首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Transition metal-catalysed C–H bond functionalisations have been extensively developed in organic and medicinal chemistry. Among these catalytic approaches, the selective activation of C(sp3)–H and C(sp2)–H bonds is particularly appealing for its remarkable synthetic versatility, yet it remains highly challenging. Herein, we demonstrate the first example of temperature-dependent selective C–H functionalisation of unactivated C(sp3)–H or C(sp2)–H bonds at remote positions through palladium catalysis using 7-pyridyl-pyrazolo[1,5-a]pyrimidine as a new directing group. At 120 °C, C(sp3)–H arylation was triggered by the chelation of a rare [6,5]-fused palladacycle, whereas at 140 °C, C(sp2)–H arylation proceeded instead through the formation of a 16-membered tetramer containing four 7-pyridyl-pyrazolo[1,5-a]pyrimidine–palladium chelation units. The subsequent mechanistic study revealed that both C–H activations shared a common 6-membered palladacycle intermediate, which was then directly transformed to either the [6,5]-fused palladacycle for C(sp3)–H activation at 120 °C or the tetramer for C(sp2)–H arylation at 140 °C with catalytic amounts of Pd(OAc)2 and AcOH. Raising the temperature from 120 °C to 140 °C can also convert the [6,5]-fused palladacycle to the tetramer with the above-mentioned catalysts, hence completing the C(sp2)–H arylation ultimately.

Unprecedented 16-membered tetramer or [6,5]-fused palladacycle, mutually shadowboxing-like transformed from the shared common intermediate, accomplishes the Pd-catalysed temperature-dependent selective arylation of C(sp2)–H or C(sp3)–H.  相似文献   

2.
Manganaelectro-catalyzed azidation of otherwise inert C(sp3)–H bonds was accomplished using most user-friendly sodium azide as the nitrogen-source. The operationally simple, resource-economic C–H azidation strategy was characterized by mild reaction conditions, no directing group, traceless electrons as the sole redox-reagent, Earth-abundant manganese as the catalyst, high functional-group compatibility and high chemoselectivity, setting the stage for late-stage azidation of bioactive compounds. Detailed mechanistic studies by experiment, spectrophotometry and cyclic voltammetry provided strong support for metal-catalyzed aliphatic radical formation, along with subsequent azidyl radical transfer within a manganese(iii/iv) manifold.

The merger of manganese-catalyzed C–H functionalization with electrosynthesis enabled C(sp3)–H azidation devoid of chemical oxidants or photochemical irradiation. Detailed mechanistic studies are supportive of a manganese(iii/iv) electrocatalysis.  相似文献   

3.
High-valent metal oxo complexes are prototypical intermediates for the activation and hydroxylation of alkyl C–H bonds. Substituting the oxo ligand with other functional groups offers the opportunity for additional C–H functionalization beyond C–O bond formation. However, few species aside from metal oxo complexes have been reported to both activate and functionalize alkyl C–H bonds. We herein report the first example of an isolated copper(iii) cyanide complex (LCuIIICN) and its C–H cyanation reactivity. We found that the redox potential (Eox) of substrates, instead of C–H bond dissociation energy, is a key determinant of the rate of PCET, suggesting an oxidative asynchronous CPET or ETPT mechanism. Among substrates with the same BDEs, those with low redox potentials transfer H atoms up to a million-fold faster. Capitalizing on this mechanistic insight, we found that LCuIIICN is highly selective for cyanation of amines, which is predisposed to oxidative asynchronous or stepwise transfer of H+/e. Our study demonstrates that the asynchronous effect of PCET is an appealing tool for controlling the selectivity of C–H functionalization.

A formal copper(iii) cyanide complex and its C–H cyanation reactivity are reported. The redox potentials of substrates, instead of C–H bond dissociation energies, were found to be the key determinant of the rates of PCET.  相似文献   

4.
A catalytic system for intramolecular C(sp2)–H and C(sp3)–H amination of substituted tetrazolopyridines has been successfully developed. The amination reactions are developed using an iron-porphyrin based catalytic system. It has been demonstrated that the same iron-porphyrin based catalytic system efficiently activates both the C(sp2)–H and C(sp3)–H bonds of the tetrazole as well as azide-featuring substrates with a high level of regioselectivity. The method exhibited an excellent functional group tolerance. The method affords three different classes of high-value N-heterocyclic scaffolds. A number of important late-stage C–H aminations have been performed to access important classes of molecules. Detailed studies (experimental and computational) showed that both the C(sp2)–H and C(sp3)–H amination reactions involve a metalloradical activation mechanism, which is different from the previously reported electro-cyclization mechanism. Collectively, this study reports the discovery of a new class of metalloradical activation modes using a base metal catalyst that should find wide application in the context of medicinal chemistry, drug discovery and industrial applications.

A catalytic system for intramolecular C(sp2)–H and C(sp3)–H amination of substituted tetrazolopyridines has been successfully developed.  相似文献   

5.
Site-selective fluorination of aliphatic C–H bonds remains synthetically challenging. While directed C–H fluorination represents the most promising approach, the limited work conducted to date has enabled just a few functional groups as the arbiters of direction. Leveraging insights gained from both computations and experimentation, we enabled the use of the ubiquitous amine functional group as a handle for the directed C–H fluorination of Csp3–H bonds. By converting primary amines to adamantoyl-based fluoroamides, site-selective C–H fluorination proceeds under the influence of a simple iron catalyst in 20 minutes. Computational studies revealed a unique reaction coordinate for the catalytic process and offer an explanation for the high site selectivity.

By converting primary amines to adamantoyl-based fluoroamides, site-selective C–H fluorination proceeds under the influence of a simple iron catalyst in 20 minutes.

