1996, 101, 3139–3156. [Google Scholar] [CrossRef]
  247. Li, B.; Liebermann, R.C. Study of the Earth’s Interior Using Measurements of Sound Velocities in Minerals by Ultrasonic Interferometry. Phys. Earth Planet. Int. 2014, 233, 135–153. [Google Scholar] [CrossRef]
  248. Bosak, A.; Serrano, J.; Krisch, M.; Watanabe, K.; Taniguchi, T.; Kanda, H. Elasticity of hexagonal boron nitride: Inelastic x-ray scattering measurements. Phys. Rev. B 2006, 73, 041402. [Google Scholar] [CrossRef]
  249. Bosak, A.; Krisch, M.; Mohr, M.; Maultzsch, J.; Thomsen, C. Elasticity of single-crystalline graphite: Inelastic x-ray scattering study. Phys. Rev. B 2007, 75, 153408. [Google Scholar] [CrossRef]
  250. Diddens, I.; Murphy, B.; Krisch, M.; Müller, M. Anisotropic Elastic Properties of Cellulose Measured Using Inelastic X-ray Scattering. Macromolecules 2008, 41, 9755–9759. [Google Scholar] [CrossRef]
  251. Kiefte, H.3; Breckon, S.; Penney, R.; Clouter, M. Elastic constants of ammonia by Brillouin spectroscopy. J. Chem. Phys. 1985, 83, 4738–4743. [Google Scholar] [CrossRef]
  252. Bass, J.D. Elasticity of grossular and spessartite garnets by Brillouin sepctroscopy. J. Geophys. Res. 1989, 94, 7621–7628. [Google Scholar] [CrossRef]
  253. Bezacier, L.; Reynard, B.; Cardon, H.; Montagnac, G.; Bass, J.D. High-pressure elasticity of serpentine and seismic properties of the hydrated mantle wedge. J. Geophys. Res. Solid Earth 2013, 118, 527–535. [Google Scholar] [CrossRef]
  254. Oliver, W.C.; Pharr, G.M. An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. J. Mater. Res. 2004, 19, 3–20. [Google Scholar] [CrossRef]
  255. Lin, J.; Shu, X.F.; Dong, J.X. The synthesis and mechanical properties of large zeolite sodalite single crystals. Stud. Surf. Sci. Catal. A 2005, 158, 231–238. [Google Scholar]
  256. Brabec, L.; Bohac, P.; Stranyanek, M.; Ctvrtlik, R.; Kocirik, M. Hardness and elastic modulus of silicalite-1 crystal twins. Microporous Mesoporous Mater. 2006, 94, 226–233. [Google Scholar] [CrossRef]
  257. Johnson, M.C.; Wang, J.; Li, Z.; Lew, C.M.; Yan, Y. Effect of calcination and polycrystallinity on mechanical properties of nanoporous MFI zeolites. Mater. Sci. Eng. A 2007, 456, 58–63. [Google Scholar] [CrossRef]
  258. Eslava, S.; Zhang, L.; Esconjauregui, S.; Yang, J.; Baklanov, K.M.; Saiz, M.R. Metal-organic framework ZIF-8 films as low-κ dielectrics in microelectronics. Chem. Mater. 2013, 25, 27–33. [Google Scholar] [CrossRef]
  259. Zeng, Z.; Tan, J.C. AFM Nanoindentation to Quantify Mechanical Properties of Nano- and Micron-Sized Crystals of a Metal-Organic Framework Material. ACS Appl. Mater. Interfaces 2017, 9, 39839–39854. [Google Scholar] [CrossRef]
  260. Boissiere, C.; Grosso, D.; Lepoutre, S.; Nicole, L.; Bruneau, A.B.; Sanchez, C. Porosity and mechanical properties of mesoporous thin films assessed by environmental ellipsometric porosimetry. Langmuir 2005, 21, 12362–12371. [Google Scholar] [CrossRef] [PubMed]
  261. Devine, S.D.; Robinson, W.H. Piezoelectric method of determining mechanical properties of a small sandwich specimen at 30 to 200 kHz. J. Appl. Phys. 1977, 48, 1437. [Google Scholar] [CrossRef]
  262. Woldenden, A.; Harmouche, M.R. Elastic constants of silver as a function of temperature. J. Mater. Sci. 1993, 28, 1015–1018. [Google Scholar] [CrossRef]
  263. Ganesan, V.V.; Dhanasekaran, M.; Thangavel, N.; Dhathathreyan, A. Elastic compliance of fibrillar assemblies in type I collagen. Biophys. Chem. 2018, 240, 15–24. [Google Scholar] [CrossRef]
  264. Kiely, E.; Zwane, R.; Fox, R.; Reilly, A.M.; Guerin, S. Density functional theory predictions of the mechanical properties of crystalline materials. CrystEngComm 2021, 23, 5697–5710. [Google Scholar] [CrossRef]
  265. Chaudhuri, M.M. The deformation stress of highly brittle explosive crystals from real contact area measurements. J. Mater. Sci. 1984, 19, 3028–3042. [Google Scholar] [CrossRef]
  266. Wang, Z.; Lambros, J.; Lobo, R.F. Micromechanical compressive response of a zeolite single crystal. J. Mater. Sci. 2002, 37, 2491–2499. [Google Scholar] [CrossRef]
  267. Singh, A.K. X-ray diffraction from solids under nonhydrostatic compression—Some recent studies. J. Phys. Chem. Solids 2004, 65, 1589–1596. [Google Scholar] [CrossRef]
  268. Singh, A.K.; Andrault, D.; Bouvier, P. X-ray diffraction from stishovite under nonhydrostatic compression to 70 GPa: Strength and elasticity across the tetragonal → orthorhombic transition. Phys. Earth Planet. Inter. 2012, 208–209, 1–10. [Google Scholar] [CrossRef]
  269. Duwal, S.; Yoo, C.S. Shear-Induced Isostructural Phase Transition and Metallization of Layered Tungsten Disulfide under Nonhydrostatic Compression. J. Phys. Chem. C 2016, 120, 5101–5107. [Google Scholar] [CrossRef]
  270. Liu, B.; Lin, L.; Gao, Y.; Ma, Y.; Zhou, P.; Han, D.; Gao, C. Metallization of Molybdenum Diselenide under Nonhydrostatic Compression. J. Phys. Chem. C 2021, 125, 5412–5416. [Google Scholar] [CrossRef]
