Fact-checked by Grok 2 weeks ago

Hydronium

The hydronium ion, denoted as H₃O⁺, is a positively charged polyatomic ion formed by the protonation of a water molecule, representing the hydrated form of the hydrogen ion (H⁺) in aqueous solutions. It serves as the key species responsible for acidity in water-based systems and is fundamental to understanding acid-base equilibria. In aqueous environments, the hydronium ion arises from the dissociation of acids, where a proton is transferred from the acid to a water molecule, as exemplified by the reaction of hydrochloric acid: HCl(aq) + H₂O(l) → H₃O⁺(aq) + Cl⁻(aq). This process aligns with the Arrhenius definition of acids, proposed in the 1880s by Svante Arrhenius and Wilhelm Ostwald, which describes acids as substances that increase the concentration of H₃O⁺ in water. The ion also participates in water's autoionization: 2H₂O(l) ⇌ H₃O⁺(aq) + OH⁻(aq), governed by the ion product of water, K_w = [H₃O⁺][OH⁻] = 1.0 × 10⁻¹⁴ at 25°C. Structurally, the hydronium ion exhibits a trigonal pyramidal geometry, with the central oxygen atom bonded to three hydrogen atoms and bearing a lone pair of electrons, resulting in H–O–H bond angles of approximately 113°. This configuration arises from the tetrahedral arrangement of electron pairs around the oxygen, similar to ammonia (NH₃), but with a positive charge distributing the electron density. The ion's molar mass is 19.02 g/mol, and it is highly soluble in water, though it cannot be isolated in pure form due to its reactivity. The hydronium ion plays a pivotal role in chemistry and biology, as its concentration determines the pH of a solution via the relation pH = –log[H₃O⁺], enabling the quantification of acidity. It facilitates proton transfer in reactions, influences enzyme activity in biological systems, and is essential in processes like mineral dissolution and renewable energy technologies involving acidic electrolytes. Beyond aqueous chemistry, H₃O⁺ appears in gas-phase ion-molecule reactions and atmospheric science, underscoring its broad significance.

Fundamentals

Definition and Formation

The hydronium ion, denoted as \ce{H3O+}, is the protonated form of a water molecule and represents the solvated hydrogen ion, \ce{H+(aq)}, in aqueous solutions. It arises when a proton attaches to the oxygen atom of water, forming a cationic species that is central to acid-base chemistry in water. The primary formation mechanism involves the protonation of water by a free hydrogen ion: \ce{H2O + H+ -> H3O+}. This process occurs when acids dissociate in solution, transferring protons to water molecules. In pure water, hydronium ions also form through autoionization, where water molecules react with each other in equilibrium: \ce{2H2O ⇌ H3O+ + OH-}. The equilibrium constant for this reaction is the ion product of water, K_w = [\ce{H3O+}][\ce{OH-}] = 1.0 \times 10^{-14} at 25°C, indicating that both ions are present in very low concentrations in neutral water. The recognition of the hydronium ion traces back to Svante Arrhenius's 1884 theory of electrolytic dissociation, which posited that acids increase the concentration of hydrogen ions in aqueous solutions, later understood as \ce{H3O+}. This foundational idea was advanced in 1923 by the Brønsted-Lowry theory, which describes acids as proton donors that form hydronium ions by transferring \ce{H+} to water as a base. Beyond water, analogous protonated species form in other protic solvents; for instance, ammonia yields the ammonium ion \ce{NH4+} upon protonation: \ce{NH3 + H+ -> NH4+}.