Due to the pervasiveness of fluorine atoms in industrially relevant small molecules, all practicing organic chemists appreciate the importance of this element. As a result of its unusual size and electronegativity, fluorine imparts unique physicochemical properties to pendant organic molecules.1 For example, the strong C–F bond can prevent biological oxidation pathways, thereby thwarting rapid clearance and potentially improving pharmacokinetics of molecules.2 Moreover, the installation of fluorine or trifluoromethyl groups, with their strong inductive effects,2 can have a profound effect on the pKa of nearby hydrogen atoms.3 These attributes, among others, have solidified the importance of fluorinated molecules in the medicinal,1–4 material,5 and agrochemical6 industries. Yet, the same unique properties that make fluorine atoms attractive chemical modifiers also make their installation difficult. Consequently, new methods for site-selective fluorine incorporation remain highly desirable.7Methods to construct Csp2–F bonds traditionally make use of the Balz–Schiemann fluorodediazonization8 and halogen exchange (“Halex” process).9 Advances in transition metal-mediated fluorination have broadened access to Csp2–F-containing molecules,10 but methods to access aliphatic fluorides remain limited. Conventional methods to make Csp3–F bonds—such as nucleophilic displacement of alkyl halides11 and deoxyfluorination12—can have limited functional group compatibility and unwanted side reactions. A more efficient route to form aliphatic C–F bonds would target the direct fluorination of Csp3–H bonds (Scheme 1).13Open in a separate windowScheme 1(a) Previous work on functional-group directed Csp3–H fluorination; (b) our approach to N-directed fluorination.Recent efforts with palladium catalysis employ conventional C–H-metallation strategies to target Csp3–H bonds for fluorination.14 Alternatively, radical H-atom abstraction can remove the transition metal from the C–H-cleavage step, thereby offering a promising approach for Csp3–H-bond functionalization.15 With undirected C–H fluorination,16 however, selectivity remains a challenge in molecules without strength-differentiated Csp3–H bonds.17 To overcome this, our group pioneered the directed fluorination of benzylic Csp3–H bonds through an iron-catalyzed process that involves 1,5 hydrogen-atom transfer (HAT) to cleave the desired Csp3–H bond.18 Since this work, other groups have demonstrated directed Csp3–H fluorination based on radical propagation that proceeds through an interrupted Hofmann–Löffler–Freytag (HLF)19 reaction (Scheme 1a). These examples employ various radical precursors such as enones,20 ketones,21 hydroperoxides,22 and carboxamides23 to direct fluorination to specific Csp3–H bonds. Since amines are ubiquitous in natural products and drugs, we sought to use amines as the building block of our directing group to achieve fluorination of unactivated Csp3–H bonds (Scheme 1b). By using amines as the starting point, one could use the approach in straightforward synthetic planning for the late-stage functionalization of remote C–H bonds.In the design phase of the project, we needed to devise a synthetically tractable N–F system that would enable 1,5-HAT and allow for fluorine transfer (Scheme 1b). To begin, we decided to examine common amine activating groups that would support 1,5-HAT while avoiding undesired radical reactions. The chosen activating group would provide the ideal steric and electronic properties to enable both N–F synthesis and N–F scission for 1,5-HAT. We first examined common acyl groups (e.g., acetyl-, benzoyl, and tosyl-based amides), but these proved unsatisfactory. For example, fluoroamide synthesis was either not achieved or low yielding, and the desired fluorine transfer proceeded with significant side reactions or returned starting material. We then turned our attention to more sterically hindered amides—which allow for higher yielding fluoroamide synthesis. For fluorine transfer, we hypothesized that the increased steric bulk could slow intermolecular H-atom transfer, thereby leading more efficient intramolecular 1,5-HAT. To that end, we were delighted that pivaloyl-based fluoroamide 1a proceeded in 64% yield to form product 2a (Scheme 2a). Interestingly, 7% of 1a underwent fluorination at the tert-butyl group of the pivaloyl—presumably through a 1,4-HAT reaction (2aa, Scheme 2a).24 The problem is further exacerbated when the pivaloyl group is homologated by one methylene—providing only 7% yield of desired 2b with 32% of the fluorination taking place on the iso-pentyl group (2bb, Scheme 2a). In an attempt to “tie back” the pivaloyl group and prevent the undesired fluorination, we employed a cyclopropylmethyl-based fluoroamide but observed no improvement.Open in a separate windowScheme 2(a) The targeted 1,5-fluorination of unactivated aliphatic C–H bonds results in partial fluorination of the amine activating group; (b) DFT studies (uM06/cc-pVTZ(-f)-LACV3P**//uM06/LACVP** level of theory) identified the competing pathways responsible for alternate fluorination; (c) DFT (uM06/cc-pVTZ(-f)-LACV3P**//uM06/LACVP** level of theory) evaluation of adamantoylamides revealed higher transition state energy for 1,4-HAT due to restricted vibrational scissoring (d) adamantoyl-activated octylamine shows no fluorination of the activating group. a 1H-NMR yield using 1,3,5-trimethoxybenzene as an internal standard. b 19F-NMR yield using 4-fluorotoluene as an internal standard.At this point, 1a proved most promising for efficient fluorine transfer, as well as being the most synthetically accessible fluoroamide. The increased steric hindrance minimizes N-sulfonylation during fluorination with NFSI, a problem that plagued the synthesis of our previously targeted fluoroamides.18 Therefore, to further investigate how to improve fluorine transfer from 1a, we decided to model H-abstraction computationally.We hypothesized that the fluorinated side product 2aa was formed after 1,4-HAT. Since 1,4-HAT is rare,24 we employed DFT (see ESI for details) to calculate the 5-membered and 6-memebered transition-states for 1,4- and 1,5-HAT, respectively. Surprisingly, we found that the barrier for 1,4 C–H abstraction in 1a was 18.7 kcal mol−1, which was only 2.6 kcal mol−1 higher in energy than the barrier calculated for 1,5 C–H abstraction in the same system (Scheme 2b). This suggested that both processes were competing at room temperature. We attributed the comparable barriers to the flexibility of the tert-butyl group, which undergoes vibrational scissoring to accommodate the C–H abstraction. The transition state distortion is modest and allows the molecule to maintain bond angles close to the ideal 109.5° (Scheme 2b). Based on this insight, we sought to limit the scissoring of the tert-butyl group and prevent the 1,4-HAT that leads to the undesired side product. After investigating several possible candidates, the underutilized adamantoyl group appeared promising. To evaluate the rigidity of adamantane, we calculated the barriers for 1,4- and 1,5-HAT for the adamantoyl-capped octylamine 1c (Scheme 2c). As expected, the barriers for 1,4- and 1,5-HAT differed significantly—with 1,4 C–H abstraction proceeding with a barrier of 25.1 kcal mol−1 and the 1,5-HAT barely changed at 16.4 kcal mol−1—an 8.7 kcal mol−1 difference. Consequently, we synthesized 1c and subjected it to the reaction conditions. Excitingly, the adamantoyl-capped system produced desired product 2c in 75% yield with no fluorination of the adamantyl group (Scheme 2d).Using the newly devised adamantoyl-based fluoroamides, the reaction conditions were optimized. While a range of metal salts, ligands, and radical initiators were evaluated, Fe(OTf)2 proved unique in catalyzing fluorine transfer with fluoroamides.18 Catalyst loading of 10 mol% allowed convenient setup and minor deviations above or below this loading had little effect on yield (see ESI). Increasing the temperature to 40 °C produced a slight increase in yield (entry 2, Table 1). Likewise, raising the temperature to 80 °C resulted in full conversion of the starting material in 20 minutes with 81% yield of the desired product (entry 3, Table 1). It should be noted that fluorine transfer occurs efficiently at a variety of temperatures with adjustments in reaction time (see ESI). Increasing the reaction concentration or changing the solvent resulted in decreased yield (entries 4 and 5, Table 1). Furthermore, the absence of Fe(OTf)2 leads to no reaction and quantitative recovery of starting material, attesting to the stability of fluoroamides and the effectiveness of Fe(OTf)2 (entry 6, Table 1).Optimization of pertinent reaction parameters
EntrySolventTemp (°C)Conc (M)TimeYielda (%)
1bDMErt0.0515 h75
2DME400.0518 h79
3 DME 80 0.05 20 min 81
4DME800.120 min73
5THE800.0520 min38
6cDME800.0520 min0
Open in a separate windowaDetermined by 1H-NMR with 1,3,5-trimethoxybenzene as an internal standard.bReaction ran inside of glovebox.cReaction ran without Fe(OTf)2.With the optimized conditions established, we evaluated the substrate scope of the reaction (Table 2). The reaction proved quite general for the fluorination of primary and secondary Csp3–H bonds (2c–l, Table 2), while tertiary Csp3–H abstraction led to greater side reactions and lower yields (2m). While all reactions resulted in complete consumption of the fluoroamide, only a singly fluorinated product is produced with the parent amide being the major side product (see ESI). The reaction proved selective for δ-fluorination even in the presence of tertiary Csp3–H bonds (e.g., 2h, 2j, and 2k), thereby demonstrating selectivity counter to C–H-bond strength. Interestingly, transannular fluorine transfer occurs with complete regioselectivity to produce 2l as the sole product. Additionally, benzylic C–H bonds can be fluorinated under these conditions (2n). The reaction also exhibits good functional group compatibility, allowing access to a variety of fluorinated motifs. In particular, the reaction proceeds in the presence of either free or protected alcohols (2o and 2p). Moreover, esters and halides are both tolerated to give fluorinated products 2q and 2r in good yield. Notably, the reaction provides access to fluorohydrin 2s—highlighting the unique ability of this methodology to access both fluorohydrins and γ-fluoroalcohols such as 2o. In addition to these examples, terminal alkene 1t works quite well giving 2t in 67% yield. Furthermore, alkene functionalizations of 2t would provide access to a diverse range of fluorinated motifs. To target difluoromethylene units with this methodology, fluoroamide 1u was prepared and subjected to the reaction conditions. Pleasingly, 2u was observed in 20% yield.Substrate scope for fluorine transfer
Open in a separate windowaIsolated yields. All reactions were run on 0.3 mmol scale unless otherwise noted.bYield reported as an average of two trials.c35 min reaction time.ddr = 1 : 3.2 when ran at room temperature for 24 h.e0.25 mmol scale.f0.18 mmol scale.g0.1 mmol scale, yield determined by 19F-NMR with 4-fluorotoluene as an internal standard.While exploring the substrate scope, we were surprised to discover that the fluoroamide N–F bond is unusually stable to a variety of common reactions. For example, fluoroamide 1o was carried through an Appel reaction, PCC oxidation, and Wittig reaction with minimal loss of the fluoroamide. With such robustness, it becomes obvious that fluoroamides could act as secondary amide protecting group—being installed and carried through a multi-step synthesis until fluorine transfer is desired. Moreover, the greater rigidity of adamantoyl-based amides relative to pivalamides offers greater stability to acid and base hydrolysis—another feature of this system. Fortunately, the amide can be cleaved using conditions reported by Charette et al. with no evidence of elimination or loss of the alkyl fluoride (see ESI).25To evaluate the differences between C–H bonds, we calculated the hypothesized minima and maxima en route to C–F bond formation for primary, secondary, and tertiary substrates (Fig. 1). To begin, we defined the start of the pathway with the fluoroamides as octahedral, high-spin Fe(OTf)2-DME complex (I).18 Ligand dissociation results in the loss of DME to form II which is 7.2 kcal mol−1 higher in energy relative to I. This ligand loss opens a coordination site that allows Fe to enter the catalytic cycle via F-abstraction from the fluoroamides. This proceeds with a barrier (II-TS) of ∼25 kcal mol−1 for all systems to form the corresponding N-based radical (III). This new N-based radical is generally about −10 kcal mol−1 from the starting materials. The 1,5-HAT proceeds through a six-membered transition state (III-TS) with 16.4, 12.6, and 9.7 kcal mol−1 barriers for primary, secondary, and tertiary substrates, respectively. This abstraction forms the corresponding C-based radicals (IV) that were −15.0, −19.9 and −22.4 kcal mol−1 relative to the starting materials for primary, secondary, and tertiary substrates, respectively. A barrierless transition allows for the abstraction of fluorine from Fe(iii)-fluoride to simultaneously furnish the products (V) and regenerate catalyst II. Interestingly, this transition seems to proceed with an intermolecular electron-transfer from the alkyl radicals to the Fe(iii) center. The overall process is highly exergonic at −53.7, −58.6, and −61.9 kcal mol−1 for primary, secondary, and tertiary substrates, respectively. We attribute the low yields for the tertiary example to rapid oxidation of the carbon radical, likely by Fe(iii), that forms a tertiary carbocation and leads to unwanted side reactions. The turnover-limiting step is the N–F abstraction by Fe (II-TS).Open in a separate windowFig. 1Computed relative Gibb''s free energies for intermediates and transition-states along the reaction pathway (uM06/cc-pVTZ(-f)-LACV3P**//uM06/LACVP** level of theory).An alternative pathway, related to the classic HLF reaction,19a,b would involve radical chain propagation. Although unlikely, we also evaluated this pathway computationally (Fig. 1). Consistent with our previous report,18 this process proceeds with an unfavorably high barrier of 30.0, 28.1, and 26.8 kcal mol−1 for primary, secondary, and tertiary substrates, respectively. Hence, this process cannot compete with the barrierless delivery of fluorine from the Fe(iii) fluoride species.In conclusion, we leveraged critical computational insights to enable the use of simple amines as a building block for the directed fluorination of C–H bonds. The reaction targets unactivated Csp3–H bonds site selectively regardless of bond strength. The reaction proceeds under mild iron catalysis that allows broad functional-group compatibility and provides access to unique fluorinated motifs. Moreover, we identified fluoroamides as surprisingly stable functional groups with likely implications for biology and materials. Mechanistic evaluation of fluorine transfer with DFT provided a detailed reaction coordinate that explains the observed reactivity. The overall reaction and mechanistic insights should provide chemists a more predictable approach to site-selective fluorination of C–H bonds.  相似文献   

