Next Article in Journal
Ligand Tuning in Cu(pyalk)2 Water Oxidation Electrocatalysis
Next Article in Special Issue
Hydrogen Release and Uptake of MgH2 Modified by Ti3CN MXene
Previous Article in Journal
Oxidation of Phospholipids by OH Radical Coordinated to Copper Amyloid-β Peptide—A Density Functional Theory Modeling
Previous Article in Special Issue
Liquid Channels Built-In Solid Magnesium Hydrides for Boosting Hydrogen Sorption
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Role of Bulk Stiffening in Reducing the Critical Temperature of the Metal-to-Hydride Phase Transition and the Hydride Stability: The Case of Zr(MoxFe1−x)2-H2

1
Unit of Nuclear Engineering, Ben Gurion University of the Negev, Beer Sheva 84105, Israel
2
Department of Materials Engineering, Ben Gurion University of the Negev, Beer Sheva 84105, Israel
*
Author to whom correspondence should be addressed.
Current address: Faculty of Engineering and Bar-Ilan Institute for Nanotechnology and Advanced Materials, Bar-Ilan University, Ramat-Gan 52900, Israel.
Inorganics 2023, 11(6), 228; https://doi.org/10.3390/inorganics11060228
Submission received: 17 April 2023 / Revised: 20 May 2023 / Accepted: 23 May 2023 / Published: 25 May 2023
(This article belongs to the Special Issue State-of-the-Art and Progress in Metal-Hydrogen Systems)

Abstract

:
This study aims to shed light on the unusual trend in the stabilities of Zr(MoxFe1−x)2, 0 ≤ x ≤ 1, hydrides. Both the rule of reversed stability and the crystal volume criterion correlate with the increased hydride stabilities from x = 0 to x = 0.5, but are in contrast with the destabilization of the end member ZrMo2 hydride. The pressure-composition isotherms of ZrMo2-H2 exhibit very wide solid solubility regions, which may be associated with diminished H–H elastic interaction, uelas. In order to discern this possibility, we measured the elastic moduli of Zr(MoxFe1−x)2, x = 0, 0.5, 1. The shear modulus, G, shows a moderate variation in this composition range, while the bulk modulus, B, increases significantly and monotonically from 148.2 GPa in ZrFe2 to 200.4 GPa in ZrMo2. The H–H elastic interaction is proportional to B and therefore its increase cannot directly account for a decrease in uelas. Therefore, we turn our attention to the volume of the hydrogen atom, v H , which usually varies in a limited range. Two coexisting phases, a Laves cubic (a = 7.826 Å) and a tetragonal (a = 5.603 Å, c = 8.081 Å) hydride phase are identified in ZrMo2H3.5, obtained by cooling to liquid nitrogen temperature at about 50 atm. The volume of the hydrogen atom in these two hydrides is estimated to be 2.2 Å3/(H atom). Some very low v H values, have been reported by other investigators. The low v H values, as well as the one derived in this work, significantly reduce uelas for ZrMo2-H2, and thus reduce the corresponding critical temperature for the metal-to-hydride phase transition, and the heat of hydride formation. We suggest that the bulk stiffening in ZrMo2 confines the corresponding hydride expansion and thus reduces the H-H elastic interaction.

