Fact-checked by Grok 2 weeks ago

Trifluoromethylation

Trifluoromethylation is the chemical process of introducing a trifluoromethyl group (–CF₃) into an organic molecule, which profoundly alters its physical, chemical, and biological properties. This modification enhances metabolic stability, lipophilicity, bioavailability, receptor binding selectivity, and dipole moments, making trifluoromethylated compounds highly valuable in drug design and synthesis. The trifluoromethyl group's electron-withdrawing nature and resistance to enzymatic degradation—contributing to its metabolic stability—position trifluoromethylation as a cornerstone of modern organic synthesis. In pharmaceuticals and agrochemicals, trifluoromethylation is widely employed to improve the efficacy and duration of action of active compounds, with numerous blockbuster drugs and pesticides featuring the –CF₃ moiety, such as (Prozac) and celecoxib (Celebrex). For instance, it boosts permeability and reduces susceptibility to oxidative metabolism, thereby extending therapeutic windows. Beyond life sciences, the process supports the development of , such as fluorinated polymers with superior thermal and chemical stability. Key methodologies for trifluoromethylation have evolved to include electrophilic, nucleophilic, and approaches, often leveraging specialized for selective C–CF₃ bond formation. Electrophilic agents, such as Umemoto's sulfonium salts, enable direct trifluoromethylation of arenes, alkenes, alkynes, and heteroatoms under mild conditions. Copper-catalyzed variants, using stable reagents like potassium (trifluoromethyl)trimethoxyborate, facilitate efficient conversion of aryl halides to trifluoromethylarenes. Advances also encompass Sandmeyer-type reactions for diazonium salts and thiolation protocols for –SCF₃ analogs, expanding the toolkit for late-stage functionalization.

Introduction

Definition and Fundamentals

Trifluoromethylation refers to the introduction of a trifluoromethyl group (-CF₃) into an via the formation of a carbon-trifluoromethyl (C-CF₃) , a process central to modifying the physicochemical properties of compounds in . This transformation is pivotal in fluorine chemistry, where the -CF₃ moiety imparts distinct electronic and steric characteristics that can enhance molecular stability and reactivity. The -CF₃ group exhibits a strong electron-withdrawing , as evidenced by its Hammett constant σ_p of 0.54, which withdraws from adjacent atoms and alters the reactivity of functional groups. This effect contributes to increased , facilitating better membrane permeation and , while also promoting metabolic stability through resistance to oxidative degradation. Sterically, the -CF₃ group is moderately bulky, akin to an isopropyl , influencing spatial arrangements in molecules without excessive hindrance. A key prerequisite in trifluoromethylation is understanding the robust nature of C-F bonds, which have a of approximately 485 kJ/mol, making them among the strongest single bonds in and contributing to the group's inertness. The electron-withdrawing properties of -CF₃ also shift the of nearby acidic protons downward; for example, has a of 0.23 compared to 4.76 for acetic acid, demonstrating enhanced acidity due to stabilization of the conjugate base. As of 2020, approximately 20% of marketed pharmaceuticals incorporate , often including the -CF₃ group, to leverage these attributes for improved therapeutic profiles. Generically, trifluoromethylation reactions proceed via the coupling of an organic substrate (R-H) with a trifluoromethyl source (CF₃-X), yielding R-CF₃ + HX, where X represents a suitable leaving group, underscoring the fundamental bond-forming step without specifying particular conditions or reagents.

Importance and Applications

Trifluoromethylation plays a pivotal role in modern drug discovery, where the incorporation of the CF₃ group enhances key pharmacokinetic properties such as lipophilicity, metabolic stability, and bioavailability, contributing to the success of numerous therapeutic agents. As of 2025, approximately 20-25% of approved pharmaceuticals contain fluorine, with the trifluoromethyl moiety being a frequent feature in many of these, reflecting its widespread adoption to optimize drug-like characteristics and potency. For instance, celecoxib, a selective COX-2 inhibitor used for treating osteoarthritis and rheumatoid arthritis, includes a CF₃ group on its pyrazole ring, which increases its binding affinity to the enzyme and improves selectivity over COX-1, thereby reducing gastrointestinal side effects compared to non-selective NSAIDs. Similarly, efavirenz, an antiretroviral medication for HIV treatment, bears a CF₃-substituted benzoxazinone structure that boosts its metabolic resistance and oral bioavailability, enabling once-daily dosing and enhancing patient compliance in long-term therapy. In 2024, the FDA approved 11 new fluorine-containing drugs, underscoring the continued importance of fluorination in pharmaceutical development. In agrochemicals, trifluoromethylation is equally vital, with the -CF₃ group present in approximately 25% of commercial pesticides to improve , environmental persistence, and target specificity. Trifloxystrobin, a strobilurin-class , exemplifies this through its trifluoromethylphenyl moiety, which facilitates broad-spectrum control of fungal pathogens like and in crops such as cereals, fruits, and by inhibiting mitochondrial respiration in fungi, offering protective and curative action with low application rates. This structural feature contributes to its high potency and reduced resistance development in target organisms. Beyond pharmaceuticals and agrochemicals, trifluoromethylation extends to materials science, where CF₃-containing perfluoroalkyl substances are integral to fluoropolymers like polytetrafluoroethylene (PTFE), imparting exceptional chemical inertness, thermal stability, and non-stick properties essential for applications in coatings, seals, and electronics. A key advantage of the CF₃ group lies in its role as a bioisosteric replacement for methyl (-CH₃) or carbonyl (-C=O) functionalities, mimicking their steric and electronic profiles while conferring resistance to oxidative metabolism by cytochrome P450 enzymes, thereby extending compound half-life and reducing toxicity in vivo. For example, replacing a methyl group with CF₃ can block enzymatic hydroxylation at the alpha position, enhancing stability without significantly altering binding interactions, as demonstrated in various lead optimization studies.

Historical Development

Early Discoveries (Pre-1980)

The introduction of the trifluoromethyl group into organic molecules began in the late with the work of Frédéric Swarts, who developed the Swarts reaction involving halogen exchange with antimony trifluoride. In 1892, Swarts reported the synthesis of the first organotrifluoromethyl compound, benzotrifluoride (C₆H₅CF₃), by treating with SbF₃, marking a foundational achievement in despite the rudimentary conditions and limited scope. Early industrial applications emerged in the 1930s with the production of chlorofluorocarbons (CFCs), known as Freons, which incorporated trifluoromethyl-like motifs through hydrofluoric acid-mediated fluorination processes. Thomas Midgley and co-workers at and commercialized dichlorodifluoromethane (Freon-12, CF₂Cl₂) in 1930 using HF as a safer alternative to SbF₃, enabling widespread use as non-toxic, non-flammable refrigerants and laying groundwork for handling fluorinated alkyl groups. These developments highlighted the potential of CF₃-containing compounds but were constrained to simple haloforms rather than direct C-C bond formation. In the 1940s, key trifluoromethylating agents became available, including iodotrifluoromethane (CF₃I), first prepared through of silver trifluoroacetate with iodine. Around the same time, bis(trifluoromethyl)mercury (Hg(CF₃)₂) was synthesized via of trifluoroacetic acid salts with mercury, serving as an early organometallic source for transferring the CF₃ group under relatively mild conditions compared to prior methods. Pioneering electrochemical approaches also emerged during this decade, with the Simons process for (ECF) enabling the production of perfluorinated compounds, including those bearing CF₃ units, though primarily for aliphatic chains via anodic oxidation in . The 1950s saw initial explorations of radical-based trifluoromethylation, notably by Robert N. Haszeldine, who demonstrated the addition of CF₃ radicals from CF₃I to . In 1949, Haszeldine reported the photoinitiated or thermally induced reaction of CF₃I with , yielding 1-iodo-2-(trifluoromethyl)ethane as the primary product: \ce{CH2=CH2 + CF3I ->[h\nu \ or \ peroxide] CF3CH2CH2I} This method established radical addition as a viable route for C-C trifluoromethylation but suffered from low selectivity and yields often below 50%, requiring high temperatures (up to 250°C) or UV irradiation. Pre-1980 trifluoromethylation efforts were plagued by inherent challenges, including the scarcity of stable CF₃ sources, incompatibility with functional groups, and demanding conditions that led to side reactions and poor efficiency. For instance, Hg(CF₃)₂ transfers proceeded in low yields due to mercury's toxicity and the agent's volatility, while radical methods with CF₃I often required excess reagent and initiators, limiting scalability. These limitations underscored the need for more accessible reagents, setting the stage for later advancements.