  271. Day, G.M.; Price, S.L.; Leslie, M. Elastic Constant Calculations for Molecular Organic Crystals. Cryst. Growth Des. 2001, 1, 13–27. [Google Scholar] [CrossRef]
  272. Han, S.S.; Goddard, W.A., III. Metal-Organic Frameworks Provide Large Negative Thermal Expansion Behavior. J. Phys. Chem. C 2007, 11, 15185–15191. [Google Scholar] [CrossRef]
  273. Wan, C.; Sun, C.C. Superior Plasticity and Tabletability of Theophylline Monohydrate. Mol. Pharm. 2019, 16, 1732–1741. [Google Scholar]
  274. Stixrude, L.; Cohen, R.E.; Singh, D.J. Iron at high pressure: Linearized-augmented-plane-wave computations in the generalized-gradient approximation. Phys. Rev. B 1994, 50, 6442–6445. [Google Scholar] [CrossRef] [PubMed]
  275. Karki, B.; Stixrude, L.; Wentzcovitch, R.M. High-pressure elastic properties of major materials of Earth’s mantle from first principles. Rev. Geophys. 2001, 39, 507–534. [Google Scholar] [CrossRef]
  276. 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] [PubMed]
  277. Chibani, S.; Coudert, F.X. Systematic exploration of the mechanical properties of 13 621 inorganic compounds. Chem. Sci. 2019, 10, 8589–8599. [Google Scholar] [CrossRef]
  278. Payne, M.C.; Teter, M.P.; Ailan, D.C.; Arias, A.; Joannopoulos, J.D. terative Minimization Techniques for ab Initio Total-Energy Calculations: Molecular Dynamics and Conjugate Gradients. Rev. Mod. Phys. 1992, 64, 1045–1097. [Google Scholar] [CrossRef]
  279. MaterialsStudio. Available online: https://3dsbiovia.com/products/collabo-rative-science/biovia-materials-studio/ (accessed on 15 June 2021).
  280. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Kristallogr. 2005, 220, 567–570. [Google Scholar] [CrossRef]
  281. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  282. Perdew, J.P.; Ruzsinszky, A.; Csonka, G.I.; Vydrov, O.A.; Scuseria, G.E.; Constantin, L.A.; Zhou, X.; Burke, K. Restoring the Density-Gradient Expansion for Exchange in Solids and Surfaces. Phys. Rev. Lett. 2008, 100, 136406. [Google Scholar] [CrossRef]
  283. Grimme, S. Semiempirical GGA-type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  284. Bonales, L.J.; Colmenero, F.; Cobos, J.; Timón, V. Spectroscopic Raman characterization of rutherfordine: A combined DFT and experimental study. Phys. Chem. Chem. Phys. 2016, 18, 16575–16584. [Google Scholar] [CrossRef]
  285. Colmenero, F.; Bonales, L.J.; Cobos, J.; Timón, V. Density Functional Theory Study of the Thermodynamic and Raman Vibrational Properties of γ-UO3 Polymorph. J. Phys. Chem. C 2017, 121, 14507–14516. [Google Scholar] [CrossRef]
  286. Colmenero, F.; Bonales, L.J.; Cobos, J.; Timón, V. Thermodynamic and Mechanical Properties of the Rutherfordine Mineral Based on Density Functional Theory.J. Phys. Chem. C 2017, 121, 5994–6001. [Google Scholar] [CrossRef]
  287. Weck, P.F.; Gordon, M.; Greathouse, J.A.; Bryan, C.E.; Meserole, S.P.; Rodriguez, M.A.; Payne, M.C.; Kim, E.J. Infrared and Raman Spectroscopy of α-ZrW2O8: A Comprehensive Density Functional Perturbation Theory and Experimental Study. J. Raman Spectrosc. 2018, 49, 1373–1384. [Google Scholar] [CrossRef]
  288. Colmenero, F.; Plášil, J.; Sejkora, J. The crystal structures and mechanical properties of the uranyl carbonate minerals roubaultite, fontanite, sharpite, widenmannite, grimselite and čejkaite. Inorg. Chem. Front. 2020, 7, 4197–4221. [Google Scholar] [CrossRef]
  289. Colmenero, F. Thermodynamic properties of the uranyl carbonate minerals roubaultite, fontanite, widenmannite, grimselite, čejkaite and bayleyite. Inorg. Chem. Front. 2020, 7, 4160–4179. [Google Scholar] [CrossRef]
  290. Colmenero, F.; Timón, V. Mechanical anomalies in mercury oxalate and the deformation of the mercury cube coordination environment under pressure. Appl. Phys. A 2021, 127, 395. [Google Scholar] [CrossRef]
  291. Troullier, N.; Martins, J.L. Efficient Pseudopotentials for Plane-Wave Calculations. Phys. Rev. B 1991, 43, 1993–2006. [Google Scholar] [CrossRef]
  292. Pfrommer, B.G.; Cote, M.; Louie, S.G.; Cohen, M.L. Relaxation of Crystals with the Quasi-Newton Method. J. Comput. Phys. 1997, 131, 233–240. [Google Scholar] [CrossRef]
  293. Downs, R.T.; Bartelmehs, K.L.; Gibbs, G.V.; Boisen, M.B. Interactive software for calculating and displaying X-ray or neutron powder diffractometer patterns of crystalline materials. Am. Mineral. 1993, 78, 1104–1107. [Google Scholar]
  294. Nye, J.F. Physical Properties of Crystals; Clarendon: Oxford, UK, 1976. [Google Scholar]
  295. Yu, R.; Zhu, J.; Ye, H.Q. Calculations of Single-Crystal Elastic Constants Made Simple. Comput. Phys. Commun. 2010, 181, 671–675. [Google Scholar] [CrossRef]
  296. Nielsen, O.H.; Martin, R.M. Quantum-mechanical theory of stress and force. Phys. Rev. B 1985, 32, 3780–3791. [Google Scholar] [CrossRef] [PubMed]
