Fact-checked by Grok 2 weeks ago

CFC

Chlorofluorocarbons (CFCs) are a family of synthetic organic compounds composed of carbon, chlorine, and fluorine atoms, characterized by their chemical stability, low toxicity, and non-flammability, which made them ideal substitutes for hazardous early-20th-century refrigerants like ammonia and sulfur dioxide. Invented in the 1920s and 1930s by American chemist Thomas Midgley Jr. at General Motors' Frigidaire division, CFCs such as dichlorodifluoromethane (CCl₂F₂, or Freon-12) revolutionized refrigeration, air conditioning, aerosol propellants, foam production, and solvent applications by enabling safe, efficient cooling systems without risk of explosion or poisoning. Their widespread adoption from the mid-20th century onward propelled industrial growth but raised environmental alarms in 1974 when chemists Mario Molina and F. Sherwood Rowland theorized, based on laboratory experiments and atmospheric modeling, that ultraviolet radiation in the stratosphere photolyzes CFCs to release chlorine atoms, which catalytically destroy ozone (O₃) molecules far more efficiently than natural processes. Empirical observations in the 1980s confirmed seasonal over , correlating with peak CFC emissions and stratospheric levels exceeding natural backgrounds by orders of magnitude, prompting the 1987 —an international treaty that phased out CFC production and consumption in developed nations by 1996 and extended controls globally. Post-phaseout monitoring by networks like NOAA has documented declining atmospheric CFC concentrations since the late , with total stratospheric peaking around 1997 and subsequently falling, aligning with predictions and contributing to recovery trends observed since 2000, including a shrinking hole. This causal evidence, derived from direct measurements rather than solely models, underscores the Protocol's efficacy, though minor illicit emissions detected since 2010 highlight ongoing compliance challenges. CFCs also possess potent properties, with potentials hundreds of times that of CO₂ over a 100-year horizon, amplifying their environmental footprint beyond effects. Despite replacement by hydrofluorocarbons (HFCs) and other alternatives, legacy CFC banks in old equipment continue slow atmospheric release, while the episode exemplifies how empirical data drove reversal of a once-celebrated technology.

Chemical Properties and Synthesis

Molecular Structure and Classes

Chlorofluorocarbons (CFCs) consist of carbon atoms bonded exclusively to and atoms, forming fully halogenated, saturated hydrocarbons without any . These compounds are typically derived from short-chain alkanes, with methane-based CFCs following the general formula CClmF4-m, where m = 1 to 3, and ethane-based CFCs following C2ClmF6-m, where m = 2 to 5. The absence of atoms results in strong, stable C–Cl and C–F bonds, contributing to their chemical inertness under normal conditions. CFCs are classified primarily by the number of carbon atoms in their backbone and identified via a standardized numbering system established for refrigerants and halocarbons. In this system, the code (e.g., CFC-xyz) is derived as follows: x (hundreds digit) = number of carbon atoms minus 1; y (tens digit) = number of hydrogen atoms plus 1 (always 0 for pure CFCs, yielding y=1); z (units digit) = number of atoms. Chlorine atoms are calculated to satisfy carbon's : 2*(C + 1) – H – F. Common one-carbon (methane-derived) examples include CFC-11 (CCl3F, ), CFC-12 (CCl2F2, ), and CFC-13 (CClF3, chlorotrifluoromethane). Two-carbon (ethane-derived) CFCs form another major class, exemplified by CFC-113 (C2Cl3F3, 1,1,2-trichloro-1,2,2-trifluoroethane), CFC-114 (C2Cl2F4), and CFC-115 (C2ClF5). These are distinguished from hydrochlorofluorocarbons (HCFCs), which incorporate hydrogen (e.g., HCFC-22, CHClF2), altering reactivity and environmental persistence. While longer-chain CFCs exist, one- and two-carbon variants dominated industrial applications due to suitable volatility and stability.

Physical and Chemical Properties

Chlorofluorocarbons (CFCs) are fully fluorinated and chlorinated aliphatic hydrocarbons that exhibit distinctive physical properties, including colorless appearance, low odor (often faint or nearly odorless), and states ranging from gases to liquids at standard ambient conditions depending on molecular size. They possess high vapor pressures, relatively high liquid densities compared to , low viscosities, and tension, facilitating their use in applications requiring volatility and flow. in is generally low (e.g., 0.1 g/100 mL for CFC-11 at 20°C), while they are highly soluble or miscible in organic solvents such as alcohols, ethers, and . Key physical properties of common CFCs are summarized below:
CompoundFormula (°C)Liquid Density (g/cm³ at ~20–25°C)Water Solubility (g/100 mL at 20°C)
CFC-11CCl₃F241.490.1
CFC-12CCl₂F₂-29.81.320.0028
CFC-113C₂Cl₃F₃47.71.56<0.01
Chemically, CFCs demonstrate exceptional stability and inertness under normal environmental conditions in the troposphere, attributed to the high bond dissociation energy of carbon-fluorine bonds (approximately 485 kJ/mol), which resists thermal and chemical attack. They are non-flammable, non-corrosive, and exhibit very low reactivity with most substances, including acids, bases, and oxidants at ambient temperatures, rendering them non-explosive and of low acute toxicity (e.g., LC50 >800,000 for CFC-12 in rats). This inertness stems from the absence of atoms and the shielding effect of , preventing free radical initiation. However, under high-energy conditions such as irradiation in the , CFCs undergo , releasing atoms. They show no significant or at Earth's surface.

Industrial Production Methods

The primary industrial production method for chlorofluorocarbons (CFCs) involves halogen exchange reactions, in which chlorine atoms in fully chlorinated derivatives are selectively replaced by atoms. This , an adaptation of the Swarts reaction originally developed by Frédéric Swarts using antimony trifluoride (SbF3), evolved industrially to employ anhydrous (HF) as the fluorinating agent due to its availability and cost-effectiveness, often in the presence of catalysts like (SbCl5) or later chromium-based oxides for vapor-phase operations. The reactions are highly exothermic and conducted in corrosion-resistant reactors, typically or nickel-lined, under controlled temperatures (around 50–150°C) and pressures to manage byproducts like (HCl) and ensure separation via . For the most prevalent CFCs, such as (CFC-11, CCl3F) and (CFC-12, CCl2F2), production begins with (CCl4) as the key precursor, sourced from methane chlorination. The core reaction is stepwise fluorination: CCl4 + HF → CCl3F + HCl for CFC-11, followed by further to CCl3F + HF → CCl2F2 + HCl for CFC-12, yielding a whose CFC-11:CFC-12 ratio (typically near 50:50) is adjusted by varying HF flow rates, activity, and residence times. Liquid-phase with SbCl5 predominated in early plants established by companies like starting in , converting the to fluorinated antimony species that facilitate fluorine-chlorine exchange while regenerating via HCl stripping. Other CFCs, such as CFC-113 (, CCl2F-CClF2), derive from or via analogous HF exchange, often in multi-stage reactors to achieve partial fluorination without over-substitution. Byproduct management includes HCl recovery for reuse in chlorination steps, and purification via to meet refrigerant-grade purity (>99.9%). These methods enabled global production to scale from thousands of tons in to peaks exceeding 500,000 tons annually by the 1980s for CFC-11 and CFC-12 combined, prior to phase-out under the . Despite phase-out, legacy processes inform destruction technologies and occasional illicit production detected in regions like .

Historical Development

Invention in the Early 20th Century

In the 1920s, mechanical refrigeration relied on dangerous working fluids such as ammonia, sulfur dioxide, and methyl chloride, which were toxic, corrosive, or flammable and contributed to fatal accidents from leaks in early domestic and commercial units. To mitigate these risks, Charles F. Kettering, General Motors' director of research, assembled a team led by mechanical engineer Thomas Midgley Jr. to identify safer alternatives through systematic testing of halogenated hydrocarbons. Midgley, along with chemists Albert L. Henne and Robert R. McNary at General Motors' Frigidaire division research laboratory, synthesized dichlorodifluoromethane (CCl₂F₂)—the first practical chlorofluorocarbon refrigerant—in 1928. This compound exhibited low toxicity, non-flammability, chemical stability, and efficient thermodynamic properties for vapor-compression cycles, outperforming prior options in safety and performance metrics. Midgley demonstrated its inertness publicly by inhaling the gas to blow out a candle, underscoring its non-reactive nature under ambient conditions. Frigidaire secured U.S. 1,886,339 for the formulation, filed in 1929 and granted in 1932, while E.I. du Pont de Nemours & Company scaled up production under the trademark , initiating commercial availability of CFC-based refrigerators by 1930. This innovation rapidly displaced hazardous in household appliances, enabling safer widespread adoption of electric without immediate evidence of long-term environmental drawbacks.