Key Milestones (1980-2020)

In 1984, Ingo Ruppert and colleagues synthesized trimethyl(trifluoromethyl)silane (TMSCF₃), marking a pivotal advancement in nucleophilic trifluoromethylation by providing a stable, storable source of the CF₃ group. This reagent, later known as the Ruppert-Prakash reagent, was activated by fluoride ions to generate a nucleophilic CF₃ species, enabling efficient addition to carbonyl compounds. In 1989, G. K. Surya Prakash demonstrated that tetrabutylammonium fluoride (TBAF) or cesium fluoride could initiate the reaction, yielding trifluoromethylated alcohols after . A representative nucleophilic trifluoromethylation proceeds as follows: \text{R-X} + \text{TMSCF}_3 + \text{F}^- \rightarrow \text{R-CF}_3 + \text{TMSF} + \text{X}^- This methodology expanded the scope to aldehydes, ketones, and other electrophiles, facilitating broader synthetic applications. During the late 1980s, radical trifluoromethylation gained traction with the introduction of sodium trifluoromethanesulfinate (CF₃SO₂Na), known as the Langlois reagent. Developed by Bernard R. Langlois in 1988, it served as a convenient precursor to the CF₃ radical under reductive conditions, such as with copper salts or dithionite, allowing trifluoromethylation of electron-deficient alkenes and aromatic systems. This reagent's mild conditions and compatibility with aqueous media made it particularly valuable for late-stage functionalization in complex molecules. Concurrently, Teruo Umemoto pioneered electrophilic trifluoromethylation using hypervalent iodine compounds in the mid-1980s. In 1986, he reported bis(trifluoromethyl)phenyl-λ³-iodane [(CF₃)₂IPh], a stable electrophilic CF₃ source capable of transferring the CF₃ group to nucleophiles like enolates and aromatics under acidic conditions. This reagent, part of a series of iodine-based transfer agents, addressed limitations of gaseous or unstable CF₃ sources, enabling selective C-CF₃ bond formation in pharmaceuticals and agrochemicals. The 1990s saw the emergence of transition metal-catalyzed trifluoromethylation, with systems enabling cross-couplings of aryl halides with CF₃ sources. Early efforts explored Pd- and Cu-mediated reactions using CF₃I or related donors, achieving modest yields, with truly catalytic protocols emerging in the early (e.g., Kotora et al. in 2001). These developments laid the groundwork for more efficient protocols in the . In the , Antonio Togni introduced benziodoxolone-based reagents in 2006, offering bench-stable electrophilic CF₃ donors with improved reactivity over prior hypervalent iodine systems. The first Togni reagent, 1-(trifluoromethyl)-3H-benzo[1,2]iodoxol-3-one, facilitated trifluoromethylation of silyl ethers, amines, and thiols under mild conditions, with high tolerance. The 2010s brought as a transformative approach, harnessing to generate CF₃ radicals from reagents like Togni or Umemoto types. Pioneering work by David A. Nagib and David W. C. MacMillan in 2011 demonstrated Ru(bpy)₃²⁺-catalyzed trifluoromethylation of heteroarenes and pharmaceuticals using CF₃SO₂Cl, achieving site-selective C-H functionalization without directing groups. This method's green credentials and orthogonality to thermal processes accelerated adoption in . Subsequent advances in the late 2010s included more robust copper-catalyzed systems tolerant of diverse functional groups. In 2015, Stephen L. Buchwald reported an efficient Cu-catalyzed trifluoromethylation of aryl and heteroaryl chlorides using a phenanthroline and (phen)CuCF₃, expanding accessibility to less reactive substrates. Nickel-catalyzed variants emerged around , offering cost-effective alternatives for C(sp²)-CF₃ bond formation. By 2020, electrochemical trifluoromethylation methods had gained traction, enabling milder conditions and broader substrate scope without sacrificial metals.

Trifluoromethylating Reagents

Nucleophilic and Ruppert-Prakash Type Reagents

The Ruppert-Prakash reagent, trimethyl(trifluoromethyl)silane (TMSCF₃), serves as a cornerstone for nucleophilic trifluoromethylation reactions due to its stability and ease of handling as a source of the CF₃ nucleophile. It is synthesized by reacting chlorotrimethylsilane with the Grignard reagent prepared from bromotrifluoromethane and magnesium in diethyl ether, followed by distillation to isolate the product in moderate yield. This method, first reported in 1984, provides a scalable route to TMSCF₃, which is commercially available as a colorless liquid. The reagent requires activation by fluoride sources to generate the reactive CF₃ species. Initial applications employed catalytic () in the presence of crown ethers for the addition to aldehydes and ketones, affording trifluoromethylated alcohols in good yields after acidic workup. Subsequent developments utilized tetrabutylammonium fluoride (TBAF) or cesium fluoride (CsF) as initiators, enabling milder conditions (often at in THF or DMF) and expanded compatibility, including non-activated ketones and imines. With imines, TMSCF₃ delivers the CF₃ group to produce α-trifluoromethyl amines, particularly effective for N-sulfinyl or N-phosphinoyl protected variants, yielding products with high diastereoselectivity in some cases. Beyond TMSCF₃, other silicon-based analogs such as triethyl(trifluoromethyl)silane or triisopropyl(trifluoromethyl)silane have been explored, offering similar nucleophilic behavior but with variations in and reactivity tuned by the silyl substituents. Sodium trifluoromethyl (NaCF₃), while a direct CF₃ anion source, is highly unstable and prone to decomposition via α-fluoride elimination, limiting its use to generation attempts that have largely been supplanted by silicon reagents. (CHF₃), an inexpensive and abundant gas, provides an alternative through with strong bases like tert-butoxide in DMF, generating a CF₃ equivalent for addition to carbonyls under mild conditions. The mechanism of TMSCF₃-mediated trifluoromethylation involves fluoride-initiated formation of a pentacoordinate siliconate , [Me₃SiF(CF₃)]⁻, which equilibrates to release free CF₃⁻; this anion then adds to the carbonyl or to form the R₂C(O⁻)CF₃. The coordinates to another TMSCF₃ , leading to O-silylation and release of the product as R₂C(OSiMe₃)CF₃, which upon yields the trifluoromethylated R₂C(OH)CF₃. This process can be represented as: \ce{R2C=O + TMSCF3 + F- ->[addition] R2C(O-)CF3 + Me3SiF} \\ \ce{R2C(O-)CF3 + TMSCF3 -> R2C(OSiMe3)CF3 + F-} \\ \ce{R2C(OSiMe3)CF3 + H2O -> R2C(OH)CF3 + Me3SiOH} These reactions proceed under mild conditions (typically 0–25 °C), exhibit broad functional group tolerance including esters and nitro groups, and avoid harsh reagents, making them suitable for complex molecule synthesis. However, challenges include the generation of silicon-containing byproducts like hexamethyldisiloxane, which require careful purification, and potential side reactions such as α-elimination under basic conditions.

