Fact-checked by Grok 2 weeks ago

Triatomic hydrogen

Triatomic hydrogen, denoted as H₃, is the simplest , consisting of three atoms arranged in an equilateral triangular with D₃h and bond lengths of approximately 0.87–0.96 in its excited states. This highly unstable species exists exclusively in metastable excited electronic states, such as the 2pA₂″ and 2sA₁′, with lifetimes ranging from picoseconds to about 700 nanoseconds before predissociating into H + H₂. Its is repulsive and unbound, leading to immediate , while the metastable excited states feature shallow potential wells (e.g., of ≈ -2.07 eV for the 2pA₂″ state relative to H + H₂), rendering H₃ transient and observable only under specific laboratory conditions like electric discharges or Rydberg state excitations. The concept of triatomic hydrogen emerged in the early 1910s when J. J. Thomson proposed its existence based on positive ray analysis, suggesting it as a carrier of 3 in discharges, though this interpretation sparked decades of debate among chemists and physicists. Experimental confirmation came in 1979 when identified its —a diffuse band near 5600 —in a discharge, marking the first direct observation of H₃ and enabling detailed spectroscopic studies. Subsequent has focused on its quantum mechanical properties, including vibrational frequencies (e.g., 3213 cm⁻¹ for symmetric stretch and 1850 cm⁻¹ for bend) and electronic transitions like the 16695 cm⁻¹ band from 2pA₂″ → 3sA₁′. In modern contexts, triatomic hydrogen serves as a benchmark for theoretical models in due to its minimalistic structure and role in understanding polyatomic dynamics. It can be generated in controlled environments such as hollow-cathode discharges or via charge transfer in fast ion beams, where metastable states are populated for lifetime measurements and Rydberg series investigations. Although rare in nature owing to its instability, H₃ may form transiently in clouds through neutralization of the more prevalent (H₃⁺) and contributes to simulations of early chemistry, including primordial gas cooling and recombination processes. Ongoing studies explore its Rydberg states and predissociation pathways, providing insights into spin-orbit coupling and electric field effects on molecular lifetimes.

Molecular structure

Geometry

The ground electronic state of neutral triatomic hydrogen, H₃, features a (PES) with a at the degenerate D_{3h} equilateral triangular geometry due to the Jahn-Teller effect from its degenerate electronic configuration. This splits the PES into an upper (partly bonding) sheet and a lower () sheet; the resides primarily on the lower sheet, where distortions along the Jahn-Teller coordinates favor bent, isosceles triangular configurations leading to rapid dissociation into H + H₂, with no stable bound minimum. calculations for the reference equilateral configuration yield H-H distances of approximately 0.87 (1.65 a.u.), though actual bond variations occur along the paths. In contrast, excited states, particularly Rydberg states, display D_{3h} symmetry with a stable equilateral triangular geometry and support bound vibrational levels. These states exhibit vibrational modes characteristic of D_{3h}, including the symmetric stretch ν₁ (A₁') and the doubly degenerate bend ν₂ (E').

Electronic configuration

The electronic configuration of neutral triatomic hydrogen (H₃) is derived from the combination of three hydrogen 1s atomic orbitals, resulting in molecular orbitals of A₁', E', and A₂'' symmetry in D₃h geometry. The lowest-energy configuration places two electrons in the fully bonding A₁' orbital and the third electron in the degenerate, non-bonding E' orbitals, yielding a ²E' ground state that is electronically degenerate. This degeneracy leads to a Jahn-Teller distortion, lowering the symmetry to C₂ᵥ and splitting the ²E' state into ²A₁ and ²B₂ components, with the lower-energy ²B₂ sheet forming the effective ground potential energy surface; however, this configuration is unbound and repulsive, contributing to the molecule's instability. In the simple Hückel model adapted for σ-bonding, the molecular orbital energies are approximated as ε₁ = α + 2β / (1 + 2S) for the A₁' bonding orbital and ε₂ = ε₃ = α - β / (1 - S) for the degenerate E' orbitals, where α is the 1s orbital energy, β the resonance integral, and S the overlap integral; the three electrons occupy these orbitals with the singly occupied E' set driving the weak bonding character. Bond order analysis for this three-center three-electron system yields an effective order of 0.5 per H-H bond, reflecting the delocalized nature of the single electron in the non-bonding E' orbitals and the absence of antibonding occupancy, which results in only marginal stabilization insufficient to prevent dissociation. The first excited states arise from promotion of the E' electron to 2p Rydberg orbitals, with the lowest such state being ²A₂″ (from 2pσ) in D₃h , which is metastable with a lifetime on the order of hundreds of nanoseconds and supports bound vibrational levels; higher ²E' states (from 2pπ) exhibit linear geometries along certain coordinates due to reduced . Stable molecular structures are observed exclusively in these metastable excited states.