Widespread Adoption Post-World War II

Following World War II, chlorofluorocarbons (CFCs) experienced accelerated adoption in consumer and industrial applications, driven by their proven safety profile—non-toxic, non-flammable, and chemically inert—amid postwar economic expansion and rising demand for modern conveniences in the United States and Europe. In refrigeration, CFC-12 (dichlorodifluoromethane, or Freon-12) solidified its dominance, powering the surge in household refrigerator production and early air conditioning units as electrification and suburbanization advanced. By the mid-1950s, CFCs had overtaken prior refrigerants like sulfur dioxide and methyl chloride in domestic markets, with mixtures of CFCs and hydrochlorofluorocarbons (HCFCs) introduced to optimize efficiency in compressors and enhance cooling capacity. After 1950, CFCs controlled the majority of domestic, automotive, and commercial refrigeration and air-conditioning sectors, reflecting their reliability in sealed systems. The sector, which employed CFC propellants (primarily CFC-11 and CFC-12) during wartime for dispersal—yielding 50 million units by 1945—rapidly commercialized for civilian use . From onward, CFCs propelled bug sprays, paints, hair conditioners, deodorants, and pharmaceuticals, enabling fine mist dispersion without flammability risks. U.S. production expanded from 4.3 million units in to substantial volumes by the , supporting diverse personal care and household products as manufacturing scaled with consumer affluence. In the late and early , CFCs extended to blowing agents for foams used in , furniture cushioning, and disposable , leveraging their low boiling points to create expanded cellular structures efficiently. Concurrently, affordable proliferated in automobiles, homes, and offices, embedding CFCs in central systems and window units. This multifaceted integration propelled global CFC production from modest postwar levels to nearly 1 million metric tons annually by the early 1970s, underscoring their role as versatile enablers of industrial and consumer innovation prior to environmental scrutiny.

Applications and Uses

Role as Refrigerants and Propellants

Chlorofluorocarbons (CFCs), particularly (CFC-12 or R-12), were developed in the late 1920s as non-toxic, non-flammable alternatives to earlier refrigerants such as , , methyl chloride, and , which posed risks of toxicity, flammability, or explosion in domestic and commercial systems. received the first U.S. patent for CFC formulations on December 31, 1928, leading to commercial of R-12 for by 1931 through a joint venture between and DuPont's Kinetic Chemical Company. These compounds exhibited favorable thermodynamic properties, including low boiling points (e.g., -29.8°C for R-12) and high of vaporization, enabling efficient in vapor-compression cycles used in household refrigerators, automotive air conditioners, and industrial chillers. By the 1950s, CFC blends with hydrochlorofluorocarbons (HCFCs) further optimized performance by enhancing volumetric capacity and reducing compression ratios in systems. Post-World War II, CFCs like (CFC-11) and R-12 saw explosive adoption in due to their , which minimized and leaks compared to predecessors, and their electrical non-conductivity, suiting hermetic compressors. , by the , over 90% of new refrigerators and air conditioners incorporated CFCs, with global production reaching millions of tons annually by the 1970s to meet demand in expanding consumer markets and industries. Their inertness under operational conditions—resisting and oxidation—allowed for simpler, more reliable system s, though long-term atmospheric persistence was not initially a design concern. As propellants, CFCs gained prominence after for their low , non-reactivity with product formulations, and ability to produce fine, uniform sprays without flammability risks associated with hydrocarbons. Starting in the late 1950s, compounds such as CFC-11, CFC-12, and CFC-114 propelled insecticides, paints, , and pharmaceuticals, with U.S. aerosol production incorporating CFCs exceeding 1 billion units by 1974. These propellants enabled precise delivery in metered-dose inhalers and , displacing earlier compressed gases like that caused inconsistent dispersion or container corrosion. By the early 1970s, CFCs accounted for nearly all non-food propellants in developed markets, valued for maintaining product sterility and shelf life through their bacteriostatic properties.

Other Commercial and Industrial Applications

CFCs, particularly (trichlorofluoromethane), were extensively employed as blowing agents in the manufacture of flexible and rigid foams used for , packaging materials, and furniture cushioning. These compounds expanded mixtures into low-density foams by releasing gas during the curing process, providing superior properties and structural versatility compared to earlier alternatives like . By the 1970s, CFC-blown foams accounted for a significant portion of rigid foam production, with global demand driven by and appliance industries requiring efficient insulating materials. CFC-113 (1,1,2-trichloro-1,2,2-trifluoroethane) found primary use as a precision solvent in the sector for and circuit boards, where its non-flammable, low-residue properties prevented and ensured with sensitive components. This application emerged prominently in the post-World War II expansion of semiconductor manufacturing, as CFCs effectively removed fluxes and contaminants without damaging metals or plastics. Industrial of metal parts for machinery also relied on CFC solvents due to their stability and evaporative efficiency, though usage declined with the introduction of aqueous alternatives in the . Minor applications included CFC-12 in certain foam expansions and as a component in specialized chemical processes, though these were overshadowed by refrigerant demands. Overall, non-refrigerant and non-propellant uses peaked in the 1980s, representing approximately 20-30% of total CFC consumption before regulatory phase-outs under the curtailed production.

Ozone Depletion Hypothesis

Theoretical Foundations (1974 Onward)

In 1974, chemists Mario J. Molina and F. Sherwood Rowland published a seminal paper in Nature proposing that chlorofluorocarbons (CFCs), such as CFCl₃ and CF₂Cl₂, posed a threat to stratospheric ozone due to their photochemical stability and eventual breakdown in the upper atmosphere. These compounds, inert in the troposphere and thus capable of reaching the stratosphere intact, would undergo photodissociation by ultraviolet radiation, releasing chlorine atoms (Cl) that act as catalysts for ozone (O₃) destruction. The theory posited a cyclic chain reaction: Cl + O₃ → ClO + O₂, followed by ClO + O → Cl + O₂, yielding a net loss of one ozone molecule per cycle while regenerating the chlorine atom for repeated catalysis. Rowland and Molina's model relied on established principles of atmospheric , including rate constants for chlorine-oxygen reactions derived from laboratory data, and estimated that unchecked CFC emissions could reduce global stratospheric by 30–50% over decades, with greater impacts at higher latitudes. They calculated chlorine release rates based on projected CFC production trends, assuming annual increases of about 10% through the , and incorporated vertical transport models showing CFCs' long atmospheric lifetimes (50–100 years). This framework highlighted the inefficiency of natural production in countering catalytic loss, as each chlorine atom could destroy thousands of O₃ molecules before sequestration into less reactive forms like HCl. Subsequent theoretical refinements from 1975 onward built on this foundation, incorporating heterogeneous chemistry and polar stratospheric clouds, though core catalytic cycles remained central. In 1974–1976, extensions by researchers like McElroy and Wofsy emphasized bromine's higher catalytic efficiency but affirmed chlorine's dominance from CFCs. By 1976, a U.S. panel endorsed the hypothesis, validating one-dimensional models predicting 2–5% annual ozone loss under rising CFC levels, supported by emerging satellite and balloon measurements of stratospheric Cl and ClO. These developments underscored the theory's reliance on verifiable reaction kinetics rather than untested assumptions, though quantitative predictions varied with uncertainties in and sinks.

Atmospheric Chemistry Mechanisms

Chlorofluorocarbons (CFCs) ascend from the to the largely intact due to their and long atmospheric lifetimes, which exceed decades for compounds like CFC-11 () and CFC-12 (). In the , with wavelengths below 220 nm photodissociates CFCs, primarily cleaving the C-Cl bond to liberate atoms (Cl•): CFCl₃ + hν → CFCl₂• + Cl•. This process, first detailed by Rowland and Molina in , initiates ozone-depleting chemistry, as free radicals react efficiently with (O₃). The atoms catalyze destruction through a null cycle that regenerates the catalyst: Cl• + O₃ → ClO• + O₂ ClO• + O• → Cl• + O₂ Net: O₃ + O• → 2 O₂ This chain allows a single Cl• to deplete thousands of molecules—estimates range from 10⁴ to 10⁵ per atom—before temporary sequestration into reservoir species like (HCl) or chlorine nitrate (ClONO₂)./Kinetics/07:_Case_Studies-_Kinetics/7.03:_Depletion_of_the_Ozone_Layer) Atomic oxygen (O•) arises from photolysis: O₃ + hν → O₂ + O•, maintaining the cycle's efficiency in sunlit conditions./Kinetics/07:_Case_Studies-_Kinetics/7.03:_Depletion_of_the_Ozone_Layer) Additional cycles amplify depletion, particularly in cold regions. The ClO dimer mechanism, prominent over , involves: ClO• + ClO• + M → Cl₂O₂ + M (M = third body) Cl₂O₂ + hν → ClOO• + Cl• ClOO• + hν → Cl• + O₂ Followed by Cl• + O₃ → ... (regenerating ClO•), yielding net 2 O₃ → 3 O₂ per dimer photolysis. Heterogeneous reactions on polar stratospheric clouds convert reservoirs to photolabile forms like Cl₂: HCl + ClONO₂ → Cl₂ + HNO₃ (surface), with subsequent Cl₂ + hν → 2 Cl•. These processes, thermodynamically favored at low temperatures (below 195 K), explain seasonal ozone minima but require activation energies consistent with observed stratospheric loading from CFC breakdown. Bromine from halons (e.g., CF₃Br) contributes synergistically via cycles like Br• + O₃ → BrO• + O₂ and BrO• + ClO• → Br• + Cl• + O₂, enhancing efficiency through interhalogen recycling, though chlorine dominates overall due to higher stratospheric burdens. Laboratory kinetics and balloon-borne measurements confirm reaction rates, with k(Cl + O₃) ≈ 1.2 × 10⁻¹¹ cm³ molecule⁻¹ s⁻¹ at 220 K, supporting model predictions of catalytic amplification.

