Fact-checked by Grok 2 weeks ago

Hydride

A hydride is a binary formed by the bonding of to another , typically a metal, non-metal, or , in which assumes a formal of −1 as the hydride anion (H⁻). These compounds encompass a diverse range of structures and properties, arising from the unique ability of to form ionic, covalent, or metallic bonds depending on the electronegativity difference with the paired element. Hydrides play critical roles in various applications, including as strong reducing agents in , materials for clean energy technologies, and components in batteries and fuel cells. Hydrides are classified into three primary categories based on their bonding characteristics and the nature of the elements involved: ionic, covalent (or molecular), and metallic (or ). Ionic hydrides form between and highly electropositive s-block metals (Groups 1 and 2), such as (LiH) and (NaH), resulting in crystalline salts with high melting points and reactivity toward to liberate gas (H₂). These are typically colorless, brittle solids used industrially as desiccants or bases. Covalent hydrides arise from 's combination with p-block non-metals or less electropositive elements, like (NH₃), (CH₄), and (PH₃), forming discrete molecules that are often gases or low-boiling liquids at and exhibit varying and hydrogen bonding. Metallic hydrides, also known as hydrides, occur with d- and f-block metals, such as (TiH₂) or palladium hydride (PdHₓ), where atoms occupy lattice voids in the metal structure, leading to materials with high absorption capacity and reversible phase s suitable for . This classification highlights the versatility of hydrides, with ongoing research focusing on complex and polymeric variants for advanced technological uses.

Fundamentals

Definition and Characteristics

Hydrides are binary compounds formed by the chemical combination of with other elements, typically electropositive metals such as or alkaline earth metals, or with nonmetals through covalent bonding; this definition excludes protonated species such as the hydronium ion (H₃O⁺). These compounds represent a broad class in , where acts as a key component bonded to elements across the periodic table, often exhibiting hydrogen in formal oxidation states of -1, 0, or +1 depending on the bonding nature. Hydrides display significant diversity in their physical states and properties, existing as gases like (CH₄) at standard conditions, liquids such as (NH₃), or solids including (NaH). This variability arises partly from hydrogen's of 2.20 on the Pauling scale, which positions it between metals and nonmetals, leading to polar bonds in many hydrides and influencing their reactivity and phase behavior. In nature, hydrides are uncommon as discrete geological compounds; for instance, (H₂O) is classified as an rather than a hydride, and rare examples like silanes occur only in trace amounts under specific volcanic or serpentinization conditions, with the first confirmed natural metal hydride, vanadium hydride (VH₂), discovered in 2019 within pyroclastic ejecta. Structurally, hydrides range from discrete molecular units, such as borane (BH₃), to extended ionic lattices like lithium hydride (LiH), reflecting the bonding interactions between hydrogen and the counterpart element. Some hydrides, particularly those involving transition metals, exhibit non-stoichiometric compositions, where the hydrogen-to-metal ratio varies due to interstitial incorporation rather than fixed stoichiometry. The term "hydride" was coined in the mid-19th century, with its first documented use around 1869, building on early 19th-century observations of hydrogen reactions with metals; for example, the preparation of alkali metal hydrides, such as lithium hydride in 1904 by direct reaction of the metal with hydrogen gas, marked key advancements in recognizing these compounds.

The Hydride Ion

The , denoted as H⁻, consists of a proton orbited by two electrons occupying the 1s orbital, resulting in a closed-shell electronic configuration isoelectronic with the neutral . This structure imparts high stability to the in isolated or lattice environments, with an of approximately 1.40 in hosts. As a strong base, H⁻ possesses a of electrons available for donation, enabling it to coordinate with Lewis acids. Its conjugate acid, dihydrogen (H₂), exhibits high basicity, with a pKa value ranging from 35 to 50 depending on the , underscoring the ion's tendency to abstract protons vigorously. The stability of H⁻ is limited in protic solutions, where it rapidly protonates to form H₂ due to its strong basicity, rendering free H⁻ ions rare outside of solid-state matrices. In contrast, the ion is thermodynamically stable within the crystal lattices of ionic hydrides, such as those of alkali metals, where electrostatic interactions with cations prevent decomposition. The of atomic hydrogen is 0.754 eV, indicating that detachment of the added electron from H⁻ to form neutral H is endothermic by this amount, contributing to the ion's relative persistence in aprotic or low-proton environments. For instance, solvated hydride species can exist transiently in liquid , though they react with the solvent to generate ions and H₂. Reactivity of H⁻ is dominated by its role as a potent and , primarily through to yield H₂, as in the : \text{H}^- + \text{H}^+ \rightarrow \text{H}_2 \quad \Delta H = -1676 \, \text{kJ/mol} This value derives from thermochemical data, including the H–H (436 /mol), the ionization potential of H (1312 /mol), and the of H (72.8 /mol). As a reductant, H⁻ transfers its hydride equivalent to oxidants, facilitating reductions in synthetic chemistry, while its nucleophilic character allows attack on electrophilic centers. Quantum mechanically, the wavefunction of H⁻ approximates that of He, described by a 1s² orbital with both electrons in a spherically symmetric , yielding a probability concentrated near the . Compared to ions, H⁻ shares a similar size with F⁻ ( ~1.33 Å) but displays greater polarizability and basicity due to its lower nuclear charge and higher near the hydrogen relative to or .

Bonding in Hydrides

Ionic Bonding

In ionic hydrides, bonding arises from the complete transfer of a from an electropositive metal atom to a , resulting in the formation of positively charged metal cations (M⁺) and hydride anions (H⁻). These oppositely charged ions are then assembled into a crystalline stabilized by electrostatic attractions. This mechanism is prevalent in hydrides of s-block metals, where the low of the metal facilitates electron donation to , which achieves a stable closed-shell configuration as H⁻. For instance, in (LiH), the interatomic Li–H distance within the rock-salt is approximately 2.04 Å, reflecting the ionic radii of Li⁺ (0.76 Å) and H⁻ (1.40 Å). The energetics of this bond formation are quantified using the Born-Haber cycle, which decomposes the overall of formation into stepwise processes: of the metal, of H₂, of the metal atom, of hydrogen, and release. For LiH, the (ΔH_f°) is -90.5 kJ/mol, with the cycle balancing endothermic steps (e.g., metal of 520 kJ/mol for Li and H–H bond of 436 kJ/mol) against the highly exothermic . This cycle underscores the dominance of in stabilizing the compound, as its magnitude often exceeds 900 kJ/mol for these small ions. Lattice energy (U), the key stabilizing factor, is calculated via the Born equation: U = -\frac{N_A M q_1 q_2}{4\pi \epsilon_0 r} where N_A is Avogadro's constant, M is the structure-dependent Madelung constant (1.748 for the rock-salt geometry common in alkali metal hydrides), q_1 and q_2 are the ion charges (±e), ε_0 is the vacuum permittivity, and r is the interionic distance. The unusually high U values (e.g., 911 kJ/mol for LiH) stem from the compact size of H⁻, which minimizes r and amplifies the Coulombic attraction compared to larger anions like halides. Theoretical models describe as an ideal electrostatic interaction assuming full charge separation, but deviations occur with increasing of the metal, introducing partial covalent character due to orbital overlap. Pauling's scale predicts predominant ionic character when the difference (Δχ) exceeds ~1.9; however, for lighter s-block hydrides like those of or (Δχ ≈ 1.2–1.0), the ionic model holds well owing to the high and minimal polarization of H⁻, whereas heavier analogs (e.g., CsH) show subtle covalency. Ionic hydrides of s-block metals are generally unstable in aqueous media, undergoing rapid to yield metal hydroxides and H₂ gas, as the H⁻ abstracts a proton from .

Covalent Bonding

Covalent bonding in hydrides involves the sharing of electron pairs between hydrogen and other atoms, forming discrete molecular structures typical of covalent hydrides. This sharing contrasts with the complete electron transfer seen in ionic hydrides involving the hydride ion (H⁻). In most cases, these bonds are two-center two-electron (2c-2e) interactions, where a pair of electrons is localized between two atoms, as exemplified by the H–H bond in the dihydrogen molecule (H₂), which has a bond dissociation energy of 436 kJ/mol. The polarity of covalent hydride bonds arises from differences in between and its bonding partner, leading to partial charges on the atoms. For instance, in (HCl), the electronegativity difference of approximately 0.9 results in a polar with bearing a partial positive charge (δ⁺ H) and a partial negative charge (δ⁻ Cl). Hybridization of atomic orbitals plays a key role in determining the of covalent hydrides. In (CH₄), the carbon atom undergoes sp³ hybridization, forming four equivalent σ with atoms arranged in a tetrahedral with bond angles of 109.5°. In unsaturated hydrides like (C₂H₂), the carbon atoms are sp hybridized, resulting in a linear structure with a carbon-carbon consisting of one σ and two π bonds formed by the overlap of unhybridized p orbitals. Certain covalent hydrides, particularly those of elements, exhibit multi-center bonding due to . In (B₂H₆), the structure features two three-center two-electron (3c-2e) bonds involving bridged hydrogen atoms between centers, allowing the molecule to achieve stability despite having only 12 valence electrons for 7 bonds. This is characteristic of hydrides like , where the central atom lacks sufficient electrons for conventional 2c-2e bonding. Theoretical frameworks provide deeper insight into these bonds. describes the H–H bond in H₂ as arising from the overlap of two 1s atomic orbitals, with between covalent and ionic contributions stabilizing the molecule, as originally proposed by Heitler and London. , applied to simple cases like the H₂⁺ ion, illustrates bonding through the formation of a σ from two 1s atomic orbitals, occupied by one , resulting in a of 0.5; the antibonding σ* orbital remains empty.