Physical properties

Stability and lifetime

Neutral triatomic hydrogen (H₃) exists as a metastable species primarily in its excited electronic states, while its ground electronic state (¹A₁') is unbound due to the zero-point vibrational energy lying above the dissociation threshold to H + H₂. This fundamental instability in the ground state arises from the shallow potential energy well in the equilateral configuration, where quantum zero-point effects prevent stable binding, leading to rapid dissociation. In the gas phase, the lifetime of neutral H₃ is less than 1 microsecond, confined to these metastable excited states formed via charge transfer reactions. The primary decay mode for metastable H₃ is spontaneous into H + H₂, occurring through predissociation where the couples to the repulsive potential surface, facilitating ultrafast fragmentation. This process is dominant over radiative decay, as evidenced by lifetimes significantly shorter than predicted for pure electronic transitions (e.g., 2p²A₂″ → 2s²A₁'). Factors such as rotational and vibrational excitation in the metastable states further influence stability, with lower rotational levels exhibiting longer lifetimes due to reduced coupling to dissociative pathways. Experimental measurements in the gas phase, using techniques like merged-beam charge exchange and photoionization, have determined lifetimes for the rotationless (N=0, K=0) level of the 2p²A₂″ at approximately 640 in the vibrational and 740 in the symmetric stretch-excited level. In matrix isolation environments, such as rare-gas solids at cryogenic temperatures, neutral H₃ exhibits lifetimes ranging from picoseconds to nanoseconds, extended slightly by the constrained but still limited by intrinsic dissociative tendencies. These short timescales underscore H₃'s transient nature, observable only under controlled conditions.

Thermodynamic parameters

The dissociation energy D_0 for the process \ce{H3 -> H + H2} is approximately 3.5 in low-lying Rydberg states of neutral , reflecting the effective binding in these metastable configurations before predissociation occurs. This value is derived from quantum-mechanical calculations and experimental data on the dynamics in Rydberg manifolds, adjusted for the H₂ dissociation energy of ≈4.5 eV relative to the three-body limit, establishing the scale of binding relative to the primary dissociation channel. The standard heat of formation \Delta H_f for neutral H3 at its potential minimum is estimated at approximately 4 kcal/mol (endothermic relative to 3/2 H₂), based on modern computational well depths of ≈2.07 below the H + H₂ asymptote. This small positive value highlights the marginal thermodynamic instability of the electronic compared to separated H and H₂ fragments, though zero-point effects render it unbound. Early calculations overestimated this at higher values like 120 due to approximate methods. Vibrational frequencies provide insight into the intramolecular forces in transient H3 structures, with the symmetric stretching mode \nu_1 at approximately 3213 cm^{-1} and the bending mode \nu_2 at approximately 1850 cm^{-1}, as computed for excited states in ab initio treatments of the equilateral triangular geometry. These frequencies, comparable to or slightly lower than those of diatomic H₂ due to delocalized bonding, contribute to the zero-point energy that influences equilibrium properties in bound states. Enthalpy and entropy for H3 are typically estimated via using the partition functions derived from its electronic, rotational, and vibrational levels, yielding a on the order of 40-60 J/· at 298 for hypothetical and an consistent with the heat of formation's marginal endothermicity. Such estimates, based on approximations, emphasize the high barrier to formation, reinforcing H3's role as a rather than a stable species.

Formation and reactivity

Synthesis methods

Neutral triatomic , H₃, is highly unstable and transient, with synthesis relying on methods that produce it for spectroscopic or dynamic studies. The primary approach involves the recombination of atoms in low-pressure gas , where atomic generated by the can form H₃ through collisions or excited states before rapid dissociation. This method was first used to observe H₃ emission spectra in or hollow cathode . Another key method is the neutralization of the (H₃⁺) via charge transfer reactions, such as with vapors or in fast ion beams, populating metastable Rydberg states of neutral H₃ for lifetime measurements and spectroscopic investigations.

Dissociation processes

The primary dissociation pathway for neutral triatomic (H₃) in its metastable 2p²A₂″ state is the unimolecular decay to a ground-state and a , represented as
\ce{H3 -> H(^{2}S) + H2(X ^{1}\Sigma_{g}^{+})}
This process occurs predominantly through predissociation, where the excited state couples to the repulsive ground electronic state via spin-orbit interaction, leading to fragmentation without a significant barrier.
The rate constant for this predissociation in the rotationless (N=0, K=0) ground vibrational state is approximately 1.6 × 10⁶ s⁻¹, corresponding to a lifetime of about 640 ns, measured under beam conditions near room temperature. Branching ratios favor production of ground-state products, with the majority of H₂ formed in low vibrational levels (v ≤ 2) and rotational states consistent with conservation of angular momentum from the parent molecule, while higher vibrational channels are minor due to energy constraints in the decay. Vibrational excitation in the symmetric stretch mode of the H₃ slightly modifies the lifetime to around 740 , but higher excess energy in the Rydberg states or excitations generally accelerates by enhancing to dissociative continua, reducing lifetimes to below 100 in some cases and increasing the efficiency of predissociation over competing radiative .