Scientific Controversies and Empirical Evidence

Mainstream Consensus on CFC Causality

The mainstream scientific consensus attributes the majority of observed stratospheric ozone depletion, including the Antarctic ozone hole, to anthropogenic emissions of chlorofluorocarbons (CFCs) and related halocarbons. This position, formalized in international assessments, holds that CFCs persist in the troposphere before diffusing into the stratosphere, where ultraviolet radiation cleaves carbon-chlorine bonds to release highly reactive chlorine atoms; these atoms then catalyze ozone destruction via cycles such as Cl + O₃ → ClO + O₂ followed by ClO + O → Cl + O₂, resulting in net ozone loss without consuming the chlorine catalyst. The theory originated in the 1974 paper by Mario Molina and F. Sherwood Rowland, who quantified that unchecked CFC emissions could deplete up to 7% of stratospheric ozone per decade by the end of the 20th century, a prediction later corroborated by atmospheric models incorporating measured CFC lifetimes of 50–100 years. Supporting observational evidence includes balloon-borne and aircraft measurements from the 1980s onward, which detected (ClO) concentrations in the exceeding 1 ppbv during springtime minima—levels attributable to CFC photolysis and activation, as natural sources like methyl chloride contribute far less stratospheric chlorine (about 0.6 ppbv versus 3.5 ppbv from CFCs at peak). NASA's 1987 Airborne Antarctic Expedition measured HCl and ClO profiles aligning with CFC-derived chlorine budgets, ruling out alternative explanations like oxides from natural variability, while ground-based confirmed the isotopic signature of consistent with industrial CFCs rather than volcanic or oceanic origins. The crystallized in reports such as the 1985 assessments, where over 90% of reviewed atmospheric chemists endorsed CFC causality based on these data. Post-Montreal Protocol (1987) trends provide retrospective causal validation: stratospheric effective equivalent chlorine peaked at approximately 3.7 ppbv around 2000, declining by 0.8% annually thereafter in parallel with a 20% reduction in ozone hole area from 1980s maxima of 25 million km², as quantified in satellite records from NASA's Total Ozone Mapping Spectrometer and instruments. The 2022 Quadrennial Ozone Assessment by the and UNEP attributes over 80% of the observed recovery to CFC controls, with model simulations showing that without phase-out, depletion would have worsened by an additional 10–20% globally by 2020. This framework underpins the awarded to Molina and Rowland in 1995, recognizing the CFC- mechanism as a cornerstone of .

Alternative Explanations and Skeptical Critiques

Skeptics of the CFC- depletion hypothesis, including atmospheric physicist S. Fred Singer, have argued that the observed ozone hole primarily results from natural meteorological and dynamical processes rather than anthropogenic halocarbons. Singer contended that polar stratospheric clouds (PSCs), which form naturally during the extreme cold of winter, facilitate heterogeneous chemical reactions leading to temporary ozone loss, independent of elevated stratospheric from CFCs; he emphasized that such depletions predate significant CFC emissions and align with historical ozone variability observed since the . Alternative explanations invoke solar variability as a key driver, with the 11-year influencing ultraviolet radiation and stratospheric temperatures, thereby modulating production and destruction rates. A technical report notes that natural fluctuations tied to solar cycles could account for much of the detected decline, as these cycles alter the balance of photochemical reactions in the without requiring CFC . Volcanic eruptions have also been proposed as contributors, releasing (HCl) gas that efficiently destroys upon stratospheric injection; the U.S. Geological Survey highlights that HCl from eruptions like in 1991 caused measurable global reductions through direct reaction with molecules. Critics argue that while mainstream models downplay natural chlorine's role due to tropospheric scavenging, incomplete understanding of transport mechanisms underestimates volcanoes' impact relative to CFCs. Empirical critiques focus on discrepancies between predictions and observations. The 1974 Molina-Rowland theory forecasted widespread, gradual global loss proportional to CFC accumulation, yet the hole emerged abruptly in spring 1979 satellite data despite steady CFC buildup since the , suggesting a non-linear, possibly natural trigger tied to vortex dynamics rather than loading. Alarmist projections of surging ultraviolet-B leading to epidemics of and have not materialized; U.S. skin cancer rates showed no statistically significant rise attributable to ozone thinning from 1970 to 2000, per analyses questioning causal linkage. Skeptics like those at the National Center for Policy Analysis assert that total column declines are minor (around 3-5% globally pre-1990s) and indistinguishable from natural variability, with CFC reductions under the correlating temporally with recovery but lacking rigorous controls for confounding factors like solar minima or reduced volcanic activity. These views, often advanced by independent researchers and policy analysts wary of regulatory overreach, highlight potential overreliance on laboratory kinetics extrapolated to complex atmospheric systems, where heterogeneous chemistry and transport dominate but remain incompletely modeled. While peer-reviewed consensus favors CFC dominance, skeptics demand direct field evidence of radicals (ClO) from CFCs—versus natural sources—driving the hole's formation, noting that activation thresholds align more closely with temperature anomalies than trends.

Observed vs. Predicted Outcomes

The Rowland-Molina hypothesis of 1974 predicted a gradual global stratospheric decline of 2–3% over the subsequent two decades if (CFC) emissions continued unchecked, based on photochemical reaction chains releasing atoms to catalytically destroy molecules. Observations from ground-based and satellite measurements confirmed a total column reduction of approximately 3% globally between the late 1970s and mid-1990s, aligning broadly with early quantitative estimates, though with greater variability influenced by quasi-biennial oscillations and solar cycles. In polar regions, predictions initially underestimated the extent of depletion due to incomplete modeling of unique Antarctic conditions, such as polar stratospheric clouds () that facilitate heterogeneous chlorine activation on ice particles during winter. The Antarctic ozone hole, defined as total column ozone below 220 Dobson units over an area greater than 2.5 million square kilometers, was first observed in September 1985 with depletions exceeding 50% in the core, far surpassing pre-1980s model projections of 20–30% seasonal loss; subsequent refinements to chemistry-climate models incorporated PSC microphysics and vortex dynamics to better reproduce this abrupt onset, which coincided with peak stratospheric CFC concentrations around 1980–1990. Post-Montreal Protocol phase-out, models forecasted stabilization of ozone trends by the early and gradual recovery to 1980 levels by 2040–2060 for the and earlier for mid-latitudes, assuming compliance and no major perturbations. Empirical data indicate a halt in global decline around 2000, with average ozone hole areas shrinking from peaks of 25–30 million square kilometers in the to the seventh-smallest on record in (peaking at under 20 million square kilometers), consistent with declining stratospheric chlorine loading derived from CFC decay. However, year-to-year variability remains high due to meteorological factors like stratospheric temperatures and planetary wave activity, with large holes in 2020–2023 attributed to colder vortex conditions rather than ODS levels. Notable model-observation discrepancies persist, particularly in the lower (15–20 km altitude), where chemistry-climate models simulate a modest recovery in mid-latitudes not yet evident in observations from 1998–2016, potentially linked to unaccounted dynamical changes, tropical intensification, or residual bank emissions. Unexpected CFC-11 emissions rises (peaking at 13–21 Gg/year around 2012–2014, likely from unreported production in eastern ) delayed projected recovery by 1–2 years in model simulations, though emissions declined sharply post-2018, restoring alignment. These gaps underscore ongoing refinements in models, with empirical monitoring via networks like NOAA's Global Monitoring Laboratory confirming overall causal linkage through (ClO) enhancements anticorrelated with loss during depletion events.

Regulatory Framework and Phase-Out

Key International Agreements

The for the Protection of the , adopted on 22 March 1985 and entering into force on 22 September 1988, established a framework for international cooperation to protect the through , monitoring, and information exchange on human activities potentially causing depletion, including emissions of substances like chlorofluorocarbons (CFCs). It did not impose binding controls but provided the basis for subsequent regulatory measures by committing parties to assess and mitigate adverse effects. By 2009, it achieved universal ratification among UN member states. Building on the , the on Substances that Deplete the was adopted on 16 September 1987 and entered into force on 1 January 1989, mandating the phase-out of production and consumption of ozone-depleting substances, with CFCs listed under Annex A Group I as primary targets due to their high ozone-depletion potential. The original protocol required developed countries to freeze CFC production at 1986 levels, achieve a 20% reduction by 1993, and a 50% reduction by 1998, while allowing developing countries (Article 5 parties) a freeze at 1995-1997 average levels with phased reductions starting later. It has achieved universal ratification by 197 countries, the first UN treaty to do so. Subsequent amendments strengthened CFC controls: the London Amendment of 1990 required complete phase-out of CFCs by 2000 in developed countries and expanded the list of controlled substances; the Amendment of 1992 accelerated the developed-country phase-out to 1996 and mandated elimination by 2010 for developing countries, with provisions for essential uses and assistance via the Multilateral Fund established in 1991. These adjustments ensured near-total global elimination of CFC production, with developed countries completing phase-out by 1996 and developing countries by 2010, though limited production for exempted uses persisted briefly. Compliance mechanisms include trade restrictions on non-parties and mandatory reporting.