Radical and Togni/Umemoto Reagents

The Langlois reagent, sodium trifluoromethanesulfinate (CF₃SO₂Na), serves as a widely used source for generating the (CF₃•) in trifluoromethylation reactions. This bench-stable, white solid is typically prepared by reducing trifluoromethanesulfonyl chloride (CF₃SO₂Cl) with aqueous , followed by precipitation and purification, although alternative routes involve the reaction of (CF₃Br) with or in . Upon activation under oxidative conditions, such as with (K₂S₂O₈) or tert-butyl , the undergoes single-electron transfer (SET) to form the CF₃SO₂• , which fragments to release CF₃• and (SO₂). The process can be represented as: \text{CF}_3\text{SO}_2\text{Na} + \text{SO}_4^{\bullet-} \rightarrow \text{CF}_3\text{SO}_2^{\bullet} + \text{SO}_4^{2-} \quad \rightarrow \quad \text{CF}_3^{\bullet} + \text{SO}_2 Metals like copper or iron salts can also facilitate this reduction, enabling efficient CF₃• addition to alkenes, arenes, and heteroarenes. Togni reagents, developed in the 2000s and 2010s, are hypervalent iodine(III) compounds based on benziodoxole scaffolds that function as precursors for CF₃• via SET processes in radical trifluoromethylation. The first-generation reagent, 1-(trifluoromethyl)-3,3-dimethyl-1,2-benziodoxole, features a pseudobinary I–CF₃ bond and is synthesized by treating 3,3-dimethyl-1,2-benziodoxole with trifluoromethyl iodide (CF₃I) in the presence of a base like silver sulfate. A second-generation variant, 1-(trifluoromethyl)-1,2-benziodoxol-3(1H)-one, incorporates a carbonyl group for enhanced stability and is prepared similarly from the parent 1-hydroxy-1,2-benziodoxol-3(1H)-one and CF₃I under basic conditions. In radical pathways, these reagents undergo SET reduction (often photocatalyzed or metal-mediated), cleaving the I–O or I–CF₃ bond to generate CF₃• while forming a stable benziodoxole byproduct; this approach has enabled trifluoromethylation of a broad range of substrates, including unactivated alkenes and C–H bonds. Umemoto reagents, introduced in the , comprise a family of electrophilic trifluoromethylating agents, including diaryliodonium bis(trifluoromethyl) s and their analogs, which can generate CF₃• through SET in contexts. The core structure, such as (trifluoromethyl)phenyliodonium bis(trifluoroacetate), is prepared by reacting iodobenzene diacetate with and CF₃I, yielding a where the CF₃ group is bound to iodine in a hypervalent . Variants like S-(trifluoromethyl)dibenzothiophenium extend this to sulfur-based systems, synthesized from dibenzothiophene and CF₃I with silver mediation. Under reductive conditions, such as with catalysts or photoredox systems, these reagents release CF₃• via homolytic cleavage, facilitating additions to ates and alkenes; a notable example is the trifluoromethylation of silyl ethers in up to 90% yield. Additional sources for CF₃ generation include trifluoromethyl iodide (CF₃I), which undergoes photolysis or copper-mediated homolysis to produce CF₃•, often applied in Kharasch-type to alkenes. Trifluoromethyl phenyl (CF₃SO₂Ph), a byproduct from Julia-Kocienski olefination, can also serve as a CF₃• precursor when reduced with metals or under photochemical conditions, though it is less common due to byproduct formation.

Electrophilic and Metal-Based

trifluoromethylation involves the use of that deliver the CF₃ group as an (CF₃⁺ equivalent), typically targeting electron-rich substrates such as enolates, aromatics, or heteroarenes. These contrast with nucleophilic sources by enabling to electron-donating centers, often under mild conditions to avoid side reactions. Seminal developments include hypervalent species and sulfonyl derivatives, though the latter can also generate radicals depending on conditions. Key examples include trifluoromethyl fluoroborate (CF₃BF₃) and triflyl (CF₃SO₂Cl), which have been employed in direct C–CF₃ bond formation. Triflyl chloride (CF₃SO₂Cl) serves as a versatile electrophilic CF₃ source, particularly in copper-mediated reactions for aryl trifluoromethylation. In a Sandmeyer-type , arenediazonium tetrafluoroborates react with CF₃SO₂Na (Langlois ) in the presence of CuBF₄·(MeCN)₄ and 2,2′;6′,2″-terpyridine ligand under mild conditions (room temperature to 50 °C in ) with tert-butyl as oxidant, affording trifluoromethylarenes in moderate to high yields (up to 85% for electron-rich substrates like 4-methoxyphenyl). This method tolerates functional groups such as esters, nitriles, ketones, nitro, and bromo, with a one-pot variant from anilines via diazotization. The reaction proceeds via copper-mediated single-electron transfer to generate the CF₃ from the sulfinate, which then reacts with the diazonium . Sodium trifluoroacetate (CF₃CO₂Na) functions as a decarboxylative CF₃ source, applicable in electrophilic contexts through oxidative to generate CF₃⁺ equivalents, particularly with Pd and directing groups. For instance, Pd(II)-catalyzed ortho-trifluoromethylation of arenes uses CF₃CO₂Na (or its acid form) as the CF₃ donor, achieving selective C–H in yields of 60–90% for substrates like 2-phenylpyridine. Under reductive or photocatalytic conditions, it can instead produce the CF₃ radical via : \ce{CF3CO2Na ->[reductive\ or\ photo] CF3^\bullet + CO2 + Na+} This dual behavior highlights its utility, though electrophilic applications prioritize oxidative setups to favor CF₃⁺ transfer. Metal-based reagents provide stable CF₃ sources for electrophilic delivery, often prepared via transmetallation from nucleophilic precursors like Ruppert–Prakash reagent (TMSCF₃). Trifluoromethylcopper (CF₃Cu) species, generated by treating CuI with TMSCF₃ and KF, act as electrophilic agents in couplings with aryl halides, yielding Ar–CF₃ in 70–95% efficiency under ligand-free conditions at 50–100 °C. Similarly, trifluoromethylzinc halides (CF₃ZnX) form via Zn insertion into CF₃I or transmetallation from Grignard reagents, enabling Negishi-type reactions with electrophiles like acyl chlorides to produce trifluoromethyl ketones (yields >80%). Trifluoromethylboronic acids (CF₃B(OH)₂) or their pinacol esters, prepared by transmetallation from CF₃Cu with B(OH)₃, serve in Suzuki–Miyaura couplings with aryl bromides using Pd catalysts, providing biaryl CF₃ products in 50–90% yields while tolerating heterocycles. These organometallics enhance selectivity for electrophilic CF₃ transfer compared to free CF₃⁺, minimizing over-fluorination. Other electrophilic variants include CF₃⁺ from CF₃BF₃, a stable salt developed by Olah et al. for direct addition to π-nucleophiles like styrene or anisole, forming CF₃-adducts in 40–70% yields under superacid conditions. This reagent underscores early efforts in isolated CF₃⁺ chemistry, influencing subsequent metal-mediated adaptations.

