Fact-checked by Grok 2 weeks ago

Trihydrogen cation

The trihydrogen cation, H₃⁺, is the simplest stable polyatomic molecular ion, comprising three protons and two electrons arranged in a planar equilateral triangular structure with D₃ₕ symmetry and a of approximately 0.90 . First identified in 1911 by J. J. Thomson through studies of gas discharges, it was initially controversial but confirmed as a key species in plasma chemistry. Its spectroscopic detection in Jupiter's in 1989 via emission lines marked the first astronomical observation, later extending to Saturn, , and the . H₃⁺ forms primarily through the radiative association of H⁺ with H₂ or collisions between H₂ and H₂⁺, making it ubiquitous in ionized environments. Despite its electronic simplicity—serving as a benchmark for ab initio quantum chemistry calculations—H₃⁺ exhibits high reactivity as a universal proton donor, reacting with any neutral species whose proton affinity exceeds 4.39 eV to form HX⁺ + H₂. This proton-transfer mechanism positions it as the cornerstone of interstellar ion-molecule chemistry, initiating the synthesis of virtually all complex molecules observed in space, from water to hydrocarbons. In diffuse interstellar clouds, its abundance probes cosmic-ray ionization rates (typically 10⁻¹⁷ to 10⁻¹⁴ s⁻¹), while in planetary atmospheres like Jupiter's aurorae, it emits rovibrational lines at temperatures above 1000 K, revealing ionospheric dynamics. Destruction occurs mainly via dissociative recombination with electrons, the dominant sink in low-density regions. H₃⁺'s role extends to laboratory plasma physics and astrochemistry, where it facilitates ortho-para conversion of H₂ in dense clouds and serves as a diagnostic for early universe conditions. Detected in regions like the with column densities up to 3 × 10¹⁵ cm⁻², it underscores the ion's prevalence as the most abundant reactive ion in the cosmos. Ongoing research leverages advanced to refine its to nanohartree accuracy, enhancing models of molecular formation in star-forming nebulae.

Historical Development

Theoretical Prediction

The trihydrogen cation, denoted as H₃⁺, was first proposed as a possible ionic species in by J. J. Thomson during his investigations of positive rays using early techniques, where he identified particles with a of 3, attributing them to a bound system of three atoms carrying a positive charge. Although initial interest in triatomic hydrogen focused more on the neutral H₃ molecule, which was ultimately deemed unstable, theoretical attention turned to the cation in amid growing quantum mechanical insights into molecular ions. In 1935, C. A. Coulson conducted a pioneering quantum mechanical analysis of H₃⁺ using , demonstrating its stability relative to dissociation into H₂ and H⁺, with a predicted bond length of approximately 0.85 . Coulson's calculation established H₃⁺ as possessing an equilateral triangular geometry with D₃ₕ symmetry, bound by a delocalized three-center two-electron system where the two electrons are shared symmetrically among the three protons. Despite these foundational predictions, H₃⁺ received little further attention until mid-20th-century laboratory experiments revived interest in its properties.

Laboratory Synthesis

The first laboratory detection of the trihydrogen cation, H₃⁺, occurred in 1911 when J. J. Thomson observed ions of 3 in a hydrogen discharge tube using positive ray analysis, an early form of . This observation was initially met with skepticism due to the unusual stability of the species, but it laid the groundwork for subsequent experimental confirmations. Theoretical predictions of H₃⁺ stability from had anticipated such a symmetric, equilateral triangular structure bound by delocalized protons. In 1925, T. R. Hogness and E. M. Lunn provided definitive confirmation through mass spectrometric studies of electron-impact ionization of H₂, identifying the key formation reaction H₂⁺ + H₂ → H₃⁺ + H with high efficiency due to its exothermicity of approximately 1.7 . This work established H₃⁺ as a stable product in controlled gas-phase environments, observable at pressures around 10⁻³ and electron energies above 15 . Subsequent mass spectrometric experiments in the mid-20th century refined these findings, demonstrating H₃⁺ dominance in plasmas. The marked a significant advancement with the development of cyclotron resonance (ICR) by J. L. Beauchamp and colleagues, which enabled precise studies of H₃⁺ stability and reactivity in trapped ensembles at low pressures (10⁻⁶ to 10⁻⁸ ). ICR techniques allowed measurement of reaction rates for H₃⁺ with neutrals like hydrocarbons, confirming its role as a with lifetimes on the order of milliseconds under ambient conditions. These methods isolated H₃⁺ by selective ejection of other ions, facilitating investigations into its proton-transfer and association reactions. Modern laboratory synthesis employs Fourier-transform ICR (FT-ICR) for high-resolution analysis of H₃⁺ reactions and lifetimes, often generating the via or impact on H₂ mixtures. FT-ICR achieves resolutions exceeding 10⁶, enabling detailed kinetic studies at temperatures down to 80 K. Complementary experiments in ion storage rings, such as CRYRING and TSR, store vibrationally relaxed H₃⁺ at cryogenic temperatures (below 10 K), revealing lifetimes extending to seconds or longer due to suppressed blackbody radiation-induced dissociation. Key experiments have quantified radiative association rates for H₃⁺ formation, such as H⁺ + H₂ → H₃⁺ + hν, measured using merged-beam techniques at relative energies near 0.1 eV, yielding rate coefficients around 10⁻¹⁶ cm³ s⁻¹ at 300 K. studies, often via ICR or collisions with He or H₂ at keV energies, have probed H₃⁺ stability, showing dissociation thresholds near 1.5 eV and cross sections scaling as E⁻¹ for low-energy impacts.

