Abstract
The mechanism of CF3 transfer from R3SiCF3 (R = Me, Et, iPr) to ketones and aldehydes, initiated by M+X– (<0.004 to 10 mol %), has been investigated by analysis of kinetics (variable-ratio stopped-flow NMR and IR), 13C/2H KIEs, LFER, addition of ligands (18-c-6, crypt-222), and density functional theory calculations. The kinetics, reaction orders, and selectivity vary substantially with reagent (R3SiCF3) and initiator (M+X–). Traces of exogenous inhibitors present in the R3SiCF3 reagents, which vary substantially in proportion and identity between batches and suppliers, also affect the kinetics. Some reactions are complete in milliseconds, others take hours, and others stall before completion. Despite these differences, a general mechanism has been elucidated in which the product alkoxide and CF3– anion act as chain carriers in an anionic chain reaction. Silyl enol ether generation competes with 1,2-addition and involves protonation of CF3– by the α-C–H of the ketone and the OH of the enol. The overarching mechanism for trifluoromethylation by R3SiCF3, in which pentacoordinate siliconate intermediates are unable to directly transfer CF3– as a nucleophile or base, rationalizes why the turnover rate (per M+X– initiator) depends on the initial concentration (but not identity) of X–, the identity (but not concentration) of M+, the identity of the R3SiCF3 reagent, and the carbonyl/R3SiCF3 ratio. It also rationalizes which R3SiCF3 reagent effects the most rapid trifluoromethylation, for a specific M+X– initiator.
Introduction
The inclusion of fluorine substituents in organic molecules is of pivotal importance to developments in, inter alia, pharmaceuticals,1 agrochemicals,2 electronics,3 materials chemistry,4 polymers,5 synthesis,6 and catalysis.7 The transfer of a formally nucleophilic CF3-moiety to an electrophile is a preeminent method for the synthesis of trifluoromethylated compounds.8 Conditions range from base-mediated reactions with fluoroform (CF3H)9 through to finely tuned borazine-based CF3 carriers recently reported by Szymczak.10 In 1989, Ruppert reported that TMSCF3 (1a)11 undergoes addition to aldehydes and ketones in the presence of 10 mol % KF.12 A faster process, using a soluble initiator (Bu4NF·xH2O; 0.6 mol %, TBAF) was reported soon after, by Prakash and Olah.13 Acidic workup affords the corresponding trifluoromethylated alcohols in good yield, Scheme 1.
This mild and selective process14 swiftly became adopted for the preparation of trifluoromethyl-carbinols,15 including enantioselective additions involving enantiopure ammonium salts as initiators.16 Indeed, over the past decade there has been an explosion of interest17 in CF3 transfer from TMSCF3 (1a) to carbon (e.g., carbonyls,14−17 imines,18 vinyl halides,19 and aromatics20) and to heteroatoms such as sulfur,21 selenium,22 phosphorus,23 boron,24 iodine,25 and bismuth.26 The formal loss of fluoride from CF3 to facilitate electrophilic TMSCF2 transfer27 or carbenoid CF2 transfer28 has also been developed, as have numerous metal-mediated and -catalyzed processes involving CF3 derived from TMSCF3 (1a).29
Despite anion-initiated trifluoromethylation by 1a having become a mainstream synthetic method,17−26 surprisingly little detail has emerged on the mechanism of CF3 transfer, under the conditions of application, Scheme 1.30 Various mechanistic dichotomies, including, inter alia, fluoride initiation versus fluoride catalysis, and siliconate versus carbanion23a pathways, have been noted by Denmark30a and by Reich,30b both of whom emphasize the lack of salient kinetic data.
Herein we report the first detailed study of the mechanism of anion-initiated CF3 transfer from TMSCF3 (1a) to ketone and aldehyde electrophiles.12,17 The in situ NMR/IR investigations include analysis of reaction kinetics, selectivity, and side reactions and the contrasting behavior of homologues triethylsilyl (TES) (1b) and triisopropylsilyl (TIPS) (1c). Throughout the investigation, the kinetic studies have both informed and been directed by density functional theory (DFT) analysis of proposed intermediates. What emerges is a nuanced kinetic landscape in which trifluoromethyl transfer proceeds via a carbanion pathway (CF3–), with the rate dictated by the identity of the electrophile, the concentration of the initiating anion, the identity of the initiator countercation, the electrophile/R3SiCF3 (1) concentration ratio, and the identity of the reagent (1a–c).
Results and Discussion
1. Prior Studies
In early studies, a termolecular anionic chain reaction (mechanism I, Scheme 2) was suggested for trifluoromethylation by 1a.13a This was later expanded to a two-step process (mechanism II), where a pentacoordinate alkoxy-siliconate (B) delivers CF3 to the ketone and in doing so liberates the O-silylated product.14 Mechanism II has been extensively adopted in the design and interpretation of asymmetric trifluoromethylation.16,29,31
In 1999, Naumann32 and Kolomeitsev and Röschenthaler33 independently reported on the reaction of a range of soluble fluoride sources (e.g., [Me4N]+F–) with TMSCF3 (1a) at low temperature. Detailed 1H, 13C, 19F, and 29Si NMR analysis identified the products as pentacoordinate complexes [Me3Si(F)(CF3)]−M+ (C) and [Me3Si(CF3)2]−M+ (D). Both complexes decompose above −20 °C.32,34 The speciation (C/D) is dependent on the stoichiometry (M+F–/1a), and the structure of D was confirmed by single-crystal X-ray diffraction. Addition of cyclohexanone at −60 °C, followed by hydrolysis, afforded the corresponding trifluoromethylated alcohol, mechanism III.32,35
In 2014, Prakash36 showed that the elusive37 trifluoromethyl anion(oid)38 can be detected in situ (13C, 19F NMR) at low temperatures after addition of KOtBu/18-crown-6 to 1a. With the much bulkier reagent TIPSCF3 (1c), the generation of ion-paired [K(18-c-6)]+[CF3]− (E) proceeds quantitatively at −78 °C over a period of 30 min. Subsequent addition of PhCOMe (11 equiv) or PhCHO (4 equiv) afforded CF3-addition products (22–68%) after quenching with H2O, mechanism IV.36 In 2015, Grushin38 demonstrated that use of crypt-222 (L, Scheme 2) facilitates generation of the free CF3– carbanion, a tetrahydrofuran (THF) solution-phase “noncovalently bound ionic species”.38 The structure of the highly air- and temperature-sensitive salt, [K(crypt-222)]+[CF3]− (E), was confirmed by single-crystal X-ray diffraction.38,39
The pioneering studies summarized above have been highly enlightening regarding the structure and stability of pentacoordinate (trifluoromethyl)siliconates (C, D)32,33 and their ability to release the trifluoromethane anion(oid) (E) under specific conditions.36,38 However, they do not yield direct detail on the kinetics and mode of transfer of CF3 from TMSCF3 (1a) to a carbonyl electrophile, using a catalytic fluoride-based initiator (M+X–), at ambient temperature.12,13
2. Preliminary Investigations
We began by studying the reaction of TMSCF3 (1a) with aldehydes and ketones in THF, chlorobenzene, and dimethylformamide (DMF). After addition of catalytic quantities (0.1 to 1 mol %) of TBAF, 19F NMR readily facilitated analysis of the proportions of residual reagent (1a) and the [1,2]-addition products. The reaction of 4-fluoroacetophenone (2) in THF at ambient temperatures proved ideal, the additional 19F nucleus allowing simultaneous analysis of reagent (1a; 0.48 M), substrate (2; 0.40 M), and product (3OTMS), Scheme 3.
Reactions were assembled manually in 5 mm NMR tubes in the glovebox prior to analysis in situ by 19F NMR. Three side-products were identified: fluoroform (CF3H), the silylenol ether (4OTMS), and a homologated addition product (5OTMS). Reactions conducted in d8-THF proceeded analogously and generated CF3H, not CF3D.40 The identity of 5OTMS, which was confirmed by independent synthesis, is consistent with difluorocyclopropanation of silylenol ether 4OTMS to generate 10, followed by a known41 anion-induced ring-opening elimination to give fluoroenone 11 and subsequent 1,2-selective12 trifluoromethylation. Addition of independently synthesized4210 to the reaction (Scheme 3) generated 5OTMS.
Reaction rates and extent of fluoroform generation (Scheme 3) were found to vary significantly between batches of TBAF (1 M, THF, ∼5 wt % H2O). Replacing TBAF with anhydrous [Bu4N][Ph3SiF2] (TBAT)43 gave more reproducible data. However, the fast turnover precluded detailed kinetic analysis; this aspect was addressed using stopped-flow methods, vide infra. Nonetheless, 19F NMR analysis revealed that CF3H is liberated in two distinct phases. The first is an initial burst of extremely rapid CF3H generation and arises from TBAT-catalyzed reaction of TMSCF3 (1a) with traces of adventitious water.44 The second phase of CF3H generation proceeds in concert with reaction of the ketone (2) and directly correlates with the rate of generation of silylenol ether (4OTMS), as confirmed by 2H-labeling (d3-2 → CF3D + d2-4OTMS). The selectivity (3OTMS versus 4OTMS) is discussed later.
3. Stability, Inhibition, and Tests for Radicals
The stability of the reaction system after complete consumption of the limiting reagent (ketone 2 or TMSCF31a) was found to depend on which one was in excess. Reactions in which 2 was in excess underwent turnover on addition of further TMSCF31a, even after a period of many hours. In contrast, for reactions where 1a was in excess, additional 2 had to be added within a few minutes to fully reinstate turnover (see SI), consistent with the known instability of pentacoordinate (trifluoromethyl)siliconates, e.g., C and D, at ambient temperatures.32−34 Further tests established that the reactions were not sensitive to exogenous water per se, as they rapidly self-dehydrated via generation of CF3H + hexamethyldisiloxane, prior to reaction of the ketone (2).45 The rates were unaffected by visible light, by exogenous product (3OTMS), and by CF3H. Deliberate sparging of the normal reaction mixture (1a/2/TBAT 0.15 mM, 0.038 mol %, THF, Scheme 3) with air caused complete inhibition of turnover, but only when a sufficient volume of CO2 (∼400 ppm) had been added to convert the active anion(s) into trifluoroacetate (i.e., [Bu4N][CF3CO2], detected by 19F NMR). Separate controls confirmed that the rate of trifluoromethylation is unaffected by CO2-scrubbed air and that [Bu4N][CF3CO2] is not effective as an initiator.
