Next Article in Journal
A Comparative Study on the K-ion Storage Behavior of Commercial Carbons
Previous Article in Journal
Significance of Alloying Elements on the Mechanical Characteristics of Mg-Based Materials for Biomedical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Calculations of the Structural, Electronic, Optical, and Mechanical Properties of 21 Pyrophosphate Crystals

1
Department of Sciences, College of Basic Education, Al Muthanna University, Samawah 66001, Iraq
2
Department of Physics and Astronomy, University of Missouri-Kansas City, Kansas City, MO 64110, USA
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(8), 1139; https://doi.org/10.3390/cryst12081139
Submission received: 19 July 2022 / Revised: 8 August 2022 / Accepted: 10 August 2022 / Published: 12 August 2022
(This article belongs to the Topic First-Principles Simulation—Nano-Theory)

Abstract

:
Pyrophosphate crystals have a wide array of applications in industrial and biomedical fields. However, fundamental understanding of their electronic structure, optical, and mechanical properties is still scattered and incomplete. In the present research, we report a comprehensive theoretical investigation of 21 pyrophosphates A2M (H2P2O7)2•2H2O with either triclinic or orthorhombic crystal structure. The molecule H2P2O7 is the dominant molecular unit, whereas A = (K, Rb, NH4, Tl), M = (Zn, Cu, Mg, Ni, Co, Mn), and H2O stand for the cation elements, transition metals, and the water molecules, respectively. The electronic structure, interatomic bonding, partial charge distribution, optical properties, and mechanical properties are investigated by first-principles calculations based on density functional theory (DFT). Most of these 21 crystals are theoretically investigated for the first time. The calculated results show a complex interplay between A, M, H2P2O7, and H2O, resulting in either metallic, half-metallic, or semi-conducting characteristics. The novel concept of total bond order density (TBOD) is used as a single quantum mechanical metric to characterize the internal cohesion of these crystals to correlate with the calculated properties, especially the mechanical properties. This work provides a large database for pyrophosphate crystals and a road map for potential applications of a wider variety of phosphates.

1. Introduction

Despite longstanding research interest, phosphate crystals have attracted a great deal of attention in recent years because of their application in novel devices. Generally speaking, phosphate crystals can be classified into three categories depending on the bonding nature of the P-O groups: orthophosphates, pyrophosphates, and polyphosphates. Among them, the pyrophosphate structure contains a P2O7 or P2H2O7 dimer. Phosphates have rich and unique crystal structure types, therefore they participate in a wide range of applications of phosphate materials, such as sodium ion metal batteries [1], nonlinear optical (NLO) crystals with short cutoff edges [2,3,4,5,6,7,8], near ultraviolet light-emitting diode (NUV LED) applications [9], and birefringent materials [10]. For example, potassium dihydrogen phosphate (KH2PO4/KDP), as a soft-brittle optical crystal, is extensively used in the construction of inertial confinement fusion (ICF) facilities [11,12]. Other types of phosphate crystals are used to reduce pollution and climate change effects due to their capacity for CO2 and CH4 sorption [13].
Inorganic acidic pyrophosphate (PPi) (or inorganic diphosphate) materials are omnipresent in nature [14,15]. These compounds belong to the family of inorganic acidic diphosphates with the very general formula A2M B2•nH2O, where A is a monovalent cation such as ammonium (NH4) or an alkaline earth cation (Na, K, and Rb), and M can be a divalent transition metal cation (Cu, Ni, Co, and Mn) or Zn2+ and Mg2+. With the acidic diphosphates, several forms of the acidic anions B can exist in the molecular structure, such as H2P2O72−, HP2O73−, and H3P2O7, causing these acidic diphosphates to exhibit important biological properties such as antitumor, antibacterial, and antifungal activity against Salmonella typhimurium, Enterococcus faecium, and Candida albicans [16]. Inorganic acidic pyrophosphates basically have two forms: crystalline and glassy forms. Both forms have attracted considerable attention as promising inorganic materials. They possess many applications in industry due to their uses as solid-state laser hosts [17], dielectric materials [18], magnetic materials [19], rechargeable lithium batteries [20], in ceramic applications [21], and in the area of gas separation [22]. Inorganic acidic pyrophosphates also have biological importance such as their use in some catalytic processes [23], or as ion exchangers [22]. Of special interest are the crystals: K2M (H2P2O7)·2H2O and (NH4)2M(H2P2O7)·2H2O where (M = Mg, Zn, Cu, Ni, Co, and Mn). These materials are polymorphic structures with three different space groups, namely triclinic P-1, orthorhombic Pnma, and monoclinic P2/m [13,24,25,26,27,28,29,30,31,32].
PPi, which can be produced through a variety of biosynthetic reactions, such as the biosynthesis of nucleotides, lipids, urea, and oxidative phosphorylation [33], is also a prevalent reactant in some metabolic pathways including fat metabolism, protein synthesis, nucleoside diphosphate sugar synthesis, and DNA and RNA synthesis [34]. For instance, orthophosphates (with orthophosphate group: PO43−), metaphosphate (with metaphosphate group: PO3), and pyrophosphate (with pyrophosphate group: P2O74−) have many biological uses [35], especially their role in some enzyme-catalyzed reaction processes [36]. The P2O74− anion is unstable in hydrous solutions and rapidly hydrolyzes to inorganic phosphate in the presence of divalent metal ions (P2O74− + H2O→2HPO42−), which promotes the splitting of both inorganic and biological pyrophosphates, especially in acidic conditions [37]. In the case of pyrophosphates, it has been found that they are very effective in certain catalytic processes such as the conversion of butane to maleic anhydride [38,39,40]. They also have a crucial role in some bioenergetic processes [41,42], such as the energy circulating in functional cells [43,44]. Moreover, they can be used as inhibitors in the in vitro formation and dissolution of apatite crystals [45].
Biological pyrophosphates hydrolyze to a family of enzymes known as pyrophosphatases (PPases) [46]. Soluble pyrophosphatases (sPPases) are very common and are found in all organisms in two forms: (i) organic and (ii) inorganic [47]. These enzymes are fundamental for the regulation of intra-cellular levels of inorganic phosphate and for the removal of pyrophosphate products of nucleotide coupling reactions [48]. There are two unrelated families of inorganic pyrophosphatases, the ubiquitous Mg-dependent family, and the Mn-dependent family, and both of them can be found in bacteria and archaea [49,50]. The soluble pyrophosphatases have several functions, including some steroid biosynthetic pathways (such as juvenile hormone synthesis), regulation of cell motility, nucleotide metabolism, and they may have a nuclear role in regulating transcription of some genes [51,52,53,54]. PPases need divalent metal cations to be more active. The efficiency of such cations as an activator decreases in the order: Mg > Zn > Co >Mn > Cd [55,56]. These enzymes can have as many as four functional divalent metal ions in the active site, with two metal ions bound as essential cofactors, whereas the third and fourth are ligated to pyrophosphate/phosphate-forming motifs [55,56].
Several theoretical and experimental investigations of some important pyrophosphate crystals have appeared in the literature in recent years. Using experimental and computational methods, some uses of pyrophosphates as water oxidation catalysts were investigated by Kim et al. [57]. Other pyrophosphate materials were investigated separately via density functional theory (DFT) [58,59,60]. Recently, another pair of complex acidic metal pyrophosphate crystals have been experimentally investigated by F. Elhafiane et al. [61,62]. The crystal structure of the orthorhombic phase of K2Zn (H2P2O7)2·2H2O, K2Cu (H2P2O7)2·2H2O, and K2Ni (H2P2O7)2·2H2O crystals was investigated experimentally for the first time by R. Essehli et al. [24]. Vibrational analysis, Raman, and infrared spectra studies of K2Zn (H2P2O7)2·2H2O and K2Cu (H2P2O7)2·2H2O crystals were reported [31,63]. The electronic structure, optical properties, and mechanical properties of the orthorhombic phase of K2Zn (H2P2O7)2·2H2O and K2Cu (H2P2O7)2·2H2O crystals and the triclinic phase of K2Mg (H2P2O7)2·2H2O, (NH4)2Mg (H2P2O7)2·2H2O, and (NH4)2Zn (H2P2O7)2·2H2O crystals were theoretically investigated recently via DFT [64,65], but they did not include the spin-polarized effect in their calculations. The ammonium diphosphates crystals: (NH4)2Ni (H2P2O7)2·2H2O, (NH4)2Co (H2P2O7)2·2H2O, and (NH4)2Mn (H2P2O7)2·2H2O were synthesized and prepared by R. Essehli et al. [26,28] and F. Capitelli et al. [27]; however, to our knowledge, no theoretical investigations have been attempted for the electronic structure, optical properties, and mechanical properties of these crystals. Even though the crystal structure and vibrational analysis of Rb2Zn (H2P2O7)2·2H2O, Rb2Mg (H2P2O7)2·2H2O, and K2Co (H2P2O7)2·2H2O crystals were experimentally investigated by R. Essehli et al. [18] and A. Lamhamdi et al. [66], their electronic structure, optical properties, and mechanical properties have not yet been theoretically investigated. The crystal structure of thallium diphosphates: Tl2M (H2P2O7)2•2H2O (M = Mg, Mn, Co, Ni, and Zn) crystals was synthesized for the first time by B. El Bali [67], but no theoretical investigations have been reported for their electronic structure, optical properties, and mechanical properties.
With the complex structures of pyrophosphate crystals and their widespread presence, the primary understanding of their structures and other properties has been a subject of many studies. However, these studies are still not as common as for other inorganic materials such as silicates and chalcogenides. With high-accuracy calculations that can be provided by quantum mechanical principles, more information on the electronic structure, molecular interaction, optical properties, and mechanical properties can be revealed. In this work, we present the results of DFT calculations of 21 pyrophosphate crystals as listed in Table 1 with more information about their experimental and calculated lattice parameters and symmetry. These 21 crystals have the same structural formula: A2M (H2P2O7)2•2H2O. They have the same pyrophosphate group, H2P2O7, and the same number of H2O molecules. The only difference between these 21 crystals is the type of the cation A (A = K, Rb, NH4, and Tl) and metallic element M (M = Mg, Zn, Cu, Ni, Co, and Mn). Some crystals under study include ferromagnetic elements (Ni, Mn, Co) in their structure. To explore the electronic and optical characteristics of these crystals, the spin-polarized DFT calculations were performed and their results are presented in a separate section. With the detailed and accurate calculations of the electronic structure, bonding, optical properties, and mechanical properties, we offer perceptible insights that give better understanding of these pyrophosphate crystals. The 21 pyrophosphate crystals are listed in Table 1 in specific order, from 1 to 21. For consistency, we follow this order in all tables, figures, and discussions. In Table 1, we give each crystal an abbreviation and this abbreviation will be used in all other tables to save space. Figure 1 depicts the structure of the crystals 1 and 2 in ball-and-stick form. Studying the set of 21 pyrophosphate crystals enables us to develop important correlations among the properties of the different crystals. For example, we use a single quantum mechanical metric—the total bond order density (TBOD), to describe the strength and internal cohesion of the crystal. Our focus will be on the bonding, mechanical properties, and optical properties, and their correlations with the TBOD. In the next section, we briefly describe the crystal structures and computational methods used in the calculation. The main results are presented and discussed in Section 3. The paper ends with a summary and some conclusions.

