Isobutane
Isobutane, systematically named 2-methylpropane, is an organic compound with the molecular formula C₄H₁₀ and the structural formula (CH₃)₃CH.[1][2] It represents one of the two structural isomers of butane, featuring a branched chain with a central tertiary carbon atom bonded to three methyl groups and one hydrogen.[1] This flammable, colorless gas occurs naturally in petroleum and natural gas deposits and is industrially produced through processes such as isomerization of n-butane or as a byproduct of petroleum refining.[3] Key applications include its use as a refrigerant under the designation R-600a in domestic appliances, owing to its thermodynamic properties and low global warming potential of approximately 3 relative to CO₂; as a non-ozone-depleting propellant in aerosol products like deodorants and foams; and as a primary alkylating agent in refinery processes to generate branched alkanes that enhance gasoline octane ratings.[4][5][6] Despite its utility, isobutane's high flammability—exhibiting an explosive range of 1.8% to 8.4% in air—necessitates rigorous safety protocols in storage, transport, and application to mitigate risks of ignition and asphyxiation in confined spaces.[7][8]Properties
Chemical Structure and Reactivity
Isobutane has the molecular formula C₄H₁₀ and is systematically named 2-methylpropane. Its molecular structure features a branched alkane chain, with a central tertiary carbon atom bonded to one hydrogen atom and three methyl (–CH₃) groups, expressed as (CH₃)₃CH. This configuration differs from n-butane (CH₃CH₂CH₂CH₃), which possesses a linear chain of four carbon atoms linked by single bonds. The branching in isobutane creates a more compact molecule, influencing its steric properties and distinguishing it from straight-chain isomers in terms of symmetry and conformational flexibility. As a saturated alkane, isobutane demonstrates high thermal and chemical stability, resisting addition reactions and electrophilic or nucleophilic attacks common to unsaturated hydrocarbons; its reactivity is dominated by homolytic cleavage of C–H bonds in free-radical processes. Primary reactions include free-radical halogenation, where bromine or chlorine substitutes hydrogen atoms, preferentially at the tertiary position due to the lower bond dissociation energy of the tertiary C–H bond (approximately 390 kJ/mol or 93 kcal/mol) compared to primary C–H bonds (410–435 kJ/mol).[9] In thermal cracking, isobutane undergoes C–C bond fission at high temperatures (above 500°C), yielding smaller alkanes, alkenes, and hydrogen, with the branched structure favoring beta-scission pathways over straight-chain variants. The steric crowding from methyl groups enhances resistance to autoignition by increasing the activation energy for radical chain propagation during combustion, resulting in a research octane number (RON) of 101–102, higher than n-butane's 93–94. This effect arises from reduced molecular packing and hindered approach of radicals, delaying the formation of branched-chain carriers that accelerate knocking in engines. Catalytic reforming, such as isomerization or dehydrogenation over platinum-based catalysts, exploits this stability to convert isobutane to higher-value products like isobutene, though requiring temperatures of 400–600°C to overcome kinetic barriers.Physical and Thermodynamic Properties
Isobutane is a colorless, odorless gas at standard temperature and pressure, with a molecular weight of 58.12 g/mol.[10] Its melting point is -159.6 °C, and its normal boiling point is -11.7 °C at 1 atm.[1] The critical temperature is 134.7 °C, above which it cannot be liquefied regardless of pressure, and the critical pressure is 36.4 bar.[1] [11] The density of isobutane gas at standard temperature and pressure (0 °C, 1 atm) is approximately 2.59 g/L, while its liquid density at the boiling point is 594 kg/m³.[8] Its vapor density relative to air is 2.01, contributing to its tendency to accumulate in low-lying areas.[3] Isobutane exhibits very low solubility in water, on the order of 0.006 g/100 mL at 20 °C, but is miscible with many organic solvents such as ethanol and ether.[8] Vapor pressure increases nonlinearly with temperature, following the Antoine equation parameters derived from experimental data, reaching about 3.1 bar at 21 °C.[12] Thermodynamically, the standard enthalpy of combustion for gaseous isobutane is -2870 kJ/mol, reflecting its high energy release upon oxidation to CO₂ and H₂O.[13] The enthalpy of vaporization at the boiling point is 21.3 kJ/mol.[14] The ideal gas heat capacity at constant pressure (C_p) is approximately 96.7 J/mol·K at 25 °C, increasing with temperature due to vibrational contributions.[13]| Property | Value | Conditions |
|---|---|---|
| Heat capacity (C_p, gas) | 96.7 J/mol·K | 25 °C, 1 atm |