  297. Colmenero, F.; Bonales, L.J.; Cobos, J.; Timón, V. Structural, mechanical and vibrational study of uranyl silicate mineral soddyite by DFT calculations. J. Solid. State Chem. 2017, 253, 249–257. [Google Scholar] [CrossRef]
  298. Colmenero, F.; Bonales, L.J.; Timón, V.; Cobos, J. Structural, mechanical and Raman spectroscopic characterization of the layered uranyl silicate mineral, uranophane-α, by density functional theory methods. Clay Miner. 2018, 53, 377–392. [Google Scholar] [CrossRef]
  299. Colmenero, F.; Cobos, J.; Timón, V. Periodic Density Functional Theory Study of the Structure, Raman Spectrum, and Mechanical Properties of Schoepite Mineral. Inorg. Chem. 2018, 57, 4470–4481. [Google Scholar] [CrossRef] [PubMed]
  300. Colmenero, F.; Fernández, A.M.; Cobos, J.; Timón, V. Becquerelite mineral phase: Crystal structure and thermodynamic and mechanical stability by using periodic DFT. RSC Adv. 2018, 8, 24599–24616. [Google Scholar] [CrossRef]
  301. Colmenero, F.; Plášil, J.; Sejkora, J. The layered uranyl silicate mineral uranophane-β: Crystal structure, mechanical properties, Raman spectrum and comparison with the α-polymorph. Dalton Trans. 2018, 48, 16722–16736. [Google Scholar] [CrossRef]
  302. Colmenero, F.; Plášil, J.; Cobos, J.; Sejkora, J.; Timón, V.; Čejka, J.; Bonales, L.J. Crystal structure, hydrogen bonding, mechanical properties and Raman spectrum of the lead uranyl silicate monohydrate mineral kasolite. RSC Adv. 2019, 9, 15323–15334. [Google Scholar] [CrossRef]
  303. Colmenero, F.; Plášil, J.; Cobos, J.; Sejkora, J.; Timón, V.; Čejka, J.; Fernández, A.M.; Petříček, V. Structural, mechanical, spectroscopic and thermodynamic characterization of the copper-uranyl tetrahydroxide mineral vandenbrandeite. RSC Adv. 2019, 9, 40708–40726. [Google Scholar] [CrossRef]
  304. Colmenero, F.; Plášil, J.; Němec, I. Uranosphaerite: Crystal structure, hydrogen bonding, mechanics, infrared and Raman spectroscopy and thermodynamics. J. Phys. Chem. Solids 2020, 141, 109400. [Google Scholar] [CrossRef]
  305. Colmenero, F.; Plášil, J.; Škácha, P. The magnesium uranyl tricarbonate octadecahydrate mineral, bayleyite: Periodic DFT study of its crystal structure, hydrogen bonding, mechanical properties and infrared spectrum. Spectrochim. Acta A 2020, 234, 118216. [Google Scholar] [CrossRef]
  306. Colmenero, F.; Plášil, J.; Timón, V.; Čejka, J. Full crystal structure, hydrogen bonding and spectroscopic, mechanical and thermodynamic properties of mineral uranopilite. RSC Adv. 2020, 10, 31947–31960. [Google Scholar] [CrossRef]
  307. Colmenero, F. Anomalous mechanical behavior of the deltic, squaric and croconic cyclic oxocarbon acids. Mater. Res. Express 2019, 6, 045610. [Google Scholar] [CrossRef]
  308. Colmenero, F. Mechanical properties of anhydrous oxalic acid and oxalic acid dihydrate. Phys. Chem. Chem. Phys. 2019, 21, 2673–2690. [Google Scholar] [CrossRef] [PubMed]
  309. Colmenero, F. Negative area compressibility in oxalic acid dihydrate. Mater. Lett. 2019, 245, 25–28. [Google Scholar] [CrossRef]
  310. Colmenero, F. Organic acids under pressure: Elastic properties, negative mechanical phenomena and pressure induced phase transitions in the lactic, maleic, succinic and citric acids. Mater. Adv. 2020, 1, 1399–1426. [Google Scholar] [CrossRef]
  311. Colmenero, F.; Cobos, J.; Timón, V. Negative linear compressibility in uranyl squarate monohydrate. J. Phys. Cond. Matter. 2019, 31, 175701. [Google Scholar] [CrossRef]
  312. Colmenero, F.; Timón, V. Extreme negative mechanical phenomena in the zinc and cadmium anhydrous metal oxalates and lead oxalate dihydrate. J. Mater. Sci. 2020, 55, 218–236. [Google Scholar] [CrossRef]
  313. Colmenero, F. Silver oxalate: Mechanical properties and extreme negative mechanical phenomena. Adv. Theor. Simul. 2019, 2, 1900040. [Google Scholar] [CrossRef]
  314. Colmenero, F.; Jiang, X.; Li, X.; Li, Y.; Lin, Z. Negative area compressibility in silver oxalate. J. Mater. Sci. 2021, 56, 269–277. [Google Scholar] [CrossRef]
  315. Baroni, S.; Giannozzi, P.; Testa, A. Elastic Constants of Crystals from Linear-Response Theory. Phys. Rev. Lett. 1987, 59, 2662–2665. [Google Scholar] [CrossRef]
  316. Wu, X.; Vanderbilt, D.; Hamann, D.R. Systematic treatment of displacements, strains, and electric fields in density-functional perturbation theory. Phys. Rev. B 2005, 72, 035105. [Google Scholar] [CrossRef]
  317. Wentzcovitch, R.M.; Wu, Z.Q.; Carrier, P. Thermodynamic properties and phase relations in mantle minerals investigated by first principles quasiharmonic theory. Rev. Mineral. Geochem. 2010, 71, 99–128. [Google Scholar] [CrossRef]
  318. Wu, Z.Q.; Wentzcovitch, R.M. Quasiharmonic thermal elasticity of crystals: An analytical approach. Phys. Rev. B 2011, 83, 184115. [Google Scholar] [CrossRef]
  319. Parrinello, J.; Rahman, A. Strain fluctuations and elastic constants. J. Chem. Phys. 1982, 76, 2662–2666. [Google Scholar] [CrossRef]