2. Materials and Methods

In this work, we used two density functional theory (DFT)-based methods to perform our calculations: the Vienna ab initio simulation package (VASP) [69] and the orthogonal linear combination of atomic orbitals method (OLCAO) [70]. VASP was used to optimize the crystal structures and calculate the mechanical properties. The one-electron orbitals and the electronic charge density were expanded in a plane wave basis set with an energy cut-off of 500 eV. The energy cut-off value was found to be the best choice, balancing the fast convergence and accurate ground state energy. We used the generalized gradient approximation (GGA) of Perdew, Burke, and Ernzerhof (PBE) [71] as the exchange and correlation potential for solving the Kohn–Sham equation. In the VASP calculation, the electronic and ionic force convergence criteria were set at 10−6 eV and 10−4 eV/Å, respectively. We used the Monkhorst scheme [72] with different k-point meshes ranging from 4 × 4 × 4 for larger crystals such as the orthorhombic phase of crystals K2Zn (H2P2O7)2·2H2O, K2Cu (H2P2O7)2·2H2O, and K2Ni (H2P2O7)2·2H2O (124 atoms) to 8 × 8 × 8 for the other smaller crystals which have triclinic space group such as (NH4)2Mg (H2P2O7)2·2H2O (31 atoms). The VASP-optimized crystal structures were used as inputs to calculate the electronic structure, interatomic bonding, and optical properties, using the in-house developed OLCAO package.
OLCAO is an all-electron method based on the local density approximation (LDA). Atomic orbitals expanded as Gaussian-type orbitals (GTO) are used in OLCAO as the basis set for expanding the solid-state wave function. The use of localized atomic orbitals in the basis expansion, which contrasts with the plane-wave expansion used in VASP, is successful for the DFT calculations of both crystalline [73,74,75,76,77,78] and non-crystalline materials [79,80], especially those with complex structures typical in biomolecular systems [81,82]. A suitably large number of Monkhorst k-points were used, (10 × 10 × 10) for all crystals, for density of states (DOS), band structure, and optical properties calculations. The Mulliken scheme [83] was used to calculate the partial charge (PC) and interatomic bonding. We define the PC of an atom as the difference between the effective charge Q* and the electronic charge of a neutral atom (Q0) in units of electron. Mathematically, ∆Q = Q0 − Q*. Negative ∆Q implies a gain of electrons (i.e., an electronegative ion), and positive ∆Q implies a loss of electrons (i.e., electropositivity). The formulae of effective charge (Qα*) and bond order (BO) values are shown in Equations (1) and (2), also called the overlap population, ραβ between any pair of atoms (α, β).
Q α * = i m , o c c j , β C i α * m C j β m S i α , j β
ρ α β = m ,   o c c i , j C i α * m C j β m S i α , j β
In the above equations, S i α . j β are the overlap integrals between the ith orbital of the αth atom and the jth orbital of the βth atom. C j β m is the eigenvector coefficients of the mth occupied band. The BO (Equation (2)) defines the relative strength of the bond between two atoms. The summation of all BO values in the crystal gives the total bond order (TBO). The total bond order density (TBOD) is obtained when TBO is normalized by the cell volume. TBOD is a single quantum mechanical metric to describe the internal cohesion of the crystal [84] and can be decomposed into partial components (PBOD) for any structural units or groups of bonded atomic pairs. The combination of using VASP and OLCAO packages with two different basis expansions is highly effective in revealing the subtle features in the material properties, especially regarding atomic scale details. More details about the methods used in the optical and mechanical properties calculations are given in the Supplementary Material.

3. Results and Discussion

3.1. Electronic Structure

Understanding the electronic structure is important for identifying the optoelectronic behaviors of the materials and their multiple uses in technological fields. For this purpose, we calculated the total density of states (TDOS) of the 21 pyrophosphate crystals. To achieve more deep and detailed information about the interaction between different atoms and the differences among the 21 crystals, we resolved the TDOS into partial density of states (PDOS) for each crystal structural unit. The TDOS and PDOS from −15 to 15 eV are displayed for all 21 crystals in Supplemental Figure S1 in the Supplementary Information (SI). The energy gaps of the 21 crystals are presented in Table 2. Most of these 21 crystals are insulators with different values of Eg. However, a closer inspection of the TDOS and PDOS shows the need for more accurate analysis. We selectively discuss some of the crystals in more detail, focusing on those crystals that have an unclear energy band gap and those that we deem to be particularly interesting. In Supplemental Figure S2, we plotted TDOS and PDOS for crystals numbered from 1 to 9, and crystals 15, 16, 18, and 21 with an energy range of −6 to 6 eV. Supplemental Figure S2 shows that there is a slight overlap of the valence bands above the Fermi level in all crystals. These bands mainly consist of 2p states of the O atom in crystals 1, 2, 8, and 9, while they mainly consist of 2p states of the O atom and 3d states of Cu and Ni atoms in crystals 3, 4, 5, and 6. In crystals 15 and 16, the valence bands cross the Fermi level and these bands mainly consist of 4s and 3d states of the Co and Mn atoms. In crystal 18, valence bands cross the Fermi level and these bands mainly consist of 2p states of O atom and 4s and 3d states of Ni atom, while in crystal 21, the valence bands that cross the Fermi level consist mainly of 2p states of O atom and 4s and 3d states of Co atom. In conclusion, the accurate analysis of the electronic structure of the discussed crystals indicates semi-metallic behavior.

3.2. Interatomic Bonding

The ability to calculate the interatomic bonding properties is a key strength of this work. Revealing and understanding the bonding properties of phosphates is crucial, especially for biological processes. For example, protein phosphorylation, which can be described in a very general picture, is a common mechanism for modifying protein function [85]; it partially depends on the bonding nature between the P atom and other atoms in DNA and RNA. Much of its use depends on the reversibility of the process, which is often mediated by protein phosphatases. Protein phosphorylation is an important post-translational modulation that is an integral part of cellular function [86,87]. The phosphorylation of any amino acid residue results in a change in charge and thus in the protein surface potential. For instance, phosphorylation of the amino acids: serine, threonine, and tyrosine, results in a change from neutral to negative (−2). Hence, there is no surprise that phosphorylation can affect protein conformation, protein–protein interactions, and biochemical pathways, and its dysregulation is connected to many disease states [88,89,90]. The interatomic interaction between atoms can be explained more clearly by showing the BO vs. BL (bond length) distribution for each crystal. The BO data can also be used to obtain the TBOD for each structural unit. BO vs. BL distribution for each crystal is shown in Supplemental Figure S3 (1–21) in the SI. In crystals 1 and 3 (the orthorhombic phase), there are six different types of bonds, while in crystal 5 (the orthorhombic phase), there are seven different types of bonds. All the remaining triclinic crystals contain five different types of bonds, except for crystal 12 ((NH4)2Zn(H2P2O7)2·2H2O), which contains six different types of bonds. The four most common bonds in all crystals are the P-O and H-O bonds, hydrogen bonds (H-H), and negligibly weak (H-O) bonds when far apart. P-O and H-O bonds can be described as polar covalent bonds. We should mention that considering these bonds as polar covalent bonds is not final, because one should keep in mind that the difference in the electronegativity value between the P atom (with electronegativity = 2.19) and the O atom (with electronegativity = 3.44), or between the H atom (with electronegativity = 2.2) and the O atom (with electronegativity = 3.44) is not small (higher than 1.2).
These 21 crystals also have some bonds that are considered to be ionic bonds, such as the Zn-O, K-O, Mg-O, Rb-O, and Tl-O bonds. The polar covalent P-O bonds are the strongest bonds, and they are only slightly different in each crystal, which implies a strong tetrahedral PO4 unit as is true in all phosphates. All P-O bonds occur at short BLs, between 1.5 Å and 1.8 Å. Uniquely, crystals 1 and 5 have very strong P-H bonds which occur at a BL of 1.5 Å. The H-O bonds in all crystals are distributed along long BLs, between 1.0 Å and 3.5 Å. The H-O bonds that occur with short BLs (about 1.0 Å) are strong bonds with a high BO value. The Zn-O, Cu-O, Ni-O, Mg-O, Co-O, and Mn-O bonds in all crystals have comparable values of BO and BL, while K-O, Rb-O, and Tl-O bonds are weak bonds with small values of BO. The unique covalent N-H bonds in crystals from 12 to 16 are very strong bonds because they are part of the intramolecular bonds in NH4. In Figure 2 and Figure 3, we plotted the BO versus BL distribution for the crystals from 1 to 6. The pie charts in Figure 4 to be shown later give the proportion contributions from diverse bond types to the total bond order. P-O bonds have the highest contribution in all crystals and are responsible for their internal cohesion. All other bonds can play their part in the crystal cohesion of pyrophosphates, including the H-O hydrogen bonds. The sum of total bond order values in the crystal when normalized by volume gives the TBOD, a very useful parameter to identify internal cohesion in pyrophosphate crystals, and they are listed in Table 2. In Figure 5, we display the TBOD for the 21 crystals in sequential order as designated earlier in Table 1. The 21 crystals were divided into four groups with four different colors: black for A2M (H2P2O7)2•2H2O(A = K), red for A2M (H2P2O7)2•2H2O(A = Rb), green for A2M (H2P2O7)2•2H2O(A = NH4), and blue for A2M (H2P2O7)2•2H2O(A = Tl). It turns out that the crystals from 12 to 16 have the highest cohesion, with the highest values of TBOD. This feature for these crystals (12–16) can be explained by the existence of unique, very strong covalent N-H bonds. In Supplemental Figure S4, we plotted the crystal structure of crystal 2 in ball-and-stick format. As can be seen, H-O bonds occur at different BLs: short BLs (about 1.0 Å), medium BLs (about 1.8 Å), and long BLs (about 3.5 Å). H-O bonds with short BLs belong to H2O molecule, while other H-O bonds with medium and long BLs probably belong to H2P2O7 molecule. There are some very weak H-O bonds with long BLs that link H2O molecule with H2P2O7 molecule.