Reaction Mechanisms

Radical Trifluoromethylation

Radical trifluoromethylation proceeds via a free-radical chain mechanism involving the electrophilic trifluoromethyl radical (•CF₃), which is highly reactive due to its electron-deficient nature. The process is governed by single-electron transfers and hydrogen atom abstractions, with reactivity influenced by SOMO-LUMO interactions between the singly occupied molecular orbital (SOMO) of •CF₃ and the lowest unoccupied molecular orbital (LUMO) of electron-rich substrates, facilitating selective addition. Initiation typically involves the generation of •CF₃ through single-electron transfer (SET) from a reductant to a trifluoromethylating agent, such as sodium trifluoromethanesulfinate (CF₃SO₂Na, Langlois reagent). In , developed prominently in the 2010s, visible light excites or polypyridyl complexes (e.g., [Ru(bpy)₃]²⁺ or [Ir(ppy)₂(dtbbpy)]⁺), enabling reductive quenching to produce •CF₃ from the reagent while regenerating the catalyst. Alternative initiations include electrochemical SET at the or thermal decomposition of like di-tert-butyl peroxide to homolyze and generate radicals that abstract from the CF₃ source. Propagation ensues with •CF₃ addition to unsaturated bonds or hydrogen abstraction, forming substrate-derived radicals that perpetuate the chain by regenerating •CF₃ or transferring electrons/hydrogen. Termination occurs via radical-radical coupling or disproportionate, yielding stable products and halting the chain. The scope encompasses addition to alkenes, arenes, and alkyl C-H bonds. For alkenes, •CF₃ adds regioselectively to the terminal carbon in an anti-Markovnikov manner, as seen in Giese-type hydrotrifluoromethylations of styrenes, affording β-CF₃ alkyl products in yields up to 95% under photoredox conditions. Arene trifluoromethylation often follows a Minisci pathway for heteroarenes, where •CF₃ adds to the protonated ring, forming a σ-complex that deprotonates to the C-H functionalized product, achieving site-selective functionalization in 70-90% yields for quinolines. For unactivated aryl C-H, •CF₃ adds to the arene ring, forming a cyclohexadienyl (σ-complex), which undergoes single-electron oxidation to the corresponding cation followed by to afford Ar-CF₃. Selectivity challenges include preventing over-trifluoromethylation from multiple •CF₃ additions, which is mitigated by using substoichiometric initiators and mild conditions to limit lifetimes. Difunctionalization strategies enhance utility by trapping the substrate with a , such as in photoredox-mediated carbotrifluoromethylation of alkenes with arylboronic acids, delivering 1,2-bifunctionalized products (e.g., Ph-CH(CF₃)-CH₂-Ar) in 85% yield without over-addition. These approaches prioritize unactivated substrates, distinguishing pathways from polarity-matched nucleophilic or electrophilic processes.00649-5.pdf)

Nucleophilic Trifluoromethylation

Nucleophilic trifluoromethylation involves the two-electron addition of a trifluoromethyl anion equivalent (CF₃⁻) to electron-deficient centers, such as carbonyls and iminiums, typically facilitated by the Ruppert-Prakash reagent, trimethyl(trifluoromethyl)silane (TMSCF₃). The mechanism proceeds via activation of TMSCF₃ by a Lewis base initiator, such as fluoride sources (e.g., KF or tetrabutylammonium fluoride, TBAF), which generates a pentacoordinate . This leads to the liberation of the CF₃⁻ , which attacks the electrophilic carbon, forming an adduct. Subsequent yields the trifluoromethylated , while the silyl group is transferred to form a or siloxide, enabling an autocatalytic cycle where the product alkoxide further activates TMSCF₃. This strategy inverts the reactivity of the CF₃ group from electrophilic to nucleophilic, overcoming the instability of direct CF₃⁻ sources like CF₃Li. The reaction scope encompasses aldehydes, ketones, and α,β-unsaturated carbonyl systems, with the latter undergoing conjugate (1,4-) addition. For aldehydes and ketones, TMSCF₃ adds efficiently to yield trifluoromethyl carbinols in high yields (typically 80–99% for aromatic substrates), as demonstrated by the representative equation for an aldehyde: \text{RCHO} + \text{CF}_3\text{SiMe}_3 + \text{KF} \rightarrow \text{RCH(OH)CF}_3 + \text{KOSiMe}_3 This process extends to iminiums for the synthesis of trifluoromethylated amines, though it requires acidic conditions to enhance electrophilicity. In α,β-unsaturated systems, such as enones or acrylates bearing electron-withdrawing groups, selective 1,4-addition predominates, providing β-trifluoromethylated products in 70–95% yields. The method is particularly effective for non-hindered, electron-deficient substrates, with seminal reports establishing its broad utility since the 1980s. Catalysis often employs Lewis bases beyond simple fluorides, including N-heterocyclic carbenes (NHCs) at 0.5 mol% loadings in DMF for enhanced rates, or phosphines like P(t-Bu)₃ (10 mol%) to promote enantioselective variants with up to 98% using chiral auxiliaries. Solvent choice significantly influences outcomes: DMF accelerates the reaction and improves yields compared to THF due to better of ionic intermediates, though THF is preferred for kinetic studies or moisture-sensitive setups. These optimizations have enabled applications in complex molecule synthesis. Despite its versatility, nucleophilic trifluoromethylation with TMSCF₃ exhibits limitations, including high sensitivity to moisture, which triggers rapid decomposition to CF₃H and siloxanes, necessitating anhydrous conditions and molecular sieves. Additionally, sterically hindered substrates, such as ortho-substituted ketones or bulky iminiums, react sluggishly or require longer trialkylsilyl analogs like TIPS CF₃, often resulting in lower conversions (e.g., 60% after 16 hours). These constraints highlight the need for careful substrate selection.