  320. Ray, J.R. Comput. Ensembles and Computer Simulation Calculation of Response Functions. Phys. Rep. 1988, 8, 109–152. [Google Scholar]
  321. Wojciechowski, K.W.; Tretiakov, K.V. Quick and accurate estimation of the elastic constants using the minimum image method. Comp. Phys. Commun. 1999, 121–122, 528–530. [Google Scholar] [CrossRef]
  322. Van Workum, K.V.; De Pablo, J. Local elastic constants in thin films of an fcc cristal. Phys. Rev. E 2003, 67, 011505. [Google Scholar] [CrossRef]
  323. Voyiatzis, E. Mechanical properties and elastic constants of atomistic systems through the stress-fluctuation formalism. Comput. Phys. Commun. 2013, 184, 27–33. [Google Scholar] [CrossRef]
  324. Marmier, A.; Lethbridge, Z.A.D.; Walton, R.I.; Smith, C.; Parker, S.C.; Evans, K.E. ElAM: A computer program for the analysis and representation of anisotropic elastic properties. Comput. Phys. Commun. 2010, 181, 2102–2115. [Google Scholar] [CrossRef]
  325. Lowenstein, W. The distribution of aluminum in the tetrahedra of silicates and aluminates. Am. Mineral. 1954, 39, 91–96. [Google Scholar]
  326. Davis, M.E.; Hathaway, P.E.; Montes, C. Zeolites and molecular sieves: Not just ordinary catalysts. Zeolites 1980, 9, 436–439. [Google Scholar] [CrossRef]
  327. Voigt, W. Lehrbuch der Kristallphysik; Teubner: Leipzig, Germany, 1962. [Google Scholar]
  328. Reuss, A.Z. Berechnung der Fließgrenze von Mischkristallen auf Grund der Plastizitätsbedingung für Einkristalle. Angew. Math. Mech. 1929, 9, 49–58. [Google Scholar] [CrossRef]
  329. Hill, R. The Elastic Behaviour of a Crystalline Aggregate. Proc. Phys. Soc. Lond. A 1952, 65, 349–354. [Google Scholar] [CrossRef]
  330. Weck, P.F.; Kim, E.; Buck, E.C. On the mechanical stability of uranyl peroxide hydrates: Implications for nuclear fuel degradation. RSC Adv. 2015, 5, 79090–79097. [Google Scholar] [CrossRef]
  331. Colmenero Ruiz, F. Theoretical Studies of the Structural, Mechanic and Raman Spectroscopic Properties of Uranyl Containing Minerals. In Minerals; Essa, K.S., Ed.; InTechOpen: London, UK, 2018; Chapter 4; pp. 65–94. [Google Scholar]
  332. Angel, R.J. Equations of State. Rev. Mineral. Geochem. 2000, 41, 35–60. [Google Scholar] [CrossRef]
  333. Lacivita, V.; D’Arco, P.; Mustapha, S.; Bernardes, D.F.; Dovesi, R.; Erba, A.; Rérat., M. Ab initio compressibility of metastable low albite: Revealing a lambda-type singularity at pressures of the Earth’s upper mantle. Phys. Chem. Miner. 2020, 47, 45. [Google Scholar] [CrossRef]
  334. Evans, K.E. Auxetic polymers: A new range of materials. Endeavour 1991, 15, 170–174. [Google Scholar] [CrossRef]
  335. Colmenero, F.; Sejkora, J.; Plášil, J. Crystal Structure, Infrared Spectrum and Elastic Anomalies in Tuperssuatsiaite. Sci. Rep. 2020, 10, 7510. [Google Scholar] [CrossRef]
  336. Baughman, R.H.; Fonseca, A.F. Straining to expand entanglements. Nat. Mater. 2015, 15, 7–8. [Google Scholar] [CrossRef] [PubMed]
  337. Bryukhanov, I.A.; Rybakov, A.A.; Larin, A.V.; Trubnikov, D.N.; Vercauteren, D.P. The role of water in the elastic properties of aluminosilicate zeolites: DFT investigation. J. Mol. Model. 2017, 23, 68. [Google Scholar] [CrossRef] [PubMed]
  338. Coasne, B.; Haines, J.; Levelut, C.; Cambon, O.; Santoro, M.; Gorelli, F.; Garbarino, G. Enhanced mechanical strength of zeolites by adsorption of guest molecules. Phys. Chem. Chem. Phys. 2011, 13, 20096–20099. [Google Scholar] [CrossRef] [PubMed]
  339. Mouhat, F.; Bousquet, D.; Boutin, A.; du Bourg, L.B.; Coudert, F.X.; Fuchs, A.H. Softening upon Adsorption in Microporous Materials: A Counterintuitive Mechanical Response. J. Phys. Chem. Lett. 2015, 6, 4265–4269. [Google Scholar] [CrossRef] [PubMed]
  340. Canepa, P.; Tan, K.; Du, Y.; Lu, H.; Chabal, Y.J.; Thonhauser, T. Structural, elastic, thermal, and electronic responses of small-molecule-loaded metal–organic framework materials. J. Mater. Chem. A 2015, 3, 986–995. [Google Scholar] [CrossRef]
  341. Hu, Y.; Navrotsky, A. Thermochemical Study of the Relative Stability of Dense and Microporous Aluminophosphate Frameworks. Chem. Mater. 1995, 7, 1816–1823. [Google Scholar] [CrossRef]
  342. Navrotsky, A.; Petrovic, I.; Hu, Y.; Chen, C.Y.; Davis, M.E. Little energetic limitation to microporous and mesoporous materials. Microporous Mater. 1995, 4, 95–98. [Google Scholar] [CrossRef]
  343. Boldyreva, E. High-Pressure Polymorphs of Molecular Solids:  When Are They Formed, and When Are They Not? Some Examples of the Role of Kinetic Control. Cryst. Growth Des. 2007, 7, 1662–1668. [Google Scholar] [CrossRef]
  344. Henson, N.J.; Cheetham, A.K.; Gale, J.D. Computational Studies of Aluminum Phosphate Polymorphs. Chem. Mater. 1995, 8, 664–670. [Google Scholar] [CrossRef]
  345. Fabbiani, M.; Polisi, M.; Fraisse, B.; Arletti, R.; Santoro, M.; Alabarse, F.; Haines, J. An in-situ x-ray diffraction and infrared spectroscopic study of the dehydration of AlPO4-54. Solid State Sci. 2020, 108, 106378. [Google Scholar] [CrossRef]
