Hydrofluoric acid is a colorless, fuming aqueous solution of hydrogen fluoride (HF), a diatomic molecule with the chemical formulaHF and a molecular weight of 20.01 g/mol.[1] It exists as a gas at room temperature but is commonly handled as a liquidsolution up to 70% concentration, exhibiting a strong irritating odor with an odor threshold of 0.03 mg/m³.[2] Despite being a weak acid with a pKa of 3.17, it forms strong hydrogen bonds due to fluorine's high electronegativity, resulting in a higher boiling point of 19.5°C compared to other hydrogen halides.[3]Chemically, hydrofluoric acid is highly reactive and corrosive, particularly to silica-containing materials like glass, which it dissolves to form silicon tetrafluoride and water.[1] It reacts exothermically with metals to produce hydrogen gas and with bases to form fluoride salts, but its low dissociation in water belies its penetrating power into organic tissues.[3]Anhydrous HF is one of the strongest known acids, capable of protonating organic compounds, while aqueous solutions are miscible with water and denser at about 1.15 g/mL for 49% concentration at 20°C.[1]Hydrofluoric acid is industrially vital, serving as a precursor for fluorochemicals such as refrigerants, Teflon, and pharmaceuticals, with about 60% of production used in fluorocarbon synthesis.[2] Key applications include glass and semiconductor etching, aluminum production, uranium refining, stainless steel pickling, and petroleum alkylation as a catalyst.[1] It is also employed in manufacturing computer screens, fluorescent bulbs, and high-octane gasoline.[4]However, hydrofluoric acid poses severe health risks due to its rapid penetration of skin and tissues, causing liquefactive necrosis, severe burns, and systemic toxicity like hypocalcemia and cardiac arrhythmias, even from small exposures.[5]Inhalation can lead to respiratory damage and pulmonary edema, while ingestion affects major organs and may be fatal; exposure limits include an 8-hour TWA of 3 ppm (OSHA PEL).[2] Prevention requires chemical-resistant protective equipment, and treatment involves immediate water rinsing followed by calcium gluconate application to bind fluoride ions.[1]
Properties
Physical properties
Hydrofluoric acid appears as a colorless, fuming liquid with a strong irritating odor, particularly at concentrations above 48%.[1]The boiling point of anhydrous hydrofluoric acid is 19.5 °C at standard pressure, whereas aqueous solutions exhibit higher boiling points that vary with concentration; for instance, the binary azeotrope consisting of 38% HF boils at 112 °C.[6][7]The density of a 48% aqueous solution is 1.15 g/cm³ at 20 °C, with values decreasing as temperature increases.[8]Hydrofluoric acid is infinitely miscible with water, ethanol, and ether, and it forms azeotropic mixtures, such as the one with water at 38% HF composition.[1]Its high volatility results in significant vapor pressure, for example, 25 mmHg at 20 °C for a 48% solution, contributing to the characteristic fuming.[9][10]Concentrated forms of hydrofluoric acid exhibit thermal stability up to approximately 100 °C, above which they decompose, releasing hydrogen fluoride gas.[11][12]
Chemical properties
Hydrofluoric acid is the aqueous solution of hydrogen fluoride, denoted as HF(aq), while the anhydrous form has the molecular formula HF. In its pure liquid state, anhydrous HF exhibits extensive hydrogen bonding due to the high electronegativity of fluorine, resulting in a structure composed of zig-zag polymeric chains of HF molecules.[13][14]The H-F bond is notably strong, with a bond dissociation energy of 565 kJ/mol, attributed to fluorine's high electronegativity. Despite this, HF readily forms the bifluorideion, [HF₂]⁻, which contains the strongest known hydrogen bond and stabilizes the species in solutions with excess fluoride.[15][16]A key reactive property of HF is its ability to etch silica-containing materials, as demonstrated by the reaction with silicon dioxide:
\ce{SiO2 + 4HF -> SiF4 + 2H2O}
This process releases silicon tetrafluoride gas and dissolves glass and quartz, making HF uniquely aggressive toward siliceous substrates compared to other acids.[17][1]HF corrodes most metals by forming corresponding metal fluorides, though it does not affect noble metals such as platinum and gold. With silicates, it forms stable complexes, including hexafluorosilicic acid (H₂SiF₆), via reactions like:
\ce{SiO2 + 6HF -> H2SiF6 + 2H2O}
Under controlled conditions, such as in the presence of catalysts, HF also reacts with organic compounds to yield fluorocarbons, facilitating the introduction of fluorine into carbon-based structures.[18][19][20]Anhydrous HF serves as a non-aqueous solvent for many ionic compounds owing to its autoprotolysis equilibrium:
\ce{3HF <=> H2F+ + HF2-}
This self-ionization produces an extremely acidic medium (H₀ ≈ -15). Unlike the other hydrogen halides (HCl, HBr, HI), which fully dissociate in water, HF remains partially undissociated due to hydrogen bonding that stabilizes the undissociated form, rendering it weaker in aqueous media but more potent as an acid in non-protic solvents.[21][22]
Acidity