However, the reactions were inhibited by addition of the persistent radical tetramethylpiperidinooxy (TEMPO). Indeed, just 0.45 mM TEMPO induced complete inhibition of the reaction of 1a with 2, initiated by 0.15 mM TBAT (Scheme 4, A). In contrast, TEMPO had a negligible impact on reactions employing TES (1b) and TIPS (1c), even when present at much higher concentrations (80 mM TEMPO); the origins of this profound difference in behavior is discussed later. Nonetheless, further tests for discrete radical intermediates46,47 were conclusively negative: 4-F-benzophenone (12) exclusively underwent 1,2-addition (Scheme 4, B),48,49 cyclopropyl ketones (6/7) reacted without any trace of competing ring-opening50 (Scheme 4, C), and competition between ketone 2 and 4-biphenyl methyl ketone for limiting TMSCF3 (1a) favored 2 (krel = 1.93).51
4. General Effects of Initiator on Rate and Selectivity
A range of initiators (M+X–) were tested and found to strongly impact the reaction outcome. In the majority of cases, the reactions initiated “instantly” and the identity of X– had no influence on the rate52 or selectivity (3OTMS/4OTMS). Specific effects were found to be dictated by the identity of the countercation (M+), Table 1.
Table 1. Examples of Effect of Initiator M+ and Reagent (1a–c) on Selectivity (3OSi/4OSi) and Rate of Trifluoromethylation of 2.
M+ | [M+X–]0, mM | TMS 1a 3/4a (time)b | TES 1b 3/4a (time)b | TIPS 1c 3/4a (time)b |
---|---|---|---|---|
[Bu4N]+ | 1.5 | 12/1 (<90 s) | 1.5/1 (<90 s) | 1/1 (30 min) |
[K]+ | 0.15 | 36/1 (<90 s) | 3.0/1c (30 min) | NR |
[K(L)]+d | 1.5 | 6.6/1 (6 min) | 2.4/1 (<90 s) | 1/1 (3.6 min) |
Selectivity 3OSi/(4OSi+CF3H) measured in situ by 19F NMR after manual assembly in an NMR tube; selectivity is independent of X–.
times indicated are for >97% conversion of 2, at 300 K.
85% conversion.
[K(L)]+ = K(crypt-222)]+, generated in situ from KOPh + crypt-222. NR = No reaction.
Reactions where M+ = K+ and Cs+ proceeded rapidly to completion, with higher selectivity for 3OTMS/4OTMS compared to Bu4N+. Reactions stalled when the cation was Li+ or Na+.53 For the K+-mediated system, the rate was strongly attenuated by addition of 18-crown-6 or crypt-222, with the latter causing turnover to become slower and less selective (3OTMS/4OTMS) than reactions initiated by TBAT (countercation Bu4N+). The identity of M+ was also found to affect the degree of charge development (ρ ranging from 1.8 to 3.0) in the ketone (R = Me, Scheme 5) at the product-determining transition state for CF3 transfer. Benzaldehydes (R = H) behaved analogously.
5. Effect of Silyl Reagent on Rate and Selectivity
To further probe the CF3 transfer process, we compared TMSCF3 (1a) with TESCF3 (1b) and TIPSCF3 (1c), Table 1. The effects of changing the reagent were counterintuitive and initially misleading regarding the mechanism of CF3 transfer, vide infra. Reactions employing TESCF3 (1b) gave lower selectivity (3OTES/4OTES ≈ 1.5/1) and proceeded very rapidly, even at low TBAT concentrations (150 μM, 0.0375 mol %; below this, reactions failed to initiate). In contrast, reactions employing TIPSCF3 (1c) proceeded very slowly, requiring high initiator concentrations to proceed efficiently (>1.5 mM, 0.375 mol %) and gave even lower selectivity (3OTIPS/4OTIPS ≈ 1/1).
Further insight was afforded by reaction of a 50/50 mixture of TMSCF3 (1a) and TESCF3 (1b), initiated by TBAT (75 μM, 0.019 mol %), Figure 1. The first 4 min of reaction is dominated by turnover of TMSCF3 (1a) to generate 3OTMS/4OTMS, and upon near-complete consumption of 1a, turnover accelerates substantially as the TESCF3 (1b) is engaged to generate 3OTES/4OTES.The data indicate that the less-hindered reagent (1a) monopolizes the anion, but undergoes slower turnover.
Under conditions where anion-induced reactions of TMS (1a), TES (1b) and TIPS (1c) with 2 could be conducted slowly enough to be accurately monitored in situ by 19F NMR, the ratios of enol/addition product (4OSi/3OSi) were all constant throughout the reaction evolution, Figure 2. A further distinction originated from the impact of the addition of crypt-222 to KOPh-initiated reactions. As noted above, for TMSCF3 (1a) the incarceration of the K+ in the crypt-222 ligand substantially attenuates the rate and selectivity. In stark contrast, for TIPSCF3 (1c), turnover is substantially accelerated by addition of crypt-222 to inhibit K+/anion pairing.
Reactions with labeled ketone (aryl-d4-2; CD3-2; 13CO-2) were also instructive, Table 2. Reaction of TMSCF3 (1a) with 2 initiated by TBAT (0.15 mM) proceeds with a very low 13C kinetic isotope effect (KIE), determined by competition with aryl-d4-2, after normalizing for the effect of aryl deuteration.
Table 2. KIEs and 2H Exchange in the Reaction of Ketone 2a,b.
CL3 (reagent) | 1a–c | 2,3OSi | kH/kD | |
---|---|---|---|---|
CH3 (13C=O) | 1a | CH3 | (k12C/k13C = 1.008)c | |
CH3 (C6D4) | 1a | CH3 | 1.038d | |
CD3 | 1a | CD3 | 6.4 | (3/4OTMS = 72/1) |
CD3 + CH3 | 1a | CD3/CH3 only | 6.1 | (rate: CF3H/CF3D) |
CD3 | 1b | CD3 | 3.1 | (3/4OTES = 4.3/1) |
CD3 + CH3 | 1b | partial CD3–n/Hn | (3/4OTES = 2.3/1) | |
CD3 | 1c | CD3 | 1.1 | (3/4OTIPS = 1.1/1) |
CD3 + CH3 | 1c | full CD3–n/Hn | 1.0 | (rate: CF3H/CF3D) |
Ketone (2/2H3-2; 0.40 M), 1a–c (0.48 M), THF, 300 K. TBAT (0.04 mol %, 0.15 mM).
Selectivity 3OSi/(4OSi + CF3H/D) and exchange measured in situ by 19F NMR analysis.
KIE determined by competition with aryl-d4-2.
2H KIE induced by aryl-deuteration, determined by competition with unlabeled 2.
In contrast, a substantial primary 2H KIE, determined from [CF3D] versus [d3-3OSi], as in Figure 2, increases the addition/enol selectivity (kH/kD = 6.4). Reactions of mixtures of 2 and d3-2 proceeded with no detectable scrambling of D/H between 2/D3-2 during turnover, provided that [1a]0 > [2]0, and again proceeded with a high KIE (kH/kD = 6.1). With TESCF3 (1b) a moderate KIE (kH/kD = 3.1) was observed, with a trace of D/H exchange between 2 and D3-2 on co-reaction, and thus into products dn-3/dn-4. With TIPSCF3 (1c) there was no significant KIE and a statistical mixture of isotopologues of dn-2/3 (n = 0–3) was evident immediately after initiation of the reaction.54
6. Variable-Ratio Stopped-Flow NMR and IR
Detailed exploration of the kinetics of the trifluoromethylation by 1a required techniques for rapid acquisition of kinetic data (some systems had formal turnover frequencies well in excess of 5000 s–1, vide infra) in a time- and material-efficient manner. Stopped-flow techniques are ideal for rapid and reproducible initiation and analysis of these reactions. However, the classic fixed-ratio dual input mode of operation (A + B; Figure 3a) requires separate solutions to be prepared for every variation in conditions. For a three-component process such as R3SiCF3 (1) + ketone 2 + initiator (M+X–), a very large number of stock solutions are required to study reactions with different concentrations of reactants and initiator.
To address this issue, we constructed a stopped-flow system, in which the delivered volumes of three solutions (A, B, C) are independently variable,55 using a computer-controlled triple stepper-motor system, Figure 3b. This setup allowed systematic analysis of the kinetics across a wide range of initial conditions, using just four stock solutions, mixing {i + iii + iv} varies [2]0; mixing {ii + iii + iv} varies [1]0; and mixing {iii + iv + THF} varies [M+X–]0, while keeping the other species constant; see SI for full details. The new system was implemented in two modes: IR and NMR.56 The former simply required adaptation of our recently developed thermostated ATR-FTIR stopped-flow cell,57 replacing the dual mixing stage with a triple mixer and a gated reaction volume. The analogous setup for variable-ratio stopped-flow NMR required bespoke construction. The principles for continuous-flow NMR recently reported by Foley et al.58 were employed for the basic design, such that the reaction vessel and associated components can be installed simply by insertion of the device into the sample transit of a standard unmodified NMR spectrometer. Nuclei premagnetization is facilitated in three independent reservoirs (A, B, C) located as close as possible to the magnetic field center, Figure 3c. The reservoirs connect at a tripodal-geometry mixer that discharges via a 0.5 mm i.d. glass capillary into a 3 mm external diameter 300 μL glass NMR flow-cell. The tube terminates at the base of the cell, with the waste outlet at the top. A fourth input to the mixer allows the system to be flushed with solvent between runs. Thermostating is achieved by passage of a heat-transfer medium (aqueous ethylene glycol), using an externally controlled recirculator, through an umbilical containing all stages of the stopped-flow circuit, except for the glass flow-cell, which is located within the spectrometer-thermostated probe head; precalibration ensures temp1 = temp2. During a typical stopped-flow (SF) “shot”, a total of 600 μL is delivered through the flow-cell at a rate of 1–2 mL s–1, fully displacing the previous contents and replacing it with 300 μL of freshly assembled reaction mixture; charging requires 70–130 ms (measured independently by UV–vis), with high-quality NMR spectra (N2-cryoprobe) achievable immediately thereafter. Control of the timing of the NMR pulse sequence is achieved by a trigger signal, sent to the spectrometer console from the computer-controlled triple stepper-motor system, immediately after the 300 μL flow-cell has been freshly charged.
7. Kinetics of Trifluoromethylation by TMSCF3 (1a) and TIPSCF3 (1c)
The kinetics of reactions initiated by M+X–, where M+ = Bu4N+, K+, and Cs+, were studied in detail by SF-IR and SF-NMR across a wide range of concentrations of 1a, 2, and [M+X–]0. For FTIR, the decay in the IR C–F stretching mode (1056 cm–1) of the TMSCF3 (1a) and the growth in C–F stretching mode (1165 cm–1) of 3OTMS were collected at scan rates of 14 or 28 s–1 with a resolution of 2 or 8 cm–1, respectively. 19F NMR analysis allowed detailed analysis of the reaction components, but was naturally more limited in terms of temporal resolution. For faster reactions, a technique involving the interleaving of a series of spectra from a sequence of stopped-flow NMR “shots” was employed, affording a higher virtual temporal resolution.