1. Introduction

The present study demonstrates a significant impact of the bulk modulus on the hydride formation of the intermetallic compounds ZrMo2 and TaV2. It would be instructive to consider at the outset some general factors determining the hydrogenation behavior of metal alloys. It is a well-known empirical rule that hydrogen-absorbing alloys or intermetallic compounds must contain at least one elementary metal forming a reversible hydride at nearly ambient conditions, e.g., Ref. [1]. The ability of the compounds to absorb hydrogen is thus determined by the metal–hydrogen interaction. In addition, the properties of the parent intermetallics provide some guidelines regarding the stability of the corresponding hydrides: (i) the rule of reversed stability—the more stable the original intermetallic, the less stable the resulting hydride [2]; (ii) the crystal volume criterion—the hydride stability increases with the increase in the original crystal volume [3]; (iii) the crystal structure—sometimes alloys of identical or nearly identical chemical compositions crystallize in a completely different lattice structure. The one, presenting interstitial sites more abundant in hydride-forming elements, forms more stable hydrides. For example, the pseudobinary U(AlxNi1−x)2 system crystallizes in two different structures, a Laves phase structure in the compositional ranges 0 ≤ x ≤ 0.2, 0.9 ≤ x ≤ 1, and a ZrNiAl-type (Fe2P-type) structure at 0.4 ≤ x ≤ 0.5. The Laves phase compounds do not absorb hydrogen up to pressures of 100 atm, while the hydrides UNiAlH2.5 and UNi1.2Al0.8H2.35 are formed at that pressure range. This difference of hydrogen absorption is attributed to the presence of uranium-richer tetrahedral sites (3U + 1(Ni,Al)) in U(AlxNi1−x)2, x = 0.4, 0.5, as compared to the (2U + 1(Ni,Al)) sites in the Laves-phase compounds, where uranium is the hydride-forming element [4,5]; and (ⅳ) we will consider in some detail the influence of the bulk elastic modulus on the critical temperature, Tc.
The metal to hydride phase transition in most hydrogen-absorbing intermetallic compounds is characterized by a critical temperature, Tc. A discontinuous transformation of the crystal lattice occurs below Tc, while above it the final hydride structure may be reached in a gradual, continuous way. This behavior imposes identical or very similar crystal structures of the original intermetallic compound and its corresponding hydride, besides the usual increase in the lattice volume upon hydrogenation. The metal to hydride phase transformation is driven by attractive hydrogen–hydrogen interaction, −u, which is composed of elastic and electronic parts [6].
u = uelas + uelec
The critical temperature, Tc, is related to the attractive H–H interaction:
Tc = u × r/4k
r is the number of interstitial sites per metal atom available for hydrogen absorption and k is Boltzmann’s constant (1.38 × 10−23 J K−1). The corresponding units of u are J.
The form of the elastic part, uelas, has been derived by several investigators, e.g., Refs. [6,7,8]:
u e l a s = 1 B B + 4 3 G B v H 2 V
B and G are the bulk and shear moduli in Pa (N/m2), respectively, v H is the excess volume occupied by one hydrogen atom in the metal matrix in m3, and V is the average volume per metal atom in m3. uelas is then obtained in J. uelas is usually presented in eV by multiplying the value in J by the conversion factor 1 J = 6.242 × 1018 eV. The volume, v H , of the hydrogen atom in the metal matrix and the term in the brackets usually do not change very much. Typical values of v H stay in the range of 2.5–3 Å3/H atom, e.g., Ref. [6]. The bulk modulus, B, is then a leading term in uelas. The physical reasoning for the increase in the long-range uelas interaction with B is that larger B induces larger strain fields in the crystal lattice around a hydrogen atom and provides better linkage between the hydrogen atoms. The attractive elastic interaction is considered to play a main role in the metal-to-hydride phase transition. Fen Li et al. have proposed a repulsive screened Coulomb interaction between the hydrogen atoms in TiCr2Hx [9]. It was recently demonstrated that the electronic interaction may be repulsive, negligible, or attractive by considering the corresponding critical temperatures of the ZrCr2-H2 (<300 K), ZrMn2-H2 (564 K), and Pd-H2 (565 K) systems in relation to their relevant elastic properties, Equations (1) and (2) [10].
This study aims at elucidating a puzzling behavior of the hydride stabilities in the Zr(MoxFe1−x)2 system, as presented at MH2018 and subsequently published [11]. Figure 1 [11] exhibits pressure-composition isotherms, demonstrating increased monotonic stability of the Zr(MoxFe1−x)2 hydrides for x = 0, 0.2, 0.5, but then this trend is reversed by a stability decrease for x = 1. Investigating these stability trends of the Zr(MoxFe1−x)2 hydrides, in view of the guidelines listed above, requires a knowledge of relevant data, namely, the heats of formation of the original intermetallic compounds, their crystal structures, lattice parameters, bulk and shear moduli, and the degree of expansion upon hydrogenation which determines v H . Most of these data appear in the literature. In order to complete the required set, elastic properties have been determined in this work. In addition, we hydrogenated ZrMo2 to the greatest possible extent under the available experimental conditions in our lab, attempting to settle some differing crystallographic results, associated with the volume of the hydrogen atom in ZrMo2 hydride(s).

2. Experimental Details

Zr(MoxFe1−x)2, x = 0, 0.5, 1, samples were prepared by arc melting high-purity (~99.9%) elements on a water-cooled copper hearth. The pellets were homogenized by melting them at least three times, turning them over after each melting. The as-cast alloys were then sealed in argon-filled quartz ampules and annealed for 72 h at 1373 K. The crystal structure of the prepared compounds was determined and verified by X-ray diffraction (XRD) in a PW 1050/70 diffractometer from Philips, utilizing Cu Kα radiation of 1.5418 Å. Parallel-faced intermetallic slices of about 2.5 mm thickness were cut using a diamond sawing disk and subsequently polished to achieve a parallelism within the range of 3 × 10−6 m. The resulting parallel-faced samples were then subjected to sound velocity measurements using the pulse echo (PE) method. The frequency of the PE piezoelectric transducers for both the longitudinal and transverse modes was 5 MHz. The sound wave velocities were determined by measuring the time needed for the acoustic waves to travel forth and back between the opposite parallel faces of the sample. The densities of the samples were analyzed using the measured structural parameters. The obtained results were then compared to the measured Archimedes densities (both wet and dry) in order to determine the porosity ratio of the samples.
ZrMo2 compound was hydrogenated at about 60 atm hydrogen pressure and room temperature in a home-built Sieverts system, equipped with 3.5, 60, and 100 atm pressure transducers. The amount of absorbed hydrogen in ZrMo2 was calculated by considering the non-ideality of the hydrogen gas. For example, the difference between the real and the ideal hydrogen pressures at 100 atm and room temperature is about 6%. The reactor chamber, with the hydrogenated sample in it, was then very slowly cooled from room to liquid nitrogen temperature in order to enhance the hydrogen absorption. The equilibrium hydrogenation pressure decreases in an exponential way vs. the decreased temperature. Thus, decreasing the sample temperature at approximately constant pressure, as most of the system is kept at ambient conditions, corresponds to significantly increasing the pressure at room temperature. Hydrogen absorption is expected to be practically none at very low temperatures, e.g., 78 K, because of extremely reduced kinetics. The very slow cooling is then intended to enable hydrogen absorption at intermediate temperatures, between 295 K and 78 K. Finally, a poisoning procedure was applied. (The purpose of the poisoning procedure is a contamination of the surface of the sample, for example by water vapor, without changing its bulk content. This contamination keeps the hydride intact and prevents its decomposition for long enough time in order to enable some experimental measurements at ambient conditions, e.g., XRD.) The hydrogen was evacuated from the reaction chamber, kept at liquid nitrogen temperature, and then the sample was exposed to the ambient atmosphere. The X-ray diffraction pattern of the air-exposed sample was immediately recorded without any additional precautions. The poisoning is considered successful provided two conditions are fulfilled: (i) the XRD pattern of the hydrogenated sample is significantly different from the pattern of the original compound; and (ii) the XRD pattern of the hydrogenated sample does not change or changes very slowly with time. These two conditions were satisfied in the present case of ZrMo2—see Section 3.1 below.