Implementation Challenges and Compliance

Implementing the phase-out of chlorofluorocarbons (CFCs) under the faced significant hurdles, particularly in developing countries classified as Article 5 nations, which received grace periods of up to 10 years for compliance but struggled with limited industrial capacity, reliance on affordable CFC-based technologies, and insufficient infrastructure for alternatives. The Multilateral Fund provided approximately $3.5 billion by 2010 to assist 147 such countries in meeting obligations through and , yet delays persisted due to economic dependencies on CFC production for and . Enforcement challenges arose from illegal production and black-market trade, which undermined phase-out efforts by supplying CFCs to regions with high demand, such as for servicing. In the United States, the Environmental Protection Agency reported confiscations of illegal CFC imports by U.S. Customs, but smuggling persisted via mislabeled shipments from non-compliant producers. A notable case involved undetected CFC-11 production in eastern , where emissions rose by about 11,000 metric tons annually around 2013–2017, detected through atmospheric monitoring and later traced to foam insulation factories evading bans. Compliance mechanisms, including mandatory reporting to the Ozone Secretariat and review by the Implementation Committee, addressed approximately 117 non-compliance cases by 2023, primarily through technical assistance rather than penalties, which were rarely imposed. Developing countries demonstrated higher rates with financial incentives, but ongoing illegal —estimated to represent up to 20–30% of HCFC supply in some markets—highlighted gaps in controls and cooperation, particularly between Article 5 and non-Article 5 nations where legal loopholes facilitated . Despite these issues, the Protocol's non-punitive approach fostered voluntary adherence, with over 99% reduction in CFC consumption achieved by , though vigilance against illicit activities remains essential.

Alternatives and Technological Transitions

Hydrochlorofluorocarbons and Hydrofluorocarbons

Hydrochlorofluorocarbons (HCFCs) emerged as interim replacements for chlorofluorocarbons (CFCs) following the Protocol's initial phase-out of CFCs, primarily in , , foam production, and solvent applications. These compounds incorporate hydrogen atoms, which reduce their (ODP) compared to CFCs by enabling more rapid degradation in the , thereby limiting chlorine release in the ; HCFCs typically exhibit ODPs of 0.01 to 0.1, versus 0.6 to 1.0 for CFCs. Despite this improvement, HCFCs still contribute to loss and have lifetimes of 1 to 20 years, shorter than CFCs' 50 to 100 years. The designated HCFCs as Class II ozone-depleting substances with a structured phase-out to allow technological adaptation while minimizing continued environmental harm. Developed countries (Article 2 parties) committed to freezing HCFC consumption at 1989 levels by 1996, achieving 65% reduction by 2010, 90% by 2015, 99.5% by 2020, and complete phase-out by January 1, 2020, except for limited servicing stocks. Developing countries (Article 5 parties) began reductions later, with a freeze at 2009-2010 averages in 2013, 10% cut by 2015, and full elimination targeted by 2030. , HCFC-22 (the most common variant) faced production bans for new equipment by 2010 and servicing allowances tapering to zero by 2030, supported by Multilateral Fund financing for compliance in lower-income nations. Global HCFC emissions peaked around 2010 and have since declined, aiding stratospheric recovery, though illegal production persists in some regions. Hydrofluorocarbons (HFCs) succeeded HCFCs as primary substitutes, valued for zero ODP since they contain no or to catalyze ozone breakdown. Widely adopted in aerosols, refrigerants (e.g., HFC-134a in ), and insulating foams, HFCs addressed concerns but introduced potent effects, with 100-year GWPs from 124 (HFC-152a) to 14,800 (HFC-23), far exceeding CO2's GWP of 1. Atmospheric concentrations of key HFCs rose rapidly post-1990s, driven by expanding cooling demand in developing economies; by 2020, HFCs accounted for about 2% of total anthropogenic on a CO2-equivalent basis. The , adopted October 15, 2016, and effective September 23, 2019, extended the to phase down HFCs globally, requiring developed countries to reduce consumption by 85% from 2011-2013 baselines by 2036, while developing countries freeze levels in 2024-2028 (or 2028-2030 for hotter climates like India's) and achieve 80-85% cuts by 2045-2047. HFC consumption trends show stabilization in developed markets due to early regulations but continued growth in , where projections without Kigali could add 0.3-0.5°C to warming by 2100; amendment compliance is forecasted to avert 0.04°C of additional warming. Transitions to lower-GWP alternatives, such as HFOs (hydrofluoroolefins) or natural refrigerants like CO2 and , are accelerating, though challenges include flammability, efficiency losses, and higher upfront costs in systems. As of 2024, over 140 parties have ratified Kigali, with monitoring via atmospheric observations confirming adherence in major emitters, though HFC-23 emissions from byproduct destruction remain a compliance concern.

Non-Halocarbon Substitutes and Innovations

Non-halocarbon substitutes for chlorofluorocarbons (CFCs) primarily encompass natural refrigerants such as hydrocarbons, , and , which contain no chlorine, fluorine, or other , thereby exhibiting zero (ODP). These alternatives emerged as viable options following the Protocol's restrictions on CFCs, offering low or negligible (GWP) compared to while supporting phase-out transitions in , , and related applications. Unlike transitional hydrochlorofluorocarbons (HCFCs) or hydrofluorocarbons (HFCs), these substances rely on established thermodynamic properties without synthetic , though their adoption has required engineering adaptations for safety and efficiency. Hydrocarbons, including (R-290) and (R-600a), have been approved by regulatory bodies for use in domestic refrigerators, small air conditioners, and heat pumps, with charge limits to mitigate flammability risks. These refrigerants demonstrate superior , with R-290 achieving up to 15% better performance than R-134a in comparable systems and R-600a offering up to 10% gains, alongside GWPs of 3 and less than 1, respectively. The U.S. Environmental Protection Agency (EPA) listed propane, isobutane, and mixtures like R-441A as acceptable substitutes for retail food refrigeration in 2011, subject to use conditions such as secondary loop designs to prevent direct leaks. adoption has expanded in and for household appliances, reducing reliance on high-GWP fluorocarbons. Ammonia (R-717) serves as a for and large-scale systems, prized for its exceptional thermodynamic —3% to 10% higher than HFC-based systems—and GWP of zero. Long utilized predating CFC dominance, systems have been modernized with indirect configurations and to address concerns, enabling deployment in and facilities. The International Institute of Ammonia notes its non-contribution to , positioning it as a direct CFC successor without atmospheric persistence issues. Innovations include hybrid ammonia-CO2 cascade systems, which enhance capacity in low-temperature applications while distributing risks. Carbon dioxide (R-744) has gained traction in transcritical and subcritical cycles for commercial refrigeration and heat pumps, particularly in supermarkets and vehicle air conditioning, due to its non-flammability, abundance, and GWP of 1. Systems employing CO2 often achieve energy savings in moderate climates through efficient heat rejection, though high-pressure components demand robust materials. The European Commission highlights CO2 alongside ammonia for chiller applications under high ambient temperatures, with real-world installations demonstrating reliability since the early 2000s. Complementary innovations involve water (R-718) in absorption chillers for air conditioning, leveraging waste heat for operation without mechanical compression, though limited to larger-scale, lower-efficiency scenarios. For propellants and blowing agents, non-halocarbon options include compressed gases like CO2 or , and hydrocarbons such as isomers, which avoid impacts entirely. These shifts, documented in industry handbooks, have minimized emissions in without compromising product performance, though hydrocarbons require controls due to . Overall, these substitutes underscore a return to pre-CFC fluids, bolstered by innovations like improved and charge optimization to offset handling challenges.

Current Status and Recent Developments

Atmospheric concentrations of chlorofluorocarbons (CFCs) are monitored through global networks such as the NOAA Global Monitoring Laboratory (GML) and the Advanced Global Atmospheric Gases Experiment (AGAGE), which collect air samples from remote baseline stations including (), Cape Grim (), and . These networks use flask sampling, in-situ , and to measure mixing ratios in parts per trillion (), with data calibrated against standards for accuracy and intercomparability. Long-term records span decades, enabling detection of annual changes as small as 1-2 for major CFCs like CFC-11 and CFC-12. Post-1987 , atmospheric abundances of long-lived CFCs began declining as emissions ceased, with CFC-11 mixing ratios peaking at approximately 270 in the mid-1990s and decreasing thereafter at rates accelerating to about 1 per year by the . CFC-12, the most abundant CFC, reached a maximum of around 540 circa 2003-2004, followed by a steady decline averaging 0.8-1.0 annually through 2020, reflecting reduced releases from existing banks and compliant phase-out. The equivalent effective chlorine (EESC), a metric aggregating -depleting potential from multiple halocarbons, peaked in the late 1990s and has fallen by about 1% per year since, as tracked by the NOAA Ozone Depleting Gas Index (ODGI). Recent through confirm continued declines, with global CFC-11 emissions dropping to 45 ± 10 gigagrams (Gg) in both and 2020, and an 18 ± 6 Gg per year (26 ± 9%) reduction from 2018 to , indicating cessation of prior anomalous sources. CFC-12 burdens remain on a downward trajectory, contributing to lower stratospheric levels, though minor increases in trace CFCs (e.g., CFC-113a at 0.01-0.02 ppt growth) are observed from non-prohibited uses like feedstocks. Overall, these trends align with model predictions under compliance, with surface abundances 20-30% below peak values for primary CFCs. Monitoring continues to refine estimates, incorporating from instruments like ACE-FTS for vertical profiles.