Electrophilic Trifluoromethylation

Electrophilic trifluoromethylation involves the direct transfer of an electrophilic trifluoromethyl cation (CF₃⁺) to electron-rich substrates, such as enolates, enol ethers, and activated aromatics, enabling the introduction of the CF₃ group under polar, two-electron mechanisms. This approach contrasts with or nucleophilic pathways by relying on hypervalent iodine or sulfonium-based that mimic classical electrophiles like in Friedel-Crafts alkylations. Developed primarily since the , these methods have gained prominence for their ability to functionalize π-systems without requiring metal in many cases, though activation is often necessary. The mechanism proceeds via activation of the electrophilic reagent to generate a reactive CF₃⁺ equivalent, followed by nucleophilic attack from the substrate and subsequent deprotonation or loss of a leaving group to restore aromaticity or planarity. For hypervalent iodine reagents like Togni's, protonation or Lewis acid coordination weakens the I–O bond, promoting CF₃⁺ transfer through a reductive elimination pathway; sulfonium salts, such as Umemoto's diaryltrifluoromethylsulfonium triflates, undergo direct SN2-like displacement at the CF₃ carbon, expelling the sulfonium leaving group. A representative reaction with aromatics illustrates this: electron-rich ArH reacts with (CF₃)₂IPh to afford ArCF₃ + PhI + CF₃H, involving initial ipso attack and rearomatization via deprotonation. Computational studies confirm that the CF₃ group transfers as a "closed-shell" electrophile, with barriers lowered by acidic activation to avoid competing homolytic cleavage. The scope encompasses electron-rich aromatics via Friedel-Crafts-type processes, silyl enol ethers to yield α-trifluoromethyl carbonyls, and heterocycles like indoles at the C2 position. For instance, indoles undergo regioselective C2-trifluoromethylation with Togni's II in the presence of trimethylsilyl triflate, affording products in 63–85% yield; silyl enol ethers of ketones react cleanly at to give β-trifluoromethylated adducts in up to 96% yield. Aromatics such as derivatives tolerate the reaction under acid promotion, though yields vary with (e.g., 42–67% for activated systems). reagents excel with enolates, delivering CF₃ to β-keto esters in 79% yield without additives. Togni's II, a shelf-stable benziodoxole derivative, is particularly versatile due to its thermal stability and ease of handling, avoiding the explosiveness of earlier salts. Typical conditions involve mild acidic or acidic promotion to enhance electrophilicity, such as HNTf₂, Zn(NTf₂)₂, or TMSOTf in at to 50 °C, minimizing side reactions. These activators facilitate CF₃⁺ release without high temperatures, and reactions often proceed in minutes to hours for enolates and enol ethers. Shelf-stable electrophiles like Togni II enable benchtop operations, with no need for inert atmospheres in many protocols. Challenges include competing protonation of substrates, which can divert electron-rich systems toward non-productive paths, particularly with hard nucleophiles or insufficient activation, leading to lower CF₃ incorporation. Regioselectivity poses issues in unsymmetrical heterocycles and aromatics, where mixtures (e.g., 2.5:1 ortho/para) arise without directing groups, necessitating optimized conditions or catalysts for indoles to favor C2 over C3. Despite these, the method's selectivity for electron-rich sites makes it complementary to other trifluoromethylation strategies. \begin{align*} &\ce{ArH + (CF3)2IPh ->[activation] ArCF3 + PhI + CF3H} \end{align*}

Catalytic and Specialized Methods

Transition Metal-Catalyzed Couplings

Transition metal-catalyzed couplings have emerged as a powerful strategy for constructing aryl- and alkyl-CF₃ bonds through the activation of C-X bonds, where X is typically a or diazonium group. These methods leverage the ability of metals like and to facilitate , transmetallation, and steps, enabling selective trifluoromethylation under controlled conditions. Early challenges, such as the instability of CF₃-metal intermediates and competing β-fluoride elimination, have been addressed through optimized ligands and CF₃ sources, allowing broad substrate scope including electron-rich and sterically hindered aryl systems. The general mechanism for palladium-catalyzed trifluoromethylation of aryl halides involves of Ar-X (X = I, Br, Cl) to Pd(0), forming an Ar-Pd(II)-X species, followed by transmetallation with a CF₃ source such as TMSCF₃ in the presence of or a boronate like CF₃B(OH)₂. then yields the Ar-CF₃ product and regenerates Pd(0). For example, Buchwald and coworkers reported a Pd-catalyzed process using (Me₂NCH₂CH₂)₂PdCl₂ (1-3 mol%) with TMSCF₃ and , achieving yields of 70-99% for diverse aryl chlorides at 80-110°C, including heteroaryl and sterically demanding substrates. This approach expanded access to Ar-CF₃ from less reactive chlorides, previously limited to iodides or bromides. Similarly, Suzuki-Miyaura-type couplings with CF₃Bpin or CF₃BF₃K have been developed; Molander et al. demonstrated Pd₂(dba)₃ (2 mol%) with XPhos coupling aryl/heteroaryl bromides and CF₃BF₃K, furnishing Ar-CF₃ in 60-95% yields under aqueous conditions at 80°C. The reaction can be represented as: \text{Ar-Br} + \text{CF}_3\text{B(OH)}_2 \xrightarrow{\text{Pd cat., base}} \text{Ar-CF}_3 + \text{Br-B(OH)}_2 Copper catalysis often proceeds via analogous steps but with lower oxidation states, particularly for Sandmeyer-type reactions. In these, Cu(I) promotes the trifluoromethylation of arenediazonium salts derived from aromatic amines. A seminal method by Li et al. uses CuI (10 mol%) and CF₃SO₂Na (Langlois reagent) at 50°C in DMSO, converting ArNH₂ to Ar-CF₃ in 50-96% yields over two steps, tolerant of nitro, carbonyl, and halide groups. The mechanism likely involves Cu(I)-mediated reduction of CF₃SO₂Na to Cu-CF₃, followed by single-electron transfer to the diazonium and recombination. For electrophilic sources like CF₃SO₂Cl, CuI (20 mol%) with phenanthroline ligands enables trifluoromethylation of aryl iodides at 100°C, yielding 70-90% Ar-CF₃ via in situ generation of Cu-CF₃. Negishi-type couplings utilize organozinc reagents for efficient transmetallation. Although direct CF₃ZnX is unstable, in situ generation from Zn and CF₃I with Pd or Cu catalysis facilitates aryl iodide trifluoromethylation. For instance, Pd(PPh₃)₄ (5 mol%) couples ArI with CF₃I/Zn in DMF at 60°C, providing Ar-CF₃ in 60-85% yields, with the mechanism involving oxidative addition, Zn-mediated CF₃ transfer, and reductive elimination. Copper variants, such as CuI (10 mol%) with bipyridine, extend this to heteroaryl halides. These methods highlight the versatility of Zn for clean CF₃ delivery in cross-couplings. Catalysts typically include Pd(0) precursors like Pd₂(dba)₃ or Pd(dppf)Cl₂ for activations, paired with bulky ligands such as , P(t-Bu)₃, or Buchwald's biarylphosphines to enhance and suppress protodemetallation. For copper systems, bidentate N-ligands like or TMEDA stabilize Cu-CF₃ intermediates and improve selectivity. Pd(II) catalysts also enable direct C-H trifluoromethylation of arenes; for example, Pd(OAc)₂ (10 %) with phenanthroline and CF₃SO₂Na at 120°C trifluoromethylates 2-phenylpyridine at the position in 80% yield, proceeding via Pd(II)/Pd(IV) cycles with directing-group assistance. Advances in the include decarboxylative couplings using CF₃CO₂Na as an inexpensive CF₃ source. Goossen et al. developed a CuI (20 mol%)-catalyzed decarboxylative trifluoromethylation of aryl iodides with CF₃CO₂Na and K₂CO₃ at 160°C in NMP, yielding Ar-CF₃ in 50-90%, where decarboxylation generates Cu-CF₃ for subsequent and elimination. This method avoids preformed CF₃ metals and accommodates electron-deficient aryls, marking a shift toward sustainable in protocols.

