Fact-checked by Grok 2 weeks ago

Propylene

Propylene, also known as propene, is a three-carbon and an unsaturated with the molecular C₃H₆ and a molecular weight of 42.08 g/mol. It appears as a colorless gas with a faint petroleum-like at , though it is often shipped and stored as a under its own . Propylene is highly flammable, with a of -47.6°C and a of -185.2°C, and it has low in (approximately 200 mg/L at 25°C) but is miscible with many organic solvents. As one of the most important , it serves primarily as a feedstock for producing , a versatile that accounts for about 70% of global propylene consumption as of 2023 and represents approximately 20% of all plastics production. Propylene is produced on an industrial scale mainly as a by-product of refining and the of feedstocks to produce , with additional production via the catalytic dehydrogenation of . Global production reached approximately 130 million metric tons annually as of 2024, with high-purity grades (99.5–99.8%) used for applications and lower grades for chemical or uses. Beyond , key derivatives include (for polyurethanes and ), (for synthetic fibers and rubbers), (for phenol and acetone), and isopropanol (for solvents and antiseptics), making it a cornerstone of the . It also finds minor applications as a additive, , and in synthetic production.

Properties

Physical properties

Propylene has the molecular C₃H₆ and a molecular weight of 42.08 g/. Its is CH₂=CH–CH₃, consisting of a three-carbon chain with a carbon-carbon between the first and second carbons, resulting in an unsymmetric arrangement due to the terminal position of the double bond and the attached . At standard conditions, propylene appears as a colorless gas with a faint petroleum-like . It has a of −185.2 °C and a of −47.6 °C at 1 . The critical temperature is 91.9 °C, and the critical is 4.62 . The density of propylene gas is 1.85 kg/m³ at 0 °C and 1 , while the density at the is 0.583 g/cm³. Its is 760 mmHg at −47.6 °C. Propylene exhibits low in (approximately 200 mg/L at 25 °C) but is soluble in organic solvents such as and . Key thermodynamic and safety-related properties include an of 457.8 °C, a of −108 °C (open cup), a heat of vaporization of 18.9 kJ/mol at the , and a of 1.76 J/g·K for the gas at 25 °C. These properties influence its handling as a under pressure in industrial applications.

Chemical properties

Propene, also known as propylene, is classified as the second simplest after and is a gaseous unsaturated featuring a single carbon-carbon in its molecular formula C₃H₆. The molecule adopts a structure where the double bond connects the first and second carbon atoms, with the third carbon bearing a , resulting in the systematic name prop-1-ene. The carbons involved in the C=C exhibit sp² hybridization, forming a trigonal planar with bond angles near 120°, while the arises from the sideways overlap of unhybridized p orbitals, creating an electron-rich cloud above and below the molecular plane that renders the nucleophilic. This facilitates reactivity toward electrophiles, enabling characteristic addition reactions at the . Although nonpolar overall due to its nature, propene possesses a small of 0.366 D, arising from the asymmetry introduced by the adjacent to the , which slightly polarizes the distribution. Propene is chemically stable under ambient conditions, resisting spontaneous , yet the imparts high reactivity toward electrophilic reagents, distinguishing it from saturated hydrocarbons like . Regarding isomerism, propene lacks geometric (cis-trans) isomers because one of the double-bonded carbons bears two identical atoms, preventing the necessary structural distinction for such ; it also exhibits no optical isomers due to the absence of a chiral center. Thermodynamically, the for gaseous propene is ΔH_f° = +20.4 kJ/mol, reflecting the endothermic nature of its formation from elements compared to alkanes, while the of formation is ΔG_f° = +62.8 kJ/mol, indicating its relative instability under standard conditions. Propene displays weak basicity at the , allowing coordination with acids, and the allylic C-H bonds exhibit mild acidity with a pK_a ≈ 43, lower than typical C-H bonds (pK_a > 50) due to stabilization of the resulting allylic anion.

History and occurrence

Historical development

Propylene was first isolated in by John Reynolds, a student of the chemist August Wilhelm von Hofmann, as the sole gaseous product resulting from the of passed over a hot wire. This discovery marked an early advancement in understanding in . In 1851, Hofmann recognized and named the compound propene, distinguishing it from other hydrocarbons based on its properties and reactivity. The of propene, CH₃CH=CH₂, was elucidated in the late amid broader progress in organic structural theory, confirming its identity as an with a terminal . A major breakthrough in propylene's industrial significance occurred in 1951, when chemists J. Paul Hogan and Robert L. Banks at accidentally produced crystalline . While investigating catalysts for converting propylene into gasoline additives, they passed propylene over a chromium oxide-silica-alumina catalyst at elevated temperatures and pressures, yielding a white, solid rather than the expected oligomers. This serendipitous finding highlighted propylene's potential for forming high-molecular-weight, crystalline polymers, though initial characterization revealed it as atactic and of limited utility. Independently, in 1954, Italian chemist advanced the field by developing stereospecific polymerization of using Ziegler-Natta catalysts, consisting of titanium trichloride and aluminum alkyls. This method produced isotactic , with regular arrangements enabling crystallinity, high strength, and properties suitable for commercial applications. Natta's innovations, building on Karl Ziegler's foundational work on olefin polymerization, earned him the 1963 , shared with Ziegler, for their contributions to stereoregular . Commercialization accelerated rapidly following these discoveries. The first large-scale production of polypropylene began in 1957, led by Montecatini in (using Natta's process), Hercules Incorporated , and in Germany. Phillips Petroleum initially commercialized from their catalyst system but shifted focus to polypropylene amid growing demand. Patent disputes delayed recognition of and Banks' contributions; their U.S. patent for crystalline polypropylene was finally granted in 1983 after over 30 years of litigation against competitors like Montecatini, affirming Phillips' priority in the atactic form while Natta's work dominated isotactic production. The 1960s saw rapid expansion of polypropylene manufacturing due to its low production costs, versatility, and superior properties compared to earlier plastics like . Global capacity grew from modest pilot-scale operations to over 500,000 tons annually by 1965. By the , production exceeded 1 million tons per year, reaching approximately 3.5 million tons by 1975, driven by applications in , textiles, and automotive parts.

Natural occurrence

Propene, commonly known as propylene (C₃H₆), is a naturally occurring present in trace amounts in various terrestrial environments. It is emitted by vegetation as part of plant metabolism, with documented occurrences in species such as Senna sophera. Propene is also released from volcanic activity and incomplete of . Additionally, certain , including marine propylene-assimilating strains like species, produce or metabolize propene as a carbon and energy source during aerobic degradation processes. In the Earth's atmosphere, propene exists at low concentrations, generally ranging from 0.1 to 5 ppb in background levels, derived mainly from biogenic emissions by and from biomass burning events. These biogenic sources contribute significantly to non-anthropogenic volatile compounds (VOCs), where propene acts as a precursor in , though its biological role remains limited to metabolic byproducts in and microbial degradation pathways. Propene occurrences are cataloged in natural products databases such as , highlighting its presence in biological systems. Geologically, propene forms as a minor constituent during the of in sedimentary rocks, appearing in trace quantities within deposits and reservoirs. Beyond , propene has been detected in settings, including the atmosphere of Saturn's moon via NASA's Cassini mission in 2013, and in the through observations of the TMC-1. It has also been identified in cometary environments, such as the dusty coma of comet 67P/Churyumov-Gerasimenko.

Production

Steam cracking

Steam cracking is the dominant industrial method for producing propylene, involving the thermal of feedstocks in the presence of to generate olefins. The process occurs in tubular reactors within cracking furnaces, where the feedstock—typically , , , or gas oil—is mixed with and heated to temperatures between 750°C and 900°C at low pressures of 0.1 to 0.3 . The dilution, maintained at a of 0.5 to 1.0 ( to by weight), serves to reduce , minimize coke formation on reactor walls, and enhance selectivity toward desired products by suppressing secondary reactions. The underlying mechanism consists of free radical chain reactions that cleave carbon-carbon bonds in the hydrocarbons. begins with the thermal dissociation of feedstock molecules into free radicals, such as the splitting of into two methyl radicals. follows as these radicals abstract or add to other molecules, forming smaller radicals and stable products like and propylene, while chain branching can amplify radical concentrations. Termination occurs when radicals combine to yield stable saturated or unsaturated hydrocarbons, including byproducts such as and aromatics. This radical pathway favors the production of olefins but also generates , , and higher hydrocarbons as side products. Yields of propylene vary by feedstock, typically ranging from 10% to 20% by weight from , with higher selectivity—up to around 40% from under optimized conditions—due to the direct dehydrogenation-like cracking of . Steam cracking accounts for approximately 50% of global propylene production as of 2023, often integrated with manufacture, where propylene serves as a valuable co-product alongside as the primary target. The process is highly -intensive, consuming 30 to 40 per ton of product, primarily due to the high-temperature furnaces and subsequent separation steps. In facilities, the cracked gas is rapidly quenched via transfer line exchangers to halt reactions, followed by compression, cooling, and to isolate propylene; no catalysts are employed, distinguishing it as a purely . These plants are commonly co-located with production units for in handling shared . Advantages include high volumes and flexibility across gaseous or feedstocks, enabling adaptation to availability. However, the is energy-consuming and emits significant CO₂, contributing to environmental challenges in operations.