Initial Astronomical Observations

The initial astronomical detection of the trihydrogen cation (H₃⁺) occurred in 1989, when its emission was identified in Jupiter's auroral . Using high-resolution at the Canada-France-Hawaii Telescope (CFHT) on , Drossart et al. observed the (2ν₂) band near 2.093 μm, revealing unexpectedly strong auroral emissions from this ion. This serendipitous finding, enabled by the laboratory spectrum measured just a decade earlier, confirmed H₃⁺ as a key component of planetary upper atmospheres and highlighted its role in ionospheric chemistry. Subsequent ground-based observations in the extended these detections to other giant planets, but the first interstellar identification came in 1996. Geballe and Oka reported H₃⁺ absorption in the fundamental ν₂ band at 3.67 μm toward two dense molecular clouds (AFGL 2136 and W33A) using the United Kingdom Infrared Telescope (UKIRT) on . These observations, which relied on precise rotational-vibrational benchmarks for line identification, demonstrated H₃⁺ abundances sufficient to initiate much of ion-molecule chemistry, with column densities on the order of 10¹⁴ cm⁻². Early detections in diffuse interstellar clouds further underscored H₃⁺'s ubiquity. In 1998, McCall et al. detected it toward the bright source Cygnus OB2 No. 12 using UKIRT, measuring a column of approximately 2 × 10¹⁴ cm⁻² and noting its persistence in low- environments where cosmic-ray sustains the . These pioneering observations overcame significant challenges from Earth's atmospheric and telluric lines in the 3-4 μm window, requiring high-altitude sites, , and cryogenic to achieve the necessary signal-to-noise ratios for weak features.

Molecular Structure

Geometry and Bonding

The trihydrogen cation, H₃⁺, possesses an equilateral triangular equilibrium geometry with D_{3h} point group symmetry, featuring three identical H–H bond lengths of approximately 0.872 Å. This structure arises from the delocalization of the two valence electrons across the three hydrogen nuclei, stabilizing the planar configuration over linear alternatives. High-accuracy ab initio calculations confirm the bond angle of exactly 60° inherent to the equilateral triangle. The bonding nature of H₃⁺ is best described as a three-center two-electron (3c–2e) bond, a classic example of electron-deficient bonding where the shared electrons provide stability without pairwise localization. In terms of molecular orbitals under D_{3h} , both valence electrons occupy the fully bonding σ orbital of a₁' (derived from the in-plane combination of 1s orbitals). The degenerate non-bonding π orbitals of e' (out-of-plane combinations) are unoccupied. The unoccupied antibonding σ* orbital of a₂'' lies higher in energy, contributing to the overall . This orbital picture explains the ion's stability and reactivity, with the electrons primarily in the bonding manifold. The of H₃⁺ features a barrierless pathway for its formation via of H₂ by H⁺, allowing efficient assembly without an activation barrier along the . of H₃⁺ back to H₂ + H⁺ requires an energy of 4.373 ± 0.021 eV, as determined experimentally from and thresholds. Modern approaches, particularly coupled-cluster theory with single, double, and perturbative triple excitations [CCSD(T)] using augmented correlation-consistent basis sets, have refined these surface details, yielding bond lengths of 0.8726 and accurate harmonic vibrational frequencies such as 3496 cm⁻¹ for the symmetric stretch (ν₁, a₁') and 1257 cm⁻¹ for the degenerate bend (ν₂, e'). These calculations achieve near-spectroscopic accuracy, underpinning the ion's structural characterization. Isotopologues of H₃⁺, such as those involving deuterium, retain the same equilateral triangular geometry with minor adjustments due to reduced mass effects.

Isotopologues

The trihydrogen cation exhibits several isotopologues, primarily H₃⁺, D₃⁺, H₂D⁺, and HD₂⁺, which arise from the substitution of hydrogen atoms with deuterium. Deuteration influences the molecular geometry through differences in zero-point energy (ZPE), with the ZPE of H₃⁺ calculated at 4361.7 cm⁻¹ compared to 3112.3 cm⁻¹ for D₃⁺, resulting in reduced vibrational amplitudes for heavier isotopologues. This leads to slight variations in vibrationally averaged bond lengths; for H₃⁺, the equilibrium bond length is 0.873 Å, while the vibrationally averaged value is 0.914 Å, and deuteration brings the averaged length for D₃⁺ closer to the equilibrium due to lower ZPE. In the , the relative abundances of these isotopologues are determined by the low cosmic D/H ratio of approximately 1.4 × 10⁻⁵, which limits deuterated forms to trace levels. Spectroscopically, the isotopologues display distinct rotational and vibrational properties; for H₃⁺, the ground-state rotational constants are B₀ = 43.51 cm⁻¹ and C₀ = 20.59 cm⁻¹, with values scaling inversely with for deuterated species like D₃⁺ and H₂D⁺. Vibrational transitions shift to lower energies upon deuteration, exemplified by the ν₂ mode lines around 3.67 μm for H₃⁺ versus the ν₁ mode at approximately 3.52 μm for H₂D⁺. The mixed isotopologues H₂D⁺ and HD₂⁺ additionally feature and nuclear isomers, influencing their rotational level populations.