6.
We report a photochemically induced, hydroxy-directed fluorination that addresses the prevailing challenge of high diastereoselectivity in this burgeoning field. Numerous simple and complex motifs showcase a spectrum of regio- and stereochemical outcomes based on the configuration of the hydroxy group. Notable examples include a long-sought switch in the selectivity of the refractory sclareolide core, an override of benzylic fluorination, and a rare case of 3,3′-difluorination. Furthermore, calculations illuminate a low barrier transition state for fluorination, supporting our notion that alcohols are engaged in coordinated reagent direction. A hydrogen bonding interaction between the innate hydroxy directing group and fluorine is also highlighted for several substrates with 19F–1H HOESY experiments, calculations, and more.

We report a photochemical, hydroxy-directed fluorination that addresses the prevailing challenge of high diastereoselectivity. Numerous motifs showcase a range of regio- and stereochemical outcomes based on the configuration of the hydroxy group.

The hydroxy (OH) group is treasured and versatile in chemistry and biology.1 Its ubiquity in nature and broad spectrum of chemical properties make it an attractive source as a potential directing group.2 The exploitation of the mild Lewis basicity exhibited by alcohols has afforded several elegant pathways for selective functionalization (e.g., Sharpless epoxidation,3 homogeneous hydrogenation,4 cross-coupling reactions,5 among others6). Recently, we reported a photochemically promoted carbonyl-directed aliphatic fluorination, and most notably, established the key role that C–H⋯O hydrogen bonds play in the success of the reaction.7 Our detailed mechanistic investigations prompt us to postulate that other Lewis basic functional groups (such as –OH) can direct fluorination in highly complementary ways.8 In this communication, we report a hydroxy-directed aliphatic fluorination method that exhibits unique directing properties and greatly expands the domain of radical fluorination into the less established realm governing high diastereoselectivity.9Our first inclination that functional groups other than carbonyls may influence fluorination regiochemical outcomes was obtained while screening substrates for our published ketone-directed radical-based method (Scheme 1).8a In this example, we surmised that oxidation of the tertiary hydroxy group on substrate 1 cannot occur and would demonstrate functional group tolerance (directing to C11, compound 2). Surprisingly, the two major regioisomers (products 3 and 4) are derivatized by Selectfluor (SF) on C12 and C16 – indicative of the freely rotating hydroxyl directing fluorination. Without an obvious explanation of how these groups could be involved in dictating regiochemistry, we continued the mechanistic study of carbonyl-directed fluorination (Scheme 2A). We established that the regioselective coordinated hydrogen atom abstraction occurs by hydrogen bonding between a strategically placed carbonyl and Selectfluor radical dication (SRD).7 However, we noted that the subsequent radical fluorination is not diastereoselective due to the locally planar nature of carbonyl groups. Thus, we posed the question: are there other directing groups that can provide both regio- and diastereoselectivity? Such a group would optimally be attached to a sp3 hybridized carbon; thus the “three dimensional” hydroxy carbon logically comes to mind as an attractive choice, and Scheme 1 illustrates the first positive hint.Open in a separate windowScheme 1Observed products for the fluorination of compound 1.Open in a separate windowScheme 2(A) Proposed mechanism, (B) β-caryophyllene alcohol hypochlorite derivative synthetic probe, (C) isodesmic relation of transition states showing the general importance of the hydroxy group to reactivity (ωB97xd/6-31+G*), and (D) 1H NMR experiment with Selectfluor and various additives at different concentrations.We began our detailed study with a simple substrate that contains a tertiary hydroxyl group. Alcohol 5 was synthesized stereoselectively by the reaction of 3-methylcyclohexanone, FeCl3, and 4-chlorophenylmagnesium bromide;10 the 4-chlorophenyl substituent allows for an uncomplicated product identification and isolation (aromatic chromophore). We sought to determine optimal reaction conditions by examination of numerous photosensitizers, bases, solvents, and light sources (7 Although we utilize cool blue LEDs (sharp cutoff ca. 400 nm), CFLs (small amount of UVB (280–315 nm) and UVA (315–400 nm)) are useable as well.11 A mild base additive was also found to neutralize adventitious HF and improve yields in the substrates indicated (
EntrySensitizer 19F yield
1None0%
2 Benzil 83%
3Benzil, no base63%
4Benzil, K2CO368%
5Benzil, CFL light source75%
65-Dibenzosuberenone15%
74,4′-Difluorobenzil63%
89,10-Phenantherenequinone71%
9Perylene8%
10Methyl benzoylformate42%
Open in a separate windowaUnless stated otherwise: substrate (0.25 mmol, 1.0 equiv.), Selectfluor (0.50 mmol, 2.0 equiv.), NaHCO3 (0.25 mmol, 1.0 equiv.), and sensitizer (0.025 mmol, 10 mol%) were dissolved in MeCN (4.0 mL) and irradiated with cool white LEDs for 14 h.Substrate scopea
Open in a separate windowaUnless otherwise specified, the substrate (0.25 mmol, 1.0 equiv.), Selectfluor (0.50 mmol, 2.0 equiv.), NaHCO3 (0.25 mmol, 1.0 equiv. or 0.0 equiv.), and benzil (0.025 mmol 10 mol%) were stirred in MeCN (4.0 mL) and irradiated with cool white LEDs for 14 h. Yields were determined by integration of 19F NMR signals relative to an internal standard and confirmed by isolation of products through column chromatography on silica gel. Yields based on recovered starting material in parentheses. Major diastereomer (with respect to C–F bond) depicted where known.b1.2 equiv. of Selectfluor used.c1.0 equiv. of NaHCO3.d0.0 equiv. of NaHCO3.e3.0 equiv. of Selectfluor used.fIncluding the monofluoride (approx. 11%) with starting material.The screening concurrently buttresses our claim that hydroxy-directed fluorination is proceeding through a mechanism involving a network of C–H⋯OH hydrogen bonds.12 Other N–F reagents (for example, N-fluorobenzenesulfonimide and N-fluoropyridinium tetrafluoroborate) do not provide the desired fluorinated product 6. The 1,3-diaxial relationship shown in Fig. 1 presents an intramolecular competition: tertiary vs. secondary C–H abstraction (O⋯H–C calculated distances: 2.62 and 2.70 Å at B3LYP 6-311++G**, respectively). The tertiary fluoride is the major product in this case.Open in a separate windowFig. 1Example of an intramolecular competition (secondary vs. tertiary C–H abstraction/fluorination) and calculated C–H⋯O distances of compound 5 (B3LYP/6-311++G**).With optimized conditions established, we assessed the site-selectivity of the method with a molecule derived from the acid catalyzed cyclization of α-caryophyllene, β-caryophyllene alcohol (commonly used as a fragrance ingredient in cosmetics, soaps, and detergents).13 When subjected to fluorination conditions, it targets the strained cyclobutane ring (substrate 7) in 52% yield (14 The hydroxy group stereochemistry is poised to direct fluorination to either the C8 or C10 positions (compound 9) due to the plane of symmetry (Fig. 3A). Moreover, we synthesized a complementary derivative through PCC oxidation followed by a Grignard reaction, thereby switching directionality of the hydroxy group (Fig. 3A) to target the C3 or C5 positions instead (compound 8). We found the resultant fluorinated products to be what one expects if engaged in coordinated hydrogen atom transfer (HAT) (55% and 40% for molecules 9 and 8) – a change in regiochemistry based on the stereochemistry of the alcohol. Additionally, only a single stereoisomer is produced for both (d.r. 99 : 1) and reinforce this study as a salient example of diastereoselective radical fluorination.Open in a separate windowFig. 