Chlorite
The chlorite ion, or chlorite, is an inorganic anion with the chemical formula ClO₂⁻. It is a monovalent oxoanion of chlorine in the +3 oxidation state and serves as the conjugate base of chlorous acid (HClO₂).[1] Chlorite adopts a bent molecular geometry similar to sulfur dioxide, with the chlorine atom bonded to two oxygen atoms and bearing a negative charge, often represented in Lewis structures with resonance between single and double bonds. Sodium chlorite (NaClO₂) is the most common chlorite salt, appearing as a white or yellowish solid that is highly soluble in water.[1] Chlorites are synthesized industrially by reducing sodium chlorate (NaClO₃) with sodium dioxide or through electrochemical methods, and they find applications as bleaching agents, disinfectants, and precursors to chlorine dioxide (ClO₂) for water treatment. Due to their oxidizing properties, chlorites pose health risks including irritation and potential toxicity upon ingestion or inhalation, and are regulated in drinking water.[2]Chemical Fundamentals
Definition and Nomenclature
The chlorite ion is a monovalent inorganic oxyanion of chlorine, with the chemical formula \ce{ClO2-}, in which the chlorine atom exhibits an oxidation state of +3. It serves as the conjugate base of chlorous acid (\ce{HClO2}), a weak and unstable oxyacid.[1][3] The nomenclature for the chlorite ion follows standard conventions for chlorine oxyanions, deriving from the parent acid chlorous acid, where the "-ous" ending indicates the lower oxidation state relative to chloric acid. The systematic IUPAC name is chlorite(1−), reflecting its anionic charge. In common usage, chlorite refers to compounds containing this ion, such as sodium chlorite (\ce{NaClO2}), named by substituting the cation for the hydrogen in the acid formula.[1][4] The chlorite ion was first described in 1843 by N. Milan through the preparation of sodium chlorite, marking its initial recognition as a distinct chemical entity. Although known since the 19th century, chlorite salts were not produced on an industrial scale until the 1930s, with the Mathieson Company initiating U.S. manufacturing in 1937. Key early references to chlorous acid stability studies emerged around this period, focusing on its tendency to disproportionate into chlorine dioxide, chlorate, and chloride ions due to inherent instability.[5][6] To avoid confusion, the chemical chlorite ion and its compounds must be distinguished from chlorite minerals, a group of phyllosilicate minerals (e.g., clinochlore) characterized by a layered sheet structure and green coloration, commonly found in metamorphic rocks.[7]Oxidation State and Basic Reactivity
In the chlorite ion (ClO₂⁻), chlorine exhibits an oxidation state of +3. This formal oxidation number is calculated by assigning each oxygen atom an oxidation state of -2, resulting in a cumulative -4 charge from the two oxygens; given the overall -1 charge of the ion, chlorine must contribute +3 to achieve electroneutrality./Descriptive_Chemistry/Elements_Organized_by_Block/Chlorine_Compounds/Chlorite) This intermediate oxidation state positions chlorite between hypochlorite (+1) and chlorate (+5) in the series of chlorine oxyanions, influencing its electronic configuration and reactivity patterns.[8] Chlorite functions as a strong oxidizing agent, driven by the relative instability of chlorine at the +3 state, which favors either reduction to chloride (Cl⁻) or oxidation to higher oxyanions such as chlorate (ClO₃⁻). In aqueous solutions, chlorite readily participates in redox reactions, often acting as an electron acceptor due to its capacity to stabilize lower oxidation states of chlorine. A key manifestation of this reactivity is the disproportionation of chlorous acid (HClO₂), the protonated form of chlorite, in acidic media, proceeding via the balanced equation: $3 \mathrm{HClO_2} \rightarrow 2 \mathrm{HClO_3} + \mathrm{HCl} This process is second-order in chlorous acid, with an observed rate constant of approximately 0.5 M⁻¹ s⁻¹ at 25°C and pH around 2, highlighting the kinetic facility of disproportionation under acidic conditions.[9] The reaction underscores chlorite's tendency to self-oxidize and self-reduce, a behavior common to intermediate oxidation states in chlorine chemistry. The oxidizing power of chlorite can be quantified by its standard reduction potential, which is approximately 0.78 V versus the standard hydrogen electrode in neutral to basic media for the half-reaction ClO₂⁻ + H₂O + 2e⁻ → Cl⁻ + 2OH⁻. This value indicates greater oxidizing strength than chlorate (around 0.63 V under similar conditions) but somewhat less than hypochlorite (0.89 V), positioning chlorite as a potent yet selectively reactive species among chlorine oxyanions.[10] In practice, this potential enables chlorite to effectively oxidize organic reductants and certain inorganic species while being susceptible to further transformation in protic environments.[8]Structure and Physical Properties