Ortho and Para Isomers

The trihydrogen cation exhibits two distinct nuclear spin , known as -H₃⁺ and -H₃⁺, arising from the identical fermionic nature of its three protons. The isomer possesses a total nuclear I = 3/2 and A₁ symmetry, while the isomer has I = 1/2 and A₂ symmetry. At high temperatures, the statistical weight ratio of to is 3:1, determined by the degeneracies of the nuclear spin states (2I + 1). The intrinsic energy difference between the ortho and para isomers is negligible, allowing both to coexist under typical conditions. However, symmetry-imposed restrictions limit the accessible rotational levels based on the requirement for the total wavefunction to have A₂ symmetry: the lowest rotational levels are thus (J,K) = (1,1) for para-H₃⁺ and (1,0) for ortho-H₃⁺, with an energy separation of 32.86 between them. Interconversion between the isomers proceeds via proton reactions, primarily H₃⁺ + H₂, or through collisions with other , but these processes are inefficient at low temperatures prevalent in environments. studies indicate that the hop-to- branching ratio (α) in the reaction decreases from 1.6 ± 0.1 at 350 to the statistical value of 0.5 ± 0.1 at 135 , reflecting slower equilibration as temperature drops. In cryogenic ion trap experiments, including those employing (FT-ICR) techniques, the ortho/para ratio equilibrates to near the high-temperature statistical limit only after times exceeding 1 hour in the presence of normal H₂ as a collision partner. These nuclear spin isomers influence spectroscopic selection rules, prohibiting direct transitions between and states and thereby segregating their rotational spectra.

Spectroscopic Properties

The absorption spectrum of the trihydrogen cation, H₃⁺, is primarily characterized by its vibrational modes, which serve as the basis for and astronomical detection. The vibrational modes are: ν₁ (a₁', symmetric stretch, band origin ≈3178 cm⁻¹, IR inactive); ν₂ (a₂'', umbrella bending, band origin ≈1151 cm⁻¹ or 8.68 μm, IR active); ν₃ (e', asymmetric stretch, band origin ≈2521 cm⁻¹ or 3.97 μm, IR active and degenerate). These modes reflect the D₃ₕ of the ion, with the ν₂ and ν₃ modes being IR-active due to changes in the . The ν₃ band, observed in the 3.5–4 μm region, contains prominent rovibrational lines such as the R(1,0) transition at 3.668 μm (2727 cm⁻¹), key for detections. Anharmonic effects in the significantly complicate the spectra, leading to perturbations in energy levels and the appearance of hot-band transitions from vibrationally excited states. These anharmonicities cause red-shifting of the fundamentals and mixing between modes, resulting in a dense rovibrational structure that requires variational calculations for accurate assignment. Hot bands, such as those involving low-lying excitations of ν₂, contribute additional intensity near the main fundamentals, particularly at elevated temperatures where population of these states increases. Line widths and intensities in the spectrum exhibit strong temperature dependence, with broadening occurring due to rotational relaxation and Doppler effects in sources, while intensities scale with Boltzmann populations of initial states. The and nuclear spin isomers of H₃⁺ influence branching ratios, as ortho-H₃⁺ (total nuclear spin I=3/2) occupies levels with K=3n and higher statistical weights, restricting certain transitions compared to para-H₃⁺ (I=1/2, K=3n±1). This distinction affects the relative strengths of lines in the ν₂ and ν₃ bands. High-resolution laboratory spectra of these modes have been recorded using storage ring facilities, notably in experiments at the Test Storage Ring (TSR) at MPIK during the 2000s, where action combined IR laser excitation with chemical probing to resolve individual rovibrational lines with sub-Doppler precision under cold conditions. These measurements provided benchmarks for anharmonic corrections and isomer-specific intensities, essential for modeling.

Other Spectroscopic Techniques

Due to its D_{3h} symmetry, the trihydrogen cation lacks a permanent , making pure rotational transitions in the regime forbidden by selection rules. Theoretical calculations have nonetheless predicted weak forbidden rotational spectra arising from higher-order effects, with transitions expected in the 100–500 GHz range corresponding to the ground-state rotational constant B \approx 35 , \mathrm{cm}^{-1}. These predictions provide benchmarks for potential detections in dense laboratory plasmas or astrophysical environments, though no direct experimental observations of pure rotational lines have been reported. Electronic of H_3^+ primarily involves to Rydberg states, with transitions occurring in the region above 100 . Synchrotron-based experiments have facilitated high-resolution photoabsorption measurements, revealing structured bands associated with excitations to np and nd Rydberg series converging to the limit. These studies highlight the of vibronic in broadening the spectra and provide quantitative cross sections for pathways, such as H_3^+ \to \mathrm{H_2}^+ + \mathrm{H}, with thresholds near 12 eV. Complementary experiments in ion storage rings using tunable UV and visible light (200–800 ) have mapped dissociation yields and lifetimes of excited states, confirming rapid predissociation on timescales for many Rydberg levels. These methods complement benchmarks by resolving rotational structure in electronic transitions.