A key component in analysis of the kinetics was the dependence of the temporal-concentration evolution of the product (3OTMS) on the concentration ratio of ketone 2 and TMSCF3 (1a). Systematic studies of initial rates using TBAT led to an empirical rate equation for turnover frequency (TOF) in which the initiator (Bu4N+X–) and ketone 2 are first order and the TMSCF3 (1a) reagent approximately inverse first order (eq 1).59 Control experiments in which the reactions were run in the presence of exogenous product (3OTMS) confirmed that it does not act as an inhibitor.
The inhibitory effect of the TMSCF3 reagent 1a (Ki1; eq 1) results in very distinctive temporal concentration profiles for the reaction, simulations of which are presented later. For example, when the initial ratio of reactants is equal ([2]0 = [1a]0), their ratio remains constant ([2]t/[1a]t = 1) throughout the reaction. What arises is an apparent pseudo-zero-order consumption of the reactants (TOF = kobs) for the majority of the reaction evolution. Conversely, when there is an excess of ketone 2 over 1a, the rate of turnover rises as a function of conversion, becoming very rapid in the final phases of reaction where [2]t/[1a]t ≫ 1.
1 |
2 |
Systematic studies using M+X– (M+ = Li+, Na+, K+, Cs+), which induce very rapid turnover, proved more challenging. Reactions where M+ = Li+ and Na+ stalled before completion and were not reproducible. KOPh and CsOPh initiated at very low concentrations, without an evident induction period, proceeded to completion, and provided reproducible kinetics. Study of the initial rates suggested higher-order dependencies on TMSCF3 ([1a]0, again inverse) and on [M+X–]0, with the ketone 2 remaining first-order. However, the reactions evolve with near-identical behavior to those initiated by TBAT (eq 1).59 The dichotomy is indicative of the presence of exogenous inhibitor(s) in low concentration in the TMSCF3 (1a) reagent that are not consumed during reaction. Increasing the initial concentration of the reagent ([1a]0) or decreasing the initiator concentration ([M+X–]0) causes a greater mole fraction of exogeneous inhibition (xEI), eq 1.59,60 Addition of [K]+[(C6F5)4B]−, to provide an additional soluble K+ source with a non-nucleophilic counteranion, had no impact on the kinetics of reactions initiated by KOPh, indicative that the rate is dependent on the initiating anion concentration and the countercation identity (but not its concentration).61 Addition of potassium-binding ligands attenuated the rates substantially, and with crypt-222, the system underwent turnover slower than with Bu4N+ (a 3 orders of magnitude rate reduction compared to free K+).
The kinetics of trifluoromethylation of 4-fluorobenzaldehyde (13) by 1a were also explored using TBAT as initiator. The aldehyde undergoes significantly faster trifluoromethylation than ketone 2 (kald/kket ≈ 80, at 21 °C), requiring lower initial TBAT concentrations and causing the traces of exogeneous inhibitor(s) in 1a to complicate the kinetics.59 Competing ketone 2 with aldehyde 13 (9/1 ratio) using stopped-flow 19F NMR to analyze the transient substrate ratio (2/13) during the first 5–30 s of reaction indicated that the relative rate of trifluoromethylation is independent of [TBAT]0 (96–384 μM) and 1a (0.08 to 0.48 M). Overall, the data are indicative that aldehyde 13 follows the same general kinetics as ketone 2, i.e., eq 1.59,60 The rate of trifluoromethylation of ketone 2 using TIPSCF3 (1c) was much slower than with 1a. Again, the kinetics were impacted by exogenous inhibitor(s) in the reagent ([1c]0), the effect of which (xEI) varied from batch to batch of 1c; see SI. Using TBAT as initiator, the reactions evolve with a first-order dependency on the initiator and on the reagent ([1c]t), with inhibition by the ketone (Ki2; eq 2). In other words, the kinetic dependencies are the opposite to that found for 1a (compare eqs 1 and 2), with reactions accelerating when there is an excess of 1c over 2. Reactions of 2 with 1c initiated by KOPh were slower than those initiated by TBAT and were accelerated on addition of crypt-222; the opposite phenomena to those observed with 1a.
8. Stopped-Flow 19F NMR Analysis of Siliconate and Alkoxide Intermediates, Exchange Dynamics with TMSCF3, and Initiator Regeneration
By use of 4-F-benzophenone (12; δF −107.0 ppm), which reacts slower than 2, and reducing the reaction temperature to 275 K, the temporal speciation of the initiator-derived species (10 mol % TBAT) was monitored using stopped-flow 19F NMR, Figure 4. The known but unstable hypervalent bis-CF3-siliconate (D; δF −63.3 ppm)32−34 is generated instantly. Integration against fluorobenzene (internal standard, δF −113.2 ppm) shows D to be present at 10 mol % and thus the predominant anion speciation. A key feature in the time series is the dynamic line-broadening in D that is constant throughout the reaction, but develops in the TMSCF3 (1a) reagent (δF −66.6 ppm) as its concentration is depleted by the overall reaction with superstoichiometric ketone 12. In parallel with this is a marked acceleration in product generation (14OTMS, δF −72.4 and −113.7 ppm), consistent with eq 1. After 6 s, the TMSCF3 (1a) is fully consumed and TBAT (δF –97.4 ppm) is regenerated from Ph3SiF/Me3SiF. The dynamic line-broadening in D/1a can be satisfactorily simulated using a three-spin exchange process in which D is in rapid dissociative equilibrium (kexch ≈ 180 s–1; ΔG⧧ ≈ 13 kcal mol–1) with 1a and a low concentration of (unobserved) [Bu4N][CF3] (E).62 At 300 K, the line-broadening is very extensive and D short-lived.
Analogous experiments using TIPSCF3 (1c) gave a very different outcome. Reactions conducted with 1c at 275 K were slow enough to be followed using ketone 2 (δF −106.7 ppm), Figure 5. The 19F NMR signal for 1c remains sharp until 2 has been fully consumed. In contrast to reactions with 1a (Figure 4) the alkoxide (3O–; δF −118.4) is present in significant concentration and exhibits dynamic line-broadening (see inset in Figure 5). The signal for ketone 2 also exhibits dynamic line-broadening, immediately after addition of TBAT. On complete consumption of 2 (∼120 s), the signals for remaining 1c and CF3H are broadened, presumably due to indirect exchange involving CF3–. After a further 300 s, 1c is fully consumed and the CF3H doublet becomes sharp again.
9. General Mechanism for Anion-Initiated CF3 Transfer from R3SiCF3 to Ketones and Aldehydes
The data outlined in Sections 2 to 8 above (see SI for full details) indicate that the M+X–-initiated trifluoromethylation of ketone 2 by TMSCF3 (1a) involves an electrophile–nucleophile reaction, in which the CF3 transfer is accompanied by M+. Enolsilane 4OTMS is also generated (≤2% when M+ = K+ and 7% when M+ = Bu4N+) with coproduct CF3H (kH/kD = 6.1). Using TIPSCF3 (1c), approximately 50% of the product is 4OTIPS and kH/kD = 1.0. Contrasting kinetic behavior is observed for 1a (eq 1) versus 1c (eq 2), with the roles of reactant for turnover and inhibitor reversed between the two systems. These disparate sets of observations can easily be misinterpreted as turnover for 1a versus 1c arising from different pathways, e.g., siliconate versus carbanion. However, analysis of the kinetics, KIEs, and DFT calculations of a wide range of potential intermediates (see SI) eventually leads to the conclusion that the two reagents elicit contrasting kinetics, selectivity (3OSi/4OSi), and KIEs, by biasing one of two extremes in a single overarching mechanism. Calculations employed the M06L/6-31+G* level of theory, which was selected from a range of other functionals and larger basis sets that were considered63 (see SI), as it provided the best quantitative agreement with experiment. All calculations were performed in Gaussian09,64 with THF solvation incorporated via a polarizable continuum model (PCM) single point at the same level of theory and with T = 298 K and pressure at 24.45 atm to achieve a 1 M standard state.65 Kinetic isotope effects were computed using the Kinisot program.66 Some of the TES- and TIPS-bearing structures required the “loose” settings during the geometry optimization, presumably because of the flat potential energy surface associated with the long Si–CF3 bonds.
The calculations permitted several possible structure types (such as hexacoordinate silicon dianions) to be excluded from consideration and also revealed pronounced differences between intermediates based on TMS, TES, and TIPS, where the increasing steric bulk substantially destabilizes the pentacoordinate anions, Figure 6. Extensive calculations were conducted to test for direct nucleophilic reactivity of the pentacoordinate anions B and D. All calculations revealed that direct transfer of CF3 from the silicon center to an electrophile requires concomitant inversion of the CF3, with a prohibitively large barrier (>100 kcal mol–1; in line with the barrier for inversion of the free CF3 anion).67 The pentacoordinate siliconate anions thus act as reservoirs, not active nucleophiles, liberating free (non-silicon-coordinated) CF3– via dissociation. The transition state for addition of the CF3– anion(oid) to the ketone formally involves movement between a nonclassical hydrogen-bonded complex and the addition product, a process that occurs with low calculated barrier (7.5 kcal mol–1) and well represents the process that occurs once the two species are in contact. The calculations support the known preference for deprotonation (kCH) in the gas phase68 and for addition (kCO) once solvation is introduced, as observed experimentally for TMSCF3 (1a). The loose addition transition state leads to a negligible 13C KIE (carbonyl) for addition, while a large primary 2H KIE is computed for C–H deprotonation. Relative rates computed from activation free energies suggest ρ = 2.0 for addition to acetophenones and a lower barrier for addition to 4-F-benzaldehyde (13) versus 2 (ΔΔG⧧ 2.6 kcal/mol; krel = 81). All of these computed values are in excellent agreement with experiment.
A general mechanism for the trifluoromethylation of ketones and aldehydes by R3SiCF3 reagents (1) in the presence of a catalytic quantity of initiator (M+X–) can thus be assembled, Figure 7. The one overarching mechanism, discussed below in the context of two extremes (Vi and Vii), rationalizes why the turnover rate (per M+X– initiator) for a given electrophile depends on the initial concentration (but not identity) of X–, the identity (but not concentration) of M+, the identity of the reagent (1a–c), and the electrophile/reagent ratio (2/1).
10. Mechanism Vi
In this regime, which describes reactions involving TMSCF3 (1a), the dominant anion speciation is the bis(trifluoromethyl) siliconate (D),32−34 generated in rapid equilibrium (K3) with CF3– (E)36,38 and 1a, as observed by NMR, Figure 4. The product-determining step (k1) involves reaction of CF3– (E) with the ketone (2) (kCO + kCH), and the reagent (1a) thus acts as a reversible inhibitor. The stronger the association of M+ with CF3– (see Section 13) and with the carbonyl oxygen, the faster the turnover rate: Bu4N+ < [K(crypt-222)]+ < [K(18-c-6)]+ < K+. The initial concentration ratios of the reactant versus the reagent dictate the temporal evolution of the reaction. When [2]0/[1a]0 = 1, pseudo-zero-order kinetics are obtained, whereas when [2]0/[1a]0 ≥ 1, the rate rises throughout the reaction, becoming very fast (asymptoting to k–3[D]) in the final stages. The kinetics of trifluoromethylation of ketone 2 by TMSCF3 (1a) can be satisfactorily simulated, Figure 8, using a truncated form of mechanism Vi that retains relationships required for TOF modulation as the temporal concentration ratio [2]t/[1a]t evolves.