Molecular Geometry
The chlorite ion (ClO₂⁻) possesses a bent V-shaped molecular geometry, with an O-Cl-O bond angle of approximately 111° and Cl-O bond lengths of about 156 pm. This configuration arises from the valence shell electron pair repulsion (VSEPR) model, where the central chlorine atom has four electron domains—two bonding pairs to oxygen atoms and two lone pairs—resulting in a tetrahedral electron-pair geometry and a bent molecular shape due to lone-pair repulsion compressing the bond angle below the ideal 109.5°. The chlorine atom exhibits sp³-like hybridization, with the four hybrid orbitals accommodating the two bonding pairs and two lone pairs, facilitating the observed three-dimensional arrangement.[11][12] The bonding in ClO₂⁻ is best represented by resonance structures that delocalize the π electrons, leading to equivalent Cl-O bonds with partial double-bond character. The two primary resonance forms depict a double bond between chlorine and one oxygen (with the other Cl-O as a single bond and formal charges of 0 on chlorine, -1 on the single-bonded oxygen, and 0 on the double-bonded oxygen in one form, and vice versa), averaging to bond orders of 1.5 for each Cl-O linkage. This delocalization stabilizes the ion and explains the bond lengths, which are shorter than a typical single Cl-O bond (approximately 170 pm) but longer than a double bond (approximately 140 pm).[13] In chlorite salts, the ion maintains its bent geometry within ionic lattices; for example, anhydrous sodium chlorite (NaClO₂) crystallizes in a monoclinic structure, where the chlorite ions are coordinated by sodium cations, preserving Cl-O bond lengths near 156 pm and O-Cl-O angles close to 111° despite packing effects.[14] Density functional theory (DFT) and ab initio quantum calculations provide detailed confirmation of the chlorite ion's geometry and dynamics. For instance, multireference configuration interaction (MRCI) computations on the ground state (¹A₁) yield a Cl-O bond length of 1.567 Å and an O-Cl-O angle of approximately 111°, values that align closely with experimental inferences from spectroscopic data and crystal structures, accounting for minor variations due to environmental effects. These models also predict vibrational frequencies, such as the symmetric Cl-O stretch at approximately 780 cm⁻¹ and asymmetric stretch near 940 cm⁻¹, which match observed infrared spectra and underscore the stability of the bent conformation over linear alternatives.[15]Spectroscopic and Thermodynamic Properties
The chlorite ion (ClO₂⁻) displays distinct vibrational signatures in its infrared (IR) and Raman spectra, arising from its C_{2v} symmetry and intra-ionic modes. The symmetric O-Cl-O stretching mode (ν₁) appears as a strong Raman band at approximately 780–790 cm⁻¹, while the asymmetric stretch (ν₃) is prominent in the IR spectrum at around 940 cm⁻¹, with the bending mode (ν₂) observed near 420–450 cm⁻¹. These assignments stem from matrix-isolation studies of alkali metal chlorites, where small cation-dependent shifts confirm the modes' ionic character.[16][17] In the ultraviolet-visible (UV-Vis) region, the chlorite ion exhibits an absorption maximum at λ_max ≈ 260 nm, attributed to ligand-to-metal charge transfer (LMCT) transitions involving promotion of electrons from oxygen-based orbitals to chlorine-centered ones. This broad band, with a molar absorptivity on the order of 150 M⁻¹ cm⁻¹, is commonly used for quantitative detection in aqueous solutions and overlaps minimally with related species like ClO₂ (λ_max ≈ 360 nm).[18] Thermodynamically, the chlorite ion participates in the reversible redox couple ClO₂⁻ / ClO₂ with a standard reduction potential E° ≈ 0.95 V (vs. SHE), indicating moderate oxidizing power suitable for applications in disinfection chemistry. The standard Gibbs free energy of formation for ClO₂⁻(aq) is ΔG_f° ≈ 26–29 kJ mol⁻¹, derived from equilibrium measurements and solubility data for ClO₂. Chlorous acid (HClO₂), the protonated form, has a pK_a of 1.94 at 25°C, reflecting its weak acidity and tendency to disproportionate in neutral or basic media.