3. Results and Discussion

3.1. Structural and Elastic Properties of Zr(MoxFe1−x)2, x = 0, 0.5, 1

ZrFe2 and Zr(Mo0.5Fe0.5)2 exhibited single Laves phases in their XRD patterns, while 95% Laves phase was obtained for ZrMo2 (Figure 2). ZrFe2 and ZrMo2 crystallized in the cubic C15 phase with lattice constants 7.074 Å and 7.589 Å, respectively. These values are in fair agreement with the published data for ZrFe2 (e.g., 7.074 Å [12]) and ZrMo2 (e.g., 7.59 Å [13]). About 5% of Mo, with lattice constant a = 3.185 Å, was present along the ZrMo2 pattern (Figure 2). The cubic crystal parameter of pure Mo is 3.147 Å [14]. The increased a value indicates the dissolution of some Zr into the molybdenum matrix. Zr(Mo0.5Fe0.5)2 presents a C14 hexagonal phase with lattice constants a = 5.173 Å and c = 8.461 Å, comparable to a = 5.172 Å and c = 8.463 Å [15] or a = 5.130 Å and c = 8.440 Å [16]. The longitudinal, vL, and transverse, vT, ultrasonic velocities, measured by the pulse-echo method, are presented in Table 1. They were used to calculate two elastic constants according to the relations C11 = vL2ρ and C44 = vT2ρ, assumed to hold for polycrystalline isotropic samples (ρ is the sample density). Bulk, B, and shear, G, moduli were derived from the relations B = C11−4 C44/3 and G = C44 (Table 1). Minor corrections of B and G were made in view of the small porosities of the samples [17,18]. The shear modulus changes moderately with the composition, x, along the Zr(MoxFe1−x)2 series, while the bulk modulus increases significantly and monotonically as a function of x. The variations of B and G are exhibited in Figure 3 and Table 1.
ZrMo2 absorbed about 2.1 H atoms per formula unit at 60 atm and room temperature, and about 3.5 H atoms per f.u. upon slow cooling to 78 K at 50 atm pressure. It should be born in mind that the hydrogenations of ZrMo2 were carried out in order to estimate the corresponding v H values. PCT (pressure-composition-temperature) isotherms of ZrMo2-H2 may be found elsewhere, e.g., Ref. [11]. The XRD pattern, recorded immediately after removing the sample from the reactor, revealed two hydride phases, cubic Laves structure with lattice constant a = 7.826 Å (26%) and tetragonal structure, space group 88, with cell parameters a = 5.603 Å, c = 8.081 Å (68%) (Figure 4). These lattice constants slowly decreased during one week of exposure to the ambient atmosphere. The crystal parameter of Mo with dissolved Zr almost did not change (a = 3.187 Å) upon hydrogenation, indicating no hydrogen absorption in it. By imposing somewhat arbitrarily identical v H for the two existing phases in ZrMo2H3.5, we obtain for the composition of these two phases ZrMo2H2.39 (MgCu2-type, a = 7.826 Å) and ZrMo2H3.98 (tetragonal, SG88, a = 5.603 Å, c = 8.081 Å). Their common v H is 2.2 Å3/(H atom). Additional v H estimations, performed in a similar way for ZrMo2H1.6 and ZrMo2H2.8, yielded values between 2.1 Å3/(H atom) and 2.2 Å3/(H atom).

3.2. Resolving the Peculiarity in the Stability Trends of the Zr(MoxFe1−x)2, x = 0, 0.1, 0.5, 1, Hydrides

As cited in the Introduction, the formation pressures of the Zr(MoxFe1−x)2 hydrides decrease between x = 0 and x = 0.5, indicating increasing stability in this compositional range. Then, this trend is unexpectedly reversed and ZrMo2 hydride is formed at higher pressures than ZrMoFe and Zr(Mo0.1Fe0.9)2 hydrides. We consider next the possible factors influencing the stability of these hydrides.

3.2.1. The Substitution of Fe by Mo

Both Fe and Mo form hydrides at very high hydrogen pressures. The equilibrium formation pressure of MoH1.1 and FeH are about 4.3 GPa at 600 °C and 3.5 GPa at 300 °C (decomposition pressure 2.2 GPa), respectively [21,22]. It is thus not expected that the hydrogen affinity properties of Mo, as compared to Fe, play a dominant role in the variation of the hydride stabilities in Zr(MoxFe1−x)2.