  346. Bennett, J.M.; Cohen, J.P.; Flanigen, E.M.; Pluth, J.J.; Smith, J.V. Aluminophosphate molecular sieve AlPO4-11: Partial refinement from powder data using a pulsed neutron source. ACS Symp. Ser. 1983, 218, 109–118. [Google Scholar] [CrossRef]
  347. Richardson, J.M.; Pluth, J.J.; Smith, J.V. Structure determination and rietveld refinement of aluminophosphate molecular sieve AIPO4-8. Acta Crystallogr. C 1987, 43, 1469–1472. [Google Scholar] [CrossRef]
  348. Ohnishi, N.; Qiu, S.; Teresaki, O.; Kajitani, T.; Hiraga, K. Hexagonal-orthorhombic phase transformation of AlPO4-5 aluminophosphate molecular sieve. Microporous Mater. 1993, 2, 73–74. [Google Scholar] [CrossRef]
  349. Ikeda, T.; Miyazawa, K.; Izumi, F.; Huang, Q.; Santoro, A.J. Structural study of the aluminophosphate AlPO4-5 by neutron powder diffraction. J. Phys. Chem. Solids 1999, 60, 1531–1535. [Google Scholar] [CrossRef]
  350. Polisi, M.; Arletti, R.; Quartieri, S.; Pastero, L.; Giacobbe, C.; Vezzalini, G. Dehydration mechanism of AlPO4-5: A high-resolution synchrotron X-ray powder diffraction study. Microporous Mesoporous Mater. 2018, 261, 137–143. [Google Scholar] [CrossRef]
  351. Xu, J.; Liu, Y.; Huang, Y.J. Ultrafast crystallization of AlPO4-5 molecular sieve in a deep eutectic solvent. Phys. Chem. C 2021, 125, 8876–8889. [Google Scholar] [CrossRef]
  352. Xu, J.; Chen, L.; Zeng, D.; Yang, J.; Zhang, M.; Ye, C.; Deng, J. Solid-state NMR of silicoaluminophosphate molecular sieves and aluminophosphate materials. J. Phys. Chem. B 2007, 111, 7105–7113. [Google Scholar] [CrossRef]
  353. Fan, F.; Feng, Z.; Sun, K.; Guo, W.; Guo, Q.; Song, Y.; Li, W.; Li, C. In situ UV Raman spectroscopic study on the synthesis mechanism of AlPO-5. Angew. Chem. Int. Ed. 2009, 121, 8899–8903. [Google Scholar] [CrossRef] [PubMed]
  354. Shi, Y.; Liu, G.; Wang, L.; Zhang, X. Ionothermal synthesis of phase pure AlPO4-5 using a series of tri-substituted imidazolium bromides. Microporous Mesoporous Mater. 2014, 193, 1–6. [Google Scholar] [CrossRef]
  355. Chen, B.; Kirby, C.W.; Huang, Y. Investigation of Crystallization of Molecular Sieve AlPO4-5 by the Dry Gel Conversion Method. J. Phys. Chem. C 2009, 113, 15868–15876. [Google Scholar] [CrossRef]
  356. Sheng, N.; Chu, Y.; Xin, S.; Wang, Q.; Yi, X.; Feng, Z.; Meng, X.; Liu, X.; Deng, F.; Xiao, F.S. Microwave Assisted Green Synthesis, Characterisation of Alanine Templated Aluminophosphate Zeolite and Study of Its Application as Adsorbent. J. Am. Chem. Soc. 2016, 138, 6171–6176. [Google Scholar] [CrossRef]
Figure 1. Views of crystal structures of (A) VPI-5; (B) ALPO-8; and (C) ALPO-31 from [0 0 1]. The images show 2 × 2 × 2 supecells of each material. Color code: Al-Green; P-Blue; O-red.
Figure 1. Views of crystal structures of (A) VPI-5; (B) ALPO-8; and (C) ALPO-31 from [0 0 1]. The images show 2 × 2 × 2 supecells of each material. Color code: Al-Green; P-Blue; O-red.
Solids 03 00032 g001
Figure 2. Views of the crystal structures (2 × 2 × 2 supecells) of ALPO-18 and ALPO-18W: (A) ALPO-18 from [0 0 1], [1 0 0] and [1 1 0]; (B) Perspective view of the 18-MR channels expanding along [0 0 1]; (C) ALPO-18W from [1 0 0]. Color code: Al-Green; P-Blue; O-red.
Figure 2. Views of the crystal structures (2 × 2 × 2 supecells) of ALPO-18 and ALPO-18W: (A) ALPO-18 from [0 0 1], [1 0 0] and [1 1 0]; (B) Perspective view of the 18-MR channels expanding along [0 0 1]; (C) ALPO-18W from [1 0 0]. Color code: Al-Green; P-Blue; O-red.
Solids 03 00032 g002
Figure 3. X-ray diffraction patterns of (A) VPI-5; (B) ALPO-8; (C) ALPO-5; (D) ALPO-18; and (E) ALPO-31, derived from the computed and experimental [4,6,9,10,12] crystal structures using CuK α radiation (λ = 1.540598 Å ).
Figure 3. X-ray diffraction patterns of (A) VPI-5; (B) ALPO-8; (C) ALPO-5; (D) ALPO-18; and (E) ALPO-31, derived from the computed and experimental [4,6,9,10,12] crystal structures using CuK α radiation (λ = 1.540598 Å ).
Solids 03 00032 g003
Figure 4. (A) Mechanical properties of VPI-5 ( P 6 3 c m ) as a function of the orientation of the applied strain: k -compressibility, E -Young modulus, G -Maximum shear modulus, and ν -Maximum Poisson’s ratio; (B) bidimensional projections on the x y plane; (C) bidimensional projections on the x z plane. The projections of the surface of minimum shear modulus are also displayed using green color in panels (B,C).
Figure 4. (A) Mechanical properties of VPI-5 ( P 6 3 c m ) as a function of the orientation of the applied strain: k -compressibility, E -Young modulus, G -Maximum shear modulus, and ν -Maximum Poisson’s ratio; (B) bidimensional projections on the x y plane; (C) bidimensional projections on the x z plane. The projections of the surface of minimum shear modulus are also displayed using green color in panels (B,C).