3.3. Partial Charge

The partial charge (PC) or the charge transfer for an atom in a crystal is the deviation of the effective charge Q* from the neutral atom charge Q0. It is a crucial part of the electronic structure. The calculated Q* on each atom in the pyrophosphate crystals is shown in Table 3. The distribution of the calculated PC for each element in each crystal is shown in Supplemental Figure S5 (1–21) in the SI. As can be seen in Supplemental Figure S5, K, Cu, Zn, Ni, Mg, Co, Rb, Mn, P, and H atoms have positive PC whereas O, Tl, and N atoms have negative PC. Mg atom has the largest positive PC among other elements. Small variations of the PC for P, H, or O are due to their locations in the structural units of H2P2O7 or H2O. P and H elements in the orthorhombic phase of K2Zn (H2P2O7)2·2H2O (crystal 1) can lose or gain charge, whereas they only lose charge in the triclinic phase (crystal 2). P element in the orthorhombic phase of K2Ni (H2P2O7)2·2H2O (crystal 5) can lose or gain charge, while it only loses charge in the triclinic phase (crystal 6).
Figure 4. The pie charts show the percentages of different bonding types for crystals from 1 to 6.
Figure 4. The pie charts show the percentages of different bonding types for crystals from 1 to 6.
Crystals 12 01139 g004
Figure 5. Distribution of TBOD for the 20 crystals. The black color is for the crystals from 1 to 8, red color is for crystals from 9 to 11, green color is for crystals from 12 to 16, and blue color is for crystals from 17 to 21.
Figure 5. Distribution of TBOD for the 20 crystals. The black color is for the crystals from 1 to 8, red color is for crystals from 9 to 11, green color is for crystals from 12 to 16, and blue color is for crystals from 17 to 21.
Crystals 12 01139 g005

3.4. Optical Properties

We have explored the optical properties of the 21 pyrophosphate crystals through the frequency-dependent dielectric function which is mainly connected with the electronic structures. The frequency-dependent dielectric function can be used to calculate many optical parameters such as refractive index, absorption, energy loss, reflectivity, and optical conductivity [91]. The complex dielectric function, which has real and imaginary parts, is given by [92]:
ε ( ω ) = ε 1 ( ω ) + i ε 2 ( ω )
The first part of the equation above gives the dispersive part, while the second part represents the absorption part of complex dielectric function. The imaginary part can be estimated by the means of density of states and momentum matrix elements associated with occupied and unoccupied states. The imaginary part can be calculated relatively easily within the one-electron random phase approximation using the OLCAO method. More information about the mathematical formulas involved in these calculations is given in Section 1 in the Supplementary Material. To investigate the linear response of a system based on the inter-band transitions, the real part of dielectric function (ε1(ω)), which is directly related to polarizability in the electromagnetic spectra, must be discussed in some detail. An intra-band transition is a transition between two electronic energy states in the same band, while an inter-band transition is a transition between two electronic energy states in two different bands. Optical dielectric functions of the 21 crystals are shown in Supplemental Figure S6 (1–21) in the SI. As can be seen in Supplemental Figure S6 (1–21), each crystal has its unique absorption features; we selectively discuss the crystals 1, 6, 8, 12, 13, 17, and 19 in some detail. The maximum value of dispersion is observed in the energy range from 6.5 eV to 8.0 eV for crystals 1, 6, 8, 12, and 13, while the maximum value of dispersion is observed in the energy range from 3.0 eV to 3.5 eV for crystals 17 and 19. After attaining its maximum value, ε1(ω) starts decreasing after a few vibrations. The dispersive part along the energy range (0–40 eV) shows a positive value due to semi-conducting behavior of crystals 1, 6, 8, 12, 13, 17, and 19. We should mention that crystals 9, 10, and 11 have negative peaks at energy E ≈ 19.0 eV, and these peaks can identify the reflection of incident photons.
The imaginary part of the dielectric function (ε2(ω)) shows the energy needed to excite electrons from the valence band to the conduction band [93]. It is implied with the absorption of incident photon. The peaks in the ε2(ω) spectra are caused by transitions from the occupied valence band (VB) states to the unoccupied conduction band (CB) states, which originally occurred due to inter-band transitions. The ε2(ω) curve has different threshold values depending on the energy gap of each crystal. Threshold values of crystals 1, 6, 8, 12, 13, 17, and 19 are 4.5 eV, 3.0 eV, 5.4 eV, 4.9 eV, 5.0 eV, 3.1 eV, and 3.2 eV, respectively. In most of the 21 crystals, the absorption spectrum is observed to diminish after 30 eV. From Supplemental Figure S6 (1–21), it can be noticed clearly that for most of the 21 crystals, the highest absorption of incident photon occurs in the energy range from 9.0 eV to 10.0 eV, which can be considered to be in the ultraviolet (UV) region. These crystals can therefore be used for protection from ultraviolet radiation. In crystals 3 and 17–21, the highest absorption of incident photon occurs in the energy range from 3.5 eV to 4.0 eV, which can be considered to be in the visible light region. The refractive index is given by:
n = ε r µ r
In Equation (4), εr is the relative permittivity or the dielectric function and µr is the relative permeability. The relative permeability (µr) is very close to unity (with the absence of magnetic effects), so that the refractive index can be approximated as n = ε r . The real part of the dielectric function ε1(ω) can be obtained from ε2(ω) using Kramers–Kronig transformation. The static refractive index n can be obtained from the square root of ε1(ω) in the zero frequency limit, i.e., n = ε 1 ( 0 ) . The calculated refractive indices are listed in Table 2. Crystals 11, 17, 20, and 21 have the highest values of n: 1.718, 1.703, 1.844, and 2.098, respectively. Another important optical parameter that can be calculated from the complex dielectric function is the energy loss function (ELF). ELFs represent the collective excitation of electrons at high frequency. The ELF is related to the loss of fast electrons [93]. Generally speaking, the peaks of the ELF are plasmon oscillations that are created due to valence band maxima and conduction band minima at different values of energy depending on the type of each crystal. The position of ELF’s main peak is interpreted as the plasmon frequency ωp, which usually occurs at the frequency when the real part of the dielectric function vanishes [ε1p) = 0)]. Below ωp, the incident waves will be mostly reflected. Above ωp, the material behaves like plasma, or all the excited electrons are nearly free electron. The calculated energy loss functions are shown in Supplemental Figure S7 in the SI, and the calculated values of ωp are listed in Table 2. Supplemental Figure S7 indicates isotropic character in the energy range from 0 to 50 eV for most crystals. The plasmon oscillations of 21 pyrophosphate crystals occur in the energy range from 20 eV to 23 eV.

3.5. Spin-Polarized Calculations

The interaction of magnetism and electronic structure characteristics in the same material may give rise to half-metallic ferromagnetism (HMF). The half-metallic materials can have metallic behavior for electrons of one spin direction while there could be semi-conducting behavior in the opposite spin direction. Half-metallic ferromagnetic materials have wide application in spintronic and spin caloritronic devices [94,95]. We have calculated the TDOS and optical dielectric functions for the crystals 3 (K2Cu-3), 4 (K2Cu-4), 5 (K2Ni-5), 6 (K2Ni-6), 7 (K2Co-7), 11 (Rb2Co-11), 14 ((NH4)2Ni-14), 15 ((NH4)2Co-15), 16 ((NH4)2Mn-16), 18 (Tl2Ni-18), 20 (Tl2Mn-20), and 21 (Tl2Co-21) by considering the spin-polarized effect within the local density approximation (LDA) using OLCAO code. Energy gaps and refractive indices are presented in Table 4. The calculated TDOS and optical dielectric functions are displayed in Supplemental Figures S8 and S9 in the SI. Positive (negative) TDOS are plotted for majority (minority) electrons. From Supplemental Figure S8, K2Cu-3 has semi-conducting behavior in the majority spin channel (spin-up), with Eg being about 1.3 eV, while it has semi-metallic behavior in the minority spin channel (spin-down) due to the slight overlap of the valence bands above the Fermi level. However, on closer inspection, K2Cu-3 has semi-metallic behavior in the majority spin channel due to a very slight overlap of the valence bands. K2Cu-4 has semi-conducting behavior in the majority spin channel with Eg being about 4.0 eV and semi-conducting behavior in the minority spin channel with Eg equal to 0.2 eV. However, on closer inspection, K2Cu-4 has semi-metallic behavior in the majority spin channel due to the slight overlap of the valence bands above Fermi level. Therefore, based on the more accurate analysis of the electronic structure of crystals 3 and 4, we say that they do not have half-metallic (HM) behavior. K2Ni-5 has semi-metallic behavior in both the majority and minority spin channels because the energy states near the valence band top are metallic bands with both occupied and unoccupied bands below and above the Fermi level.
TDOS for K2Ni-6 is clearly identical for both the majority and minority spin channels. K2Co-7 has semi-conducting behavior in the majority spin channel with Eg being about 0.6 eV, while it has semi-metallic behavior in the minority spin channel due to the overlap of the valence bands above the Fermi level. Rb2Co-11 has semi-metallic behavior in the majority spin channel due to the slight overlap of the valence bands above the Fermi level, while it has semi-conducting behavior in the minority spin channel with Eg being about 0.2 eV. TDOS for (NH4)2Ni-14 is identical for both the majority and minority spin channels. (NH4)2Co-15 has semi-conducting behavior in the majority and minority spin channels, however, on closer inspection, it appears that (NH4)2Co-15 has semi-metallic behavior in the minority spin channel due to the slight overlap of the valence bands above Fermi level. (NH4)2Mn-16 has semi-conducting behavior in the majority spin channel with Eg being about 0.2 eV, while it has semi-metallic behavior in the minority spin channel because the energy states near the valence band top are metallic bands with both occupied and unoccupied band below and above the Fermi level. Tl2Co-21 has semi-metallic behavior in the majority spin channel and semi-conducting behavior in the majority spin channel with clear Eg. In conclusion, our calculations showed that the crystals 3, 4, 5, 7, 11, 15, 16, 20, and 21 have half-metallic (HM) behavior, while the crystals 6, 14, and 18 have identical electronic behavior.