3.2.2. Trends of the Heats of Formation and the Crystal Volumes in the Zr(MoxFe1−x)2 System

Table 2 presents the crystal volumes per formula unit (f.u.), V, and theoretically derived heats of formation, ∆Hf, of the original, non-hydrogenated Zr(MoxFe1−x)2 intermetallics, x = 0, 0.1, 0.5, 1. The theoretically determined ∆Hf of 0.281 eV/(metal atom) for ZrFe2 is in good agreement with a corresponding experimental value of 0.26 eV/(metal atom), e.g., Ref. [23]. ∆Hf values of Zr(MoxFe1−x)2, x = 0.1, 0.5 are obtained by interpolation.
The heats of formation of the intermetallic compounds decrease with x from 0.281 eV/(metal atom) for ZrFe2 to 0.139 eV/(metal atom) for ZrMo2, while the corresponding crystal volumes increase from 44.058 Å3/(f.u.) to 54.634 Å3/(f.u.) (see Table 2). Thus, according to both the rule of reverse stability and the crystal volume criterion (see (i) and (ii) in the Introduction), the hydride stabilities should increase with x. These trends comply with the behavior for 0 ≤ x ≤ 0.5, but disagree with the decreased stability of ZrMo2 hydride.

3.2.3. Elastic Properties and H–H Elastic Interaction in Zr(MoxFe1−x)2-H2

As already noted, the H–H interaction influences the critical temperature, Tc, according to Equation (1). It also contributes to the heat of hydrogen solution or hydride formation [6]. We find two indications for the reduction in the attractive, elastically mediated H–H interaction.
(a) From inspection of Figure 1, the very wide solubility region of ZrMo2-H2 indicates a decrease in Tc with respect to the rest of the hydrogenated Zr(MoxFe1−x)2 compounds, x = 0, 0.1, 0.5. This occurs in spite of the monotonic increase in the bulk modulus, B, from 148.2 GPa for ZrFe2 to 200.4 GPa for ZrMo2, expected in turn to increase Tc.
(b) The increased hydrogenation pressures in ZrMo2-H2 with regard to Zr(MoxFe1−x)2-H2, x = 0.1, 0.5, indicate a corresponding destabilization of ZrMo2Hx. Indeed, the absolute enthalpy of ZrMo2 hydride formation, 22 kJ/(mole H2), is significantly lower than those of Zr(MoxFe1−x)2 hydrides, 29.5 kJ/(mole H2) for x = 0.5, and 26 kJ/(mole H2) for x = 0.1. It even approaches the 21.3 kJ/(mole H2) for ZrFe2 hydride [11]. This destabilization behavior suggests that the attractive (negative) contribution of the H–H elastic interaction to the enthalpy of hydride formation or to the enthalpy of hydrogen solubility in ZrMo2 weakens with the increase in hydrogen content (see the similar case of TaV2-H2 below).
As the variation in B of the Zr(MoxFe1−x)2 compounds cannot directly account for the observed trend in the critical temperatures of the metal-to-hydride phase transitions and the heats of hydride formation at x > 0.5, we inspect another term that appears in a squared form in Equation (2), namely, v H . A priori, v H is not supposed to significantly affect uelas according to Equation (2), as v H usually acquires values in quite a narrow range, e.g., Ref. [6]. We find hereafter surprising experimentally derived v H values. Lushnikov et al. report that a Laves ZrMo2 compound with lattice constant a = 7.591 Å undergoes a tetragonal distortion upon deuteration to ZrMo2D4 with lattice constants a = 5.496 Å and c = 7.986 Å [27]. The authors note that such a tetragonal distortion is a known phenomenon in hydrogenated Laves phase compounds [29]. It is also worth noting that the composition of the interstitial sites does not change upon the tetragonal distortion. The deuterium atoms occupy (2Zr + 2Mo) tetrahedral interstitial sites [27], which are the preferable (2A + 2B) sites for hydrogen occupation in AB2 Laves phase compounds [30]. A straightforward evaluation of the volume of the hydrogen atom in ZrMo2D4 (4 f.u./unit cell), using also the above data for ZrMo2 (8 f.u./unit cell), yields v H = 1.4 Å3/(H atom). This is the lowest reported v H value in intermetallic and even in binary hydrides [6] (pp. 105–109). Supporting evidence for this hydrogen atomic volume may be found in the reported cell parameter a = 7.698 Å of ZrMo2H1.4 [16]. Utilizing a = 7.598 Å for ZrMo2 by the same authors [31], v H = 1.6 Å3/(H atom) is evaluated. It should be noted that a cell parameter a = 7.548 Å is reported for ZrMo2 in [16], which is very different from other reported values [13,27], as well as from the value, reported later, by the same authors [31]. Nonetheless, v H of 2.33 Å3/(H atom) is obtained for ZrMo2H1.4 by employing a = 7.548 Å for ZrMo2, although the stated v H in [16] is 2.6 Å3/(H atom). Additional v H values, reported for various hydrogen compositions in ZrMo2, span between 3.2 and 4.5 Å3/(H atom) [31]. We recommend adopting v H = 2.2 Å3/(H atom) for the ZrMo2 hydrides following the experimental estimations of this work. It is worthwhile noting that we found larger lattice constants of the tetragonal hydride phase (see Section 3.1) than those cited above. The latter lattice constants were obtained in the course of a seemingly very comprehensive neutron diffraction experiment, under much more extreme conditions of 2500 atm hydrogen pressure and a temperature of 78 K [27]. Table 2 presents the relevant information for the Zr(MoxFe1−x)2 compounds and their hydrides. The small values of v H affect uelas very significantly. Table 2 lists the calculated values of uelas according to Equation (2), utilizing the measured bulk and shear moduli (Table 1), and the V and v H volumes (Table 2). The ZrMo2-H2 system clearly exhibits the smallest v H and uelas. This may explain the decreased Tc in this system according to Equation (1). The hydride stability would also be decreased as uelas provides a smaller negative contribution to the heat of hydride formation of ZrMo2 with regard to ZrFe2 and ZrMoFe. A most probable reason for the decrease in v H , and hence in uelas, in the Zr(MoxFe1−x)2 system is the large bulk modulus of ZrMo2. We suggest that the large bulk modulus of ZrMo2 confines the expansion of the metal lattice upon hydrogenation by restricting the H atoms in smaller interstitial volumes, v H . The variation of v H as a function of B, plotted in Figure 5, supports this suggestion. The depressed expansion of the crystal lattice mitigates the strain fields around a hydrogen atom and thus reduces the elastically mediated interaction between the hydrogen atoms.
Elastic moduli rarely play such a significant role in the hydrogenation properties of intermetallic compounds. We note that a very similar behavior to that of ZrMo2-H2 is demonstrated by the C15 compound TaV2. The experimentally derived bulk modulus of C15 TaV2 varies between 200 GPa at low temperatures and 194 GPa at room temperature [32]. These B values of TaV2 are very close to the 200.4 GPa bulk modulus of ZrMo2, obtained in this work (see Table 1 and Figure 5). The volume of the hydrogen atom in hydrogenated TaV2 is 2.09 Å3/(H atom) [33] in the low range of v H values, and in close proximity to v H of 2.2 Å3/(H atom), estimated here for the hydrogenated ZrMo2 (see Table 2 and Figure 5). Accordingly, TaV2-H2 exhibits a broad solid hydrogen solubility region without a discontinuous metal-to-hydride phase transition even at temperatures as low as 195 K and pressures of 1000 atm [33]. In addition, the heat of hydrogen solution decreases with the increase in H content, i.e., becomes less exothermic. This may be regarded as surmounting the attractive elastic H–H interaction by some repulsive, probably electronic H–H interaction. It is worthwhile noting that the heats of hydrogen solution in metal–hydrogen systems usually become more exothermic with the increase in H content, e.g., Refs. [6,34]. We suggest that, in resemblance to ZrMo2-H2, the large bulk modulus of TaV2 confines the hydrogen atoms into small interstitial volumes and consequently reduces the H–H elastic interaction, as indicated by the broad solubility region and the lack of metal to hydride phase transition in TaV2-H2. Another example of a significant role of elastic moduli is the shear stiffening in Zr(AlxM1−x)2, M=Fe, Co [19,35]. It prevents hydrogen absorption in ZrAl2 even at extremely high H2 pressures of 40 GPa [36], although Al substitution drastically stabilizes the hydrogen absorption of ZrFe2 and ZrCo2 for small x values [37,38].