Unexplained Emissions Increases

In , atmospheric observations revealed an unexpected slowdown in the decline of CFC-11 () concentrations, with global emissions increasing by an estimated 13 ± 5 gigagrams per year—or approximately 25 ± 13%—since 2012, contrary to expectations under the Protocol's phase-out. This rise, initially unexplained, was later attributed primarily to unreported production in eastern , where facilities violated production quotas for foam-blowing applications, accounting for 40-60% of the anomaly based on isotopic and regional modeling. Following international scrutiny and enforcement actions by Chinese authorities starting in , CFC-11 emissions declined sharply, returning to pre-2012 trends by 2020-2021, with annual emissions dropping by over 10,000 tonnes per year from their peak. However, a 2025 indicated that stockpiled (banked) CFC-11 from historical production continues to contribute to persistent emissions, averaging 17 (range: 10-20) kilotonnes per year from 2010 to 2022, exceeding model predictions and delaying recovery timelines while adding to burdens. These banked releases, primarily from legacy equipment and delayed destruction, represent an unforeseen ongoing source, as emissions did not align with expected bank depletion rates under compliance scenarios. More recently, emissions of five minor regulated CFCs—CFC-13, CFC-112a, CFC-113a, CFC-114a, and CFC-115—rose 2.6-fold between and , based on global flask sampling networks like AGAGE and NOAA. While CFC-113a, CFC-114a, and CFC-115 may stem partly from allowable trace byproducts in (HFC) manufacturing processes (exempted under provisions if unintentional and minimized), sources for CFC-13 and CFC-112a remain largely unexplained, with no legitimate uses; potential origins include inadvertent releases from plasma arc destruction of CFCs, aluminum smelting, or unreported side production tied to past non-compliance in regions like . In ozone-depletion terms, the combined 2020 emissions of these five equated to about one-tenth the impact of contemporaneous CFC-11 releases, posing minimal short-term risk to overall stratospheric recovery but underscoring gaps in emission reporting and byproduct controls. Ongoing monitoring is essential, as unchecked growth could compound delays in returning the to 1980 levels.

Broader Impacts and Assessments

Environmental and Climate Effects

Chlorofluorocarbons (CFCs) primarily exert environmental effects through their role in . Upon release into the atmosphere, CFCs ascend intact to the due to their , where ultraviolet radiation photodissociates them, liberating atoms. Each atom initiates a : Cl + O₃ → ClO + O₂, followed by ClO + O → Cl + O₂, resulting in the net destruction of (O₃) molecules without consuming the catalyst. Laboratory studies confirm that a single atom can destroy up to 100,000 molecules before being neutralized, amplifying the depletion efficiency. Empirical observations link CFC emissions to widespread ozone loss, particularly the Antarctic ozone hole, first detected in 1985 via ground-based and satellite measurements showing springtime column reductions exceeding 50% below pre-1970s baselines. Stratospheric chlorine levels, derived from CFC breakdown, peaked at approximately 3.66 (ppb) in 1993, correlating with maximum depletion rates; by 2020, total stratospheric chlorine had declined 11.5% to 3.24 ppb, reflecting reduced CFC inputs post-1987 . This depletion increased ultraviolet-B (UV-B) radiation reaching Earth's surface by 10-20% in affected regions, elevating risks to marine , amphibian populations, and terrestrial ecosystems via DNA damage and reduced productivity, as documented in field studies. Ozone recovery trends substantiate the causal role of CFCs: satellite data from NASA's mission indicate a 20% reduction in ozone loss between 2005 and 2016, directly attributable to declining CFC concentrations, with the 2024 ozone hole ranking as the seventh-smallest since phase-out began. However, unexpected CFC-11 emissions rises since 2013, estimated at 13 gigagrams per year, have introduced delays, potentially slowing full recovery by years if unchecked, though current impacts on depletion rates remain modest. In terms of climate effects, CFCs act as potent gases by absorbing infrared radiation in the and , with global warming potentials (GWPs) far exceeding over 100-year horizons. CFC-11 (CCl₃F) has a 100-year GWP of 4,660 to 6,230 relative to CO₂, while CFC-12 (CCl₂F₂) ranges from 10,200 to 12,500, based on efficiencies and atmospheric lifetimes exceeding 50 years. Pre-phase-out emissions contributed significantly to , accounting for roughly 10-15% of total forcing in the late , exacerbating tropospheric warming. Ozone depletion induced by CFCs also influences climate indirectly: stratospheric cooling from reduced ozone absorption of solar radiation alters circulation patterns, potentially intensifying the and contributing to tropospheric warming trends observed since the 1980s. Declining CFC levels have mitigated further forcing, with projected avoidance of 0.5-1°C additional warming by 2100 due to the , though legacy atmospheric burdens persist, with current concentrations around 0.23 ppb for CFC-11 and 0.53 ppb for CFC-12. Unregulated emissions could amplify these effects, underscoring the dual ozone-climate linkage.

Health, Safety, and Economic Consequences

Chlorofluorocarbons (CFCs) exhibit low acute toxicity in typical environmental or occupational exposures, with symptoms of overexposure including dizziness, central nervous system depression, and cardiac arrhythmias at concentrations exceeding 11% by volume. Prolonged or high-level inhalation can lead to respiratory irritation, nausea upon ingestion, and in extreme cases, fatalities from asphyxiation or arrhythmia, as documented in incidents involving CFC-113 leaks during industrial maintenance. Occupational studies have linked unprotected CFC exposure in refrigeration workers to cardiotoxicity, manifesting as arrhythmias, though such risks are mitigated by ventilation and exposure limits. Indirect health consequences arise primarily from CFCs' role in stratospheric ozone depletion, which elevates ultraviolet B (UVB) radiation reaching Earth's surface, increasing incidences of skin cancers (including melanoma), cataracts, and immunosuppression. Epidemiological data indicate that without interventions like the Montreal Protocol, UVB increases could have caused millions of additional skin cancer cases annually by the mid-21st century, with children and fair-skinned populations at highest risk due to cumulative exposure. Safety concerns with CFCs in applications like and center on handling hazards rather than inherent flammability, as CFCs are non-flammable and chemically stable under normal conditions. CFC releases can cause upon skin contact, and confined-space leaks pose asphyxiation risks from oxygen displacement, prompting regulatory guidelines for atmospheric monitoring during repairs. Aerosol propellants presented minimal direct safety issues for consumers, but pressurized system failures could generate irritating vapors. The economic consequences of CFC phaseout under the involved upfront transition costs for industries, including refrigerant retrofits and alternative development, estimated in billions for compliance in developed nations by the . However, cost-benefit analyses project net global benefits exceeding $2 trillion by 2060 from averted health damages, agricultural losses, and materials degradation due to reduced UV exposure, far outweighing phaseout expenditures. Developing countries received multilateral funding via the Multilateral Fund to offset costs, enabling technology transfers that spurred innovations in hydrofluorocarbons and other substitutes, though some sectors like foam manufacturing faced short-term disruptions.