3.6. Mechanical Properties

Although pyrophosphate crystals have many structural applications, their mechanical properties are much less studied compared to other properties. Here, we have calculated the mechanical properties of the 21 pyrophosphate crystals, using the method described in Section 2 in the Supplementary Material. The calculated elastic constants are listed in Table S2 in the SI, which provides a profusion of information about the stability, stiffness, brittleness, ductility, and anisotropy of these crystals. C11, C22, and C33 coefficients reflect the isotropic elasticity of crystals, and they have very close values in the cubic crystals. C11, C22, and C33 are associated with unidirectional compression along the principal x, y, and z directions [96]. Synonymously, C11, C22, and C33 represent the resistance of a crystal to the deformation along x, y, and z directions. Large C11, C22, and C33 elastic constants of a material indicate that it is very incompressible under uniaxial stress along x, y, or z axes, respectively. The value of C11 for triclinic crystals: 2, 4, and 6 is much larger than C11 of the orthorhombic crystals: 1, 3, and 5, which indicates that triclinic crystals: 2, 4, and 6 are much more incompressible under uniaxial stress along x than the orthorhombic ones. The values of the elastic constants depend on the crystal structure (lattice parameters) and the strength of the bonds between the elements. Table S2 shows that C33 is higher than C11 and C22 in crystal 1, which implies that this crystal is more compressible along x- and y-axes than along the z-axis. C22 is higher than C11 and C33 in crystal 3, which implies that this crystal is more compressible along x- and z-axes than along the y-axis. C11 is higher than C22 and C33 in crystal 2 and crystals 4–21, so these crystals are more compressible along y- and z-axes than along the x-axis. The 21 crystals have lower symmetry and a larger number of independent elastic constants (Cij). For the orthorhombic crystal structure, the mechanical stability criteria are given by [97,98]:
C11 > 0; C22 > 0; C33 > 0; C44 > 0; C55 > 0; C66 > 0; C11C22 > C212; C11 + C22 − 2C12 > 0
The calculated elastic constants for the orthorhombic phase crystals: 1, 3, and 5, fulfill the stability criteria above. For the other triclinic crystals which have even lower symmetry and a larger number of independent elastic constants than the orthorhombic crystals, the mechanical stability criterion requires all values of Cij to be positive [97]. All elastic constant values are presented in Tables S2–S4 in the SI. From these values, we conclude that the crystals 4, 6, 7, 9, 11, 12, 14, 15, 18, and 21 do not match this criterion and they are mechanically unstable crystals.
From the elastic coefficient Cij and compliance tensor Sij (Sij = 1/Cij), the other mechanical parameters: bulk modulus (K), Young’s modulus (E), shear modulus (G), and Poisson’s ratio (η) can be obtained under the Voigt–Reuss–Hill (VRH) approximation for polycrystals as explained in Section 2 in the Supplementary Material [99,100]. These parameters are listed in Table S1. Generally speaking, Young’s modulus can measure the stiffness of the materials, or Young’s modulus measures the change in length. Bulk modulus is a measure of resistance to compressibility, or bulk modulus measures the change in volume when a pressure is applied. Shear modulus represents the resistance against shear distortion. Supplemental Figure S10a–d show the distribution of the bulk modulus, Young’s modulus, shear modulus, and Poisson’s ration (η) of the 21 crystals. The 21 crystals were divided into four groups regarding the A element, similar to Figure 5 for the TBOD. We can notice from Supplemental Figure S10a,b that the crystals 12–16 have the highest Young’s modulus and bulk modulus. As can be seen in Table S1, most pyrophosphate crystals have a bulk modulus less than 30 GPa, except for crystals 4 and 12–16. In general, if the obtained bulk moduli are quite small, i.e., <30 GPa, the materials can be classified as relatively soft materials with high compressibility. Another useful parameter that can be used to classify the brittle and ductile behaviors of a material is the ratio of the shear modulus to the bulk modulus, G/K. It is an empirical relationship related to the plastic and elastic properties of the material, and it is called Pugh’s modulus ratio [101,102]. G/K values for the 21 crystals are represented in Table S1. According to Pugh’s criterion, crystals with G/K larger than 0.571 tend to be brittle and those with G/K less than 0.571 tend to be more ductile [102,103,104,105]. The crystals 2, 7, 8, 10, and 12 have G/K larger than 0.571, which indicates that these five crystals tend to be more brittle. All remaining crystals have G/K less than 0.571, which indicates that these crystals tend to be more ductile. Further, the Vickers hardness (HV) formula was used to calculate the hardness values of pyrophosphate crystals under study. Vickers hardness formula, which was formulated by Chen et al. [106] and Tian et al. [107], can be derived from the elastic constants, as shown earlier [108,109,110]. In this work, we used the formula of Tian et al. [107]:
H V = 0.92 ( G K ) 1.137 G 0.708
It is assumed that materials with Vickers hardness larger than 40 GPa can be classified as super-hard materials [111]. The calculated values of HV of all 21 crystals are presented in Table S1 in the SI. As can be seen in this table, HV has value less than 10 GPa for all crystals, so these 21 crystals cannot be considered as hard materials, or they are soft materials.
The elastic anisotropy parameter can be characterized by the universal anisotropic index AU [112]:
A U = 5 G V G R + K V K R 6
where GV, GR, KV, and VR are the shear and bulk modulus’ Voigt and Reuss estimates respectively. The calculated values of AU are given in Table S5. If AU has a value of unity, the crystal exhibits an isotropic characteristic, while values other than unity represent varying degrees of anisotropy [113]. After calculation, AU for all 21 crystals varies from unity (in different degrees), indicating anisotropic behavior for all crystals. However, crystals 5 and 12 show weak anisotropy. The percentage of anisotropy in the compression and shear is another way of measuring the elastic anisotropy. They are given by [114]:
A c o m p = K V K R K V + K R
A s h e a r = G V G R G V + G R
The calculated values of Acomp and Ashear are given in Table S5. The value of zero for Acomp and Ashear indicates isotropic elastic behavior, while the variation from zero refers to varying degrees of anisotropy, and a value of 1 (100%) is the largest possible anisotropy. To characterize the materials as ductile or brittle, Frantsevich’s rule of Poisson’s ratio [115] can be applied here. According to this rule, if the Poisson’s ratio (η) is less than 0.26, the material tends to be brittle in nature; otherwise, the material will be ductile. As can be seen in Table S1, most crystals have η higher than 0.26, so they can be considered as ductile materials, except for crystals 2, 7, 8, 10, and 12, which have η less than 0.26, so they can be considered as brittle materials, and this result was exactly the same when we applied the Pugh’s ratio rule. To reveal the correlation between the electronic structure/bonding characteristics and the mechanical properties of the pyrophosphate crystals, we plotted bulk modulus (Figure 6a), Young’s modulus (Figure 6b), and shear modulus (Figure 6c) versus TBOD for the 21 crystals. As can be seen in these figures, E, K, and G steadily increase with the TBOD, which allows us to conclude that the mechanical properties of these 21 crystals are strongly correlated to the strength of the interatomic bonds which are collectively represented by the single metric TBOD.

4. Conclusions

We have studied 21 pyrophosphate crystals in the framework of density functional theory. By using first-principles calculations, the electronic structure, bonding, charge transfer, optical properties, and mechanical properties were calculated. Electronic structure calculations show the semi-metallic nature of some crystals and semi-conducting behavior for others. Bonding characteristics show that the two most common bonds in all crystal are the P-O and H-O bonds. The polar covalent bonds P-O are found to be the strongest bonds in most crystals, expect for crystals numbered from 12 to 16, which contain N-H bonds that are a little stronger than P-O bonds. P-O and N-H bonds come from H2P2O7 and NH4 units. The metal-oxygen bonds: Zn-O, Cu-O, Ni-O, Mg-O, Co-O, and Mn-O in all crystals are comparable. NH4 is a unique group of atoms that replace the metallic ion (M) in the structure formula: A2M (H2P2O7)2•2H2O, and this replacement caused the crystals 12 to 16 to have the largest values of TBOD. The electronic structure and bonding in these crystals appear to be quite similar, but there are subtle differences here and there due to the differences in their compositions in different metallic elements and crystal geometry. The optical dispersion analysis of complex dielectric function ε, energy loss function ELF, and refractive index n was also conducted in the energy range of 0–40 eV for ε and 0–50 eV for ELF. The calculated optical results of all investigated materials show that the highest absorption of incident photon occurs in the ultraviolet (UV) region, and these crystals have medium values of n ranging from 1.446 to 2.098. Spin-polarized DFT calculations of the electronic structure were conducted for the crystals with (M = Cu, Ni, Co, Mn) to check if there are half-metallic ferromagnetic characteristics. Our calculations show that crystals 3, 4, 5, 7, 11, 15, 16, 20, and 21 have half-metallic (HM) behavior, while the crystals 6, 14, and 18 have identical electronic behavior in both spin-up and spin-down directions of electrons. Mechanical properties calculations show that most of these 21 crystals are ductile, soft, and anisotropic materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst12081139/s1, Table S1: Young’s modulus (E), bulk modulus (K), shear modulus (G), Poisson’s ratio (η), Pugh’s ratio (G/K), and Vicker’s hardness(HV) for the 21 crystals; Table S2: The calculated elastic constants Cij (GPa) for the 21 crystals; Table S3: The calculated elastic constants Cij (GPa) for the crystals 2, 4, 6, and the crystals from 7 to 21; Table S4: The calculated elastic constants Cij (GPa) for the crystals 2, 4, 6, and the crystals from 7 to 21; Table S5: The calculated bulk and shear modulus’ Voigt and Reuss estimates (KV, KR, GV, GR), the universal anisotropic index (AU), and the percentage of anisotropy in the compression and shear for the 21 crystals (Acomp, Ashear); Figure S1: Calculated partial density of states (PDOS) for the 21 crystals; Figure S2: Calculated TDOS and PDOS of crystals from 1 to 9, and crystals: 15, 16, 18, and 21; Figure S3: Calculated BO vs BL for the 21 crystals; Figure S4: Crystal structure of crystals 2 in ball and stick format. Crystal 2 with triclinic phase; Figure S5: Calculated partial charge distribution in the 21 pyrophosphate crystals; Figure S6: Calculated optical dielectric functions for the 21 crystals; Figure S7: Calculated energy lose function (ELF) for the 21 crystals; Figure S8: Calculated TDOS with spin-polarized effect for the crystals 3, 4, 5, 6, 11, 14, 15, 16, 18, 20, and 21. The abbreviation #_2 means that this figure belong to the same # crystal but in different range of energy. For example, 3_1 and 3_2 figures represent the same crystal; Figure S9: Calculated spin –up and spin-down refractive indices of crystals: 3, 4, 11, 15, 16, 20, and 21. The abbreviation #_2 means that this figure belong to the same # crystal but in different range of energy. For example, 3_1 and 3_2 figures represent the same crystal; Figure S10: (a–d) Distribution of bulk modulus (a), Young’s modulus (b), shear modulus (c), and Poisson’s ratio (d) for the 21 crystals. The black color is for the crystals from 1 to 8, red color is for crystals from 9 to 11, green color is for crystals from 12 to 16, and blue color is for crystals from 17 to 21. Ref. [116] is cited in Supplementary Materials.