4. Summary and Conclusions

The elastic moduli of Zr(MoxFe1−x)2, x = 0, 0.5, 1, as well as hydrogen absorption in ZrMo2, were measured in an attempt to shed light on the unusual trend of hydride stabilities in this system. It seems that the bulk modulus plays a dual role in that context. Initially, B participates in generating attractive elastic interaction between the hydrogen atoms. Then, the significant increase in B in ZrMo2 confines the expansion of the metal matrix and restrains the hydrogen atoms in smaller interstitial volumes. Consequently, the attractive H–H interaction is reduced, demonstrated by a sharp decrease in the critical temperature for metal-to-hydride phase transition and the hydride stability in the ZrMo2-H2 system. We also suggest a similar explanation for the lack of a discontinuous metal-to-hydride phase transition in the Laves phase TaV2-H2 system. It would be of interest to find a criterion that determines at which point B converts from a stabilizing to destabilizing agent during hydrogen absorption in intermetallic systems.

Author Contributions

Conceptualization, I.J.; methodology, D.B. and M.B.; validation, D.B. and M.B.; formal analysis, D.B.; investigation, D.B. and M.B.; writing—original draft preparation, I.J.; writing—review and editing, D.B., R.Z.S. and I.J.; supervision, I.J. and R.Z.S.; funding acquisition, I.J. and R.Z.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partly funded by the Israel Science Foundation grant number 745/15.