Olefin conversion technology

Olefin conversion technology (OCT) involves the catalytic metathesis of lighter olefins, such as ethylene, with heavier olefins, like 2-butene, to produce propylene on purpose. This process, also known as olefin disproportionation, rearranges the carbon-carbon double bonds in a reversible, equilibrium-limited reaction represented by: \ce{C2H4 + C4H8 <=> 2 C3H6} The reaction typically employs a 1:1 molar ratio of ethylene to 2-butene, with excess ethylene often used to shift the equilibrium toward propylene formation and suppress side reactions like self-metathesis of ethylene to ethane. Supported metal oxide catalysts, such as oxide on alumina (WO₃/Al₂O₃) or oxide on silica (MoO₃/SiO₂), facilitate the metathesis at temperatures of 200–400 °C and pressures of 2–3 . These heterogeneous catalysts operate in fixed-bed reactors, where the reaction mixture passes through the catalyst bed, followed by product separation and of unreacted olefins to maximize conversion. Early developments include the Triolefin from the 1960s, which originally converted to and 2-butene using WO₃/SiO₂ catalysts, but the reversible nature of metathesis enabled its adaptation for propylene production in modern OCT implementations by companies like Lummus and Toyo . The process achieves propylene selectivity up to 90–95%, with per-pass 2-butene conversion exceeding 60%, depending on activity and operating conditions. Globally, OCT contributes approximately 3% to propylene supply as of 2019, often integrated with crackers to utilize byproduct and butenes, thereby enhancing overall plant efficiency. Its advantages include operational flexibility to adjust olefin ratios and lower requirements compared to cracking processes, as the metathesis step is exothermic and needs no external heat input. However, deactivation by poisons such as , oxygenates, or compounds necessitates careful feedstock pretreatment and periodic regeneration.

Fluid catalytic cracking

Fluid catalytic cracking (FCC) is a key refinery process that generates propylene as a valuable during the conversion of heavy feedstocks, such as gas or atmospheric residues, into lighter fractions like and (LPG). The employs zeolite-based catalysts, primarily Y-type ( Y with Si/ ratio of 5-8), in a fluidized-bed reactor system consisting of a riser reactor and a regenerator. In the riser, preheated feedstock (around 300°C) is injected and contacts hot catalyst particles (500-540°C) at temperatures of 480-550°C and pressures of 0.1-0.3 , with catalyst-to- ratios of 5-9, enabling rapid and cracking within a short contact time of 1-5 seconds. The endothermic cracking reaction is balanced by heat supplied from combustion in the regenerator, which operates at approximately 700°C. The mechanism of propylene formation in FCC proceeds via acid-catalyzed cracking on Brønsted acid sites of the catalyst, generating () intermediates from the heavy hydrocarbons. These intermediates undergo β-scission and hydrogen transfer reactions, yielding propylene (=) at 15-25 wt% of the feed in maximum olefins mode, alongside other olefins and paraffins. The riser design promotes short residence times to minimize overcracking and secondary reactions, enhancing selectivity toward light olefins like propylene. Incorporation of additives (at 25-50% loading relative to the base catalyst) further boosts propylene yields by promoting secondary cracking of gasoline-range hydrocarbons into C3-C4 olefins, particularly in maximum olefins (MOG) mode, where yields can reach up to 15 wt% propylene increment. In integration, FCC units process 2,000-10,000 tons of feed per day, with propylene recovered from the LPG stream post-separation from and other products via columns. This coproduction aligns FCC with demands, as the primary output is transportation , while propylene serves needs. Globally, FCC accounts for approximately 30% of propylene production, providing a significant but variable supply tied to refining operations. The advantages of propylene via FCC include its scalability in existing refinery infrastructure and flexibility to adjust yields based on market demand for fuels versus chemicals, facilitated by catalyst additives and operating conditions. However, disadvantages arise from its byproduct status, leading to yield variability (influenced by feedstock quality and refinery economics) and challenges like catalyst deactivation by (0.8-1.3 wt%) or metal contaminants, necessitating continuous regeneration.

Other commercialized methods

Global propylene production exceeded 130 million metric tons in 2023, with on-purpose routes like PDH and MTO growing to approximately 20-25% combined share as of 2024. Propane dehydrogenation (PDH) represents a key on-purpose method for propylene production, involving the endothermic conversion of propane (C₃H₈) to propylene (C₃H₆) and hydrogen (H₂) at temperatures of 550–650 °C and low pressure. This non-oxidative process typically employs chromium oxide (Cr₂O₃) supported on alumina catalysts for fixed-bed operations or platinum-tin (Pt/Sn) bimetallic catalysts on alumina for more flexible configurations, achieving propane conversions of 40–45% and propylene selectivities around 85%, resulting in yields of 30–50%. Commercial implementations include the Catofin process, a cyclic fixed-bed technology licensed by Lummus Technology, which uses multiple reactors in regeneration cycles, and the Oleflex process by UOP Honeywell, featuring a continuous moving-bed design with radial-flow reactors for improved efficiency and lower energy use. These technologies account for approximately 10–15% of global propylene production capacity, with significant growth in regions like China and the US driven by abundant propane from shale gas. The -to-olefins (MTO) process provides another established route, converting derived from , , or into light olefins, including propylene, over silicoaluminophosphate (SAPO-34) catalysts in a fluidized-bed reactor at 400–500 °C. This dual-cycle mechanism proceeds via initial dehydration to and water, followed by pool intermediates that favor olefin formation, yielding and propylene selectivities combined at around 80%, with propylene comprising 35–50% of the olefin output depending on catalyst modifications. Widely commercialized in since the first DMTO plant in in 2010, the process has expanded to over two dozen units with a total capacity exceeding 20 million tons per annum of and propylene as of 2024, leveraging 's coal-to- infrastructure to supplement traditional oil-based routes. Bio-based propylene production remains an emerging but minor commercial pathway, typically involving the of sugars to isopropanol followed by catalytic , or indirect routes via to alcohols and subsequent upgrading, contributing less than 1% to the global market. For instance, processes like those developed by companies such as explore microbial conversion of renewable feedstocks to isopropanol, which is then dehydrated over acid catalysts to yield propylene, offering a renewable alternative with potential carbon-negative footprints when integrated with gas . Prereforming via of is a less common method, primarily explored for generation but adaptable for selective propylene formation through controlled oxygen addition, though it faces scalability issues due to side reactions producing and CO₂. These on-purpose methods offer advantages such as reduced dependence on crude oil by utilizing abundant , coal-derived , or , while enabling higher propylene yields compared to byproduct routes like . However, challenges include high capital and operational costs—particularly for PDH's energy-intensive heating and catalyst regeneration—and management, as the coproduced H₂ requires separation or valorization to maintain economics, alongside environmental concerns from CO₂ emissions in oxidative variants. MTO and bio-routes mitigate some oil reliance but contend with feedstock price volatility and lower current scalability.

Commercial aspects

Market overview

Global propylene production reached approximately 130 million metric tons in 2024, with projections estimating an increase to around 134 million tons in 2025, driven by capacity expansions primarily in Asia. The market is expected to grow at a (CAGR) of 2.6% from 2025 to 2030, reflecting steady demand amid new on-purpose production facilities offsetting traditional supply sources. The propylene market was valued at USD 116.23 billion in 2025 and is forecasted to reach USD 162.03 billion by 2034, expanding at a CAGR of 3.76% during this period. Demand is predominantly driven by the sector, with accounting for about 60% of consumption due to its widespread use in , automotive, and consumer goods. Regionally, holds the largest share as both producer and consumer, commanding around 30% of the global market, while the and each represent approximately 20%, supported by established infrastructures. Key players in propylene production include Industries, , and , which collectively dominate global supply through integrated and cracking operations. The relies heavily on oil-derived feedstocks for about 70% of production via naphtha-based , 20% from via propane dehydrogenation, and 10% from coal-to-olefins routes, particularly in . In the United States, polymer-grade propylene (PGP) spot prices averaged $0.30–0.40 per pound in 2025, exhibiting volatility influenced by crude oil and naphtha fluctuations, as well as unplanned outages at refineries and crackers. International trade features significant exports from the and to meet global needs, with the region supplying over 40% of seaborne volumes. However, 2025 has seen shortages in the market due to supply constraints from maintenance turnarounds and limited availability, prompting higher imports. Post-2020 trends include a shift toward propane dehydrogenation (PDH) and methanol-to-olefins (MTO) technologies to counter oil price volatility, alongside initiatives that are gradually reducing reliance on energy-intensive processes.