Metallic Bonding

In metallic hydrides, also known as interstitial hydrides, hydrogen atoms occupy octahedral or tetrahedral interstitial sites within the close-packed metal , leading to a mechanism characterized by delocalized electrons. The metal atoms form a positively charged , while contributes as nearly bare protons; the conduction electrons from the metal, augmented by those from , form an itinerant "electron gas" that binds the structure through electrostatic attraction, similar to pure metals but modified by 's presence. This delocalization is exemplified in palladium hydride (PdH_x), where acts as a proton embedded in the , with x typically ranging up to about 0.6 in the stable phase, enhancing expansion and electronic conductivity without forming discrete bonds. Band theory provides a framework for understanding these interactions, where hydrogen perturbs the host metal's electronic structure by hybridizing its 1s orbitals with the metal's s-p and d bands, often introducing new bands below the Fermi level. In transition metal hydrides, the metal s-band lowers significantly (e.g., by ~3 eV in PdH_x), creating a peak in the density of states that accommodates additional electrons from hydrogen, while d-bands remain largely unaffected at low concentrations. This leads to phase transitions, such as the α-to-β transition in the Pd-H system around x \approx 0.6, where the α-phase (low hydrogen solubility, x < 0.02) features localized Pd-H covalent-like interactions with hydrogen 1s electrons below the Fermi level, transitioning to the β-phase with delocalized conduction electrons contributing to σ-bonding bands. Solubility limits arise from band filling; for instance, the α-phase's low capacity stems from electrons occupying states that minimize energy up to a critical concentration, beyond which the β-phase with expanded lattice accommodates more hydrogen. Non-stoichiometry is a hallmark of metallic hydrides, allowing variable hydrogen content without fixed ratios, as hydrogen randomly occupies interstitial sites up to a solubility limit before forming ordered hydride phases. For example, exhibits compositions in the range TiH_{1.5-2}, where the δ-phase accommodates hydrogen variably between 1.73 and 1.99 atoms per titanium, influenced by temperature and pressure, leading to disordered occupancy at intermediate loadings and ordered arrangements in hydride precipitates. This variability arises from the delocalized nature of the electrons, enabling flexible site filling without disrupting the overall metallic cohesion. Theoretical models elucidate these behaviors through concepts like the electron-to-atom (e/a) ratio, which governs phase stability analogous to in alloys; in hydrides, hydrogen donation increases the e/a ratio, filling bands and promoting dihydride formation when it reaches ~7-8 for early transition metals. Pseudopotential approximations further model H-metal interactions by treating hydrogen's core as a weak pseudopotential, allowing efficient computation of band structures and binding energies; for instance, in iron-series hydrides, these methods reveal that hydrogen stabilizes octahedral sites by screening metal d-electrons, with binding energies scaling with the pseudopotential form factor.

Types of Hydrides

Ionic Hydrides

Ionic hydrides, also known as saline hydrides, are binary compounds formed primarily by the reaction of hydrogen with highly electropositive s-block metals such as alkali and alkaline earth elements, resulting in the hydride anion (H⁻) paired with metal cations. These compounds exhibit predominantly ionic bonding, characterized by strong electrostatic interactions between the small, highly basic H⁻ ions and the metal cations, leading to high lattice energies and stability in solid form. Unlike covalent or metallic hydrides, ionic hydrides are typically white, crystalline solids that are nonvolatile and insoluble in non-protic solvents, though they display extreme reactivity toward water and other protic compounds due to the nucleophilic nature of H⁻./Descriptive_Chemistry/Elements_Organized_by_Block/1_s-Block_Elements/Group__1%3A_The_Alkali_Metals/2Reactions_of_the_Group_1_Elements/Group_1_Compounds) The synthesis of ionic hydrides generally involves the direct combination of the elemental metal with hydrogen gas under controlled high-temperature conditions to overcome kinetic barriers and ensure complete reaction. For instance, alkali metal hydrides like (NaH) are produced via the reaction 2Na + H₂ → 2NaH, typically carried out at temperatures between 300°C and 400°C in an inert atmosphere to prevent oxidation, yielding nearly quantitative conversion after several hours. Alkaline earth hydrides, such as (CaH₂), follow a similar direct hydrogenation route but require higher temperatures, around 400–500°C, due to the greater stability of the metal lattice: Ca + H₂ → CaH₂. Alternative methods include electrolysis of fused metal halides in the presence of hydrogen or reactions involving metal amides, though these are less common for simple ionic hydrides and are reserved for specialized preparations. In terms of crystal structure, most ionic hydrides of alkali metals adopt the rock-salt (NaCl-type) lattice, a face-centered cubic arrangement where each metal cation is octahedrally coordinated by six H⁻ anions, and vice versa, reflecting the similar ionic radii of H⁻ (approximately 140 pm) and halides. For NaH, this manifests as a cubic unit cell with a lattice parameter of 4.88 Å at room temperature, contributing to its high density (1.36 g/cm³) and mechanical stability. In contrast, some rare earth ionic hydrides, such as lanthanum dihydride (LaH₂), exhibit a fluorite (CaF₂-type) structure, where La³⁺ cations occupy a face-centered cubic lattice with H⁻ anions in tetrahedral sites, resulting in a more open framework with a lattice parameter around 5.55 Å. These structural motifs underscore the ionic character, with minimal covalent contributions except in borderline cases like aluminum hydride (AlH₃), which forms a polymeric network with partial ionic bonding but is often considered transitional to covalent hydrides. The physical and chemical behaviors of ionic hydrides are dominated by their strong ionic lattices, leading to elevated melting points; for example, lithium hydride (LiH) melts at 680°C without decomposition, the highest among alkali hydrides, due to its particularly high lattice energy from the small Li⁺ cation. Chemically, they act as powerful reducing agents and bases, reacting vigorously with protic solvents to liberate hydrogen gas; the hydrolysis of NaH proceeds as NaH + H₂O → NaOH + H₂, an exothermic process with ΔH = -83.6 kJ/mol for solid NaOH product, driven by the protonation of H⁻ to form H₂. Thermal decomposition occurs at high temperatures, reversing the synthesis: for instance, 2LiH → 2Li + H₂ above 900°C under vacuum, with stability decreasing down the alkali metal group due to increasing cation size and weaker lattice energies./Descriptive_Chemistry/Elements_Organized_by_Block/1_s-Block_Elements/Group__1%3A_The_Alkali_Metals/2Reactions_of_the_Group_1_Elements/Group_1_Compounds) Representative examples include alkali hydrides like , valued for its thermal stability in hydrogen storage applications, and NaH, widely used in organic synthesis as a base; alkaline earth variants such as CaH₂ serve as desiccants and reducing agents in metallurgy.

Covalent Hydrides

Covalent hydrides are molecular compounds formed primarily by elements of groups 14 through 17 of the periodic table, where hydrogen shares electrons with the central atom through covalent bonds, often with significant polarity due to electronegativity differences. In group 14, exemplifies a stable, nonpolar hydride with strong C-H bonds, while and show decreasing bond strengths down the group owing to larger atomic sizes and weaker orbital overlap. Group 15 hydrides, such as , exhibit basic character from the lone pair on nitrogen, with bond polarity increasing up the group; is less basic and more volatile. In group 16, displays acidic properties in its proton-donating ability, contrasting with the weaker acidity of , and bond strengths diminish down the group. Group 17 hydrides like are highly polar, with strong H-F bonds enabling hydrogen bonding, while heavier analogs such as show reduced polarity and acidity. These hydrides generally exhibit high volatility and low melting points due to their molecular nature, often existing as gases or low-boiling liquids at room temperature. For instance, (PH₃) is a gas with a boiling point of -87.7°C, and (H₂S) boils at -60°C, reflecting weak intermolecular forces in the absence of hydrogen bonding. However, hydrogen bonding significantly elevates boiling points in (-33°C), (100°C), and (19.5°C), leading to liquid states under ambient conditions despite their small molecular sizes. Synthesis of covalent hydrides often involves hydrolysis or reduction methods tailored to the element's reactivity. Silane (SiH₄), for example, is prepared by the reaction of magnesium silicide (Mg₂Si) with hydrochloric acid: Mg₂Si + 4 HCl → 2 MgCl₂ + SiH₄, yielding a colorless, pyrophoric gas used in semiconductor production. Heavier group 14 analogs like germane (GeH₄) are similarly synthesized but display greater instability, decomposing thermally above approximately 280°C to form germanium and hydrogen, though it is handled under controlled conditions due to pyrophoricity, limiting its practical handling compared to the more stable SiH₄. Certain covalent hydrides deviate from simple mononuclear structures, including electron-deficient variants like boron hydrides. Diborane (B₂H₆) features a bridged structure with three-center two-electron B-H-B bonds, resulting in only 12 valence electrons for 14 needed in a conventional Lewis sense, which imparts high reactivity and volatility (boiling point -92.5°C). Complex variants such as lithium aluminum hydride (LiAlH₄) adopt a polymeric solid-state structure, comprising a three-dimensional network of AlH₄ tetrahedra corner-shared with LiH₅ trigonal bipyramids, enhancing its utility as a selective reducing agent in synthesis.