Photocatalytic and Light-Driven Approaches

Photocatalytic trifluoromethylation leverages visible light to activate trifluoromethylating reagents through , enabling the formation of under mild conditions. These methods typically involve complexes such as Ru(bpy)₃Cl₂ or Ir(ppy)₃ as photocatalysts, which undergo photoexcitation to facilitate single-electron transfer (SET) processes. Upon visible light irradiation, the excited photocatalyst donates an electron to the trifluoromethyl source, generating a that adds to substrates, followed by subsequent oxidation and steps to afford the trifluoromethylated product. An alternative mechanism employs electron donor-acceptor (EDA) complexes, where ground-state charge-transfer interactions between an electron-rich donor (e.g., an or ) and the electrophilic CF₃ reagent (e.g., Togni's reagent) form a photoactive complex; visible light excitation then induces SET to produce the without requiring a metal catalyst. A landmark development occurred in 2011 when Nagib and MacMillan reported the direct C-H trifluoromethylation of unactivated arenes and heteroarenes using CF₃SO₂Cl as the reagent, Ru(bpy)₃Cl₂ as the photocatalyst, and a household light bulb as the visible light source. The reaction proceeds via SET from the photoexcited Ru(II)* to CF₃SO₂Cl, yielding the CF₃ radical, which undergoes from the arene C-H bond, followed by re-aromatization. Representative substrates include electron-rich arenes like (yielding 4-(trifluoromethyl) in good yield) and heteroarenes such as indoles and pyrroles, demonstrating broad scope with operational simplicity at . The general transformation can be represented as: \text{R-H} + \text{CF}_3\text{SO}_2\text{Cl} \xrightarrow{\text{h}\nu / \text{Ru(bpy)}_3\text{Cl}_2} \text{R-CF}_3 + \text{SO}_2 + \text{HCl} This approach marked a shift toward radical-mediated C-H activation, avoiding harsh conditions typical of earlier thermal methods. In the 2020s, metal-free variants using organic dyes and EDA complexes have gained prominence for their cost-effectiveness and reduced toxicity. For instance, in 2021, Wang and coworkers introduced a general organocatalytic system employing dithiocarbamate or xanthate salts as donors to form EDA complexes with Togni's reagent, enabling visible-light-driven trifluoromethylation of heteroarenes and silyl enol ethers with yields up to 95%. These systems extend to alkene difunctionalization, such as the hydrotrifluoromethylation of styrenes, where the CF₃ radical adds across the double bond, followed by protonation. Scope includes unactivated alkanes via selective C-H abstraction, though primarily for benzylic or allylic positions, and late-stage modification of pharmaceuticals like naproxen derivatives. These light-driven approaches offer significant advantages, including mild reaction conditions (, aqueous or organic solvents), high site-selectivity for C-H bonds, and compatibility with complex molecules for late-stage functionalization in . For example, direct trifluoromethylation of the cetirizine enhances its metabolic stability without protecting groups. Compared to traditional methods, minimizes byproduct formation and energy input, promoting greener synthesis while maintaining efficiency across diverse substrates like five- and six-membered heterocycles.

Asymmetric Trifluoromethylation

Asymmetric trifluoromethylation encompasses enantioselective methods that introduce the CF₃ group into prochiral substrates to generate stereogenic centers, primarily through nucleophilic and pathways using chiral catalysts. Chiral bases, such as derivatives, activate nucleophilic reagents like TMSCF₃ for addition to carbonyl and acceptors, often in combination with sources to generate chiral initiators. These organocatalytic approaches achieve high enantioselectivities, typically exceeding 90% , by forming transient chiral pairs that direct stereoselective CF₃ transfer. Complementarily, catalysts with chiral ligands, such as bisoxazoline () complexes of , enable -based trifluoromethylation, where the metal coordinates the radical intermediate to enforce asymmetry. The scope of these methods includes the conversion of prochiral ketones to trifluoromethylated tertiary alcohols and aldehydes to secondary alcohols bearing a CF₃ group at the carbinol carbon. For instance, aromatic aldehydes undergo nucleophilic trifluoromethylation under cinchona/TMAF catalysis to afford enantioenriched products. \mathrm{RCHO + TMSCF_3 + chiral\ F^- \rightarrow (R)-\mathrm{RCH(OH)CF_3} Imines are also viable electrophiles, yielding α-trifluoromethylated amines with utility in pharmaceutical synthesis. Enantioselectivities are quantified via chiral HPLC, with absolute configurations often confirmed by X-ray crystallography of derivatized products. Seminal work by Shibata and coworkers in the 2010s established cinchona alkaloid ammonium salts with TMAF as effective catalysts for nucleophilic trifluoromethylation. In 2007, aryl ketones were trifluoromethylated to tetrasubstituted alcohols in up to 94% using 5 % quinine-derived bromide and TMSCF₃ in THF at . This protocol extended to cyclic ketones with 82–92% . A 2010 advancement applied bis-cinchona catalysts to alkynyl ketones, delivering propargylic alcohols in 88–96% , enabling access to bioactive motifs. These methods highlight the role of the chiral in modulating the nucleophilic attack via steric differentiation. For radical processes, chiral copper catalysis with ligands has been pivotal. In 2017, and colleagues reported an intermolecular trifluoromethylarylation of styrenes using CuI, (S)-Ph-, CF₃I, and arylboronic acids, generating 1,1-diarylmethanes with benzylic CF₃ in 80–95% yield and 82–92% via enantioselective arylation of transient benzylic radicals. The ligand's rigidity ensures high enantiocontrol in the radical capture step. In the 2020s, advancements include dual-catalytic strategies for allylic trifluoromethylation. A 2022 nickel-catalyzed reductive trifluoroalkylation of alkenyl halides with CF₃Br, using NiBr₂•DME and a chiral 3,5-dimethylphenyl-oxazoline ligand (13 mol%), produced α-trifluoromethylated allylic amides in 70–99% yield and up to 97% at low temperatures (-4 °C) in THF/DMP solvent with as reductant. This method constructs allylic stereocenters with broad substrate tolerance, including late-stage modification of drugs like ibuprofen (95% ).