Hydrofluoric acid (HF) is classified as a weak acid in dilute aqueous solutions, characterized by a pKa of 3.17 at 25 °C and an acid dissociation constant K_a = 6.8 \times 10^{-4}. This indicates partial dissociation, where only about 8% of HF molecules ionize in a 0.1 M solution. The equilibrium is represented as:\text{HF} \rightleftharpoons \text{H}^+ + \text{F}^-The limited dissociation arises from the strong H–F bond (bond dissociation energy of 569 kJ/mol), driven by fluorine's high electronegativity, combined with the small size of the F⁻ ion, which leads to strong but entropically unfavorable hydrogen bonding in its hydration shell. The standard Gibbs free energy change for dissociation, \Delta G^\circ \approx 18.1 kJ/mol at 25 °C, reflects this endergonic process, where enthalpic gains from bond breaking are offset by a negative entropy change due to the structured hydration of ions, particularly F⁻.Compared to other hydrogen halides (HX: HCl, HBr, HI), HF is the weakest acid because the H–F bond strength dominates over the solvation stabilization of the conjugate base, unlike larger halides where weaker bonds and better ion solvation enhance acidity (pKa values: HCl ≈ -7, HBr ≈ -9, HI ≈ -10). In concentrated aqueous solutions (> ~5 M), HF's effective acidity increases through ion pairing, forming bifluoride ions [HF₂]⁻, which reduces the activity of free F⁻ and shifts the equilibrium toward greater proton release, making it behave more strongly. In non-aqueous solvents or its anhydrous form, HF functions as a strong acid and superacid, fully ionizing due to the absence of water's leveling effect and its ability to solvate protons efficiently.The HF/F⁻ buffer system maintains stable pH around 3.2 and ensures availability of free fluoride ions for complexation or reactions, despite the weak initial dissociation; this buffering capacity is utilized in analytical chemistry for controlling fluoridespeciation.
Production
Industrial production
Hydrofluoric acid is produced on an industrial scale primarily through the reaction of acid-grade fluorspar (calcium fluoride, CaF₂, containing at least 97% CaF₂) with concentrated sulfuric acid. The process involves heating the mixture to 200–300 °C in rotary kilns or, less commonly, fluidized bed reactors, where the endothermic reaction generates hydrogen fluoride gas and calcium sulfate as a solid byproduct:
\ce{CaF2 + H2SO4 -> 2HF + CaSO4}
This reaction requires external heating to maintain temperature and proceeds over 30–60 minutes, with the solid calcium sulfate (gypsum) separated as waste material that must be managed due to its volume and environmental impact. The process is energy-intensive owing to the endothermic nature of the reaction, which absorbs approximately 1,400–2,500 kJ per kg of HF produced.[23][24][25]The gaseous HF is then captured, condensed, and purified via distillation to remove impurities such as sulfur dioxide (SO₂), residual sulfuric acid (H₂SO₄), water, and trace fluorides like SiF₄ or BF₃. Distillation columns separate low-boiling impurities (e.g., SO₂) in overhead streams and high-boiling ones (e.g., H₂SO₄) in bottoms, yielding high-purity HF that can be hydrated to form aqueous solutions or further processed to anhydrous form through fractional distillation under controlled conditions to achieve 99.9% purity or higher.[26][27]As of 2024, global production capacity for hydrofluoric acid exceeds 3 million metric tons per year, with actual output around 2–3 million tons annually to meet demand. Major producing countries include China (over 60% of exports), Mexico, and the United States, where capacity is approximately 220,000 metric tons per year as of 2022.[28][29][30][31]Alternative production routes, though less prevalent than the fluorspar method, include derivation from fluoroapatite in phosphate rock processing, where fluorosilicic acid (H₂SiF₆) byproduct is decomposed with sulfuric acid to yield HF, and recycling fluoride wastes from aluminum smelting operations. These methods contribute a smaller share of supply but support sustainability by utilizing industrial byproducts. Industrial hydrofluoric acid is typically marketed as 48–70% aqueous solutions for general use, while anhydrous HF is produced via additional distillation for specialized applications requiring high purity.[32][26]
Laboratory preparation
In laboratory settings, hydrofluoric acid is prepared on a small scale, typically ranging from grams to liters, to support research or educational applications, with strict emphasis on ventilation systems and neutralization setups to manage hazardous vapors.[33] A common method involves the thermal decomposition of potassium bifluoride (KHF₂), where the solid is heated to release hydrogen fluoride gas, which is then absorbed into water to form the aqueous acid.\text{KHF}_2 \xrightarrow{\text{heat}} \text{KF} + \text{HF (g)}This decomposition occurs quantitatively at temperatures above 200°C, producing pure HF gas that can be collected and dissolved in distilled water under controlled conditions.[34] The process requires apparatus made from corrosion-resistant materials such as platinum or polytetrafluoroethylene (PTFE, commonly known as Teflon) to prevent reaction with the highly corrosive HF, and distillation is often conducted under reduced pressure to lower boiling points and minimize equipment stress.[33]An alternative approach utilizes the reaction of sodium fluoride (NaF) with concentrated sulfuric acid in glass-free apparatus to generate HF gas, followed by condensation or absorption.$2\text{NaF} + \text{H}_2\text{SO}_4 \rightarrow \text{Na}_2\text{SO}_4 + 2\text{HF}This method avoids the formation of insoluble byproducts like calcium sulfate encountered in other preparations and is suitable for producing small quantities of HF, though it demands similar safety measures including plastic or metal equipment to avert glasscorrosion.[33]For obtaining anhydrous HF, aqueous solutions are dried by mixing with concentrated sulfuric acid, which forms a water-sulfuric acid azeotrope that is distilled off, leaving pure HF that can then be redistilled under reduced pressure.[35] Historically, early laboratory preparations involved distilling HF directly from fluorspar (CaF₂) with sulfuric acid, a method pioneered by Frémy in the mid-19th century through heating potassium bifluoride, but it has been largely supplanted due to significant hazards from uncontrolled gas evolution and equipment failure.