[8][19] Alkali metal chlorites demonstrate high aqueous solubility, facilitating their use in solution-based processes. For instance, sodium chlorite (NaClO₂) dissolves at 75.8 g/100 mL at 25°C, increasing to 122 g/100 mL at 60°C, while potassium chlorite (KClO₂) exhibits comparable or slightly lower solubility around 50–70 g/100 mL under similar conditions, with both showing endothermic dissolution trends. This high solubility underscores the ionic nature of the chlorite anion and its hydration in water.[20]Chlorite Compounds
Chlorous Acid
Chlorous acid, with the chemical formula \ce{HClO2}, serves as the parent oxoacid of the chlorite ion and exists primarily in dilute aqueous solutions due to its inherent instability. It behaves as a weak acid, characterized by a pKa value of 1.94 at 25°C, indicating partial dissociation in water to produce \ce{H+} and \ce{ClO2-}. The acid decomposes rapidly via disproportionation, yielding chlorate (\ce{ClO3-}) and chloride (\ce{Cl-}) ions as primary products, which limits its persistence in solution. This decomposition follows the overall stoichiometry where multiple molecules of \ce{HClO2} convert to higher and lower oxidation states of chlorine. In concentrated solutions, the half-life is less than 1 minute, while it persists longer only in highly dilute forms below 0.1 M, where the second-order rate constant for disproportionation is approximately 0.0275 M⁻¹ s⁻¹ at pH 1.94.[21][22] Preparation of chlorous acid typically involves the reduction of chlorine dioxide (\ce{ClO2}) with a suitable reducing agent or the acidification of barium chlorite (\ce{Ba(ClO2)2}) using dilute sulfuric acid, followed by filtration to remove barium sulfate precipitate. Early laboratory efforts to isolate the pure acid, dating back to the 1930s, underscored the significant challenges posed by its rapid decomposition, preventing stable isolation beyond transient aqueous species.[22] A key aspect of its reactivity is the disproportionation reaction: \ce{2 HClO2 -> HOCl + H+ + ClO3-} ] This process contributes to the acid's instability and is central to its behavior in solution. Chlorous acid relates to chlorite salts as their protonated form, though the salts offer greater stability for practical applications.[](https://cdnsciencepub.com/doi/10.1139/v68-335) ### Metal Chlorites and Salts Sodium chlorite (NaClO₂) is the most commonly used metal chlorite, available in technical grades with approximately 80% purity and appearing as slightly hygroscopic white crystals or flakes that do not readily cake. These crystals exhibit triclinic leaflet morphology in their trihydrate form, which transitions to the [anhydrous](/page/Anhydrous) state upon heating to 38°C or storage in a [desiccator](/page/Desiccator). The compound is highly soluble in [water](/page/Water), with a [solubility](/page/Solubility) of 64 g per 100 g at 17°C, and slightly soluble in [methanol](/page/Methanol). It decomposes at 180–200°C, releasing [heat](/page/Heat) and potentially forming [explosive](/page/Explosive) mixtures with organic materials. Potassium chlorite (KClO₂) is an orthorhombic compound ([space group](/page/Space_group) Cmcm) that is less stable than its sodium analog, decomposing within hours at [room temperature](/page/Room_temperature).[](https://journals.iucr.org/c/issues/2005/02/00/bc1064/bc1064.pdf) It shares similar reactivity as a strong oxidizer but is infrequently used due to its thermal instability.[](https://www.smolecule.com/products/s655170) Lithium chlorite (LiClO₂), for comparison, adopts a tetragonal structure ([space group](/page/Space_group) P4₂/ncm) and exhibits greater stability among [alkali metal](/page/Alkali_metal) chlorites.[](https://journals.iucr.org/c/issues/2005/02/00/bc1064/bc1064.pdf) Calcium chlorite (Ca(ClO₂)₂) forms a white granular solid that decomposes upon contact with water to yield [calcium hydroxide](/page/Calcium_hydroxide) and [chlorous acid](/page/Chlorous_acid) ($\ce{HClO2}$), rendering it insoluble in aqueous media and highly reactive. This alkaline earth chlorite acts as a strong oxidizer, igniting with substances like [potassium thiocyanate](/page/Potassium_thiocyanate) and reacting explosively with [chlorine](/page/Chlorine) or [ammonia](/page/Ammonia), with prolonged heating leading to container rupture.