| Enthalpy of formation (Δ_f H°) | -135 kJ/mol | 298 K, gas |
| Standard entropy (S°) | 250 J/mol·K | 298 K, gas |
Production
Industrial Synthesis
Isobutane is primarily synthesized industrially via the catalytic isomerization of n-butane in dedicated refinery units known as butamer processes. This skeletal rearrangement converts straight-chain n-butane (C4H10 linear) into branched isobutane (2-methylpropane) using bifunctional catalysts that combine metal sites for dehydrogenation/hydrogenation and acid sites for carbon skeleton branching. Common catalysts include platinum supported on chlorinated alumina (Pt/Al2O3) or zeolite-based systems like mordenite, operating in the presence of hydrogen to suppress cracking and coke formation.[15][16][17] The reaction proceeds at temperatures of 140–250°C and pressures of 100–1000 psig, conditions that favor the endothermic isomerization equilibrium toward isobutane, with lower temperatures enhancing selectivity despite slower kinetics. Yields typically approach thermodynamic equilibrium limits of 45–50 wt% isobutane from pure n-butane feeds, optimized by recycling unreacted n-butane and separating products via fractional distillation in a deisobutanizer column; overall process efficiencies exceed 90% recovery of butane isomers with minimal byproducts like propane or pentanes under controlled space velocities. Energy inputs involve heating for reaction and compression, with recent catalyst designs reducing required severities to lower operational costs and emissions.[18][15][19] Commercial development of these processes accelerated post-1940s, exemplified by early patents for aluminum chloride or supported phosphoric acid catalysis, enabling scalable production of high-purity isobutane (>95%) for petrochemical feedstocks and supplanting inconsistent natural occurrences in crude oil or gas streams. This refinery-centric shift addressed surging demand for isobutane in alkylation units producing high-octane gasoline components, with isomerization capacity expanding alongside shale gas booms to balance n-butane surpluses.[20][16] Advancements since the 2010s emphasize eco-friendly, non-chlorinated catalysts such as sulfated metal oxides (e.g., SI-3 or ISOMALK-3 systems), which mitigate corrosion and effluent treatment needs while maintaining or improving yields through enhanced acid site stability and reduced deactivation rates. These innovations lower energy consumption per ton of isobutane by enabling milder conditions and longer catalyst cycles, aligning with emission reduction goals in modern refineries.[15][21]Sources and Extraction
Isobutane occurs naturally as a component of natural gas and crude oil, primarily within natural gas liquids (NGLs) derived from wet natural gas fields. In raw natural gas, its concentration is typically low, ranging from trace amounts to 0.02%, though richer gas fields yield higher NGL content where isobutane constitutes 6-7% of the processed NGL barrel.[22][16] Extraction begins with NGL recovery from natural gas streams via cryogenic processes like turbo-expansion or absorption, capturing 75-90% of butane components, followed by preliminary separation to isolate heavier hydrocarbons from methane and lighter gases.[23][24] Further purification employs fractional distillation in multi-stage towers, leveraging boiling point differences (isobutane at -11.7°C versus n-butane at -0.5°C) to achieve separation without relying on synthetic isomerization for natural sources. This process is economically viable due to the abundance of associated natural gas in regions like U.S. shale plays and Middle Eastern fields, where operational costs remain low relative to output volumes. Refinery gas streams from crude oil processing also contribute, though NGLs dominate primary recovery.[25][26] Global isobutane supply is led by U.S. producers, who accounted for significant butane exports reaching record levels in 2024, and Middle Eastern nations leveraging vast gas reserves. Production economics favor these areas, with market expansion projected at a 6.1% CAGR through the 2030s, fueled by LPG blending demand amid steady natural gas output growth. High-purity grades (99%+ for specialized uses) are obtained via refined fractionation, ensuring minimal contaminants like n-butane or pentanes.[27][28][29][1]Applications
Refrigeration and Heat Transfer