Author Contributions

W.-Y.C. conceived and directed the project. S.H. performed the calculations. S.H. made all figures. S.H., P.R. and W.-Y.C. drafted the paper. All authors participated in the discussion and interpretation of the results. All authors edited and proofread the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This research used the resources of the National Energy Research Scientific Computing Center supported by DOE under Contract No. DE-AC03-76SF00098 and the Research Computing Support Services (RCSS) of the University of Missouri System.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fang, W.; An, Z.; Xu, J.; Zhao, H.; Zhang, J. Superior performance of Na7V4(P2O7)4PO4 in sodium ion batteries. RSC Adv. 2018, 8, 21224–21228. [Google Scholar] [CrossRef] [PubMed]
  2. Shen, Y.; Zhao, S.; Zhao, B.; Ji, C.; Li, L.; Sun, Z.; Hong, M.; Luo, J. Strong Nonlinear-Optical Response in the Pyrophosphate CsLiCdP2O7 with a Short Cutoff Edge. Inorg. Chem. 2016, 55, 11626–11629. [Google Scholar] [CrossRef] [PubMed]
  3. Yu, H.; Young, J.; Wu, H.; Zhang, W.; Rondinelli, J.M.; Halasyamani, P.S. M4Mg4(P2O7)3 (M = K, Rb): Structural Engineering of Pyrophosphates for Nonlinear Optical Applications. Chem. Mater. 2017, 29, 1845–1855. [Google Scholar] [CrossRef]
  4. Yu, H.; Cantwell, J.; Wu, H.; Zhang, W.; Poeppelmeier, K.R.; Halasyamani, P.S. Top-Seeded Solution Crystal Growth, Morphology, Optical and Thermal Properties of Ba3(ZnB5O10)PO4. Cryst. Growth Des. 2016, 16, 3976–3982. [Google Scholar] [CrossRef]
  5. Wu, C.; Li, L.; Yang, G.; Song, J.; Yan, B.; Humphrey, M.G.; Zhang, L.; Shao, J.; Zhang, C. Low-temperature-flux syntheses of ultraviolet-transparent borophosphates Na4MB2P3O13(M = Rb, Cs) exhibiting a second-harmonic generation response. Dalton Trans. 2017, 46, 12605–12611. [Google Scholar] [CrossRef]
  6. Jiang, J.-H.; Zhang, L.-C.; Huang, Y.-X.; Sun, Z.-M.; Pan, Y.; Mi, J.-X. KB(PO4)F: A novel acentric deep-ultraviolet material. Dalton Trans. 2017, 46, 1677–1683. [Google Scholar] [CrossRef]
  7. Zhou, Y.; Cao, L.; Lin, C.; Luo, M.; Yan, T.; Ye, N.; Cheng, W. AMgPO4·6H2O (A = Rb, Cs): Strong SHG responses originated from orderly PO4 groups. J. Mater. Chem. C 2016, 4, 9219–9226. [Google Scholar] [CrossRef]
  8. Li, L.; Wang, Y.; Lei, B.-H.; Han, S.; Yang, Z.; Poeppelmeier, K.R.; Pan, S. A New Deep-Ultraviolet Transparent Orthophosphate LiCs2PO4 with Large Second Harmonic Generation Response. J. Am. Chem. Soc. 2016, 138, 9101–9104. [Google Scholar] [CrossRef]
  9. Zhang, X.; Mo, F.; Zhou, L.; Gong, M. Properties-structure relationship research on LiCaPO4: Eu2+ as blue phosphor for NUV LED application. J. Alloy. Compd. 2013, 575, 314–318. [Google Scholar] [CrossRef]
  10. Dong, W.; Sun, Y.; Yao, Q.; Wang, Q.; Wen, H.; Li, J.; Xu, X.; Wang, J. Te3O3(PO4)2: A phosphate crystal with large birefringence activated by the highly distorted [TeO5] group and antiparallel [PO4] pseudo-layer. J. Mater. Chem. C 2020, 8, 9585–9592. [Google Scholar] [CrossRef]
  11. Liu, Q.; Liao, Z.; Axinte, D. Temperature effect on the material removal mechanism of soft-brittle crystals at nano/micron scale. Int. J. Mach. Tools Manuf. 2020, 159, 103620. [Google Scholar] [CrossRef]
  12. Cheng, J.; Xiao, Y.; Liu, Q.; Yang, H.; Zhao, L.; Chen, M.; Tan, J.; Liao, W.; Chen, J.; Yuan, X. Effect of surface scallop tool marks generated in micro-milling repairing process on the optical performance of potassium dihydrogen phosphate crystal. Mater. Des. 2018, 157, 447–456. [Google Scholar] [CrossRef]
  13. Essehli, R.; Sabri, S.; El-Mellouhi, F.; Aïssa, B.; Ben Yahia, H.; Altamash, T.; Khraisheh, M.; Amhamed, A.; El Bali, B. Single crystal structure, vibrational spectroscopy, gas sorption and antimicrobial properties of a new inorganic acidic diphosphates material (NH4)2Mg(H2P2O7)2 2H2O. Sci. Rep. 2020, 10, 8909. [Google Scholar] [CrossRef]
  14. Walker, J.E. ATP Synthesis by Rotary Catalysis (Nobel lecture). Angew. Chem. Int. Ed. Engl. 1998, 37, 2308–2319. [Google Scholar] [CrossRef]
  15. Boyer, P.D. The recognition of binding changes of catalytic sites of the ATP synthase…… was followed by the postulate of a rotational catalysis that has now been Energy, Life, and ATP (Nobel Lecture). Angew. Chem. Int. Ed. 1998, 37, 2296–2307. [Google Scholar] [CrossRef]
  16. Ben Saad, A.; Ratel-Ramond, N.; Rzaigui, M.; Akriche, S. A New Photoluminescent Co(II)-Diphosphate Cluster Templated by Fampridine Cation: Synthesis and Biophysicochemical Evaluation. J. Clust. Sci. 2016, 27, 657–670. [Google Scholar] [CrossRef]
  17. Jouini, A.; Gâcon, J.; Ferid, M.; Trabelsi-Ayadi, M. Luminescence and scintillation properties of praseodymium poly and diphosphates. Opt. Mater. 2003, 24, 175–180. [Google Scholar] [CrossRef]
  18. Essehli, R.; El Bali, B.; Lachkar, M.; Cruciani, G. Two new acidic diphosphates Rb2M(H2P2O7)2·2H2O (M = Zn and Mg): Crystal structures and vibrational study. J. Alloy. Compd. 2010, 492, 358–362. [Google Scholar] [CrossRef]
  19. Boonchom, B.; Youngme, S.; Danvirutai, C. A rapid co-precipitation and non-isothermal decomposition kinetics of new binary Mn0.5Co0.5(H2PO4)2·2H2O. Solid State Sci. 2008, 10, 129–136. [Google Scholar] [CrossRef]
  20. Uebou, Y.; Okada, S.; Egashira, M.; Yamaki, J.-I. Cathode properties of pyrophosphates for rechargeable lithium batteries. Solid State Ion. 2002, 148, 323–328. [Google Scholar] [CrossRef]
  21. Kitsugi, T. Transmission electron microscopy observations at the interface of bone and four types of calcium phosphate ceramics with different calcium/phosphorus molar ratios. Biomaterials 1995, 16, 1101–1107. [Google Scholar] [CrossRef]
  22. Cheetham, A.K.; Férey, G.; Loiseau, T. Open-framework inorganic materials. Angew. Chem. Int. Ed. 1999, 38, 3268–3292. [Google Scholar] [CrossRef]
  23. Popescu, I.; Wu, Y.; Granger, P.; Marcu, I.-C. An in situ electrical conductivity study of LaCoFe perovskite-based catalysts in correlation with the total oxidation of methane. Appl. Catal. A Gen. 2014, 485, 20–27. [Google Scholar] [CrossRef]
  24. Essehli, R.; El Bali, B.; Tahiri, A.A.; Lachkar, M.; Manoun, B.; Dušek, M.; Fejfarová, K. K2M (H2P2O7)2 2H2O (M = Ni, Cu, Zn): Orthorhombic forms and Raman spectra. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 2005, 61, i120–i124. [Google Scholar] [CrossRef]
  25. Capitelli, F.; Ii, R.K.; Ii, M.H.; Ii, A.E.; Hmad, S.; Ii, H. Crystal structure and vibrational spectroscopy of the new acidic diphosphate(NH4)2Zn(H2P2O7)2·2H2O. Z. Kristallogr. 2005, 220, 25–30. [Google Scholar] [CrossRef]
  26. Essehli, R.; Lachkar, M.; Svoboda, I.; Fuess, H.; El Bali, B. (NH4)2[Ni(H2P2O7)2(H2O)2]. Acta Crystallogr. Sect. E Struct. Rep. Online 2005, 61, i64–i66. [Google Scholar] [CrossRef]
  27. Capitelli, F.; Brouzi, K.; Harcharras, M.; Ennaciri, A.; Moliterni, A.G.G.; Bertolasi, V. Two new ammonium diphosphates: Crystal structure. Z. Für Krist.-Cryst. Mater. 2004, 219, 93–98. [Google Scholar] [CrossRef]
  28. Essehli, R.; Lachkar, M.; Svoboda, I.; Fuess, H.; El Bali, B. (NH4)2[Co(H2P2O7)2(H2O)2]. Acta Crystallogr. Sect. E Struct. Rep. Online 2005, 61, 61–63. [Google Scholar] [CrossRef]
  29. Harcharras, M.; Capitelli, F.; Ennaciri, A.; Brouzi, K.; Moliterni, A.; Mattei, G.; Bertolasi, V. Synthesis, X-ray crystal structure and vibrational spectroscopy of the acidic pyrophosphate KMg0.5H2P2O7·H2O. J. Solid State Chem. 2003, 176, 27–32. [Google Scholar] [CrossRef]
  30. Tahiri, A.A.; Ouarsal, R.; Lachkar, M.; Zavalij, P.Y.; El Bali, B. Dipotassium zinc bis (dihydrogendiphosphate) dihydrate, K2Zn(H2P2O7)2·2H2O. Acta Crystallogr. Sect. E Struct. Rep. Online 2003, 59, i50–i52. [Google Scholar] [CrossRef]
  31. Khaoulaf, R.; Ennaciri, A.; Ezzaafrani, M.; Capitelli, F. Structure and Vibrational Spectra of a New Acidic Diphosphate K2Cu(H2P2O7)2·2H2O. Phosphorus Sulfur Silicon Relat. Elem. 2013, 188, 1038–1052. [Google Scholar] [CrossRef]
  32. Tahiri, A.A.; Messouri, I.; Lachkar, M.; Zavalij, P.Y.; Glaum, R.; El Bali, B.; Rachid, O. Dipotassium nickel(II) bis(dihydrogendiphosphate) dihydrate, K2Ni(H2P2O7)2·2H2O. Acta Crystallogr. Sect. E Struct. Rep. Online 2003, 60, 5–7. [Google Scholar] [CrossRef]
  33. Xu, J.-Y.; Tian, J.-L.; Zhang, Q.-W.; Zhao, J.; Yan, S.-P.; Liao, D.-Z. Tetranuclear pyrophosphate-bridged Cu(II) complex with 2,2′-dipyridylamine: Crystal structure, spectroscopy and magnetism. Inorg. Chem. Commun. 2008, 11, 69–72. [Google Scholar] [CrossRef]
  34. Edition, S. Geomicrobiology, 6th ed.; CRC Press: Boca Raton, FL, USA.
  35. Knight, W.B.; Dunaway-Mariano, D.; Ransom, S.C.; Villafranca, J.J. Investigations of the metal ion-binding sites of yeast inorganic pyrophosphatase. J. Biol. Chem. 1984, 259, 2886–2895. [Google Scholar] [CrossRef]