Spectroscopy

Spectral features

The ultraviolet spectral features of triatomic hydrogen consist of broad, continuous bands spanning 200–400 nm, exhibiting two pronounced maxima and arising from radiative between Rydberg states and the ground ^2E' state following predissociation. These features reflect the molecule's unstable nature, with the broadness attributed to rapid dissociation dynamics in the s. Visible spectra display diffuse rotational structure in parallel bands near 560 nm and 602.5 nm, as well as perpendicular bands near 710 nm, corresponding to the $2p ^2A_2'' \rightarrow ^2E' from the first . The diffuse character results from vibronic predissociation, limiting resolution and confirming the short-lived 's role in the observable signatures. Infrared spectra lack stable absorption due to H3's short lifetimes ranging from picoseconds to about 700 ns, but transient emission signals have been captured in time-resolved studies using neutralized ion beams, revealing Rydberg–Rydberg transitions such as those from 3p ^2A_1' to mixed higher states around 3600 cm⁻¹. These transient features provide glimpses into vibrational modes otherwise inaccessible. Raman scattering yields weak signals from symmetric stretching and bending modes, observable only under specialized conditions like high-density plasmas, owing to the molecule's low concentration and rapid decay. Recent advancements in femtosecond laser have enabled detection of predissociation lines in the UV-visible range, elucidating time-dependent dynamics of excited-state decay pathways.

Energy levels

The energy levels of triatomic hydrogen are characterized by quantized vibrational, rotational, and electronic states, reflecting its nominal planar D3h geometry in bound excited states, though the ground electronic state is unbound and exists primarily as resonances or in excited Rydberg states. The vibrational levels exhibit significant anharmonic effects due to the shallow in excited states, leading to Fermi resonances between the symmetric stretch mode ν1 and the overtone of the degenerate bending mode 2ν2, which mixes these states and alters the expected harmonic frequencies. These resonances are crucial for understanding the short lifetime and predissociation pathways of . For rotational levels, in its configurations—relevant for bent distortions or excited states—has rotational constants approximately A ≈ 21 cm⁻¹ and B ≈ 44 cm⁻¹, consistent with its oblate-like character and moments of . These constants govern the energy spacing in rotational sublevels, with the symmetric top structure (in D3h) giving rise to K quantum number projections, though in C2v leads to splitting of degenerate levels. The is given by E_rot = B J (J+1) + (A - B) K^2, where J is the and K is the along the . The electronic states of H3 include a ground state of ^2E' symmetry and low-lying excited states such as the 2p ^2A_2'' state, where Renner-Teller coupling plays a key role in the ground state. This coupling arises from vibronic interactions in the degenerate bending mode of the ^2E' state, lifting the degeneracy at non-linear geometries and resulting in upper and lower components of the . The Renner-Teller parameter k_RT quantifies the linear strength, leading to avoided crossings and enhanced predissociation rates in these states. Transitions between these energy levels follow specific selection rules for electric dipole allowed processes. For vibrational-rotational transitions, the rules are = ±1 for fundamental changes in vibrational quantum numbers and ΔJ = 0, ±1 for rotational, with ΔK = 0 for parallel bands and ΔK = ±1 for perpendicular bands in the symmetric top . These rules determine the observable spectral branches (, , ) in or visible spectra, though vibronic coupling can relax some restrictions in excited electronic states.