Isobutane, designated as refrigerant R-600a, serves as a hydrocarbon working fluid in vapor-compression refrigeration cycles, particularly in domestic appliances, due to its thermodynamic properties that yield a coefficient of performance (COP) superior to that of R-134a under comparable conditions.[30][31] Its boiling point of -11.7°C and critical temperature of 134.7°C enable efficient heat absorption and rejection, with volumetric cooling capacity closely matching R-134a while requiring smaller compressor displacements.[32] Adopted in household refrigerators since the late 1990s following the phaseout of CFCs and early HCFCs, R-600a now powers over 75% of global domestic refrigerator production as of 2020, with more than one billion units in use by 2018.[33][34] R-600a's environmental profile, featuring a global warming potential (GWP) of 3 and zero ozone depletion potential (ODP), aligns with regulatory mandates such as the U.S. EPA's technology transition rules under the AIM Act, which from January 1, 2025, prohibit new refrigeration equipment using HFCs with GWP exceeding 700 in applicable sectors.[35][36] This low GWP facilitates compliance without sacrificing cycle efficiency, as empirical tests demonstrate R-600a systems achieving up to 10-20% lower energy consumption than R-134a equivalents in household applications.[30] However, its classification as an A3 refrigerant under ASHRAE Standard 34—indicating higher flammability—necessitates charge restrictions to mitigate ignition risks from leaks, with international standards like IEC 60335-2-24 limiting household units to 150 grams.[37] Safety protocols under ASHRAE 15, including leak detection and ventilation requirements, have enabled safe real-world deployment, with incident data from millions of units showing negligible fire risks when charges remain below limits and systems incorporate flame-arresting components. Recent advancements, including IEC 60335-2-89 revisions in 2019 raising commercial self-contained system limits to 500 grams for A3 refrigerants, have spurred adoption in small-scale commercial refrigeration from 2023 onward, offsetting higher initial safety engineering costs through reduced lifecycle energy and emissions.[38] Empirical assessments confirm that these expansions maintain risk levels comparable to non-flammable alternatives when causal factors like leak propagation and ignition probability are controlled via design.[39]Fuels, Propellants, and Solvents
Isobutane serves as a key component in liquefied petroleum gas (LPG), typically comprising part of the hydrocarbon mixture alongside propane and n-butane, enabling efficient storage and combustion for heating and portable fuel applications.[40] Its branched molecular structure contributes to a higher vapor pressure than n-butane—approximately 3.4 bar at 20°C—facilitating reliable vaporization and delivery in fuel systems, particularly in colder conditions where linear isomers may underperform.[41] This property arises from weaker intermolecular forces in the compact, tetrahedral configuration, promoting phase transition without external heating.[1] In gasoline production, isobutane is primarily employed in alkylation processes, reacting with olefins to yield alkylate—a high-octane blending stock with research octane numbers (RON) exceeding 90, often approaching 95-100, which mitigates engine knock by slowing flame propagation compared to straight-chain hydrocarbons.[16] [42] The anti-knock benefit stems from the molecule's tertiary carbon, which resists autoignition under compression, allowing higher compression ratios in spark-ignition engines for improved efficiency.[43] Blending isobutane-derived components elevates overall pool octane, supporting premium fuels rated above 100 RON while complying with volatility standards like Reid vapor pressure limits.[44] As an aerosol propellant, isobutane has gained prominence since the 1990s phaseout of chlorofluorocarbons (CFCs) and hydro chlorofluorocarbons (HCFCs), offering zero ozone depletion potential and a global warming potential (GWP) of zero, unlike HFC alternatives.[45] [46] Its vapor pressure enables fine droplet atomization in sprays and foams, with consistent performance across formulations due to rapid expansion upon release.[47] Market data indicate sustained growth, with the global isobutane sector projected to reach $21.15 billion in 2025, driven by adoption in personal care and household products seeking low-impact alternatives amid HFC restrictions.[28] Hydrocarbon propellants like isobutane now dominate non-medical aerosols, with the segment expanding at a 6.9% CAGR through 2034 for eco-compliant foams and sprays.[48] In solvent applications, isobutane's non-polar nature—evidenced by its low dielectric constant—allows selective dissolution of lipophilic compounds in extractions, such as terpenes and cannabinoids from plant material, yielding purer isolates upon evaporation.[49] [50] However, its high volatility (boiling point -11.7°C) limits use to closed-loop systems, where rapid recovery minimizes residue but heightens flammability risks during handling.[1] This makes it suitable for precision cleaning in industrial settings, though safer for targeted, low-residue tasks rather than broad-surface applications.[51]Chemical Intermediates and Feedstocks