3Examples of hydroxy group stereochemical switches.In the midst of characterizing compound 9, we uncovered a noteworthy hydrogen bonding interaction. Firstly, our plan was to identify the –OH peak within the 1H NMR spectrum and determine if there is a through-space interaction with fluorine in the 19F–1H HOESY NMR spectrum (ultimately aiding in assigning the stereochemistry of the fluorine).15 At first glance, no peaks were immediately discernible as the –OH; however, when a stoichiometric amount of H2O is added, it becomes apparent that the –OH group and geminal proton to the hydroxy peaks broaden by rapid proton exchange (Fig. 2A). Upon closer examination of the dry 1H NMR spectrum, the –OH peak appears to be a sharp doublet of doublets: one bond coupling to the geminal C–H proton of 9 Hz and one of the largest reported through-space couplings to fluorine of 20 Hz. The 19F–1H HOESY spectrum also supports our regio- and stereochemical assignment – a strong interaction between fluorine and Ha, Hb, and Hd, as well as no apparent interaction with Hc and He (Fig. 2B). Consequently, we postulate that intramolecular hydrogen bonding is responsible for the considerable coupling constant. This conclusion is also supported by calculations at B3LYP/6-311++G** (Fig. 2C): the O–H–F angle is given as 140° and F⋯H–O bond distance is 1.97 Å.Open in a separate windowFig. 2(A) Top spectrum (pink) has broadened peaks due to adventitious H2O in solution. (B) Strong interaction observed between the installed fluorine and designated hydroxy proton in the 19F–1H HOESY NMR spectrum. (C) Calculated structure for compound 9 at B3LYP/6-311++G* revealing the hydroxy proton aiming toward the fluorine.Appreciating the complexity and biological significance of steroids,16 we derivatized dehydroepiandrosterone to afford fluorinated substrate 10 (42%; d.r. 99 : 1). Computational modeling assisted in verifying that the β-hydroxy group targets the C12 position (B3LYP/6-311++G**); furthermore, the β-fluoro isomer is the major product (validated by NOESY, 1H, and 19F NMR). Additionally, we subjected 17α-hydroxyprogesterone (endogenous progestogen steroid hormone17) to fluorination conditions and found the α-fluoro product (11) as the major diastereomer in 55% yield (99 : 1 d.r.). To investigate further the notion of coordinated fluorination and explanation of the observed stereoisomers (e.g., β-hydroxy/β-fluoro and α-hydroxy/α-fluoro), we calculated a simplified system comparing the fluorination of 1-propyl radical and γ-propanol radical (Scheme 2C). The reaction can be distilled into two key steps: a site-selective HAT, followed by a diastereoselective fluorination reaction. The following isodesmic relation (ωB97xd/6-31+G*, −7.63 kcal mol−1) illustrates the stabilizing energetic role that the hydroxy group plays in commanding diastereoselectivity. The transition states represent low barrier processes; a solvent dielectric was necessary to find saddle points.Additionally, a simple Protein Data Bank (PDB) survey showed numerous intermolecular close contacts between hydroxy groups and H–C–+NR3 moieties.18 What is more, solutions of Selectfluor with various alcohols at different concentrations reveal characteristic H–C–+NR3 downfield chemical shifts in the 1H NMR spectra (Scheme 2D).19 Both of these observations buttress the claim of a putative hydrogen bonding interaction between Selectfluor and the hydroxy group.We theorize that the regioselective HAT step proceeds similarly to the reported carbonyl-directed pathway (Scheme 2A) involving Selectfluor radical cation coordination (considering the likenesses in conditions and aforementioned Lewis basicity logic). Alternatively, one can imagine the reaction proceeding through a Barton20 or Hofmann–Löffler–Freytag21 style mechanism. To probe this possibility, we employed a β-caryophyllene alcohol hypochlorite derivative to form the alkoxy radical directly, and found that under standard conditions there is complex fragmentation and nonselective fluorination (Scheme 2B). Lastly, we compared the hydroxy versus carbonyl group SF coordination computationally. The carbonyl group is preferred to bind to SF through nonclassical C–H⋯O hydrogen bonds preferentially over the hydroxy group, as the following isodesmic relation shows (acetone and t-BuOH as models; ωB97xd/6-31+G*, −3.81 kcal mol−1), but, once again, rigidity and propinquity are ultimately more important factors in determining directing effects (Scheme 3).Open in a separate windowScheme 3Isodesmic equation comparing carbonyl versus hydroxy group Selectfluor coordination.The tetrahedral nature of hydroxy groups provides unique access to previously unobtainable sites. For example, we compared menthol and an alkylated congener to form products 12 and 13 (Fig. 3B). The hydroxy group in the precursor to 12 is in the equatorial position, mandating the exocyclic isopropyl group as the reactive site (40% yield).22 In the precursor to 13, the methyl and isopropyl substituent lock the hydroxy group into the axial position, targeting its endocyclic tertiary site through a 1,3-diaxial relationship to afford fluorinated product in 57% yield (d.r. 99 : 1). In all, the comparison showcases the versatility in directing ability, offering a choice of regio- and stereoselectivity based on the stereochemistry of the hydroxy group. The directing system only necessitates two features based on our results: (1) the hydroxy group must be either secondary or tertiary (primary tends to favor oxidation) and (2) the oxygen atom must be within the range of 2.4–3.2 Å of the targeted secondary or tertiary hydrogen.Among the several biologically active compounds we screened, caratol derivatives 14 and 15 were found to be attractive candidates that reveal directed fluorination to an exocyclic isopropyl group (23).24 After extraction, isolation, and derivatization, molecules 14 and 15 are afforded in 65% and 83% yield (25 Groves,9f Britton,26 and others.27 The derived alcohol finally overrode this natural tendency and directed to the predicted position in 56% (d.r. 99 : 1) (product 16). Smaller amounts of competitive polar effect fluorination were observed at the C2 and C3 positions, highlighting how challenging a problem the functionalization of the sclareolide core presents.28,29An altered dihydroactinidiolide was found to participate in the fluorination through a 1,3-diaxial guided HAT and fluorination in 55% yield (product 17, d.r. 99 : 1). We next modeled several more substrates that participated in similar 1,3 relationships; however, each exhibited a variation from one another (e.g., ring size or fused aromatic ring). Products 19 and 18 displayed the reaction''s capability to direct to the desired positions with an expanded (65%; d.r. 99 : 1) and reduced (45%; d.r. 99 : 1) ring system when compared to the previous 6-membered ring examples. Additionally, we examined a methylated α-tetralone derivative. The desired 3-fluoro product 20 forms in 43% yield (d.r. 99 : 1), overriding benzylic fluorination (Scheme 4).30 Under identical conditions α-tetralone provides 4-fluorotetralone in 48% yield. In similar motif, 1-phenylindanol, we intentionally targeted the benzylic position in a 90% and 10 : 1 d.r. (product 21). Unlike the methylated α-tetralone derivative, the geometry of the starting material calculated at B3LYP/6-311++G** shows the hydroxy group is not truly axial and is 4.30 Å from the targeted C–H bond, explaining the dip in diastereoselectivity.Open in a separate windowScheme 4Comparing fluorination outcomes for different functional groups.Next, we examined an isomer of borneol that is widely used in perfumery, fenchol.31 The secondary alcohol displays a diastereoselective fluorination in 38% (d.r. 99 : 1) (product 22). Our last designed motif was ideally constructed to have a doubly-directing effect. Our observations show that a well-positioned hydroxy group not only provides sequential regioselective hydrogen atom abstraction but also displays a powerful demonstration of Selectfluor guidance to afford the cis-difluoro product (23) in 33% yield (85% brsm, d.r. 99 : 1). Spectroscopically (1H, 13C, and 19F NMR), the product possesses apparent Cs symmetry and showcases close interactions (e.g., diagnostic couplings and chemical shifts). cis-Polyfluorocycloalkanes are of intense current interest in materials chemistry, wherein faces of differing polarity can complement one another.32All in all, this photochemical hydroxy-directed fluorination report represents one of the first steps in commanding diastereoselectivity within the field of radical fluorination. An ability to dictate regio- and stereoselectivity is demonstrated in a variety of substrates by simply switching the stereochemistry of the hydroxy group. Computations support the key role of Selectfluor coordination to the key hydroxy group in the fluorination step. Future studies will seek to uncover other compatible Lewis basic functional groups, expanding further the versatility of radical fluorination.  相似文献   