Metallic Hydrides

Metallic hydrides, also referred to as interstitial hydrides, form through the absorption of gas into the crystal lattices of transition metals, primarily d-block elements, where hydrogen atoms occupy octahedral or tetrahedral interstitial sites without significantly altering the host metal structure. This process occurs reversibly under moderate conditions for certain metals, such as palladium, where the reaction \ce{Pd + 0.5 H2 -> PdH_{0.7}} proceeds at and , resulting in up to 70% occupancy of octahedral sites. The formation is driven by the exothermic interaction between hydrogen and the metal, leading to non-stoichiometric compounds that exhibit metallic conductivity and properties intermediate between alloys and true hydrides. The of hydride formation are characterized by pressure-composition isotherms (PCI curves), which plot the equilibrium against the hydrogen-to-metal ratio (H/M) at constant , revealing the and desorption behavior. These isotherms typically feature flat plateaus corresponding to two-phase coexistence regions, where uptake occurs at nearly constant until a is complete, followed by steep rises or falls indicating single-phase solid solutions. For instance, in the palladium- system, PCI curves demonstrate a plateau around 0.02 at 25°C for the \alpha \to \beta , enabling precise control of loading. , observed as a difference between and desorption branches, arises from and interface pinning during phase changes. Phase diagrams for metallic hydrides, such as those for the Pd-H , delineate regions of the low-hydrogen \alpha- (a dilute with H/M < 0.02) and the high-hydrogen \beta- (a hydride with H/M up to 0.7), separated by a two-phase coexistence area that widens with decreasing temperature. The \beta- involves significant lattice expansion (up to 10% volume increase), contributing to the observed hysteresis and potential mechanical instability. In intermetallic alloys, similar diagrams apply; for example, LaNi_5H_6 forms a \beta- hydride with a plateau pressure of about 2 bar at room temperature, making it suitable for practical applications due to its reversible cycling over thousands of absorptions. Representative examples illustrate behavioral differences across groups. In groups 4 and 5, hydrides like and (up to ) are brittle, with the \delta-phase exhibiting high hardness but poor reversibility, requiring temperatures above 400°C for hydrogen release due to strong metal-hydrogen bonding. Conversely, group 10 metals like palladium form reversible hydrides, with (x \approx 0.6-0.7) showing rapid absorption/desorption kinetics and minimal degradation over cycles, attributed to weaker hydride stability. These properties stem from electronic factors, such as d-band filling, influencing hydrogen solubility and phase stability. Engineering challenges in metallic hydrides include hydrogen embrittlement, particularly in steels, where absorbed hydrogen diffuses to crack tips, reducing fracture toughness and promoting brittle intergranular or quasi-cleavage failure under tensile stress. This phenomenon has led to catastrophic failures in high-strength pipelines and structures exposed to hydrogen-rich environments, with susceptibility increasing with steel yield strength above 1000 MPa. Hydrogen diffusion coefficients provide insight into embrittlement rates; in palladium, the value is approximately $1.3 \times 10^{-7} cm²/s at room temperature, facilitating rapid ingress but also enabling modeling of permeation barriers. Mitigation strategies, such as alloying with elements like Nb or coatings, aim to lower solubility and enhance resistance without compromising hydride utility.

Transition Metal Hydride Complexes

Transition metal hydride complexes are coordination compounds in which hydrogen atoms are bound to transition metal centers, typically as hydride ligands (H⁻) or dihydrogen units (η²-H₂), distinguishing them from bulk metallic hydrides by their molecular, soluble nature. These complexes play crucial roles in catalysis and hydrogen activation, exhibiting diverse structures and reactivity due to the d-block metals' ability to engage in multiple bonding interactions. The structures of transition metal hydride complexes include terminal M-H bonds, where a single hydride is directly attached to one metal atom, as seen in the classic example HCo(CO)₄, first synthesized in 1937, or the polyhydride [FeH₄(PPh₃)₃]⁻. Bridged hydrides feature M-H-M motifs, with the hydrogen shared between two metals, commonly observed in dinuclear species and enzymatic active sites like those in NiFe hydrogenases. Additionally, side-on η²-H₂ complexes coordinate dihydrogen as a two-electron donor, exemplified by [Ru(H₂)(NH₂NEt₂)₃]²⁺, discovered by Kubas in 1984, where the H-H bond length elongates slightly from the free H₂ value. These structural variations allow for fluxional behavior and interconversion between forms. Bonding in these complexes primarily involves σ-donation from the hydride ligand to the empty metal orbital, complemented by π-backbonding from filled metal d-orbitals to the hydride's antibonding orbital, which strengthens the M-H interaction and influences reactivity. Many stable complexes adhere to the 18-electron rule, achieving an octet plus d10 configuration, as in Mn(CO)₅H. Formation often proceeds via oxidative addition of H₂ to a low-valent metal precursor, such as the reaction of Mn₂(CO)₁₀ with H₂ to yield Mn(CO)₅H, a process that increases the metal's oxidation state by two units while splitting the H-H bond. This mode contrasts with heterolytic pathways but underscores the role of backbonding in stabilizing the product. Synthesis of transition metal hydride complexes commonly involves the addition of H₂ under pressure to coordinatively unsaturated precursors, as in the formation of η²-H₂ complexes, or protonation of metalate anions with acids, generating terminal hydrides like those from [Re(CO)₅]⁻. Alternative routes include reaction with hydride donors such as NaBH₄ or LiAlH₄. These complexes are generally air-sensitive and require inert atmospheres for stability, with second- and third-row metals forming more robust M-H bonds due to better orbital overlap. The acidity of transition metal hydrides varies significantly with the metal, ligands, and oxidation state, ranging from strongly acidic (protic character) in electron-poor late metals to basic (hydridic character) in early metals. For instance, HCo(CO)₄ exhibits a pKₐ of approximately 8.4 in acetonitrile, reflecting its enhanced acidity due to π-acceptor CO ligands that deplete electron density from the metal and thus the hydride. In contrast, complexes like Cp₂ReH are far less acidic, with pKₐ values around 20 or higher. This tunability arises from the partial positive charge on hydrogen in protic hydrides, enabling proton-transfer reactions, while hydridic ones act as nucleophiles; estimation methods using ligand acidity constants provide predictive power for pKₐ values across solvents.

Mixed Anion Hydrides

Mixed anion hydrides are crystalline compounds in which the hydride anion (H⁻) coexists with other anions, such as oxide (O²⁻) or nitride (N³⁻), within the same lattice, often leading to novel structural and electronic properties due to the differing ionic sizes and electronegativities of the anions. These materials frequently adopt perovskite or fluorite-related structures that accommodate the mixed anions in ordered or disordered arrangements. Representative examples include oxyhydrides like lanthanum oxyhydride (LaHO), which possesses a fluorite superstructure, and barium scandium oxyhydride (BaScO₂H), a cubic perovskite where H⁻ occupies specific interstitial sites. Other notable oxyhydrides, such as strontium vanadium oxyhydride (SrVO₂H), feature perovskite lattices with H⁻ in apical positions alongside O²⁻. Examples involving nitrides include calcium chromium nitride hydride (Ca₃CrN₃H), a mixed-anion catalyst structure, and barium titanium oxynitride hydrides like BaTiO₃₋ₓN₂ₓ/₃Hᵧ, which incorporate N³⁻, O²⁻, and H⁻. Synthesis of mixed anion hydrides typically requires conditions that stabilize the reactive H⁻ anion against decomposition, such as high-pressure reactions or topochemical reductions. For instance, BaScO₂H is synthesized by reacting BaO and Sc₂O₃ with CaH₂ under 5 GPa and 1000 °C to facilitate hydride incorporation without H₂ loss. Similarly, LaHO can be prepared via solid-state reaction of La₂O₃ with CaH₂ at high pressure (around 5 GPa) and elevated temperatures, promoting the exchange and diffusion of anions. Topochemical methods, involving low-temperature (300–600 °C) reduction of parent oxides or nitrides with metal hydrides like CaH₂ under hydrogen atmospheres, are also common for oxyhydrides and nitride hydrides, enabling controlled anion substitution. For nitride-containing variants like Ca₃CrN₃H, solid-state metathesis or high-temperature reactions under inert conditions yield the mixed phases. These compounds exhibit distinctive properties arising from the interplay of anions, including high ionic conductivity driven by the smaller size and greater mobility of H⁻ compared to O²⁻ in similar lattices, enabling pure hydride-ion conduction in materials like LaHO and SrVO₂H. Oxyhydrides such as LaHO demonstrate stability under reducing conditions but are susceptible to hydrolysis, releasing H₂ upon exposure to water. In perovskite oxyhydrides like BaScO₂H and SrVO₂H, local anion ordering contributes to insulating behavior and unique magnetic properties, including room-temperature antiferromagnetism in the latter. Nitride hydrides, exemplified by Ca₃CrN₃H, show enhanced vibrational dynamics of H⁻, supporting catalytic applications. Recent developments since 2010 have focused on oxyhydrides as proton or hydride conductors, with discoveries like SrVO₂H (reported in 2016) revealing ordered anion arrangements that boost H⁻ transport for potential use in solid-state ionic devices. Advances in high-pressure synthesis have expanded the family, including yttrium-based oxyhydrides with tunable multi-anion chemistries for photocatalysis and energy storage. As of 2025, advances include Ba–Si orthosilicate oxynitride–hydride catalysts for ammonia synthesis via anion vacancies and high-conductivity hydroborates like Na₃(BH₄)(B₁₂H₁₂) for solid-state ionic devices.

Properties

Physical Properties

Hydrides exhibit diverse physical states depending on their bonding type. Ionic hydrides, formed by alkali and alkaline earth metals, are typically white, crystalline solids with densities ranging from approximately 0.8 to 2.5 g/cm³; for example, (LiH) has a density of 0.82 g/cm³, while (NaH) is 1.39 g/cm³. Covalent hydrides, involving nonmetals, often exist as gases or low-boiling liquids at room temperature, with very low densities; (CH₄), a representative covalent hydride, has a gas density of about 0.0007 g/cm³ at standard temperature and pressure. Metallic hydrides, interstitial compounds in transition metals, are solids with densities similar to the parent metal but slightly reduced due to lattice expansion; (PdH_{0.6}) has a density around 11 g/cm³. Thermal properties of hydrides vary significantly across classes. Ionic hydrides generally have high melting points due to strong electrostatic interactions, such as LiH melting at 688°C. In contrast, covalent hydrides display low boiling points, exemplified by ammonia (NH₃) at -33°C. Specific heat capacities are moderate; LiH, for instance, has a value of 3.51 J/g·K at room temperature. Metallic hydrides show thermal expansion influenced by hydrogen insertion, with interstitial hydrides like PdH exhibiting coefficients around 14 × 10^{-6} K^{-1} at room temperature, augmented by hydrogen-induced lattice changes. Mechanically, ionic hydrides are hard and brittle, fracturing under shear due to ion misalignment in their lattice structures. Metallic hydrides often display increased brittleness compared to pure metals, attributed to hydrogen-embrittlement effects. Optically, some hydrides like lithium deuteride (LiD) are transparent in the visible spectrum, enabling applications in optical diagnostics. Hydrogen absorption in metallic hydrides induces significant volume expansion, typically 10-20% for full hydride formation, as seen in palladium systems where the lattice swells by about 10% at PdH_{0.5}. Spectroscopic techniques reveal key structural features of hydrides. Infrared (IR) spectroscopy identifies metal-hydrogen (M-H) stretching vibrations in the 1500-2200 cm^{-1} range, with terminal hydrides often appearing near 2000 cm^{-1}. Nuclear magnetic resonance (NMR) spectroscopy, particularly {}^1H NMR, probes hydrogen environments, distinguishing hydride signals in complex systems and enabling quantification in metal hydrides under high pressure.