Isobutane serves as a primary feedstock in the alkylation process, where it reacts with C3–C5 olefins such as propylene or butenes in the presence of strong acid catalysts like sulfuric acid or hydrofluoric acid to yield high-octane alkylate, predominantly consisting of trimethylpentanes such as isooctane (2,2,4-trimethylpentane).[5] This reaction proceeds via carbocation intermediates, with kinetics favoring hydride transfer from isobutane to olefin-derived cations, achieving selectivities to desired C8 alkylates of 90–95% under optimized conditions, though side reactions like polymerization reduce overall yields to 1.5–2.0 barrels of alkylate per barrel of olefin feedstock.[52] Industrial processes, such as those using HF catalysis, demonstrate higher isobutane efficiency (lower consumption per unit alkylate) compared to H2SO4, with consumption rates 6–15% lower for equivalent production.[52] Alkylation units leverage refinery-derived isobutane streams, enabling scalable output of gasoline blending components with octane numbers exceeding 92 RON.[16] Dehydrogenation of isobutane to isobutylene occurs via catalytic processes like Honeywell UOP's C4 Oleflex technology, which employs platinum-tin catalysts on alumina in adiabatic reactors at 550–600°C, yielding up to 85% isobutylene selectivity at 50–60% conversion per pass due to favorable endothermic kinetics and continuous catalyst regeneration.[53] The resulting isobutylene is further converted to oxygenates like methyl tert-butyl ether (MTBE) or ethyl tert-butyl ether (ETBE) by reacting with methanol or ethanol, with industrial plants achieving capacities of 400,000–500,000 metric tons per year from refinery C4 fractions.[54] Isobutylene also feeds polymerization to polyisobutylene, used in adhesives and sealants, with chain-growth kinetics controlled by cationic initiators to produce high-molecular-weight polymers scalable from dehydrogenation outputs.[55] Liquid-phase autoxidation of isobutane with molecular oxygen produces tert-butyl hydroperoxide (TBHP) non-catalytically at 120–140°C and 25–35 bar, initiating radical chain reactions that achieve TBHP selectivities of 60–70% at 10–15% conversion, limited by consecutive decomposition to tert-butanol and acetone.[56] This intermediate enables epoxidation routes, such as in propylene oxide synthesis, where TBHP reacts with propylene over titanium-silicalite catalysts, highlighting isobutane's role in oxygenated chemical chains derivable from abundant alkane feedstocks.[57] Catalytic cracking of isobutane over zeolites like HZSM-5 or Fe-modified variants generates propylene through carbenium ion mechanisms at 500–600°C, with yields up to 40–50 wt% propylene at high space velocities, though commercial adoption remains limited compared to naphtha cracking due to lower overall economics from lighter feeds.[58] For butyl rubber production, dehydrogenated isobutylene copolymerizes with 1–3% isoprene via cationic polymerization in methyl chloride at –95°C, initiated by AlCl3 complexes, yielding elastomers with molecular weights of 200,000–500,000 g/mol and scalability tied to isobutane refinery availability.[59]Safety and Health Considerations
Toxicity and Exposure Risks
Isobutane demonstrates low acute toxicity in mammalian species, with inhalation LC50 values exceeding 658,000 mg/m³ over 4 hours in rats and greater than 570,000 ppm over 15 minutes, indicating that lethality requires extremely high concentrations primarily due to oxygen displacement rather than inherent chemical toxicity.[60][61] As a simple asphyxiant, isobutane poses risks when its volume fraction surpasses 10% in air, reducing oxygen levels below 19.5% and potentially causing central nervous system depression, dizziness, unconsciousness, or death through hypoxia, though effects may onset at concentrations above 1% (10,000 ppm) with symptoms like headache and narcosis.[62][1] Direct contact yields minimal adverse effects, with no significant skin or eye irritation observed in standard tests, and oral or dermal LD50 values exceeding 2,000 mg/kg in rats, underscoring its low inherent toxicity beyond asphyxiation hazards.[63][64] Occupational exposure guidelines reflect this profile, including a NIOSH recommended exposure limit of 800 ppm (1,900 mg/m³) as an 10-hour time-weighted average and an ACGIH threshold limit value of 1,000 ppm, both predicated on preventing central nervous system impairment and oxygen displacement rather than organ-specific damage.[65][66] Chronic exposure at ambient or occupational levels shows negligible systemic effects, with animal studies revealing no evidence of carcinogenicity, mutagenicity, or reproductive toxicity for the pure compound, though European classifications under certain harmonized criteria may apply to commercial grades containing trace impurities like 1,3-butadiene, a known carcinogen not intrinsic to isobutane.[67][68] Recent evaluations, including the Cosmetic Ingredient Review's 2023 assessment deeming isobutane safe in cosmetics at concentrations up to 98% and the European Food Safety Authority's 2025 re-evaluation confirming no consumer safety concerns for its use as a food additive propellant, affirm its low risk profile in formulated products under typical exposure scenarios.[61][69]Flammability and Operational Hazards