Research and innovations

Recent research in propylene production emphasizes sustainable methods to reduce reliance on fossil fuels and minimize carbon emissions. One promising approach is CO₂-assisted oxidative dehydrogenation (ODH) of , which utilizes CO₂ as an oxidant to enhance selectivity and efficiency. Studies have demonstrated effective using composite metal oxides supported on , such as 10% MxOy-TiO₂ (where M includes Zr, Ce, Ca, Cr, or Ga), achieving propylene yields up to 20% under mild conditions while consuming CO₂ and mitigating . Additionally, gold-supported catalysts on Y-doped ceria have shown high activity for this process, promoting without excessive formation. For bio-based routes, hydrogenolysis of glycerol derivatives—derived from waste —offers a renewable pathway to bio-propylene. A novel catalyst developed in enables efficient of glycerol-derived to propylene with high selectivity, leveraging abundant byproducts. Catalyst advancements focus on improving selectivity and stability in key reactions. Single-site catalysts, including metallocene-inspired designs, enhance for propylene production by providing uniform active sites that boost selectivity toward desired olefins, as seen in metathesis systems. In propane dehydrogenation (PDH), bimetallic Pt-Sn catalysts outperform monometallic Pt, delivering higher propylene selectivity (up to 95%) and turnover rates due to alloying effects that suppress side reactions. Although AI-optimized PDH remains exploratory, advanced computational screening has identified promoter-modified zeolites achieving propylene yields exceeding 50% in lab-scale tests by tuning acid-base sites. Electrochemical methods are gaining traction for direct propene valorization, particularly oxidation to (PO) or to fuels, with emphasis on selectivity control from 2022 onward. Pd-based s enable propene to PO with faradaic efficiencies over 80%, where under anodic potentials tunes product distribution toward products. Facet-specific Ag₃PO₄ catalysts achieve near-100% selectivity to PO at ambient conditions, highlighting the role of crystal orientation in suppressing over-oxidation. Recent 2024-2025 studies on V-activated systems in electrode assemblies further improve PO yields by stabilizing intermediates during electro-epoxidation. Efforts toward a include propylene from plastic waste via , converting into recoverable monomers and olefins. processes yield up to 70% liquid products containing propylene precursors, enabling chemical to close the loop on waste. regulations, such as the Packaging and Packaging Waste Regulation, drive this transition by mandating 55% of plastic packaging by 2030 and full recyclability of all plastics, incentivizing advanced technologies to meet net-zero goals. Emerging technologies target integrated, low-emission synthesis. A 2022 breakthrough enables direct conversion of propane to PO using inert supports like boron nitride (BN) or SiO₂ at elevated temperatures, yielding PO with minimal CO₂ byproduct through gas-phase radical mechanisms. The hydrogen peroxide to propylene oxide (HPPO) process has seen scale-up, with pilot plants achieving 1 kt/a PO production using titanium silicalite-1 catalysts, offering >99% selectivity and water as the sole byproduct. Despite progress, challenges persist in bio-based propylene, where costs remain approximately twice those of routes due to feedstock variability and catalyst durability issues. Post-2020 research prioritizes , with bio-routes reducing GHG footprints by 45% compared to baselines, though full decarbonization requires electrification and carbon capture .

Uses

In polymer industry

The polymer industry represents the largest application for propylene, accounting for approximately 70% of global propylene consumption. (PP), the primary derived from propylene, is produced through the of propylene monomers, yielding a versatile with high tensile strength, excellent chemical resistance, and good fatigue resistance. Isotactic polypropylene, the most common form, is synthesized using Ziegler-Natta catalysts or metallocene catalysts, which enable stereospecific to achieve the desired crystalline structure. Global PP production reached about 70 million metric tons in 2024, driven by demand in various sectors. Polypropylene is available in several types tailored to specific needs, including homopolymers and copolymers. Homopolymer PP, consisting solely of propylene units, offers rigidity and is widely used in applications like rigid , such as containers and crates, due to its high stiffness and thermal stability. Copolymers, incorporating or other monomers, enhance resistance and flexibility; for instance, block copolymers are employed in automotive components like bumpers and interior parts, where under stress is essential. Commercial production of PP typically occurs via gas-phase or processes, where propylene is reacted in the presence of catalysts under controlled and conditions to form granules. Beyond standard , propylene is incorporated into other polymers such as polyallomers, which are copolymers of propylene and designed for improved clarity and impact strength compared to pure polypropylene, finding use in devices and transparent packaging. Another key derivative is , a terpolymer that includes propylene alongside and a for cross-linking, valued for its weather resistance and elasticity in applications like seals, hoses, and tires. In terms of end uses, polypropylene dominates in , which accounts for over 50% of demand through films, bottles, and flexible containers that benefit from its and barrier properties. The automotive sector consumes around 20% for components such as bumpers, dashboards, and under-the-hood parts, leveraging PP's and recyclability. Textiles and fibers represent about 15% of usage, including carpets, ropes, and non-woven fabrics for apparel and products, where PP's and processability are advantageous. Consumer goods, such as items and supplies, further expand its applications, underscoring propylene's central role in polymer-based materials.

In chemical synthesis

Propylene serves as a vital feedstock in the synthesis of numerous industrial chemicals and intermediates, comprising the remaining approximately 30% of global consumption. These derivatives find applications in sectors such as pharmaceuticals, detergents, and fuel additives, underscoring propylene's versatility in the petrochemical industry. A primary derivative is (), which accounts for approximately 8% of propylene demand and was produced at a global scale of about 10 million tonnes in 2024. is synthesized via the , where propylene reacts with to yield propylene chlorohydrin as an intermediate, subsequently hydrolyzed with to form , or through the -based HPPO process, which directly epoxidizes propylene using as the oxidant, generating only as a byproduct for improved environmental efficiency. serves as a precursor for polyether polyols used in foams and for propylene glycols obtained via its hydration, which are employed in de-icing fluids, , and . Acrylonitrile, another key product, is manufactured through the Sohio process, a catalytic ammoxidation of with and air over a phosphomolybdate catalyst in a fluidized-bed reactor, achieving high selectivity and accounting for a significant portion of propylene's non-polymer uses. This compound is essential for producing acrylic fibers, resins, and nylon precursors like . Cumene production consumes about 12% of propylene and involves the acid-catalyzed of with propylene, typically using catalysts in liquid-phase processes to yield (isopropylbenzene) with high selectivity. is then oxidized to produce phenol and acetone, critical intermediates for resins, plastics, and solvents. Isopropyl alcohol is derived from the direct hydration of propylene, where high-purity propylene reacts with water over a catalyst like or in either indirect (via sulfate esters) or direct vapor-phase processes, yielding the alcohol used as a , , and component. Other notable derivatives include , produced by the high-temperature chlorination of propylene, which serves as a precursor for used in resins and glycidyl ethers; , obtained via the selective of propylene over metal oxide catalysts like bismuth molybdate, acting as an intermediate for and ; and additional propylene glycols for broader applications in pharmaceuticals and detergents. As of 2025, demand for propylene in sustainable applications, such as bio-propylene for , is growing due to environmental regulations.

Reactions

Transition metal complexes

Propylene coordinates to s through its π-orbitals, forming η²-alkene complexes with metals such as , , and . In these complexes, the C=C binds side-on to the metal center, with the propylene acting as a two-electron donor . A representative example is the trichloro(propene)platinate(II) anion, [PtCl₃(η²-C₃H₆)]⁻, the propene analog of , which can be isolated as the salt K[PtCl₃(η²-C₃H₆)]·H₂O. Similar unstable π-complexes form with and , often observed at low temperatures via matrix isolation techniques. The bonding in propylene transition metal complexes is rationalized by the Dewar-Chatt-Duncanson model, involving synergistic σ-donation from the filled π-orbital of propylene to an empty metal orbital and π-backbonding from metal d-orbitals to the antibonding π* orbital of the . This interaction increases on the , weakening the C=C bond and imparting partial single-bond character. The stability of these complexes depends on the metal's electron richness and the alkene's substitution; late s like Pt and Pd favor stronger back-donation, enhancing complex formation. The weakened C=C bond manifests in spectroscopic signatures, including a red shift in the IR stretching frequency (Δν ≈ 100–200 cm⁻¹ lower than free propylene's ν_{C=C} at 1645 cm⁻¹), reflecting reduced . For instance, in analogous platinum-ethylene complexes like , the ν_{C=C} appears at 1518 cm⁻¹, a shift of 105 cm⁻¹, and similar perturbations occur in propylene complexes. In ¹H NMR spectra, the vinyl protons of coordinated propylene shift upfield (typically by 1–3 relative to free propylene at δ 4.9–5.9 ), due to the deshielding of the metal-alkene interaction and increased sp³ hybridization at the carbons. These complexes are foundational in , serving as precursors for propylene oligomerization via or centers, where the η²-bound facilitates subsequent transformations like migratory insertion. In copolymerization contexts, the coordinated propylene undergoes migratory insertion into metal-alkyl bonds, enabling incorporation into chains, though detailed mechanisms are beyond the scope of coordination chemistry here.