Chemical Properties

Hydrides exhibit distinct reactivity patterns depending on their type, with ionic hydrides undergoing rapid, exothermic hydrolysis upon contact with water. For instance, reacts vigorously according to the equation CaH₂ + 2H₂O → Ca(OH)₂ + 2H₂, releasing hydrogen gas and heat that may ignite the evolved hydrogen. In contrast, covalent hydrides hydrolyze more slowly; (B₂H₆), for example, decomposes in water to form (H₃BO₃) and hydrogen gas, though the reaction is still exothermic and can be violent in moist air. Many hydrides are highly air-sensitive due to their strong reducing nature, stemming from the hydride ion's low standard reduction potential (E° ≈ -2.25 V for H₂ + e⁻ → H⁻), which makes them prone to oxidation. Lithium aluminum hydride (LiAlH₄), a representative complex hydride, ignites spontaneously upon exposure to moist air or water, generating flammable hydrogen gas and potentially exploding due to frictional heat or static discharge. This flammability underscores their utility as powerful reducing agents but necessitates inert handling conditions. Thermal decomposition is a key reactivity mode for many hydrides, often reversible and exploited in hydrogen storage applications. Magnesium hydride (MgH₂) decomposes above approximately 300°C under 1 bar H₂ pressure via MgH₂ → Mg + H₂, allowing reabsorption of hydrogen upon cooling, though the process is kinetically sluggish without catalysts. Certain transition metal hydride complexes undergo photochemical decomposition, where irradiation induces dissociative loss of H₂ or hydride transfer, facilitating photocatalytic hydrogen evolution. Stability of hydrides generally increases with the electropositivity of the metal, as more electropositive elements form stronger ionic bonds with H⁻, enhancing resistance to decomposition; for example, alkali metal hydrides are more stable than those of less electropositive transition metals. Impurities, such as trace oxides or halides, can catalyze decomposition by lowering activation energies and reducing onset temperatures, thereby accelerating hydrogen release in storage systems.

Isotopic Variants

Hydrides incorporating hydrogen isotopes other than protium (^1H) exhibit distinct behaviors due to differences in atomic mass, which influence bond strengths, vibrational frequencies, and reaction kinetics. Deuterium (^2H or D) and tritium (^3H or T) are the primary heavy isotopes relevant to hydride chemistry, with deuterium being stable and tritium radioactive with a half-life of 12.32 years via beta decay. The increased mass of these isotopes leads to stronger bonds compared to protium counterparts; for instance, the bond dissociation energy of D_2 is 443 kJ/mol, higher than that of H_2 at 436 kJ/mol, arising from reduced zero-point vibrational energy in the heavier isotope. Deuterides and tritides are prepared through methods that leverage isotopic exchange or direct synthesis under controlled conditions to achieve high purity. Deuterides are commonly prepared by direct reaction of the metal with deuterium gas; for example, lithium deuteride (LiD) is synthesized by heating lithium metal in an atmosphere of deuterium gas. Electrolysis of water enriched in heavy isotopes is another key technique for concentrating , often used in the production of heavy water (D_2O) as a precursor for deuteride synthesis, exploiting the slight isotopic fractionation during the process. Tritides, like uranium tritide (UT_3), are typically formed by exposing uranium metal to gas at moderate temperatures, absorbing up to three equivalents per uranium atom to form the stoichiometric tritide phase. The properties of isotopic hydrides differ notably from protium-based ones due to mass-dependent effects on intermolecular forces and reaction rates. Deuterides generally display higher melting and boiling points; for example, D_2O has a melting point of 3.82 °C compared to 0 °C for H_2O, attributed to stronger hydrogen bonding from lower zero-point vibrations. A prominent feature is the kinetic isotope effect (KIE), where reactions involving deuterium transfer occur more slowly than with protium, with typical primary KIE values of k_H/k_D around 7 for hydrogen atom or proton transfers, stemming from the higher activation energy required to stretch the heavier D-X bond. In tritides, the radioactive decay of T to ^3He introduces additional challenges, such as helium accumulation leading to structural swelling and potential release of tritium gas. These isotopic variants find specialized applications leveraging their unique nuclear and physical traits. Lithium deuteride (LiD) serves as an effective neutron moderator in nuclear reactors due to its high hydrogen density and low neutron absorption cross-section, slowing fast neutrons without significant capture, as demonstrated in hydride reflector designs. Tritides like UT_3 are employed in tritium storage and handling systems for fusion research, where the beta decay produces ^3He, but the material's stability allows controlled release; notably, the D-T fusion reaction, involving deuterium and tritium, releases 17.6 MeV per fusion event and powers experimental reactors like ITER. Such uses highlight the role of isotopic hydrides in advancing nuclear technologies beyond conventional protium systems.

Applications

Reducing Agents and Bases

Hydrides serve as versatile reducing agents in organic synthesis, particularly ionic and covalent variants that deliver hydride ions to electrophilic centers. Sodium borohydride (NaBH₄) is a mild, selective reductant commonly used to convert aldehydes and ketones to primary and secondary alcohols, respectively, without affecting esters or carboxylic acids under ambient conditions. For instance, in the reduction of benzoin, NaBH₄ adds a hydride to the carbonyl carbon, forming a stable alkoxide intermediate that protonates upon workup to yield hydrobenzoin. In contrast, lithium aluminum hydride (LiAlH₄) is a stronger reducing agent capable of fully reducing esters to primary alcohols by delivering multiple hydrides, cleaving the C-O bond in the process. This reagent transforms ethyl acetate, for example, into ethanol via sequential hydride additions and elimination steps. Beyond reduction, certain metal hydrides function as strong bases in synthetic applications by deprotonating weakly acidic sites. Sodium hydride (NaH) effectively deprotonates carbon acids to form enolates, facilitating reactions like alkylation in organic synthesis. For example, treatment of acetone with NaH generates the sodium enolate, which can then react with alkyl halides to produce α-substituted ketones. Potassium hydride (KH), an even stronger base due to the larger potassium cation, is employed in the anionic polymerization of epoxides such as propylene oxide, initiating ring-opening by abstracting a proton from the monomer or solvent. This leads to polyether chains with controlled microstructure, as KH's high reactivity ensures rapid initiation at room temperature in tetrahydrofuran. The reactivity of these hydrides stems from hydride transfer mechanisms, which can proceed concertedly or stepwise depending on the substrate and conditions. In NaBH₄ reductions, hydride delivery to the carbonyl is typically a concerted process involving nucleophilic attack and simultaneous proton transfer, forming a tetrahedral intermediate. LiAlH₄ reductions of esters involve stepwise hydride additions, with the first hydride forming an aldehyde intermediate that is further reduced in situ. Side reactions, such as over-reduction, can occur with excess reagent or protic impurities, leading to pinacol coupling in ketones or cleavage in sensitive functional groups. Historically, sodium hydride was first synthesized in 1907 by through the direct combination of sodium metal and hydrogen gas under high pressure and temperature, marking an early milestone in metal hydride chemistry. Its industrial production scaled up significantly after World War II, driven by demand for base-catalyzed condensations and alkylations in pharmaceutical and polymer manufacturing, with processes involving molten sodium hydrogenation in inert media. This development enabled widespread adoption of NaH in laboratory and large-scale synthesis.