References

  1. [1]
    Chlorofluorocarbons and Ozone Depletion
    Apr 18, 2017 · Molina discovered that chlorofluorocarbons (CFCs) could deplete Earth's atmospheric ozone layer, which blocks the sun's damaging ultraviolet ...Missing: peer- | Show results with:peer-
  2. [2]
    [PDF] Atmospheric histories and emissions of chlorofluorocarbons CFC-13 ...
    Jan 25, 2018 · Chlorofluorocarbons (CFCs) are very stable man-made com- pounds known to destroy stratospheric ozone. For this reason they were regulated for ...
  3. [3]
    Scientific Assessment of Ozone Depletion 2018: Executive Summary
    The Montreal Protocol has led to decreases in ozone-depleting substances and the start of ozone recovery, with the Antarctic hole recovering, and upper ...<|control11|><|separator|>
  4. [4]
    Ozone-Depleting Substances | US EPA
    Mar 25, 2025 · chlorofluorocarbons chlorofluorocarbons Gases covered under the 1987 Montreal Protocol and used for refrigeration, air conditioning ...Missing: evidence | Show results with:evidence
  5. [5]
    Joint inference of CFC lifetimes and banks suggests previously ... - NIH
    May 18, 2021 · Chlorofluorocarbons (CFCs) are harmful ozone depleting substances and greenhouse gases. CFC production was phased-out under the Montreal ...
  6. [6]
    [PDF] Global increase of ozone-depleting chlorofluorocarbons from 2010 ...
    Chlorofluorocarbons (CFCs) are industrially produced compounds that were commonly used as refrigerants, aerosol propellants, sol- vents and foam-blowing agents.
  7. [7]
    [PDF] 1 Chlorofluorocarbons (CFCs) described as "miracle chemicals ...
    CFCs are halocarbons that contain only the elements carbon, chlorine, and fluorine. The most common CFCs are small molecules containing only one or two carbon ...Missing: classes | Show results with:classes
  8. [8]
    Chlorofluorocarbons (CFCs) - Global Monitoring Laboratory - NOAA
    CFCs are classified as halocarbons, a class of compounds that contain atoms of carbon and halogen atoms. Individual CFC molecules are labeled with a unique ...
  9. [9]
    [PDF] Naming Convention for CFCs & Halons - UMD
    CFCs are named using i, j, and k. i is carbon atoms minus 1, j is hydrogen atoms plus 1, k is fluorine atoms. Chlorine is calculated as 2(C+1) - H - F.
  10. [10]
    3-I-1 NUMBERING SYSTEM FOR CFCS, HCFCS, AND HALONS
    For CFCs/HCFCs, the 1's digit is fluorine, 10's is hydrogen+1, 100's is carbon-1. Halons use 1's to 1000's for Br, Cl, F, and C atoms.
  11. [11]
    Cfc 113 | Cl2FCCClF2 | CID 6428 - PubChem - NIH
    Cfc 113 | Cl2FCCClF2 or C2Cl3F3 | CID 6428 - structure, chemical names, physical and chemical properties, classification, patents, literature, ...
  12. [12]
    Dichlorodifluoromethane | CCl2F2 | CID 6391 - PubChem - NIH
    Solubility of water in dichlorodifluoromethane, 0.009 wt% at 25 °C. Flick, E.W. ... Pressurized refrigerant or liquid fluorocarbons with a low boiling point ...
  13. [13]
    [PDF] Chlorinated fluorocarbons (CFC) - California Air Resources Board
    CFC-113 is a colorless, volatile gas with a nearly odorless smell. It is soluble in alcohol, ether, and benzene (Merck,. 1989). Physical Properties of CFC-113.
  14. [14]
    [PDF] Fully Halogenated Chlorofluorocarbons - Columbia University
    They are usually characterized by high vapour pressure and density and low viscosity, surface tension, refractive index, and solu- bility in water.
  15. [15]
  16. [16]
    Chlorofluorocarbons Background - USGS Groundwater Dating Lab
    Jun 15, 2016 · CFCs are nonflammable, noncorrosive, nonexplosive, very low in toxicity, and have physical properties conducive to a wide range of industrial ...
  17. [17]
    CFC-12 - ChemBK
    CFC-12 - Physico-chemical Properties ; Molar Mass, 88.0043 ; Density, 1.315g/cm ; Melting Point, -158℃ ; Water Solubility, Insoluble. 0.028 g/100 mL ; Vapor Presure ...<|control11|><|separator|>
  18. [18]
    CFC11 - Molecule of the Month - June 2019 (HTML version)
    Indeed, this became known as the Swarts reaction. It seems to have been used by Midgley to make CFCs in the 1920s. SbF3Br2 + CCl4 CFCl3 + SbF2Br2Cl. Other ...
  19. [19]
    Swarts reaction mechanism - BYJU'S
    The Swarts reaction mechanism is quite simple – the metal fluorine bond is broken and a new bond is formed between carbon and fluorine. The displaced chlorine ...
  20. [20]
    Chlorofluorocarbon - an overview | ScienceDirect Topics
    Chlorofluorocarbons (CFCs) are man-made organic compounds which are produced for a range of industrial and domestic purposes (Rowland, 1990).
  21. [21]
    Investigation of East Asian Emissions of CFC-11 Using Atmospheric ...
    CFC-11 has historically been produced via fluorination of CCl4 to produce a mixture of CFC-11 (CCl3F) and CFC-12 (CCl2F2). The production ratio has typically ...
  22. [22]
    [PDF] Background Information - Overview - CFC-11 emissions
    A mix of CFC-11 and CFC-12 is produced, with the proportion of CFC-12 and CFC- 11 controlled by varying the operating conditions.
  23. [23]
    History of Air Conditioning - Department of Energy
    Jul 20, 2015 · Chlorofluorocarbon refrigerants created. 1928. Thomas Midgley, Albert Henne and Robert McNary of General Motors synthesize chlorofluorocarbon ( ...Missing: Freon | Show results with:Freon
  24. [24]
    (R)Evolution of Refrigerants | Journal of Chemical & Engineering Data
    Jul 2, 2020 · Through the 1930s, 1940s, and 1950s additional CFCs and HCFCs (hydrochlorofluorocarbons) were developed and commercialized. Significant among ...
  25. [25]
    The History of Freon - ThoughtCo
    Mar 4, 2019 · In 1928, Midgley and Kettering invented a "miracle compound" called Freon. Frigidaire received the first patent, US#1,886,339, for the formula ...Missing: synthesis | Show results with:synthesis
  26. [26]
    Air Conditioning and Refrigeration Timeline | ashrae.org
    Clorofluorocarbon refrigerants synthesized by General Motors Research Lab team of Thomas Midgley, Albert Henne and Robert McNary for Frigidaire. Announced ...
  27. [27]
    The Refrigerant Story: From R-22 to R-410A | Goodman Manufacturing
    In 1928, Thomas Midgley, Albert Henne and Robert McNary created chlorofluorocarbon (CFC) refrigerants.
  28. [28]
    The Brilliant Inventor Who Made Two of History's Biggest Mistakes
    Mar 17, 2023 · A century ago, Thomas Midgley Jr. was responsible for two phenomenally destructive innovations. What can we learn from them today?
  29. [29]
    Midgley Introduces Dichlorodifluoromethane as a Refrigerant Gas
    Thomas Midgley, Jr., introduced dichlorodifluoro- methane as a safe refrigerant gas for domestic refrigerators, leading to rapid growth in the acceptance of ...
  30. [30]
    [PDF] Aerosol History
    American soldiers used it to fight insect borne diseases in the pacific; 50 million units were produced during World War II. The aerosol package expanded soon ...
  31. [31]
    History of aerosol development-sihai-aerosolcan.com
    Jan 11, 2023 · Soon after the war, aerosol packaging expanded. in 1947, 4.3 million units were produced for civilian use. Production grew steadily and soon ...
  32. [32]
    Chlorofluorocarbon Refrigerant - an overview | ScienceDirect Topics
    One year later, the dichlorodifluoromethane (R-12) were introduced as commercial refrigerant used in refrigerators followed by R-11 a year later. Since ...<|control11|><|separator|>
  33. [33]
    (R)Evolution of Refrigerants - PMC - NIH
    The CFCs and HCFCs came to dominate most refrigeration and air-conditioning applications, and this period lasted roughly 60 years, into the 1990s. Ammonia in ...
  34. [34]
    The History of Therapeutic Aerosols: A Chronological Review - PMC
    In 1987, the Montreal Protocol was signed and called for the elimination of CFC propellants with a date of January 1996 eventually agreed upon. Although orally ...
  35. [35]
    Questions and Answers About SNAP | US EPA
    CFCs are commonly used as refrigerants, solvents, and foam blowing agents. The most common CFCs are CFC-11, CFC-12, CFC-113, CFC-114, and CFC-115. The ozone ...
  36. [36]
    Chlorofluorocarbon - an overview | ScienceDirect Topics
    Chlorofluorocarbons have been used as refrigerants, fire extinguishing agents, local anesthetics, aerosol propellants, blowing agents for foams, chemical ...Missing: applications besides
  37. [37]
    The Chlorofluorocarbon Problem - Resources Magazine
    Feb 1, 1983 · These uses range all the way from hair sprays and pesticides to making polyurethane foams and degreasing parts for heavy machinery.Missing: applications besides
  38. [38]
    Working Toward World Without CFCs - The Donella Meadows Project
    CFCs also turned out to be ideal solvents for cleaning and degreasing electronic circuit boards. They are great gases for blowing stuff out of aerosol cans and ...
  39. [39]
    Chlorofluorocarbon - an overview | ScienceDirect Topics
    They have also found widespread use as foam blowing agents, aerosol propellants, and chemical solvents (see Table 2). Global production of CFCs reached a ...
  40. [40]
    Ozone Layer Protection and You
    CFC-11 and CFC-12 are also used as blowing agents in the manufacture of foams for home furnishing, insulation and packaging.
  41. [41]
    Stratospheric sink for chlorofluoromethanes: chlorine atom ... - Nature
    Jun 28, 1974 · Molina, M., Rowland, F. Stratospheric sink for chlorofluoromethanes: chlorine atom-catalysed destruction of ozone. Nature 249, 810–812 (1974).
  42. [42]
    [PDF] Mario J. Molina - Nobel Lecture
    In early 1974 F. Sherwood (Sherry) Rowland and I proposed that chloro- fluorocarbons (CFCs) would decompose in the stratosphere, releasing chlo- rine atoms ...<|separator|>
  43. [43]
    [PDF] Chlorofluorocarbons and Ozone Depletion
    A National Historic Chemical Landmark​​ Sherwood Rowland and Mario J. Molina discovered that chlorofluorocarbons (CFCs) could deplete Earth's atmospheric ozone ...Missing: definition | Show results with:definition