Ionic variants

H3+ cation

The triatomic hydrogen cation, denoted as H₃⁺, serves as the stable ionic form of triatomic hydrogen, contrasting with the highly unstable neutral H₃ radical. In its ground electronic state, H₃⁺ exhibits an equilateral triangular geometry with D_{3h} point group symmetry and an equilibrium bond length of approximately 0.87 Å between hydrogen atoms. This structure arises from the delocalization of the two valence electrons across the three protons, forming a symmetric molecular ion with aromatic-like stability due to its 2e⁻/3c² bonding motif. A primary formation pathway for H₃⁺ involves the exothermic protonation of molecular hydrogen: H₂ + H⁺ → H₃⁺, which releases approximately 4.4 eV of energy and serves as a cornerstone of ion-molecule reaction chains in various environments. This reaction is highly efficient at low temperatures, with a rate constant near the collision limit, making H₃⁺ a key initiator of interstellar chemistry by acting as a proton donor to neutral species. The stability of H₃⁺ is underscored by its dissociation energy of about 4.37 eV to H₂ + H⁺, which exceeds that of many diatomic ions and enables persistence in dilute, cold conditions. In the interstellar medium (ISM), H₃⁺ ranks as the most abundant molecular ion after H₂, with column densities often reaching 10¹⁴ cm⁻² in diffuse clouds, where it drives the synthesis of complex organics despite low densities (n ∼ 10–100 cm⁻³). Its longevity in the ISM stems from slow recombination with electrons and limited photodissociation, owing to an excitation energy of 19.3 eV that places absorptions beyond typical cosmic ray-induced UV fluxes. Recent investigations in 2025 have revealed alternative formation routes for H₃⁺ via double of methyl halides (CH₃X, where X = F, Cl, Br, I) and (CH₃CN, CH₃NC), proceeding through a H₂ mechanism that ejects a neutral H₂ fragment to abstract a proton, yielding H₃⁺ with yields up to 20% in femtosecond experiments. These pathways highlight H₃⁺ association with H₂ in transient clusters, such as H₃⁺(H₂)_n (n=1–10), where by H₂ molecules stabilizes the ion with binding energies of 10–20 meV per H₂, influencing reactivity in and astrophysical simulations.

Anionic forms

The anionic form of triatomic , H₃⁻, is theoretically predicted to adopt a linear asymmetric , resembling a (H₂) weakly bound to a anion (H⁻). In this configuration, the H-H within the H₂ subunit is approximately 1.421 (0.753 ), nearly identical to that of isolated H₂, while the distance from the H⁻ to the center of the H₂ is about 6.069 (3.21 ), indicative of a van der Waals-type interaction. Quantum chemical calculations, including coupled-cluster methods, reveal a shallow well for the electronic (¹Σ⁺), with a well depth (Dₑ) of approximately 401 cm⁻¹ (0.050 ). The zero-point-corrected (D₀) for the vibrational is around 70 cm⁻¹ (0.009 ), supporting only nine bound vibrational levels and underscoring the high instability of the species, which readily dissociates into H₂ + H⁻. This contrasts sharply with the stable, equilateral triangular structure of the H₃⁺ cation. Theoretically, H₃⁻ can form through electron attachment to neutral triatomic hydrogen (H₃ + e⁻ → H₃⁻) or via the association of H₂⁻ with an atomic hydrogen (H₂⁻ + H → H₃⁻). Despite these predictions, no confirmed experimental detection of H₃⁻ has been achieved as of 2025, with studies limited to computational explorations of its and spectroscopic properties.

Theoretical investigations

Computational approaches

Ab initio methods form the cornerstone of theoretical modeling for neutral triatomic hydrogen (H₃), providing insights into its fleeting geometry and energetics. The Hartree-Fock (HF) method serves as the starting point, offering a mean-field approximation to the electronic structure but neglecting electron correlation, which leads to overestimated bond lengths and energies for this weakly bound system. To account for correlation effects, second-order Møller-Plesset perturbation theory (MP2) improves upon HF by including pairwise electron interactions, yielding more accurate equilibrium geometries. The "gold standard" coupled-cluster method with singles, doubles, and perturbative triples excitations, CCSD(T), further refines these results through systematic inclusion of higher-order correlations, optimizing the ground-state (^2E') geometry as an unstable with bond lengths of approximately 1.97 a.u. and an energy of about 2.05 eV above the H + H₂ dissociation limit. High-accuracy basis sets are crucial for reliable calculations on H₃, given its delocalized electrons and proximity to . Augmented correlation-consistent polarized quadruple-zeta (aug-cc-pVQZ) basis sets, comprising 5s4p3d2f functions per with diffuse primitives, achieve near-complete basis set limits for , reducing errors to below 0.1 kcal/ compared to experimental benchmarks for related systems. These basis sets balance computational cost and precision, enabling convergence in optimizations and evaluations essential for capturing the shallow of neutral H₃. Quantum dynamics methods, particularly wavepacket propagation, elucidate the ultrafast dissociation of neutral H₃ by solving the time-dependent Schrödinger equation on multi-dimensional potential surfaces. In studies of the short-lived 3²A'(2sa'₁) excited state, three-dimensional wavepacket dynamics on coupled ground- and excited-state surfaces reveal three-body predissociation pathways, with initial wavepackets evolving to produce correlated momentum distributions among the three H fragments, highlighting nonadiabatic couplings and rotational effects in D₃ and H₃ isotopomers. These simulations, typically employing split-operator or Chebyshev propagation techniques, provide lifetimes on the order of femtoseconds and validate experimental momentum correlations. Recent 2025 advancements incorporate potentials to enable real-time simulations of neutral H₃ dynamics, trained on sparse datasets from hydrogen atom transfer reactions like H + H₂ exchange. models, such as those using invariant representations, approximate CCSD(T)-level energies and forces with root-mean-square deviations below 0.5 kcal/mol, facilitating long-timescale trajectory calculations that were previously prohibitive due to high computational demands. These potentials enhance accessibility for exploring pathways and derived properties like infrared spectra.