Hydrogen Storage and Energy

Hydrides play a crucial role in hydrogen storage for energy applications, offering compact, solid-state solutions that exceed the volumetric density of compressed or liquid hydrogen while approaching gravimetric targets for onboard vehicle use. Metallic hydrides, such as (MgH₂), and complex hydrides, like (NaAlH₄), enable reversible hydrogen absorption and desorption through exothermic and endothermic reactions, respectively, making them suitable for fuel cell vehicles and stationary power systems. These materials address the need for safe, efficient storage amid growing demand for clean energy carriers, with research focusing on overcoming thermodynamic and kinetic barriers to meet practical deployment goals. Reversible hydrogen storage in hydrides occurs via interstitial absorption in metallic lattices or multistep dehydrogenation in complex systems. For instance, absorbs up to 7.6 wt% hydrogen, forming a hydride phase under moderate pressure, but requires temperatures around 300°C for desorption due to its high stability. Kinetics in MgH₂ have been enhanced through doping with transition metals like titanium or nickel, reducing activation energies and enabling faster hydrogen release at lower temperatures, such as below 250°C after multiple cycles. Similarly, offers 5.6 wt% reversible capacity through a two-step process: NaAlH₄ → Na₃AlH₆ + Al + 1.5 H₂ (3.7 wt%) followed by Na₃AlH₆ → 3 Na + Al + 3 H₂ (1.9 wt%), with desorption initiating at approximately 100-150°C under ambient pressure. Catalysts like titanium chloride have historically improved NaAlH₄ reversibility, maintaining over 90% capacity retention after dozens of cycles. Key performance metrics for hydride storage systems align with U.S. Department of Energy (DOE) targets for light-duty vehicles, which specify 5.5 wt% gravimetric and 0.040 kg H₂/L volumetric capacity by 2025, progressing to ultimate goals of 6.5 wt% and 0.050 kg H₂/L. MgH₂ exceeds the gravimetric target but falls short volumetrically at around 0.110 kg H₂/L in bulk form, while NaAlH₄ meets both at 5.6 wt% and higher densities due to its complex structure. Desorption temperatures remain a benchmark, with DOE emphasizing systems operable below 85°C for automotive refueling, though current hydrides like alanates operate at 100-200°C, necessitating heat management strategies. Cycle life targets exceed 5,000 absorptions/desorption events with minimal capacity fade, a metric where doped MgH₂ achieves over 92% retention after 10 cycles. In energy applications, hydrides integrate with fuel cells by providing on-demand hydrogen for proton-exchange membrane systems, enhancing vehicle range without high-pressure tanks. Complex hydrides like serve as solid electrolytes in all-solid-state lithium batteries, offering ionic conductivities up to 10⁻³ S/cm at room temperature when stabilized with additives like ZrO₂, enabling safer, higher-energy-density alternatives to liquid electrolytes with capacities exceeding 300 mAh/g. Hydride-based anodes, such as binary mixtures of and , support reversible lithium storage in full cells, demonstrating stable cycling over 100 charge-discharge events at potentials above 1 V. These advancements position hydrides in hybrid energy storage for electric vehicles and grid-scale batteries. Recent progress from 2020-2025 emphasizes nanoconfinement to accelerate kinetics, where hydrides are infiltrated into nanoporous scaffolds like carbon aerogels or metal-organic frameworks, reducing particle sizes below 5 nm and lowering desorption barriers by up to 50%. For , confinement in graphene scaffolds yields hydrogen release at 150°C with 6 wt% capacity retention after 100 cycles, surpassing bulk performance. NaAlH₄ nanoconfined in mesoporous silica achieves full reversibility at 80°C, aligning closer to DOE operability targets. These techniques also mitigate sintering, preserving microstructure during cycling. Despite advancements, challenges persist in hydride systems, including limited cycle life from particle agglomeration and phase segregation, which can reduce capacity by 20-30% after 50 cycles without stabilization. Safety concerns arise from uncontrolled exothermic hydrogen release, potentially leading to thermal runaway, though hydrides' solid form inherently lowers explosion risks compared to gaseous storage. Ongoing research targets durable scaffolds and alloying to extend life beyond 1,000 cycles while ensuring controlled desorption for practical energy integration.

Catalytic and Synthetic Uses

Hydrides play a pivotal role in catalytic hydrogenation processes, particularly through transition metal complexes that facilitate the addition of hydrogen to unsaturated substrates. Wilkinson's catalyst, chlorotris(triphenylphosphine)rhodium(I) [RhCl(PPh₃)₃], exemplifies this application by enabling the homogeneous hydrogenation of alkenes under mild conditions, typically at room temperature and atmospheric pressure. The catalytic cycle begins with the oxidative addition of dihydrogen (H₂) to the rhodium center, forming a dihydride intermediate, followed by coordination of the alkene and migratory insertion of the hydride into the metal-alkene bond to generate an alkyl hydride species. This step is succeeded by a second oxidative addition of H₂ and reductive elimination of the alkane product, regenerating the catalyst. Such mechanisms highlight the transient hydride species as key intermediates that drive high selectivity and efficiency, with turnover frequencies for alkene hydrogenation reaching up to 650 h⁻¹ under standard conditions. In synthetic applications, hydrides serve as reagents in specialized reductions beyond simple hydrogenations. Triethylsilane (Et₃SiH), an organosilicon hydride, acts as a mild hydrogen atom donor in radical chain reductions, often replacing toxic tributyltin hydride (Bu₃SnH) for dehalogenations and azide reductions to amines. The mechanism involves initiation by a radical source, such as AIBN, generating a silyl radical that propagates the chain by abstracting hydrogen from additional Et₃SiH while transferring the radical to the substrate. Similarly, borane-tetrahydrofuran complex (BH₃·THF) is employed in hydroboration reactions, where the B-H bond adds across alkenes in an anti-Markovnikov, syn fashion to form organoboranes, which can be oxidized to alcohols. The addition proceeds via a four-center transition state, with boron attaching to the less substituted carbon, underscoring the hydride's role in directing regioselectivity. Industrial processes indirectly leverage hydride chemistry. Recent advancements in ruthenium-loaded metal hydrides have enabled lower-temperature ammonia synthesis by stabilizing hydride ions that enhance N₂ dissociation. In asymmetric hydrogenation, rhodium complexes with chiral diphosphine ligands, pioneered by Knowles, achieve enantioselective reductions of prochiral alkenes via hydride intermediates, with turnover numbers exceeding 10⁴ in optimized systems. This work, shared in the 2001 Nobel Prize with Noyori's ruthenium-based systems, has transformed pharmaceutical synthesis by enabling >99% enantiomeric excess in large-scale productions.

Nomenclature

Naming Conventions

Hydrides are named according to systematic IUPAC rules that distinguish between compounds, molecular structures, and coordination complexes, while also retaining certain common names for widespread use. For hydrides involving metals, the nomenclature employs the format "element hydride," where the metal name precedes "hydride," as seen in for NaH. In contrast, hydrides of nonmetals are named as "hydrogen element," such as for HCl, reflecting the compositional order based on . Molecular hydrides, which form discrete molecules, typically follow substitutive derived from parent hydride names, replacing atoms with substituents as needed; for instance, serves as the parent name for CH₄. For coordination compounds or anionic complexes, additive is applied, using prefixes like "tetrahydro" to indicate ligands, as in tetrahydroaluminate for [AlH₄]⁻. Common names coexist with systematic ones for simplicity and historical reasons, particularly for well-known compounds; is the retained name for H₂O rather than the systematic oxidane, and is used for SiH₄ instead of silane being strictly substitutive. Polyhydrides, where multiple hydrogen atoms are present, incorporate stoichiometric prefixes to specify composition, such as calcium dihydride for CaH₂. Isotopic variants may employ terms like deuteride in place of hydride when substitutes .

Precedence Rules

In the nomenclature of mixed-anion compounds containing hydrides, IUPAC establishes a hierarchy of anions based on electronegativity and the order specified in Table VI of the recommendations, where oxygen-containing anions take precedence over hydride (H⁻). For instance, in oxohydrides, the compound H₂O is named dihydrogen oxide or oxidane, treating the oxide (O²⁻) as the parent anion rather than hydride, with hydrogen atoms as substituents. Similarly, halogens precede hydride in precedence due to their higher electronegativity, as seen in binary compounds like HCl named hydrogen chloride, where chloride is the anionic component. For compounds with multiple anions, the parent structure is selected based on the senior anion according to this , with lower-precedence anions named as substituents using additive . In mixed anion hydrides, such as BaScO₂H, the name is hydride, where the serves as the parent and hydride as the . This approach ensures the primary anionic component reflects the compound's dominant structural or chemical character, with hydride often appearing alphabetically or as a prefix in the full name. In coordination compounds, hydride acts as a named with the prefix "hydrido-", ordered alphabetically among other ligands and following rules for anionic endings in "-ido". For example, the complex RhCl(H)(PPh₃)₂ is named chlorohydridobis(triphenylphosphine)(I), where "hydrido" precedes other ligands like "chlorido" in alphabetical listing, and the metal is indicated in . For isotopically labeled hydrides, IUPAC specifies notation using nuclide symbols in square brackets for specific substitution or superscripts for general cases, with (²H) and (³H) following standard rules but preferring numerical superscripts over symbols like D or T. An example is [²H]methane for deuterated (CH₃D or variants), while tritium-labeled compounds use (³H₁) or similar precise locants to denote position, ensuring clarity in spectroscopic or mechanistic contexts. Exceptions to these precedence rules allow retention of historical names for well-established compounds, such as for NH₃ instead of azane, to maintain continuity in despite the systematic preference for parent hydride names.