11. Mechanism Vii
In this regime, which describes reactions involving TIPSCF3 (1c), the dominant anion speciation is a combination of the product alkoxide (3O–), the enolate anion (4O–), and MX. Ketone 2 can reversibly H-bond (see F in Figure 7) with oxy-anions 3/4O–, as observed by NMR, Figure 5, leading to inhibition (K4).69 When [1c]0/[2]0 = 1 pseudo-zero-order kinetics are observed; reactions in which [1c]0/[2]0 > 1 exhibit accelerating rate in the last stages of reaction. The more strongly bound M+ to [3/4O–], the slower the reaction with 1c, leading to rates increasing in the series K+ < [K(18-c-6)]+ < Bu4N+ < [K(crypt-222)]+, i.e., the opposite order to Vi. When the nonenolizable ketone 4-F-benzophenone 12 is employed, the kinetics show clean pseudo-first-order decay in 1c (see SI), with no inhibition by 12 (i.e., mechanism Vii, where K4 = 0, and eq 2, where Ki2 = 0).
12. Competing Enolization
Also shown in Figure 7 is the generation of the enol ether (4OSi) and CF3H from ketone 2, for which the selectivity (4OSi/3OSi) is dependent on M+ and the reagent (1a–c), Table 1. The major pathway for generation of 4OTMS in mechanism Vi is via C–H deprotonation (kCH) with an attendant large primary 2H-KIE.70,71 In contrast, for mechanism Vii, the significant concentration of [3/4O–] allows keto–enol equilibrium (pKenol ≈ 8)72 in 2 to be approached, with attendant intermolecular scrambling of 2H between ketone methyl groups. Deprotonation (kOH) of the enol (2enol) is predicted (DFT) to be of very low barrier and thus proceed with a negligible 2H-KIE.70 Despite their different origins (kCH versus kOH) mechanisms Vi and Vii both lead to 4OSi/3OSi ratios that are independent of the concentration of reactants (1, 2) and constant throughout the reaction, Figure 2.
13. Cation–CF3 Interactions
The interactions between the CF3– anion (free and Si-bound) and the counter-cations K+ and Me4N+ (as a model for Bu4N+) were explored computationally, with multidentate CF3 interactions found to be favored, e.g. Figure 9; see SI for details.
The indirect transfer of CF3 from reagent 1a to the ketone/aldehyde, i.e., via a silicon-free carbanion E, has implications for the mode by which enantioselective catalysis can be achieved using chiral ammonium initiators, e.g., cinchonidinium salts. The CF3–anion binding modes found computationally for Me4N+ (Figure 9i) show how an ammonium cation might simultaneously interact with a CF3– anion and control a developing alkoxide anion, Figure 9ii. Mechanism Vi contrasts most,16d,16e,31 but not all,16f prior interpretations, where mechanisms II/III (Scheme 2) involving CF3-siliconates bearing the initiating (C) or propagating (B) anion, are proposed to play key roles in the enantioselective trifluoromethylation step.
14. Broader Mechanistic Aspects
The mechanistic features elucidated in the current study extend beyond carbonyl trifluoromethylation. A number of corollaries follow for generic anion-initiated trifluoromethylation of an electrophile (E) by 1a, or deprotonation (R–H),73 via pathways analogous to mechanism Vi, and where [E, R–H]0 ≫ [M+X–]0. Thus, the initiator (M+X–) affects the rate of reaction in a number of ways. [X–]0 sets the initial concentration of the siliconate ([D]0 = (1 – xEI)[X–]0),60 which, in the absence of endogenous inhibitors, is essentially constant throughout the reaction. The insurmountable barrier for CF3 inversion67 means that, independent of the identity of the electrophile, E, or proton donor, R–H, the siliconate is unable to effect direct anionic trifluoromethyl transfer, Figure 10i. In all cases, the reaction must proceed via a dissociative pathway, Figure 10ii, in which M+ plays a key role: the stronger the association of M+ with CF3, the more favorable k–3. In contrast, efficient regeneration of the siliconate (k2, Figure 7) is favored by weaker interactions between M+ and the anionic coproduct from trifluoromethyl transfer (CF3–E–; R–; or products thereof). When the anion is unable to react with 1a, stoichiometric initiation by [M+X–] is required.14−26
15. Exogenous Inhibition
Trifluoromethylations initiated by low concentrations of (M+X–) are highly sensitive to traces of exogenous inhibitor(s). Species that generate an anion (LG–) of insufficient reactivity toward 1a to propagate will terminate the anionic chain reaction, Figure 10iii. In a series of control experiments, additives of the form Z-LG, (Z = H, R3Si, LG = Cl, Br) were found to function as powerful inhibitors for the anion-initiated reaction of 2 with 1a. For example, the trifluoromethylation of 2 (0.4 M) initiated by 150 μM TBAT ceases immediately on addition of 150 μM TMSCl; see SI. Slower-onset irreversible inhibition is effected by the more hindered TIPSCl, which also inhibits the reaction of TIPSCF3 (1c). Competing consumption of 1a is effected by other species in low concentrations, including CCl4 (Cl transfer),74 Cl3CH (deprotonation/Cl transfer),73b and TMS–OH (deprotonation), but without significant chain termination. There was no detectable inhibition by dichloroethane (DCE), CH2Cl2,73b TMS-O-TMS, Ph3SiF, Me3SiCF2H, or MeCN.73c
In our experience, a diverse range of inhibitors and competitors (e.g., CCl4 and CHCl3) are present, in low concentrations and variable proportions, in commercial samples of TMSCF3 (1a). This leads to substantial differences in reaction outcome, depending on the supplier. For example, comparison of the reaction of 2 (0.40 M) with five samples of distilled 1a (0.48 M) revealed that the concentration of initiator (TBAT, KOPh) required to effect >99% conversion of 2 ranged from 30 μM to 2.0 mM (0.0075 to 0.5 mol %); see SI.
A major difference found between reactions involving reagent 1a versus 1b,c is the impact of the persistent radical, TEMPO, which powerfully inhibits reactions involving 1a, Scheme 4A. The difference in behavior toward TEMPO cannot arise from oxidation of the CF3 anion (E), as this is a common intermediate to all three reagents (1a–c), and the partitioning of E between reaction with the ketone (2) versus TEMPO will be constant across the series, i.e., independent of the provenance of the carbanion E. Since the major difference between reagents 1a and 1c under the conditions of the reaction is the dominant anion speciation (D; mechanism Vi, 1a, versus alkoxides 3O–/4O–, mechanism Vii, 1c), this suggests that reaction of siliconate D with TEMPO is responsible for the inhibition. We were unable to identify any products in situ or by quenching, arising from TEMPO under the standard reaction conditions; see SI. While siliconates of type D are also generated from 1b and 1c, they (a) are only present at low concentration or as transient species, thus reducing their net rate of reaction with TEMPO, and (b) may be more resistant to reaction with TEMPO due to their greater steric bulk.
Conclusions
The trifluoromethylation of ketones and aldehydes by TMSCF3 (1a), initiated by catalytic fluoride ion, has been employed in synthesis for three decades.17 Previous mechanistic work has focused on stoichiometric reactions of R3SiCF3 (1a,c) with anions at low temperatures, generating unstable trifluoromethyl siliconates (C, D)32−34 and carbanion(oids) (E),36−38 depending on conditions. Which of these two pathways is followed in catalytic reactions at ambient temperature has been a long-standing mechanistic dichotomy.30 A variable-ratio stopped-flow NMR/IR approach (Figure 3) has been developed to facilitate time- and material-efficient analysis of a wide range of initiator (M+X–) and reactant concentrations. Change of reagent from TMSCF3 (1a) to TIPSCF3 (1c) has a profound impact on the reaction. For example, the conversion of 4-F-acetophenone (2, 0.4 M) to 3OTMS by equimolar 1a in THF at ambient temperature takes <125 ms to complete using 0.1 mol % KOPh initiator and generates <2% of silylenol ether 4OTMS, whereas with TIPSCF3 (1c) and 3.75 mol % KOPh, the reaction proceeds to just 60% conversion in 16 h and generates 50% 4OTIPS. The rates of reaction are strongly affected by traces of inhibitors present in the reagents (1), especially at the low concentrations of initiator (M+X–) employed for the fastest reacting systems; see eqs 1 and 2.59,60 Nonetheless, while these render misleading initial rate data, study of the full reaction time-course, e.g., Figure 8, provides a coherent kinetics analysis.
A unified mechanism (V) for the reaction of R3SiCF3 reagents (1a–c) with ketones and aldehydes under conditions of catalytic anionic initiator (M+X–) is presented in Figure 7. The work confirms that the carbanion36−38 mechanism prevails under conditions of application (Scheme 1). Mechanism V allows a number of initially confusing observations to be rationalized. The main difference between use of TMSCF3 (1a) versus TIPSCF3 (1c) reagents is an inversion in the major anion speciation in the overall anionic chain reaction. This inversion leads to opposing influences of electrophile and silicon reagent (mechanisms Vi and Vii) and to keto–enol equilibration (2/2enol) with 1c (Vii). When TBAT is used as initiator,75 TESCF3 (1b) effects the most rapid trifluoromethylation in the series 1a–c. The increased steric bulk in 1b reduces reagent inhibition (K3) relative to 1a, without the substantial kinetic penalty in k2 experienced by 1c. These factors shift the reaction with 1b closer to an “ideal” catalytic cycle in which the intermediates are all connected by low TS barriers, with reduced off-cycle speciation. A consequence of adding TMSCF3 (1a) to TESCF3 (1b) is therefore to strongly inhibit turnover of 1b until all of 1a has been consumed, Figure 1.