Uses
Synthesis of organofluorine compounds
Hydrofluoric acid (HF) plays a pivotal role in the synthesis of organofluorine compounds by facilitating the formation of carbon-fluorine bonds through direct fluorination, catalysis, or as a reagent in key transformations. These processes are essential for producing materials used in pharmaceuticals, refrigerants, and polymers, where the strong C-F bond imparts stability and unique properties. Anhydrous HF, in particular, serves as both a fluorinating agent and a solvent due to its superacidic nature, enabling reactions that are challenging in conventional media.[36]In the production of chlorofluorocarbons (CFCs), HF is employed in the vapor-phase catalytic displacement of chlorine atoms in chlorocarbons, such as the reaction of HF with carbon tetrachloride (CCl₄) to yield trichlorofluoromethane (CFCl₃) and HCl:\mathrm{CCl_4 + HF \rightarrow CFCl_3 + HCl}This process, a variant of the Swarts reaction, was historically significant for manufacturing refrigerants and propellants before regulatory restrictions.[37] The reaction typically requires catalysts like antimony chlorofluorides to enhance selectivity and efficiency.[2]A key application involves the synthesis of polytetrafluoroethylene (PTFE), known commercially as Teflon, whose monomer tetrafluoroethylene (TFE) is derived from chloroform and HF. Chloroform (CHCl₃) undergoes stepwise fluorination with HF to form chlorodifluoromethane (CHClF₂), which is then pyrolyzed at high temperatures (around 700–900°C) to produce TFE:\mathrm{CHCl_3 + 2HF \rightarrow CHClF_2 + 2HCl}, \quad \mathrm{2CHClF_2 \rightarrow CF_2=CF_2 + 2HCl}This multi-step process highlights HF's role in introducing fluorine atoms to aliphatic precursors, yielding a polymer renowned for its chemical inertness and non-stick properties.[38]In pharmaceutical synthesis, HF is indirectly involved in the Balz-Schiemann reaction, a classical method for preparing aryl fluorides from aromatic amines. The process begins with diazotization of the amine to form an aryldiazonium salt, followed by treatment with tetrafluoroboric acid (HBF₄, prepared from HF and boric acid) to isolate the tetrafluoroborate salt (ArN₂⁺ BF₄⁻). Thermal decomposition then affords the aryl fluoride:\mathrm{ArN_2^+ BF_4^- \rightarrow ArF + N_2 + BF_3}This reaction is widely used to introduce fluorine into drug scaffolds, enhancing metabolic stability, as seen in compounds like fluoroquinolone antibiotics. Yields typically range from 40–70%, with modern variants improving efficiency through solvent optimization.[39]For refrigerants, HF is crucial in producing hydrofluorocarbons (HFCs), such as 1,1,1,2-tetrafluoroethane (HFC-134a), a common CFC replacement. The synthesis starts with the hydrofluorination of trichloroethylene (ClCH=CCl₂) using HF, often in a catalyzed vapor-phase process, proceeding through intermediates like 1-chloro-1,1,2-trifluoroethane to the final product:\mathrm{ClCH=CCl_2 + 4HF \rightarrow CF_3CH_2F + 3HCl}This multi-stage reaction, developed in the late 1980s, achieves high selectivity (>95%) under controlled conditions and supports large-scale production for automotive air conditioning.[40]Anhydrous HF also acts as a solvent for electrophilic fluorination reactions, particularly in superacid systems combined with Lewis acids like SbF₅, enabling the fluorination of aromatic and aliphatic substrates via electrophilic aromatic substitution or addition to unsaturated bonds. For instance, it facilitates the direct fluorination of benzene to fluorobenzene under electrochemical conditions, where HF provides both the medium and fluoride source. These methods are valued for their ability to introduce fluorine at specific positions, though they require specialized equipment due to HF's corrosivity.The transition from CFCs to HFCs was driven by the 1987 Montreal Protocol, which phased out ozone-depleting substances, prompting the chemical industry to develop HF-based processes for HFCs that avoid chlorine while maintaining thermodynamic performance. This shift increased global HF demand for organofluorine production, with HFC-134a becoming a flagship example of environmentally adapted synthesis.[41]
Production of inorganic fluorides
Hydrofluoric acid (HF) serves as a key reagent in the industrial synthesis of various inorganic fluorides, enabling the fluorination of metal and non-metal compounds for applications in aluminum production, water treatment, nuclear fuels, and electronics. These processes typically involve neutralization or direct reaction of HF with oxides, hydroxides, or carbonates, often under controlled conditions to manage the acid's corrosivity and ensure high purity. The resulting fluorides are essential in ceramics for fluxing agents, in electronics for plasma etching precursors, and in catalysts for industrial reactions.[42]Aluminum fluoride (AlF₃) is produced on a large scale by neutralizing hydrofluoric acid with aluminum hydroxide, following the reaction:\ce{3HF + Al(OH)3 -> AlF3 + 3H2O}This wet process yields a product used primarily as an electrolyte additive in aluminum smelting via the Hall-Héroult method, where it lowers the melting point of cryolite and improves conductivity. Production is predominantly from HF-based routes, supporting the aluminum industry's demand for high-purity grades.