[](https://www.webqc.org/balanced-equation-Ca(ClO2)2+H2O=Ca(OH)2+HClO2) Heavy metal chlorites, such as silver chlorite (AgClO₂), are notably hazardous due to their explosive nature; silver chlorite is classified as a forbidden material for transport owing to its sensitivity to shock and heat.[](https://pubchem.ncbi.nlm.nih.gov/compound/9855611) These compounds generally exhibit lower stability compared to alkali variants, with decomposition often triggered at lower temperatures. Many metal chlorites exist in both anhydrous and hydrated forms, with phase transitions driven by dehydration; for instance, sodium chlorite trihydrate loses water at 38°C to form the anhydrous phase, altering its crystal lattice from triclinic to a more compact structure. Hydrated forms tend to be more stable under ambient conditions but undergo endothermic transitions upon heating, impacting solubility and reactivity. ## Synthesis and Preparation ### Industrial Production Methods Chlorite minerals are not synthesized industrially on a large scale due to their abundance in natural deposits. Instead, they are obtained through mining and processing of chlorite-bearing rocks, such as greenschists, phyllites, and schists. Extraction typically involves open-pit or underground mining in regions with metamorphic or altered igneous formations, followed by crushing, grinding, and beneficiation to concentrate the mineral. Flotation or magnetic separation may be used to isolate chlorite from associated minerals like quartz or feldspar. The processed chlorite is then used as a raw material in applications such as fillers in ceramics, paints, and drilling muds, or as crushed stone for construction. Major production occurs in countries with significant metamorphic rock resources, including the United States, Canada, and Brazil, though specific annual volumes are not widely tracked due to its status as a common accessory mineral rather than a primary commodity.[](https://geology.com/minerals/chlorite.shtml)[](https://www.mindat.org/min-1016.html) ### Laboratory Synthesis Routes Laboratory synthesis of chlorite minerals is primarily experimental, aimed at understanding [geological formation](/page/Geological_formation) processes rather than practical production. One common method involves [hydrothermal synthesis](/page/Hydrothermal_synthesis), where precursor materials like [smectite](/page/Smectite) clays or magnesium-iron silicates are subjected to elevated temperatures (200–400°C) and pressures (0.1–1 GPa) in aqueous solutions containing magnesium, iron, and aluminum ions. For example, clinochlore can be synthesized by reacting [serpentine](/page/The_Serpentine) or [talc](/page/Talc) with aluminum [hydroxide](/page/Hydroxide) under hydrothermal conditions at 300°C for several days, promoting the formation of the characteristic 1:1 layer structure with interlayer hydroxide sheets.[](https://pubs.geoscienceworld.org/msa/ammin/article-abstract/43/7-8/707/541385/Synthesis-of-the-chlorites-and-their-structural)[](https://hal.univ-lorraine.fr/hal-01876608/document) Another route uses solid-state reactions, heating mixtures of oxides (e.g., MgO, Fe₂O₃, Al₂O₃, SiO₂) with [water vapor](/page/Water_vapor) at 500–700°C to mimic metamorphic conditions, though this often yields impure products requiring [X-ray](/page/X-ray) [diffraction](/page/Diffraction) for verification. Yields vary (50–80%) depending on composition and conditions, with challenges in achieving perfect crystallinity due to the mineral's sensitivity to exact cation ratios. These syntheses confirm chlorite's stability in greenschist facies environments and aid petrological studies, but pure synthetic chlorite is not commercially produced.[](https://www.sciencedirect.com/science/article/pii/S0070457109700093) ## Applications and Uses ### Bleaching and Disinfection Sodium chlorite (NaClO₂) serves as a chlorine-free alternative to elemental [chlorine](/page/Chlorine) (Cl₂) in bleaching processes for [textile](/page/Textile)s and [pulp](/page/Pulp), offering effective whitening without significant fiber degradation or AOX formation. In [textile](/page/Textile) applications, it is applied directly in baths at concentrations of 0.1–0.5% under mildly alkaline conditions ([pH](/page/PH) 9–10), typically at temperatures of 70–90°C for 1–2 hours, enabling high whiteness levels suitable for subsequent [dyeing](/page/Dyeing).