Solids 03 00032 g004
Figure 5. (A) Mechanical properties of ALPO-8 ( C m c 2 1 ) as a function of the orientation of the applied strain: k -compressibility, E -Young modulus, G -Maximum shear modulus, and ν -Maximum Poisson’s ratio; (B) bidimensional projections on the x y plane; (C) bidimensional projections on the x z plane. The projections of the surface of minimum shear modulus are also displayed using green color in panels (B,C).
Figure 5. (A) Mechanical properties of ALPO-8 ( C m c 2 1 ) as a function of the orientation of the applied strain: k -compressibility, E -Young modulus, G -Maximum shear modulus, and ν -Maximum Poisson’s ratio; (B) bidimensional projections on the x y plane; (C) bidimensional projections on the x z plane. The projections of the surface of minimum shear modulus are also displayed using green color in panels (B,C).
Solids 03 00032 g005
Figure 6. (A) Mechanical properties of ALPO-5 ( P 6 c c ) as a function of the orientation of the applied strain: k -compressibility, E -Young modulus, G -Maximum shear modulus, ν -Maximum Poisson’s ratio; (B) bidimensional projections on the x y plane; (C) bidimensional projections on the x z plane. The projections of the surface of minimum shear modulus are also displayed using green color in panels (B,C).
Figure 6. (A) Mechanical properties of ALPO-5 ( P 6 c c ) as a function of the orientation of the applied strain: k -compressibility, E -Young modulus, G -Maximum shear modulus, ν -Maximum Poisson’s ratio; (B) bidimensional projections on the x y plane; (C) bidimensional projections on the x z plane. The projections of the surface of minimum shear modulus are also displayed using green color in panels (B,C).
Solids 03 00032 g006
Figure 7. (A) Mechanical properties of ALPO-18 ( C 2 / c ) as a function of the orientation of the applied strain: k -compressibility, E -Young modulus, G -Maximum shear modulus, and ν -Maximum Poisson’s ratio; (B) bidimensional projections on the x y plane; (C) bidimensional projections on the x z plane. The projections of the surface of minimum shear modulus are also displayed using green color in panels (B,C).
Figure 7. (A) Mechanical properties of ALPO-18 ( C 2 / c ) as a function of the orientation of the applied strain: k -compressibility, E -Young modulus, G -Maximum shear modulus, and ν -Maximum Poisson’s ratio; (B) bidimensional projections on the x y plane; (C) bidimensional projections on the x z plane. The projections of the surface of minimum shear modulus are also displayed using green color in panels (B,C).
Solids 03 00032 g007
Figure 8. (A) Mechanical properties of ALPO-31 ( R 3 ¯ h ) as a function of the orientation of the applied strain: k -compressibility, E -Young modulus, G -Maximum shear modulus, and ν -Maximum Poisson’s ratio; (B) bidimensional projections on the x y plane; (C) bidimensional projections on the x z plane. The projections of the surface of minimum shear modulus are also displayed using green color in panels (B,C).
Figure 8. (A) Mechanical properties of ALPO-31 ( R 3 ¯ h ) as a function of the orientation of the applied strain: k -compressibility, E -Young modulus, G -Maximum shear modulus, and ν -Maximum Poisson’s ratio; (B) bidimensional projections on the x y plane; (C) bidimensional projections on the x z plane. The projections of the surface of minimum shear modulus are also displayed using green color in panels (B,C).
Solids 03 00032 g008
Figure 9. Computed unit cell volume (A) and lattice parameters a =   b (B) and c (C) of VPI-5 ( P 6 3 c m ) for different isotropic pressures. The volumetric compressibilities (D) and the linear compressibilities along a (E) and c (F) directions are shown in the panels of the right-hand side. The blue horizontal lines in panels (E,F) mark k l = 10 TPa 1 .
Figure 9. Computed unit cell volume (A) and lattice parameters a =   b (B) and c (C) of VPI-5 ( P 6 3 c m ) for different isotropic pressures. The volumetric compressibilities (D) and the linear compressibilities along a (E) and c (F) directions are shown in the panels of the right-hand side. The blue horizontal lines in panels (E,F) mark k l = 10 TPa 1 .
Solids 03 00032 g009
Figure 10. Computed unit cell volume (A) and lattice parameters (BD) of ALPO-8 under different external isotropic pressures. The volumetric compressibilities (E) and the linear compressibilities along a , b and c (FH) directions are shown in the panels of the right-hand side. The blue horizontal lines in panels (FH) mark k l = 10.0 TPa 1 .
Figure 10. Computed unit cell volume (A) and lattice parameters (BD) of ALPO-8 under different external isotropic pressures. The volumetric compressibilities (E) and the linear compressibilities along a , b and c (FH) directions are shown in the panels of the right-hand side. The blue horizontal lines in panels (FH) mark k l = 10.0 TPa 1 .
Solids 03 00032 g010
Figure 11. Computed unit cell volume (A) and lattice parameters a =   b (B) and c (C) of ALPO-5 ( P 6 c c ) for different isotropic pressures. The volumetric compressibilities (D) and the linear compressibilities along a (E) and c (F) directions are shown in the panels of the right-hand side. The blue horizontal lines in panels (E,F) mark k l = 10.0 TPa 1 .
Figure 11. Computed unit cell volume (A) and lattice parameters a =   b (B) and c (C) of ALPO-5 ( P 6 c c ) for different isotropic pressures. The volumetric compressibilities (D) and the linear compressibilities along a (E) and c (F) directions are shown in the panels of the right-hand side. The blue horizontal lines in panels (E,F) mark k l = 10.0 TPa 1 .
Solids 03 00032 g011
Figure 12. Computed unit cell volume (A) and lattice parameters a =   b (B) and c (C) of ALPO-31 ( R 3 ¯ h ) for different isotropic pressures. The volumetric compressibilities (D) and the linear compressibilities along a (E) and c (F) directions are shown in the panels of the right-hand side. The blue horizontal lines in panels (E,F) mark k l = 10.0 TPa 1 .
Figure 12. Computed unit cell volume (A) and lattice parameters a =   b (B) and c (C) of ALPO-31 ( R 3 ¯ h ) for different isotropic pressures. The volumetric compressibilities (D) and the linear compressibilities along a (E) and c (F) directions are shown in the panels of the right-hand side. The blue horizontal lines in panels (E,F) mark k l = 10.0 TPa 1 .