Chemical Reactivity

Formation Mechanisms

The primary mechanism for the formation of the trihydrogen cation (H₃⁺) in gas-phase environments such as and laboratory plasmas is the barrierless and between and molecular :
\ce{H2+ + H2 -> H3+ + H}
This process releases approximately 1.9 eV of energy and proceeds at near the collision rate, with a measured rate constant of approximately $2 \times 10^{-9} cm³ s⁻¹ at 300 K. The reaction's efficiency stems from the strong interaction between the ionic H₂⁺ and the polarizable H₂ molecule, making it dominant where H₂⁺ and H₂ are abundant.
Laboratory measurements of this rate constant have been conducted using flowing afterglow and selected-ion flow tube (SIFT) techniques, which enable precise control of densities, temperatures, and gas flows to isolate bimolecular . These apparatuses, pioneered in the and refined over decades, confirm the reaction's temperature dependence as k(T) \approx 2.1 \times 10^{-9} (T/300)^{0.2} cm³ s⁻¹, highlighting its relevance across a wide range of conditions from to astrophysical regimes. An alternative formation route, significant in dilute regions where three-body collisions are negligible, is the radiative association of H⁺ with H₂:
\ce{H+ + H2 -> H3+ + h\nu}
This photon-emitting process stabilizes the nascent H₃⁺ complex and becomes relevant in low-density regions, with an estimated rate coefficient on the order of $10^{-16} cm³ s⁻¹, though direct measurements remain challenging due to the weak radiative yield.
Recent investigations (2021–2025) have uncovered novel pathways involving complex precursors. In 2021, an anomalous surface-mediated formation was demonstrated on water-adsorbed surfaces under intense irradiation, where bimolecular water interactions yield H₃⁺ without organic intermediaries, suggesting potential roles in catalytic or environments. A mechanism in the double ionization of , identified by researchers at , involves transient H₂ ejection followed by proton abstraction, contributing to H₃⁺ production in strong-field conditions. Extending this, a 2025 study on methyl dissociations revealed similar dynamics upon double ionization, where substituents modulate the efficiency of H₂ and subsequent H₃⁺ formation, with yields varying by up to 50% across halides. These discoveries highlight H₃⁺ formation's versatility beyond simple ion-molecule encounters.

Destruction Pathways

The primary destruction pathway for the trihydrogen cation (H₃⁺) in low-density environments is recombination with , yielding neutral products such as H₂ + H or three H atoms:
\ce{H3+ + e^- -> H2 + H}
or
\ce{H3+ + e^- -> 3H}.
This process proceeds via capture of the into a resonant state followed by rapid , with a of approximately $3 \times 10^{-7} (T/300)^{-0.5} cm³ s⁻¹, where T is the in . The decreases mildly with increasing due to reduced cross sections at higher collision energies, making recombination dominant in cold, electron-rich regions.
Proton transfer reactions with abundant neutrals represent another key loss channel, particularly in environments with trace . A representative example is the reaction with :
\ce{H3+ + HD -> H2D+ + H2},
which is exothermic by approximately 220 K at interstellar temperatures below 100 K, favoring the forward direction and suppressing the reverse reaction. This pathway initiates sequential deuteration in ion-molecule chains, converting H₃⁺ to heavier isotopologues like H₂D⁺ and beyond, with near-gas-kinetic rates on the order of 10^{-9} cm³ s⁻¹. The exothermicity enhances at low temperatures due to ortho-para nuclear spin effects, making it significant when D/H ratios exceed statistical expectations (e.g., >1/4 in depleted regions).
At elevated collision energies, such as in laboratory beams exceeding keV scales, H₃⁺ undergoes collisional dissociation with helium:
\ce{H3+ + He -> H2 + H+ + He}.
This charge-transfer dissociation arises from electronic excitation during close encounters, with cross sections peaking around 4–5 keV center-of-mass energies and leading to fragment kinetic energy distributions consistent with polar dissociation channels. Recent investigations (circa 2023) into H₃⁺ interactions with helium clusters in cryogenic traps highlight how solvation stabilizes the ion against spontaneous dissociation but can enhance loss via cluster-induced charge migration or vibrational relaxation in ultracold matrices near 1 K.

Role in Astrophysics

Detection in Planetary Atmospheres

The trihydrogen cation (H₃⁺) is particularly abundant in the auroral ionospheres of , where it dominates the positive ion composition above altitudes of approximately 200 km and plays a key role in ionospheric chemistry by facilitating charge transfer and . Observations using the Infrared Space Observatory (ISO) in 1996, followed by extensive ground-based from telescopes such as the Infrared Telescope (UKIRT) and the (VLT) through the 2000s and 2010s, have mapped H₃⁺ emissions in these regions, revealing column densities on the order of 10¹⁵ cm⁻². These detections rely on the strong rovibrational emission lines of H₃⁺ in the 3–4 μm wavelength range, which arise from ν₂ band transitions excited by particle precipitation. In Saturn's auroral atmosphere, H₃⁺ was initially identified through ground-based near-infrared observations in the early 1990s, with the Cassini spacecraft's Visual and Infrared Mapping Spectrometer (VIMS) providing detailed spectral confirmation during the 2000s, particularly from 2004 to 2017, highlighting its role as the primary ion in the upper ionosphere. Emissions from H₃⁺ in Saturn's polar regions indicate ionospheric temperatures of 500–800 K, lower than Jupiter's due to reduced precipitation energy, but still reflecting significant magnetospheric coupling. Ground-based infrared observations of in the 2010s, using instruments like North/GNIRS, IRTF/SpeX, and VLT/CRIRES, detected faint H₃⁺ emissions from the dayside , confirming its presence as a minor but chemically important species with column densities below 10¹⁴ cm⁻² and temperatures around 500–600 K. For , ground-based efforts in the 2010s, including Keck/NIRSPEC , established upper limits on H₃⁺ abundance rather than direct detections, suggesting low production rates due to the planet's distance from . Recent (JWST) observations in 2024–2025, utilizing the NIRSpec instrument, have refined H₃⁺ vertical profiles in the gas giants, revealing dynamic auroral structures and time-variable emissions on with column densities varying rapidly over seconds to minutes. On Saturn, these data highlight sub-auroral ionospheric features, while on , JWST confirmed the first definitive H₃⁺ detection in 2023 (analyzed in 2025 publications), with a globally averaged column density of (7.2 ± 1.4) × 10¹⁴ m⁻² and localized auroral enhancements up to twice that value at temperatures of 358 ± 8 K. In 's auroral zones, H₃⁺ emission line ratios from such high-resolution spectra yield ionospheric temperatures up to 1000 K, underscoring the ion's sensitivity to energetic electron precipitation.