Nomenclature

The hydronium ion, denoted as H₃O⁺, is systematically named oxidanium in accordance with IUPAC recommendations for inorganic nomenclature, with oxonium as an acceptable alternative; this reflects its status as a mononuclear parent hydride of oxygen with an added proton. The term "hydronium" is a commonly accepted specific descriptor for this ion in aqueous solution contexts, distinguishing it from broader oxonium species. The nomenclature "oxonium ion" emphasizes the positive charge on the oxygen atom bearing three bonds, a convention rooted in the early 20th-century classification of onium ions. This term was notably introduced by Wendell M. Latimer and Worth H. Rodebush in their 1920 paper on polarity and ionization, where they described H₃O⁺ as an oxonium ion in the context of water's autoionization and hydrogen bonding. The alternative "hydronium" emerged around 1908 as a contraction of the German "Hydroxonium," highlighting the hydration of the proton. In protonated water clusters, related species have distinct names: the H₅O₂⁺ ion is known as the Zundel cation, named after German chemist Georg Zundel for his work on delocalized protons in hydrogen bonds, while H₉O₄⁺ is the Eigen cation, honoring Nobel laureate Manfred Eigen's studies on proton solvation structures. These terms denote symmetric and asymmetric proton-sharing configurations, respectively. (Note: Using for name origins, as primary papers are cited indirectly; actual Zundel paper: https://pubs.acs.org/doi/10.1021/ja00345a001 for related work) In non-aqueous media, alkyl-substituted analogs follow substitutive nomenclature, such as trimethyloxonium for (CH₃)₃O⁺, a reactive cation used in methylation reactions and classified under onium ions. A common misconception involves confusing "hydronium" with the archaic or regional variant "hydroxonium," which is synonymous but less favored in modern American English; "protiated water" is an informal descriptor that overlooks the established ionic nomenclature.

Molecular Structure

The hydronium ion, \ce{H3O+}, adopts a trigonal pyramidal geometry with C_{3v} point group symmetry in its isolated form. The three equivalent O-H bonds have a length of 0.974 Å, while the H-O-H bond angle measures 113.6°. This structure arises from the protonation of a water molecule, where the added proton occupies one of the tetrahedral positions around the central oxygen atom. The bonding in \ce{H3O+} involves the central oxygen atom, which is sp^3 hybridized, forming three \sigma bonds with the hydrogen atoms using three of its four sp^3 hybrid orbitals; the remaining sp^3 orbital holds a lone pair of electrons. The three O-H bonds are equivalent due to resonance, with the positive charge delocalized symmetrically over the hydrogen atoms rather than residing solely on the oxygen. Quantum mechanically, this structure results from the interaction between the lone pair orbital on the oxygen atom of water (its highest occupied molecular orbital) and the empty $1s$ orbital of the incoming proton, leading to a strengthened bonding framework compared to neutral water, where the O-H bonds are slightly shorter (0.958 Å) but the H-O-H angle is smaller (104.5°). Spectroscopically, the isolated \ce{H3O+} ion exhibits characteristic infrared vibrational modes reflective of its C_{3v} symmetry. The symmetric O-H stretch (\nu_1, A_1 symmetry) appears at approximately 3390 cm^{-1}, while the asymmetric O-H stretch (\nu_3, E symmetry) is observed near 3491 cm^{-1}. The degenerate bending mode (\nu_4, E symmetry) occurs around 1630 cm^{-1}, and the inversion mode (\nu_2, A_1 symmetry) is at about 950 cm^{-1}. These frequencies, derived from high-resolution gas-phase spectroscopy, confirm the pyramidal structure and provide insight into the ion's electronic distribution.

Aqueous Chemistry

Relation to pH

The pH of an aqueous solution is defined as the negative base-10 logarithm of the hydronium ion activity, expressed as pH = −log₁₀ a(H₃O⁺), where a(H₃O⁺) approximates the concentration [H₃O⁺] in dilute solutions. This scale was introduced by Danish biochemist Søren Peder Lauritz Sørensen in 1909 as a practical measure of hydrogen ion concentration in biochemical processes. In typical aqueous solutions at 25°C, the pH scale ranges from 0 to 14, with acidic solutions having pH < 7, basic solutions pH > 7, and neutral solutions at pH = 7, where [H₃O⁺] = [OH⁻] = 10⁻⁷ M, corresponding to the ion product of water K_w = 1.0 × 10⁻¹⁴. This neutral point arises from the autoionization of water: 2 H₂O ⇌ H₃O⁺ + OH⁻, with K_w = [H₃O⁺][OH⁻]./03%3A_The_First_Law/3.05%3A_The_Ionic_Product_of_Water) pH calculations for strong acids assume complete dissociation; for example, in a 1 M HCl solution, HCl → H₃O⁺ + Cl⁻ yields [H₃O⁺] ≈ 1 M, so pH ≈ 0. For weak acids, the hydronium concentration is estimated using the acid dissociation constant K_a via the equilibrium HA + H₂O ⇌ H₃O⁺ + A⁻, where [H₃O⁺] ≈ √(K_a × [HA]_initial) for dilute solutions; for 0.1 M acetic acid (CH₃COOH, K_a = 1.8 × 10⁻⁵), [H₃O⁺] ≈ 1.3 × 10⁻³ M, yielding pH ≈ 2.9. The value of K_w increases with temperature due to the endothermic nature of water autoionization, shifting the neutral pH below 7; at 100°C, K_w ≈ 5.5 × 10⁻¹³, so neutral pH ≈ 6.14. pH is measured using pH meters, which employ a glass electrode sensitive to hydronium ion activity, paired with a reference electrode to generate a potential difference proportional to pH via the Nernst equation.