Advanced Topics

Trifluoromethyl Cation Chemistry

The trifluoromethyl cation (CF₃⁺) is a highly reactive and elusive electrophilic species characterized by its planar structure and significant instability in condensed phases due to the strong electron-withdrawing effects of the three atoms, which destabilize the positive charge on the central carbon. Despite extensive efforts, CF₃⁺ cannot be stabilized in media such as HSO₃F/SbF₅ () or with common Lewis acids like SbF₅, as experimental and theoretical studies indicate that no known counterions or solvents can sufficiently isolate it without decomposition. Its generation has primarily been achieved in the gas phase through of precursors like CF₃Cl, CF₃Br, or C₂F₄, where ion yield curves confirm its existence with a heat of formation of approximately 98 kcal/mol at 298 K, reflecting its high energy content relative to typical alkyl carbocations. In gas-phase studies, CF₃⁺ exhibits potent electrophilic reactivity, particularly toward π-systems like s, where it undergoes initial addition to form a β-trifluoromethyl intermediate. For instance, the reaction with ethene (CH₂=CH₂) proceeds via nucleophilic attack by the on CF₃⁺, yielding an excited CF₃CH₂CH₂⁺ that rapidly rearranges. This instability drives competing pathways: elimination of to form C₃H₃F₂⁺ (branching ratio 60%) or hydride leading to C₂H₃⁺ + CHF₃ (branching ratio 40%), with an overall reaction efficiency of 0.64 at 298 K. The observed HF loss highlights the cation's tendency to favor fluorination cascades over stable trifluoromethylation products, often resulting in difluorinated or rearranged alkenyl cations rather than direct CF₃ incorporation. Similar addition-elimination behavior is noted with other small s, underscoring CF₃⁺'s role as a transient prone to fragmentation. Computational analyses further elucidate CF₃⁺'s instability, revealing that while it benefits from partial hyperconjugative stabilization from fluorine lone pairs, the inductive withdrawal dominates, rendering it less stable than less fluorinated analogs like the difluoromethyl cation (CHF₂⁺) or protonated trifluoromethane (CF₃H⁺). calculations show that the reaction CH₃⁺ + CHF₃ → CF₃⁺ + CH₄ is slightly exothermic by 3.5 kcal/mol, indicating modest stabilization relative to the methyl cation (CH₃⁺), but overall, CF₃⁺ remains a high-energy unsuitable for . These insights from gas-phase experiments and quantum chemical modeling have limited practical applications for CF₃⁺ to mechanistic investigations in , where it serves as a precursor for probing organic reactivity, and rare fluorination cascades in specialized environments.

Recent Advances (2020-2025)

Recent advances in trifluoromethylation have emphasized sustainable, mild, and stereoselective methods, leveraging photocatalysis, biocatalysis, and mechanochemistry to address limitations in traditional approaches. A key development involves photocatalytic decarboxylative trifluoromethylation using trifluoroacetic acid (TFA) and its derivatives, which enables efficient C-CF₃ bond formation under visible light without harsh reagents. For instance, reviews from 2023–2025 highlight systems employing electron donor-acceptor (EDA) complexes or ligand-to-metal charge transfer (LMCT) mechanisms with TFA salts, achieving high yields in arene and alkene functionalizations. A notable example utilizes blue LED irradiation with quinuclidine as a hydrogen atom transfer (HAT) catalyst to generate alkyl-CF₃ products from aliphatic precursors, demonstrating broad substrate tolerance and scalability. Biocatalytic strategies have emerged as transformative for enantioselective hydrotrifluoromethylation, particularly using engineered flavin-dependent enzymes. In 2025, researchers developed ground-state flavin-dependent enzymes, such as variants of old yellow enzyme 1 (OYE1-N251M), to catalyze radical-mediated addition of CF₃ groups to alkenes with trifluoromethyl thianthrenium salts as donors. This approach yields anti-Markovnikov CF₃ adducts on styrenes with exceptional stereocontrol, achieving up to 98:2 enantiomeric ratios (e.r.) and >95% in many cases, with yields exceeding 90% under aqueous conditions. The reaction proceeds as follows: \text{Alkene} + \text{CF}_3\text{–source (via enzyme)} \rightarrow \text{anti-Markovnikov CF}_3 \text{ adduct} This method highlights the integration of enzyme engineering with radical chemistry for precise fluorination in complex molecules. Mechanochemical innovations, including piezoelectric radical generation, offer green alternatives by avoiding solvents and external energy inputs. A 2025 study demonstrated ball-milling with piezoelectric materials (e.g., BaTiO₃) to generate CF₃• radicals from hypervalent iodine reagents like Togni II under mechanical stress, enabling C(sp²)–H trifluoromethylation of enamides and acrylamides with up to 88% yields and full stereoretention. This solvent-free process serves as a sustainable counterpart to traditional photolytic or thermal activations, minimizing waste while maintaining efficiency. Although earlier work (2020) explored CF₃I with piezoelectrics for arene C-H trifluoromethylation, recent adaptations emphasize broader applicability. Additional breakthroughs include light-driven difunctionalizations for polyfluorinated motifs. Complementing this, 2025 reviews detail metallaphotoredox strategies for synthesizing trifluoromethylated aliphatic amines, advancing access to bioactive fluorinated heterocycles. These developments underscore a shift toward eco-friendly, selective trifluoromethylation for pharmaceutical and materials applications.