  36. Haromy, T.P.; Knight, W.B.; Dunaway-Mariano, D.; Sundaralingam, M. Structure of Bidentate Tetraammine (pyrophosphato) chromium (III) Dihydrate. Acta Cryst. 1984, C40, 223–226. [Google Scholar]
  37. Halonen, P.; Baykov, A.A.; Goldman, A.; Lahti, R.; Cooperman, B.S. Single-turnover kinetics of Saccharomyces cerevisiae inorganic pyrophosphatase. Biochemistry 2002, 41, 12025–12031. [Google Scholar] [CrossRef] [PubMed]
  38. Datta, A.; Agarwal, M.; Dasgupta, S. Novel vanadium phosphate phases as catalysts for selective oxidation. J. Chem. Sci. 2002, 114, 379–390. [Google Scholar] [CrossRef]
  39. Chen, B.; Munson, E.J. Investigation of the Mechanism of n-Butane Oxidation on Vanadium Phosphorus Oxide Catalysts: Evidence from Isotopic Labeling Studies. J. Am. Chem. Soc. 2002, 124, 1638–1652. [Google Scholar] [CrossRef]
  40. Ballarini, N.; Cavani, F.; Cortelli, C.; Ligi, S.; Pierelli, F.; Trifirò, F.; Fumagalli, C.; Mazzoni, G.; Monti, T. VPO catalyst for n-butane oxidation to maleic anhydride: A goal achieved, or a still open challenge? Top. Catal. 2006, 38, 147–156. [Google Scholar] [CrossRef]
  41. Papa, S. Mitochondrial oxidative phosphorylation changes in the life span. Molecular aspects and physiopathological implications. Biochim. Biophys. Acta-Bioenerg. 1996, 1276, 87–105. [Google Scholar] [CrossRef]
  42. Lipscomb, W.N.; Sträter, N. Recent Advances in Zinc Enzymology. Chem. Rev. 1996, 96, 2375–2434. [Google Scholar] [CrossRef] [PubMed]
  43. Kulaev, I.S.; Vagabov, V.M.; Kulakovskaya, T.V. The Biochemistry of Inorganic Polyphosphates, 2nd ed.; Wiley: Hoboken, NJ, USA, 2004. [Google Scholar] [CrossRef]
  44. Arnstein, H. The biochemistry of the nucleic acids. 10th edition. FEBS Lett. 1987, 219, 486. [Google Scholar] [CrossRef]
  45. Kajander, T.; Kellosalo, J.; Goldman, A. Inorganic pyrophosphatases: One substrate, three mechanisms. FEBS Lett. 2013, 587, 1863–1869. [Google Scholar] [CrossRef] [PubMed]
  46. Baykov, A.A.; Cooperman, B.S.; Goldman, A.; Lahti, R. Cytoplasmic Inorganic Pyrophosphatase. Prog. Mol. Subcell. Biol. 1999, 23, 127–150. [Google Scholar] [CrossRef] [PubMed]
  47. Motta, L.S.; Ramos, I.B.; Gomes, F.M.; de Souza, W.; Champagne, D.E.; Santiago, M.F.; Docampo, R.; Miranda, K.; Machado, E.A. Proton-pyrophosphatase and polyphosphate in acidocalcisome-like vesicles from oocytes and eggs of Periplaneta americana. Insect Biochem. Mol. Biol. 2009, 39, 198–206. [Google Scholar] [CrossRef]
  48. Wilcox, D.E. Binuclear Metallohydrolases. Chem. Rev. 1996, 96, 2435–2458. [Google Scholar] [CrossRef]
  49. Shintani, T.; Uchiumi, T.; Yonezawa, T.; Salminen, A.; Baykov, A.A.; Lahti, R.; Hachimori, A. Cloning and expression of a unique inorganic pyrophosphatase from Bacillus subtilis: Evidence for a new family of enzymes. FEBS Lett. 1998, 439, 263–266. [Google Scholar] [CrossRef]
  50. Kuhn, N.J.; Wadeson, A.; Ward, S.; Young, T.W. Methanococcus jannaschii ORF mj0608 Codes for a Class C Inorganic Pyrophosphatase Protected by Co2+ or Mn2+ Ions against Fluoride Inhibition. Arch. Biochem. Biophys. 2000, 379, 292–298. [Google Scholar] [CrossRef]
  51. Moreira, B.L.M.; Medeiros, L.C.A.S.; Miranda, K.; De Souza, W.; Hentschel, J.; Plattner, H.; Barrabin, H. Kinetics of pyrophosphate-driven proton uptake by acidocalcisomes of Leptomonas wallacei. Biochem. Biophys. Res. Commun. 2005, 334, 1206–1213. [Google Scholar] [CrossRef]
  52. Mimura, H.; Nakanishi, Y.; Maeshima, M. Oligomerization of H+-pyrophosphatase and its structural and functional consequences. Biochim. Biophys. Acta-Bioenerg. 2005, 1708, 393–403. [Google Scholar] [CrossRef]
  53. Westfall, D.; Aboushadi, N.; Shackelford, J.E.; Krisans, S.K. Metabolism of Farnesol: Phosphorylation of Farnesol by Rat Liver Microsomal and Peroxisomal Fractions. Biochem. Biophys. Res. Commun. 1997, 230, 562–568. [Google Scholar] [CrossRef]
  54. Gdula, D.A.; Sandaltzopoulos, R.; Tsukiyama, T.; Ossipow, V.; Wu, C. Inorganic pyrophosphatase is a component of the Drosophila nucleosome remodeling factor complex. Genes Dev. 1998, 12, 3206–3216. [Google Scholar] [CrossRef] [PubMed]
  55. Harutyunyan, E.H.; Kuranova, I.P.; Vainshtein, B.K.; Hohne, W.E.; Lamzin, V.; Dauter, Z.; Teplyakov, A.V.; Wilson, K. X-ray Structure of Yeast Inorganic Pyrophosphatase Complexed with Manganese and Phosphate. JBIC J. Biol. Inorg. Chem. 1996, 239, 220–228. [Google Scholar] [CrossRef] [PubMed]
  56. Gómez-García, M.R.; Losada, M.; Serrano, A. A novel subfamily of monomeric inorganic pyrophosphatases in photosynthetic eukaryotes. Biochem. J. 2006, 395, 211–221. [Google Scholar] [CrossRef]
  57. Kim, H.; Park, J.; Park, I.; Jin, K.; Jerng, S.E.; Kim, S.H.; Nam, K.T.; Kang, K. Coordination tuning of cobalt phosphates towards efficient water oxidation catalyst. Nat. Commun. 2015, 6, 8253. [Google Scholar] [CrossRef] [PubMed]
  58. Zhang, Y.; Cheng, W.; Wu, D.; Zhang, H.; Chen, D.; Gong, Y.; Kan, Z. Crystal and Band Structures, Bonding, and Optical Properties of Solid Compounds of Alkaline Indium (III) Pyrophosphates MInP2O7 (M = Na, K, Rb, Cs). Chem. Mater. 2004, 16, 4150–4159. [Google Scholar] [CrossRef]
  59. Witko, M.; Tokarz, R.; Haber, J.; Hermann, K. Electronic structure of vanadyl pyrophosphate: Cluster model studies. J. Mol. Catal. A Chem. 2001, 166, 59–72. [Google Scholar] [CrossRef]
  60. Xiang, H.; Feng, Z.; Zhou, Y. Ab initio computations of electronic, mechanical, lattice dynamical and thermal properties of ZrP2O7. J. Eur. Ceram. Soc. 2014, 34, 1809–1818. [Google Scholar] [CrossRef]
  61. Elhafiane, F.; Khaoulaf, R.; Harcharras, M.; Brouzi, K. Synthesis, crystal structure, vibrational spectroscopy and electrochemical investigations of a new acidic metal pyrophosphate NiK1.18N0.82(H2P2O7)2.2H2O. J. Mol. Struct. 2021, 1245, 131234. [Google Scholar] [CrossRef]
  62. Elhafiane, F.Z.; Brouzi, K.; Khaoulaf, R.; Rghioui, L.; Harcharras, M. New acidic pyrophosphate CoK1.078N0.922(H2P2O7)2.2H2O. Crystal structure and vibrational spectroscopy. Phosphorus Sulfur Silicon Relat. Elem. 2019, 195, 173–179. [Google Scholar] [CrossRef]
  63. Khaoulaf, R.; Ezzaafrani, M.; Ennaciri, A.; Harcharras, M.; Capitelli, F. Vibrational Study of Dipotassium Zinc Bis(Dihydrogendiphosphate) Dihydrate, K2ZN(H2P2O7)2·2H2O. Phosphorus Sulfur Silicon Relat. Elem. 2012, 187, 1367–1376. [Google Scholar] [CrossRef]
  64. Khaoulaf, R.; Adhikari, P.; Harcharras, M.; Brouzi, K.; Ez-Zahraouy, H.; Ching, W.-Y. Atomic-Scale Understanding of Structure and Properties of Complex Pyrophosphate Crystals by First-Principles Calculations. Appl. Sci. 2019, 9, 840. [Google Scholar] [CrossRef]
  65. Adhikari, P.; Khaoulaf, R.; Ez-zahraouy, H.; Ching, W. Subject Category: Subject Areas: Complex interplay of interatomic bonding in a K2Mg (H2P2O7)2 2H2O. R. Soc. Open sci. 2017, 4, 170982. [Google Scholar] [CrossRef]
  66. Lamhamdi, A.; Essehli, R.; El Bali, B.; Dusek, M.; Fejfarov, K. The triclinic form of dipotassium cobalt(II) bis(dihydrogendiphosphate) dihydrate. Acta Crystallogr. Sect. E Struct. Rep. Online 2011, 67, i37–i38. [Google Scholar] [CrossRef]
  67. El Bali, B.; Capitelli, F.; Alaoui, A.T.; Lachkar, M.; da Silva, I.; Alvarez-Larena, A.; Piniella, J.F. New thallium diphosphates Tl2 Me(H2P2O7)2 2H2O, Me = Mg, Mn, Co, Ni and Zn. Synthesis, single crystal X-ray structures and powder X-ray structure of the Mg phase. Z. Fur Krist. 2008, 223, 448–455. [Google Scholar] [CrossRef]
  68. Deyab, M.A.; Essehli, R.; El Bali, B.; Lachkar, M. Fabrication and evaluation of Rb2Co(H2P2O7)2·2H2O/waterborne polyurethane nanocomposite coating for corrosion protection aspects. RSC Adv. 2017, 7, 55074–55080. [Google Scholar] [CrossRef]
  69. Kresse, G. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis se. Phys. Rev. B. 1996, 54, 11169. [Google Scholar] [CrossRef]
  70. Mo, Y.; Rulis, P.; Ching, W.Y. Electronic structure and optical conductivities of 20 MAX-phase compounds. Phys. Rev. B-Condens. Matter Mater. Phys. 2012, 86, 165122. [Google Scholar] [CrossRef]
  71. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  72. Monkhorst, H.J.; Pack, J.D. Special points for Brillonin-zone integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  73. Hasan, S.; Adhikari, P.; Baral, K.; Ching, W.-Y. Conspicuous interatomic bonding in chalcogenide crystals and implications on electronic, optical, and elastic properties. AIP Adv. 2020, 10, 075216. [Google Scholar] [CrossRef]
  74. Hasan, S.; Baral, K.; Li, N.; Ching, W.-Y. Structural and physical properties of 99 complex bulk chalcogenides crystals using first-principles calculations. Sci. Rep. 2021, 11, 9921. [Google Scholar] [CrossRef] [PubMed]