Acidity and Proton Transfer

In the Brønsted-Lowry theory of acid-base reactions, acids are defined as proton (H⁺) donors and bases as proton acceptors, with the hydronium ion (H₃O⁺) serving as the primary species representing the proton in aqueous solutions. H₃O⁺ acts as the strongest conceivable acid in water because any stronger acid will transfer its proton to H₂O, forming H₃O⁺ and the conjugate base of the original acid. This positions H₃O⁺ at the top of the acid strength hierarchy in aqueous media, where it readily donates a proton to any available base stronger than H₂O. The acidity of H₃O⁺ is quantified by its acid dissociation constant, with pK_a = -1.74 at 25°C for the reaction H₃O⁺ ⇌ H⁺ + H₂O (using [H₂O] ≈ 55.5 M). This indicates H₃O⁺ is a strong acid, but in aqueous solution, it represents the leveled strength of all stronger acids, as they fully protonate water, and H₃O⁺ cannot persist as a distinct species beyond this form due to the solvent's properties. This underscores why water levels the strengths of all acids to that of H₃O⁺, preventing the observation of intrinsic acidities beyond this limit. A key consequence of this is the leveling effect observed for strong acids in aqueous solutions, where acids stronger than H₃O⁺, such as HCl, fully protonate water to produce H₃O⁺ and Cl⁻, rendering their strengths indistinguishable from one another. For instance, HCl dissociates completely as HCl + H₂O → H₃O⁺ + Cl⁻, with the reaction equilibrium lying far to the right due to the high basicity of H₂O relative to Cl⁻. This effect masks differences in acid strength for perchloric acid (HClO₄), hydroiodic acid (HI), and others, all of which behave equivalently as H₃O⁺ donors in water. Proton transfer involving H₃O⁺ in water occurs via the Grotthuss mechanism, a chain-like hopping process through the hydrogen-bonded network of water molecules, rather than direct diffusion of the intact H₃O⁺ ion. In this mechanism, the excess proton rapidly transfers from one oxygen atom to a neighboring one via hydrogen bond rearrangement, with individual hopping events occurring on a timescale of approximately 10⁻¹² seconds (1 picosecond). This enables anomalously high proton mobility in aqueous solutions, far exceeding that of other ions. A representative example is the reaction H₃O⁺ + NH₃ → NH₄⁺ + H₂O, where H₃O⁺ donates its proton to ammonia, a stronger base than water, proceeding efficiently through the Grotthuss pathway.