References

  1. [1]
    Hydrides compounds for electrochemical applications - ScienceDirect
    Hydrides are inorganic compounds in which hydrogen is combined with other elements, mainly metallic or semi-metallic. Based on their chemical bonding, hydrides ...
  2. [2]
    Phenomena of hydrides | Journal of Applied Physics - AIP Publishing
    Nov 8, 2022 · Hydrides are interstitial compounds in which hydrogen atoms occupy voids in the metal lattice. During the formation of hydrides, hydrogen is ...
  3. [3]
    Recent advances in metal hydrides for clean energy applications
    Jun 7, 2013 · Metal hydrides are comprised of metal atoms and hydrogen forming chemical bonds. Many metals react with hydrogen at elevated temperatures and ...<|control11|><|separator|>
  4. [4]
    Sodium hydride - PubChem - NIH
    Sodium hydride appears as a silvery to whitish powder or slurry in oil, used to make other chemicals. It is also known as NaH.
  5. [5]
    Hydrides - Chemistry LibreTexts
    Jun 30, 2023 · Hydrides are classified into three major groups, depending on what elements the hydrogen bonds to. The three major groups are covalent, ionic, and metallic ...
  6. [6]
    Hydrides | Thermo Fisher Scientific - US
    Hydrides can be formed with almost all main-group elements, but inorganic hydrides are usually classified iinto four main groups: covalent, ionic, transition ...
  7. [7]
    Hydride - an overview | ScienceDirect Topics
    Hydride is defined as a class of chemically diverse compounds that contain hydrogen in the negative oxidation state, often utilized in material synthesis as ...
  8. [8]
    WebElements Periodic Table » Hydrogen » electronegativity
    The first scale of electronegativity was developed by Linus Pauling and on his scale hydrogen has a value of 2.20 on a scale running from from about 0.7 (an ...
  9. [9]
    The occurrence and geoscience of natural hydrogen
    This review clearly shows that molecular hydrogen is present in every type of environment on Earth and often plays an important role in many natural processes.
  10. [10]
    Hydride Phase - an overview | ScienceDirect Topics
    A hydride phase is defined as a high concentration phase of a solid solution that forms after reaching the solubility limit at a given temperature, ...
  11. [11]
    HYDRIDE Definition & Meaning - Merriam-Webster
    Sep 21, 2025 · The meaning of HYDRIDE is a compound of hydrogen with a more electropositive element or group ... Word History. First Known Use. 1869, in the ...
  12. [12]
    The Development of Metal Hydride Chemistry - jstor
    CONJECTURAL AND ACTUAL HYDRIDES OF ALKALI METALS. Before actual metal hydrides were discovered, they had a peculiar kind of conjectural precedents. Humphry ...
  13. [13]
    [PDF] 19720066808.pdf - NASA Technical Reports Server (NTRS)
    lithium hydride should be anionic, taking on an electron to form the hydride ion, H-, with the two electron stable con- figuration of the helium atom. Hence ...
  14. [14]
    Hydride ions in oxide hosts hidden by hydroxide ions - PMC
    We demonstrate that the electron density near the hydrogen nucleus in an OH− ion (formally H+ state) exceeds that in an H− ion. This behaviour is the opposite ...Missing: reactivity | Show results with:reactivity
  15. [15]
    [PDF] pKa Values in DMSO Compilation (by Reich and Bordwell)
    Equilibrium pKa Table (DMSO Solvent and Reference). 22.433. Ph3P +-CH2-X. Ph3P ... H2 (400.4)1. NH3 (399.6)1. HO-H (390.8)1. MeO-H (380.6)2. F-H (371.5)1. MeS ...
  16. [16]
    Occurrence, Preparation, and Compounds of Hydrogen – Chemistry
    The compounds formed are crystalline, ionic hydrides that contain the hydride anion, H−, a strong reducing agent and a strong base, which reacts vigorously with ...<|control11|><|separator|>
  17. [17]
    Hydrogen atom - the NIST WebBook
    Electron affinity determinations ; 0.754189, LPES, Lykke, Murray, et al., 1991, Reported: 6082.99±0.15 cm-1, or 0.754195(18) eV; B ; 0.776 ± 0.020, PD, Feldman, ...
  18. [18]
    Hydrogen ionic conductors and ammonia conversions
    Feb 10, 2023 · Conversely protonic conductors are well suited to direct ammonia fuel cell use. Hydride ions can be very mobile and are strongly reducing.
  19. [19]
  20. [20]
  21. [21]
    Coordination of LiH Molecules to Mo≣Mo Bonds - PubMed Central
    The first is almost coincident with the average Mo–H–Mo bond lengths determined by neutron diffraction, while the second is somewhat longer than the 1.60 Å ...
  22. [22]
    [PDF] A Lattice Energy Calculation for LiH - CSB+SJU
    The Born-Haber analysis shown below yields a lattice energy of 912 kJ/mol. Thus, the calculated result of the proposed model is in error by 8%.Missing: cycle | Show results with:cycle
  23. [23]
    Lattice Energies (E - Wired Chemist
    Lattice Energies (E. ) ; LiH, 906 ; LiF, 1009 ; LiCl, 829 ; LiBr, 789.
  24. [24]
    Linus Pauling's Development of an Electronegativity Scale
    ΨA+B¯ stands for the energy contribution arising from what Pauling calls the "additional ionic character of the bond." He shows both possible cases of ionic ...
  25. [25]
    Bond Energies - Chemistry LibreTexts
    Apr 3, 2025 · For example the average bond energy of O-H in H2O is 464 kJ/mol. This is due to the fact that the H-OH bond requires 498.7 kJ/mol to dissociate ...
  26. [26]
    6.1: Electronegativity and Polarity - Chemistry LibreTexts
    Oct 27, 2020 · The absolute values of the electronegativity differences between the atoms in the bonds H–H, H–Cl, and Na–Cl are 0 (nonpolar), 0.9 (polar ...Missing: HCl | Show results with:HCl
  27. [27]
    1.6: sp³ Hybrid Orbitals and the Structure of Methane
    Jan 28, 2023 · The text explains the structure of methane (CH4) using the concept of sp3 hybridization of the central carbon atom.
  28. [28]
    1.9: sp Hybrid Orbitals and the Structure of Acetylene
    Jan 28, 2023 · Acetylene is said to have three sigma bonds and two pi bonds. The carbon-carbon triple bond in acetylene is the shortest (120 pm) and the strongest (965 kJ/mol)
  29. [29]
    6.4: Boron Hydrides - Chemistry LibreTexts
    May 3, 2023 · Diborane represents the archetypal electron deficient dimeric compound, of which Al2Me3 is also a member of this class of electron deficient ...
  30. [30]
    The three-center, two-electron chemical bond - ACS Publications
    Three-center two-electron bonds in the boranes B2H6 and B3H8− from the quantum interference perspective. Theoretical Chemistry Accounts 2020, 139 (8) https ...
  31. [31]
    9.7: Molecular Orbitals - Chemistry LibreTexts
    Jul 7, 2023 · (a) The H2+ ion, (b) the He2+ ion, and (c) the He2 molecule are shown here. for the H2+ ion the H has one unpaired electron in the 1s orbital so ...
  32. [32]
    [PDF] Electronic structure of transition metal hydrides
    For years the electronic structure of metal-hydride systems has been discussed within one- electron band structure theories. This type of theory works well ...
  33. [33]
    Metal hydrides—structure and band structure - ResearchGate
    Aug 6, 2025 · The original transition metal s-band hybridizes with the bonding combination of the two hydrogens, the implications being that for compounds ...
  34. [34]
  35. [35]
    Review The relation between the electron to atom ratio and some ...
    Aug 7, 2025 · Review The relation between the electron to atom ratio ... A survey of theoretical models for primary metallic solutions. Article. Dec 1957; J ...
  36. [36]
    The Production of Sodium Hydride and Some of its Reactions
    Holt ( 17 ) in 1903 described the preparation of sodium hydride. The reaction of hydrogen with sodium can be made to go to completion rapidly by several methods ...
  37. [37]
    Preparation of alkali metal hydrides by mechanical alloying
    There are two methods, based on the reaction of hydrogen on the molten metal in the presence of a temperature gradient: the metal is heated at around 250°C, ...
  38. [38]
    Direct Synthesis of Complex Metal Hydrides | Inorganic Chemistry
    Rehydrogenation of Dehydrogenated NaAlH4 at Low Temperature ... Hydrogenation properties of complex alkali metal hydrides fabricated by mechano-chemical synthesis ...
  39. [39]
    mp-23870: NaH (Cubic, Fm-3m, 225) - Materials Project
    mp-23870, NaH, is a cubic material with a 3.77 eV band gap, non-magnetic, and has a rock salt structure with a lattice of 4.77 Å.
  40. [40]
    mp-24153: LaH2 (cubic, Fm-3m, 225) - Materials Project
    LaH2 is Fluorite structured and crystallizes in the cubic Fm-3m space group. The structure is three-dimensional. La2+ is bonded in a body-centered cubic ...
  41. [41]
    A First-Principles Study of the Electronic Structure and Stability of a ...
    The AlH4 tetrahedral structure is confirmed by the large bonding peaks between Al−H1, Al−H2, Al−H3, and Al−H4 in the hydride (Figure 3). Comparing the PDOSs of ...Missing: LaH2 | Show results with:LaH2<|separator|>
  42. [42]
    Lithium hydride | LiH | CID 62714 - PubChem
    5 Melting Point. 1256 °F (EPA, 1998). U.S. Environmental Protection Agency. 1998. Extremely Hazardous Substances (EHS) Chemical Profiles and Emergency First ...
  43. [43]
    Sodium hydroxide - the NIST WebBook
    Enthalpy of formation of solid at standard conditions. ΔrH°, Enthalpy of reaction at standard conditions. Data from NIST Standard Reference Database 69: NIST ...
  44. [44]
    How do you compare the boiling points of SiH4 PH3 H2S and HCl?
    The boiling point of SiH4: -111.7 °C. The boiling point of PH3: -87.7 °C. The boiling point of H2S: -60 °C. The boiling point of HCl: -85.05 °C.
  45. [45]
    Understanding Silanes Made Easy; Everything You need to Know ...
    ... Mg2Si + 4 HCl → 2 MgCl2 + SiH4. It is also prepared from metallurgical grade silicon in a two-step process. First, silicon is treated with hydrogen chloride ...
  46. [46]
    GERMANE - CAMEO Chemicals - NOAA