Polymerization

Propylene undergoes coordination-insertion polymerization to form (), a major , primarily using heterogeneous Ziegler-Natta catalysts composed of (TiCl₄) and trialkylaluminum (AlR₃, where R is typically ethyl) supported on (MgCl₂). The follows the Cossee model, involving sequential coordination of the propylene to a vacant site on the titanium center, followed by migratory insertion into the Ti-C bond of the growing polymer chain. Isotactic selectivity, which yields highly crystalline PP with over 90% isotactic triads, arises from steric interactions between the of propylene and the chiral around the Ti active site, directing the monomer to insert in a regioregular and stereoregular manner. Metallocene catalysts, typically zirconium-based single-site systems such as rac-[ethylenebis(indenyl)]zirconium dichloride (rac-Et(Ind)₂ZrCl₂) activated by methylaluminoxane (MAO), enable precise control over microstructure, producing atactic, syndiotactic, or isotactic depending on the symmetry. These homogeneous or supported catalysts facilitate uniform active sites, leading to narrower molecular weight distributions and tailored properties compared to traditional Ziegler-Natta systems. The overall reaction is represented as: n \ \ce{C3H6} \rightarrow \ce{[-CH2-CH(CH3)-]_n} Polymerization conditions typically involve temperatures of 50–100 °C and pressures of 1–10 atm to balance activity and polymer morphology, with stereochemistry control ensuring high crystallinity in isotactic PP (>90%). Kinetically, the propagation rate constant (k_p) for propylene insertion is approximately 10³ L/mol·s at 40–70 °C, dominated by the coordination-insertion steps, while chain transfer via β-hydride elimination or hydrogenolysis limits molecular weights to 10⁵–10⁶ g/mol. Industrial variants include slurry processes using liquid propylene as solvent and gas-phase methods like the UNIPOL process, which employs a fluidized-bed reactor for efficient heat removal and high productivity. Post-2020 advancements feature late-transition metal catalysts, such as Ni- and Pd-based systems with diimine ligands, enabling the synthesis of branched PP with enhanced processability through controlled chain walking mechanisms.

Oligomerization

Oligomerization of involves the controlled catalytic coupling of units to form short-chain oligomers, typically with chain lengths of fewer than 10 units, producing valuable to C12 hydrocarbons used in fuels and chemicals. This process differs from by terminating chain growth early to yield discrete, low-molecular-weight products rather than high polymers. Commercial methods primarily employ - or titanium-based catalysts for selective dimerization and trimerization, with the Dimersol process developed by the Institut Français du Pétrole (IFP) serving as a key example for producing olefins from propylene. In the Dimersol process, a homogeneous -phosphine catalyst activated by aluminum alkyls facilitates the liquid-phase dimerization under mild conditions, converting propylene into isohexene fractions suitable for blending. The mechanism of propylene oligomerization follows the Cossee-Arlman pathway, involving successive migratory insertions of propylene into a metal-alkyl bond, followed by β-hydride elimination to release the oligomer and regenerate the active site. This insertion mechanism allows for regioselectivity, often favoring 2,1-insertion in nickel systems, which directs the formation of branched products and influences selectivity toward specific dimers or trimers. For instance, nickel catalysts can achieve high selectivity to trimers like nonenes, while titanium systems may emphasize linear or branched C6-C9 species depending on ligand design. Catalyst coordination, such as bidentate phosphine ligands in Ni complexes, stabilizes intermediates and enhances selectivity, as explored in related transition metal catalysis. Key products from propylene oligomerization include dimers such as 2-methyl-1-pentene and 4-methyl-1-pentene, which serve as branched olefins for additives due to their high contributions after . Trimers, represented by the $3 \ce{C3H6} \rightarrow \ce{C9H18}, yield nonene isomers valuable as alpha-olefins. Yields are typically high, with selectivities exceeding 80-95% for desired dimers in optimized systems, minimizing higher oligomers. These processes operate at moderate temperatures of 50-150 °C and mild pressures (often 1-30 bar) to maintain liquid-phase conditions and control exothermicity. Applications of propylene oligomers extend to synthetic lubricants, where C6-C12 fractions provide modifiers and base stocks with superior thermal stability, and to alpha-olefins for detergents and . These uses are overshadowed by dominant applications in polymers and bulk chemicals.

Safety, environment, and handling

Health and safety

Propylene exhibits low , with an LC50 greater than 65,000 for a 4-hour exposure in rats, indicating it is not highly poisonous under normal conditions. As a simple asphyxiant, it can displace oxygen in confined spaces, leading to symptoms such as , , and at concentrations above 10% in air, potentially causing if oxygen levels fall below 19.5%. At high concentrations, propylene vapors may irritate the eyes, skin, and , while direct contact with the liquefied form can cause or freeze burns due to rapid cooling. Regarding chronic effects, propylene is not classified as a carcinogen by the International Agency for Research on Cancer (IARC Group 3, unclassifiable as to carcinogenicity to humans). Prolonged exposure to elevated levels may result in central nervous system depression, including drowsiness and fatigue, though significant long-term health impacts are uncommon at occupational exposure limits. The Occupational Safety and Health Administration (OSHA) permissible exposure limit (PEL) is 1,000 ppm as an 8-hour time-weighted average (TWA), with recommendations from the American Conference of Governmental Industrial Hygienists (ACGIH) setting a threshold limit value (TLV) at 500 ppm TWA to minimize risks. The primary route of exposure is of the gas, though and with the liquid form poses risks of cryogenic injury. Propylene is extremely flammable, with a lower limit (LEL) of 2.0% and an upper limit (UEL) of 11.1% in air, an of 457°C, and a of -104°C, creating significant hazards in confined spaces or during leaks. To mitigate these risks, propylene should be handled in well-ventilated areas to prevent oxygen displacement and vapor accumulation, using explosion-proof equipment and non-sparking tools. (PPE) including safety goggles, insulated gloves, and protective clothing is essential, particularly when handling the ; respiratory protection such as (SCBA) is required in oxygen-deficient or high-concentration environments. The (NFPA) rates propylene as Health 1 (slight hazard), Flammability 4 (extreme danger), and Instability 0 (minimal reactivity under normal conditions). Incidents involving propylene are rare but can be severe, as leaks may form mixtures with air, leading to fires or blasts if ignited; historical cases highlight the importance of and emergency response protocols.

Environmental impact

Propylene, as a (VOC), contributes to the formation of photochemical and through reactions with atmospheric oxidants. In refineries, it accounts for a notable portion of VOC emissions, typically comprising 5-10% of total refinery VOC releases, with global industrial emissions estimated at approximately 1-2 million tons per year based on production scales and loss factors. These emissions primarily arise from sources during production, storage, and transport, exacerbating urban air quality issues. In the environment, propylene exhibits high volatility, characterized by a constant of 0.196 atm·m³/mol, facilitating rapid partitioning from water to air. It degrades quickly in the atmosphere via reaction with hydroxyl (OH) radicals, with an estimated of about 5 hours to 1-2 days, minimizing long-term persistence. Bioaccumulation potential is low due to its rapid atmospheric removal and limited in . Direct toxicity to wildlife is generally low; acute aquatic toxicity tests show LC50 values greater than 50 mg/L for fish, indicating minimal harm at typical environmental concentrations. However, indirect effects occur through smog formation, which can impair ecosystems by altering atmospheric chemistry and deposition patterns. Regulatory frameworks address propylene's environmental footprint. In the United States, the Environmental Protection Agency regulates it as a VOC under the Clean Air Act, imposing controls on emissions from stationary sources to curb ozone precursors. Under the European Union's REACH regulation, propylene is registered with requirements for emission limits and risk assessments to protect air and water quality. Post-2020 developments include the International Maritime Organization's approval of a net-zero framework for shipping in April 2025, including fuel standards and GHG pricing, which aim to reduce indirect emissions from energy-intensive propylene supply chains like steam cracking; however, adoption was postponed in October 2025, with negotiations to resume in 2026. Emerging carbon pricing mechanisms further influence production economics, incentivizing lower-emission pathways. Mitigation strategies focus on reducing emissions and carbon intensity. Technologies for flare gas recovery and carbon capture can decrease associated CO₂ releases, while process optimizations limit VOC venting. Life-cycle assessments indicate that conventional propylene production emits approximately 1.5-2.5 tons of CO₂ equivalent per ton, primarily from energy inputs in cracking or dehydrogenation processes. Recent attention has grown on managing methane copollutants in propane dehydrogenation (PDH) routes, where upstream leaks and side reactions contribute to potent GHG emissions, prompting calls for carbon taxes and improved catalysts to curb these impacts.