References

  1. [1]
    New Electrophilic Trifluoromethylating Agents - ACS Publications
    As a consequence the introduction of a trifluoromethyl group into organic molecules often changes their physical, chemical, and physiological properties.
  2. [2]
    Di- and Trifluoromethyl(thiol)ations - Ruhr-Universität Bochum
    Aug 10, 2021 · Trifluoromethyl-substituted compounds often display an improved metabolic stability, a better receptor binding selectivity, increased ...<|control11|><|separator|>
  3. [3]
    Recent advances in trifluoromethylation of organic compounds ...
    This review highlights recent developments in the direct introduction of a trifluoromethyl group into organic compounds with Umemoto's reagents.
  4. [4]
  5. [5]
    Common Bond Energies (D - Wired Chemist
    Common Bond Energies (D. ) and Bond Lengths (r). Hydrogen. Bond, D (kJ/mol), r (pm). H-H, 432, 74. H-B, 389, 119. H-C, 411, 109. H-Si, 318, 148. H-Ge, 288, 153.
  6. [6]
    Contribution of Organofluorine Compounds to Pharmaceuticals
    In recent years, an estimated 20% of the marketed drugs have been fluoro-pharmaceuticals.Introduction · Chemotypes of the Fluoro... · Chiral Fluoro-Pharmaceuticals
  7. [7]
    The Role of Trifluoromethyl and Trifluoromethoxy Groups in ...
    Jul 18, 2025 · In addition, the trifluoromethyl group's unique combination of high electronegativity, electron-withdrawing effect, and lipophilicity makes it ...Missing: Hammett | Show results with:Hammett
  8. [8]
    FDA-Approved Trifluoromethyl Group-Containing Drugs - MDPI
    Trifluoromethylation refers to any chemical process that adds a TFM group to an organic product. The product was then purified from the insoluble crude product ...<|control11|><|separator|>
  9. [9]
    Celecoxib: Uses, Interactions, Mechanism of Action | DrugBank Online
    Celecoxib is an NSAID used to treat osteoarthritis, rheumatoid arthritis, acute pain, menstrual symptoms, and to reduce polyps is familial adenomatous ...
  10. [10]
    Efavirenz Dissolution Enhancement I: Co-Micronization - MDPI
    ... EFV processing to enhance the dissolution profile. It can lead to greater drug bioavailability, as Vogt and colleagues [7] demonstrated for fenofibrate.<|separator|>
  11. [11]
    Manufacturing Approaches of New Halogenated Agrochemicals
    Feb 17, 2022 · From these, around 29 % of the new halogenated agrochemicals contain a trifluoromethyl (F3C) group, approximately 10 % a difluoromethyl (F2HC) ...Missing: CF3 | Show results with:CF3
  12. [12]
    [PDF] trifloxystrobin (213)
    Trifloxystrobin's IUPAC name is methyl (E)-methoxyimino-{(E)-α-[1-(α,α,α-trifluoro- m-tolyl)ethylideneaminooxy]-o-tolyl}acetate, with a molecular formula of C ...
  13. [13]
    Novel Synthetic Method and Insecticidal Evaluation of CF 3
    Feb 22, 2025 · The trifluoromethyl (CF3) group is widely incorporated in pharmaceuticals and agrochemicals due to its notable electronegativity, high ...Missing: percentage | Show results with:percentage
  14. [14]
    An overview of the uses of per- and polyfluoroalkyl substances (PFAS)
    Oct 30, 2020 · Here we present a systematic description of more than 200 uses of PFAS and the individual substances associated with each of them (over 1400 PFAS in total).
  15. [15]
    Recent Advances and Outlook for the Isosteric Replacement of ...
    The CYP450 and peroxidase enzyme systems are generally considered the most important groups of enzymes involved in bioactivation, producing either electrophilic ...
  16. [16]
    METABOLISM OF FLUORINE-CONTAINING DRUGS
    Metabolic ac- tivation by the cytochrome P450 system is generally blocked at the carbon to which fluorine is attached. Miller & Miller (53) used fluorine ...
  17. [17]
    Full article: Recent advances in green fluorine chemistry
    In the early 1930s, HF was found to be an alternative to SbF3 in Swarts reaction. This approach has been used for the industrial production of Freons such as ...
  18. [18]
    Overview on the history of organofluorine chemistry from the ...
    This paper overviews the historical development of organofluorine chemistry especially from the viewpoint of material industry.
  19. [19]
    603. The reactions of fluorocarbon radicals. Part I ... - RSC Publishing
    The reaction of iodotrifluoromethane with ethylene and tetrafluoroethylene. R. N. Haszeldine, J. Chem. Soc., 1949, 2856 DOI: 10.1039/JR9490002856. To request ...Missing: trifluoromethylation CF3I
  20. [20]
  21. [21]
  22. [22]
  23. [23]
  24. [24]
  25. [25]
    Radical trifluoromethylation - RSC Publishing
    Missing: mechanism | Show results with:mechanism
  26. [26]
    Selective Radical Trifluoromethylation of Native Residues in Proteins
    Jan 5, 2018 · We reveal a mild method for direct protein radical trifluoromethylation at native residues as a strategy for symmetric-multifluorine incorporation on mg scales ...
  27. [27]
  28. [28]
  29. [29]
    Electrophilic Trifluoromethylation by Use of Hypervalent Iodine ...
    This review article is concerned with the use of compounds 1 and 2 (Figure 1) as effective and very versatile reagents for trifluoromethylation reactions of ...
  30. [30]
    Shelf-stable electrophilic trifluoromethylating reagents: A brief ...
    In this review, we wish to briefly provide a historical perspective of the development of so-called “shelf-stable electrophilic trifluoromethylating reagents”.
  31. [31]
    Reactivity of Electrophilic Trifluoromethylating Reagents - 2024
    Feb 28, 2024 · Only highly reactive C- and S-centered nucleophiles were used in uncatalyzed reactions with the Umemoto reagent II 1 c (Scheme 3).Results And Discussion · Kinetic Studies · Correlation Analysis
  32. [32]
    Transition-Metal-Catalyzed Trifluoromethylation of Aryl Halides - 2010
    Sep 28, 2010 · Making inroads into ArCF3: Recent advances in the copper- and palladium-catalyzed cross-coupling of aryl halides and the trifluoromethyl anion, ...
  33. [33]
    Copper-Promoted Sandmeyer Trifluoromethylation Reaction
    The study introduces a copper-promoted Sandmeyer trifluoromethylation reaction for converting aromatic amines into trifluoromethylated arenes and heteroarenes ...
  34. [34]
    Active Trifluoromethylating Agents from Well-Defined Copper(I)
    Copper-Promoted Sandmeyer Trifluoromethylation Reaction. Journal of the American Chemical Society 2013, 135 (23) , 8436-8439. https://doi.org/10.1021 ...
  35. [35]
    Trifluoromethylation of arenes and heteroarenes by means of photoredox catalysis - Nature
    ### Summary of Photocatalytic Trifluoromethylation Using CF₃SO₂Cl
  36. [36]
    A General Organocatalytic System for Electron Donor–Acceptor ...
    Jul 28, 2021 · We report herein a modular class of organic catalysts that, acting as donors, can readily form photoactive electron donor–acceptor (EDA) complexes with a ...
  37. [37]
    Existence of the Halocarbonyl and Trifluoromethyl Cations in the Condensed Phase
    ### Summary of Trifluoromethyl Cation (CF₃⁺) Information
  38. [38]
    On the heats of formation of trifluoromethyl radical CF 3 and its cation
    Jan 1, 1997 · The CF+ and CF 3 + fragment ion yield curves from C2F4 have been remeasured by photoionization mass spectrometry.
  39. [39]
    Selected Ion Flow Tube Study of the Gas-Phase Reactions of CF+ ...
    Jul 14, 2012 · We study how the degree of fluorine substitution for hydrogen atoms in ethene affects its reactivity in the gas phase.
  40. [40]
    Ionic Lewis superacids in the gas phase. Part 3. Reactions of ...
    The trifluoromethyl ion CF3+ is evaluated as a chemical ionization (CI) precursor in a compact Fourier Transform Ion Cyclotron Resonance (FTICR) mass ...
  41. [41]
    Structures and stabilities of C2H2F3+ cations
    Despite the strong a-electron withdrawing effect of fluorine,. CF3+ is stabilized relative to CH3+: the reaction CH3+ +. CHF3 -, CF3+ + CH4 is exothermic by ...<|control11|><|separator|>
  42. [42]
    Advances in photocatalytic research on decarboxylative ... - Frontiers
    May 13, 2025 · This review systematically summarizes advancements in photocatalytic decarboxylative trifluoromethylation using TFA and its derivatives over the past decade.
  43. [43]
    Ground-state flavin-dependent enzymes catalyzed enantioselective ...
    Two engineered flavin-dependent enzymes are successfully developed to catalyze stereoselective hydrotrifluoromethylation and trifluoromethyl-alkyl cross- ...
  44. [44]
    Ball-milling and piezoelectric materials enabled radical ...
    Under mechanochemical compression, the use of 0.5 to 1.0 equivalents of piezoelectric materials enabled the solid-state formation of the key CF3 radical ...Missing: generation CF3I
  45. [45]
    Solid‐State Radical C−H Trifluoromethylation Reactions Using Ball ...
    Sep 11, 2020 · Piezoelectricity has been used to generate trifluoromethyl (CF3) radicals for the mechanochemical C−H trifluoromethylation of aromatic compounds ...Missing: CF3I | Show results with:CF3I
  46. [46]
    Review Recent advances in direct 1,2-bis-trifluoro-methylation ...
    Review. Recent advances in direct 1,2-bis-trifluoro-methylation, -methoxylation, -methylthiolation, and –methylselenylation of unsaturated hydrocarbons.
  47. [47]