Potential energy surfaces

The potential energy surface (PES) for the ground electronic state of triatomic hydrogen H3, which governs the H + H2 exchange reaction, features a in the collinear with D∞h , located at an H-H of approximately 1.82 a_0 and a classical barrier height of 0.425 relative to separated H + H2 reactants. This represents the for dissociation into H + H2 products, with the linear minimizing the energy along the reaction path. The collinear approach of the incoming H atom to exhibits the lowest barrier, approximately 0.42 , facilitating direct abstraction-exchange dynamics at and suprathermal energies. In comparison, the approach (C_{2v} ), where the H atom attacks orthogonal to the H2 axis, faces a substantially higher barrier of about 1.1 due to increased repulsion in the compact triatomic region, rendering it less favorable and shifting reactivity toward collinear geometries. The PES is commonly parameterized as V(R_1, R_2, \theta), where R_1 and R_2 are the two H-H internuclear distances and \theta is the angle between them, allowing representation of the full three-dimensional landscape from asymptotic H + regions to the central barrier. Fitted analytic forms, such as the BKMP2 potential, decompose V into two-body (diatomic) and interaction terms for global accuracy, providing sub-chemical accuracy (rms error < 0.1 kcal/mol) across the surface based on extensive data. To capture vibrational effects near the and improve dynamical predictions, anharmonic corrections are essential, incorporating cubic and quartic terms in the Taylor expansion of the PES around stationary points; these terms account for mode-mode couplings and reduce errors in and frequencies by up to 20% compared to approximations.

Occurrence and detection

Astrophysical contexts

Neutral triatomic hydrogen (H₃) forms transiently in diffuse clouds through the recombination reaction H + H₂ → H₃, which is part of the broader process H + H + H → H₂ + H where the third body stabilizes the system. However, H₃ is highly unstable in its and dissociates rapidly, with lifetimes typically below 1 μs due to its unbound nature and predissociative pathways. In the low-density environment of diffuse clouds (n_H ≈ 10–100 cm⁻³), the stabilization rate is inefficient, limiting H₃ to fleeting intermediates that do not accumulate to observable levels. The short lifetime of H₃ precludes its detection in interstellar spectra, as it lacks stable rovibrational states suitable for or lines in the observable range. Upper limits on its abundance have been inferred from observations of sightlines through diffuse clouds, where no signatures appear despite sensitive searches for potential electronic transitions; these limits suggest column densities at least three orders of magnitude below those of H₃⁺. In contrast to the pervasive H₃⁺ ion, which drives much of the gas-phase chemistry in these regions, neutral H₃ plays no significant steady-state role. In chemistry, acts as a key intermediate in the formation of molecular during the of metal-free halos in the early . At densities exceeding 10⁸ cm⁻³ and temperatures around 800–1000 K, the recombination H + H + H → H₃ → H₂ + H becomes efficient, enabling H₂ buildup essential for cooling and fragmentation leading to the first . Quantum dynamical calculations incorporating Jahn-Teller in H₃ enhance the coefficients by up to 12% at 300 K, underscoring its transient but crucial role before . Recent astrochemical models indicate that H₃ contributes only marginally to in photon-dominated regions, primarily via of transient H₃ followed by rapid reactions, but its low abundance renders this pathway negligible compared to the dominant cosmic-ray initiated route.

Laboratory observations

The first laboratory detection of triatomic hydrogen (H₃) occurred in 1979 through observation of its in a hollow cathode discharge. Herzberg and coworkers identified parallel bands near 5600 and 6025 for H₃, along with analogous features for D₃, arising from transitions between Rydberg states and the repulsive . These diffuse bands provided initial evidence for the transient existence of H₃ in excited electronic states, with rotational structure consistent with a planar geometry. Further characterization of metastable H₃ followed in the late and early using charge exchange reactions to generate from H₃⁺ ions. In 1990, the lifetime of metastable H₃ was determined to be approximately 1 μs via in a time-of-flight mass spectrometer, where H₃ formed by H₃⁺ reaction with cesium vapor was ionized to yield m/z = 3 signals, confirming the precursor. This approach isolated the dynamics of the metastable state, revealing radiative decay pathways from Rydberg levels. In 1993, UV emission spectra from isotopic variants of neutral H₃, produced in a fast molecular beam via impact of H₂, provided evidence for the double-sheeted nature of the ground-state . The observed isotopic shifts in the emission bands supported theoretical predictions of two electronically coupled sheets, enabling transient H₃ survival on timescales before .