Isobutane exhibits high flammability, classified under the NFPA 704 system with a flammability rating of 4, indicating it burns readily and vaporizes completely at normal ambient temperatures and pressures.[7] Its autoignition temperature is 460 °C, above which spontaneous combustion can occur in air without an external spark.[70] The lower explosive limit (LEL) is 1.8% by volume in air, and the upper explosive limit (UEL) is 8.4%, defining the concentration range where ignition can propagate a flame.[70] [7] Upon combustion, isobutane releases approximately 49.5 MJ/kg of energy, slightly lower than propane's 50.5 MJ/kg, resulting in comparatively reduced thermal output per unit mass during deflagration events.[71] Operational hazards primarily arise from unintended leaks forming ignitable mixtures, particularly in enclosed spaces, where static discharge or hot surfaces can serve as ignition sources. Safe handling protocols emphasize adequate ventilation to disperse vapors below the LEL and electrical grounding of equipment and containers to mitigate static spark risks, as ungrounded transfers can generate electrostatic discharges capable of ignition.[72] [73] In refrigeration systems, where isobutane serves as a working fluid, international standards such as IEC 60335-2-89 specify charge limits (e.g., up to 150 g for certain domestic units), component spacing to prevent arc ignition, and fault-tolerant designs to contain leaks.[74] Quantitative risk assessments reveal that explosion incidents remain rare in controlled applications, with ignition frequencies estimated at less than 10^{-7} per appliance-year for hydrocarbon-charged refrigerators under normal operation, far below theoretical hazards due to dilution effects and absence of competent ignition sources in typical domestic environments.[75] Empirical data from field deployments of isobutane-based systems corroborate this, showing negligible explosion rates attributable to adherence to engineering mitigations rather than inherent material volatility alone. These findings underscore that overemphasis on flammability often overlooks causal factors like system integrity and procedural compliance in assessing practical dangers.Environmental Impact
Atmospheric Effects and GWP
Isobutane undergoes rapid atmospheric degradation primarily through reaction with hydroxyl (OH) radicals in the troposphere, yielding an estimated lifetime of 7 days (range: 5.2–10.7 days).[76] This short residence time minimizes its accumulation and direct radiative forcing, resulting in a 100-year global warming potential (GWP) of approximately 3 relative to CO₂.[77] In contrast, hydrofluorocarbon (HFC) refrigerants such as R-410A exhibit GWPs exceeding 2,000, driven by their longer atmospheric persistence and stronger infrared absorption.[35] As a volatile organic compound (VOC), emitted isobutane serves as a precursor to tropospheric ozone formation via photochemical oxidation in the presence of nitrogen oxides (NOx), potentially exacerbating local smog under high-insolation conditions.[78] Its photochemical ozone creation potential (POCP) is moderate compared to alkenes but higher than that of many fluorocarbons, which degrade into products with limited radical propagation.[79] Nonetheless, lifecycle analyses indicate that isobutane's deployment in low-emission, closed-loop systems—such as refrigeration—yields net climate benefits over HFCs, owing to the latter's dominant GWP contributions from even minor leaks.[80] Empirical data underscore hydrocarbons' role as viable low-GWP alternatives, aligning with assessments favoring reduced reliance on persistent synthetics.[81]Regulatory Framework and Usage Limits