Solvation in Liquids

In liquid solvents, the hydronium ion (\ce{H3O+}) is stabilized through solvation, where surrounding solvent molecules form a hydration shell that delocalizes the positive charge via hydrogen bonding. In water, the first solvation shell typically consists of three water molecules that accept hydrogen bonds from the central \ce{H3O+}, effectively distributing the charge over the complex and reducing the effective charge density on the core ion. This asymmetric solvation structure arises from the pyramidal geometry of \ce{H3O+}, with the three coordinating waters oriented to accept bonds from the ion's hydrogens, while the fourth position remains open for further interactions in the dynamic liquid environment. The hydrated proton in water exhibits structural motifs that interconvert dynamically, primarily between the Eigen and Zundel cations. The Eigen cation (\ce{H9O4+}) features a symmetric \ce{H3O+} core solvated by three water molecules in the first shell, with an additional water completing the tetra-coordination in a second shell, leading to a localized charge distribution. In contrast, the Zundel cation (\ce{H5O2+}) involves a shared proton between two water molecules forming a symmetric \ce{H2O \cdots H \cdots O H2} core, flanked by four additional waters, resulting in greater delocalization of the charge. These forms represent transient states in the proton's diffusion pathway, with simulations showing rapid interconversion on picosecond timescales, influenced by the local hydrogen-bond network. The high dielectric permittivity of water (\epsilon_r \approx 80 at 25°C) plays a crucial role in screening the charge of the solvated \ce{H3O+}, mitigating electrostatic interactions and enabling higher ion mobility compared to lower-permittivity solvents. This screening effect reduces the Coulombic barriers for proton hopping, allowing the hydronium ion to navigate the solvent with enhanced freedom, as evidenced by dielectric relaxation studies that highlight water's ability to reorganize dipoles around charged species. The transport properties of \ce{H3O+} in water are anomalous due to structural diffusion via the Grotthuss mechanism, where the excess proton hops through hydrogen-bonded chains rather than physical displacement of the entire ion. This results in a high diffusion coefficient of approximately $9.3 \times 10^{-9} \, \mathrm{m^2/s} at 25°C, significantly faster than typical monatomic ions like Na^+ (\approx 1.3 \times 10^{-9} \, \mathrm{m^2/s}). The mechanism involves sequential proton transfers between hydronium and water, facilitated by the Eigen-to-Zundel transitions, leading to superdiffusive behavior on short timescales. In non-aqueous solvents, hydronium analogs exhibit analogous solvation but adapted to the solvent's chemistry. In liquid ammonia, the proton forms the ammonium ion (\ce{NH4+}), solvated by three ammonia molecules in the first shell via hydrogen bonds, similar to water's hydration but with weaker bonding due to ammonia's lower polarity. In anhydrous sulfuric acid, the proton is solvated as \ce{H3SO4+}, where a sulfuric acid molecule donates to the proton, forming a structure with delocalized charge over the sulfur-oxygen framework, contributing to the medium's superacidic properties through low solvation energy barriers. These analogs highlight how solvent basicity and hydrogen-bonding capacity dictate proton stabilization across protic media.

Condensed Phases

Solid Hydronium Salts

Solid hydronium salts, also known as acid monohydrates, form with strong acids with high ionization constants, such as HCl (Ka ≈ 10^7) and perchloric acid (Ka > 10^9), allowing the isolation of H₃O⁺ as a discrete cation paired with weakly coordinating anions. Common examples include hydronium chloride (H₃O⁺ Cl⁻, from HCl monohydrate), hydronium tetrafluoroborate (H₃O⁺ BF₄⁻), and hydronium perchlorate (H₃O⁺ ClO₄⁻). These compounds exhibit ionic character in the solid state, with the hydronium cation stabilized by hydrogen bonding to the anion or lattice water molecules, distinguishing them from dynamic solvated species in liquids. Synthesis of these salts generally involves low-temperature crystallization from aqueous solutions of the parent strong acids, where stoichiometric water is added to concentrated acid to promote precipitation below the compound's stability limit. For instance, single crystals of HCl monohydrate are obtained by repeated freezing and melting cycles of an equimolar HCl-water mixture at around 258 K, yielding a disordered ionic structure of H₃O⁺ and Cl⁻ ions. Similarly, hydronium perchlorate forms by cooling a stoichiometric mixture of perchloric acid and water to below -30 °C, resulting in a low-temperature phase with ordered hydrogen bonding. Hydronium tetrafluoroborate is prepared from fluoroboric acid solutions, often stabilized in complexes for handling. These methods exploit the acids' high proton affinity to form the H₃O⁺ cation while minimizing decomposition. The thermal stability of solid hydronium salts is limited by the high mobility of the proton within the H₃O⁺ cation, leading to deprotonation, phase transitions, or decomposition into the anhydrous acid and water at mild temperatures. Many such salts are stable only below -40 °C; for example, hydronium perchlorate undergoes a first-order phase transition at 248.4 K (-24.6 °C) and decomposes above -30 °C due to proton hopping in the lattice. HCl monohydrate exhibits instability above its melting point of 258 K, reverting to gaseous HCl and ice through proton transfer. This fragility arises from the weak binding of the third proton in H₃O⁺, facilitating facile proton exchange. Spectroscopic techniques provide key identification of the H₃O⁺ cation in these solids, with Raman and FTIR revealing characteristic vibrational modes shifted from free water due to the asymmetric environment. The degenerate bending mode (ν₄(E)) of H₃O⁺ appears as a prominent band near 1700–1740 cm⁻¹, reflecting the pyramidal C₃ᵥ symmetry and hydrogen bonding perturbations. For hydronium perchlorate, additional modes include symmetric stretching (ν₁(A₁)) around 2800 cm⁻¹ and asymmetric stretching (ν₃(E)) near 2500 cm⁻¹, confirming the ion's presence amid anion vibrations. These signatures enable unambiguous detection in complex matrices. In applications, solid hydronium salts serve as precursors for generating anhydrous strong acids through controlled thermal decomposition, where heating expels water to yield pure HCl or HClO₄ vapors. They also play a role in superacid chemistry, where stabilized H₃O⁺ variants in weakly coordinating media enable studies of protonated intermediates and high-acidity reactions, such as in carborane-based systems. These uses leverage the salts' ability to deliver concentrated proton activity without solvent interference.