Historical background

Theoretical predictions

In the mid-1930s, early quantum mechanical calculations using suggested that the neutral triatomic molecule, H₃, was unstable. Joseph O. Hirschfelder, Henry Eyring, and Norman Rosen performed variational calculations on the energy of H₃ in linear and bent configurations, finding that the was insufficient to form a stable compared to three separated atoms. Their work, building on the Heitler-London approach for diatomic , indicated that H₃ dissociated readily, with no minimum in the for the . Subsequent quantum calculations in reinforced this view of an unbound . Applications of the Heitler-London valence bond method to H₃ configurations showed that the lacked sufficient exchange stabilization to overcome repulsion in the neutral form, leading to predictions of instability under normal conditions. These efforts highlighted the challenges in extending simple or valence bond models to triatomic systems, where symmetry and delocalization played critical roles. In 1937, Hermann Arthur Jahn and published their theorem on the instability of polyatomic molecules in degenerate electronic states, which had direct implications for H₃. The theorem predicted that electronic degeneracy in symmetric configurations, such as the for H₃, would lead to distortions via vibronic coupling to lower the energy, preventing stable equilibrium in degenerate states. These considerations contributed to theoretical doubts about the persistence of H₃ in such states. Pre-1950 semi-empirical models provided nuanced insights into potential . Hirschfelder's 1938 analysis using a semi-empirical scheme, incorporating and integrals, predicted a shallow energy well for certain collision complexes of H₃, suggesting short-lived metastable states during atomic recombination despite overall instability. These models, calibrated against known diatomic potentials, indicated that while the was unbound, transient configurations could exhibit temporary stability before .

Experimental milestones

The hypothesis of neutral triatomic hydrogen emerged in the early from J. J. Thomson's positive ray analysis in hydrogen discharges, where he interpreted particles of 3 as H₃ molecules. This proposal initiated decades of debate and experimental efforts to detect H₃, including spectroscopic searches in the and , though quantum theoretical predictions of instability and the later of tritium as the mass-3 isotope shifted interpretations. The first laboratory observation of the spectrum of neutral triatomic hydrogen, H₃, was reported in 1979 by Gerhard Herzberg and colleagues using a hollow-cathode electric discharge through low-pressure molecular hydrogen gas. The experiment revealed diffuse emission bands near 5600 Å and 6025 Å, identified as originating from Rydberg-Rydberg transitions in the excited states of H₃, marking the initial empirical confirmation of this transient species after decades of theoretical speculation. Building on this, Herzberg and his team conducted high-resolution spectroscopy in the early 1980s to resolve the rotational structure of these bands, confirming the equilateral triangular (D_{3h}) geometry of H₃ in its Rydberg states and distinguishing it from potential contaminants like H₂. Key analyses included parallel bands observed in emission spectra of H₃ and its deuterated isotopomer D₃, which provided precise molecular constants and vibrational frequencies, solidifying the structural assignment. Additional infrared emission studies further elucidated the energy levels and transition moments, demonstrating H₃'s short lifetime in the ground state due to predissociation.