  75. Dharmawardhana, C.; Bakare, M.; Misra, A.; Ching, W. Nature of Interatomic Bonding in Controlling the Mechanical Properties of Calcium Silicate Hydrates. J. Am. Ceram. Soc. 2016, 99, 2120–2130. [Google Scholar] [CrossRef]
  76. Poudel, L.; Tamerler, C.; Misra, A.; Ching, W.-Y. Atomic-Scale Quantification of Interfacial Binding between Peptides and Inorganic Crystals: The Case of Calcium Carbonate Binding Peptide on Aragonite. J. Phys. Chem. C 2017, 121, 28354–28363. [Google Scholar] [CrossRef]
  77. San, S.; Li, N.; Tao, Y.; Zhang, W.; Ching, W.-Y. Understanding the atomic and electronic origin of mechanical property in thaumasite and ettringite mineral crystals. J. Am. Ceram. Soc. 2018, 101, 5177–5187. [Google Scholar] [CrossRef]
  78. Hasan, S.; San, S.; Baral, K.; Li, N.; Rulis, P.; Ching, W.-Y. First-Principles Calculations of Thermoelectric Transport Properties of Quaternary and Ternary Bulk Chalcogenide Crystals. Materials 2022, 15, 2843. [Google Scholar] [CrossRef] [PubMed]
  79. Hunca, B.; Dharmawardhana, C.; Sakidja, R.; Ching, W.-Y. Ab initio calculations of thermomechanical properties and electronic structure of vitreloy Zr41.2Ti13.8Cu12.5Ni10Be22.5. Phys. Rev. B 2016, 94, 144207. [Google Scholar] [CrossRef]
  80. Ching, W.; Yoshiya, M.; Adhikari, P.; Rulis, P.; Ikuhara, Y.; Tanaka, I. First-principles study in an inter-granular glassy film model of silicon nitride. J. Am. Ceram. Soc. 2018, 101, 2673–2688. [Google Scholar] [CrossRef]
  81. Poudel, L.; Twarock, R.; Steinmetz, N.F.; Podgornik, R.; Ching, W.-Y. Impact of Hydrogen Bonding in the Binding Site between Capsid Protein and MS2 Bacteriophage ssRNA. J. Phys. Chem. B 2017, 121, 6321–6330. [Google Scholar] [CrossRef]
  82. Adhikari, P.; Li, N.; Shin, M.; Steinmetz, N.F.; Twarock, R.; Podgornik, R.; Ching, W.-Y. Intra- and intermolecular atomic-scale interactions in the receptor binding domain of SARS-CoV-2 spike protein: Implication for ACE2 receptor binding. Phys. Chem. Chem. Phys. 2020, 22, 18272–18283. [Google Scholar] [CrossRef]
  83. Mulliken, R.S. Electronic population analysis on LCAO-MO molecular wave functions. J. Chem. Phys. 1955, 23, 1833–1840. [Google Scholar] [CrossRef]
  84. Dharmawardhana, C.C.; Misra, A.; Ching, W.-Y. Quantum Mechanical Metric for Internal Cohesion in Cement Crystals. Sci. Rep. 2014, 4, 7332. [Google Scholar] [CrossRef]
  85. Schulte, J.E.; Goulian, M. The Phosphohistidine Phosphatase SixA Targets a Phosphotransferase System. mBio 2018, 9, e01666-18. [Google Scholar] [CrossRef] [PubMed]
  86. Hunter, T. A journey from phosphotyrosine to phosphohistidine and beyond. Mol. Cell 2022, 82, 2190–2200. [Google Scholar] [CrossRef] [PubMed]
  87. Gohla, A. Do metabolic HAD phosphatases moonlight as protein phosphatases? Biochim. Biophys. Acta-Mol. Cell Res. 2018, 1866, 153–166. [Google Scholar] [CrossRef] [PubMed]
  88. Makwana, M.V.; Muimo, R.; Jackson, R.F. Advances in development of new tools for the study of phosphohistidine. Lab. Investig. 2017, 98, 291–303. [Google Scholar] [CrossRef]
  89. Hindupur, S.K.; Colombi, M.; Fuhs, S.R.; Matter, M.; Guri, Y.; Adam, K.; Cornu, M.; Piscuoglio, S.; Ng, C.K.Y.; Betz, C.; et al. The protein histidine phosphatase LHPP is a tumour suppressor. Nature 2018, 555, 678–682. [Google Scholar] [CrossRef]
  90. Schulte, J.E.; Roggiani, M.; Shi, H.; Zhu, J.; Goulian, M. The phosphohistidine phosphatase SixA dephosphorylates the phosphocarrier NPr. J. Biol. Chem. 2021, 296, 100090. [Google Scholar] [CrossRef]
  91. Williams, W.S. The thermal conductivity of metallic ceramics. JOM 1998, 50, 62–66. [Google Scholar] [CrossRef]
  92. Madsen, G.K.H.; Blaha, P.; Schwarz, K.; Sjöstedt, E.; Nordström, L. Efficient linearization of the augmented plane-wave method. Phys. Rev. B-Condens. Matter Mater. Phys. 2001, 64, 195134. [Google Scholar] [CrossRef]
  93. Irfan, M.; Azam, S.; Dahshan, A.; El Bakkali, I.; Nouneh, K. First-principles study of opto-electronic and thermoelectric properties of SrCdSnX4 (X=S, Se, Te) alkali metal chalcogenides. Comput. Condens. Matter 2021, 30, e00625. [Google Scholar] [CrossRef]
  94. Žutić, I.; Fabian, J.; Das Sarma, S. Spintronics: Fundamentals and applications. Rev. Mod. Phys. 2004, 76, 323. [Google Scholar] [CrossRef]
  95. Boehnke, A.; Martens, U.; Sterwerf, C.; Niesen, A.; Huebner, T.; Von Der Ehe, M.; Meinert, M.; Kuschel, T.; Thomas, A.; Heiliger, C.; et al. Large magneto-Seebeck effect in magnetic tunnel junctions with half-metallic Heusler electrodes. Nat. Commun. 2017, 8, 1626. [Google Scholar] [CrossRef] [PubMed]
  96. Deng, H. Theoretical prediction of the structural, electronic, mechanical and thermodynamic properties of the binary α-As 2 Te 3 and β-As 2 Te 3. J. Alloy. Compd. 2016, 656, 695–701. [Google Scholar] [CrossRef]
  97. Mouhat, F.; Coudert, F.-X. Necessary and sufficient elastic stability conditions in various crystal systems. Phys. Rev. B-Condens. Matter Mater. Phys. 2014, 90, 224104. [Google Scholar] [CrossRef]
  98. Beckstein, O.; Klepeis, J.E.; Hart, G.L.W.; Pankratov, O. First-principles elastic constants and electronic structure ofα−Pt2Siand PtSi. Phys. Rev. B-Condens. Matter Mater. Phys. 2001, 63, 134112. [Google Scholar] [CrossRef]
  99. Reuss, A. Berechnung der Fließgrenze von Mischkristallen auf Grund der Plastizitätsbedingung für Einkristalle. ZAMM-J. Appl. Math. Mech./Z. Für Angew. Math. Und Mech. 1929, 9, 49–58. [Google Scholar] [CrossRef]
  100. Hill, R. The Elastic Behaviour of a Crystalline Aggregate. Proc. Phys. Soc. Sect. A 1952, 65, 349. [Google Scholar] [CrossRef]
  101. Pugh, S.F. XCII. Relations between the elastic moduli and the plastic properties of polycrystalline pure metals. Philos. Mag. Ser. 1954, 45, 823–843. [Google Scholar] [CrossRef]
  102. Varshney, D.; Jain, S.; Shriya, S.; Khenata, R. High-pressure and temperature-induced structural, elastic, and thermodynamical properties of strontium chalcogenides. J. Theor. Appl. Phys. 2016, 10, 163–193. [Google Scholar] [CrossRef]
  103. Ali, M.A.; Hossain, M.A.; Rayhan, M.A.; Hossain, M.M.; Uddin, M.M.; Roknuzzaman, M.; Ostrikov, K.; Islam, A.K.M.A.; Naqib, S.H. First-principles study of elastic, electronic, optical and thermoelectric properties of newly synthesized K2Cu2GeS4 chalcogenide. J. Alloy. Compd. 2019, 781, 37–46. [Google Scholar] [CrossRef]
  104. Benkabou, A.; Bouafia, H.; Sahli, B.; Abidri, B.; Ameri, M.; Hiadsi, S.; Rached, D.; Bouhafs, B.; Benkhettou, N.; Al-Douri, Y. Structural, elastic, electronic and thermodynamic investigations of neptunium chalcogenides: First-principles calculations. Chin. J. Phys. 2016, 54, 33–41. [Google Scholar] [CrossRef]
  105. Heciri, D.; Belkhir, H.; Belghit, R.; Bouhafs, B.; Khenata, R.; Ahmed, R.; Bouhemadou, A.; Ouahrani, T.; Wang, X.; Bin Omran, S. Insight into the structural, elastic and electronic properties of tetragonal inter-alkali metal chalcogenides CsNaX (X=S, Se, and Te) from first-principles calculations. Mater. Chem. Phys. 2019, 221, 125–137. [Google Scholar] [CrossRef]
  106. 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]
  107. Tian, Y.; Xu, B.; Zhao, Z. Microscopic theory of hardness and design of novel superhard crystals. Int. J. Refract. Met. Hard Mater. 2012, 33, 93–106. [Google Scholar] [CrossRef]
  108. Liu, Z.; Zhou, X.; Gall, D.; Khare, S. First-principles investigation of the structural, mechanical and electronic properties of the NbO-structured 3d, 4d and 5d transition metal nitrides. Comput. Mater. Sci. 2014, 84, 365–373. [Google Scholar] [CrossRef]
  109. Liu, Z.T.Y.; Zhou, X.; Khare, S.V.; Gall, D. Structural, mechanical and electronic properties of 3d transition metal nitrides in cubic zincblende, rocksalt and cesium chloride structures: A first-principles investigation. J. Phys. Condens. Matter 2014, 26, 025404. [Google Scholar] [CrossRef]
  110. Zhou, X.; Gall, D.; Khare, S.V. Mechanical properties and electronic structure of anti-ReO3 structured cubic nitrides, M3N, of d block transition metals M: An ab initio study. J. Alloy. Compd. 2014, 595, 80–86. [Google Scholar] [CrossRef]
  111. Luan, X.; Qin, H.; Liu, F.; Dai, Z.; Yi, Y.; Li, Q. The Mechanical Properties and Elastic Anisotropies of Cubic Ni3Al from First Principles Calculations. Crystals 2018, 8, 307. [Google Scholar] [CrossRef]
  112. Ranganathan, S.I.; Ostoja-Starzewski, M. Universal Elastic Anisotropy Index. Phys. Rev. Lett. 2008, 101, 055504. [Google Scholar] [CrossRef]
  113. Sun, L.; Gao, Y.; Xiao, B.; Li, Y.; Wang, G. Anisotropic elastic and thermal properties of titanium borides by first-principles calculations. J. Alloy. Compd. 2013, 579, 457–467. [Google Scholar] [CrossRef]