Detection in Interstellar Clouds

The trihydrogen cation (H₃⁺) exhibits high abundances in diffuse clouds (n < 10³ cm⁻³), where it serves as a key tracer of cosmic-ray ionization due to its simple formation and destruction chemistry. Column densities in these environments typically range from 10¹⁴ to 10¹⁵ cm⁻², as measured via absorption in the ν₁ vibrational band near 3.7 μm using high-resolution infrared spectroscopy on telescopes like UKIRT and Keck from the 1990s to the 2020s. For instance, toward the sightline to Cygnus OB2 No. 12, a column density of (3.8 ± 0.6) × 10¹⁴ cm⁻² was detected, implying a path length through low-density gas exceeding 100 pc rather than elevated H₃⁺ concentrations. Similarly, observations toward ζ Persei revealed a column density of approximately 4 × 10¹⁴ cm⁻², attributed to an enhanced cosmic-ray flux about 40 times higher than average, with destruction primarily by electron recombination in these low-density regimes. These findings highlight H₃⁺'s persistence in diffuse media despite UV radiation, owing to its formation via H₂⁺ + H₂ and protection within sightlines. In contrast, H₃⁺ abundances are significantly lower in dense molecular clouds (n ≥ 10⁴ cm⁻³) because of rapid neutralization through proton-transfer reactions with abundant species like CO and H₂O, limiting column densities to ~10¹³–10¹⁴ cm⁻². Detections in such environments were achieved using infrared absorption against background continuum sources, confirming H₃⁺'s role as an initiator of even in shielded regions. These observations underscore the ion's sensitivity to cloud density, with lower levels reflecting efficient destruction pathways unavailable in diffuse settings. Isotopologue ratios, particularly H₂D⁺/H₃⁺, have been probed in interstellar clouds to investigate deuterium fractionation and the ortho/para ratio of H₂, which influences ion formation and stability. In dense clouds like , where H₂D⁺ is more readily observed via rotational lines, these ratios reveal enhanced deuterium incorporation at low temperatures (~10 K), providing constraints on the primordial D/H ratio and H₂ spin state. Recent advancements have extended H₃⁺ detections to photon-dominated regions (PDRs), where UV irradiation from nearby stars drives ion chemistry. Using and in 2024, H₃⁺ was identified in the irradiated disk d203-506 within the , a prototypical PDR environment, with column densities indicating FUV-driven formation alongside cosmic-ray contributions. Such observations in PDRs like the vicinity emphasize H₃⁺'s utility in mapping ionization layers at cloud edges.