Acknowledgments

This research was partly supported by the Israel Science Foundation (grant No. 745/15). One of the authors (IJ) is grateful to S. Mitrokhin for granting access to his presentation at MH2018, as well as for turning attention to the unusual stability trends of the Zr(MoxFe1−x)2 hydrides.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Reilly, J.J. Chemistry of intermetallic hydrides, BNL report 46778, presented at the Symposium for Hydrogen Storage Materials, Battery and Electrochemistry. In Proceedings of the 180th Meeting of the Electrochemical Society, Phoenix, AZ, USA, 13–18 October 1991. [Google Scholar]
  2. van Mal, H.H.; Buschow, K.H.J.; Miedema, A.R. Hydrogen absorption in LaNi5 and related compounds: Experimental observations and their explanation. J. Less-Common Met. 1974, 35, 65–76. [Google Scholar] [CrossRef]
  3. Lartigue, C. Etude Structurale et Thermodynamique du Systeme LaNi5−XMnX—Hydrogene. Ph.D. Thesis, Universite Paris VI, Paris, France, 1984. [Google Scholar]
  4. Jacob, I.; Hadari, Z.; Reilly, J.J. Hydrogen absorption in ANiAl (A = Zr, Y, U). J. Less-Common Met. 1984, 103, 123–127. [Google Scholar] [CrossRef]
  5. Biderman, S.; Jacob, I.; Mintz, M.H.; Hadari, Z. Analysis of the hydrogen absorption in the U(AlxNi1−x)2 system. Trans. Nucl. Soc. Isr. 1982, 10, 129–132. [Google Scholar]
  6. Fukai, Y. The Metal-Hydrogen System: Basic Bulk Properties, 2nd ed.; Springer: Berlin/Heidelberg, Germany, 2005. [Google Scholar]
  7. Alefeld, G. Phase transitions of hydrogen in metals due to elastic interaction. Ber. Bunsenges. Phys. Chem. 1972, 76, 746–755. [Google Scholar]
  8. Wagner, H. Elastic Interaction and phase transitionin in coherent metal-hydrogen systems. In Hydrogen in Metals I; Alefeld, G., Völkl, J., Eds.; Springer: Berlin/Heidelberg, Germany, 1978; pp. 5–51. [Google Scholar]
  9. Li, F.; Zhao, J.; Tian, D.; Zhang, H.; Ke, X.; Johansson, B. Hydrogen storage behavior in C15 Laves compound TiCr2 by first principles. J. Appl. Phys. 2009, 105, 043707. [Google Scholar] [CrossRef]
  10. Babai, D.; Bereznitsky, M.; Shneck, R.Z.; Jacob, I. The effect of Pd on hydride formation in Zr(PdxM1−x)2 intermetallics where M is a 3d element. J. Alloys Compd. 2021, 889, 161503. [Google Scholar] [CrossRef]
  11. Mitrokhin, S.; Verbetsky, V. Peculiarities of Hydrogen Interaction with Alloys of ZrFe2-ZrMo2 System. Int. J. Hydrogen Energy 2019, 44, 29166–29169. [Google Scholar] [CrossRef]
  12. Muraoka, Y.; Shiga, M.; Nakamura, Y. Magnetic properties and Mössbauer effect of A(Fe1−xBx)2 (A = Y or Zr, B = Al or Ni) Laves phase intermetallic compounds. Phys. Status Solidi 1977, 42, 369–374. [Google Scholar] [CrossRef]
  13. Domagala, R.F.; McPherson, D.J.; Hansen, M. Systems Zirconium-Molybdenum and Zirconium-Wolfram. JOM 1953, 5, 73–79. [Google Scholar] [CrossRef]
  14. Straumanis, M.E.; Shodhan, R.P. Lattice Constants, Thermal Expansion Coefficients and Densities of Molybdenum and the Solubility of Sulphur, Selenium and Tellurium in it at 1100 °C. Z. Metallkd. 1968, 59, 492–495. [Google Scholar] [CrossRef]
  15. Yartys, V.A.; Burnasheva, V.V.; Fadeeva, N.V.; Solov’ev, S.P.; Semenenko, K.N. The crystal structure of the deuteride ZrMoFeD2.6. Sov. Phys. Crystallogr. (Transl. Krist.) 1982, 27, 540–543. [Google Scholar]
  16. Semenenko, K.N.; Verbetskii, V.N.; Mitrokhin, S.V.; Burnasheva, V.V. Investigation of the interaction with hydrogen of Zirconium intermetallic compounds crystallised in Laves phase structure types. Russ. J. Inorg. Chem. 1980, 25, 961–964, Translated from Zhurnal Neorgamcheskoi Khimil 1980, 25, 1731–1736. [Google Scholar]
  17. Mackenzie, J.K. The elastic constants of a solid containing spherical holes. Proc. Phys. Soc. B 1950, 63, 2–11. [Google Scholar] [CrossRef]
  18. Masi, L.; Borchi, E.; de Gennaro, S. Porosity behavior of ultrasonic velocities in polycrystalline Y-B-C-O. J. Phys. D Appl. Phys. 1996, 29, 2015–2019. [Google Scholar] [CrossRef]
  19. Willis, F.; Leisure, R.G.; Jacob, I. Elastic moduli of the Laves-phase pseudobinary compounds Zr(AlxFe1−x)2 as determined by ultrasonic measurements. Phys. Rev. B 1994, 50, 13792–13794. [Google Scholar] [CrossRef]