Storage and handling

Propylene is typically stored as a liquefied compressed gas under its own or as a cryogenic liquid for larger volumes. In the compressed form, it is maintained at temperatures between -40°C and 50°C, with s ranging from 8 to 15 at ambient conditions to keep it in the liquid state. For industrial-scale storage, refrigerated systems at approximately -50°C are used to handle large volumes, reducing the required and enhancing . Suitable materials for storage and handling include and , which provide adequate compatibility without significant degradation. Copper and should be avoided due to the risk of embrittlement and potential reactions with the . Containers must be DOT-approved cylinders or tanks equipped with safety relief valves to prevent over-pressurization. For large-scale refrigerated storage, insulated tanks are employed to maintain the low temperatures. During handling, all must be properly grounded and bonded to prevent buildup, which could ignite the flammable gas. Leak detection systems utilizing (IR) sensors are recommended to monitor for releases, and areas should be ventilated to keep propylene concentrations below 25% of the lower limit (LEL) to mitigate risks. Non-sparking tools and are essential to avoid ignition sources. For transportation, propylene is classified under UN 1077 as a refrigerated and is shipped by rail or vessel in insulated tanks designed for cryogenic service. In the event of a spill, isolation distances of 100 meters for small spills and up to 800 meters for large spills are required to protect personnel and surrounding areas. In emergencies, systems should be purged with an such as to safely depressurize and eliminate residual propylene before maintenance. For fires involving storage or transport, water spray should be applied from a safe distance to cool exposed containers and control the blaze, but direct streams on leaks must be avoided to prevent spreading the flammable vapor cloud.