[43][44][45]Sodium fluoride (NaF) is manufactured through the neutralization of hydrofluoric acid with sodium hydroxide or sodium carbonate:\ce{HF + NaOH -> NaF + H2O}The resulting salt is centrifuged, washed, and dried to achieve pharmaceutical or technical grades. It finds widespread use in water fluoridation to prevent dental caries and in toothpaste formulations as an anti-cavity agent, with production emphasizing low-impurity variants for public health applications.[42]Cryolite (Na₃AlF₆) is synthesized by first forming hexafluoroaluminate acid from HF and aluminum hydroxide, followed by reaction with sodium carbonate:\ce{Al(OH)3 + 3HF -> H3AlF6 + 3H2O}, \quad \ce{H3AlF6 + 3Na2CO3 -> 2Na3AlF6 + 3CO2 + 3H2O}This synthetic cryolite substitutes for scarce natural deposits and is vital in the Hall-Héroult process as the primary solvent for alumina dissolution during aluminum electrolysis. Industrial production focuses on granular forms to optimize electrolytic performance and reduce energy consumption in smelters.[46]Uranium tetrafluoride (UF₄), a precursor for nuclear fuel, is prepared by hydrofluorination of uranium dioxide with anhydrous HF gas in a fluidized bed reactor:\ce{UO2 + 4HF -> UF4 + 2H2O}The process operates under nitrogen dilution to facilitate gas-solid contact and remove water vapor, yielding green salt (UF₄) that is subsequently converted to UF₆ for enrichment. This method ensures high conversion efficiency in nuclear fuel cycles, with stringent controls for radiological safety.[47]Silicon tetrafluoride (SiF₄) is generated as a volatile product from the reaction of hydrofluoric acid with silicon dioxide, often during quartzetching or recycling processes:\ce{SiO2 + 4HF -> SiF4 + 2H2O}Captured as a byproduct in semiconductormanufacturing, SiF₄ is recycled by hydrolysis to recover HF and silica, minimizing waste in electronics production. Its gaseous nature allows easy purification for use in chemical vapor deposition and optical fiber synthesis.Nitrogen trifluoride (NF₃) for electronics is produced via electrolysis of a molten mixture of ammonium fluoride, potassium fluoride, and hydrofluoric acid (NH₄F–KF–HF), achieving purities up to 99.99% essential for plasma cleaning in semiconductor fabrication. This electrolytic route avoids direct fluorination hazards and supports the demand for ultra-high-purity gases in microelectronics.[48]
Etching and cleaning
Hydrofluoric acid plays a critical role in semiconductor manufacturing, where buffered hydrofluoric acid (BHF)—a mixture of hydrofluoric acid and ammonium fluoride (NH₄F)—is used for the isotropic etching of silicon dioxide (SiO₂) layers on microchips. This process selectively removes oxide films to define circuit patterns, with typical etch rates for thermal SiO₂ reaching approximately 100 nm/min in a 5:1 BHF solution (5 parts 40% NH₄F to 1 part 49% HF). The buffering stabilizes the pH and fluoride ion concentration, ensuring uniform etching and preventing undercutting of photoresist masks, which is essential for precise fabrication in integrated circuits.[49][50]In glass processing, hydrofluoric acid etches surfaces to produce frosted or matte finishes for decorative items, lighting, and architectural elements, typically using concentrations of 5-10% HF applied via immersion or gel formulations to control the depth and uniformity of the etch. This reaction dissolves the silica network in glass (SiO₂ + 4HF → SiF₄ + 2H₂O), creating a light-diffusing texture without compromising structural integrity. For industrial cleaning, hydrofluoric acid removes rust and oxide scales from metals, as in the reaction Fe₂O₃ + 6HF → 2FeF₃ + 3H₂O, and dissolves silicate-based deposits in boiler systems, often in inhibited formulations to protect underlying alloys.[1][51][52]Specific applications include texturing silicon wafers in solar panel production to create anti-reflective surfaces, where acidic mixtures containing hydrofluoric acid roughen multicrystalline silicon for improved light trapping and efficiency. In phosphate fertilizer processing, hydrofluoric acid facilitates silica removal from phosphate rock by converting impurities to soluble fluorosilicates, enhancing product purity during wet acidulation. Safety protocols emphasize diluting hydrofluoric acid to 1-5% for etching and cleaning operations to minimize exposure risks, followed by neutralization with hydrated lime (Ca(OH)₂) to form insoluble calcium fluoride (CaF₂) for safe disposal: 2HF + Ca(OH)₂ → CaF₂ + 2H₂O. Demand from the electronics industry, particularly semiconductors and display manufacturing, represents a significant portion of global hydrofluoric acid production.[53][54][55][56][57]
Petroleum refining