    It is flammable. The gas is heavier than air and a flame can flash back to the source of leak very easily. It is toxic by inhalation. Prolonged exposure of the ...
  47. [47]
    mp-27653: LiAlH4 (Monoclinic, P2_1/c, 14) - Materials Project
    LiAlH₄ crystallizes in the monoclinic P2₁/c space group. Li¹⁺ is bonded to five H¹⁻ atoms to form LiH₅ trigonal bipyramids that share corners with five ...
  48. [48]
    Formation and electronic properties of palladium hydrides ... - Nature
    Jun 14, 2017 · The limit of absorption at normal pressures is PdH0.7, indicating that approximately 70% of the octahedral holes are occupied. Hydrogen ...
  49. [49]
    Thermodynamic properties of the Pd-H system - ScienceDirect
    Among all noble metals, Pd is the only element that absorbs hydrogen at ambient temperature and pressure forming an interstitial metallic hydride PdH0.67 [70].
  50. [50]
    Pressure-Composition-Isotherm - an overview | ScienceDirect Topics
    The PCI isotherm indicates the dependence between the hydrogen equilibrium pressure and the amount of hydrogen incorporated into the solid phase at a fixed ...
  51. [51]
    NIST-MSEL Hydrogen Storage Program: Research - Introduction
    The thermodynamic aspects of the hydride formation from gaseous hydrogen can be described by pressure-composition temperature (PCT) curves, which is also called ...
  52. [52]
    Hysteresis in the palladium + hydrogen system - Journals
    The co-existence of both α- and β-phases over a range of pressures is probably associated with the large volume change accompanying the transition and a ...
  53. [53]
    [PDF] Interface pinning causes the hysteresis of the hydride transformation ...
    Jan 26, 2022 · Hysteresis in metal hydrides manifests as a difference in the hydrogen absorption and desorption pressures (pressure hystere- sis) and as a ...
  54. [54]
    LaNi 5 H 6 and Similar Alloys for Hydrogen Storage - Stanford
    Nov 10, 2017 · This article shall focus on the use of LaNi 5 H 6 and similar alloys as a storage system for hydrogen where LaNi 5 H6 is an intermetallic hydride with hydrogen ...
  55. [55]
    [PDF] Room Temperature Metal Hydrides for Stationary and Heat Storage ...
    Apr 9, 2021 · For example, TiH2 requires temperatures up to 800◦C to fully release hydrogen (Suwarno et al., 2012). For stationary applications, hydrogen ...
  56. [56]
    [PDF] Understanding the TiH(2-x)/TiOy System at Elevated Temperature
    Titanium hydride is much more brittle than metallic titanium, and thus is more readily milled into powder. Following thermal cleaning and/or washing steps ...
  57. [57]
    A Review of Group V Transition Metals - Niobium, Vanadium and ...
    Aug 7, 2025 · Thermodynamic data show that the niobium-hydrogen solid solution is more stable than the vanadium- or tantalum-hydrogen solid solutions.
  58. [58]
    Hydrogen Embrittlement of Steel - Industrial Metallurgists
    Hydrogen embrittlement is a metal's loss of ductility and load bearing capability due to hydrogen absorption, causing cracking and fracture. Hydrogen diffuses ...
  59. [59]
    [PDF] The Mechanism of Hydrogen Embrittlement in Steel
    Hydrogen embrittlement occurs when hydrogen atoms diffuse into steel, causing micro-cracks to form and grow due to hydrogen pressure, affecting crack ...
  60. [60]
    The Adsorption and Diffusion of Electrolytic Hydrogen in Palladium
    The diffusion constant for a hydrogen poor ac-palladium has been found to be 1*30 + 0*20 x 10-7 cm2 s-' at room temperature.
  61. [61]
    Introduction: Metal Hydrides | Chemical Reviews - ACS Publications
    Aug 10, 2016 · The first molecular “hydrides” were probably H2Fe(CO)4 (1931) and HCo(CO)4 (1937), both discovered by Hieber and co-workers.
  62. [62]
    [PDF] Metal Hydride Complexes
    Early transition metal hydrides tend to carry significant negative charge on the H atom whereas later more electronegative transition metals favour a more ...<|control11|><|separator|>
  63. [63]
  64. [64]
  65. [65]
    Early transition metal hydride complexes: Synthesis and reactivity
    Aug 5, 2025 · This review describes the synthesis and reactivity studies of early metal (Sc, Y, Ti, Zr, Hf, V, Nb, Ta) hydride compounds that have ...Missing: acidity | Show results with:acidity
  66. [66]
    Estimating the Acidity of Transition Metal Hydride and Dihydrogen ...
    Jan 10, 2014 · It involves adding acidity constants AL for each of the ligands in the 5-, 6-, 7-, or 8-coordinate conjugate base complex of the hydride or ...
  67. [67]
    Expanding frontiers in materials chemistry and physics with multiple ...
    Feb 22, 2018 · Oxyhydrides (oxide hydrides), containing oxide and negatively charged hydride (H–) anions, are rare but can be remarkable materials. Several ...
  68. [68]
    New chemistry of transition metal oxyhydrides - PMC - NIH
    2. Synthesis of oxyhydrides. The synthesis of oxyhydrides requires special care, given the unusual properties of the hydride anion. Preparation typically ...
  69. [69]
    Configuration and Dynamics of Hydride Ions in the Nitride-Hydride ...
    Dec 11, 2024 · We report results on the configuration and vibrational dynamics of hydride ions (H – ) in the novel mixed-anion, nitride-hydride, catalyst Ca 3 CrN 3 H
  70. [70]
    [PDF] Hydride-ion transport mechanisms in oxyhydrides and nitride ...
    Another example is the synthesis of the nitride–hydride BaTiO3–x N2x/3 ... mixed-anion compounds, which can give rise to emergent macroscopic properties.
  71. [71]
  72. [72]
    Exploring Multi-Anion Chemistry in Yttrium Oxyhydrides: Solid-State ...
    Jul 17, 2023 · (7) Metal hydrides are well-known for their hydrogen storage properties, (8,9) whereas rare earth oxyhydrides are utilized as a photocatalyst ...
  73. [73]
    Lithium hydride | 7580-67-8 - ChemicalBook
    Aug 19, 2025 · Lithium hydride (HLi) has a melting point of 680°C, density of 0.82 g/mL, is white to gray, and is air & moisture sensitive. It is a danger ...
  74. [74]
    Sodium Hydride | AMERICAN ELEMENTS ®
    NaH. MDL Number. MFCD00003471. EC No.: 231-587-3. Email SDS. ×. Email address ... Density, 1.396 g/cm3. Solubility in H2O, Reacts. Exact Mass, 23.997595.
  75. [75]
    Methane - Wikipedia
    Density. 0.657 kg/m3 (gas, 25 °C, 1 atm); 0.717 kg/m3 (gas, 0 °C, 1 atm) · 422.8 g/L (liquid, −162 °C) ; Melting point, −182.456 °C (−296.421 °F; 90.694 K).Methane clathrate · Atmospheric methane · Aerobic methane production
  76. [76]
    Ammonia | NH3 | CID 222 - PubChem - NIH
    While this substance does not burn or ignite readily (autoignition temp: 1204 °F), containers of ammonia may explode in the heat of a fire. For small fires ...
  77. [77]
    LiH properties
    The compound melts at 688.7 °C without decomposition under hydrogen atmosphere. Thermal decomposition occurs between 900–1000 °C, producing lithium metal and ...
  78. [78]
    Molecular dynamics studies of fundamental bulk properties of ...
    Jun 14, 2018 · In addition, the room-temperature palladium thermal expansion coefficient determined in Eq. (1), α ≈ 14 × 10−6 K−1, is quite close to the ...
  79. [79]
    3.10: Some Properties of Ionic Compounds - Chemistry LibreTexts
    Aug 8, 2022 · Ionic compounds have high melting points. Ionic compounds are hard and brittle. Ionic compounds dissociate into ions when dissolved in water.Missing: hydrides | Show results with:hydrides
  80. [80]
    Hydride | Properties, Reactions & Uses - Britannica
    Oct 1, 2025 · Hydride, any of a class of chemical compounds in which hydrogen is combined with another element. Three basic types of hydrides—saline ...
  81. [81]
    Shock compression experiments on Lithium Deuteride (LiD) single ...
    Dec 21, 2016 · Since the LiD sample is transparent in the visible spectrum, the 532 nm laser light could pass through the target holder and reflect off the ...
  82. [82]
    Volume changes during hydrogen absorption in metals - IOPscience
    The semi-empirical model of Griessen and Driessen (1984) which relates the heat of formation Delta H of a metal hydride to the difference of the Fermi energy EF ...
  83. [83]
    [PDF] The Hydrides of Palladium
    The original volume of a palladium speci- men expands by about 10 per cent when. H/Pd attains a value of about 0.5 but, in con- trast to the other transition ...
  84. [84]
    Transition metal hydride - Wikipedia
    History. An ill-defined copper hydride had been described in the 1844 as resulting from treatment of copper salts with hypophosphorous acid. It was ...
  85. [85]
    First-Principles Calculation of 1H NMR Chemical Shifts of Complex ...
    Nov 23, 2020 · A new strategy by computational 1 H NMR spectroscopy was used to predict and support the characterization of complex metal polyhydrides.Introduction · Results and Discussion · Conclusions · Computational Methods
  86. [86]
    Calcium hydride
    ### Summary of Hydrolysis with Water for Calcium Hydride
  87. [87]
    Diborane
    ### Summary of Diborane Hydrolysis Reaction with Water
  88. [88]
    Lithium aluminum hydride | AlH4.Li | CID 28112 - PubChem
    IDENTIFICATION AND USE: Lithium aluminum hydride is a gray-white monoclinic crystalline material which is soluble in ether, tetrahydrofuran, dimethylcellosolve; ...
  89. [89]
    Recent progress in thermodynamic and kinetics modification of ...
    Studies have demonstrated that magnesium hydride initiates dehydrogenation at a minimum of 287 °C under a pressure of 1 atm, and achieving complete ...<|separator|>
  90. [90]
    Photochemistry of Transition Metal Hydrides | Chemical Reviews
    Jul 6, 2016 · Dihydrogen complexes typically lose H2 photochemically. The majority of photochemical reactions are likely to be dissociative, but hydride ...PHOTOCHEMICAL... · SPECIALIST... · Metal Hydride Photochemistry...
  91. [91]
    Low-temperature regeneration of metal hydride after impurity gas ...