  20. Turkdal, N.; Deligoz, E.; Ozisik, H.; Ozisik, H.B. First-principles studies of the structural, elastic, and lattice dynamical properties of ZrMo2 and HfMo2. Phase Transit. 2017, 90, 598–609. [Google Scholar] [CrossRef]
  21. Abramov, S.N.; Antonov, V.E.; Bulychev, B.M.; Fedotov, V.K.; Kulakov, V.I.; Matveev, D.V.; Sholin, I.A.; Tkacz, M. T-P phase diagrams of the Mo-H system revisited. J. Alloys Compd. 2016, 672, 623–629. [Google Scholar] [CrossRef]
  22. Tkacz, M. Thermodynamic properties of iron hydride. J. Alloys Compd. 2002, 330–332, 25–28. [Google Scholar] [CrossRef]
  23. Klein, R.; Jacob, I.; O’Hare, P.A.G.; Goldberg, R.N. Solution-calorimetric determination of the standard molar enthalpies of formation of the pseudobinary compounds Zr(AlxFe1-x)2 at the temperature 298.15 K. J. Chem. Thermodyn. 1994, 26, 599–608. [Google Scholar] [CrossRef]
  24. Zotov, T.; Movlaev, E.; Mitrokhin, S.; Verbetsky, V. Interaction in (Ti,Sc)Fe2-H2 (Zr,Sc)Fe2-H2 systems. J. Alloys Compd. 2008, 459, 220–224. [Google Scholar] [CrossRef]
  25. Zotov, T.A.; Sivov, R.B.; Movlaev, E.A.; Mitrokhin, S.V.; Verbetsky, V.N. IMC hydrides with high hydrogen dissociation pressure. J. Alloys Compd. 2011, 509, S839–S843. [Google Scholar] [CrossRef]
  26. The Materials Project. Available online: https://materialsproject.org/materials/mp-1718/ (accessed on 9 April 2023).
  27. Lushnikov, S.A.; Movlaev, E.A.; Verbetsky, V.N.; Somenlov, V.A.; Agafonov, S.S. Interaction of ZrMo2 with hydrogen at high pressure. Int. J. Hydrogen Energy 2017, 42, 29166–29169. [Google Scholar] [CrossRef]
  28. The Materials Project. Available online: https://materialsproject.org/materials/mp-2049/ (accessed on 9 April 2023).
  29. Irodova, A.V.; Glazkov, V.P.; Somenkov, V.A.; Shilstein, S.S. Hydrogen ordering in the cubic Laves phase HfV2. J. Less-Common Met. 1981, 77, 89–98. [Google Scholar] [CrossRef]
  30. Jacob, I.; Bloch, J.M.; Shaltiel, D.; Davidov, D. On the occupation of interstitial sites by hydrogen atoms in intermetallic hydrides: A quantitative model. Solid State Commun. 1980, 35, 155. [Google Scholar] [CrossRef]
  31. Semenenko, K.N.; Verbetskii, V.N.; Pilchenko, V.A. Interaction of ZrMo2 with hydrogen at low temperatures. Mosc. Univ. Chem. Bull. (Transl. Vestn. Mosk. Univ. Ser. 2) 1986, 41, 131–133. [Google Scholar]
  32. Foster, K.; Hightower, J.E.; Leisure, R.G.; Skripov, A.V. Elastic moduli of the C15 Laves-phase materials TaV2, TaV2H(D)x and ZrCr2. Philos. Mag. B 2000, 80, 1667–1679. [Google Scholar] [CrossRef]
  33. Lynch, J.F. The solution of hydrogen in TaV2. J. Phys. Chem. Solids 1981, 42, 411–419. [Google Scholar] [CrossRef]
  34. Bereznitsky, M.; Mogilyanski, D.; Jacob, I. Destabilizing effect of Al substitution on hydrogen absorption in Zr(AlxV1−x)2. J. Alloys Compd. 2012, 542, 213–217. [Google Scholar] [CrossRef]
  35. Jacob, I.; Bereznitsky, M.; Yeheskel, O.; Leisure, R.G. Role of shear stiffening in reducing hydrogenation in intermetallic compounds. Appl. Phys. Lett. 2006, 89, 201909. [Google Scholar] [CrossRef]
  36. Machida, A. Unpublished Results.
  37. Bereznitsky, M.; Jacob, I.; Bloch, J.; Mintz, M.H. Thermodynamic and structural aspects of hydrogen absorption in the Zr(AlxCo1−x)2 system. J. Alloys Compd. 2002, 346, 217. [Google Scholar] [CrossRef]
  38. Bereznitsky, M.; Jacob, I.; Bloch, J.; Mintz, M.H. Thermodynamic and structural aspects of hydrogen absorption in the Zr(AlxFe1−x)2 system. J. Alloys Compd. 2003, 351, 180. [Google Scholar] [CrossRef]
Figure 1. Pressure-composition isotherms of Zr(MoxFe1−x)2-H2 systems, x = 0, 0.1. 0.5, 1 at the indicated temperatures. (Reprinted with permission from Ref. [11] under license 5531441401625 by Elsevier).
Figure 1. Pressure-composition isotherms of Zr(MoxFe1−x)2-H2 systems, x = 0, 0.1. 0.5, 1 at the indicated temperatures. (Reprinted with permission from Ref. [11] under license 5531441401625 by Elsevier).
Inorganics 11 00228 g001
Figure 2. XRD patterns of Zr(MoxFe1−x)2, x = 0, 0.5, 1. Single-phased cubic C15 and hexagonal C14 Laves structures were found for ZrFe2 (blue pattern) and ZrMoFe (red pattern), respectively. For ZrMo2, 95% C15 structure was found with 5% Mo (▼) (green pattern).