In petroleum refining, hydrofluoric acid (HF) serves as a key catalyst in the alkylation process, which combines isobutane with light olefins such as propylene or butenes to produce high-octane alkylate, primarily branched C8 hydrocarbons like 2,2,4-trimethylpentane (C8H18).[58] This reaction occurs in the liquid phase at moderate temperatures of 20–40 °C and elevated pressures, where HF promotes the formation of carbocations from the olefins, facilitating their combination with isobutane to yield branched-chain products that enhance gasoline octane ratings to 95 or higher.[58] The process is highly selective for desirable isomers, contributing to cleaner-burning, high-performance fuels.[59]The two predominant commercial HF alkylation technologies are the Phillips process and the UOP process, both featuring a reactor where hydrocarbons and anhydrous HF are intimately mixed, followed by separation in a settler and HF regeneration through fractionation and purification to remove water and contaminants.[60] In these systems, HF is continuously recycled, achieving over 99% recovery rates with minimal byproducts such as propane and n-butane, which are separated via distillation.[61] Worldwide, approximately 50 HF alkylation units operate, primarily in the United States where about 48 facilities produce roughly 1 million barrels per day of alkylate, representing a significant portion of global high-octane gasoline blending stock.[62][63]Compared to sulfuric acid (H2SO4) alkylation, HF offers advantages including lower overall acid consumption, reduced sensitivity to temperature variations, easier regeneration without the need for frequent acid replacement, and less equipment corrosion in optimized systems.[58] However, HF's extreme toxicity and potential for vapor cloud formation have prompted a phasedown, driven by heightened regulations in the 2010s, such as EPA risk assessments and state-level mandates for enhanced mitigation like water neutralization systems.[64] These concerns were amplified following the 1987 Texas City incident, where a major HF release from a Marathon refinery exposed thousands and led to stricter environmental controls and inventory limits at HF units.[65]Emerging alternatives to liquid HF include solid acid catalysts, which eliminate handling risks associated with corrosive acids and are under continued development, though broader commercial adoption remains limited as of 2025.[66] In February 2025, environmental groups petitioned the U.S. EPA to prohibit HF use in refineries under the Toxic Substances Control Act due to ongoing safety risks, adding to regulatory scrutiny.[67]
Health and safety
Toxicity and hazards
Hydrofluoric acid (HF) is exceptionally hazardous due to its dual corrosive and systemic toxic effects, primarily from the penetration and dissociation of fluoride ions into biological tissues. Upon dermal contact, HF rapidly diffuses through the skin and underlying tissues, often without immediate pain, allowing deep penetration before dissociation occurs. The released fluoride ions (F⁻) strongly bind to divalent cations such as calcium (Ca²⁺) and magnesium (Mg²⁺), precipitating them as insoluble salts and depleting essential electrolytes in cells and extracellular fluids. This leads to cellular disruption, liquefactive necrosis, and progressive tissue destruction that can extend to bone if untreated.[14]Systemic effects from significant absorption, particularly dermal exposures exceeding 2-3% of total body surface area with concentrated solutions, include severe hypocalcemia and hypomagnesemia, which can trigger life-threatening cardiac arrhythmias, renal failure, and multi-organ dysfunction. HF has an acute oral LD50 of approximately 25–50 mg/kg in rats (GHS Category 2), underscoring its high toxicity via ingestion and potential for rapid systemic uptake.[68]Inhalation of HF vapor causes immediate irritation to the respiratory tract, potentially progressing to laryngeal spasm, bronchitis, and delayed-onset pulmonary edema due to corrosive damage and fluid accumulation in the lungs. The fuming nature of HF in humid air increases the risk of inadvertent inhalation during handling. Ocular exposure results in severe corneal ulceration and opacification, often leading to permanent vision impairment or blindness from rapid penetration and ion binding in eye tissues.[69][14]Chronic low-level exposure through repeated dermal contact or inhalation can accumulate fluoride, resulting in fluorosis—a condition marked by skeletal fluorosis with bone densification, brittleness, joint stiffness, and potential dental mottling. The carcinogenicity of HF remains unclassifiable; inorganic fluorides employed in drinking water, encompassing HF-derived compounds, are designated Group 3 by the International Agency for Research on Cancer (IARC), indicating inadequate evidence for human carcinogenicity.[70][71]In contrast to other mineral acids like sulfuric or hydrochloric acid, which typically produce coagulative necrosis that confines damage to the surface, HF uniquely causes delayed pain onset—sometimes up to 24 hours post-exposure—and facilitates deeper tissue liquefaction through ongoing fluoride diffusion and electrolyte sequestration.[14][72]Regulatory exposure limits reflect HF's potency: the Occupational Safety and Health Administration (OSHA) establishes a permissible exposure limit (PEL) of 3 ppm as a ceiling value, while the National Institute for Occupational Safety and Health (NIOSH) defines the immediately dangerous to life or health (IDLH) concentration at 30 ppm.[73]
Treatment and first aid