The overarching mechanism (V, Figure 7) for anion-initiated reactions of R3SiCF3 (1) with ketones and aldehydes should prove of utility in their application in synthesis. For example, in the context of the design and analysis of enantioselective trifluoromethylation processes,16,29,30a,31 mechanism V shows that control must be achieved by the CF3–/[M]+ ion pair, Figure 9ii, and not by a siliconate intermediate. Moreover, the key mechanistic features of the anion-initiated reactions of 1 with carbonyl compounds (Figure 7) translate to reactions of 1 with other electrophiles (E)29−31 and proton donors (R–H to generate R–),73Figure 10. Thus, all processes in which siliconate D or analogous species formally acts as a nucleophilic or basic source of CF3 must proceed via a dissociative pathway (Figure 10ii). Siliconate D is inherently unstable and decomposes at ambient temperature to generate, interalia, complex perfluorocarbanions.34,38a The rate of anionic chain transfer, as dictated by the reactivity of the electrophile (E)29−31 or carbon acid (R–H)73 toward CF3–, as well as the presence of species able to attenuate decomposition (e.g., via CF2 capture, 4OSi →10, Scheme 3), controls the formal lifetime of D and in turn the minimum loading of initiator (M+X–) that will be required to achieve complete conversion of substrate. Moreover, traces of exogenous inhibitor(s) (e.g., Z–LG, Figure 10iii) ubiquitous in R3SiCF3 reagents (1) act to reduce the net active anion in the chain reaction, again increasing the requisite loading of initiator (M+X–). Compounds employed in synthetic routes to reagents 1a–c, e.g., TMSCl and TIPSCl,11 function as powerful inhibitors. However, the identity and effect of the inhibitors in reagents 1a–c vary substantially from batch to batch and between commercial suppliers (see SI). Electrophiles or carbon acids (R–H) that react with CF3– to ultimately generate an anion of inherently low reactivity toward 1 require a stoichiometric initiator to proceed to completion.14−26
Acknowledgments
The research leading to these results has received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007–2013)/ERC grant agreement no. 340163. The Carnegie Trust provided a collaborative research grant. C.P.J. thanks the EC for an International Outgoing Fellowship (PIOF-GA-2013-627695). We thank Veronica Forcina (Edinburgh, UK) and Prof. Dusan Uhrin (Edinburgh, UK) for assistance with stopped-flow NMR and dynamics and Dr. Per-Ola Norrby (AstraZeneca, Sweden), Prof. Robert Mulvey (Strathclyde, UK) and Dr. Mark Crimmin (Imperial, UK) for valuable mechanistic discussions in the early phases of this work.
Supporting Information Available
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.8b06777.
Additional discussion, experimental procedures, further kinetic data and analysis, characterization data, and NMR spectra (PDF)
Author Contributions
⊥ C. P. Johnston and T. H. West contributed equally to this work.
The authors declare no competing financial interest.
Supplementary Material
References
- Selected reviews:; a Purser S.; Moore P. R.; Swallow S.; Gouverneur V. Fluorine in medicinal chemistry. Chem. Soc. Rev. 2008, 37, 320–330. 10.1039/B610213C. [DOI] [PubMed] [Google Scholar]; b Gillis E. P.; Eastman K. J.; Hill M. D.; Donnelly D. J.; Meanwell N. A. Applications of Fluorine in Medicinal Chemistry. J. Med. Chem. 2015, 58, 8315–8359. 10.1021/acs.jmedchem.5b00258. [DOI] [PubMed] [Google Scholar]; c Zhou Y.; Wang J.; Gu Z.; Wang S.; Zhu W.; Aceña J. L.; Soloshonok V. A.; Izawa K.; Liu H. Next Generation of Fluorine-Containing Pharmaceuticals, Compounds Currently in Phase II–III Clinical Trials of Major Pharmaceutical Companies: New Structural Trends and Therapeutic Areas. Chem. Rev. 2016, 116, 422–518. 10.1021/acs.chemrev.5b00392. [DOI] [PubMed] [Google Scholar]
- Selected recent reviews:; a Fujiwara T.; O’Hagan D. Successful fluorine-containing herbicide agrochemicals. J. Fluorine Chem. 2014, 167, 16–29. 10.1016/j.jfluchem.2014.06.014. [DOI] [Google Scholar]; b Jeschke P. The unique role of halogen substituents in the design of modern agrochemicals. Pest Manage. Sci. 2010, 66, 10–27. 10.1002/ps.1829. [DOI] [PubMed] [Google Scholar]
- Babudri F.; Farinola G. M.; Naso F.; Ragni R. Fluorinated organic materials for electronic and optoelectronic applications: the role of the fluorine atom. Chem. Commun. 2007, 1003–1022. 10.1039/B611336B. [DOI] [PubMed] [Google Scholar]
- Berger R.; Resnati G.; Metrangolo P.; Weber E.; Hulliger J. Organic fluorine compounds: a great opportunity for enhanced materials properties. Chem. Soc. Rev. 2011, 40, 3496–3508. 10.1039/c0cs00221f. [DOI] [PubMed] [Google Scholar]
- Fluorinated Polymers (Vols. 1 and 2); Ameduri B.; Sawada H., Eds.; Royal Society of Chemistry: Cambridge, U.K., 2017. [Google Scholar]
- Ni C.; Hu J. The unique fluorine effects in organic reactions: recent facts and insights into fluoroalkylations. Chem. Soc. Rev. 2016, 45, 5441–5454. 10.1039/C6CS00351F. [DOI] [PubMed] [Google Scholar]
- Vincent J. M. Fluorous Catalysis: From the Origin to Recent Advances. Top. Curr. Chem. 2011, 308, 153–174. 10.1007/128_2011_241. [DOI] [PubMed] [Google Scholar]
- Langlois B. R.; Billard T.; Roussel S. Nucleophilic trifluoromethylation: Some recent reagents and their stereoselective aspects. J. Fluorine Chem. 2005, 126, 173–179. 10.1016/j.jfluchem.2004.11.007. [DOI] [Google Scholar]
- See for example:; a Zhang Y.; Fujiu M.; Serizawa H.; Mikami K. Organocatalysis approach to trifluoromethylation with fluoroform. J. Fluorine Chem. 2013, 156, 367–371. 10.1016/j.jfluchem.2013.07.018. [DOI] [Google Scholar]; b Musio B.; Gala E.; Ley S. V. Real-Time Spectroscopic Analysis Enabling Quantitative and Safe Consumption of Fluoroform during Nucleophilic Trifluoromethylation in Flow. ACS Sustainable Chem. Eng. 2018, 6, 1489–1495. 10.1021/acssuschemeng.7b04012. [DOI] [Google Scholar]
- a Geri J. B.; Szymczak N. K. Recyclable Trifluoromethylation Reagents from Fluoroform. J. Am. Chem. Soc. 2017, 139, 9811–9814. 10.1021/jacs.7b05408. [DOI] [PubMed] [Google Scholar]; b Geri J. B.; Wade Wolfe M. M.; Szymczak N. K. Borazine-CF3–Adducts for Rapid, Room Temperature, and Broad Scope Trifluoromethylation. Angew. Chem., Int. Ed. 2018, 57, 1381–1385. 10.1002/anie.201711316. [DOI] [PubMed] [Google Scholar]
- a Ruppert I.; Schlich K.; Volbach W.; Die Ersten CF3-Substituierten Organyl(Chlor)Silane. Tetrahedron Lett. 1984, 25, 2195–2198. 10.1016/S0040-4039(01)80208-2. [DOI] [Google Scholar]; b Ramaiah P.; Krishnamurti R.; Prakash G. K. S. 1-Trifluoromethyl-1-cyclohexanol. Org. Synth. 1995, 72, 232–236. 10.1002/0471264180.os072.28. [DOI] [Google Scholar]; c Prakash G. K. S.; Jog P. V.; Batamack P. T. D.; Olah G. A. Taming of Fluoroform: Direct Nucleophilic Trifluoromethylation of Si, B, S, and C Centers. Science 2012, 338, 1324–1327. 10.1126/science.1227859. [DOI] [PubMed] [Google Scholar]; d Pawelke G. Tetrakis(dimethylamino)ethylene/trifluoroiodomethane. a Specific Novel Trifluoromethylating agent. J. Fluorine Chem. 1989, 42, 429–433. 10.1016/S0022-1139(00)83934-2. [DOI] [Google Scholar]; e Eaborn C.; Griffiths R. W.; Pidcock A. Further Studies on Reactions of Organic Halides With Disilanes Catalysed by Transition Metal Complexes. J. Organomet. Chem. 1982, 225, 331–341. 10.1016/S0022-328X(00)86833-3. [DOI] [Google Scholar]
- a Kruse A.; Siegemund G.; Schumann A.; Ruppert I.. A process for the production of perfluoroalkyl compounds, and the pentafluoroethyl-trimethylsilane. German Pat. DE3805534, 1989; priority date February 23, 1988.