Figure 2. XRD patterns of Zr(MoxFe1−x)2, x = 0, 0.5, 1. Single-phased cubic C15 and hexagonal C14 Laves structures were found for ZrFe2 (blue pattern) and ZrMoFe (red pattern), respectively. For ZrMo2, 95% C15 structure was found with 5% Mo (▼) (green pattern).
Inorganics 11 00228 g002
Figure 3. Bulk, B, and shear, G, moduli in Zr(MoxFe1−x)2 intermetallic compounds.
Figure 3. Bulk, B, and shear, G, moduli in Zr(MoxFe1−x)2 intermetallic compounds.
Inorganics 11 00228 g003
Figure 4. XRD pattern of ZrMo2H3.5 (upper plot) reveals the existence of two hydride phases (see also text). The pattern of ZrMo2 before hydrogenation is shown in the lower plot for comparison.
Figure 4. XRD pattern of ZrMo2H3.5 (upper plot) reveals the existence of two hydride phases (see also text). The pattern of ZrMo2 before hydrogenation is shown in the lower plot for comparison.
Inorganics 11 00228 g004
Figure 5. Variation of the hydrogen atomic volume, v H , in the hydrogenated Zr(MoxFe1−x)2, x = 0, 0.5, 1, compounds as a function of the bulk moduli, B, of the corresponding original intermetallics.
Figure 5. Variation of the hydrogen atomic volume, v H , in the hydrogenated Zr(MoxFe1−x)2, x = 0, 0.5, 1, compounds as a function of the bulk moduli, B, of the corresponding original intermetallics.
Inorganics 11 00228 g005
Table 1. Lattice constants, theoretical densities, porosities, longitudinal, vL, and transverse, vT, ultrasonic velocities, shear, G, and bulk, B, elastic moduli in the Laves phase Zr(MoxFe1−x)2 system. Experimental B and G values of Zr(Al0.04Fe0.96)2 and theoretical B and G values of ZrMo2 are also presented. The numbers in the parentheses present the estimated errors of the last significant digits.
Table 1. Lattice constants, theoretical densities, porosities, longitudinal, vL, and transverse, vT, ultrasonic velocities, shear, G, and bulk, B, elastic moduli in the Laves phase Zr(MoxFe1−x)2 system. Experimental B and G values of Zr(Al0.04Fe0.96)2 and theoretical B and G values of ZrMo2 are also presented. The numbers in the parentheses present the estimated errors of the last significant digits.
xa [Å]c [Å]Theoretical Density [kg/m3]Porosity
[%]
vL [m/s]vT [m/s]G [GPa]B [GPa]
0 *
[19]
75.8148
07.074 76140.95674 (9)3142 (4)75.6 (3)148.2 (9)
0.55.1738.46182970.85337 (2)2608 (1)56.3 (2)164.2 (4)
17.589 86050.75733 (16)2762 (14)66.1 (8)200.4 (2)
1
[20]
7.594 57.3196.5
* Zr(Al0.04Fe0.96)2.
Table 2. Unit cell volumes, V, heats of intermetallic formation, ∆Hf, hydrogen contents, hydrogen atomic volumes, v H , and elastic interactions, uelas, for Zr(MoxFe1−x)2 and the corresponding hydrides.
Table 2. Unit cell volumes, V, heats of intermetallic formation, ∆Hf, hydrogen contents, hydrogen atomic volumes, v H , and elastic interactions, uelas, for Zr(MoxFe1−x)2 and the corresponding hydrides.
xV [Å3/f.u.]∆Hf [eV/Atom]H Content
[H Atoms/f.u.]
v H
[Å3/H Atom]
uelas [eV]
0 [24,25]44.0580.281 [26]3.543.020.23
0.1 [25]44.8050.2673.542.99
0.548.089 [16]0.212.52.900.17
49.013 [15]2.62.740.15
154.677 [27]0.139 [28]41.410.042
54.829 [16]1.41.570.052
54.634 this work3.52.20.102
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jacob, I.; Babai, D.; Bereznitsky, M.; Shneck, R.Z. The Role of Bulk Stiffening in Reducing the Critical Temperature of the Metal-to-Hydride Phase Transition and the Hydride Stability: The Case of Zr(MoxFe1−x)2-H2. Inorganics 2023, 11, 228. https://doi.org/10.3390/inorganics11060228

AMA Style

Jacob I, Babai D, Bereznitsky M, Shneck RZ. The Role of Bulk Stiffening in Reducing the Critical Temperature of the Metal-to-Hydride Phase Transition and the Hydride Stability: The Case of Zr(MoxFe1−x)2-H2. Inorganics. 2023; 11(6):228. https://doi.org/10.3390/inorganics11060228

Chicago/Turabian Style

Jacob, Isaac, Dotan Babai, Matvey Bereznitsky, and Roni Z. Shneck. 2023. "The Role of Bulk Stiffening in Reducing the Critical Temperature of the Metal-to-Hydride Phase Transition and the Hydride Stability: The Case of Zr(MoxFe1−x)2-H2" Inorganics 11, no. 6: 228. https://doi.org/10.3390/inorganics11060228

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