Immediate first aid for hydrofluoric acid (HF) exposure focuses on rapid decontamination to minimize fluorideion penetration and subsequent administration of calcium-based antidotes to counteract systemic effects by binding free fluoride ions.[69] Workplaces handling HF must maintain calcium gluconate gel as a mandatory first aid component, per occupational safety standards requiring its availability for prompt topical application.[74]For skin exposure, the initial step is to rinse the affected area with copious lukewarm water for 15-30 minutes to dilute and remove the acid, followed immediately by application of 2.5-10% calcium gluconate gel or jelly, massaged into the skin every 10-15 minutes until pain subsides, as this binds fluoride ions locally.[69][75] In severe cases involving large surface areas (>20% body surface) or signs of systemic toxicity such as hypocalcemia, intravenous administration of 10-20 mL of 10% calcium gluconate solution is indicated, with continuous electrocardiogram (ECG) monitoring for arrhythmias like QT prolongation.[69][76]Eye exposure requires immediate irrigation with normal saline or a calcium lactate solution for at least 20-30 minutes, using 1-2 liters of fluid while holding eyelids open, followed by topical 1% calcium gluconate drops if corneal damage is suspected; urgent ophthalmology consultation is essential for ongoing evaluation and potential subconjunctival calcium injection.[69][77]For inhalation exposure, provide supplemental oxygen therapy to support respiration and prevent pulmonary edema, while nebulizing 2.5% calcium gluconate may be attempted; benzodiazepines such as lorazepam are used to control any resultant seizures.[69] In cases of ingestion, administer 4-12 ounces of water or milk if the patient is conscious to dilute the acid, but avoid gastric lavage due to the high risk of esophageal or gastric perforation from the corrosive nature of HF.[69][76]In hospital settings, protocols include serial monitoring of serum calcium and magnesium levels every 6 hours, along with ECG and vital signs assessment for at least 24 hours; hemodialysis is employed for severe renal failure or refractory hypocalcemia, as outlined in updated CDC/ATSDR guidelines from the 2020s.[69][78] All exposed individuals should receive evaluation even if asymptomatic, given the potential for delayed toxicity.[79]
Handling and regulations
Hydrofluoric acid (HF) requires specialized storage to prevent corrosion and accidental release due to its reactivity with many materials. It should be stored in containers lined with polyethylene or polytetrafluoroethylene (PTFE, commonly known as Teflon) to ensure compatibility and avoid degradation. Storage areas must be kept away from sources of water or moisture to minimize the risk of leaks or reactions that could generate hazardous fumes, and facilities should maintain ventilation sufficient to keep airborne concentrations below 3 parts per million (ppm), aligning with occupational exposure limits. Secondary containment, such as spill trays or bunds, is essential in all storage locations to capture potential leaks.[70][79]Personal protective equipment (PPE) is critical for safe handling of HF to protect against its severe corrosive and toxic effects. Gloves made from neoprene or Viton provide effective resistance to permeation for extended periods, while face shields and chemical-resistant aprons or suits are required to shield the face, eyes, and body. Cotton clothing must be avoided, as it readily absorbs HF and can prolong skin contact, exacerbating injury. All PPE should be inspected regularly and stored separately from work areas to prevent contamination.[79][80][81]Transportation of HF is governed by international and national regulations to mitigate risks during shipping. Aqueous solutions of HF (up to 60% concentration) are classified under United Nations number UN 1790 as a Class 8 corrosive substance, requiring appropriate placards, labeling, and packaging in corrosion-resistant containers such as steel drums with protective linings. Anhydrous HF, handled as hydrogen fluoride gas, falls under UN 1056 and is subject to additional restrictions due to its toxicity (Class 2.3) and corrosivity. Shippers must comply with Department of Transportation (DOT) rules, including segregation from incompatible materials and emergency response information.[82][83]In the event of a spill, immediate containment and neutralization are necessary to limit exposure and environmental impact. Small spills should be neutralized using lime (calcium oxide) or soda ash (sodium carbonate) to form less hazardous fluoride salts, followed by absorption with inert materials like vermiculite or sand before disposal. Larger spills require evacuation and professional response teams. HF waste, including spill residues, is designated under EPA hazardous waste code D002 for its corrosivity (pH ≤2), and must be managed through permitted treatment, storage, and disposal facilities.[84][85][86]HF is subject to stringent regulatory oversight in multiple jurisdictions to ensure safe management and minimize environmental release. In the United States, it is listed on the Toxic Substances Control Act (TSCA) inventory, with ongoing evaluations of its risks in industrial applications. Under the EU's REACH regulation, HF is classified as a substance of very high concern due to its severe health hazards, imposing authorization and restriction requirements for uses like alkylation in refineries. Emissions are controlled under the US Clean Air Act as a hazardous air pollutant, mandating maximum achievable control technology for sources exceeding thresholds. As of 2025, environmental groups have petitioned the EPA under TSCA Section 21 to prohibit HF use in domestic oil refining, citing unreasonable risks, leading to litigation and potential phaseout rules.Workers handling HF, particularly in high-risk processes like refineryalkylation units, must undergo comprehensive training as required by OSHA's Process Safety Management standard (29 CFR 1910.119). This includes initial and refresher training every three years on HF-specific hazards, safe work practices, emergency procedures, and process equipment integrity. Facilities must develop written safety programs, conduct hazard analyses, and ensure mechanical integrity to prevent releases.[87][88][89]