References

  1. [1]
    Propylene
    **Summary of Propylene (CID 8252 - PubChem):**
  2. [2]
    PROPYLENE - CAMEO Chemicals - NOAA
    Propylene is a colorless gas with a faint petroleum like odor. It is shipped as a liquefied gas under its own vapor pressure.
  3. [3]
    Propylene - Some Industrial Chemicals - NCBI Bookshelf
    Propylene is produced primarily as a by-product of petroleum refining and of ethylene production by steam cracking of hydrocarbon feedstocks.Missing: formula | Show results with:formula
  4. [4]
    [PDF] Propylene - Hazardous Substance Fact Sheet
    Propylene is a colorless gas with a slight odor, or a liquid under pressure. It is used in the production of many organic chemicals including resins, plastics, ...
  5. [5]
    Propene - the NIST WebBook
    Formula: C3H · Molecular weight: 42.0797 · IUPAC Standard InChI: InChI=1S/C3H6/c1-3-2/h3H,1H2,2H3 Copy · IUPAC Standard InChIKey: QQONPFPTGQHPMA-UHFFFAOYSA-N CopyMissing: physical | Show results with:physical
  6. [6]
    [PDF] Safety Data Sheet - NOVA Chemicals
    Mar 13, 2020 · Melting point/freezing point: -185 °C (-301 °F). Initial boiling point and boiling range: -48 °C (-54 °F). Flash Point: -108 °C (-162 °F).
  7. [7]
  8. [8]
    [PDF] PROPYLENE - CAMEO Chemicals
    Propene. Keep people away. Shut off ignition sources ... 9.5 Critical Temperature: 197.2°F = 91.8°C = 365°K. 9.6 Critical Pressure: 670 psia = 45.6 atm = 4.62.Missing: reliable | Show results with:reliable
  9. [9]
    Chemical Properties of Propene (CAS 115-07-1) - Cheméo
    Physical Properties ; ΔcH° · -2057.80 ± 1.10, kJ/mol ; ΔcH° · -2057.70 ± 0.60, kJ/mol ; μ, 0.40, debye ; LFL, 2.00, % in Air ...Physical Properties · Temperature Dependent... · Datasets · Correlations
  10. [10]
    Propene - an overview | ScienceDirect Topics
    Solubility in water, 446 mg/L at 20 − 25 °C: soluble in ethanol, diethyl ether ; Partition coefficients: ; Log K · 1.77 ; Log K · 2.34 − 2.7 ; Vapor pressure at 25 °C ...
  11. [11]
    None
    ### Physical Properties of Propylene (Section 9: Physical and Chemical Properties)
  12. [12]
    Propene
    **Summary of Thermodynamic Data for Propene:**
  13. [13]
  14. [14]
    Dipole Moment | Rosamonte's Physical Chemistry Website
    Dipole moment This page provides a list of dipole moment for about 800 molecules ... Propene, 0.366 ± 0.001. C3H6Br2, 1,2-Dibromopropane, [1.2]. C3H6Cl2, 1,2 ...
  15. [15]
    [PDF] values to know 1. Protonated carbonyl pKa
    Protonated carbonyl pKa = -7 Other important pKa's. 2. Protonated ... Alkene: vinyl 45-50; allylic 43. 16. Alkane pKa = above 50. Examples: H. H. H. H.Missing: propene | Show results with:propene
  16. [16]
    Introduction | SpringerLink
    Jul 22, 2018 · Propylene was then discovered by Captain John W. Reynolds 14 in 1850 [49], who decomposed amyl alcohol (pentanol, C5H12O) by passing its vapor ...
  17. [17]
    Researches on a New Class of Alcohols - jstor
    of propylene, C6 H, Br, discovered previously by CAHOURS, REYNOLDS and HOFMANN, in ... Propylene. ... Allylic alcohol, the history of which we have endeavoured to ...
  18. [18]
    Poylpropylene and High-Density Polyethylene - National Historic ...
    Banks of Phillips Petroleum Company discovered polypropylene, a high-melting crystalline aliphatic hydrocarbon. This discovery led to the development of a new ...Missing: 1850 Reynolds
  19. [19]
    [PDF] Giulio Natta - Nobel Lecture
    At the beginning of 1954 we succeded in polymerizing propylene, other α-olefins, and styrene; thus we obtained polymers having very different properties from ...
  20. [20]
    Polypropylene - The Plastics Historical Society
    Dec 6, 2016 · Isotactic polypropylene was discovered in 1954 by the Italian chemist Guilio Natta and his assistant Paulo Chini, working in conjunction ...Missing: 1850 Hofmann Reynolds<|separator|>
  21. [21]
    A Most Invented Invention | Invention & Technology Magazine
    Hoechst, Montecatini, and Hercules all built pilot plants to work out the intricacies of manufacturing polypropylene and began full-scale production in 1957.
  22. [22]
    [PDF] Screening Assessment for 1-Propene Chemical Abstracts Service ...
    Propene is a naturally occurring gas that is emitted from many plants, and is a component in natural gas, volcanoes, and from incomplete biomass combustion.
  23. [23]
    Genetic and Physiological Characteristics of a Novel Marine ...
    In the present study, we isolated propylene-assimilating bacteria from seawater and revealed their biological characteristics. ... role in propylene metabolism.
  24. [24]
    Secondary organic aerosol formation from propylene irradiations in ...
    The concentration of propylene in the atmosphere is in the 10−1∼101 ppb levels (Chang et al., 2005), which is similar to other LMW VOCs, such as ethylene ...
  25. [25]
    The Human Exposure Potential from Propylene Releases to the ...
    Propylene is produced via the steam cracking of petroleum feedstocks such as natural gas liquids and gas oils [3]. There are also numerous pyrogenic and ...Missing: terrestrial | Show results with:terrestrial
  26. [26]
    Natural catalytic activity in a marine shale for generating natural gas
    Apr 21, 2010 · A Cretaceous Mowry shale catalysed the dimerization of propylene (C3H6) to methyl cyclopentane (MCP, C6H12) and n-hexane (n-C6, C6H14) at 50°C ...
  27. [27]
    Propylene on Titan - NASA Scientific Visualization Studio
    Sep 30, 2013 · Now, thanks to NASA's Cassini spacecraft, scientists have detected propylene on Titan ... NASA's Webb Telescope has discovered a new molecule in ...Missing: interstellar Sagittarius B2 ALMA comets
  28. [28]
    Oxygen-bearing organic molecules in comet 67P's dusty coma
    Propylene oxide may or may not be marginally present at the expense of some acetone and propanal in comet 67P. Moreover, it is almost indistinguishable from ...
  29. [29]
    A Review on the Production of Light Olefins Using Steam Cracking ...
    The simplest free radical mechanism can be observed for ethane cracking, where ethane splits into two methyl radicals in the chain initiation step (Equation (1)) ...
  30. [30]
    Life Cycle Assessment and Multiobjective Optimization for Steam ...
    Apr 27, 2022 · This work aims to optimize the steam cracking process to increase profits and reduce emissions. The scope of LCA in this paper is the cracking ...
  31. [31]
    [PDF] 3 The Propylene - eere.energy.gov
    Here, propylene with a purity of greater than 99.5 percent is fed to a continuously stirred reactor along with hydrogen and catalyst. In the reactor, the ...Missing: composition | Show results with:composition
  32. [32]
    Ethylene and Propylene Market Size, Share & Trends, 2033
    Steam cracking remains the dominant production method, accounting for over 70% of global ethylene capacity and over 60% for propylene. The market is tightly ...
  33. [33]
    Review of Electric Cracking of Hydrocarbons - ACS Publications
    The specific energy consumption (SEC) for the production of ethylene from naphtha is in the range of 26–31 GJ/ton of ethylene. (3) Best cases are given in Table ...
  34. [34]
    Metathesis for maximum propylene - DigitalRefining
    If the alpha olefin is ethylene and the secondary olefin is 2-butene, two propylene molecules are formed. If the reaction is between an alpha olefin (1-butene ...
  35. [35]
    Propagation of Olefin Metathesis to Propene on WO3 Catalysts
    This structural conversion can greatly reduce surface strain on the inactive methyl groups and stabilize the six-membered C5 (oxa)metallacycle intermediate.
  36. [36]
    Tungsten Oxide on Silica Catalyst for Phillips' Triolefin Process
    A study of the properties of catalysts for propylene metathesis by means of the temperature programmed reaction method. Materials Chemistry and Physics 1989 ...
  37. [37]
    Olefins Conversion Technology (OCT) - Toyo Engineering
    The OCT process is an excellent process in terms of investment cost and utility consumption compared to other propylene boosting processes such as catalytic ...Missing: WO3/ Al2O3<|separator|>
  38. [38]
    OLEFINS CONVERSION TECHNOLOGY APPLICATIONS - OnePetro
    The per-pass conversion of butylene is greater than 60% with overall selectivity to propylene exceeding 90%. The product from the metathesis reactor ...Missing: yield | Show results with:yield
  39. [39]
    Turning Ethylene Into Propylene - C&EN - American Chemical Society
    Apr 23, 2007 · With metathesis technology, 1 mole of butylene and 1 mole of ethylene yields 2 moles of propylene. More than 90% of the world's propylene is made as a ...
  40. [40]
    [PDF] Environmental and Safety Considerations for Olefins Conversion Unit
    Mar 16, 2023 · Integration of CDIsis® technology with the OCT process allows an economical and cost- effective solution to enhance propylene production.
  41. [41]
    Fluid Catalytic Cracking - an overview | ScienceDirect Topics
    Fluid catalytic cracking (FCC) is defined as a process used in oil refineries to convert high molecular weight hydrocarbons into lower boiling components, ...
  42. [42]
    Fluid catalytic cracking: recent developments on the grand old lady ...
    We will discuss several trends in FCC catalysis research by focusing on ways to improve the zeolite structure stability, propylene selectivity and the overall ...
  43. [43]
    [PDF] Catalyst Residence Time Distributions in Riser Reactors for Catalytic ...
    Literature references2 indicate that most FCC risers are designed to operate with a bulk flow gas residence time of 2–3 seconds, hence these flow rates and the ...
  44. [44]
    Increasing FCC propylene - DigitalRefining
    ZSM-5-based catalyst additives increase LPG yields, especially propylene, in the FCCU by cracking gasoline-range hydrocarbons.
  45. [45]
    Maximizing propylene production via FCC technology
    Mar 22, 2015 · A large proportion of propylene is produced by steam cracking (SC) of light naphtha and during the fluid catalytic cracking (FCC) process.
  46. [46]
    [PDF] Propane Dehydrogenation (PDH) | Aranca
    There are two major steps in the PDH process: 1) propane feed dehydrogenation, and 2) product separation (removal of hydrogen). The process is highly ...
  47. [47]
    Beyond platinum for propane dehydrogenation - Nature
    Jun 23, 2025 · Propane dehydrogenation (PDH) is an industrial alternative to conventional oil-based cracking methods for propylene production.
  48. [48]
    [PDF] Propylene Analytics - Argus Media
    May 31, 2024 · But steam cracking will remain the primary form of propylene production at the end of the forecast period, accounting for 41pc of capacity in ...
  49. [49]
    Methanol to Olefins (MTO): From Fundamentals to Commercialization
    Significantly, SAPO-34 templated by DIPA shows good MTO catalytic performance with an ethene plus propene selectivity of 85.8%. Further extending this method to ...FUNDAMENTAL STUDY ON... · SYNTHESIS AND... · DMTO PROCESS...
  50. [50]
    Insights into the methanol to olefins (MTO) performance of SAPO-34 ...
    The MTO performance of SAPO-34 under the stripper conditions was studied, and the influences of the coke from heavy oil cracking were analyzed.
  51. [51]