  44. [44]
    Rowland & Molina Suggest that CFCs Deplete the Ozone Layer
    McElroy and Wofsy extended the work of Rowland and Molina by showing that bromine atoms were even more effective catalysts for ozone loss than chlorine atoms ...Missing: paper | Show results with:paper
  45. [45]
    Full article: Stratospheric ozone, global warming, and the principle of ...
    Between 1975 and 1985, scientists debated but strongly supported the Molina–Rowland theory of stratospheric ozone depletion, although estimates of likely ozone ...
  46. [46]
    Stratospheric ozone, global warming, and the principle of ... - PubMed
    In 1974, Mario Molina and F. Sherwood Rowland warned that chlorofluorocarbons (CFCs) could destroy the stratospheric ozone layer that protects Earth from ...Missing: paper | Show results with:paper
  47. [47]
    Press release: The 1995 Nobel Prize in Chemistry - NobelPrize.org
    Molina and Rowland realised that the chemically inert CFC could gradually be transported up to the ozone layer, there to be met by such intensive ultra-violet ...
  48. [48]
    Stratospheric ozone – depletion, recovery and new challenges - ACP
    Mar 1, 2024 · We summarise current important and well-established open issues related to the depletion of stratospheric ozone and discuss some newly emerging challenges.
  49. [49]
    Thermodynamics and Reaction Mechanisms of Ozone Depletion
    Feb 3, 2025 · This paper investigates the chemical processes involved in ozone (O₃) depletion, with a focus on the role of chlorofluorocarbons (CFCs) in the ...Missing: peer- | Show results with:peer-
  50. [50]
    [PDF] Scientific ASSeSSment of ozone Depletion: 2022
    NOx chemistry controls ozone amounts in the stratosphere and troposphere. In ... CFC-11 emissions led to scientific investigations and policy responses ...
  51. [51]
    Stratospheric ozone depletion - PMC - NIH
    The growing emissions of synthetic chlorofluorocarbon molecules cause a significant diminution in the ozone content of the stratosphere.Missing: peer- | Show results with:peer-
  52. [52]
    [PDF] Studies on Ozone Destruction Expand Beyond Antarctic
    May 30, 1988 · The evidence now is irrefutable: Chlorine from chlorofluorocarbons. (CFCs) destroys ozone in the strato-.
  53. [53]
    NASA Data Aids Ozone Hole's Journey to Recovery
    Apr 15, 2020 · The data disproved other theories and gave scientists evidence that CFCs were causing the ozone hole. Strahan and her colleagues' data would not ...
  54. [54]
    how science and international cooperation closed the ozone hole
    May 14, 2025 · In 1974, the chemists Mario Molina and Sherwood Rowland had highlighted the perils of chlorofluorocarbons (CFCs) for stratospheric ozone.
  55. [55]
    NASA Study: First Direct Proof of Ozone Hole Recovery Due to ...
    Jan 4, 2018 · This study is the first to use measurements of the chemical composition inside the ozone hole to confirm that not only is ozone depletion ...
  56. [56]
    [PDF] Scientific Assessment of Ozone Depletion 2022
    ... CFCs plus HCFCs are comparable to those of HFCs. Reductions in the future emissions of CFCs and HCFCs requires addressing releases from banks and continuing ...
  57. [57]
    [PDF] Should We Worry About Ozone? - National Center for Policy Analysis
    If the ozone layer is being depleted, it is not clear that CFCs are the major cause or that reducing their use now will stop or even slow depletion.
  58. [58]
    Ozone Depletion's Lessons for Global Warming - The Heartland ...
    Fred Singer, president of the Science and Environmental Policy Project and longtime critic of ozone alarmism. According to Singer and others, no cause and ...
  59. [59]
    [PDF] WHAT COULD BE CAUSING GLOBAL OZONE DEPLETION?
    o Natural effects related to the variability of solar cycles may also be responsible for an observed ozone decline. The analyses implicitly assume perfect ...
  60. [60]
    Volcanoes affect atmospheric ozone, our friend and foe - USGS.gov
    Volcanoes play an interesting role in the destruction of ozone. For instance, hydrogen chloride, a common volcanic gas, efficiently destroys ozone.
  61. [61]
    [PDF] Sorting Out Ozone - Fraser Institute
    Scientists think that there may be a decline in ozone caused by some chemicals, but this decline is so small that it is hard to distinguish from natural changes ...
  62. [62]
    Ozone: The Hole Truth | The Heritage Foundation
    Sep 14, 2007 · The lurid predictions of ozone depletion-induced skin cancer epidemics, ecosystem destruction and others haven't come true.
  63. [63]
    Scientific Assessment of Ozone Depletion 2022: Executive Summary
    The Montreal Protocol's control of ODSs stimulated the development of replacement substances, firstly HCFCs and then HFCs, in a number of industrial sectors.
  64. [64]
    Understanding Model‐Observation Discrepancies in Satellite ...
    Dec 14, 2022 · We examine multiple factors in the representation of satellite-retrieved atmospheric temperature diagnostics in historical simulations of climate change
  65. [65]
    World of Change: Antarctic Ozone Hole - NASA Earth Observatory
    In the early 1980s, scientists began to realize that CFCs were creating a thin spot—a hole—in the ozone layer over Antarctica every spring.
  66. [66]
    2024 Antarctic ozone hole ranks 7th-smallest since recovery began
    Oct 30, 2024 · NOAA and NASA scientists project the ozone layer could fully recover by 2066. During the peak of ozone depletion season from September 7 through ...
  67. [67]
    Inconsistencies between chemistry–climate models and ... - ACP
    Aug 20, 2020 · We show that CCMs display an ozone decline similar to that observed in the tropics over 1998–2016, likely driven by an increase in tropical upwelling.
  68. [68]
    A decline in global CFC-11 emissions during 2018−2019 - Nature
    Feb 10, 2021 · We find that global CFC-11 emissions decreased by 18 ± 6 gigagrams per year (26 ± 9 per cent; one standard deviation) from 2018 to 2019.
  69. [69]
    The Vienna Convention for the Protection of the Ozone Layer
    The Convention aimed to promote cooperation among nations by exchanging information on the effects of human activities on the ozone layer. In doing so, the ...
  70. [70]
    The Montreal Protocol on Substances that Deplete the Ozone Layer
    This phase-out plan includes both the production and consumption of ozone-depleting substances. The landmark agreement was signed in 1987 and entered into force ...Treaties · Summary of control measures... · Introduction
  71. [71]
    International Treaties and Cooperation about the Protection of ... - EPA
    Mar 25, 2025 · The Vienna Convention. Adopted in 1985, The Vienna Convention for the Protection of the Ozone Layer is the precursor to the Montreal Protocol.
  72. [72]
    The Montreal Protocol on Substances that Deplete the Ozone Layer
    Mar 20, 2025 · The Montreal Protocol is signed by 197 countries – the first treaty in the history of the United Nations to achieve universal ratification.
  73. [73]
    Article 5: Special situation of developing countries | Ozone Secretariat
    The Montreal Protocol on Substances that Deplete the Ozone Layer · Article 5: Special situation of developing countries.Missing: cases | Show results with:cases
  74. [74]
    Vienna Convention and its Montreal Protocol
    Through the Fund, approximately US$3.5 billion has been provided to date to assist 147 developing countries to meet their Montreal Protocol obligations. The ...
  75. [75]
    Ozone-Depleting Substances on the Black Market | US EPA
    Sep 23, 2025 · The most immediate consequence of possessing illegal CFCs/HCFCs is having them confiscated. The U.S. Customs Service may confiscate any goods ...Missing: difficulties | Show results with:difficulties
  76. [76]
    How an Illicit Chemical Is Jeopardizing Recovery of the Ozone Layer
    May 23, 2019 · New research finds a troubling rise in emissions of a banned chemical that is being routinely used in factories in northeast China.Missing: difficulties | Show results with:difficulties
  77. [77]
    [PDF] The Montreal Protocol compliance mechanism
    • Special situation of developing countries: • Up to 10-year grace period ... • Approximately 117 non-compliance cases (resolved to date) available here.
  78. [78]
    [PDF] Illegal Trade in Ozone Depleting Substances
    As long as trade in ODS between Article 5 and non-Article 5 countries continues to be legal, controlling illegal trade will be a challenge that could prove to ...
  79. [79]
    About Montreal Protocol - UNEP
    The Montreal Protocol on Substances that Deplete the Ozone Layer is the landmark multilateral environmental agreement that regulates the production and ...
  80. [80]
    Phaseout of Class II Ozone-Depleting Substances | US EPA
    Aug 7, 2025 · Although 34 HCFCs are subject to the phaseout, only a few were commonly used. Historically, the most widely used included HCFC-22 (usually as a ...
  81. [81]
    [PDF] Are there controls on the production of ozone-depleting substances?
    The Montreal Protocol provides for the use of hydrochlorofluorocarbons (HCFCs) as transitional, short- term substitute compounds for ODSs with higher ODPs, such ...
  82. [82]
    What Is the Phaseout of Ozone-Depleting Substances? | US EPA
    Jul 30, 2025 · The Montreal Protocol on Substances That Deplete the Ozone Layer and its amendments control the phaseout of ODS production and use.
  83. [83]
    A class of ozone-depleting chemicals is declining, thanks to the ...
    Jun 13, 2024 · Developing countries agreed to start their phase-out process in 2013 and with a complete phase-out of HCFCs targeted by 2030.
  84. [84]
    [PDF] EPA Phaseout of Class II ODS
    Class II ODS are HCFCs, with a complete phaseout by 2030. HCFC usage must be 90% below baseline in 2015 and 99.5% in 2020.
  85. [85]
    An International Climate Success Story: HCFCs