    Aug 5, 2025 · In this study, the effects of gaseous impurities like CO2, CH4, and O2 on the hydrogen cycling behavior of commercial Ti-Fe-Mn were carefully ...
  92. [92]
    The bond dissociation energy of H2 , D2 and T2 follow the order (D ...
    Oct 14, 2022 · Bond dissociation energy of T 2 is 446.9 kJ mol -1 D 2 is 443.4 kJ mol -1 and H 2 is 435.9 kJ mol -1 Therefore the answer is b.
  93. [93]
    124-125. Lithium Hydride, LiH, and Lithium Deuteride, LiD
    M. ithium hydride and lithium deuteride are prepared by heating lithium metal in an atmosphere of hydrogen or deuterium.
  94. [94]
  95. [95]
    Deuterium Oxide | H2O | CID 24602 - PubChem - NIH
    3.2.2 Color / Form. Colorless liquid ; 3.2.3 Odor. Odorless ; 3.2.4 Boiling Point. 101.42 °C ; 3.2.5 Melting Point. 3.81 °C; triple point 3.82 °C ; 3.2.6 Density.
  96. [96]
    Hydrides as neutron moderator and reflector materials - ScienceDirect
    Hydrides as neutron moderator and reflector materials☆. Author links ... Lithium hydride (LiH) and lithium deuteride (LiD). Anal. Chem., 28 (1956), p ...
  97. [97]
    [PDF] 5-reduction.pdf - Harvard
    Sodium Borohydride: NaBH4. • Sodium borohydride reduces aldehydes and ketones to the corresponding alcohols at or below. 25 °C. Under these conditions, esters ...
  98. [98]
    [PDF] USING HYDROGEN AS A NUCLEOPHILE IN HYDRIDE ...
    The mechanism of the hydride attack on a carbonyl carbon shown below demonstrates how these reagents in general work. The hydride-delivering agent must approach ...
  99. [99]
    [PDF] Sodium Borohydride Reduction of Benzoin Introduction
    The carbonyl bond is reduced by the formal addition of H2 across the C=O π-bond (Figure 1). Note the hydrogen atom from the reducing agent (red) is attached to ...
  100. [100]
    Ch20: Reduction of Esters using LiAlH4 to 1o alcohols
    Carboxylic esters are reduced give 2 alcohols, one from the alcohol portion of the ester and a 1o alcohol from the reduction of the carboxylate portion. · Esters ...
  101. [101]
    Reduction of Organic Compounds by Lithium Aluminum Hydride. I ...
    A safety guidance document for lithium aluminum hydride (LAH) reduction: a resource for developing specific SOPs on LAH manipulations.<|separator|>
  102. [102]
    Esters can be reduced to 1° alcohols using \(LiAlH_4\) - Chemistry
    Jan 22, 2023 · Esters can be converted to 1o alcohols using LiAlH4, while sodium borohydride ( N ⁢ a ⁢ B ⁢ H 4 ) is not a strong enough reducing agent to ...
  103. [103]
    3.3: Enolate Addition- Aldol reactions - Chemistry LibreTexts
    Oct 4, 2022 · Some examples of compounds used as very strong bases are sodium hydride (NaH) ... sodium hydride (NaH), would result in complete enolate formation.
  104. [104]
    enolate ions - csbsju
    A very strong base, like sodium amide (NaNH2), butyllithium (CH3CH2CH2CH2Li, or BuLi) or sodium hydride (NaH), would result in complete enolate formation.<|separator|>
  105. [105]
    Ring-opening polymerization of monosubstituted oxiranes in the ...
    Nov 20, 2019 · Potassium hydride is active initiator of anionic polymerization of monosubstituted oxiranes, such as propylene oxide, allyl glycidyl ether and ...
  106. [106]
    A Study of the Polymerization of Propylene Oxide Catalyzed by ...
    Ring-opening polymerization of monosubstituted oxiranes in the presence of potassium hydride: determination of initiation course and structure of ...
  107. [107]
    Stepwise Hydride Transfer in a Biological System: Insights into ... - NIH
    Hydride transfer plays a crucial role in a wide range of biological systems. However, its mode of action (concerted or stepwise) is still under debate.
  108. [108]
    LiAlH4 and NaBH4 Carbonyl Reduction Mechanism - Chemistry Steps
    The hydride ion reacts with the carbonyl group, which, in turn, is also a polar covalent bond, and the presence of the π bond makes the addition of H– possible:.
  109. [109]
    Sodium Borohydride (NaBH4) As A Reagent In Organic Chemistry
    Aug 12, 2011 · Using sodium borohydride for the reduction of unsaturated ketones sometimes results in the side reaction of conjugate reduction, i.e. the ...
  110. [110]
    [PDF] Henri Moissan - Revistas UNAM
    Moissan prepared the hydrides of potassium, sodium, cesium, and rubidium and showed that they did not conduct electricity and that the hydrogen was supposed to ...
  111. [111]
    Manufacture, Handling, and Uses of Sodium Hydride
    Jul 22, 2009 · The principal use of sodium hydride is to carry out condensation and alkylation reactions which proceed through the formation of a carbanion (base-catalyzed).
  112. [112]
    Hydrogen Storage | Department of Energy
    Specific system targets include the following: 1.5 kWh/kg system (4.5 wt.% hydrogen); 1.0 kWh/L system (0.030 kg hydrogen/L); $10/kWh ($333/kg stored hydrogen ...
  113. [113]
    Hydrogen Storage Performance of Mg/MgH2 and Its Improvement ...
    Feb 14, 2023 · The reversible hydrogen storage capacity of Mg/MgH2 can reach approximately 7.6 wt.%, which satisfies DOE's regulations [1,5,6,34]. However, the ...
  114. [114]
    Improved Hydrogen Storage Kinetic Properties of MgH2 with NiO ...
    Dec 3, 2024 · In terms of hydrogen storage capacity, it was found that MgH2 + NC exhibited hydrogen absorption capacities of 2.76, 5.89, and 6.24 wt % H2 ...
  115. [115]
    Sodium Alanates | Hydrogen Link
    Alkali alanates, in particular sodium alanate NaAlH4, combine high, reversible hydrogen capacity (up to 5.6 weight %) with low cost and commonly available ...
  116. [116]
    [PDF] Effect of Particle Size on Hydrogen Release from Sodium Alanate ...
    One such ma- terial, sodium alanate (NaAlH4), reversibly stores 5.6% hydrogen by weight in a two- step reaction: At 1 atm, the equilibrium temperatures for ...
  117. [117]
    DOE Technical Targets for Onboard Hydrogen Storage for Light ...
    This table summarizes technical performance targets for hydrogen storage systems onboard light-duty vehicles.
  118. [118]
    Regulation of Kinetic Properties of Chemical Hydrogen Absorption ...
    In the cycle-stability test of the composite system, the hydrogen storage capacity of MgH2 can still reach more than 92% after the end of the 10th cycle, and ...<|separator|>
  119. [119]
    Hydrogen Storage Technology, and Its Challenges: A Review - MDPI
    However, they face challenges like operating at high temperatures and pressures, slow reaction rates, and material stability issues. Researchers are ...
  120. [120]
    LiBH4 as a Solid-State Electrolyte for Li and Li-Ion Batteries: A Review
    May 12, 2023 · In this paper, the methods used to enhance the conductivity of LiBH 4 , a potential electrolyte for the construction of solid-state batteries, are summarized.
  121. [121]
    Solid-state lithium-ion battery employing LiBH4–ZrO2 as a solid ...
    May 29, 2025 · LiBH4 has been extensively investigated as a solid-state electrolyte for Li-ion battery applications. The.
  122. [122]
    Recent Development in Nanoconfined Hydrides for Energy Storage
    Innovative methods to enhance the properties of solid hydrogen storage materials based on hydrides through nanoconfinement: A review. J. Supercrit. Fluids ...
  123. [123]
    A critical review of hydrogen storage: toward the nanoconfinement of ...
    Oct 1, 2024 · 5, metal hydrides can generally be classified into two major groups: simple metal hydrides and complex metal hydrides. Simple metal hydrides are ...
  124. [124]
    Recent advances in the nanoconfinement of Mg-related hydrogen ...
    Aug 6, 2025 · In this paper, we will review the recent advances in the nanoconfinement of Mg-related hydrogen storage materials by loading Mg particles on ...
  125. [125]
    Recent challenges and development of technical and ...
    Jun 12, 2024 · In this comprehensive review paper, we have undertaken the task of categorising and evaluating various hydrogen storage technologies across three different ...
  126. [126]
    Metal Hydride Storage | H2tools | Hydrogen Tools
    Metals that form hydrides can reversibly absorb and release hydrogen when appropriate temperatures and pressures are applied.<|control11|><|separator|>
  127. [127]
    Observation of New Intermediates in Hydrogenation Catalyzed by ...
    A novel Wilkinson's type silica supported polymer catalyst: Insights from solid-state NMR and hyperpolarization techniques.
  128. [128]
    [PDF] Conventional Catalytic cycle for hydrogenation with Wilkinson's ...
    The first step of this catalytic cycle is the cleavage of a PPh3 to generate the active form of the catalyst followed by oxidative addition of dihydrogen. Page ...
  129. [129]
    [PDF] triethylsilane - MSU Chemistry
    Radical Chain Reductions. Triethylsilane can replace toxic and difficult to remove organotin reagents for synthetic reductions under radical chain conditions.
  130. [130]
    Radical Reduction of Aromatic Azides to Amines with Triethylsilane
    Under radical conditions, tributyltin hydride performs effective reduction of both aromatic and aliphatic azides, but the reductions of the aromatic azides are ...Experimental Section · Supporting Information Available · References
  131. [131]
    Essential role of hydride ion in ruthenium-based ammonia synthesis ...
    Ruthenium-loaded metal hydrides with hydrogen vacancies function as efficient catalysts for ammonia synthesis under low temperature and low pressure ...
  132. [132]
    [PDF] William S. Knowles - Nobel Lecture
    The key to asymmetric hydrogenation is the structure of the chiral ligand. The phosphines are prepared by a multi-step route and are quite expensive, but ...
  133. [133]
    [PDF] Nomenclature of Inorganic Chemistry | IUPAC
    Chemical nomenclature must evolve to reflect the needs of the community that makes use of it. In particular, nomenclature must be created to describe new ...