In the United States, the Environmental Protection Agency's Significant New Alternatives Policy (SNAP) program approves isobutane (R-600a) as an acceptable substitute refrigerant in specific applications, such as household refrigerators, freezers, and certain retail food refrigeration systems, subject to use conditions that limit charge sizes to mitigate flammability risks—typically up to 57 grams in household units and higher in commercial settings with requirements for leak detection and ventilation.[82] [83] These approvals persist despite isobutane's A3 flammability classification, reflecting empirical data from decades of global use showing rare incidents; for instance, millions of hydrocarbon-equipped appliances in Europe and Japan have recorded minimal fires or explosions attributable to refrigerant leaks when standards are followed.[84] The ongoing HFC phasedown under the American Innovation and Manufacturing Act, with restrictions on high-GWP refrigerants effective January 1, 2025, is accelerating adoption of hydrocarbons like isobutane in new equipment, though flammability clauses continue to impose design constraints that may exceed risks evidenced by low incident rates.[85] In the European Union, the F-Gas Regulation (EU) 2024/573 phases down HFCs through quotas and bans on high-GWP uses, promoting low-GWP alternatives like isobutane in refrigeration and heat pumps, while the Renewable Energy Directive (RED II) incentivizes efficient, low-impact systems that align with hydrocarbon deployment.[86] [87] Charge limits for flammable hydrocarbons remain stringent under harmonized standards, such as up to 150 grams in domestic systems and approximately 2.6 kilograms in commercial refrigerated display cabinets with secondary loop designs or enhanced safety measures, derived from probabilistic modeling of ignition scenarios rather than aggregated incident data, which indicates hydrocarbons have caused only isolated injuries globally over 30 years of widespread application.[74] [88] Recent updates, including increased charge allowances via revised IEC 60335-2-89, support broader use in heat pumps, yet compliance with explosion-proof components and risk assessments elevates upfront costs, potentially hindering market penetration despite favorable thermodynamics and near-zero GWP.[89] Internationally, ISO 5149 outlines safety requirements for refrigerating systems using hydrocarbons, with 2023-2025 revisions incorporating flammability research to expand applicability in heat pumps and process cooling, allowing higher charges (up to several kilograms) in ventilated or sealed systems based on occupancy and leak propagation models.[90] [91] These standards emphasize empirical risk mitigation over blanket prohibitions, aligning with safety records where hydrocarbon incidents remain statistically negligible—fewer than one major event per million units annually—suggesting that charge caps rooted in conservative assumptions may impose unnecessary economic barriers, such as added sensor and enclosure expenses, without proportional safety gains.[92] [93]Nomenclature and Isomers
Naming Conventions
The systematic International Union of Pure and Applied Chemistry (IUPAC) name for isobutane is 2-methylpropane, denoting a three-carbon propane backbone with a methyl group attached to the central carbon atom.[1][10] This nomenclature adheres to rules for unbranched and branched alkanes, prioritizing the longest continuous chain while identifying substituents./Alkanes/Nomenclature_of_Alkanes) The common name isobutane incorporates the "iso-" prefix, traditionally applied to denote the branched isomer of alkanes with a single methyl branch at the second carbon position, reflecting its structural deviation from the straight-chain n-butane despite sharing the C₄H₁₀ molecular formula./Alkanes/Nomenclature_of_Alkanes)[94] This naming convention, retained for practical use, has origins in early organic chemistry where "iso-" signified compounds with identical empirical formulas but different properties due to isomerism.[94] Isobutane's Chemical Abstracts Service (CAS) registry number is 75-28-5, and its molar mass is 58.12 g/mol.[1][10] In commercial and industrial contexts, the term "isobutane" specifies the pure branched isomer, differentiated from generic "butane" designations that typically refer to n-butane or blended hydrocarbon mixtures for fuels and propellants.[95] This distinction ensures clarity in applications requiring the specific physical properties of the branched structure, such as lower boiling point compared to n-butane mixtures.[95]Comparison to n-Butane
Isobutane, or 2-methylpropane, features a branched carbon skeleton with a central carbon atom bonded to one hydrogen and three methyl groups, contrasting with the linear chain of n-butane, which consists of four carbons in a straight sequence. This structural branching causally reduces intermolecular van der Waals forces due to lower surface area contact, resulting in distinct physical properties. Specifically, isobutane has a boiling point of -11.7°C compared to -0.5°C for n-butane, enabling easier vaporization and separation via fractional distillation.[95] Liquid density at 20°C is lower for isobutane (0.549 g/cm³) than n-butane (0.579 g/cm³), attributable to the compact, spherical shape hindering efficient molecular packing in the liquid phase.[96]| Property | Isobutane | n-Butane |
|---|---|---|
| Boiling Point (°C) | -11.7 | -0.5 |
| Liquid Density (g/cm³ at 20°C) | 0.549 | 0.579 |
| Research Octane Number (RON) | 102 | 94 |