7.
Diverse strategies for transition metal catalyzed distal C(sp3)–H functionalizations     
Jayabrata Das  Srimanta Guin  Debabrata Maiti 《Chemical science》2020,11(40):10887
Transition metal catalyzed C(sp3)–H functionalization is a rapidly growing field. Despite severe challenges, distal C–H functionalizations of aliphatic molecules by overriding proximal positions have witnessed tremendous progress. While usage of stoichiometric directing groups played a crucial role, reactions with catalytic transient directing groups or methods without any directing groups are gaining more attention due to their practicality. Various innovative strategies, slowly but steadily, circumvented issues related to remote functionalizations of aliphatic molecules. A systematic compilation has been presented here to provide insights into the recent developments and future challenges in the field. The Present perspective is expected to open up a new dimension and provide an avenue for deep insights into the distal C(sp3)–H functionalizations that could be applied routinely in various pharmaceutical and agrochemical industries.

Transition metal catalyzed C(sp3)–H functionalization is a rapidly growing field.  相似文献   

8.
Site-selective coupling of remote C(sp3)–H/meta-C(sp2)–H bonds enabled by Ru/photoredox dual catalysis and mechanistic studies     
Hong-Chao Liu  Xiangtao Kong  Xiao-Ping Gong  Yuke Li  Zhi-Jie Niu  Xue-Ya Gou  Xue-Song Li  Yu-Zhao Wang  Wei-Yu Shi  Yan-Chong Huang  Xue-Yuan Liu  Yong-Min Liang 《Chemical science》2022,13(18):5382
Construction of C(sp2)–C(sp3) bonds via regioselective coupling of C(sp2)–H/C(sp3)–H bonds is challenging due to the low reactivity and regioselectivity of C–H bonds. Here, a novel photoinduced Ru/photocatalyst-cocatalyzed regioselective cross-dehydrogenative coupling of dual remote C–H bonds, including inert γ-C(sp3)–H bonds in amides and meta-C(sp2)–H bonds in arenes, to construct meta-alkylated arenes has been accomplished. This metallaphotoredox-enabled site-selective coupling between remote inert C(sp3)–H bonds and meta-C(sp2)–H bonds is characterized by its unique site-selectivity, redox-neutral conditions, broad substrate scope and wide use of late-stage functionalization of bioactive molecules. Moreover, this reaction represents a novel case of regioselective cross-dehydrogenative coupling of unactivated alkanes and arenes via a new catalytic process and provides a new strategy for meta-functionalized arenes under mild reaction conditions. Density functional theory (DFT) calculations and control experiments explained the site-selectivity and the detailed mechanism of this reaction.

A novel photoinduced Ru/photocatalyst-cocatalyzed regioselective cross-dehydrogenative coupling of dual remote C–H bonds, including inert γ-C(sp3)–H bonds in amides and meta-C(sp2)–H bonds in arenes, to construct meta-alkylated arenes has been accomplished.  相似文献   

9.
Iodoetherification of unactivated alkenes catalyzed by diphosphine palladium(II) complexes     
Todd A. Doroski 《Tetrahedron letters》2009,50(36):5162-5164
A palladium-catalyzed intramolecular iodoetherification of alkenes is reported. The reaction is efficient and highly diastereoselective for disubstituted alkenes. The tether length between the alcohol and alkene can be varied to produce tetrahydrofuran and tetrahydropyran rings. Diphosphine palladium(II) salts are highly active catalysts enabling future studies on the development of an enantioselective process.  相似文献   

10.
Efficient alkyl ether synthesis via palladium-catalyzed, picolinamide-directed alkoxylation of unactivated C(sp3)-H and C(sp2)-H bonds at remote positions     
Zhang SY  He G  Zhao Y  Wright K  Nack WA  Chen G 《Journal of the American Chemical Society》2012,134(17):7313-7316
We report the efficient synthesis of alkyl ethers by the functionalization of unactivated sp(3)- and sp(2)-hybridized C-H bonds. In the Pd(OAc)(2)-catalyzed, PhI(OAc)(2)-mediated reaction system, picolinamide-protected amine substrates undergo facile alkoxylation at the γ or δ positions with a range of alcohols, including t-BuOH, to give alkoxylated products. This method features a relatively broad substrate scope for amines and alcohols, inexpensive reagents, and convenient operating conditions. This method highlights the emerging value of unactivated C-H bonds, particularly the C(sp(3))-H bond of methyl groups, as functional groups in organic synthesis.  相似文献   