[](https://www.mdpi.com/2071-1050/16/20/9084) For [pulp](/page/Pulp) bleaching, similar dosages in alkaline media achieve delignification with reduced environmental impact compared to chlorine-based methods, preserving [pulp](/page/Pulp) brightness and strength.[](https://www.oxy.com/siteassets/documents/chemicals/products/other-essentials/sodium-chlorite-handbook.pdf) In disinfection, [sodium chlorite](/page/Sodium_chlorite), often as acidified sodium chlorite (ASC), provides antimicrobial action in [water treatment](/page/Water_treatment) through the generation of [reactive oxygen species](/page/Reactive_oxygen_species) (ROS) that damage microbial cell membranes and proteins. For [drinking water](/page/Drinking_water), regulatory limits permit chlorite residuals up to 1 mg/L to ensure pathogen inactivation while minimizing byproducts, as established by the U.S. EPA.[](https://wwwn.cdc.gov/TSP/PHS/PHS.aspx?phsid=580&toxid=108) This approach effectively targets bacteria like *[Escherichia coli](/page/Escherichia_coli)* and viruses, with ROS-mediated oxidation disrupting cellular integrity. Sodium chlorite is incorporated into oral care products such as mouthwashes and toothpastes as an anti-plaque agent, functioning via localized ROS production to inhibit bacterial adhesion and biofilm formation on dental surfaces. Approved formulations maintain concentrations below 0.1% to balance efficacy and safety, with clinical studies showing reduced plaque indices and gingival inflammation comparable to chlorhexidine.[](https://pmc.ncbi.nlm.nih.gov/articles/PMC9349900/) Recent developments from 2024–2025 highlight sodium chlorite's potential in hospital settings for biofilm disruption, particularly through ASC formulations that eradicate persistent microbial biofilms at concentrations as low as 0.05% (500 ppm). In vitro studies demonstrate its ROS-driven efficacy against hospital-relevant pathogens like *Staphylococcus aureus* and *Pseudomonas aeruginosa* in biofilms, supporting applications in surface and water system decontamination to curb healthcare-associated infections. ### Generation of Chlorine Dioxide Chlorite ions, typically in the form of [sodium chlorite](/page/Sodium_chlorite) (NaClO₂), serve as a primary precursor for generating [chlorine dioxide](/page/Chlorine_dioxide) (ClO₂), a potent [oxidizing agent](/page/Oxidizing_agent) used in industrial applications. The most established method is acid activation, where chlorite reacts with a strong acid such as [hydrochloric acid](/page/Hydrochloric_acid) (HCl) under controlled conditions to produce ClO₂ gas. This process is widely adopted in industrial settings due to its simplicity and high yield. The balanced reaction is: \[ 5 \mathrm{NaClO_2} + 4 \mathrm{HCl} \rightarrow 4 \mathrm{ClO_2} + 5 \mathrm{NaCl} + 2 \mathrm{H_2O} This equation represents the industrial standard, achieving up to 95% conversion efficiency when performed with precise stoichiometry and temperature control around 20–25°C.[23][24] Electrochemical methods have emerged as an advanced alternative, particularly for selective and sustainable production. A notable development involves the use of a Ti₄O₇ ceramic anode in an undivided electrolytic cell, which oxidizes chlorite ions (ClO₂⁻) to ClO₂ with exceptional selectivity. In a 25 mM chlorite solution, this approach yields 99% selectivity to ClO₂ at a current density of 10 mA cm⁻², minimizing unwanted oxidation products. The process operates at ambient conditions and avoids chemical additives, making it suitable for on-demand generation. This 2025 innovation highlights the potential for electrochemical systems to replace traditional chemical activations in environmentally sensitive applications.[25] On-site generation systems utilizing chlorite-based methods are essential for applications like paper pulp bleaching and water sterilization, where ClO₂ must be produced immediately prior to use due to its instability. These systems typically employ acid-chlorite or electrochemical reactors to deliver ClO₂ at concentrations of 1–10 g L⁻¹, with overall efficiencies exceeding 95% based on chlorite conversion. For instance, in pulp bleaching, on-site generators ensure consistent ClO₂ supply to achieve lignin removal without excessive fiber degradation, while in sterilization, they provide targeted dosing for microbial control in water systems. Such setups reduce transportation risks associated with ClO₂ storage and enable scalable production from small modular units to large industrial plants.