History
Discovery and early development
Hydrofluoric acid was first isolated in 1771 by Swedish chemist Carl Wilhelm Scheele through the distillation of fluorspar (calcium fluoride, CaF₂) with concentrated sulfuric acid, producing a corrosive liquid that etched glass vessels.[90] Scheele noted its unique ability to dissolve silicates, distinguishing it from other mineral acids, though he did not fully characterize its composition.[91]In the early 19th century, British chemist Humphry Davy, building on André-Marie Ampère's hypothesis, proposed that fluoric acid consisted of hydrogen combined with a novel element he termed "fluorine," based on its etymological root from fluorspar.[90] This naming reflected early observations of its glass-corroding properties, which complicated handling and led to experiments in lead or platinum vessels as alternatives to glass.[92]Advances in the mid-19th century included French chemist Edmond Frémy's 1856 preparation of anhydrous hydrofluoric acid by distilling potassium bifluoride, yielding a purer form free of water that intensified its reactivity.[93] This work paved the way for Henri Moissan's 1886 electrolytic isolation of elemental fluorine from a molten mixture of potassium fluoride and hydrofluoric acid, confirming the acid's structure as hydrogen fluoride (HF) and earning Moissan the 1906 Nobel Prize in Chemistry.[91]Early applications were limited by the acid's extreme hazards, including severe burns and toxicity, but it found niche use in etching gemstones such as aquamarine (a variety of beryl) to create decorative frosted effects, often on a small scale in artisanal settings.[94] Pre-20th century production remained artisanal, primarily via Scheele's fluorspar-sulfuric acid method or occasional alternatives like cryolite decomposition, yielding modest quantities for laboratory and craft purposes.[95]Key scientific milestones included Leopold Gmelin's 1819 Handbuch der theoretischen Chemie, which systematically documented hydrofluoric acid's properties and reactions, establishing it as a foundational reference in inorganic chemistry.[96] These developments underscored the acid's dual role as a powerful reagent and perilous substance, shaping cautious early research protocols.
Industrial adoption and notable incidents
Hydrofluoric acid (HF) saw significant industrial adoption in the early 20th century, particularly in the production of chlorofluorocarbons (CFCs) such as Freon, which was developed by Thomas Midgley Jr. in 1928 as a non-toxic refrigerant alternative.[97] HF served as a key fluorinating agent in the synthesis of these compounds, enabling the rapid commercialization of refrigeration and air conditioning technologies.[98] During World War II, HF played a critical role in the Manhattan Project for uranium enrichment, where it was used to convert uranium oxide to uranium tetrafluoride and subsequently to uranium hexafluoride gas for gaseous diffusion processes.[99]Post-1950s, HF use expanded dramatically in petroleum refining through alkylation processes to produce high-octane gasoline. Phillips Petroleum Company patented and commercialized the first HF alkylation unit in 1942, marking a shift from sulfuric acid-based methods due to HF's efficiency and recyclability, which spurred widespread adoption in U.S. refineries during the post-war economic boom.[100] By the 1970s, surging demand in the electronics industry further drove HF consumption, as it became essential for etching silicon wafers and cleaning semiconductor surfaces in the burgeoning integrated circuit manufacturing sector.[101]Several notable incidents underscored HF's hazards and influenced industrial practices. On October 30, 1987, a construction accident at the Marathon Petroleum refinery in Texas City, Texas, ruptured an HF storage tank, releasing a toxic vapor cloud that affected nearby residents and workers, resulting in over 700 medical treatments for respiratory irritation and burns, with thousands evacuated.[64] In June 2019, explosions at the Philadelphia Energy Solutions refinery released approximately 5,000 pounds of HF, along with a 38,000-pound vessel fragment that landed off-site, prompting evacuations and highlighting vulnerabilities in aging infrastructure, though no fatalities occurred.[102]More recently, between October 2021 and June 2024, multiple hydrogen fluoride releases occurred at the Honeywell Geismar facility in Louisiana due to heat exchanger ruptures, affecting workers and prompting a CSB investigation. The final report, released on May 27, 2025, highlighted systemic safety failures.