11.
C(sp3)–H oxygenation via alkoxypalladium(ii) species: an update for the mechanism     
Shuaizhong Zhang  Jinquan Zhang  Hongbin Zou 《Chemical science》2022,13(5):1298
Pd-catalyzed C(sp3)–H oxygenation has emerged as an attractive strategy for organic synthesis. The most commonly proposed mechanism involves C(sp3)–H activation followed by oxidative addition of an oxygen electrophile to give an alkylpalladium(iv) species and further C(sp3)–O reductive elimination. In the present study of γ-C(sp3)–H acyloxylation of amine derivatives, we show a different mechanism when tert-butyl hydroperoxide (TBHP) is used as an oxidant—namely, a bimetallic oxidative addition-oxo-insertion process. This catalytic model results in an alkoxypalladium(ii) intermediate from which acyloxylation and alkoxylation products are formed. Experimental and computational studies, including isolation of the putative post-oxo-insertion alkoxypalladium(ii) intermediates, support this mechanistic model. Density functional theory reveals that the classical alkylpalladium(iv) oxidative addition pathway is higher in energy than the bimetallic oxo-insertion pathway. Further kinetic studies revealed second-order dependence on [Pd] and first-order on [TBHP], which is consistent with DFT analysis. This procedure is compatible with a wide range of acids and alcohols for γ-C(sp3)–H oxygenation. Preliminary functional group transformations of the products underscore the great potential of this protocol for structural manipulation.

Alkoxypalladium(ii) species lead to γ-C(sp3)–H acyloxylation and alkoxylation products using tert-butyl hydroperoxide as the oxidant.  相似文献   

12.
Copper(ii) ketimides in sp3 C–H amination     
Isuri U. Jayasooriya  Abolghasem Bakhoda  Rachel Palmer  Kristi Ng  Nour L. Khachemoune  Jeffery A. Bertke  Timothy H. Warren 《Chemical science》2021,12(47):15733
Commercially available benzophenone imine (HN Created by potrace 1.16, written by Peter Selinger 2001-2019 CPh2) reacts with β-diketiminato copper(ii) tert-butoxide complexes [CuII]–OtBu to form isolable copper(ii) ketimides [CuII]–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CPh2. Structural characterization of the three coordinate copper(ii) ketimide [Me3NN]Cu–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CPh2 reveals a short Cu-Nketimide distance (1.700(2) Å) with a nearly linear Cu–N–C linkage (178.9(2)°). Copper(ii) ketimides [CuII]–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CPh2 readily capture alkyl radicals R˙ (PhCH(˙)Me and Cy˙) to form the corresponding R–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CPh2 products in a process that competes with N–N coupling of copper(ii) ketimides [CuII]–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CPh2 to form the azine Ph2C Created by potrace 1.16, written by Peter Selinger 2001-2019 N–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CPh2. Copper(ii) ketimides [CuII]–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CAr2 serve as intermediates in catalytic sp3 C–H amination of substrates R–H with ketimines HN Created by potrace 1.16, written by Peter Selinger 2001-2019 CAr2 and tBuOOtBu as oxidant to form N-alkyl ketimines R–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CAr2. This protocol enables the use of unactivated sp3 C–H bonds to give R–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CAr2 products easily converted to primary amines R–NH2via simple acidic deprotection.

Commercially available benzophenone imine (HN Created by potrace 1.16, written by Peter Selinger 2001-2019 CPh2) reacts with β-diketiminato copper(ii) tert-butoxide complexes [CuII]–OtBu to form isolable copper(ii) ketimides [CuII]–N Created by potrace 1.16, written by Peter Selinger 2001-2019 CPh2 that serve as intermediates in catalytic sp3 C−H amination via radical relay.  相似文献   

13.
Transition metal-catalyzed arylation of unactivated C(sp3)-H bonds     
Baudoin O 《Chemical Society reviews》2011,40(10):4902-4911
Transition-metal-catalyzed C-H bond arylation has recently emerged as a powerful tool for the functionalization of organic molecules that may complement or even replace traditional catalytic cross-couplings. While many efforts have focused on the arylation of arenes and heteroarenes in the past two decades, less studies have been devoted to the arylation of nonacidic C-H bonds of alkyl groups. This tutorial review highlights recent work in this active area.  相似文献   

14.
Merging C(sp3)–H activation with DNA-encoding     
Zhoulong Fan  Shuai Zhao  Tao Liu  Peng-Xiang Shen  Zi-Ning Cui  Zhe Zhuang  Qian Shao  Jason S. Chen  Anokha S. Ratnayake  Mark E. Flanagan  Dominik K. Klmel  David W. Piotrowski  Paul Richardson  Jin-Quan Yu 《Chemical science》2020,11(45):12282
DNA-encoded library (DEL) technology has the potential to dramatically expedite hit identification in drug discovery owing to its ability to perform protein affinity selection with millions or billions of molecules in a few experiments. To expand the molecular diversity of DEL, it is critical to develop different types of DNA-encoded transformations that produce billions of molecules with distinct molecular scaffolds. Sequential functionalization of multiple C–H bonds provides a unique avenue for creating diversity and complexity from simple starting materials. However, the use of water as solvent, the presence of DNA, and the extremely low concentration of DNA-encoded coupling partners (0.001 M) have hampered the development of DNA-encoded C(sp3)–H activation reactions. Herein, we report the realization of palladium-catalyzed C(sp3)–H arylation of aliphatic carboxylic acids, amides and ketones with DNA-encoded aryl iodides in water. Notably, the present method enables the use of alternative sets of monofunctional building blocks, providing a linchpin to facilitate further setup for DELs. Furthermore, the C–H arylation chemistry enabled the on-DNA synthesis of structurally-diverse scaffolds containing enriched C(sp3) character, chiral centers, cyclopropane, cyclobutane, and heterocycles.

DNA-compatible C(sp3)–H activation reactions of aliphatic carboxylic acids, amides, and ketones were developed for efficient access to DEL synthesis.  相似文献   

15.
Modular allylation of C(sp3)–H bonds by combining decatungstate photocatalysis and HWE olefination in flow     
Luca Capaldo  Stefano Bonciolini  Antonio Pulcinella  Manuel Nuo  Timothy Noël 《Chemical science》2022,13(24):7325
The late-stage introduction of allyl groups provides an opportunity to synthetic organic chemists for subsequent diversification, furnishing a rapid access to new chemical space. Here, we report the development of a modular synthetic sequence for the allylation of strong aliphatic C(sp3)–H bonds. Our sequence features the merger of two distinct steps to accomplish this goal, including a photocatalytic Hydrogen Atom Transfer and an ensuing Horner–Wadsworth–Emmons (HWE) reaction. This practical protocol enables the modular and scalable allylation of valuable building blocks and has been applied to structurally complex molecules.

We report a flow platform for the modular allylation of strong aliphatic C(sp3)–H bonds based on the merger of photocatalytic HAT and a HWE reaction. This approach enables both early- and late-stage diversification of various hydroalkanes.