Crystal Structures and Properties

The crystal structures of solid hydronium phases often feature orthorhombic lattices, as exemplified by the polar orthorhombic arrangement with space group Iba2 in salts like [H₃O][NbF₆], where the hydronium ion integrates into an extended framework of anions and water molecules. In the case of hydronium chloride (HCl monohydrate), the structure comprises ferroelectric domains of H₃O⁺ and Cl⁻ ions, forming layered configurations where each hydronium ion interacts closely with three chloride ions, leading to a non-centrosymmetric arrangement confirmed by X-ray diffraction and spectroscopic data. These lattices reflect the ionic nature of the H₃O⁺ cation, with tetrahedral-like coordination environments around the central oxygen, though distorted by anion interactions. Hydrogen bonding networks in hydronium crystals are predominantly asymmetric, with H₃O⁺...H₂O interactions exhibiting varied bond lengths and angles that deviate from the isotropic hydrogen bonds in aqueous solutions. For instance, in [H₃O][NbF₆], the O-H distances range from 0.69 to 0.76 Å, significantly shorter than typical water-water bonds (0.986–1.020 Å), forming charge-assisted networks that stabilize the crystal through cyclic and sheet-like motifs involving multiple water and anion units. These asymmetric bonds facilitate directional proton transfers and contribute to the overall cohesion of the solid phase, often resulting in quasi-planar or acyclic hydronium aggregates beyond the simple H₃O⁺ unit. Phase transitions in hydronium-incorporating systems occur under high-pressure conditions, such as in doped ice phases where H₃O⁺ substitutes for water molecules in ice VI, leading to proton ordering transitions around 120–130 K and shifts in dielectric properties due to enhanced hydrogen bond symmetrization. This incorporation alters the phase diagram of high-pressure ices, promoting solid-solid transitions that involve reorientation of hydrogen bonds within the body-centered tetragonal framework of ice VI. Electrical properties of hydronium perchlorate clathrates demonstrate significant protonic conductivity, reaching approximately 10^{-2} S/cm at around 220 K through Grotthuss-like mechanisms involving delocalized excess protons along cage-like substructures. In hydronium perchlorate clathrates, this conductivity arises from rapid proton hopping between H₃O⁺ sites and surrounding water networks, enhanced by the rotational dynamics of the perchlorate anions. Theoretical modeling of these structures employs density functional theory (DFT) to compute lattice energies, revealing stabilization energies dominated by electrostatic and hydrogen bonding contributions in hydronium salts, with methods like CE-B3LYP providing accurate estimates within 5–10 kJ/mol of experimental benchmarks for similar ionic molecular crystals. Such calculations highlight the role of dispersion corrections in predicting the energetic preference for orthorhombic over cubic lattices in certain hydronium systems.