References

  1. [1]
    H3 properties
    Triatomic hydrogen (H₃) represents the simplest possible triatomic molecule, consisting of three hydrogen atoms arranged in an equilateral triangular ...
  2. [2]
    A Controversial Molecule: The Early History of Triatomic Hydrogen
    Sep 13, 2011 · The hypothesis of a triatomic hydrogen molecule H3 originated in the early 1910s in connection with J. J. Thomson's experiments on positive ...
  3. [3]
    A spectrum of triatomic hydrogen | The Journal of Chemical Physics
    May 15, 1979 · The possibility of HD whose excited states have B values In the range 15–24 cm −1 can easily be eliminated since in discharges in mixtures ...
  4. [4]
    Rydberg States of Triatomic Hydrogen and Deuterium
    The triatomic hydrogen ion (H 3 +) has spurred tremendous interest in astrophysics in recent decades, and Rydberg states of H3 have also maintained an important ...Introduction · Method · 3pπ Rydberg States of D3 · References
  5. [5]
    Triatomic hydrogen ion generation in a low-pressure gas discharge
    Triatomic hydrogen and deuterium ions are generated in a hollow-cathode discharge. •. Plasma ion composition is determined by discharge current and working ...
  6. [6]
    Experimental and quantum-chemical studies on the three-particle ...
    Dec 27, 2005 · The state preparation involves metastable triatomic hydrogen molecules which are produced in a fast (3 keV) beam by charge transfer ...
  7. [7]
    Measurement of the lifetime of metastable triatomic hydrogen
    We conclude that spin-orbit-coupling induced predissociation may indeed be one decay path that could shorten the lifetime of metastable H3. As a second ...
  8. [8]
    [PDF] METASTABILITY AND RYDBERG STATES OF TRIATOMIC ... - HAL
    Abstract : The np,nd and nf Rydberg series of H3 have been studied by one- or two- photon excitation from the lowest metastable state of H3 : B 2p 24.
  9. [9]
    A Semiempirical and Ab Initio Study of a Simple Jahn–Teller System
    Karplus. Department of Chemistry, Harvard University, Cambridge, Massachusetts. J. Chem. Phys. 49, 5163–5178 (1968). https://doi.org/10.1063/1.1670017. Article ...
  10. [10]
    [PDF] Dynamics on the ground-state potential surfaces of H3 and its ...
    These maxima could be attributed to the Jahn-Teller struc- ture of the ground-state potential surface that has two ener- getically clearly separated sheets @8#.
  11. [11]
    The electronic emission spectrum of triatomic hydrogen. I. Parallel ...
    A spectrum of triatomic hydrogen and deuterium was first discovered by means of an emission band with diffuse rotational structure near 5600 Å. An ...
  12. [12]
  13. [13]
    [PDF] Three-body dissociation dynamics of the low-lying Rydberg states of ...
    Oct 8, 2004 · sociative 2p 2E0 ground-state is doubly degenerate in D3h symmetry and undergoes a Jahn-Teller distortion, lifting the degeneracy. This ...
  14. [14]
  15. [15]
    [PDF] Rydberg States of H3 and HeH as Potential Coolants for Primordial ...
    ... neutral H3 molecule, which then collides with He to form a “very long-lived Rydberg state with high orbital momentum.” The rotationally excited H3 and D3 ...<|control11|><|separator|>
  16. [16]
  17. [17]
    Predissociation of H3 n=2 Rydberg states - AIP Publishing
    This energy consists of 1.0 eV of elec- tronic energy plus the average rovibrational energy. The average W values in the three-body distributions were ob-.
  18. [18]
    I. Calculation of Energy of H 3 Molecule - AIP Publishing
    The activation energies are for the reaction H+H2→H2+H, and are calculated with the same type of approximation for H2 and H3. All of the difficult three‐center ...
  19. [19]
    I. Calculation of Energy of H3 Molecule - Semantic Scholar
    The molecular-orbital method has been applied to a study of in its ground state and excited levels, and the relative importance of the perturbation and ...
  20. [20]
    The Energy of the Triatomic Hydrogen Molecule and Ion, V.
    Dec 1, 2024 · 3 The triatomic hydrogen ion has its stable configuration intermediate between a right and an equilateral triangle.Missing: geometry | Show results with:geometry
  21. [21]
  22. [22]
    Improved infrared and visible emission spectra of the H3 and D3 ...
    In the present work, infrared and visible emission spectra of H3 and D3 have been measured using a corona discharge source of the type described by Droege and ...
  23. [23]
    The photolysis of ammonia | Proceedings of the Royal Society of ...
    There are two general explanations of the discrepancy: (1) the hydrogen atoms are not produced as fast as that calculated on the assumption that every ammonia ...Missing: H3 | Show results with:H3
  24. [24]
    Electron-impact excitation of molecular hydrogen | Phys. Rev. A
    Feb 6, 2017 · Accurate electron-impact electronic excitation cross sections of molecular hydrogen are important for modeling various plasmas. The applications ...Missing: H2 H3
  25. [25]
    Unique Applications of para-Hydrogen Matrix Isolation to ...
    The matrix-isolation technique using noble-gas matrix hosts (Ne, Ar, Kr, Xe, and N2) has been widely employed to preserve unstable species for spectroscopic or ...Missing: H3 | Show results with:H3
  26. [26]
    Two-Stage Pulsed Laser Ablation for the Production of Ag@TiO2 ...
    This study aims to synthesize Ag@TiO2 core–shell nanoparticles using a two-stage pulsed laser ablation method for potential biomedical applications.
  27. [27]
  28. [28]