Modern drug discovery programs capitalize increasingly on the application of late-stage functionalization methodologies to accelerate the lead optimization phase.1,2 Such strategies allow for the rapid and cost-efficient3,4 diversification of the parent molecule by exploiting native functionalities (e.g., C–H bonds), thus effectively avoiding the need to redesign its entire synthetic route to access new leads.5–7 More specifically, the late-stage decoration of organic molecules with multipurpose functional groups would provide new points of entry for subsequent diversification.8 Such a strategy could be particularly convenient when it is realized via a chemo- and regioselective functionalization of C–H bonds in the absence of any proximal directing or activating groups.7 However, while C(sp2)–H activation has been extensively investigated, the direct functionalization of C(sp3)–H bonds remains challenging and is often narrow in scope.9 Recently, photocatalytic Hydrogen Atom Transfer (HAT) has been exploited to enable the late-stage functionalization of C(sp3)–H bonds, showing remarkable levels of regioselectivity even in complex molecules (Scheme 1A).10 In HAT photocatalysis, a catalyst converts light energy into chemical energy for the homolytic cleavage of strong aliphatic C–H bonds.Open in a separate windowScheme 1Allylation of C(sp3)–H bonds. (A) Photocatalytic HAT enables late-stage functionalization of structurally complex molecules. (B) Reported approaches for the photocatalyzed radical allylation of organic molecules. (C) A telescoped approach for the modular allylation of C(sp3)–H bonds (this work).Especially, the decatungstate anion ([W10O32]4−) has shown remarkable selectivity for specific C(sp3)–H bonds, governed by an intricate balance between steric and electronic interactions.9,11,12We envisioned that the regioselective introduction of an allyl moiety onto hydrocarbon frameworks would be particularly useful as it provides a convenient branching point for further late-stage synthetic exploitation.13 To install such moieties, radical allylation has manifested itself as a valuable strategy. One approach relies on the use of transition metal complexes to activate a substrate containing an allylic leaving group to afford a π-allyl complex, which is then suited to trap a C-centered radical (Scheme 1B).14 This strategy can engage a diverse set of allyl coupling partners but typically requires purposely designed radical precursors, which prevents the direct allylation of unactivated C(sp3)–H bonds.Another tactic exploits radicofugal groups X (e.g., X = halide, SO2R, SnR3, etc) in the allylic position to afford the desired product via a radical addition–fragmentation process (Scheme 1B).15–28 However, while synthetically useful, this transformation is not suitable for the synthesis of densely functionalized allylic functionalities.Seeking to address these challenges, we sought to develop a robust and versatile synthetic platform for the allylation of strong aliphatic C(sp3)–H bonds. Hereto, a modular synthetic sequence is preferred in which the allylic moiety is assembled in a stepwise fashion, enabling the rapid generation of structurally diverse analogues. Specifically, our sequence features the merger of two distinct synthetic steps to accomplish this goal (Scheme 1C). First, we planned to activate C(sp3)–H bonds via decatungstate-catalyzed Hydrogen Atom Transfer29,30 and subsequently trap the resulting C-centered radical with a vinyl phosphonate.31,32 The ensuing radical addition product serves as a suitable linchpin for the second step, in which a classical Horner–Wadsworth–Emmons (HWE) olefination33,34 is able to deliver the targeted allylated compounds. In order to streamline these two steps, we reasoned that a telescoped flow protocol where the reactions are performed in tandem without the need for tedious purification of intermediates would be indispensable not only to accelerate access to these valuable building blocks but also to ensure facile scalability.35–37 Herein, we report the successful realization of a flow platform enabling the allylation of a wide range of unactivated hydrocarbons.Our investigations commenced with the decatungstate-enabled hydroalkylation of ethyl 2-(diethoxyphosphoryl)acrylate (2) using cyclohexane as the H-donor (see ESI, Table S1). Following a careful optimization of different reaction parameters, we found that the photocatalytic radical addition performed optimal in continuous-flow using a commercially available Vapourtec UV-150 photochemical reactor (PFA (perfluoroalkoxy) capillary, ID: 0.75 mm; V = 3.06 mL, flow rate = 0.612 mL min−1, tr = 5 min) equipped with a 60 W UV-A LED light source (λ = 365 nm), which matches the measured absorption spectrum of decatungstate. A 65% NMR yield (64% after isolation) was obtained for the targeted hydroalkylated compound when a CH3CN solution of the acrylate (0.1 M), cyclohexane (20 equivalents) and tetrabutylammonium decatungstate (TBADT, (Bu4N)4[W10O32]) as the photocatalyst (1 mol%)38–46 was irradiated for 5 minutes (see ESI, Table S1, Entry 9). Other HAT photocatalysts, such as Eosin Y,47 anthraquinone,48 5,7,12,14-pentacenetetrone28 and 9-fluorenone49 were also evaluated, but failed to deliver the targeted product. Interestingly, benzophenone50,51 showed a comparable activity to the decatungstate anion, although only when used at high catalyst loading (20 mol%, 68% NMR yield). Due to the lower extinction coefficient of benzophenone compared to TBADT (<200 vs. 13 500 M−1 cm−1),52,53 and its known tendency to dimerize to form benzopinacol upon UV-A irradiation, we selected TBADT as the best photocatalyst for the targeted hydroalkylation reaction. Notably, this transformation is quite general and a diverse set of alkylphosphonates (3) could be readily isolated and characterized (see ESI, Section 7). A mechanistic study confirmed the radical nature of the process (see ESI, Section 5), where HAT is likely to occur during the rate-determining step (KIE = 1.9).Next, the obtained alkylphosphonates were subjected to the successive HWE olefination (Scheme 2). A telescoped flow approach was developed in which the two individual steps were connected in a single streamlined flow process without intermediate purification. We selected 1,3-benzodioxole (1a), a common moiety in many medicinally-relevant molecules, as the H-donor and exposed it to the photocatalytic reaction conditions. Upon exiting the photochemical reactor, the reaction mixture containing the alkylphosphonate is merged with a stream containing paraformaldehyde (3 equiv.) and lithium tert-butoxide (1.1 equiv.) in tetrahydrofuran. The combined reaction mixture is subsequently introduced into a second capillary microreactor (PFA, ID: 0.75 mm; V = 7.1 mL; tr = 5 min; T = 40 °C) and, after only 5 minutes of residence time, the targeted C(sp3)–H allylated product 4 could be obtained in 80% overall NMR yield (70% after isolation). Interestingly, the reaction performed decently also with 1 equivalent of 1a (65% NMR yield). Notably, the tactical combination of these two steps in flow results in a very efficient and operationally simple protocol, delivering these coveted scaffolds in only 10 minutes overall reaction time. As another benefit, the flow process could be readily scaled to produce 10 mmol of the desired compound 4 (1.52 g, 65% isolated yield, Scheme 2) without the need for tedious reoptimization of the reaction conditions, which is typically associated with batch-type scale up procedures.Open in a separate windowScheme 2Scope of the modular allylation of strong aliphatic C–H bonds with (deuterated) paraformaldehyde. Yields are given over two steps. For further experimental details see the SI. a For (CH2O)n: 0.23 M aldehyde and 0.084 M LiOtBu solution in tetrahydrofuran; flow rate = 0.802 mL min−1; tR = 5 min. For (CD2O)n: 0.11 M aldehyde and 0.084 M LiOtBu solution in tetrahydrofuran; flow rate = 0.802 mL min−1; tR = 8 min. b TBADT was used 5 mol%.This telescoped strategy could be subsequently applied to a wide variety of hydrogen atom donors 1 (Scheme 2). Activated substrates, such as hydrocarbon scaffolds with α-to-O C(sp3)–H bonds (5–7), were regioselectively allylated in yields ranging from 49–66% over two steps. Similarly, substrates containing α-to-S (8 and 9) and α-to-N (10–13) C(sp3)–H bonds were functionalized without difficulty (52–70% overall yield). Allylic functional groups could also be appended to activated benzylic positions (14, 32% overall yield).Finally, even strong, non-activated aliphatic C–H bonds could be readily allylated using our approach (15–19, 44–53% overall yield).To further demonstrate the potential of this operationally facile approach to introduce allylic functional groups, we wondered whether paraformaldehyde-d2 could be used in the HWE step. Such a straightforward, regioselective introduction of deuterium atoms in organic molecules would be of tremendous importance for mechanistic,54,55 spectroscopic and tracer studies.56 Using our two-step flow protocol, the analogous deutero-allylated compound 4-d2 was isolated in 68% yield, perfectly matching the result obtained for the non-deuterated version 4. Similarly, N-Boc piperidinone and N-methyl-2-pyrrolidone were competent substrates for this protocol affording the deuterated products 20 and 21 in 44% and 52% yield, respectively. Finally, in an effort to demonstrate the applicability of this method to the late-stage functionalization of medicinally relevant molecules, we subjected biologically active molecules to our two-step flow protocol: the terpenoid ambroxide (22, 40% yield) and the nootropic drug aniracetam (23, 21% yield) could be decorated with a deuterated allylic moiety.In a similar vein, we turned our attention to introduce aromatic and aliphatic aldehydes in the second step, yielding trisubstituted allylic moieties, which are particularly challenging to synthesize via traditional photocatalyzed radical allylation approaches (Scheme 1B). By exploiting our modular protocol, a virtually limitless array of substituents can be systematically introduced (Scheme 3). In most cases, prolonged reaction times were required to obtain full conversion. In particular, electron-deficient aldehydes were convenient substrates for a fully telescoped manifold, where the flow exiting the photoreactor was directly merged with a stream containing the aldehyde and the base (see e.g., 26–30, 35–40). The HWE step required 30 minutes residence time and the temperature was kept at 40 °C. We found that a range of pyridine-derived nicotinaldehydes and heteroaromatic aldehydes (35–41) were ideal substrates for this approach as well. As for electron-neutral and -rich carbonyl compounds, the HWE step required considerably longer reaction times and thus a fed-batch approach was found to be more practical (e.g., 25, 31). Here, the reaction stream exiting the photoreactor was directly dosed into a stirring solution of aldehyde and base. It is important to stress that a fully telescoped approach was still possible in these cases, however higher reaction temperature (60 °C) and a back-pressure regulator (BPR, 2.8 bar) were needed to obtain full conversion within 1 hour (e.g., 24, 33, 45). Another general observation that could be made is that the presence of ortho-substituents resulted in higher E : Z ratios (e.g., 28–31, 33 and 40).Open in a separate windowScheme 3Scope of the modular allylation of strong aliphatic C–H bonds with aromatic and aliphatic aldehydes. Yields are given over two steps. For the experimental details of the fed-batch procedure see GP4 in the ESI, while for fully telescoped approach see GP5. a Reactions were carried out on a 0.5 mmol scale and yields refer to isolated products, E : Z ratios were measured by 1H-NMR. b Reaction performed according to GP5, but the HWE step required 60 °C, a BPR (2.8 bar) and 1 hour residence time. c Reaction time: 16 h. d Reaction performed via general procedure GP6 in the ESI.Next, we turned to investigate different classes of hydrogen donors, such as hydrocarbons (43, 43%), (thio)ethers (44–45, 47–68%), protected amines (46, 51%) and amides (47, 55%): all proved to be competent reaction partners. In all cases, the reaction performed well, delivering densely functionalized alkenes in good yields and stereoselectivity.It is important to note that it would be extremely challenging to access either of these motifs with the current radical allylation methodologies (Scheme 1B). Unfortunately, all attempts to engage ketones in the HWE step did not afford the desired fully-substituted olefins.Interestingly, our protocol was also amenable to aliphatic aldehydes containing enolizable positions (48–52, 57–71% yield). The use of protected piperidine-4-carboxaldehydes allowed to obtain the corresponding allylated products 51 and 52 in excellent yields (60–68%) and with good diastereomeric ratios. In addition, medicinal agents and natural products containing carbonyls, such as acetyl-protected helicin, citronellal and indomethacin aldehyde derivatives, were also reactive delivering the targeted olefins in synthetically useful yields (53–55, 20–63%). This proves the potential of this strategy to rapidly diversify double bonds.Next, the importance of the ester moiety as electron-withdrawing group (EWG) in the substrates to enable the targeted transformations was evaluated (Scheme 4A). Thus, we synthesized different vinyl phosphonates (2′–2′′′) and found that all of them performed well (40–68% 1H-NMR yield) in the photocatalytic radical hydroalkylation. We then tested our streamlined process with benzaldehyde (GP4) to study the effect of the EWG on the diastereomeric ratio in the final allylated compound. The cyano group-bearing substrate furnished the targeted compound 56 with an excellent diasteroselectivity; however, a poor mass balance was observed (22% yield despite full conversion of 3′). In contrast, products 57 and 58 (EWG : COR) were not formed, with a complete recovery of 3′′ and 3′′′. Interestingly, we found that compound 2′′′′ could serve as a suitable radical trap as well (Scheme 4B). Using 1a as coupling partner, the targeted hydroalkylation product was obtained in excellent yield (3′′′′, 90% by 1H-NMR). A solvent switch and a stronger base (nBuLi, n-butyl lithium) were however required to induce the subsequent HWE step yielding styrenes 59–61 in good yields after isolation (see GP7 in the ESI).57,58Open in a separate windowScheme 4(A) Effect of the EWG on the diasteroselectivity in the final allylated product; (B) synthesis of densely functionalized styrenes by exploiting phenyl-substituted vinyl phosphonate 2′′′′; (C) examples of further diversification of compound 4, including olefin reduction, ester reduction, Giese-type radical addition and Mizoroki–Heck coupling. a Full conversion of 3′ was observed. b Full recovery of the alkyl phosphonates.The regioselective and late-stage installation of allylic groups opens up innumerable possibilities for further diversification.13 As an illustration of this synthetic potential, we explored diverse conditions for the conversion of 4 into functionalized derivatives (Scheme 4C). The olefin and the ester functionalities could be orthogonally reduced by exploiting different reduction conditions, yielding compounds 62 (70%) and 63 (62%), respectively.59,60 Moreover, compound 4 was an ideal substrate for another Giese-type radical addition using decatungstate-photocatalyzed HAT (64, 62%). Finally, product 65 could be obtained via a classical Mizoroki–Heck-type coupling (60%).61  相似文献   