    Jun 13, 2024 · The successful global emissions reductions of HCFCs through the Montreal Protocol provide several key lessons for engaging in other emission ...Missing: substitutes | Show results with:substitutes
  86. [86]
    Hydrofluorocarbons (HFCs) | Climate & Clean Air Coalition
    HFCs were introduced to replace ozone depleting substances and the growing demand for cooling. ... HFC-23 has a global warming potential (GWP) that is 14,800 ...
  87. [87]
    Hydrofluorocarbons saved the ozone layer but are warming the earth
    Dec 13, 2021 · While HFCs do not contribute to ozone depletion, they have global warming potentials (GWP) at the magnitude of hundreds to thousands of times higher than that ...
  88. [88]
    What Are Hydrofluorocarbons? — EIA
    HFCs or hydrofluorocarbons, are super greenhouse gases because with a global warming potential hundreds to thousands of times more potent than CO2.
  89. [89]
    New research estimates the future emissions of potent greenhouse ...
    Jun 2, 2022 · In 2019, the Kigali Amendment to the Montreal Protocol came into force with the goal of limiting the use of HFCs globally, and currently, ...
  90. [90]
    Projections of hydrofluorocarbon (HFC) emissions and the resulting ...
    May 10, 2022 · Warming from HFCs is additionally limited by the Kigali Amendment controls to a contribution of about 0.04 ∘C by 2100.
  91. [91]
    A Study on the Impacts of HFC Consumption Trends in Article 5 ...
    Aug 19, 2022 · This report presents the findings of the HFC consumption assessment of Article 5 countries in the Montreal Protocol.
  92. [92]
    Climate-friendly alternatives to HFCs
    Most of the HFCs are used as refrigerants in refrigeration and air conditioning (RAC) equipment, but also as blowing agents, aerosol propellants and solvents.Missing: besides | Show results with:besides
  93. [93]
    Emissions of HFC-23 do not reflect commitments made under ... - NIH
    Dec 21, 2024 · Our findings suggest that HFC-23 emissions have not been destroyed to the extent reported by the Parties since the implementation of the Kigali Amendment.
  94. [94]
    CFCs and their substitutes in stratospheric ozone depletion.
    Besides recycling, the demand for CFCs has been further reduced by the use of substitutes. Some applications, for example degreasing of metals and cleaning ...Missing: explanations | Show results with:explanations
  95. [95]
    Environment-Friendly Refrigerants for Sustainable Refrigeration and ...
    Sep 1, 2023 · Ammonia, carbon dioxide, hydrocarbons are natural refrigerants that possess a negligible ODP and exhibit significantly lower or zero GWP ...<|separator|>
  96. [96]
    Hydrocarbons Refrigerants Isobutane (R600a) and Propane (R290)
    Jun 15, 2023 · Compared to R134a, the use of R290 (propane) can improve efficiency by up to 15%, and R600a (isobutane) can improve efficiency by up to 10%.
  97. [97]
    Alternative Refrigerants to Replace Hydrofluorocarbons in ...
    The three hydrocarbon refrigerants approved as acceptable substitutes, with use conditions, are propane, isobutane, and a chemical known as R-441A.
  98. [98]
    All News Releases By Date - EPA
    Dec 14, 2011 · The three hydrocarbon refrigerants approved as acceptable substitutes, with use conditions, are propane, isobutane, and a chemical known as R- ...
  99. [99]
    Ammonia Vs HFC/HCFC- The guide to finding a better refrigerant
    An ammonia-based industrial refrigeration system are 3 to 10% more efficient in performance than the systems that use HFCs/HCFCs as refrigerants.
  100. [100]
    Environmental Advantages - IIAR
    No. Ammonia does not harm atmospheric ozone. Ammonia is a natural refrigerant. It is not a halocarbon like many of the synthetic refrigerants on the market.Missing: non- | Show results with:non-
  101. [101]
    [PDF] not-in-kind alternatives to high global warming hfcs | c2es
    Ammonia. R-717. B2L. 2L. Used in indirect systems, but energy efficiency can be reduced if liquid secondary refrigerants are used. Carbon. Dioxide. R-744. A1. 1.Missing: halocarbon | Show results with:halocarbon
  102. [102]
    [PDF] Ozone-Depleting Substances: Alternatives
    Other solvent uses of CFC-113, methyl chloroform, and carbon tetrachloride. Specialized applications of ozone-depleting solvents include drying of com-.<|control11|><|separator|>
  103. [103]
    Ozone Depleting Gas Index (ODGI) - Global Monitoring Laboratory
    The ODGI is an index that helps to quantify the health of the Earth's ozone layer, and the amount of ozone depleting substances in the atmosphere.
  104. [104]
    A decrease in radiative forcing and equivalent effective chlorine from ...
    Jun 11, 2024 · The global total radiative forcing of HCFCs in 2023 is 61.78, 61.71 or 61.75 mW m−2 using measurements from AGAGE, NOAA or both networks, where ...<|separator|>
  105. [105]
    Historical trend of ozone-depleting substances and ...
    Jan 1, 2023 · Large discrepancies and unexplained variations are seen in the time series of HCFC-142b, CFC-12 and CCl4, necessitating further optimization of ...
  106. [106]
    An unexpected and persistent increase in global emissions of ozone ...
    May 16, 2018 · A simple model analysis of our findings suggests an increase in CFC-11 emissions of 13 ± 5 gigagrams per year (25 ± 13 per cent) since 2012.
  107. [107]
    China identified as source of unexpected rise in CFC emissions
    May 22, 2019 · At least 40-60% of the increase in emissions of a key ozone-depleting chemical reported last year can be traced back to sources in eastern China.
  108. [108]
    Emissions of a banned ozone-depleting gas are back on the decline
    Feb 10, 2021 · In 2018, a research team led by Montzka announced that from 2014 to 2016, emissions of CFC-11 had increased by more than 13,000 tonnes per year ...
  109. [109]
    Banked CFC-11 contributes to an unforeseen emission rise and sets ...
    Sep 29, 2025 · An unexpected rise of trichlorofluoromethane (CFC-11) emissions has undermined the efforts behind the Montreal Protocol.
  110. [110]
  111. [111]
    Scientific Assessment of Ozone Depletion: 1994 - NOAA CSL
    What is the Evidence that Stratospheric Ozone is Destroyed by Chlorine and Bromine? Laboratory studies show that chlorine (Cl) reacts very rapidly with ozone.
  112. [112]
    Delay in recovery of the Antarctic ozone hole from unexpected CFC ...
    Dec 19, 2019 · The recently discovered increase in CFC-11 emissions of ~ 13 Gg yr−1 may delay recovery. So far the impact on ozone is small, but if these ...Missing: empirical | Show results with:empirical
  113. [113]
    [PDF] IPCC Global Warming Potential Values - GHG Protocol
    Aug 7, 2024 · IPCC Global Warming Potential (GWP) values relative to CO2. Common ... CFC-11. CCl3F. 4,750. 4,660. 6,230. CFC-12. CCl2F2. 10,900. 10,200. 12,500.
  114. [114]
    What is the concentration of CFCs in the atmosphere, and how much ...
    Feb 13, 2023 · “Global warming potential” (GWP) measures a greenhouse gas's ability to trap heat for 100 years. Carbon dioxide has a global warming potential ...
  115. [115]
    Understanding Global Warming Potentials | US EPA
    Jan 16, 2025 · Nitrous Oxide (N2O) has a GWP 273 times that of CO2 for a 100-year timescale. N2O emitted today remains in the atmosphere for more than 100 ...
  116. [116]
    Study: The ozone hole is healing, thanks to global reduction of CFCs
    Mar 5, 2025 · A new MIT-led study confirms that the Antarctic ozone layer is healing, as a direct result of global efforts to reduce ozone-depleting substances.
  117. [117]
    Chlorofluorocarbons, fully halogenated (EHC 113, 1990) - INCHEM
    Atmospheric levels of CFC-11 and CFC-12 increased steadily up to 1985, when combined levels for these two compounds in the USA were 9120 ng/m3 in urban/suburban ...
  118. [118]
    Preventing Death from Excessive Exposure to Chlorofluorocarbon ...
    These workers were apparently unaware that CFC-113 might generate vapor concentrations sufficient to cause death by cardiac arrhythmia or asphyxiation. NIOSH ...Missing: direct | Show results with:direct
  119. [119]
    Cardiotoxicity of Freon among refrigeration services workers
    Jul 13, 2009 · Unprotected occupational exposure to chlorofluorocarbons can induce cardiotoxicity in the form of cardiac arrhythmias. The role of ...
  120. [120]
    Basic Ozone Layer Science | US EPA
    Mar 5, 2025 · Gaseous CFCs can deplete the ozone layer when they slowly rise into the stratosphere, are broken down by strong ultraviolet radiation, release ...
  121. [121]
    Environment and health: 3. Ozone depletion and ultraviolet radiation
    The health effects that are believed to be affected by an increased flux in UV radiation include skin cancer, cataract and cellular immunity; the decreased ...
  122. [122]
    Children suffer most from the effects of ozone depletion
    Sep 16, 2003 · “We know that by reducing over-exposure of children and adolescents to the sun, we can substantially reduce the risk of contracting skin cancers ...
  123. [123]
    Refrigerant Safety | US EPA
    Mar 12, 2025 · The risks associated with the use of refrigerants in refrigeration and air-conditioning equipment can include toxicity, flammability, asphyxiation, and ...
  124. [124]
    [PDF] Hazards during the Repair and Maintenance of Refrigeration ...
    Liquid releases can form aerosols that may tend to accumulate at low points. Therefore, it is important to take atmospheric readings at various locations within ...
  125. [125]
    Chlorofluorocarbon (CFC) - Canada.ca
    Jun 2, 2025 · Chlorofluorocarbons (CFCs) are organic and synthetic (ie, human-made) chemicals consisting of carbon, chlorine and fluorine.Missing: effects | Show results with:effects
  126. [126]
    [PDF] THE MONTREAL PROTOCOL AND THE GREEN ECONOMY
    Available cost benefit studies on the impacts of the Montreal Protocol have concluded that the benefits of ODS phase-out will far outweigh the costs with.<|control11|><|separator|>