Solids 03 00032 g012
Figure 13. Computed unit cell volume (A) and lattice parameters (BD) of ALPO-18 under different external isotropic pressures. The volumetric compressibilities (E) and the linear compressibilities along a , b and c (FH) directions are shown in the subgraphs of the right-hand side. The blue horizontal lines in panels (FH) mark k l = 10.0 TPa 1 .
Figure 13. Computed unit cell volume (A) and lattice parameters (BD) of ALPO-18 under different external isotropic pressures. The volumetric compressibilities (E) and the linear compressibilities along a , b and c (FH) directions are shown in the subgraphs of the right-hand side. The blue horizontal lines in panels (FH) mark k l = 10.0 TPa 1 .
Solids 03 00032 g013
Figure 14. Computed unit cell volume (A) and lattice parameters (BD) of ALPO-18 under different uniaxial pressures applied along the direction of minimum compressibility, [1 0 0]. The volumetric compressibilities are shown in panel (E) The green horizontal line in panel (E) mark k V = 0.0 TPa 1 .
Figure 14. Computed unit cell volume (A) and lattice parameters (BD) of ALPO-18 under different uniaxial pressures applied along the direction of minimum compressibility, [1 0 0]. The volumetric compressibilities are shown in panel (E) The green horizontal line in panel (E) mark k V = 0.0 TPa 1 .
Solids 03 00032 g014
Figure 15. Two contiguous 8-MR channels expanding along [0 0 1] in the crystal structure of ALPO-18 under increasing isotropic pressures: 1.0, 1.75, 2.0, 2.25, and 2.50 GPa. The meaning of the width of a channel ( ω c h ) as measured by the distance between two opposite oxygen atoms is illustrated in the structure at P = 1.0 GPa. The values of ω c h are 6.453, 6.722, 7.049, 7.551, and 7.829 Å , respectively.
Figure 15. Two contiguous 8-MR channels expanding along [0 0 1] in the crystal structure of ALPO-18 under increasing isotropic pressures: 1.0, 1.75, 2.0, 2.25, and 2.50 GPa. The meaning of the width of a channel ( ω c h ) as measured by the distance between two opposite oxygen atoms is illustrated in the structure at P = 1.0 GPa. The values of ω c h are 6.453, 6.722, 7.049, 7.551, and 7.829 Å , respectively.
Solids 03 00032 g015
Figure 16. Comparison of the computed a lattice parameter of ALPO-18 under the effect of different external isotropic pressures using the PBEsol functional and the PBE functional supplemented with dispersion corrections.
Figure 16. Comparison of the computed a lattice parameter of ALPO-18 under the effect of different external isotropic pressures using the PBEsol functional and the PBE functional supplemented with dispersion corrections.
Solids 03 00032 g016
Figure 17. Comparison of the unit-cell volumes (A) and enthalpies (B) of the P 63 c m 2 and C 1 m 1 2 crystal structures of VPI-5 under different isotropic pressures.
Figure 17. Comparison of the unit-cell volumes (A) and enthalpies (B) of the P 63 c m 2 and C 1 m 1 2 crystal structures of VPI-5 under different isotropic pressures.
Solids 03 00032 g017
Figure 18. Comparison of the unit-cell volumes (A) and enthalpies (B) of the P cc 2 and P 6 c c crystal structures of ALPO-5 under different isotropic pressures.
Figure 18. Comparison of the unit-cell volumes (A) and enthalpies (B) of the P cc 2 and P 6 c c crystal structures of ALPO-5 under different isotropic pressures.
Solids 03 00032 g018
Table 1. Computed and experimental unit-cell parameters of the selected aluminophosphate materials.
Table 1. Computed and experimental unit-cell parameters of the selected aluminophosphate materials.
Parameter α   ( Å ) b   ( Å ) c   ( Å ) α   ( deg ) β   ( deg ) γ   ( deg ) Vol .   ( Å 3 ) ρ   ( g / cm 3 )
VPI-5 ( P 6 3 c m )
PBE18.566518.56658.533090.090.0120.02547.36301.431
PBE + Disp18.525318.52538.523490.090.0120.02533.23301.439
PBEsol18.543018.54308.539890.090.0120.02542.94951.433
Exp [4]18.6005(6)18.6005(6)8.3664(4)90.090.0120.02506.79311.455
ALPO-8 ( C m c 2 1 )
PBE33.589914.72808.533490.090.090.04221.55571.727
PBE + Disp33.395414.69158.522290.090.090.04181.19941.744
PBEsol33.517514.72238.539890.090.090.04213.97461.730
Exp [6]33.29(2)14.76(2)8.257(4)90.090.090.04057.16281.797
ALPO-5 ( P 6 c c )
PBE13.867113.86718.535090.090.0120.01421.36701.710
PBE + Disp13.840113.84018.526190.090.0120.01414.35651.718
PBEsol13.864513.86458.538290.090.0120.01421.36421.710
Exp [9]13.718(1)13.718(1)8.4526(5)90.090.0120.01377.53471.765
ALPO-18 ( C 2 / c )
PBE13.570412.677318.469990.090.0190.03177.48701.529
PBE + Disp13.556112.661318.444290.090.0290.03165.70841.535
PBEsol13.578812.670518.459290.090.0190.03175.90011.530
Exp [10]13.7114(1)12.7314(1)18.5703(1)90.090.01(1)90.03241.73021.500
ALPO-18 W   ( P 1 )
PBE9.32769.411318.337688.2491.7788.971607.93911.957
PBE + Disp9.25829.351318.244887.9991.8189.321577.67931.995
PBEsol9.24879.372318.250788.6691.9388.541580.11871.992
Exp [11]9.2519.36218.42890.8996.3590.871585.79711.985
ALPO-31 ( R 3 ¯ h )
PBE20.972420.97245.077890.090.0120.01934.22421.884
PBE + Disp20.923020.92305.068390.090.0120.01921.50461.897
PBEsol20.963520.96355.076790.090.0120.01932.15671.886
Exp [12]20.827(1)20.827(1)5.003(1)90.090.0120.01879.37981.940
Table 2. Computed elastic constants of selected aluminophosphate materials. All of the values are given in GPa.