- a Prakash G. K. S.; Krishnamurti R.; Olah G. A. Synthetic methods and reactions. 141. Fluoride-induced trifluoromethylation of carbonyl compounds with trifluoromethyltrimethylsilane (TMS-CF3). A trifluoromethide equivalent. J. Am. Chem. Soc. 1989, 111, 393–395. 10.1021/ja00183a073. [DOI] [Google Scholar]; See also:; b Stahly G. P.; Bell D. R. A New Method for Synthesis of Trifluoromethyl-Substituted Phenols and Anilines. J. Org. Chem. 1989, 54, 2873–2877. 10.1021/jo00273a020. [DOI] [Google Scholar]; c Dubuffet T.; Sauvêtre R.; Normant J. F. Reaction Des Fluorosilyloxirannes Avec Les Electrophiles En Presence D’une Quantite Catalytique D’ion Fluorure. Tetrahedron Lett. 1988, 29, 5923–5924. 10.1016/S0040-4039(00)82227-3. [DOI] [Google Scholar]; For analogous fluoride-initiated nucleophilic transfer of CF2H, see:; d Chen D.; Ni C.; Zhao Y.; Cai X.; Li X.; Xiao P.; Hu J. Bis(difluoromethyl)trimethylsilicate Anion: A Key Intermediate in Nucleophilic Difluoromethylation of Enolizable Ketones with Me3SiCF2H. Angew. Chem., Int. Ed. 2016, 55, 12632–12636. 10.1002/anie.201605280. [DOI] [PubMed] [Google Scholar]
- a Prakash G. K. S.; Yudin A. K. Perfluoroalkylation with Organosilicon Reagents. Chem. Rev. 1997, 97, 757–786. 10.1021/cr9408991. [DOI] [PubMed] [Google Scholar]; b Singh R. P.; Shreeve J. M. Nucleophilic Trifluoromethylation Reactions of Organic Compounds with (Trifluoromethyl)trimethylsilane. Tetrahedron 2000, 56, 7613–7632. 10.1016/S0040-4020(00)00550-0. [DOI] [PubMed] [Google Scholar]; c Gawronski J.; Wascinska N.; Gajewy J. Recent Progress in Lewis Base Activation and Control of Stereoselectivity in the Additions of Trimethylsilyl Nucleophiles. Chem. Rev. 2008, 108, 5227–5252. 10.1021/cr800421c. [DOI] [PubMed] [Google Scholar]
- a Singh R. P.; Cao G.; Kirchmeier R. L.; Shreeve J. M. Cesium Fluoride Catalyzed Trifluoromethylation of Esters, Aldehydes, and Ketones with (Trifluoromethyl)trimethylsilane. J. Org. Chem. 1999, 64, 2873–2876. 10.1021/jo982494c. [DOI] [PubMed] [Google Scholar]; b Krishnamturi R.; Bellew D. R.; Prakash G. K. S. Preparation of trifluoromethyl and other perfluoroalkyl compounds with (perfluoroalkyl)trimethylsilanes. J. Org. Chem. 1991, 56, 984–989. 10.1021/jo00003a017. [DOI] [Google Scholar]; c Prakash G. K. S.; Panja C.; Vaghoo H.; Surampudi V.; Kultyshev R.; Mandal M.; Rasul G.; Mathew T.; Olah G. A. Facile Synthesis of TMS-Protected Trifluoromethylated Alcohols Using Trifluoromethyltrimethylsilane (TMSCF3) and Various Nucleophilic Catalysts in DMF. J. Org. Chem. 2006, 71, 6806–6813. 10.1021/jo060835d. [DOI] [PubMed] [Google Scholar]
- a Mizuta S.; Shibata N.; Akiti S.; Fujimoto H.; Nakamura S.; Toru T. Cinchona Alkaloids/TMAF Combination-Catalyzed Nucleophilic Enantioselective Trifluoromethylation of Aryl Ketones. Org. Lett. 2007, 9, 3707–3710. 10.1021/ol701791r. [DOI] [PubMed] [Google Scholar]; b Mizuta S.; Shibata N.; Hibino M.; Nagano S.; Nakamura S.; Toru T. Ammonium bromides/KF catalyzed trifluoromethylation of carbonyl compounds with (trifluoromethyl)trimethylsilane and its application in the enantioselective trifluoromethylation reaction. Tetrahedron 2007, 63, 8521–8528. 10.1016/j.tet.2007.05.024. [DOI] [Google Scholar]; c Nagao H.; Kawano Y.; Mukaiyama T. Enantioselective Trifluoromethylation of Ketones with (Trifluoromethyl)trimethylsilane Catalyzed by Chiral Quaternary Ammonium Phenoxides. Bull. Chem. Soc. Jpn. 2007, 80, 2406–2412. 10.1246/bcsj.80.2406. [DOI] [Google Scholar]; d Hu X.; Wang J.; Wei L.; Lin L.; Liu X.; Feng X. Cinchona alkaloid-derived quaternary ammonium salt combined with NaH: a facile catalyst system for the asymmetric trifluoromethylation of ketones. Tetrahedron Lett. 2009, 50, 4378–4380. 10.1016/j.tetlet.2009.05.041. [DOI] [Google Scholar]; e Zhao H.; Qin B.; Liu X.; Feng X. Enantioselective trifluoromethylation of aromatic aldehydes catalyzed by combinatorial catalysts. Tetrahedron 2007, 63, 2682–2686. 10.1016/j.tet.2007.04.068. [DOI] [Google Scholar]; f Caron S.; Do N. M.; Arpin P.; Larivee A. Enantioselective Addition of a Trifluoromethyl Anion to Aryl Ketones and Aldehydes. Synthesis 2003, 2003, 1693–1698. 10.1055/s-2003-40889. [DOI] [Google Scholar]; g Kawai H.; Mizuta S.; Tokunaga E.; Shibata N. Cinchona alkaloid/TMAF combination: Enantioselective trifluoromethylation of aryl aldehydes. J. Fluorine Chem. 2013, 152, 46–50. 10.1016/j.jfluchem.2013.01.032. [DOI] [Google Scholar]; h Okusu S.; Hirano K.; Yasuda Y.; Tokunaga E.; Shibata N. Flow Trifluoromethylation of Carbonyl Compounds by Ruppert–Prakash Reagent and its Application for Pharmaceuticals, Efavirenz and HSD-016. RSC Adv. 2016, 6, 82716–82720. 10.1039/C6RA19790F. [DOI] [Google Scholar]
- For example, a substructure search on 14/06/2018 for the addition of TMSCF3 (1a, exact reagent) to C=O (any carbonyl) → C(OA)-CF3 (A = any atom) using Reaxys/SciFinder gave 434/450 primary research papers and 631/811 patents.
- Dilman A. D.; Levin V. V. Nucleophilic Trifluoromethylation of C=N Bonds. Eur. J. Org. Chem. 2011, 2011, 831–841. 10.1002/ejoc.201001558. [DOI] [Google Scholar]
- a Jin G.; Zhang X.; Fu D.; Dai W.; Cao S. Mild and metal-free trifluoromethylation and pentafluoroethylation of gem-difluoroalkenes with TMSCF3 and TMSCF2CF3. Tetrahedron 2015, 71, 7892–7899. 10.1016/j.tet.2015.08.016. [DOI] [Google Scholar]; b Ono T. Synthesis of highly branched perfluoroolefins that are super-congested via multi-substitution of trifluoromethyl groups: Trifluoromethylation of hexafluoropropene trimers with Ruppert-Prakash reagent. J. Fluorine Chem. 2017, 196, 128–134. 10.1016/j.jfluchem.2016.12.001. [DOI] [Google Scholar]
- a Furin G. G.; Bardin V. V. (Polyfluoroorganyl) trimethylsilanes in syntheses of fluoroorganic compounds. J. Fluorine Chem. 1991, 54, 241–241. 10.1016/S0022-1139(00)83751-3. [DOI] [Google Scholar]; b Bardin V. V.; Kolomeitsev A. A.; Furin G. G.; Yagupol’skii Y. L. Trifluoromethylation of perfluoroaromatic compounds using trifluoromethyltrimethylsilane in the presence of fluoride ions. Bull. Acad. Sci. USSR, Div. Chem. Sci. 1990, 39, 1539–1539. 10.1007/BF00957892. [DOI] [Google Scholar]
- a Kolomeitsev A. A.; Movchun V. N.; Kondratenko N. V.; Yagupolski Y. L. A Convenient Route to Aryl Trifluoromethyl Sulfones by Fluoride-Catalyzed Cross-Coupling of Arenesulfonyl Fluorides with (Trifluoromethyl)trimethylsilane and (Trifluoromethyl)trimethylstannane. Synthesis 1990, 1990, 1151–1152. 10.1055/s-1990-27121. [DOI] [Google Scholar]; b Kowalczyk R.; Edmunds A. J.; Hall R. G.; Bolm C. Synthesis of CF3-Substituted Sulfoximines from Sulfonimidoyl Fluorides. Org. Lett. 2011, 13, 768–771. 10.1021/ol103030w. [DOI] [PubMed] [Google Scholar]; c Gouault-Bironneau S.; Timoshenko V. M.; Grellepois F.; Portella C. Thiophilic nucleophilic trifluoromethylation of α-substituted dithioesters. Access to S-trifluoromethyl ketene dithioacetals and their reactivity with electrophilic species. J. Fluorine Chem. 2012, 134, 164–171. 10.1016/j.jfluchem.2011.02.016. [DOI] [Google Scholar]; d Timoshenko V. M.; Portella C. Domino nucleophilic trifluoromethylations of alkyl perfluorodithioesters. J. Fluorine Chem. 2009, 130, 586–590. 10.1016/j.jfluchem.2009.04.004. [DOI] [Google Scholar]; e Yagupolskii L. M.; Matsnev A. V.; Orlova R. K.; Deryabkin B. G.; Yagupolskii Y. L. A new method for the synthesis of trifluoromethylating agents—Diaryltrifluoromethylsulfonium salts. J. Fluorine Chem. 2008, 129, 131–136. 10.1016/j.jfluchem.2007.10.001. [DOI] [Google Scholar]
- a Billard T.; Langlois B. R.; Large S. Synthesis of Trifluoromethyl Selenides. Phosphorus, Sulfur Silicon Relat. Elem. 1998, 136, 521–524. 10.1080/10426509808545987. [DOI] [Google Scholar]; b Tyrra W.; Naumann D.; Yagupolskii Y. L. Stable trifluoromethylselenates(0), [A]SeCF3—synthesis, characterizations and properties. J. Fluorine Chem. 2003, 123, 183–187. 10.1016/S0022-1139(03)00118-0. [DOI] [Google Scholar]
- a Tworowska I.; Dabkowski W.; Michalski J. Synthesis of Tri- and Tetracoordinate Phosphorus Compounds Containing a PCF3 Group by Nucleophilic Trifluoromethylation of the Corresponding PF Compounds. Angew. Chem., Int. Ed. 2001, 40, 2898–2900. . [DOI] [PubMed] [Google Scholar]; b Pane P.; Naumann D.; Hoge B. Cyanide initiated perfluoroorganylations with perfluoroorgano silicon compounds. J. Fluorine Chem. 2001, 112, 283–286. 10.1016/S0022-1139(01)00513-9. [DOI] [Google Scholar]; c Murphy-Jolly M. B.; Lewis L. C.; Caffyn A. J. M. The synthesis of tris(perfluoroalkyl)phosphines. Chem. Commun. 2005, 4479–4480. 10.1039/b507752d. [DOI] [PubMed] [Google Scholar]
- a Molander G. A.; Hoag B. P. Improved Synthesis of Potassium (Trifluoromethyl)trifluoroborate [K(CF3BF3)]. Organometallics 2003, 22, 3313–3315. 10.1021/om0302645. [DOI] [Google Scholar]; b Kolomeitsev A. A.; Kadyrov A. A.; Szczepkowska-Sztolcman J.; Milewska M.; Koroniak H.; Bissky G.; Barten J. A.; Röschenthaler G.-V. Perfluoroalkyl borates and boronic esters: new promising partners for Suzuki and Petasis reactions. Tetrahedron Lett. 2003, 44, 8273–8277. 10.1016/j.tetlet.2003.09.072. [DOI] [Google Scholar]