  114. Ravindran, P.; Fast, L.; Korzhavyi, P.; Johansson, B.; Wills, J.; Eriksson, O. Density functional theory for calculation of elastic properties of orthorhombic crystals: Application to TiSi2. J. Appl. Phys. 1998, 84, 4891–4904. [Google Scholar] [CrossRef]
  115. Thakur, V.; Pagare, G. Theoretical calculations of elastic, mechanical and thermal properties of REPt3 (RE = Sc, Y and Lu) intermetallic compounds based on DFT. Indian J. Phys. 2018, 92, 1503–1513. [Google Scholar] [CrossRef]
  116. Yao, H.; Ouyang, L.; Ching, W.-Y. Ab Initio Calculation of Elastic Constants of Ceramic Crystals. J. Am. Ceram. Soc. 2007, 90, 3194–3204. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of crystals 1 and 2 in ball-and-stick format: (a) is for crystal 1 (K2Zn(H2P2O7)·2H2O) with orthorhombic phase and (b) is for crystal 2 (K2Zn(H2P2O7)·2H2O) with triclinic phase.
Figure 1. Crystal structure of crystals 1 and 2 in ball-and-stick format: (a) is for crystal 1 (K2Zn(H2P2O7)·2H2O) with orthorhombic phase and (b) is for crystal 2 (K2Zn(H2P2O7)·2H2O) with triclinic phase.
Crystals 12 01139 g001
Figure 2. Distribution of calculated BO vs. BL for 1 and 2 crystals.
Figure 2. Distribution of calculated BO vs. BL for 1 and 2 crystals.
Crystals 12 01139 g002
Figure 3. Distribution of calculated BO vs. BL for crystals from 3 to 6.
Figure 3. Distribution of calculated BO vs. BL for crystals from 3 to 6.
Crystals 12 01139 g003
Figure 6. Distribution of bulk modulus (a), Young’s modulus (b), and shear modulus (c) versus TBOD for the 21 crystals.
Figure 6. Distribution of bulk modulus (a), Young’s modulus (b), and shear modulus (c) versus TBOD for the 21 crystals.
Crystals 12 01139 g006
Table 1. Experimental and calculated lattice parameters for the 21 crystals.
Table 1. Experimental and calculated lattice parameters for the 21 crystals.
#CrystalAbbreviationSpace Groupa, b, c(Å), α, β, γ (Degree)
(Experimental)
a, b, c(Å), α, β, γ (Degree)
(Our Calculations)
1K2Zn(H2P2O7)2·2H2OK2Zn-1Ortho,
Pnma(62)
9.770, 10.980, 13.420,
90.0, 90.0, 90.0 [24]
10.385, 11.229, 13.566,
90.0, 90.0, 90.0
2K2Zn(H2P2O7)2·2H2OK2Zn-2Tric, P-1(2)6.827, 7.333, 7.570,
80.753, 72.547, 83.442 [30]
6.972, 7.466, 7.630,
80.754, 73.419, 84.039
3K2Cu(H2P2O7)2·2H2OK2Cu-3Ortho,
Pnma(62)
9.899, 10.781, 13.401,
90.0, 90.0, 90.0 [24]
9.835, 11.662, 14.083,
90.0, 90.0, 90.0
4K2Cu(H2P2O7)2·2H2OK2Cu-4Tric, P-1(2)6.856, 7.314, 7.557,
81.028, 72.327, 83.697 [31]
7.074, 7.424, 7.604,
78.940, 71.785, 83.500
5K2Ni(H2P2O7)2·2H2OK2Ni-5Ortho,
Pnma(62)
9.912, 10.774, 13.422,
90.0, 90.0, 90.0 [24]
10.815, 11.132, 14.568,
90.0, 90.0, 90.0
6K2Ni(H2P2O7)2·2H2OK2Ni-6Tric, P-1(2)6.855, 7.312, 7.561,
81.012, 72.301, 83.691 [32]
7.153, 7.382, 7.658,
77.933, 70.398, 83.217
7K2Co(H2P2O7)2·2H2OK2Co-7Tric, P-1(2)6.874, 7.357, 7.614,
80.740, 72.397, 83.484 [66]
7.087, 7.647, 7.548,
77.860, 67.386, 87.327
8K2Mg(H2P2O7)2·2H2OK2Mg-8Tric, P-1(2)6.857, 7.362, 7.620,
81.044, 72.248, 83.314 [29]
6.968, 7.478, 7.662,
81.170, 73.303, 83.516
9Rb2Zn(H2P2O7)2·2H2ORb2Zn-9Tric, P-1(2)6.957, 7.362, 7.794,
81.851, 70.622, 86.263 [18]
7.068, 7.486, 7.813,
81.840, 71.764, 86.863
10Rb2Mg(H2P2O7)2·2H2ORb2Mg-10Tric, P-1(2)6.955, 7.375, 7.812,
81.986, 70.275, 85.988 [18]
7.054, 7.488, 7.835,
82.119, 71.265, 86.504
11Rb2Co(H2P2O7)2·2H2ORb2Co-11Tric, P-1(2)6.980, 7.370, 7.816,
81.740, 70.350, 86.340 [68]
7.071, 7.359, 7.709,
82.405, 71.758, 87.859
12(NH4)2Zn(H2P2O7)2·2H2O(NH4)2Zn-12Tric, P-1(2)7.003, 7.330, 7.789,
81.229, 71.064, 88.172 [25]
7.330, 6.930, 7.732,
84.653, 71.551, 88.116
13(NH4)2Mg(H2P2O7)2·2H2O(NH4)2Mg-13Tric, P-1(2)7.076, 7.431, 7.894,
81.420, 70.900, 87.780 [13]
7.138, 7.432, 7.818,
81.516, 71.444, 89.924
14(NH4)2Ni(H2P2O7)2·2H2O(NH4)2Ni-14Tric, P-1(2)7.034, 7.321, 7.792,
81.530, 70.910, 88.210 [26]
7.266, 7.338, 7.917,
79.058, 68.138, 88.466
15(NH4)2Co(H2P2O7)2·2H2O(NH4)2Co-15Tric, P-1(2)7.067, 7.368, 7.852,
81.230, 70.680, 88.540 [28]
7.220, 7.286, 7.804,
80.824, 69.610, 89.739
16(NH4)2Mn(H2P2O7)2·2H2O(NH4)2Mn-16Tric, P-1(2)7.003, 7.440, 7.877,
80.444, 71.359, 87.408 [27]
7.144, 7.322, 7.739,
81.831, 71.537, 90.547
17Tl2Zn(H2P2O7)2·2H2OTl2Zn-17Tric, P-1(2)6.963, 7.363, 7.740,
81.590, 71.441, 86.323 [67]
7.078, 7.523, 7.707,
81.487, 73.016, 86.740
18Tl2Ni(H2P2O7)2·2H2OTl2Ni-18Tric, P-1(2)6.966, 7.308, 7.702,
81.801, 71.185, 86.528 [67]
7.089, 7.432, 7.617,
81.500, 73.213, 87.502
19Tl2Mg(H2P2O7)2·2H2OTl2Mg-19Tric, P-1(2)6.962, 7.364, 7.771,
81.801, 70.860, 86.137 [67]
7.066, 7.519, 7.723,
81.825, 72.726, 86.339
20Tl2Mn(H2P2O7)2·2H2OTl2Mn-20Tric, P-1(2)6.958, 7.475, 7.825,
80.723, 71.912, 85.661 [67]
7.095, 7.439, 7.625,
81.901, 73.026, 86.945
21Tl2Co(H2P2O7)2·2H2OTl2Co-21Tric, P-1(2)6.977, 7.363, 7.769,
81.421, 71.114, 86.424 [67]
7.096, 7.399, 7.597,
81.827, 73.038, 87.597
Table 2. Energy band gap (Eg), total bond order density (TBOD), refractive index (n), and plasma frequency (ωp) for the 21 crystals.
Table 2. Energy band gap (Eg), total bond order density (TBOD), refractive index (n), and plasma frequency (ωp) for the 21 crystals.
#Crystal# of AtomsEg (eV)TBOD (e3)nωp (eV)
1K2Zn-11244.70.01941.48323.10
2K2Zn-2315.00.02171.46622.10
3K2Cu-3124-0.0188--
4K2Cu-4313.80.01221.53323.00
5K2Ni-5124-0.0086--
6K2Ni-631-0.02201.50023.10
7K2Co-731-0.0222--
8K2Mg-8315.40.02171.44922.50
9Rb2Zn-9314.90.02071.45022.00
10Rb2Mg-10315.10.02091.45621.50
11Rb2Co-11310.60.02151.71822.05
12(NH4)2Zn-12394.60.02881.48121.90
13(NH4)2Mg-13395.10.02761.44621.80
14(NH4)2Ni-14391.20.02801.48321.00
15(NH4)2Co-1539-0.02841.50021.90
16(NH4)2Mn-16390.60.02861.50321.80
17Tl2Zn-17313.050.02021.70321.10
18Tl2Ni-18312.050.0202--
19Tl2Mg-19313.20.02031.68821.10
20Tl2Mn-20310.10.02101.84420.50
21Tl2Co-21310.50.02092.09821.30
Table 3. Effective charge (Q*) for each atom in the 21 crystals.
Table 3. Effective charge (Q*) for each atom in the 21 crystals.
#CrystalQ*(in e)
1K2Zn-111.092(Zn), 6.127(K), 2.992(P), 6.836(O), 0.665(H)
2K2Zn-211.132 (Zn), 6.197(K), 2.807(P), 6.918(O), 0.569(H)
3K2Cu-310.269(Cu), 6.114(K), 2.792(P), 6.873(O), 0.670(H)
4K2Cu-410.349(Cu), 6.202(K), 2.809(P), 6.904(O), 0.569(H)
5K2Ni-59.378(Ni), 6.142(K), 3.044(P), 6.814(O), 0.643(H)
6K2Ni-69.344(Ni), 6.214(K), 2.804(P), 6.905(O), 0.567(H)
7K2Co-78.348(Co), 6.232(K), 2.799(P), 6.906(O), 0.562(H)
8K2Mg-80.826(Mg), 6.199(K), 2.816(P), 6.934(O), 0.571(H)
9Rb2Zn-911.125(Zn), 6.191(Rb), 2.812(P), 6.920(O), 0.565(H)
10Rb2Mg-100.825(Mg), 6.190(Rb), 2.821(P), 6.936(O), 0.568(H)
11Rb2Co-118.358(Co), 6.196(Rb), 2.817(P), 6.904(O), 0.566(H)
12(NH4)2Zn-1211.125(Zn), 5.803(N), 2.801(P), 6.921(O), 0.583(H)
13(NH4)2Mg-130.794(Mg), 5.807(N), 2.805(P), 6.942(O), 0.581(H)
14(NH4)2Ni-149.318(Ni), 5.816(N), 2.789(P), 6.911(O), 0.582(H)
15(NH4)2Co-158.332(Co), 5.809(N), 2.795(P), 6.910(O), 0.583(H)
16(NH4)2Mn-166.349(Mn), 5.801(N), 2.797(P), 6.910(O), 0.581(H)
17Tl2Zn-1711.124(Zn), 12.296(Tl), 2.801(P), 6.911(O), 0.563(H)
18Tl2Ni-189.322(Ni), 12.298(Tl), 2.806(P), 6.899(O), 0.559(H)
19Tl2Mg-190.814(Mg), 12.301(Tl), 2.809(P), 6.927(O), 0.565(H)
20Tl2Mn-206.346(Mn), 12.306(Tl), 2.804(P), 6.895(O), 0.562(H)
21Tl2Co-218.348(Co), 12.303(Tl), 2.807(P), 6.895(O), 0.562(H)
Table 4. Calculated spin-up and spin-down energy gap and refractive index.
Table 4. Calculated spin-up and spin-down energy gap and refractive index.
#CrystalEg (eV)n
3K2Cu-31.39 (spin-up)1.342 (spin-up), 1.483 (spin-down)
4K2Cu-43.8 (spin-up), 0.2 (spin-down)1.265 (spin-up), 1.342 (spin-down)
7K2Co-70.7 (spin-up)1.273 (spin-up)
11Rb2Co-110.6 (spin-down)1.483 (spin-up), 1.295 (spin-down)
15(NH4)2Co-150.4 (spin-up)1.265 (spin-up, spin-down)
16(NH4)2Mn-160.1 (spin-up)1.265 (spin-up, spin-down)
20Tl2Mn-200.1 (spin-up)1.449 (spin-up), 1.643 (spin-down)
21Tl2Co-210.6 (spin-down)1.789 (spin-up), 1.449 (spin-down)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hasan, S.; Rulis, P.; Ching, W.-Y. First-Principles Calculations of the Structural, Electronic, Optical, and Mechanical Properties of 21 Pyrophosphate Crystals. Crystals 2022, 12, 1139. https://doi.org/10.3390/cryst12081139

AMA Style

Hasan S, Rulis P, Ching W-Y. First-Principles Calculations of the Structural, Electronic, Optical, and Mechanical Properties of 21 Pyrophosphate Crystals. Crystals. 2022; 12(8):1139. https://doi.org/10.3390/cryst12081139

Chicago/Turabian Style

Hasan, Sahib, Paul Rulis, and Wai-Yim Ching. 2022. "First-Principles Calculations of the Structural, Electronic, Optical, and Mechanical Properties of 21 Pyrophosphate Crystals" Crystals 12, no. 8: 1139. https://doi.org/10.3390/cryst12081139

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop