Fact-checked by Grok 2 weeks ago

Titanium hydride

Titanium hydride is a inorganic compound composed of and , with the TiH₂, typically existing as a non-stoichiometric δ-phase (TiHx, where 1.45 < x < 1.99) that appears as a stable grey-black powder. It crystallizes in a cubic ( Fm&bar;3m), where Ti²⁺ ions are coordinated to eight H⁻ ions at a of 1.91 , resulting in a of approximately 3.87 g/cm³. The compound is insoluble in and decomposes at temperatures above 375–400°C, releasing gas without . Titanium hydride forms through the reaction of metal with gas at elevated temperatures (typically 480–650°C) and pressures, often as part of the Kroll process where the resulting brittle material is crushed into powder; it can also be produced electrochemically or via . Upon exposure to air, a thin passivating layer (3–50 nm) develops on its surface, which influences release kinetics and can dissolve into an oxyhydride phase at 400–500°C. Mechanically, the δ-hydride phase exhibits low (approximately 2.2 MPa·m¹/²) and a Young's modulus of about 125 GPa, contributing to in . diffusion is slower in the hydride phase (~4 × 10⁻¹² cm²/s) compared to the matrix (~2.6 × 10⁻¹⁰ cm²/s), affecting phase stability and formation at grain boundaries. Titanium hydride finds diverse applications due to its capacity (up to 4 wt% reversible ) and catalytic properties, serving as an additive in to reduce oxide impurities and as a for producing metal foams. It enhances sorption in complex metal hydrides like MgH₂ and NaAlH₄, and acts as a catalyst in synthesis owing to hydrogen-deficient surfaces with high vacancy activity. In and additive manufacturing, its controlled dehydrogenation provides precise performance tuning and oxide reduction, while its stability makes it suitable for structural applications in and biomedical implants, though hydride formation must be managed to mitigate embrittlement.

Chemical Identity and Properties

Composition and Stoichiometry

Titanium hydride is represented by the general formula \ce{TiH_x}, where x ranges from 0 to approximately 2, reflecting its non-stoichiometric nature across various phases in the Ti-H system. This variability arises from hydrogen atoms occupying interstitial tetrahedral sites within the titanium lattice, allowing partial occupancy that leads to compositions from nearly pure titanium (x \approx 0) to near \ce{TiH_2} (x \approx 2). Common forms include the delta phase, denoted as \ce{TiH_{1.5-1.75}}, which is widely studied for its stability and applications, followed by the epsilon phase \ce{TiH_{1.75-2.0}}. Historically, titanium hydride was often denoted simply as \ce{TiH_2}, assuming a stoichiometric dihydride, but phase diagram assessments revealed the need for the more precise \ce{TiH_x} notation to account for compositional ranges and phase equilibria. The Ti-H phase diagram at delineates key phases based on hydrogen-to-titanium ratios: the alpha phase (\alpha, hexagonal close-packed structure with low hydrogen content, x < 0.3); the beta phase (\beta, body-centered cubic with interstitial , up to x \approx 1.5 at elevated temperatures such as 800°C); the gamma phase (\gamma, face-centered tetragonal, x \approx 0.8-1.0); the delta phase (\delta, face-centered cubic, x \approx 1.5-1.75); and the epsilon phase (\varepsilon, face-centered tetragonal, x \approx 1.75-2.0). These boundaries, established through thermodynamic assessments, highlight the progression from solid solution phases to hydride precipitates as concentration increases. The non-stoichiometric compositions influence the thermodynamics of hydride formation, particularly through the partial molar free energy of hydrogen, which determines phase stability and the characteristic plateaus in pressure-composition isotherms. For instance, in the delta phase, deviations from stoichiometry result in lower hydrogen occupancy, affecting the enthalpy and entropy of absorption and leading to two-phase coexistence regions with distinct equilibrium pressures. This variability enables reversible hydrogen storage but also complicates precise control in synthesis, as phase boundaries shift with temperature and pressure.

Physical Properties

Titanium hydride appears as a brittle, gray-black metallic powder or solid, with fine powders exhibiting pyrophoric behavior upon exposure to air. The density of the delta phase ranges from 3.75 to 3.91 g/cm³, depending on the hydrogen stoichiometry. Titanium hydride decomposes at approximately 400–500 °C without reaching a melting point, rendering a boiling point inapplicable. The material demonstrates increased hardness due to hydride formation, with Vickers hardness values of 330–400 HV, substantially higher than pure titanium (~150-200 HV). Its thermal conductivity is low, around 12 W/m·K at 400–500 °C for fine powders (-325 mesh), and the specific heat capacity is approximately 0.5 J/g·K at room temperature. Titanium hydride is insoluble in water but reacts with acids.

Crystal Structure

Titanium hydride exhibits several distinct crystal phases depending on concentration, temperature, and pressure, each characterized by specific atomic arrangements that influence hydrogen occupancy and overall material properties. The alpha (α) phase corresponds to low hydrogen content (H/Ti < 0.1), where hydrogen atoms dissolve interstitially in the hexagonal close-packed (hcp) lattice of titanium without forming a separate hydride structure. In this phase, hydrogen primarily occupies tetrahedral sites, with minor occupancy in octahedral sites, leading to slight lattice expansion. The lattice parameters are approximately a = 0.295 nm and c = 0.468 nm, closely matching those of pure α-Ti but with minor distortions due to hydrogen incorporation. At intermediate hydrogen concentrations and elevated temperatures, the beta (β) phase emerges, featuring a body-centered cubic (bcc) structure with hydrogen atoms distributed as interstitials in the lattice, accommodating up to H/Ti ≈ 1.5 at ~800°C. This phase accommodates higher hydrogen solubility compared to α-Ti, with lattice parameter a ≈ 0.330 nm, reflecting the expanded bcc framework of β-Ti modified by hydride formation. The γ phase, often metastable and observed at H/Ti ≈ 0.8–1.0, adopts a face-centered tetragonal (fct) structure resulting from a tetragonal distortion of the fcc lattice, with lattice parameters a ≈ 0.42 nm and c ≈ 0.47 nm (c/a > 1). Hydrogen occupies ordered tetrahedral sites in this distorted arrangement, contributing to its needle-like morphology in micrographs. The delta (δ) phase, stable at higher hydrogen contents (H/Ti ≈ 1.5–1.75), features a face-centered cubic (fcc) (CaF₂) structure ( Fm¯3m), where atoms form a fcc sublattice and occupies tetrahedral interstices, often with vacancies; the epsilon (ε) extends this to H/Ti ≈ 1.75–2.0 with face-centered tetragonal distortion. For near-stoichiometric TiH₂ (ε ), the lattice parameters show slight tetragonal elongation (a ≈ 0.445 nm, c ≈ 0.444 nm). diffraction () patterns of these s reveal characteristic peaks: the α shows hcp reflections (e.g., (002) at ~40° 2θ), β exhibits bcc peaks (e.g., () at ~39° 2θ), γ displays split fct lines, and δ presents fcc-like patterns with broadening from vacancies (e.g., (111) at ~36° 2θ). transitions, such as α to δ, occur around 150°C under sufficient pressure, as evidenced by in situ showing peak shifts and new reflections during hydriding. Defect structures in the δ primarily involve hydrogen vacancies, resulting in compositions TiH_{2-x} (0 < x < 0.5), which disrupt long-range ordering and lead to local lattice relaxations observable in XRD line broadening. These vacancies enhance diffusivity but can induce strain that affects phase stability, with ordering transitions suppressed at higher vacancy concentrations. Stoichiometry ranges, such as H/Ti > 1.5 for δ-phase dominance, further define these structural features.

Synthesis and Production

Direct Hydriding Methods

Direct hydriding involves the reaction of metal with gas to form titanium hydride, represented as \ce{Ti + \frac{x}{2} H2 -> TiH_x}, where x typically ranges from 1 to 2, and the process is exothermic with a \Delta H_f \approx -130 kJ/mol for TiH_2 at 323 K. This method relies on exposing metal or powder to a atmosphere, heating it to 480–650°C under pressures of 0.1–10 to facilitate absorption and hydride formation. An initial incubation period precedes , during which diffuses into the lattice before precipitates form, often lasting several minutes depending on conditions. As hydrogen exposure increases, the process progresses through distinct phases: starting from the alpha phase (α-Ti), with hydrogen absorption leading to β-phase (β-Ti) stabilization due to high diffusivity in the body-centered cubic structure, and further precipitation of the delta (δ) hydride phase (TiH_x, 1.5 < x < 2), with details on these crystal structures provided in the Crystal Structure section. Key factors influencing the hydriding rate include particle size, where finer titanium particles enable faster hydriding due to increased surface area and nucleation sites; material purity, as oxygen or iron impurities can hinder hydrogen diffusion and adsorption; and precise temperature control to prevent premature embrittlement from uneven hydride distribution. The direct hydriding method was first reported in the early 1950s through low-pressure studies of the titanium-hydrogen system, with subsequent development driven by nuclear applications requiring stable hydride materials for moderation and storage. It has since become the standard industrial approach for producing commercial TiH_2 powder, achieving hydride conversion yields up to 99% and hydrogen contents of 3.8–4.0 wt%.

Alternative Production Routes

One prominent alternative route to direct hydriding involves the magnesiothermic reduction of titanium tetrachloride (TiCl₄) with magnesium metal in a hydrogen atmosphere, following the idealized reaction TiCl₄ + 2 Mg + H₂ → TiH₂ + 2 MgCl₂. A 2023 investigation demonstrated this process at temperatures of 1073–1173 K with TiCl₄ feeding rates of 19.16–57.42 g·min⁻¹, producing a mixture of titanium and TiH_{1.5} powder with an oxygen concentration as low as 0.116 mass% after cooling in H₂ gas, enabling high-purity output suitable for powder metallurgy. This method leverages readily available TiCl₄ precursors to bypass high-temperature elemental titanium handling. Titanium hydride can also be synthesized from titanium intermediates via mechanical or thermal processes. Ball milling of titanium powder under a hydrogen atmosphere at room temperature rapidly forms TiH₂ with a body-centered tetragonal crystal structure, as shown in foundational experiments where hydrogenation occurred within hours without external heating, yielding fine hydride particles for subsequent applications. Thermal decomposition of titanium alkoxides, such as titanium isopropoxide, in a hydrogen environment has been reported to generate hydride phases, though this approach remains less optimized and typically requires controlled heating to facilitate hydrogen incorporation during precursor breakdown. Additionally, reactions of TiCl₄ with magnesium hydride (MgH₂) in H₂ at relatively low temperatures (around 500 °C) produce TiH₂ powder after purification steps like washing, offering thermodynamic favorability over traditional reductions. Electrolytic routes employ molten salt electrodeposition to deposit titanium, with hydrogen incorporation occurring either during electrolysis or in a follow-up hydriding step. In NaCl-KCl-based molten salts, titanium is electrodeposited from Ti(III) ions at potentials around -1.5 V vs. a reference electrode, and subsequent exposure to hydrogen at elevated temperatures (e.g., 400–600 °C) forms , achieving hydride contents up to with controlled particle morphology. This method integrates reduction and potential in-situ hydrogenation, minimizing oxygen contamination from aqueous media. Recent innovations from 2022 to 2025 emphasize enhanced precursor-based and assisted processes for improved scalability. For instance, the 2023 magnesiothermic approach highlights direct TiH₂ powder generation from TiCl₄, reducing steps in titanium powder production chains. Plasma-assisted techniques have emerged for hydriding titanium alloys, where low-pressure plasma enhances hydrogen diffusion into surfaces at lower temperatures (below 600 °C), forming uniform hydride layers in alloys like Ti-6Al-4V for advanced material processing, though full-scale adoption remains exploratory. In 2025, self-propagating high-temperature synthesis (SHS) hydrogenation was applied to Ti-6Al-4V chips to produce fine TiH₂ powder, enabling recycling of machining waste into usable material. These methods offer key advantages over direct hydriding of elemental titanium, including operation at lower effective temperatures (often below 1000 K), production of finer particle sizes in the 1–10 μm range for better sinterability, and diminished risk of hydrogen embrittlement during synthesis due to controlled precursor reactions. However, challenges persist in impurity management, such as chlorine residues from TiCl₄-based processes that require additional leaching, and achieving industrial scalability, where energy efficiency and byproduct separation limit widespread implementation.

Chemical Reactivity and Stability

Reactions with Other Substances

Titanium hydride reacts with water, potentially violently or explosively, producing titanium oxide and hydrogen gas (e.g., TiH₂ + 2 H₂O → TiO₂ + 3 H₂), especially in fine powder form; the process accelerates significantly with catalysts or elevated temperatures. This hydrolysis accelerates significantly in the presence of acids, where the compound dissolves while liberating hydrogen. For instance, in hydrochloric acid, the reaction produces Ti³⁺ salts and hydrogen, approximately as TiH₂ + 3 HCl → TiCl₃ + 2.5 H₂ (with partial exchange in deuterated media), vigorous in concentrated solutions (e.g., 15 M HCl), completing within hours and yielding a deuterium/hydrogen ratio consistent with partial exchange in deuterated media. Similar behavior occurs with sulfuric acid, where dissolution releases hydrogen and forms soluble titanium salts, though rates depend on acid concentration and temperature. Titanium hydride reacts exothermically with halogens, such as fluorine, forming titanium halides and hydrogen halides, as exemplified by TiH₂ + 2 F₂ → TiF₄ + 2 HF; this reactivity stems from its strong reducing nature and contributes to its utility in getter applications for scavenging reactive gases in vacuum systems. The compound's interaction with oxidizing agents like halogens is rapid and energetic, even with weaker oxidants, underscoring its role in pyrotechnic formulations where controlled combustion is desired. Upon exposure to air, titanium hydride spontaneously develops a passivating surface layer of titanium oxide, typically 3–50 nm thick, which incorporates hydrogen to form non-stoichiometric TiO_{2-x}H_y phases that stabilize the material against further oxidation at room temperature. However, fine powders are pyrophoric in dust form, igniting spontaneously upon contact with air due to rapid oxidation. Bulk material shows pyrophoric ignition above 400°C, where the oxide layer dissolves, leading to oxyhydride formation and accelerated hydrogen release. In interactions with metals and alloys, titanium hydride serves as an effective hydrogen source for doping, particularly in Ti-Mn-based hydrogen storage systems, where it facilitates hydride formation and enhances absorption/desorption kinetics; for example, TiMn_{1.5} alloys achieve near-optimal performance with hydrogen capacities around 1.8 wt% at room temperature, benefiting from such doping to improve activation and reversibility. Recent surface studies from 2021 highlight the dynamics of H-Ti interactions, revealing low barriers for hydrogen adsorption (no dissociation barrier on TiH₂(111) surfaces, exothermic by 2.92 eV) and diffusion (0.03 eV from surface to subsurface), which promote the formation of TiH_{2-x}O_y oxyhydride interfaces under ambient oxidative conditions and underscore titanium hydride's catalytic role in gas-solid reactions.

Thermal Decomposition and Stability

Titanium hydride undergoes thermal decomposition via the endothermic reaction \ce{TiH_x -> Ti + (x/2) H2}, where x typically ranges from 1.5 to 2, releasing gas. In or inert atmospheres, decomposition initiates around 400–450°C, with significant release occurring above 450°C and near-complete conversion to metallic by 500–600°C, though full dehydrogenation may extend to 800°C depending on particle size and conditions. The of this follow a desorption mechanism, characterized by the \frac{d[\ce{H}]}{dt} = -k [\ce{H}]^n where n \approx 1, reflecting the surface-controlled release of . energies for the process vary between 102 and 160 kJ/mol across different phases and conditions, with values around 120–150 kJ/mol commonly reported for the δ-phase . Isothermal studies at 400–660°C reveal sigmoidal transformation curves, where the time for 50% dehydrogenation decreases exponentially with , consistent with Arrhenius . During dehydrogenation, titanium hydride exhibits sequential phase reversals as hydrogen content decreases: the δ-phase (face-centered cubic TiH_{1.5–2}) transitions to a mixed δ + β phase, followed by β (body-centered cubic titanium), and finally to the α-phase (hexagonal close-packed titanium) upon near-complete hydrogen loss. This process is accompanied by hysteresis in the absorption-desorption cycle, where desorption temperatures exceed absorption thresholds due to kinetic barriers and lattice strain. In inert atmospheres, titanium hydride remains stable up to approximately 300°C without significant , but it begins to oxidize in air above 200°C, leading to surface formation and accelerated loss. This limited thermal stability arises from the compound's sensitivity to oxygen, which promotes exothermic reactions during heating. The process enables titanium hydride to serve as a source of high-purity (>99.9% yield) upon heating to 500°C in systems, a method employed in settings since the 1950s for precise dosing and getter applications. Recent modeling efforts, including 2022 studies on hydride stability in , have utilized to predict phase equilibria and barriers under elevated temperatures, aiding the design of materials with enhanced thermal resilience. As of 2025, studies have further detailed phase transformation mechanisms during dehydrogenation, assessed TiH₂ dust ignition and risks for , and demonstrated its use in electrocatalysis for via lattice transfer.

Applications and Uses

Traditional Industrial Applications

Titanium hydride is widely utilized as a in the production of lightweight titanium foams for applications, leveraging its to release gas that expands the material during processing. This results in porous structures with densities typically ranging from 0.5 to 1.5 g/cm³, offering a favorable strength-to-weight ratio suitable for components in and . In , titanium hydride functions as a reactive fuel in inflators and signal flares, where it combines with oxidizers to generate rapid evolution and intense heat for reliable ignition and gas generation. This application has been established since the 1980s, particularly in initiator squibs and igniters for systems and military devices. As a controlled hydrogen source, provides ultra-pure gas for fabrication processes and laboratory experiments, while also serving as a getter material in systems to scavenge residual gases and maintain high integrity. In , it acts as a reductant to facilitate the of by reducing surface oxides on powder particles, enhancing densification and mechanical properties.

Emerging and Advanced Uses

Titanium hydride has garnered attention in applications, particularly within Ti-Mn alloys that enable reversible uptake. These alloys demonstrate capacities of approximately 1.5-2 wt% , with fast absorption and desorption kinetics operable at moderate temperatures of 100-200°C, as evidenced in 2022 studies on their activation and cycling performance. Such properties position Ti-Mn systems as promising for fuel cells, where stoichiometric tuning of the hydride enhances reversibility without compromising overall . In lithium-ion battery development, TiH₂ serves as an anode material in prototypes that leverage its conversion mechanism for elevated capacities. A 2022 review highlights TiH₂-based electrodes achieving up to 300 mAh/g, surpassing traditional anodes in specific scenarios due to hydride decomposition facilitating lithium insertion. This improvement stems from the material's high theoretical volumetric capacity and compatibility with solid electrolytes, though challenges in cycle stability persist in prototype testing. Additive manufacturing exploits powders to fabricate 3D-printed components with tailored . The hydride-dehydride process during and post-treatment allows precise control over pore distribution, yielding structures with porosities up to 40 vol% suitable for lightweight or biomedical scaffolds. evolution from the hydride phase during enhances powder flowability and reduces defects, enabling complex geometries unattainable with pure powders. Post-2010 research has explored titanium hydride as a in nuclear reactors, capitalizing on 's effective thermalization of neutrons. Its high hydrogen density supports compact, high-temperature designs in micro-reactors, where hydrided offers stability under compared to moderators. Patent developments since 2023 further refine hydrided metal moderators, including titanium variants, for enhanced safety in advanced systems. Emerging biomedical applications involve titanium hydride formation for biocompatible coatings on implants, improving through controlled surface hydriding. 2024-2025 studies on hybrid Mg-Ti implants demonstrate that induces hydride layers that enhance resistance and bioactivity without toxicity, promoting in orthopedic devices. The for titanium hydride, particularly in forms, is projected to grow at a 5-7% CAGR through 2030, fueled by demand in green energy sectors like and batteries. As of 2025, the size is estimated at USD 1.0-1.2 billion, with 2025 reports underscoring this expansion, attributing it to innovations addressing storage needs.

Impacts on Titanium Materials

Hydrogen Embrittlement Mechanisms

Hydrogen diffuses into the primarily through sites, with faster in the β phase compared to the α phase due to its body-centered cubic structure providing more available positions. Upon reaching , typically exceeding the solubility limit, hydrogen precipitates as brittle hydrides (TiH_x, where x ≈ 1.5–2.0), forming plate-like or needle-shaped structures known as hydride platelets. These platelets are inherently brittle and serve as stress concentrators, initiating microcracks that propagate via along the hydride-matrix interface or through the hydride itself, leading to delayed cracking. The threshold for such cracking (K_H) in is approximately 10–20 MPa·m^{1/2}, below which stable crack growth is suppressed. The embrittlement process unfolds in distinct stages: initial hydrogen absorption and diffusion into the lattice, followed by hydride precipitation at sites of high stress or defects; subsequent stress concentration around the growing hydrides induces localized plasticity or decohesion; and finally, fracture initiation and propagation in a brittle mode, contrasting with the ductile failure typical of unhydrogenated titanium. This sequence contrasts with ductile titanium behavior, where deformation is accommodated by twinning and slip without brittle phase formation. Several factors influence the severity of in . Lower temperatures, particularly below 100°C, exacerbate the issue by reducing in the α and promoting , while higher temperatures facilitate but may delay . Critical concentrations above 30 initiate formation, with embrittlement becoming pronounced at levels exceeding 100–200 , leading to significant loss. Alloying elements and microstructure play key roles, with α- dominant alloys (e.g., commercially pure ) being more susceptible than β-stabilized alloys due to the α 's lower tolerance and propensity for . Hydrogen embrittlement in titanium arises from two primary types: internal hydrogen absorbed from environmental sources during processing or service, and residual hydrogen present from manufacturing, both contributing to hydride formation. Hydrides typically orient along the basal planes of the α-titanium hexagonal close-packed lattice, aligning with {10\overline{1}0} or {10\overline{1}7} habit planes, which directs crack paths and amplifies anisotropy in embrittlement. Recent modeling efforts, particularly in , have advanced understanding through finite element simulations coupled with phase-field methods to predict hydride growth and decohesion. These models integrate , mechanical stress, and microstructural features like grain boundaries, revealing how elastic strains drive preferential and of hydrides at stress concentrations. In the , undetected hydriding contributed to several failures in components, such as blades and structural parts in high-performance , highlighting the need for stringent control in processing and service environments.

Mitigation and Material Effects

To mitigate in materials, is a primary prevention strategy, involving heating the in a to 600–700°C for several hours to desorb dissolved and prevent formation. This process can reduce levels to the low tens of parts per million by weight (ppmw) in α+β , effectively restoring and when performed as part of stress-relief annealing. Alloying with elements such as () or () enhances by increasing solubility in the lattice without promoting brittle precipitation; for instance, Nb additions stabilize the β-phase, allowing higher interstitial retention while minimizing embrittlement. Pd alloying similarly creates trapping sites that facilitate and release without phase instability, as observed in Ti-Pd systems designed for controlled environments. Detection of hydrogen content is crucial for quality control, with non-destructive methods like ultrasonic testing identifying hydride platelets through acoustic impedance changes, and permeation techniques measuring diffusion rates to quantify absorbed hydrogen. Safe hydrogen thresholds vary by application, but levels below 10 ppm are targeted for ultra-high-performance components to avoid any risk of hydride nucleation, while general aerospace limits are set at 150 ppm maximum per ASTM B348 standards. Beyond embrittlement, titanium hydride formation induces significant material hardening, with yield strength increases of 20–50% due to the higher modulus and stress-bearing capacity of hydride phases (e.g., δ-hydride yielding up to 680 MPa versus ~400–500 MPa for unhydrided α-titanium). However, this comes at the cost of severe ductility loss, often reducing elongation to below 5% as hydrides act as brittle inclusions that promote early fracture under tensile loads. In titanium alloys, β-phase dominant structures (body-centered cubic lattice) exhibit reduced susceptibility compared to α-phase, as the open bcc arrangement accommodates more dissolved hydrogen interstitially without rapid hydride precipitation. Recent advancements, such as 2023 developments in TiN coatings applied via magnetron sputtering, further mitigate effects by enhancing corrosion resistance and blocking hydrogen ingress in aggressive environments like seawater or acidic media. Industry standards like ASTM B348 enforce strict hydrogen control (maximum 150 ) for aerospace parts to ensure structural integrity under cyclic loading. Hydride-induced alterations broadly impact , with hydrided components experiencing 50–80% reductions in cyclic life due to accelerated crack initiation at hydride-matrix interfaces, as evidenced in near-β alloys where cycles drop from over 17,000 to as low as 6,600 at elevated levels.

References

  1. [1]
    [PDF] Understanding the TiH(2-x)/TiOy System at Elevated Temperature
    Titanium hydride is thus used to reduce oxide content in powder metallurgy and additive manufacturing applications, as a blowing agent to make metal foams, and ...
  2. [2]
    TITANIUM HYDRIDE CAS#: 7704-98-5 - ChemicalBook
    Melting point, >400 °C (dec.) ; Density, 3.91 g/mL at 25 °C (lit.) ; bulk density, 1400kg/m3 ; Flash point, >400°C ; storage temp. no restrictions.
  3. [3]
    mp-24161: TiH2 (Cubic, Fm-3m, 225) - Materials Project
    TiH₂ is Fluorite structured and crystallizes in the cubic Fm̅3m space group. Ti²⁺ is bonded in a body-centered cubic geometry to eight equivalent H¹⁻ atoms.
  4. [4]
    WebElements Periodic Table » Titanium » titanium dihydride
    Titanium dihydride ; Class: hydride ; Colour: grey-black ; Appearance: solid ; Melting point: 450°C (decomposes) ; Density: 3750 kg m ...Missing: physical | Show results with:physical
  5. [5]
    Hydriding of titanium: Recent trends and perspectives in advanced ...
    This perspective article introduces the key unanswered questions and challenges in understanding hydride formation in Ti alloys.
  6. [6]
    Titanium Hydride Nanoplates Enable 5 wt% of Reversible Hydrogen ...
    Titanium hydride, TiH2, with Ti being already in low valent state and containing hydrogen itself, is expected to be a better candidate of dopant in comparison ...
  7. [7]
    Surface Properties of the Hydrogen–Titanium System - PMC - NIH
    Titanium is an excellent getter material, catalyzes gas–solid reactions such as hydrogen absorption in lightweight metal hydrides and complex metal hydrides.
  8. [8]
  9. [9]
  10. [10]
  11. [11]
    Titanium Hydride (TiH₂) - Reade Advanced Materials
    Titanium Hydride, TiH₂, is a brittle, metallic-gray solid that is stable at room temperature and inert to water and most chemical reagents.
  12. [12]
    TITANIUM HYDRIDE - CAMEO Chemicals - NOAA
    A white to light colored solid. Insoluble in water and denser than water. Contact may irritate skin, eyes and mucous membranes. May be toxic by ingestion.
  13. [13]
    Titanium hydride (TiH2) - ChemBK
    Titanium hydride (TiH2) - Physico-chemical Properties ; Molar Mass, 47.87 ; Density, 3.91 g/mL at 25 °C (lit.) ; Melting Point, >400 °C (dec.) ; Flash Point, >400°C.
  14. [14]
    Titanium Hydride TiH2 Powder, 5um, 99.5%
    Titanium hydride TiH2 Melting point: 350 °C (662 °F; 623 K) approximately. Titanium hydride TiH2 Molar mass: 49.88 g/mol TiH2. Titanium hydride TiH2 Purity: ...
  15. [15]
    Mechanical properties of titanium hydride - ResearchGate
    Aug 7, 2025 · The Vickers hardness study has already been shown that the δ-titanium hydride is approximately 30% harder than that of the single-phase (α) Ti.Missing: conductivity | Show results with:conductivity
  16. [16]
    Relative density, porosity, and Vickers hardness of the TiH2 samples...
    Beta Titanium Alloys Produced from Titanium Hydride: Effect of Alloying Elements on Titanium Hydride Decomposition ... values between 330–400 Hv. Maximum ...
  17. [17]
    Thermal properties of titanium hydrides | Request PDF - ResearchGate
    Aug 7, 2025 · Interestingly, the TiH2 (−325 mesh) sample had a thermal conductivity of about 12 W/(mK) at 400-500 °C, and the plot is represented in Figure 6.
  18. [18]
    titanium dihydride - the NIST WebBook
    Solid Phase Heat Capacity (Shomate Equation) ; 298. to 1000. · -18.62650 · 190.9860 · -142.4470 · 38.31300 ; 1000. to 2000. · 117.8740 · -60.43550 · 33.68230 · -6.548070.
  19. [19]
    Titanium hydride | H2Ti | CID 197094 - PubChem - NIH
    Titanium hydride appears as a white to light colored solid. Insoluble in water and denser than water. Contact may irritate skin, eyes and mucous membranes.
  20. [20]
    Titanium Hydride - ESPI Metals
    Solubility in H2O: Soluble. Appearance and Odor ... Incompatibility (Materials to Avoid): Oxidizing agents, acids, halogens, halocarbons, water/moisture.
  21. [21]
    [PDF] Impact of hydrogen on the mechanical behavior under low cycle ...
    Jul 2, 2025 · Hydride Composition Crystal structure Lattice parameters Morphology Volume dilatation γ. TiH. f.c.t. a = 0.42 c = 0.47. Fine needles. 12.5-14% δ.
  22. [22]
    First-principles investigation of metal-hydride phase stability: The Ti ...
    Aug 29, 2007 · Within the hcp and fcc hosts, hydrogen can reside in either octahedral or tetrahedral sites. For each Ti atom in the hcp and fcc structures, ...
  23. [23]
    Phase Transformations α → β → γ → δ → ε in Titanium Hydride TiH ...
    Aug 10, 2025 · It is known that hydrogenation of commercially pure titanium can reduce the temperature of α → β polymorphic transformation [32] . Further ...
  24. [24]
    Effect of temperature and hydrogen concentration on the lattice ...
    In pure titanium, the lattice parameter a = 0.331 nm at 900°C and the thermal expansion coefficient κT = 13.8×10-6 °C-1 over the temperature range 900°C to 1000 ...
  25. [25]
    Solute hydrogen and hydride phase implications on the plasticity of ...
    Jun 12, 2017 · As long as Ti-α and Zr-α accept hydrogen in solid solution, it is randomly distributed to energetically favourable tetrahedral sites [8,9].
  26. [26]
    mp-24726: TiH2 (Tetragonal, I4/mmm, 139) - Materials Project
    TiH₂ is Fluorite structured and crystallizes in the tetragonal I4/mmm space group. Ti²⁺ is bonded in a body-centered cubic geometry to eight equivalent H¹⁻ ...
  27. [27]
    Kinetics of thermal decomposition of titanium hydride powder using ...
    Apr 6, 2005 · The delta-to-alpha phase transformation was followed in situ by HTXRD at temperatures varying from room temperature up to 1000 °C. The ...
  28. [28]
    Hydrogen diffusion in titanium dihydrides from first principles
    T i H x . They demonstrated that, at a given temperature, hydrogen diffusion rate is proportional to the concentration of structural vacancies, which equals · ( ...Missing: delta parameters
  29. [29]
    Calorimetrically measured enthalpies for the reaction of H2(D2) (g ...
    ΔH 0f (TiH1.5) = −98.4 kJ and ΔH 0f (TiH2) = −130.3 kJ evaluated at 323 K. No differences in enthalpies were found between H and D.
  30. [30]
    Microstructural evolution and formation mechanism of FCC titanium ...
    Aug 7, 2025 · Microstructural evolution and formation mechanism of FCC titanium hydride in Ti–6Al–4V–xH alloys ... The incubation period exists at the ...
  31. [31]
    The Titanium—Hydrogen System and Titanium Hydride. I. Low ...
    Chemistry of Metal Hydrides as Related to Their Applications in Nuclear Technology. 1968, 119-164. https://doi.org/10.1016/B978-1-4832-3215-7.50009-5.Missing: historical | Show results with:historical
  32. [32]
    (PDF) SHS hydrogenation of titanium: Some structural and kinetic ...
    Aug 6, 2025 · This work investigates experimentally the production of titanium hydride ... pure titanium hydride containing about 4 wt% of hydrogen. Study of ...
  33. [33]
    Production of Titanium Hydride Powder from Titanium Tetrachloride ...
    This study investigated the process of producing Ti hydride powder directly from titanium tetrachloride (TiCl4) using magnesium (Mg) metal and hydrogen gas (H2) ...Missing: TiCl4 H2
  34. [34]
    Formation of titanium hydride at room temperature by ball milling
    Titanium hydride with a body centred tetragonal structure was rapidly synthesized at room temperature by ball milling titanium powders under a hydrogen ...
  35. [35]
    Synthesis of Ti Powder from the Reduction of TiCl 4 with Metal ...
    A novel method was proposed to produce TiH 2 from the reduction of titanium tetrachloride (TiCl 4 ) with magnesium hydride (MgH 2 ) in the hydrogen (H 2 ) ...
  36. [36]
    Titanium Hydride Powder Market Insights and Forecast to 2033
    Sep 19, 2025 · Titanium Hydride Powder Market size is estimated to be USD 150 Million in 2024 and is expected to reach USD 250 Million by 2033 at a CAGR of 6.5 ...Missing: 2030 | Show results with:2030
  37. [37]
    [PDF] AD 403 803 - DTIC
    The reaction of the hydride with hydrochloric acid solution is in no case ... PdHn + nCe(IV) = Pd + nCe. + n/2 H 2 . The re sults obtained with cerium hydride and ...
  38. [38]
    Ti–Mn hydrogen storage alloys: from properties to applications
    Dec 14, 2022 · Ti–Mn hydrogen storage alloys have the characteristics of relatively high hydrogen storage capacity, easy activation, fast hydrogen absorption and desorption ...
  39. [39]
  40. [40]
    [PDF] Kinetics of Thermal Decomposition of Titanium Hydride Powder ...
    Apr 6, 2005 · The results of TGA show that the decomposition of the titanium hydride becomes significant at about 450 °C. Above 500 °C the decomposition is ...
  41. [41]
  42. [42]
    Thermal decomposition of titanium hydride and its application to low ...
    Thermal decomposition behavior of titanium hydride (TiH2) powder at temperatures between 250 and 330·C (523 and 603 K) has been studied from both thermodynamic ...<|separator|>
  43. [43]
    Dehydrogenation of Titanium-Hydride Powder and Role of This ...
    Phase transformations, volume changes, and hydrogen release kinetics under dehydrogenation of compacted titanium-hydride powder are studied.
  44. [44]
    Thermal stability of titanium hydride modified by the electrochemical ...
    Oct 21, 2020 · This study aims to address the poor thermal stability of titanium hydride. Surface microstructural observations, differential thermal analysis, and electron- ...
  45. [45]
    Thermal characterization of titanium hydride in thermal oxidation ...
    We investigated the thermal behavior of titanium hydride with different particle size at the different atmospheres using the thermal analysis techniques.
  46. [46]
    [PDF] Investigation of Foamed Metals for Application on Space Capsules
    Titanium hydride was then added to break down the oxide film at sintering temperatures and permit metal-to-metal sintering. The sintered products again were ...
  47. [47]
    Designing compressive properties of titanium foams - ResearchGate
    Aug 7, 2025 · Titanium foams having various densities between 20 and 70% were manufactured and systematically characterized. Pore sizes varying from a few ...
  48. [48]
    Titanium Hydride, Grade P | Albemarle
    Titanium hydride powders find application in both pyrotechnic and metallurgical areas. They are utilized in initiator squibs and igniters.
  49. [49]
    Critical changes in the ignition and combustion characteristics of ...
    Titanium hydride potassium perchlorate (THPP), which is one of the most commonly utilized pyrotechnic initiators, can fail or deviate from the desired ...
  50. [50]
    Passivation of the Pyrotechnic Fuels Titanium and Titanium Hydride ...
    Passivation of the Pyrotechnic Fuels Titanium and Titanium Hydride in Oxidative Environments. · Publication Date, 1978 · Personal Author, Quinn, R. K. · Page ...
  51. [51]
    Titanium Hydride Powder - AG materials Inc.
    Therefore, it can be used as a high-purity hydrogen reserving material or as a foaming agent for aluminum foam manufacturing process. APPLICATIONS Titanium ...
  52. [52]
    Use of titanium hydride in the sintering of parts from an atomized ...
    During sintering the presence of a titanium hydride addition in a high-zinc brass powder promotes the reduction of the starting oxide films, prevents furth.Missing: reductant | Show results with:reductant
  53. [53]
    Grey Black Titanium Hydride Powder Used As Electric Vacuum Getter
    Titanium hydride is a chemical substance, usually is dark gray powder similar to metal. It is one of titanium smelting intermediate product in the chemical and ...Missing: halogens | Show results with:halogens
  54. [54]
    TITANIUM HYDRIDE POWDER - Atlantic Equipment Engineers
    Titanium hydride dark gray / black powder that is insuluble in water. TI 251 can be used in the production of sintered magnets, powdered metals, and titanium ...
  55. [55]