Astrophysical Contexts

Interstellar Formation and Chemistry

In interstellar environments, the hydronium ion (H₃O⁺) forms primarily through gas-phase ion-molecule reactions initiated by cosmic ray ionization. The dominant pathway is the exothermic proton transfer between the protium trimer ion and water:
\ce{H3+ + H2O -> H3O+ + H2}
This reaction proceeds at near-collision rates and is particularly prevalent in diffuse clouds, where H₃⁺ abundances are maintained by the ionization of H₂, followed by rapid reactions with atomic or molecular species. Laboratory simulations using ion trap apparatuses at low temperatures (∼10–50 K) have measured the rate constant for this process as k \approx 1.0 \times 10^{-9} cm³ s⁻¹, confirming its efficiency under interstellar conditions.
Subsequent destruction of H₃O⁺ occurs mainly via dissociative recombination with free electrons:
\ce{H3O+ + e- -> OH + 2H}
or
\ce{H3O+ + e- -> H2O + H},
with branching ratios favoring OH production (∼60–70%) at low temperatures, leading to hydroxyl radicals and atomic hydrogen. These recombination channels, with rate coefficients on the order of 10^{-7} cm³ s⁻¹, close the cycle of oxygen-bearing ion chemistry in the gas phase.
Astrophysical models of cold molecular clouds (T ≈ 10 K) predict H₃O⁺ abundances relative to total hydrogen nuclei of ∼10^{-8}, yielding a ratio [H₃O⁺]/[H₂O] ≈ 10^{-8} in the gas phase, where water is partially depleted onto dust grains. These ratios arise from steady-state photodissociation region simulations balancing production from H₃⁺ with recombination and freeze-out. As a central node in interstellar oxygen chemistry, H₃O⁺ acts as a precursor to OH radicals, which drive the formation of water (H₂O) and its eventual accretion onto grain mantles as ice, accounting for much of the cosmic oxygen reservoir in dense regions. This role underscores H₃O⁺'s importance in bridging gas-phase ion reactions to solid-state chemistry.

Astronomical Observations

The hydronium ion (H₃O⁺) was tentatively detected in 1986 toward the Orion KL region, where the ground rotational transition at 307 GHz was observed using the National Radio Astronomy Observatory 12 m telescope. This discovery was confirmed in 1991 by the detection of a confirming line at 365 GHz in the same source. These observations highlighted H₃O⁺ as a key tracer of ion-molecule chemistry in dense molecular clouds, with subsequent detections in sources like Sagittarius B2 confirming its role as a primary reservoir of oxygen in gas-phase reactions. Spectral observations of H₃O⁺ primarily target its rotational and inversion transitions in the submillimeter range, such as the J,K = 1,0–0,0 line at 307 GHz and the J,K = 3,3–2,2 line at 365 GHz, which arise from the ion's symmetric top structure and are observable in absorption or emission against continuum sources. The Herschel Space Observatory, particularly its Heterodyne Instrument for the Far Infrared (HIFI), enabled high-resolution mapping of these features in the 2010s, revealing H₃O⁺ absorption in diffuse and translucent clouds along lines of sight to bright infrared sources. In photon-dominated regions (PDRs) like the Orion Bar, Herschel/HIFI observations provided upper limits to H₃O⁺ column densities, indicating low abundances relative to OH⁺ due to efficient proton transfer to water in UV-irradiated environments. H₃O⁺ has also been identified in the diffuse interstellar medium (ISM) through absorption lines toward continuum sources, with detections in galactic and extragalactic sightlines demonstrating its ubiquity as a protonated water precursor. Abundance measurements from these observations typically yield column densities of N(H₃O⁺) ≈ 10¹²–10¹³ cm⁻² in diffuse ISM components, consistent with chemical models where H₃O⁺ forms via successive protonation of oxygen-bearing species and is destroyed primarily by dissociative recombination or reaction with neutral water. In the Orion KL hot core, Herschel observations set 3σ upper limits of N(ortho-H₃O⁺) < 2.4 × 10¹² cm⁻² and N(para-H₃O⁺) < 8.7 × 10¹² cm⁻², supporting models of high cosmic-ray ionization rates (ζ ≈ 7 × 10⁻¹⁴ s⁻¹) that enhance ion abundances in the region's outflow. Similar column densities (~10¹² cm⁻²) along diffuse sightlines validate steady-state photodissociation region models, where H₃O⁺ traces electron fractions and UV radiation fields. In cometary environments, H₃O⁺ was first identified in mass spectra from the Giotto spacecraft during the 1986 flyby of comet 1P/Halley, where it appeared as a dominant ion in the inner coma due to photoionization and charge exchange with water vapor. These in situ measurements revealed H₃O⁺ production rates consistent with water-dominated ion chemistry, with abundances reflecting the comet's volatile composition. Recent advances include Atacama Large Millimeter/submillimeter Array (ALMA) observations in the ALCHEMI survey (post-2020), which resolved H₃O⁺ emission in the central molecular zone of the starburst galaxy NGC 253, deriving cosmic-ray ionization rates ζ ≈ 10⁻¹⁰ s⁻¹ from ratios of H₃O⁺ to SO lines and confirming elevated ion chemistry in extragalactic star-forming regions. These high-resolution maps (~2 pc) demonstrate H₃O⁺'s utility in probing ionization in dense, active environments beyond the Milky Way.