Modeling and Predictions

Theoretical models for the trihydrogen cation (H₃⁺) in astrophysical environments primarily rely on steady-state approximations to predict its abundances by equating formation and destruction rates. In low-density (diffuse) interstellar clouds, where cosmic-ray ionization dominates and electron recombination governs destruction, the production of H₃⁺ occurs via the sequence of cosmic-ray ionization of H₂ followed by rapid reaction with H₂ to form H₃⁺, yielding a production rate of ζ n(H₂), where ζ is the cosmic-ray ionization rate (typically ~10⁻¹⁷ s⁻¹) and n(H₂) is the molecular hydrogen density. Destruction is governed by dissociative recombination with electrons, H₃⁺ + e⁻ → H + H + H, with rate coefficient α_recomb ≈ 3 × 10⁻⁷ (T/300)^{-0.5} cm³ s⁻¹. The steady-state abundance thus simplifies to n(\ce{H3+}) = \frac{\zeta n(\ce{H2})}{\alpha_\text{recomb} n(e^-)}, where n(e⁻) is the electron density, often tied to metallicities or ionization balance. This model predicts H₃⁺ abundances of ~10⁻⁸ to 10⁻⁵ relative to H₂ in shielded regions, dropping to lower values in UV-exposed zones due to increased photodissociation and electron densities from photoionization. In dense interstellar clouds, destruction is primarily by proton-transfer reactions with abundant neutrals (e.g., CO, H₂O) with rate coefficients k_pt ≈ 2 × 10^{-9} cm³ s⁻¹, leading to n(H₃⁺) ≈ ζ / k_pt ≈ 10^{-9} to 10^{-8} cm^{-3}, largely independent of density. In photodissociation regions (PDRs) and diffuse clouds, more comprehensive models incorporate spatial gradients in UV radiation and density. The UCL_PDR code, a one-dimensional model solving time-dependent or steady-state chemistry with over 1700 reactions, computes H₃⁺ abundances as a function of visual extinction A_V, showing peaks near the H/H₂ transition (A_V ≈ 1–2 mag) where ionization balances recombination. These models predict higher H₃⁺ in high-density (n_H > 10⁴ cm⁻³) PDRs with intense radiation fields (G₀ > 10³), reaching fractional abundances up to 10⁻⁷. Incorporation of ortho/para nuclear spin isomers for H₃⁺ and H₂, along with isotopologues like H₂D⁺, refines predictions by accounting for spin-dependent reaction rates and fractionation effects; for instance, para-H₃⁺ forms preferentially at low temperatures (<30 K), influencing observed rotational populations in cold clouds. Steady-state ortho/para ratios approach statistical limits (1:2) in equilibrated plasmas but deviate in non-equilibrium conditions, as modeled with statistical weights for formation channels. Historical discrepancies between modeled and observed H₃⁺ column densities in diffuse clouds (often underpredicting by factors of 2–10) have been partially resolved by incorporating alternative formation pathways identified in recent laboratory studies. In 2025, investigations revealed roaming mechanisms in doubly ionized methanol (CH₃OH²⁺) and methyl halides (e.g., CH₃Cl²⁺), where an H₂ fragment abstracts a proton from a roaming counterpart, yielding H₃⁺ with efficiencies up to 67% relative to primary channels. These pathways, proceeding on singlet potential energy surfaces with timescales ~200–350 fs, provide additional sources in ion-rich interstellar environments, potentially boosting overall H₃⁺ yields by 20–50% under conditions with abundant organics, thus aligning models with observations in UV-irradiated regions. Predictions for H₃⁺ in exoplanet atmospheres emphasize its role as a thermospheric coolant via infrared emission lines (e.g., ν₂ band at 3–4 μm), particularly in hot Jupiters receiving high stellar EUV flux. Non-equilibrium photochemical models forecast H₃⁺ volume mixing ratios of 10⁻⁹ to 10⁻⁶ in the upper atmospheres of close-in gas giants like HD 209458b, with emission luminosities 10²–10⁵ times Jupiter's (~10¹⁶–10¹⁹ W), enabling detection as a tracer of ionospheric dynamics and escape processes. Future observations with the Extremely Large Telescope (ELT), expected operational in the 2030s, are anticipated to resolve H₃⁺ lines in high-resolution near-IR spectra (R > 10⁵), facilitating direct mapping of auroral and thermospheric structures in ~10–20 nearby exoplanets.