[103]These events contributed to stricter safety measures, including the development of OSHA's Process Safety Management standard in 1992, which was partly motivated by HF-related risks and the Clean Air Act Amendments of 1990.[104] The amendments also prompted EPA assessments, such as the 1993 report on HF accident risks, fueling ongoing discussions about phasing out HF in refining to mitigate potential catastrophic releases.[64] Globally, HF production has grown substantially since the mid-20th century, from modest scales in the 1950s to over 3 million metric tons annually by the 2020s, largely driven by demand for fluoropolymers and electronics applications.[105] Regulatory frameworks evolved accordingly, with HF registered under the EU's REACH regulation effective from 2007, requiring detailed safety data and risk management.[106] In the U.S., the EPA has conducted ongoing risk evaluations, including a 2022 recommendation from the Chemical Safety and Hazard Investigation Board to prioritize HF under TSCA for risk evaluation.[107]
Cultural references
Hydrofluoric acid (HF) has gained notoriety in popular media for its extreme corrosiveness, most prominently in the television series Breaking Bad (2008–2013), where characters use it to dissolve human remains in a bathtub, leading to the acid eating through the porcelain and floor. This depiction, while dramatized, heightened public awareness of HF's ability to penetrate tissues and cause severe, delayed burns, though scientific analysis later clarified that concentrated HF cannot fully dissolve organic matter as rapidly or completely as shown. The scene's impact extended to educational discussions on chemical hazards, influencing viewer perceptions of acids in crime narratives.[108][109]In literature, HF appears in forensic and thriller contexts, such as Jeff Lindsay's Dexter book series (2004–2015), where it is referenced in methods for evidence disposal, perpetuating myths about its efficacy in "melting" bodies without trace. These portrayals often exaggerate HF's speed and totality of dissolution for narrative tension, drawing from real chemical properties but amplifying them into plot devices that explore moral and ethical dilemmas in crime fiction. Such references have contributed to broader cultural fascination with acids in forensic science, blending fact with fiction to underscore the substance's dangers.[110][111]Documentaries and films from the 2010s have further explored HF's risks through educational lenses, including the MythBusters episode in 2013 testing the Breaking Bad acid bath scenario, which demonstrated that HF fails to dissolve a pig carcass (as a human proxy) or breach a bathtub within realistic timeframes, using enhanced acids like sulfuric for comparison. Educational YouTube videos since 2010, such as analyses by chemistry educators, have dissected these scenes to correct misconceptions while highlighting HF's unique toxicity, often tying briefly to its real hazards like bone dissolution and systemic poisoning that inspire such fiction. These productions have served as informal public service announcements, promoting safer chemical understanding amid entertainment.[109][112]Cultural misconceptions portray HF as an "instant death acid" capable of immediate, total destruction, fueled by urban legends and media tropes that ignore its weak acidity and slower action compared to stronger corrosives like hydrochloric acid. These myths, often spread via online forums and anecdotal tales, overestimate HF's body-dissolving prowess, leading to unnecessary fear; CDC fact sheets counter this by emphasizing that while HF causes insidious, penetrating burns without initial pain, it requires specific medical intervention like calcium gluconate rather than evoking instant lethality. Such clarifications help demystify HF, distinguishing its fluoride ion toxicity from sensationalized invincibility.[113][5]In art and scientific illustration, HF has historical ties to 19th-century glass etching techniques, where it was used to create intricate cameo designs by selectively dissolving layers of crystal, as seen in works like the 1867 Greek Revival vase at the Petit Palais, showcasing its role in producing frosted, decorative surfaces. This application transformed HF from an industrial tool into an artistic medium, enabling precise patterns in luxury glassware during the Victorian era. Modern biohazard art incorporates HF motifs symbolically, appearing in safety posters and digital illustrations that depict corrosive warnings alongside hazard symbols, reflecting contemporary concerns over chemical spills in creative expressions of environmental risk. Art safety guidelines highlight HF's presence in studio materials, urging precautions to prevent its dangers from infiltrating artistic practice.[114][115]As of 2025, podcasts on chemical safety have addressed HF in the wake of recent industrial incidents, such as the January episode of Chemical Processing's process safety series discussing recurring accidents involving corrosive releases and the need for enhanced emergency protocols. These discussions, including analyses of thermal breakdowns producing HF byproducts in refrigerants on the Industrial Strength Podcast, emphasize organizational learning from spills to mitigate public health threats, bridging cultural awareness with preventive strategies.[116][117]