    Research and application of Ti–Mn-based hydrogen storage alloys
    Feb 28, 2023 · Ti–Mn-based hydrogen storage alloys are considered to be one of the most promising hydrogen storage alloys for proton exchange membrane fuel cell applications.
  56. [56]
    [PDF] Metallic and complex hydride-based electrochemical storage of ...
    May 30, 2023 · Metal hydrides are attractive anodes for Li-ion batteries thanks to their high specific and volumetric capacities along with their low potential ...
  57. [57]
    A Hybrid Intercalation and Conversion Mechanism for Reversible ...
    Feb 6, 2025 · (25,32,33) By fabricating lithium batteries with silicane-based anodes, Deng et al. reported an initial cycle capacity of 934 mAh/g and a ...
  58. [58]
    Study of printability and porosity formation in laser powder bed ...
    We studied the cost-efficient hydride-dehydride (HDH) Ti-6Al-4 V powder in L-PBF and show that the in-part porosity can be controlled to produce nearly fully ...
  59. [59]
    Powder Pressing, Shaping and 3D-Printing of Titanium Parts From ...
    Titanium hydride powder offers a cost-effective alternative to process high performance titanium parts. Hydrogen release during thermal treatment helps to ...
  60. [60]
    [PDF] Advanced Moderation Module for High-Temperature Micro-Reactor ...
    Aug 31, 2020 · If higher hydrogen partial pressure in the moderator enclosure is acceptable, zirconium and titanium hydrides may also be used. Yttrium hydride ...
  61. [61]
    Hydrided moderators for nuclear reactors (Patent) - OSTI.GOV
    Oct 31, 2023 · High temperature moderators for nuclear reactors and processes for their production are disclosed. The moderators include at least one hydrided metal.Missing: titanium post- 2010
  62. [62]
    Three-Dimensional Distribution of Titanium Hydrides After ... - MDPI
    Titanium hydride formation was observed in a region approximately 120 µm away from the titanium surface and correlates with the amount of absorbed hydrogen. We ...
  63. [63]
    Titanium Hydride Powder Market Size, Trends, Insights, & Market ...
    Rating 4.9 (59) According to recent data from the U.S. Geological Survey, titanium production reached approximately 150,000 metric tons in 2022, reflecting the growing interest ...Missing: annual pre- 2020
  64. [64]
    Hydrogen Storage Alloys Market Size and Statistics - 2035 - Fact.MR
    Hydrogen storage alloy market is valued at $3431.8 Mn in 2024 is set to grow at an 8.5% CAGR, reaching $8352.3 Mn by 2035, U.S. expected 10.3% by 2035.
  65. [65]
    The hydrogen embrittlement of titanium-based alloys | JOM
    This paper addresses the hydrogen embrittlement of titanium-based alloys. The hydrogen-titanium interaction is reviewed, including the solubility of hydrogen ...
  66. [66]
    [PDF] Hydriding of Titanium. - DTIC
    To determine the critical potential of hydride formation as a function of solution activity, solution temperature, applied strain, and surface conditions. • To ...
  67. [67]
    [PDF] Hydriding of Titanium. - DTIC
    The solubility of hydrogen in titanium varies with temperature, pressure, chemical purity, alloy composition, stress and deformation. Room temperature ...Missing: particle | Show results with:particle
  68. [68]
    Hydrogen Embrittlement in Titanium and Its Alloys - J-Stage
    It is considered that hydrogen embrittlement appears due to hydrogen diffusion to crack tips and hydride formation there during testing.<|separator|>
  69. [69]
    Determination of orientation relationships between FCC-hydride and ...
    Mar 15, 2020 · The present work aimed at exploring thoroughly the ORs and their interface planes between the fcc-hydride and the α-titanium using the stereographic ...
  70. [70]
    [PDF] 19760021477.pdf - NASA Technical Reports Server (NTRS)
    Blackburn, M. J. and Smyrl, W. H.: Stress Corrosion and Hydrogen. Embrittlement Critical Review, Titanium Sci. Technol., Proc. Int. Conf., 2nd, 2577-2609 (1973) ...
  71. [71]
    Hydrogen in Ti and Zr alloys: industrial perspective, failure modes ...
    Jun 12, 2017 · This paper aims to provide an overview of the role hydrogen has to play through the material life cycle.
  72. [72]
    Effects of Nb and Mo alloying on resistance to hydrogen ...
    Jul 3, 2020 · The complex alloying of Nb and Mo improves the resistance; however, the underlying microstructural evolutions and synergistic mechanisms are not ...
  73. [73]
    Hydrogen separation using coated titanium alloys - Google Patents
    the palladium coating allowed hydrogen to diffuse into the alloy but that it accumulated in the body of the alloy in regions called "traps". The accumulated ...
  74. [74]
    [PDF] nondestructive detection of hydrides and alpha-case in titanium alloys
    Current specifications limit hydrogen content to 50 to 150 ppm depending on the alloy and mill product. Above these contents, hydrides may form which can ...Missing: permeation | Show results with:permeation
  75. [75]
    Effect of hydrogen on corrosion resistance and fatigue properties of ...
    For most common titanium materials, the ASTM B348–10 standards require a maximum hydrogen content of 150 ppm [15], [16]. At room temperature, the maximum ...
  76. [76]
    [PDF] Effect of Hydride on the Mechanical Property of PT-7M Titanium Alloy
    PT-7M titanium alloys showed average yield strength values of 350 MPa with little ... Effect of hydrogen content on Vickers Hardness of PT-7M titanium alloy.
  77. [77]
    The hydrogen embrittlement of titanium-based alloys - ResearchGate
    Aug 7, 2025 · This paper addresses the hydrogen embrittlement of titanium-based alloys. The hydrogen-titanium interaction is reviewed, including the solubility of hydrogen ...
  78. [78]
    Preventing Hydrogen Embrittlement: The Role of Barrier Coatings ...
    May 17, 2023 · Hydrogen barrier coatings are expected to enable steels, which are susceptible to hydrogen embrittlement, specifically cost-effective low alloy- ...