References

  1. [1]
    A simple approximate solution for the H3+ ion - Wiley Online Library
    Sep 27, 2022 · Experimentally, gas-phase is of equilateral geometry with a H H bond length of R ≈ 0.90 Å and it has a Raman-active breathing mode at ν1 = ...
  2. [2]
    Interstellar H3+ - PNAS
    Within a decade and a half of the discovery of H3+ by J. J. Thomson in 1911 (1), it was clear that H3+ was the most abundant ion in ordinary hydrogen discharges ...
  3. [3]
  4. [4]
    Hydrogen molecular ions: H3+, H5+ and beyond - PubMed Central
    Aug 5, 2019 · The H3+ molecular ion is the simplest stable molecular ion of hydrogen. It is rapidly formed by collisions between H2 and H+2. ... Use of H3+ to ...
  5. [5]
    Chemistry, astronomy and physics of H 3 + - Journals
    Nov 13, 2012 · The great developments in the chemistry, astronomy and physics of H3+ since 2006, which have led to this Royal Society Theo Murphy Meeting, ...
  6. [6]
    [PDF] Rays of Positive Electricity
    By Sir J. J. THOMSON *. FIND that the investigation of the Positive Rays or. Canalstrahlen is made much easier by using very large vessels ...
  7. [7]
    To be or not to be: the early history of H 3 and H 3 - Journals
    Nov 13, 2012 · Triatomic hydrogen became a major research area only after 1980, but its history goes back to J. J. Thomson's discovery in 1911.
  8. [8]
    Trihydrogen is the most important ion you've never heard of - C&EN
    Jun 8, 2019 · The highly reactive H3+, however, is primed to donate a proton to anything it stumbles into. The little molecule enriches the chemistry of the ...
  9. [9]
    Analysis of the mechanism of reaction of H3+ [tritium ion] with ...
    Photoionization and ion cyclotron resonance studies of the ion chemistry of ethylene oxide. ... T.B. McMahon, P.G. Miasek, J.L. Beauchamp. Trapped ion ...
  10. [10]
    Ion Cyclotron Resonance Power Absorption: Collision Frequencies ...
    Sep 1, 1971 · Pulsed ICR techniques are used to obtain the instantaneous power absorption curves for N 2 + , CO 2 + ⁠, and H 3 + ions in their parent gases as ...
  11. [11]
    Destruction Rate of H3+ by Low-Energy Electrons Measured in a ...
    As an attempt to resolve this uncertainty, the cross section for dissociative recombination of vibrationally cold H3+ has been measured with an ion storage ring ...
  12. [12]
    Chemistry, astronomy and physics of H3+ - PMC - PubMed Central
    The great developments in the chemistry, astronomy and physics of H3+ since 2006, which have led to this Royal Society Theo Murphy Meeting, are reviewed.Missing: trihydrogen cation
  13. [13]
    Reactions of H 20+ and D 20+ with Molecular Hydrogen. I. Proton ...
    Nov 15, 1972 · Frankevich and Tal'rose observed Reaction (1), but their failure to observe the exothermic reaction. (2) resulted in the range 2.65<PA(H2) ...
  14. [14]
    Detection of H 3 + on Jupiter - Nature
    Aug 17, 1989 · In the magnetosphere, the H+3 ion, probably of ionospheric origin, was detected in situ by the Voyagers4. Infrared emissions were observed in ...Abstract · Author Information · Authors And Affiliations
  15. [15]
  16. [16]
    Neural Network Representation of Three-State Quasidiabatic ...
    Mar 1, 2021 · bond length of H3+ is predicted to be 0.8732(2) Å. 45. Röhse, R ... (ab initio) diabatic PESs of H3+. We have depicted the convergence ...
  17. [17]
    An accurate, global, ab initio potential energy surface for the H+ 3 ...
    Aug 10, 2025 · gies have an absolute accuracy of about 0.05 cm. at high energies we use the ab initio data of Schinke et al. 75 000 cm. up to 70 000 c mabove ...
  18. [18]
    Theoretical calculations of the hyperfine structure in the spectra of ...
    The present paper reports a purely theoretical calculation of the hyperfine structure in the rotation-vibration spectra of H+3, D+3, H2 D+ and HD+.Missing: cation isotopologues geometry<|control11|><|separator|>
  19. [19]
    Zero point energy of H - Astronomy & Astrophysics (A&A)
    Zero point energy of H. +. 3 and its deuterated isotopomers. J. Ramanlal, O. L. Polyansky. , and J. Tennyson. Department of Physics and Astronomy, University ...Missing: H3+ | Show results with:H3+
  20. [20]
    Ab Initio calculation of near-equilibrium potential ... - AIP Publishing
    The equilibrium bond length ofH3+ is predicted to be 0.8732(2) A. I. INTRODUCTION. H3+' the simplest polyatomic molecule, has been at- tracting the interest of ...
  21. [21]
    H3 +: Ab initio calculation of the vibration spectrum - AIP Publishing
    Nov 1, 1976 · The ground state is found to have a zero‐point energy of 4345 cm−1 and a vibrationally averaged geometry of R1=R2=0.91396 Å, ϑ=60.0012°, where ϑ ...Missing: D3+ | Show results with:D3+
  22. [22]
    A High-precision Survey of the D/H Ratio in the Nearby Interstellar ...
    Mar 24, 2023 · We present high signal-to-noise ratio measurements of the H i Lyα absorption line toward 16 Galactic targets that are at distances between approximately 190 ...
  23. [23]
    On the rovibrational levels of the H3+ and H2D+ molecules
    Aug 10, 2025 · For H3 + calculated ground state rotational constants are B0 = 43.51 (43-57) cm -1, Co = 20.59 (20-71) cm -1, Dj ° = 0.04 (0.05) cm -1, Djn° = - ...
  24. [24]
    L: high-resolution line lists of H3+, H2D+, D2H+, and D3+
    As symmetric tops, the molecules H 3 + and D 3 + are defined by two primary vibrational modes, symmetric stretching (ν1) and bending (ν2). The bending mode is ...
  25. [25]
  26. [26]
  27. [27]
    The vibration-rotation spectrum and anharmonic potential of H3+
    Some qualitative effects of anharmonicity on the spectra of H3+ and D3+ between low vibrational levels are described. Using large-basis vibration-rotation ...
  28. [28]
    Action spectroscopy and temperature diagnostics of - AIP Publishing
    Infrared absorption spectroscopy of few hundred H3+ ions trapped in a 22-pole ion trap is presented using chemical probing as a sensitive detection ...Missing: MPIK | Show results with:MPIK
  29. [29]
    Electronic structure of Rydberg states of triatomic hydrogen, neon ...
    Theoretical calculations on electronic transitions for H3, including Rydberg and transition state spectra. The Journal of Chemical Physics 1988, 89 (11) ...Missing: H3+ | Show results with:H3+
  30. [30]
    Mechanisms and time-resolved dynamics for trihydrogen cation (H3 ...
    Jul 5, 2017 · Trihydrogen cation formation from organic molecules is one such case that requires multiple bonds to break and form.Missing: length | Show results with:length