- Matousek V.; Pietrasiak E.; Schwenk R.; Togni A. One-Pot Synthesis of Hypervalent Iodine Reagents for Electrophilic Trifluoromethylation. J. Org. Chem. 2013, 78, 6763–6768. 10.1021/jo400774u. [DOI] [PubMed] [Google Scholar]
- Tyrra W.; Naumann D.; Kirij N. V.; Kolomeitsev A. A.; Yagupolskii Y. L. The first alkyl bismuthates: tris(trifluoromethyl)fluoro- and tetrakis(trifluoromethyl)-bismuthate. J. Chem. Soc., Dalton Trans. 1999, 657–658. 10.1039/a900308h. [DOI] [Google Scholar]
- a Prakash G. K. S.; Krishnamoorthy S.; Kar S.; Olah G. A. Direct S-difluoromethylation of thiols using the Ruppert–Prakash reagent. J. Fluorine Chem. 2015, 180, 186–191. 10.1016/j.jfluchem.2015.09.011. [DOI] [Google Scholar]; b Hashimoto R.; Iida T.; Aikawa K.; Ito S.; Mikami K. Direct α-Siladifluoromethylation of Lithium Enolates with Ruppert-Prakash Reagent via C–F Bond Activation. Chem. - Eur. J. 2014, 20, 2750–2754. 10.1002/chem.201304473. [DOI] [PubMed] [Google Scholar]; c Ito S.; Kato N.; Mikami K. Stable (sila)difluoromethylboranes via C–F activation of fluoroform derivatives. Chem. Commun. 2017, 53, 5546–5548. 10.1039/C7CC02327H. [DOI] [PubMed] [Google Scholar]
- Ni C.; Hu J. Recent Advances in the Synthetic Application of Difluorocarbene. Synthesis 2014, 46, 842–863. 10.1055/s-0033-1340856. [DOI] [Google Scholar]
- a Liu X.; Xu C.; Wang M.; Liu Q. Trifluoromethyltrimethylsilane Nucleophilic Trifluoromethylation and Beyond. Chem. Rev. 2015, 115, 683–730. 10.1021/cr400473a. [DOI] [PubMed] [Google Scholar]; b Takeshi Komiyama T.; Minami Y.; Hiyama T. Recent Advances in Transition-Metal-Catalyzed Synthetic Transformations of Organosilicon Reagents. ACS Catal. 2017, 7, 631–651. 10.1021/acscatal.6b02374. [DOI] [Google Scholar]
- a Denmark S. E.; Buetner G. L. Lewis Base Catalysis in Organic Synthesis. Angew. Chem., Int. Ed. 2008, 47, 1560–1638. 10.1002/anie.200604943. [DOI] [PubMed] [Google Scholar]; b Reich H. J.Mechanism of C-Si Bond Cleavage Using Lewis Bases (n→σ*). In Lewis Base Catalysis in Organic Synthesis, 1st ed.; Vedejs E.; Denmark S. E., Eds.; Wiley VCH: Weinheim, Germany, 2016; Vol. 1, pp 251–253. [Google Scholar]
- For leading references see:; a Yang X.; Wu T.; Phipps R. J.; Toste F. D. Advances in Catalytic Enantioselective Fluorination, Mono-, Di-, and Trifluoromethylation, and Trifluoromethylthiolation Reactions. Chem. Rev. 2015, 115, 826–870. 10.1021/cr500277b. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Shibata N.; Mizuta S.; Kawai H. Recent advances in enantioselective trifluoromethylation reactions. Tetrahedron: Asymmetry 2008, 19, 2633–2644. 10.1016/j.tetasy.2008.11.011. [DOI] [Google Scholar]
- Maggiarosa N.; Tyrra W.; Naumann D.; Kirij N. V.; Yagupolskii Y. L. Me3Si(CF3)F]− and [Me3Si(CF3) 2]−: Reactive Intermediates in Fluoride-Initiated Trifluoromethylation with Me3SiCF3 – An NMR Study. Angew. Chem., Int. Ed. 1999, 38, 2252–2253. . [DOI] [PubMed] [Google Scholar]
- Kolomeitsev A.; Bissky G.; Lork E.; Movchun V.; Rusanov E.; Kirsch P.; Röschenthaler G.-V. Different fluoride anion sources and (trifluoromethyl)trimethylsilane: molecular structure of tris(dimethylamino)sulfonium bis(trifluoromethyl)trimethylsiliconate, the first isolated pentacoordinate silicon species with five Si–C bonds. Chem. Commun. 1999, 1017–1018. 10.1039/a901953g. [DOI] [Google Scholar]
- Tyrra W.; Kremlev M. M.; Naumann D.; Scherer H.; Schmidt H.; Hoge B.; Pantenburg I.; Yagupolskii Y. L. How Trimethyl(trifluoromethyl)silane Reacts with Itself in the Presence of Naked Fluoride—A One-Pot Synthesis of Bis([15]crown-5)cesium 1,1,1,3,5,5,5-Heptafluoro-2,4-bis(trifluoromethyl)pentenide. Chem. - Eur. J. 2005, 11, 6514–6518. 10.1002/chem.200500799. [DOI] [PubMed] [Google Scholar]
- Kotun S. P.; Anderson J. D. O.; DesMarteau D. D. Fluorinated Tertiary Alcohols and Alkoxides from Nucleophilic Trifluoromethylation of Carbonyl Compounds. J. Org. Chem. 1992, 57, 1124–1131. 10.1021/jo00030a018. [DOI] [Google Scholar]
- Prakash G. K. S.; Wang F.; Zhang Z.; Haiges R.; Rahm M.; Christe K. O.; Mathew T.; Olah G. A. Long-Lived Trifluoromethanide Anion: A Key Intermediate in Nucleophilic Trifluoromethylations. Angew. Chem., Int. Ed. 2014, 53, 11575–11578. 10.1002/anie.201406505. [DOI] [PubMed] [Google Scholar]
- Santschi N.; Gilmour R. The (Not So) Ephemeral Trifluoromethanide Anion. Angew. Chem., Int. Ed. 2014, 53, 11414–11415. 10.1002/anie.201408013. [DOI] [PubMed] [Google Scholar]
- a Lishchynskyi A.; Miloserdov F. M.; Martin E.; Benet-Buchholz J.; Escudero-Adán E. C.; Konovalov A. I.; Grushin V. V. The Trifluoromethyl Anion. Angew. Chem., Int. Ed. 2015, 54, 15289–15293. 10.1002/anie.201507356. [DOI] [PubMed] [Google Scholar]; b Miloserdov F. M.; Konovalov A. I.; Martin E.; Benet-Buchholz J.; Escudero-Adán E. C.; Lishchynskyi A.; Grushin V. V. The Trifluoromethyl Anion: Evidence for [K(crypt-222)]+CF3–. Helv. Chim. Acta 2017, 100, e1700032. 10.1002/hlca.201700032. [DOI] [PubMed] [Google Scholar]; c Harlow R. L.; Benet-Buchholz J.; Miloserdov F. M.; Konovalov A. I.; Marshall W. J.; Martin E.; Benet-Buchholz J.; Escudero-Adán E. C.; Martin E.; Lishchynskyi A.; Grushin V. V. On the Structure of [K(crypt-222)]+CF3–. Helv. Chim. Acta 2018, 101, e1800015. 10.1002/hlca.201800015. [DOI] [Google Scholar]
- The X-ray structural analysis of E (ref (38a)) was challenged:Becker S.; Müller P. On the Crystal Structure Analysis of [K(crypt-222)]+CF3- a Reinterpretation: No Proof for the Trifluoromethanide Ion. Chem. - Eur. J. 2017, 23, 7081–7086. 10.1002/chem.201700554. [DOI] [PubMed] [Google Scholar]; The original interpretation was comprehensively defended; see ref (38c).
- The CF3H contained traces of CF3D (0.1% 19F NMR) likely from adventitious DOH/D2O in the d8-THF.
- Song X.; Chang J.; Zhu D.; Li J.; Xu C.; Liu Q.; Wang M. Catalytic Domino Reaction of Ketones/Aldehydes with Me3SiCF2Br for the Synthesis of α-Fluoroenones/α-Fluoroenals. Org. Lett. 2015, 17, 1712–1715. 10.1021/acs.orglett.5b00488. [DOI] [PubMed] [Google Scholar]
- Song X.; Xu C.; Du D.; Zhao Z.; Zhu D.; Wang M. Ring-Opening Diarylation of Siloxydifluorocyclopropanes by Ag(I) Catalysis: Stereoselective Construction of 2-Fluoroallylic Scaffold. Org. Lett. 2017, 19, 6542–6545. 10.1021/acs.orglett.7b03254. [DOI] [PubMed] [Google Scholar]
- Pilcher A. S.; DeShong P. Utilization of Tetrabutylammonium Triphenyldifluorosilicate as a Fluoride Source for Silicon–Carbon Bond Cleavage. J. Org. Chem. 1996, 61, 6901–6905. 10.1021/jo960922d. [DOI] [PubMed] [Google Scholar]
- Water may be liberated from the NMR tube surface, the reagents (1a, 2; freshly distilled), the TBAT (anhydrous solid; THF solutions prepared and stored in a glovebox), or the THF (∼8 ppm H2O, Karl Fischer titration).
- Turnover rates were affected only by the impact on TMSCF3/ketone ratio resulting from the rapid prior consumption of TMSCF3 by the H2O.
- For a review of CF3 transfer involving CF3 radicals, see:Studer A. A ’Renaissance’ in Radical Trifluoromethylation. Angew. Chem., Int. Ed. 2012, 51, 8950–8958. 10.1002/anie.201202624. [DOI] [PubMed] [Google Scholar]
- For an example of CF3 radical character transfer from TMSCF3 via AgCF3 intermediates, see:Ye Y.; Lee S. H.; Sanford M. S. Silver-Mediated Trifluoromethylation of Arenes Using TMSCF3. Org. Lett. 2011, 13, 5464–5467. and references therein 10.1021/ol202174a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Competing 1,4- and 1,6-addition to the aryl ring are characteristic of SET-type mechanisms for nucleophile addition to benzophenones; see:Holm T.; Crossland I. Mechanism of the Grignard Addition Reaction.. Acta Chem. Scand. 1971, 25, 59–69. and references therein 10.3891/acta.chem.scand.25-0059. [DOI] [Google Scholar]
- A near-continuum sequential SET pathway is possible; see e.g.:Eisch J. J. Single Electron Transfers in the Reactions of Carbanions. Res. Chem. Intermed. 1996, 22, 145–187. 10.1163/156856796X00593. [DOI] [Google Scholar]
- a Newcombe M.; Johnson C. C.; Manek M. B.; Varick T. R. Picosecond radical kinetics. Ring openings of phenyl-substituted cyclopropylcarbinyl radicals. J. Am. Chem. Soc. 1992, 114, 10915–10921. 10.1021/ja00053a031. [DOI] [Google Scholar]; b Miyazoe H.; Yamago S.; Yoshida J. Novel Group-Transfer Three-Component Coupling of Silyltellurides, Carbonyl Compounds, and Isocyanides. Angew. Chem., Int. Ed. 2000, 39, 3669–3671. . [DOI] [PubMed] [Google Scholar]
- Biphenyl is significantly more stabilizing for radical pathways than 4-fluorophenyl (σC•, +0.46 and −0.06 respectively), see:Creary X. Super Radical Stabilizers. Acc. Chem. Res. 2006, 39, 761–771. 10.1021/ar0680724. [DOI] [PubMed] [Google Scholar]
- For example, BuN+X– where X– = Ph3SiF2–, F–, HO–, PhO–, AcO–, and 3O–, behaved identically within experimental error. However when X– was less nucleophilic, e.g., BzO–, or 3,5-(CF3)C6H3O–, significant induction periods were evident. Induction periods were extreme with KOC(CF3)3.