Table 2. Computed elastic constants of selected aluminophosphate materials. All of the values are given in GPa.
i j C i j
VPI-5ALPO-8ALPO-5ALPO-18ALPO-31
P 6 3 c m C m c 2 1 P 6 c c C 2 / c R 3 ¯ h
1185.1782.23124.23105.68109.15
2285.17102.00124.2398.33109.15
33158.27187.62192.56105.82138.05
4419.1023.9426.2413.9930.81
5519.1025.9626.2418.7130.81
6621.7923.3528.2228.9023.01
1241.5953.0167.8068.6263.13
1343.6648.0559.7057.1761.58
140.00.00.00.05.76
150.00.00.0−2.420.46
160.00.00.00.00.0
2343.6651.1759.7052.1361.58
240.00.00.00.0−5.76
250.00.00.0−1.73−0.46
260.00.00.00.00.0
340.00.00.00.00.0
350.00.00.03.160.0
360.00.00.00.00.0
450.00.00.00.00.0
460.00.00.0−1.21−0.46
560.00.00.00.05.76
Table 3. Computed mechanical properties of selected aluminophosphate materials. The values of the bulk, shear and Young’s moduli ( B , G and E ) are given in in GPa. The single-crystal bulk moduli ( B s c ) are also given in the last row of the table for comparison.
Table 3. Computed mechanical properties of selected aluminophosphate materials. The values of the bulk, shear and Young’s moduli ( B , G and E ) are given in in GPa. The single-crystal bulk moduli ( B s c ) are also given in the last row of the table for comparison.
PropertyVPI-5ALPO-8ALPO-5ALPO-18ALPO-31
P 6 3 c m C m c 2 1 P 6 c c C 2 / c R 3 ¯ h
B Bulk modulus60.49 ± 0.5668.27 ± 0.8788.22 ± 0.6073.56 ± 0.9580.17 ± 1.16
G Shear modulus22.6725.5630.2419.3426.41
E Young’s modulus60.4668.1881.4353.3571.39
ν Poisson’s ratio0.330.330.350.380.35
D Ductility index2.662.672.923.803.04
D I Intrinsic ductility index0.370.430.511.020.45
H Hardness index0.941.221.130.630.70
A U Universal anisotropy0.660.830.490.460.36
B s c Bulk modulus (SC)60.74 ± 0.8172.60 ± 0.4287.27 ± 1.3673.85 ± 2.9290.14 ± 1.56
Table 4. Calculated volumetric compressibilities and linear compressibilities at zero pressure for VPI-5, ALPO-8, ALPO-5, ALPO-18 and ALPO-31.
Table 4. Calculated volumetric compressibilities and linear compressibilities at zero pressure for VPI-5, ALPO-8, ALPO-5, ALPO-18 and ALPO-31.
Material k V ( T P a 1 ) k a ( T P a 1 ) k b ( T P a 1 ) k c ( T P a 1 )
VPI-5 ( P 6 3 c m )16.466.956.952.56
ALPO-8 ( C m c 2 1 )13.776.935.041.81
ALPO-5 ( P 6 c c ) 11.454.614.612.25
ALPO-18 ( C 2 / c ) 13.543.405.174.97
ALPO-31 ( R 3 ¯ h )11.094.434.432.23
Table 5. Computed elastic constants of VPI-5 ( C 1 m 1 ), ALPO-5 ( P c c 2 ), and ALPO-18W ( P 1 ). All of the values are given in GPa.
Table 5. Computed elastic constants of VPI-5 ( C 1 m 1 ), ALPO-5 ( P c c 2 ), and ALPO-18W ( P 1 ). All of the values are given in GPa.
i j C i j
VPI-5ALPO-5ALPO-18W
C 1 m 1 P c c 2 P 1
1169.66120.9092.30
2268.50127.95103.17
3379.68193.34103.75
4418.7526.3025.58
5519.6126.0424.06
6619.3825.6125.81
1231.5367.0137.53
1314.0460.7542.27
140.00.0−2.52
15−0.050.0−5.10
160.00.07.90
2312.6062.5341.38
240.00.06.67
25−0.200.0−0.47
260.00.03.03
340.00.00.50
350.420.0−3.83
360.00.0−1.19
450.00.01.85
460.200.00.58
560.00.01.37
Table 6. Computed mechanical properties of VPI-5 ( C 1 m 1 ), ALPO-5 ( P c c 2 ) and ALPO-18W ( P 1 ). The values of the bulk, shear and Young’s moduli ( B , G and E ) are given in in GPa. The single-crystal bulk moduli ( B s c ) are also given in the last row of the table for comparison.
Table 6. Computed mechanical properties of VPI-5 ( C 1 m 1 ), ALPO-5 ( P c c 2 ) and ALPO-18W ( P 1 ). The values of the bulk, shear and Young’s moduli ( B , G and E ) are given in in GPa. The single-crystal bulk moduli ( B s c ) are also given in the last row of the table for comparison.
PropertyVPI-5ALPO-5ALPO-18W
C 1 m 1 P c c 2 P 1
B 37.06 ± 0.8188.62 ± 0.3758.82 ± 0.94
G 20.9829.5825.81
E 54.2479.8767.56
ν 0.260.350.31
D 1.773.002.28
D I 0.240.510.18
H 3.101.022.11
A U 0.290.500.24
B s c 33.85 ± 1.9087.71 ± 0.2758.80 ± 1.84
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Colmenero, F.; Lobato, Á.; Timón, V. Mechanical Characterization of Anhydrous Microporous Aluminophosphate Materials: Tridimensional Incompressibility, Ductility, Isotropy and Negative Linear Compressibility. Solids 2022, 3, 457-499. https://doi.org/10.3390/solids3030032

AMA Style

Colmenero F, Lobato Á, Timón V. Mechanical Characterization of Anhydrous Microporous Aluminophosphate Materials: Tridimensional Incompressibility, Ductility, Isotropy and Negative Linear Compressibility. Solids. 2022; 3(3):457-499. https://doi.org/10.3390/solids3030032

Chicago/Turabian Style

Colmenero, Francisco, Álvaro Lobato, and Vicente Timón. 2022. "Mechanical Characterization of Anhydrous Microporous Aluminophosphate Materials: Tridimensional Incompressibility, Ductility, Isotropy and Negative Linear Compressibility" Solids 3, no. 3: 457-499. https://doi.org/10.3390/solids3030032

Article Metrics

Back to TopTop