    DMTO: A Sustainable Methanol-to-Olefins Technology - Engineering
    MTO catalysts are typically made in the form of microspheres with controllable physical properties via the spray-drying method, with SAPO-34 molecular sieves ...
  52. [52]
    [PDF] Bio-Based Chemicals - A 2020 Update - IEA Bioenergy
    Routes for bio-based production of propylene. Raw Material. Process. Ethanol ... DSM and Roquette take next step in Reverdia bio-succinic acid joint venture.
  53. [53]
    Carbon-negative production of acetone and isopropanol by gas ...
    Feb 21, 2022 · Here we describe the development of a carbon-negative fermentation route to producing the industrially important chemicals acetone and isopropanol from ...<|separator|>
  54. [54]
    Operating envelope of a short contact time fuel reformer for propane ...
    The present work combines thermodynamic modeling of propane catalytic partial oxidation (cPOx) and experimental performance of a Precision Combustion Inc. (PCI) ...
  55. [55]
    [PDF] Filling the Propylene Gap – Shaping the Future with On-Purpose ...
    Propane dehydrogenation is a simple process with one feed (propane) that is converted to one primary product (propylene) with the option to use the by-product ...Missing: Catofin | Show results with:Catofin<|separator|>
  56. [56]
    Report evaluates emerging 'on-purpose' propylene production ...
    The shift to ethane feedstocks has created a shortfall in the supply of propylene, which is essential for plastics production.
  57. [57]
    Propylene Price Index, Chart, Trend and Forecast 2025
    Propylene Industry Analysis ... The global propylene industry size reached 130.29 Million Tonnes in 2024. By 2033, IMARC Group expects the market to reach 210.96 ...Propylene Prices October... · Market Overview Q3 Ending... · Propylene Industry Analysis
  58. [58]
    Global Propylene Industry Report 2025: China has Emerged as the ...
    May 23, 2025 · The global propylene market, projected to grow at a 2.6% CAGR from 2025 to 2030, sees China as the leading consumer. In 2024, China, the USA ...
  59. [59]
    Propylene Market Size To Attain USD 162.03 Billion By 2034
    May 19, 2025 · The global propylene market size is estimated at USD 116.23 billion in 2025 and is predicted to surpass around USD 162.03 billion by 2034, expanding at a CAGR ...
  60. [60]
    Propylene Market Size, Share & Trends | Industry Outlook [2032]
    The global propylene market size was valued at USD 121.34 billion in 2023 and is projected to grow from USD 127.85 billion in 2024 to USD 194.62 billion by ...
  61. [61]
    Top Propylene Companies & How to Compare Them (2025)
    Oct 11, 2025 · Sinopec: Major Asian producer with significant regional influence and capacity. ... Large-scale, global supply: ExxonMobil, Shell, LyondellBasell.
  62. [62]
    Propylene: through the downcycle and beyond | Wood Mackenzie
    Aug 11, 2022 · The propylene industry is heading for a downturn. Capacity additions, driven primarily by massive investment in China, will outweigh demand growth through to ...
  63. [63]
    PGP Polymer Grade Propylene (PCW) Financial Futures - CME Group
    PGP Polymer Grade Propylene (PCW) Financial Futures - Quotes ; NOV 2025. PGGX5 · 0.24938 ; DEC 2025. PGGZ5 · 0.26000 ; JAN 2026. PGGF6 · 0.28250 ; FEB 2026. PGGG6.
  64. [64]
    US Gulf Coast Propylene Supply Dynamics Shifting to More ... - OPIS
    Mar 20, 2025 · US Gulf Coast propylene market fundamentals in early 2025 have been indicative of more seasonal volatility.
  65. [65]
    Americas Plastics Mid-Year Outlook 2025 - ICIS
    US polypropylene (PP) is expected to continue to be less volatile in the rest of 2025, following 2024 where prices changed every month. Tariffs should have ...
  66. [66]
    Viewpoint: US PGP prices set to rise in 2025 | Latest Market News
    Dec 26, 2024 · US spot polymer-grade propylene (PGP) prices are likely to increase into 2025, driven largely by several planned limitations on supply.Missing: volatility | Show results with:volatility
  67. [67]
    Propane Dehydrogenation (PDH) to Propylene - Data Insights Market
    Rating 4.8 (1,980) Oct 24, 2025 · This shift is attributed to the growing demand for propylene that outpaces its supply from steam crackers, especially with the rise of light ...
  68. [68]
    Propylene Production via Oxidative Dehydrogenation of Propane ...
    The CO2-assisted oxidative dehydrogenation of propane (ODP) was investigated over titania based composite metal oxides, 10% MxOy-TiO2 (M: Zr, Ce, Ca, Cr, Ga).
  69. [69]
    [PDF] CO2-ASSISTED OXIDATIVE DEHYDROGENATION OF PROPANE ...
    3 sept 2025 · TUD-1 Catalyst for Oxidative Dehydrogenation of Propane to Propylene with CO2 ... assisted by CO2 over Y-doped ceria supported Au catalysts ...Falta(n): bio- waste
  70. [70]
    New catalyst developed for sustainable propylene production from ...
    Sep 24, 2024 · Researchers at Osaka Metropolitan University have developed a new catalyst that efficiently converts a derivative of glycerol into bio-based propylene.
  71. [71]
    [PDF] Preparation and Characterization of New Supported Catalysts ...
    Sep 16, 2025 · the production of single site catalysts for a variety of reactions (for example alkane metathesis, alkane dehydrogenation, olefin metathesis ...
  72. [72]
    Propylene Synthesis: Recent Advances in the Use of Pt-Based ...
    Compared to the monometallic Pt catalyst, the Pt-Sn bimetallic catalysts showed improved turnover rates, propylene selectivity and stability in PDH reaction [21] ...
  73. [73]
    Advanced monometallic and bimetallic catalysts for energy efficient ...
    This review highlights emerging trends in catalyst design focused on reducing activation energy barriers and enhancing selectivity for propylene in propane ...
  74. [74]
    Reaction Mechanism and Selectivity Tuning of Propene Oxidation at ...
    Nov 7, 2022 · We report the mechanistic studies of propene electrooxidation on PdO/C and Pd/C catalysts, considering that the Pd-based catalyst is one of the most promising ...Missing: electroreduction | Show results with:electroreduction
  75. [75]
    Facet-dependent electrooxidation of propylene into ... - Nature
    Feb 17, 2022 · In this work, we develop an efficient electrocatalyst of Ag 3 PO 4 for the electrooxidation of propylene into propylene oxide.
  76. [76]
    V activated electro-epoxidation catalyst in membrane electrode ...
    Apr 1, 2025 · Reaction mechanism and selectivity tuning of propene oxidation at the electrochemical interface. J. Am. Chem. Soc. 144, 20895–20902 (2022) ...
  77. [77]
    [PDF] Chemical Recycling in a Circular Economy for Plastics
    Under current European regulations, pyrolysis-based chemical recycling (Py-CR)5 is cur- rently the only route capable of “at-scale”6 processing of flexible ...
  78. [78]
    The path to circularity in Europe: recognising the value of waste - ICIS
    The EU has set a target for 50% of plastic packaging to be recycled by 2025, rising to 55% by 2030. The new Packaging and Packaging Waste Regulation (PPWR), ...
  79. [79]
    Green synthesis of propylene oxide directly from propane - Nature
    Dec 13, 2022 · This process involves the reaction of propane with oxygen to produce propylene oxide as the target product. The rest of the considered reaction ...
  80. [80]
    Review and perspectives on TS-1 catalyzed propylene epoxidation
    Green and efficient epoxidation of propylene with hydrogen peroxide (HPPO process) catalyzed by hollow TS-1 zeolite: A. 1.0 kt/a pilot-scale study. Chem. Eng ...
  81. [81]
    Towards Sustainable Production of Bio-Based Ethylene Glycol
    Aug 6, 2025 · The review also discusses the current challenges in scaling up PEF production, such as the high cost of FDCA, the need for efficient ...
  82. [82]
    Emerging bio-based products have nearly half the GHG footprint of ...
    Feb 7, 2024 · Bio-based products emitted 45% less GHG than fossil counterparts, with 80 of 98 products having lower GHG footprint, but none reaching net-zero.
  83. [83]
    [PDF] Primary chemicals industry net-zero tracker
    The primary chemicals industry, relying on fossil fuels, accounts for 2.5% of global CO2e emissions. Absolute emissions rose 6% (2019-2023), but intensity is ...Missing: 2x 2020<|control11|><|separator|>
  84. [84]
    [PDF] Argus report sample
    Polypropylene remains the largest consumer of propylene, accounting for 70pc of the total consumption, followed by propylene oxide (7pc) and acrylonitrile. (6pc) ...
  85. [85]
    Ziegler-Natta Polypropylene - an overview | ScienceDirect Topics
    The development of metallocene catalysts has resulted in improved properties for packaging applications, including greater puncture resistance, higher impact ...
  86. [86]
    Ziegler-Natta and metallocene catalyzed isotactic polypropylene
    Materials catalyzed using the Ziegler-Natta catalysts typically have broad molecular weight distributions, high peak melting temperatures, a heterogeneous ...
  87. [87]
    Polypropylene Market Size, Share, Analysis and Forecast 2035
    The global Polypropylene (PP) market stood at approximately 70 million tonnes in 2024 and is expected to grow at a healthy CAGR of 4.44% during the forecast ...
  88. [88]
    What are the differences between PP homopolymer and copolymer?
    Jun 28, 2024 · PP homopolymer is often used in plastic products, fibers, films and other fields, while PP copolymer is often used in products in the ...
  89. [89]
    Understanding the Distinctive Characteristics of Copolymer and ...
    Apr 11, 2024 · Homopolymer propylene consists of only propylene monomers, while copolymer propylene combines propylene with other monomers, leading to ...
  90. [90]
    Polypropylene, slurry process and gas phase process | GLAD
    The two main technologies for producing polypropylene are slurry and gas phase. In bulk slurry technology, the polymerization reaction takes place in liquid ...
  91. [91]
    Ethylene-Propylene Copolymer - Polyallomer - AZoM
    May 9, 2001 · Better resistance to stress cracking than Polyethylene or Polypropylene. Relative advantages, disadvantages and applications are listed ...
  92. [92]
    All About EPDM Rubber – Properties, Applications and Uses
    Aug 28, 2025 · Finally, we will highlight the wide array of applications where EPDM is commonly utilized, from automotive components and industrial hoses to ...
  93. [93]
    Polypropylene Market Size, Forecast 2030 | Research Report
    Sep 19, 2025 · The Polypropylene Market is expected to reach 97.30 million tons in 2025 and grow at a CAGR of 5.78% to reach 128.86 million tons by 2030.
  94. [94]
    Research and application of polypropylene: a review - PMC - NIH
    Jan 2, 2024 · Its use is quickly expanding in various sectors, including packaging, textiles, automotive, and consumer products.
  95. [95]
    Polypropylene Market Size, Share | Analysis Report [2032]
    The global polypropylene market size was $83.85 billion in 2023 & is projected to grow from $88.24 billion in 2024 to $134.12 billion in 2032, ...
  96. [96]
    Propylene - an overview | ScienceDirect Topics
    Propylene is a reactive olefin that can be obtained by steam cracking of higher molecular weight hydrocarbon derivatives, where it is coproduced with ethylene.Missing: physical | Show results with:physical