- Luo G.; Luo Y.; Qu J. Direct nucleophilic trifluoromethylation using fluoroform: a theoretical mechanistic investigation and insight into the effect of alkali metal cations. New J. Chem. 2013, 37, 3274–3280. 10.1039/c3nj00686g. [DOI] [Google Scholar]
- Control experiments confirmed that TBAT induces negligible H/D exchange between 2 and D3-2 over the reaction period, whereas 3OK induces complete scrambling in under 95 s.
- For examples of mechanical variable ratio mixing for stopped-flow kinetics see:Goez M. Inexpensive Pre-Mixer For Stopped-Flow Apparatus. J. Phys. E: Sci. Instrum. 1988, 21, 440–442. and references therein 10.1088/0022-3735/21/5/003. [DOI] [Google Scholar]
- For recent developments see:; a Dunn A. L.; Landis C. R. Progress toward reaction monitoring at variable temperatures: a new stopped-flow NMR probe design. Magn. Reson. Chem. 2017, 55, 329–336. 10.1002/mrc.4538. [DOI] [PubMed] [Google Scholar]; b Thomas A. A.; Denmark S. E.; Ernest L.. Eliel, a Physical Organic Chemist with the Right Tool for the Job: Rapid Injection Nuclear Magnetic Resonance. In Stereochemistry and Global Connectivity: The Legacy of Ernest L. Eliel, Cheng H. N., Ed.; American Chemical Society: Washington DC, 2017; Vol. 2, pp 105–134. [Google Scholar]
- Cox P. A.; Reid M.; Leach A. G.; Campbell A. D.; King E. J.; Lloyd-Jones G. C. Base-Catalyzed Aryl-B(OH)2 Protodeboronation Revisited: From Concerted Proton Transfer to Liberation of a Transient Aryl Anion. J. Am. Chem. Soc. 2017, 139, 13156–13165. 10.1021/jacs.7b07444. [DOI] [PubMed] [Google Scholar]
- Foley D. A.; Bez E.; Codina A.; Colson K. L.; Fey M.; Krull R.; Piroli D.; Zell M. T.; Marquez B. L. NMR Flow Tube for Online NMR Reaction Monitoring. Anal. Chem. 2014, 86, 12008–12013. 10.1021/ac502300q. [DOI] [PubMed] [Google Scholar]
- Plots of initial rate versus [TBAT] for reaction of 2 and 13 with 1a both have a nonzero x-axis intercept (0.03 mM) with curvature evident at low [TBAT] concentrations, suggesting the inhibitor(s) are present at <0.02% 1a, in the specific batches of commercial reagents that were employed; see SI.
- In eqs 1 and 2 and elsewhere, (1 – xEI) represents the mole fraction of active anion relative to total anion [M+X–]0. Based on a 1:1 inhibition mode, xEI = (KEI[I]/(1 + KEI[I])) with [I]0 = xi[1]0, where xi = mole fraction inhibitor in reagent 1. Experimental data (see SI) suggest KEI is substantially greater with 1c.
- This contrasts with the borazine systems recently developed by Szymczak (see ref (10)), where addition of KBARF profoundly accelerates CF3-transfer rates.
- Simulations were conducted using the three-spin parametrization in WNDNMR, with the rates and frequencies as indicated in Figure 4 (kex = 2k–ex; where k–ex is the apparent 19F nuclei exchange rate through reassociation). The fraction of CF3 present as E was arbitrarily set to 0.1%, with the remaining 99.9% partitioned between D and 1a as required to fit. The chemical shift of CF3[M] is reported as δF −18.7 ppm (see ref (38), and −17.2 ppm, see ref (36)), depending on the identity of [M].
- Hariharan P. C.; Pople J. A. The influence of polarization functions on molecular orbital hydrogenation energies. Theor. Chim. Acta 1973, 28, 213–222. 10.1007/BF00533485. [DOI] [Google Scholar]
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; Nakatsuji H.; Caricato M.; Li X.; Hratchian H. P.; Izmaylov A. F.; Bloino J.; Zheng G.; Sonnenberg J. L.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M.; Heyd J. J.; Brothers E.; Kudin K. N.; Staroverov V. N.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Rega N.; Millam J. M.; Klene M.; Knox J. E.; Cross J. B.; Bakken V.; Adamo C.; Jaramillo J.; Gomperts R.; Stratmann R. E.; Yazyev O.; Austin A. J.; Cammi R.; Pomelli C.; Ochterski J. W.; Martin R. L.; Morokuma K.; Zakrzewski V. G.; Voth G. A.; Salvador P.; Dannenberg J. J.; Dapprich S.; Daniels A. D.; Farkas Ö.; Foresman J. B.; Ortiz J. V.; Cioslowski J.; Fox D. J.. Gaussian 09; Gaussian, Inc.: Wallingford, CT, 2009.
- a Zhao Y.; Truhlar D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. 10.1007/s00214-007-0310-x. [DOI] [Google Scholar]; b Zhao Y.; Truhlar D. G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem. Res. 2008, 41, 157–167. 10.1021/ar700111a. [DOI] [PubMed] [Google Scholar]; c Tomasi J.; Mennucci B.; Cammi R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. 10.1021/cr9904009. [DOI] [PubMed] [Google Scholar]
- a Brown S. B.; Dewar M. J. S.; Ford G. P.; Nelson D. J.; Rzepa H. S. Ground states of molecules. 51. MNDO (modified neglect of diatomic overlap) calculations of kinetic isotope effects. J. Am. Chem. Soc. 1978, 100, 7832–7836. 10.1021/ja00493a008. [DOI] [Google Scholar]; b Dewar M. J. S.; Olivella S.; Rzepa H. S. Ground states of molecules. 49. MINDO/3 study of the retro-Diels-Alder reaction of cyclohexene. J. Am. Chem. Soc. 1978, 100, 5650–5659. 10.1021/ja00486a013. [DOI] [Google Scholar]; c Rzepa H. S.KINISOT. A basic program to calculate kinetic isotope effects using normal coordinate analysis of transition state and reactants; http://doi.org/10.5281/zenodo.19272, 2015.; d Paton R. S.Kinisot: v 1.0.0 public API for Kinisot.py. http://doi.org/10.5281/zenodo.60082, 2016.
- Han H.; Alday B.; Shuman N. S.; Wiens J. P.; Troe J.; Viggiano A. A.; Guo H. Calculations of the active mode and energetic barrier to electron attachment to CF3 and comparison with kinetic modeling of experimental results. Phys. Chem. Chem. Phys. 2016, 18, 31064–31071. 10.1039/C6CP05867A. [DOI] [PubMed] [Google Scholar]
- McDonald R. N.; Chowdhury A. K. Nucleophilic reactions of trifluoromethyl ion (F3C-) at sp2 and sp3 carbon in the gas phase. Characterization of carbonyl addition adducts. J. Am. Chem. Soc. 1983, 105, 7267–7271. 10.1021/ja00363a010. [DOI] [Google Scholar]
- For a range of anion–ketone interactions, including H-bonded adducts and aldolate products, see:Kolonko K. J.; Reich H. J. Stabilization of Ketone and Aldehyde Enols by Formation of Hydrogen Bonds to Phosphazene Enolates and Their Aldol Products. J. Am. Chem. Soc. 2008, 130, 9668–9669. 10.1021/ja804221x. [DOI] [PubMed] [Google Scholar]; The lowest energy of these was an enolate-ketone H-bonded adduct (F).
- Thibblin A.; Jencks W. P. Unstable Carbanions. General Acid Catalysis of the Cleavage of 1-Phenylcyclopropanol and 1-Phenyl-2-arylcyclopropanol Anions. J. Am. Chem. Soc. 1979, 101, 4963–4973. 10.1021/ja00511a028. [DOI] [Google Scholar]
- The 2H-KIE for α-C–H deprotonation of d3-2 by LDA.LiOR is kH/kD = 6.3; with (LDA)2 it is 1.7:Kolonko K. J.; Wherrit D. J.; Reich H. J. Mechanistic Studies of the Lithium Enolate of 4-Fluoroacetophenone: Rapid-Injection NMR Study of Enolate Formation, Dynamics, and Aldol Reactivity. J. Am. Chem. Soc. 2011, 133, 16774–16777. 10.1021/ja207218f. [DOI] [PubMed] [Google Scholar]
- Chiang Y.; Kresge A. J.; Wirz J. Flash-Photolytic Generation of Acetophenone Enol. The Keto-Enol Equilibrium Constant and pKa of Acetophenone in Aqueous Solution. J. Am. Chem. Soc. 1984, 106, 6392–6395. 10.1021/ja00333a049. [DOI] [Google Scholar]
- a Sasaki M.; Kondo Y. Deprotonative C–H Silylation of Functionalized Arenes and Heteroarenes Using Trifluoromethyltrialkylsilane with Fluoride. Org. Lett. 2015, 17, 848–851. 10.1021/ol503671b. [DOI] [PubMed] [Google Scholar]; b Behr J.-B.; Chavaria D.; Plantier-Royon R. Trifluoromethide as a Strong Base: [CF3–] Mediates Dichloromethylation of Nitrones by Proton Abstraction from the Solvent. J. Org. Chem. 2013, 78, 11477–11482. 10.1021/jo402028r. [DOI] [PubMed] [Google Scholar]; c Adams D. J.; Clark J. H.; Hansen L. B.; Sanders V. C.; Tavener S. J. Reaction of Tetramethylammonium Fluoride with Trifluoromethyltrimethylsilane. J. Fluorine Chem. 1998, 92, 123–125. 10.1016/S0022-1139(98)00273-5. [DOI] [Google Scholar]; d Yoshimatsu M.; Kuribayashi M. A Novel Utilization of Trifluoromethanide as a Base: a Convenient Synthesis of Trimethylsilylacetylene. J. Chem. Soc., Perkin Trans. 1 2001, 1256–1257. 10.1039/b101537k. [DOI] [Google Scholar]; e Nozawa-Kumada K.; Osawa S.; Sasaki M.; Chataigner I.; Shigeno M.; Kondo Y. Deprotonative Silylation of Aromatic C–H Bonds Mediated by a Combination of Trifluoromethyltrialkylsilane and Fluoride. J. Org. Chem. 2017, 82, 9487–9496. 10.1021/acs.joc.7b01525. [DOI] [PubMed] [Google Scholar]
- Zefirov N. S.; Makhon’kov D. I. X-Philic Reactions. Chem. Rev. 1982, 82, 615–624. 10.1021/cr00052a004. [DOI] [Google Scholar]
- The kinetics for reactions of 2 (0.4 M) with 1b (0.48 M) are again complicated by inhibitors. When initiated by 0.3 mM KOPh, the reaction is pseudo-zero-order throughout; this result is consistent with a number of mechanisms, including for example turnover and inhibition by 1b or rate-limiting dissociation of CF3 from dominant anion Bb. The faster rates of reaction with Bu4N+ versus K+ as counterion suggest mechanism Vii dominates.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.