  31. [31]
    Thirty years of H + 3 astronomy - Physical Review Link Manager
    Aug 3, 2020 · The general consensus before the detection of H 3 + was that it was particles precipitating from this torus that gave rise to the main ...
  32. [32]
    H3++H2 isotopic system at low temperatures: Microcanonical model ...
    Apr 22, 2009 · State-to-state thermal rate coefficients for reactions of all H3++H2 isotopic variants are derived and compared to new experimental data.
  33. [33]
    A flowing afterglow selected ion flow tube (FA/SIFT) comparison of ...
    Aug 6, 2025 · A flowing afterglow-selected ion flow drift tube has been used to measure the rate coefficients and product ion distributions for reactions of ...
  34. [34]
    Flowing-afterglow study of electron-ion recombination of para– - H 3 +
    Jul 22, 2015 · The measurements with SA-CRDS at 77 K showed that the binary recombination rate coefficient pαbin for para–H 3 + is at 77 K at least three ...
  35. [35]
    Radiative association of H+ H2 - experimental study - IOP Science
    This clearly indicates domination of radiative association at low hydrogen densities. The obtained dependence is fitted with function keff = kr + k3[H2]. 1.Missing: H3+ | Show results with:H3+
  36. [36]
    Anomalous formation of trihydrogen cations from water on ... - PubMed
    Jun 22, 2021 · We present results on forming [Formula: see text] from bimolecular reactions that involve only an inorganic molecule, namely water, without the presence of any ...Missing: surface H3+ tungsten
  37. [37]
    H2 roaming chemistry and the formation of H3+ from organic ...
    Dec 5, 2018 · The fundamental mechanistic findings presented here explore the formation of H 3 + , the most important ion in interstellar chemistry, through H 2 roaming ...
  38. [38]
    Factors governing H 3 + formation from methyl halogens and ...
    Jan 6, 2025 · The formation of H 3 + ions within the interstellar medium proceeds via the Hogness and Lunn mechanism, H 2 + + H 2 → H 3 + + H , where H 2 + is ...Missing: H2 + | Show results with:H2 +
  39. [39]
  40. [40]
    The Deuteration Ladder of H3+ - NASA ADS
    Deuteration of H3+ begins with the reaction H3+ + HD -> H2D+ + H2 which is exothermic by 220 K (therefore, the backward reaction is slow in the cold ISM) and ...Missing: → | Show results with:→
  41. [41]
    Physical conditions in the central molecular zone inferred by H 3
    molecules are formed in their para (1, 1)/ortho (1, 0) ground states, in the ratio of their statistical weights. Figure 9 shows the N(3, 3)/N(1, 1) ratios ...
  42. [42]
    The astrochemistry of H 3 + - Journals
    For example, the reaction of H3+ with atomic O leads, eventually, to the production of water, while the reaction with HD leads to the production of a wide ...
  43. [43]
  44. [44]
    [PDF] arXiv:2306.12056v1 [physics.atm-clus] 21 Jun 2023
    Jun 21, 2023 · The mass spectra of ionized Cyt and Thy clusters and microhy- drated Cyt/Thy clusters in helium droplets deviate from clusters formed in a ...
  45. [45]
    Spectro-imaging observations of H3+ on Jupiter - PubMed
    The spectra include a total of 14 H3+ lines, including two hot lines from the 3v2-v2 band, detected on Jupiter for the first time. They can be used to determine ...Missing: ISO satellite
  46. [46]
    Mapping the H 3 + aurora on Saturn from Cassini/VIMS observations
    Previously, H3+ was detected on Saturn from the ground in spectroscopy by Geballe et al (ApJ, 1993), and found to be much fainter than Jupiter emission. The ...
  47. [47]
    Local-time averaged maps of H3+ emission, temperature and ion ...
    Aug 5, 2019 · We present Keck-NIRSPEC observations of Saturn's H3+ aurora taken over a period of a month, in support of the Cassini mission's 'Grand ...
  48. [48]
    Post-equinoctial observations of the ionosphere of Uranus
    We analyse ground-based infrared observations of H 3 + emission from the upper atmosphere of Uranus using Gemini North/GNIRS, NASA IRTF/SpeX and VLT/CRIRES.
  49. [49]
    JWST observations of Jupiter's time variable H3+ auroral emissions
    We show that during transient brightening events column density rapidly increases (over ~30s) while the temperature decreases, suggesting that brightening is ...Missing: cation detection ISO telescope
  50. [50]
    Discovery of H 3 + and infrared aurorae at Neptune with JWST
    Mar 26, 2025 · Here, using observations from the James Webb Space Telescope, we detect H 3 + at Neptune, as well as distinct infrared southern auroral ...
  51. [51]
    Mapping H3+ Temperatures in Jupiter's Northern Auroral ...
    Jul 10, 2018 · We present a detailed study of the H 3 + auroral emissions at Jupiter, using data taken on 31 December 2012 with the long-slit Echelle spectrometer CRIRES (ESO ...3 Data Analysis · 4 Results · 5 DiscussionMissing: ISO | Show results with:ISO
  52. [52]
    [1203.6813] Using deuterated H3+ and other molecular species to ...
    Abstract page for arXiv paper 1203.6813: Using deuterated H3+ and other molecular species to understand the formation of stars and planets.Missing: Thirty years astronomy<|control11|><|separator|>
  53. [53]
    PDRs4All XV. CH radical and H3+ molecular ion in the irradiated ...
    tion, and the vibrational excitation temperature. ... ν1, ν2, ν3, ν4. 1,2,3. SH ν1. 17. HeH. + ν1. 14. HCO. + ... ro-vibrational transitions are ∆J = 0 or ±1, and ...
  54. [54]
    H 3 + in diffuse interstellar gas - Oxford Academic
    (1996) for five lines of sight, a ratio that agrees with the value of 1.3×10−4 found for ζ Ophiuchi (Snow & Witt 1996). For oxygen we adopt 3.3×10−4 (Snow ...
  55. [55]
    None
    Below is a merged summary of the UCL_PDR code modeling of H3+, including ortho/para states and isotopologues, consolidating all information from the provided segments into a single, dense response. To maximize detail retention, I’ve organized key information into a table where appropriate, supplemented by narrative text for clarity and completeness.
  56. [56]
    The ortho : para ratio of H 3 + in laboratory and astrophysical plasmas
    Nov 13, 2012 · In diffuse molecular clouds, the nuclear spin temperature of (approx. 30 K) is much lower than the cloud kinetic temperature (approx. 70 K).
  57. [57]
    A CRIRES-search for H3+ emission from the hot Jupiter atmosphere ...
    We calculate the emission limit for the H3+ line at 3953.0 nm [Q(1,0)] to be 8.32 × 1018 W and a limit of 5.34 × 1018 W for the line at 3985.5 nm [Q(3,0)].