    Product state distributions in the dissociation of H3 (n=2,3) Rydberg ...
    Sep 1, 1996 · Dissociation of the 2s 2A1′, 2p 2A2″, 3s 2A1′, and 3d 2E″ Rydberg states of the H3 molecule is investigated using a fast neutral beam ...
  29. [29]
    Rydberg states of triatomic hydrogen(H3): application of neutralized ...
    Rydberg states of triatomic hydrogen(H3): application of neutralized ion beam techniques | Accounts of Chemical Research. Molecular Modeling: Principles and ...
  30. [30]
    Intramolecular Vibrational Energy Redistribution in the Reaction H 3 ...
    Oct 7, 2024 · ... 2ν2, ν1, and ν3 states, were obsd. to derive the rotational, centrifugal distortion, and l-type doubling consts. The Fermi resonance between ...
  31. [31]
    [PDF] Analysis of the Rotational−Vibrational States of the Molecular Ion H3
    Nov 5, 2013 · makes it possible to exploit a C2v(M) symmetry of the H3. + molecule. Although the full symmetry of H3. + is D3h(M), the. C2v(M) symmetry ...
  32. [32]
    [PDF] Topological Study of the H3 ++ Molecular System - arXiv
    The numerical study centers on the three lower states of these molecular systems and consequently we refer to two kinds of conical intersections (ci): one kind ...Missing: ground | Show results with:ground
  33. [33]
    Measurement of vibrational frequencies of the H3 molecule using ...
    Aug 6, 2025 · Triatomic hydrogen in the lowest rotational level of the 2p2A2' state is long-lived (∼80 μs). This molecule is readily formed in charge transfer ...
  34. [34]
    Finite temperature quantum statistics of H3+ molecular ion
    The equilibrium geometry of the H 3 + ion in its ground state is an equilateral triangle D 3 h for which the internuclear equilibrium distance is R = 1.65 a 0 ⁠ ...
  35. [35]
    A simple approximate solution for the H3+ ion - Wiley Online Library
    Sep 27, 2022 · This Tutorial Review guides the student step-by-step toward the variational solution of the trihydrogen cation (H3+) in its equilateral ...
  36. [36]
    Interstellar H3+ - PNAS
    Protonated molecular hydrogen, H 3 + , is the simplest polyatomic molecule. It is the most abundantly produced interstellar molecule, next only to H 2.
  37. [37]
    Experimental determination of the H3+ bond dissociation energy
    Aug 8, 2025 · The results of recent photoionization and photodissociation studies of the H3 molecule are used to establish an experimental value for the H3+ ion bond ...
  38. [38]
    Factors of $${\rm H}_{3}^{+}$$ Formation from Methyl Halogens
    Jan 6, 2025 · The H 3 + formation dynamics via a roaming mechanism considered in the present study proceeds mainly on the lowest singlet states of the CH3X2+ ...Missing: association H2
  39. [39]
  40. [40]
    (PDF) Bound states, resonances, and formation of the H3^- anion
    Only A1 and A2 vibrational symmetries are shown. The curves of the E symmetry are very similar to the A 1 and A 2 symmetries for energies below 4000 cm −1 ( ...
  41. [41]
    [PDF] The Electronic Structure of H3+, H3 and H3 - RIULL Principal
    Sep 7, 2021 · In the case of H3. –. , the ab initio calculations revealed a linear equilibrium configuration in the ground state. Hence, for the reference ...<|control11|><|separator|>
  42. [42]
    An accurate three‐dimensional potential energy surface for H3
    Mar 1, 1978 · An accurate three‐dimensional potential energy surface for H 3 has been obtained by the configuration interaction (CI) method.
  43. [43]
    [PDF] Potential Energy Surfaces and Elastic and Inelastic Scattering," DG ...
    "H+H2: Potential Energy Surfaces and Elastic and Inelastic Scattering," D. G. Truhlar and R. E. Wyatt, in Advances in Chemical Physics, Vol. 36, edited by I ...
  44. [44]
    A refined H3 potential energy surface - AIP Publishing
    M. Karplus. ,. J. Chem. Phys. 49. ,. 5163. (. 1968. ). Google ...
  45. [45]
    Spectra and lifetimes of triatomic hydrogen molecules
    Excited neutral triatomic hydrogen molecules were formed by charge transfer to beams of , , D2H+, and H2D+ ions. The density of the molecules generated was ...
  46. [46]
    [PDF] Jahn-Teller effect in three-body recombination of hydrogen atoms
    The x-axis is in log- scale. Channels that converge to the H + H + H dissociation limit at the energy E = 0 are three-body channels (dashed lines).<|separator|>
  47. [47]
    H3 +: The initiator of interstellar chemistry - ResearchGate
    Aug 6, 2025 · The H 3 + molecule plays a crucial role in interstellar chemistry due to its efficiency as a proton donor in H 3 + + X → H 2 + HX + ion-molecule ...
  48. [48]
    Experimental evidence for the double‐sheetedness of the ground ...
    Oct 20, 1993 · The uv spectra of isotopic variants of the neutral triatomic hydrogen molecule were measured as they are emitted by a fast molecular beam.
  49. [49]
    Stability of polyatomic molecules in degenerate electronic states
    We investigate the conditions under which a polyatomic molecule can have a stable equilibrium configuration when its electronic state has orbital degeneracy.
  50. [50]
    The electronic emission spectrum of triatomic hydrogen. III. Infrared ...
    Measurement of the lifetime of metastable triatomic hydrogen. Go to Citation Crossref Google Scholar. 44. Infrared laser spectroscopy of H2 and D2 Rydberg ...
  51. [51]
    How a triatomic molecule works off excess energy | RIKEN
    Jul 17, 2025 · A resonance effect can significantly affect how a three-atom molecule cools down when excited, RIKEN physicists have found.