[26][27] Effective byproduct control is critical in chlorite-based ClO₂ generation, as side reactions can form chlorate (ClO₃⁻), a less desirable oxidant. Chlorate formation arises from over-oxidation of chlorite or ClO₂ decomposition, but it can be minimized through precise pH management during the reaction. Optimal performance occurs at pH 2–3, where the acid activation proceeds rapidly and selectively toward ClO₂, limiting chlorate yields to below 5% under stoichiometric conditions. This pH range is achieved by metering acid addition and monitoring the reaction mixture, ensuring high-purity ClO₂ output while complying with process efficiency standards.[28]Health, Safety, and Environmental Aspects
Toxicity and Exposure Risks
Chlorite-group minerals are generally considered low in toxicity and are classified as nuisance dusts rather than hazardous substances. Acute exposure through ingestion or dermal contact is unlikely to cause severe effects, with estimated LD50 values exceeding 5,000 mg/kg for both oral and dermal routes in animal models.[29] Inhalation of dust may lead to mechanical irritation of the respiratory tract, eyes, and skin, causing temporary discomfort such as coughing, redness, or itching, similar to other silicate minerals. Prolonged exposure to high levels of respirable dust can contribute to respiratory issues, though chlorite itself is not associated with silicosis or asbestosis unless contaminated with silica or fibrous minerals.[30] No evidence links chlorite minerals to carcinogenicity, reproductive toxicity, or mutagenicity; the International Agency for Research on Cancer (IARC) does not classify them as carcinogenic. In occupational settings, exposure is regulated under general dust limits, with the Occupational Safety and Health Administration (OSHA) setting a permissible exposure limit (PEL) of 5 mg/m³ for respirable dust (as Al₂O₃ equivalent) and 15 mg/m³ for total dust, and the American Conference of Governmental Industrial Hygienists (ACGIH) recommending a threshold limit value (TLV) of 10 mg/m³ as time-weighted average (TWA).[31] These limits aim to prevent irritation and cumulative lung effects from mineral dusts.Regulatory Standards and Mitigation
Regulatory oversight for chlorite minerals focuses on general mineral dust and mining safety rather than substance-specific standards, as they pose minimal unique risks. In the United States, the Mine Safety and Health Administration (MSHA) enforces dust control in mining operations under 30 CFR Part 56/57, requiring ventilation, wetting agents, and personal protective equipment (PPE) like respirators (NIOSH-approved) to keep exposures below PELs. The Environmental Protection Agency (EPA) does not list chlorite minerals as priority pollutants under the Clean Water Act or Clean Air Act, reflecting their low solubility and environmental persistence concerns.[32] Environmentally, chlorite extraction as a byproduct in aggregate or ornamental stone mining has limited direct impacts, primarily habitat disruption, soil erosion, and water sedimentation from site operations, akin to other non-metallic mineral mining. No specific ecological toxicity is reported; chlorite minerals are inert in aquatic systems and do not bioaccumulate. Mitigation involves best practices such as revegetation of disturbed lands, sediment control basins, and compliance with the National Pollutant Discharge Elimination System (NPDES) for wastewater. In the European Union, the REACH regulation assesses minerals like chlorite under aggregate registrations, with no harmonized hazard classifications as of 2025. Industry guidelines emphasize dust suppression and spill containment to minimize releases.[33]Related Chlorine Chemistry
Comparison to Other Oxyanions
The chlorine oxyanions form a homologous series derived from chlorine in various oxidation states, ranging from chloride at -1 to perchlorate at +7. These species exhibit distinct chemical behaviors influenced by the central chlorine atom's oxidation state and the number of bound oxygen atoms. Chlorite (ClO₂⁻), with chlorine in the +3 state, occupies an intermediate position in this series, bridging the more reactive lower-oxidation-state anions and the more stable higher ones.| Formula | Name | Oxidation State of Cl |
|---|---|---|
| Cl⁻ | Chloride | -1 |
| ClO⁻ | Hypochlorite | +1 |
| ClO₂⁻ | Chlorite | +3 |
| ClO₃⁻ | Chlorate | +5 |
| ClO₄⁻ | Perchlorate | +7 |