  97. [97]
    Propylene Oxide Routes Take Off - C&EN
    Oct 9, 2006 · The oldest route, the chlorohydrin process, uses chlorine and propylene to produce propylene chlorohydrin as an intermediate.
  98. [98]
    HPPO Technology - Evonik Active Oxygens
    In the HPPO process, hydrogen peroxide (H2O2) is used as the oxidizing agent to oxidize propylene to propylene oxide (PO) with only water as a co-product. The ...
  99. [99]
    Propylene Oxide(PO) hydration method of Propylene Glycol
    Aug 26, 2024 · The PO hydration method produces highly pure(≥99.5%) propylene glycol. It involves the addition of water to propylene oxide, followed by purification steps.<|control11|><|separator|>
  100. [100]
    Sohio Acrylonitrile Process - American Chemical Society
    Today the Sohio Acrylonitrile Process is utililzed in over 90% of the world's acrylonitrile production, representing plants in sixteen countries worldwide.Brief History of Acrylonitrile · Discovery of the Sohio Process · Sohio Gambles with...
  101. [101]
    Advances in the catalytic production of acrylonitrile - ScienceDirect
    Jan 18, 2024 · Currently, ACN is mainly produced using the commercial propylene-based process (i.e., the SOHIO process). Driven by the climbing cost of ...
  102. [102]
    Propene (Propylene) - The Essential Chemical Industry
    Propene (often known as propylene), like ethene, is a very important building block for a large number of chemicals, including the addition polymer, ...
  103. [103]
    Cumene Benzene Alkylation | ExxonMobil Product Solutions
    Cumene alkylation uses ExxonMobil catalysts to convert benzene and propylene to cumene in a liquid phase, with a small amount of cumene further alkylated.
  104. [104]
    Design of a Process for Production of Isopropyl Alcohol by Hydration ...
    Design of a Process for Production of Isopropyl Alcohol by Hydration of Propylene in a Catalytic Distillation Column. A novel process flow sheet has been ...
  105. [105]
    [PDF] Allyl Chloride Production - Intratec.us
    The vast majority of Allyl Chloride produced on commercial scale is made from chlorine and propylene, via high-temperature chlorination processes. Examples of ...
  106. [106]
    Selective Oxidation of Propylene to Acrolein over Silver Molybdate ...
    May 9, 2022 · The selective oxidation of propylene is industrially carried out (1−3) to produce acrolein and propylene oxide (PPO) as important precursors for ...
  107. [107]
    I. Infrared spectra of propene complexes of some atomic metals
    ... propene analog of Zeise's salt K(C3H6PtCl3) · H2O. Only an unstable, black, charge-transfer complex is obtained from magnesium, and its infrared spectrum is ...<|control11|><|separator|>
  108. [108]
    Infrared Spectra and Normal Coordinate Analysis of Metal-Olefin ...
    Infrared Spectra and Normal Coordinate Analysis of Metal-Olefin Complexes. I. Zeise's Salt Potassium Trichloro(ethylene)platinate(II) Monohydrate1. Click to ...Missing: analog | Show results with:analog
  109. [109]
    Group 10 Metal Complexes Supported by Pincer Ligands with an ...
    The 1H NMR spectrum of 12 shows the olefinic protons at 5.91 ppm, shifted upfield compared to the corresponding protons of the free ligand. All four methine ...
  110. [110]
    New α-diimine nickel complexes—Synthesis and catalysis of alkene ...
    Three new α-diimine complexes of Ni are synthesized. The complexes are used as catalyst precursors for oligomerization of ethylene and propylene.
  111. [111]
    Mechanism of Propylene Polymerization with MgCl 2 -Supported ...
    Mechanism of Propylene Polymerization with MgCl2-Supported Ziegler–Natta Catalysts Based on Counting of Active Centers: The Role of External Electron Donor.
  112. [112]
    [PDF] Zielger-Natta Polymerization - Denmark Group
    • Polymerization of ethylene and propylene n. Toluene, 23 °C, 0.5 h. 1.4 mM ... Mechanism of Ziegler-Natta Polymerization. • Cossee mechanism (1960). Key ...
  113. [113]
    The Influence of Ziegler-Natta and Metallocene Catalysts on ...
    In fact, isotactic Ziegler-Natta PPs can be considered as a mixture of very different types of chains: short atactic chains are present even in most isotactic ...
  114. [114]
    Metallocene Catalysts Initiate New Era In Polymer Synthesis
    Germany's Hoechst also has demonstrated metallocene catalysts for producing isotactic or syndiotactic polypropylene. Other firms working in the metallocene ...
  115. [115]
    Effect of temperature on the isospecific propylene polymerization ...
    1/MAO has a very high activity for propylene polymerization at 50–70 °C and 1 atm propylene pressure. A beneficial feature of catalyst 1 is its higher ...
  116. [116]
    A Kinetic Study of Ziegler-Natta Propylene Polymerization
    Aug 7, 2025 · The kinetics of the propylene polymerization with TiCl3-AlEt3 has been investigated in the temperature range (30-70°C.), the pressure range ...
  117. [117]
    [PDF] Kinetic modeling of slurry propylene polymerization using rac-ET(Ind ...
    Jan 24, 2006 · A propagation rate constant at 40°C of 365. L/(mol s) for the first site and 362 L/(mol s) for the second site was reported. Nele et al.
  118. [118]
    UNIPOL ® PP Process Technology - W.R. Grace
    Gas phase polymerization is a process used to produce polypropylene (PP) by the polymerization of propylene gas in a fluidized bed reactor using a powdered ...
  119. [119]
    Recent Advances in Controlled Production of Long‐Chain Branched ...
    Mar 15, 2024 · 3 Late Transition Metal (Ni/Pd) Based Catalysts. The development of diimine nickel and palladium catalysts in the mid-90s by Brookhart and ...
  120. [120]
    The Dimerization and Oligomerization of Alkenes Catalyzed with ...
    Another way to synthesize oligomers involves transition metal catalysis (Zr, Ti, Hf, Ni, Co, Fe, V, and Ta) in which the Cossee–Arlman mechanism is implemented ...
  121. [121]
    Dimersol process for dimerizing propylene into a gasoline ...
    Apr 28, 1980 · Institut Francaise du Petrole's Dimersol process, which is a catalyzed, liquid-phase process for dimerizing C/sub 3//C/sub 4/ olefins, is being used at Total ...Missing: IFP oligomerization
  122. [122]
  123. [123]
    Validation of the Cossee–Arlman mechanism for propylene ...
    In this work, steady state propylene oligomerization rates and selectivities were measured in the absence of an activator on nickel functionalized UiO-66 metal ...
  124. [124]
    Catalyzed Propylene Dimerization: Cossee–Arlman Versus ...
    May 20, 2021 · 2-Methyl-1-pentene (2M1P), another dimerization product of propylene, is also suitable to produce branched copolymers and for the synthesis ...
  125. [125]
    Zirconium and Titanium Complexes Supported by Tridentate LX 2 ...
    When activated with MAO, the titanium bis(phenolate)furan and bis(phenolate)thiophene systems were found to promote propylene oligomerization. ACS Publications.
  126. [126]
    Highly selective propylene dimerization catalyzed by C1‐symmetric ...
    Apr 20, 2014 · ... dimerization of propylene in toluene at 100°C to afford 2-methyl-1-pentene with high selectivities up to 95.7–98.4% and moderate activities ...
  127. [127]
    US20020103406A1 - Production of olefin dimers and oligomers
    37, No. 4 (1995), describe a nickel based catalyst system known as the Dimersol® process. This process is useful for dimerizing or oligomerizing a variety of ...
  128. [128]
    Toward sustainable propylene: A comparison of current and future ...
    This review aims to provide and analyze many possible routes for the production of propylene using sustainable resources.Missing: ppb | Show results with:ppb
  129. [129]
    [PDF] Propylene
    Oct 23, 2018 · Eye Contact: Contact with gas/liquid escaping the container can cause frostbite, freeze burns, and permanent eye damage. Page 3. Propylene SDS ...
  130. [130]
    Propylene (IARC Summary & Evaluation, Volume 60, 1994)
    Aug 26, 1997 · International Agency for Research on Cancer (IARC) ... Propylene is not classifiable as to its carcinogenicity to humans (Group 3).
  131. [131]
    Summary Review of the Health Effects Associated with Propylene
    Propylene can be released to the environment as emissions from incinerators burning hazardous waste.Missing: Senna sophera
  132. [132]
    Environmental impacts and emission profiles of volatile organic ...
    May 3, 2025 · This study evaluates the emissions of volatile organic compounds (VOCs) from petroleum refinery operations, focusing on quantifying the VOCs emission inventory.
  133. [133]
    Volatile Organic Compounds (VOCs) Emitted from Petroleum and ...
    Aug 7, 2025 · Volatile Organic Compounds (VOCs) Emitted from Petroleum and their Influence on Photochemical Smog Formation in the Atmosphere. January 2012 ...<|separator|>
  134. [134]
    [DOC] Propylene SIAR - OECD Existing Chemicals Database
    In urban and polluted air, propylene concentrations have been reported to range between 0.35 to 32 ppb. Cigarette smoke releases propylene: each cigarette ...
  135. [135]
    [PDF] Safety Data Sheet - BASF
    Toxicity. Aquatic toxicity. Assessment of aquatic toxicity: Acutely harmful for aquatic organisms. Toxicity to fish. LC50 (96 h) 51.32 mg/l, Fish (calculated).
  136. [136]
    Propene - Substance Information - ECHA - European Union
    This substance is registered under the REACH Regulation and is manufactured in and / or imported to the European Economic Area, at ≥ 10 000 000 tonnes per ...
  137. [137]
    IMO approves net-zero regulations for global shipping
    Apr 11, 2025 · Draft regulations will set mandatory marine fuel standard and GHG emissions pricing for shipping to address climate change.
  138. [138]
    Volatile chemical product emissions enhance ozone and ... - PNAS
    Aug 2, 2021 · We show that VCPs are as important to ozone production as fossil fuel VOCs and that the chemistry of VCPs can have significant impacts on model simulations.Results And Discussion · Vcp Emissions Observed... · Modeling Ozone During An...
  139. [139]
    Life cycle assessment and carbon substance flow analysis of ...
    Aug 15, 2023 · Most flare gas recovery technologies lead to a CO2 emission reduction compared to flaring, except gas-to-power. However, no full life cycle ...
  140. [140]
    Life cycle assessment of primary energy demand and greenhouse ...
    Oct 1, 2017 · Life cycle GHG emissions via four pathways are 1.60, 2.06, 12.16, 9.23 tCO2 eq/t propylene, respectively.Missing: CO2 | Show results with:CO2
  141. [141]
    Carbon tax to phase out coal as propylene feedstock and control ...
    Aug 10, 2024 · Our study provides spatiotemporal insights into the carbon tax for China's propylene sector to decarbonize and control upstream methane emissions.
  142. [142]
    [PDF] Propylene - Linde
    Wear self-contained breathing apparatus when entering area unless atmosphere is proven to be safe. Ensure adequate ventilation. Stop leak if safe to do so. If ...
  143. [143]
    [PDF] Propylene - Gas Innovations
    Jul 30, 2024 · Additional Information About This Product: No data available. NFPA Ratings: 0= Minimal Hazard. 1= Slight Hazard. 2= Moderate Hazard. 3= Serious ...