16.
Kinetic resolution of sulfur-stereogenic sulfoximines by Pd(ii)–MPAA catalyzed C–H arylation and olefination     
Kallol Mukherjee  Nicolas Grimblat  Somratan Sau  Koushik Ghosh  Majji Shankar  Vincent Gandon  Akhila K. Sahoo 《Chemical science》2021,12(44):14863
A direct Pd(ii)-catalyzed kinetic resolution of heteroaryl-enabled sulfoximines through an ortho-C–H alkenylation/arylation of arenes has been developed. The coordination of the sulfoximine pyridyl-motif and the chiral amino acid MPAA ligand to the Pd(ii)-catalyst controls the enantio-discriminating C(aryl)–H activation. This method provides access to a wide range of enantiomerically enriched unreacted aryl-pyridyl-sulfoximine precursors and C(aryl)–H alkenylation/arylation products in good yields with high enantioselectivity (up to >99% ee), and selectivity factor up to >200. The coordination preference of the directing group, ligand effect, geometry constraints, and the transient six-membered concerted-metalation–deprotonation species dictate the stereoselectivity; DFT studies validate this hypothesis.

A Pd/MPAA catalysed KR of heteroaryl substituted sulfoximines through C–H alkenylation and arylation (up to >99% ee) is developed. In-depth DFT studies uncover the salient features.  相似文献   

17.
Highly regioselective carbonylation of unactivated C(sp3)-H bonds by ruthenium carbonyl     
Hasegawa N  Charra V  Inoue S  Fukumoto Y  Chatani N 《Journal of the American Chemical Society》2011,133(21):8070-8073
The regioselective carbonylation of unactivated C(sp(3))-H bonds of aliphatic amides was achieved using Ru(3)(CO)(12) as a catalyst. The presence of a 2-pyridinylmethylamine moiety in the amide is crucial for a successful reaction. The reaction shows a preference for C-H bonds of methyl groups as opposed to methylene C-H bonds and tolerates a variety of functional groups. The stoichiometric reaction of an amide with Ru(3)(CO)(12) gave a dinuclear ruthenium complex in which the 2-pyridinylmethylamino moiety was coordinated to the ruthenium center in an N,N manner.  相似文献   

18.
Mild electrophilic trifluoromethylation of secondary and primary aryl- and alkylphosphines using hypervalent iodine(iii)-CF(3) reagents     
Eisenberger P  Kieltsch I  Armanino N  Togni A 《Chemical communications (Cambridge, England)》2008,(13):1575-1577
A direct, mild and efficient trifluoromethylation of primary and secondary phosphines is achieved with easily accessible, cheap hypervalent iodine compounds acting as electrophilic CF(3)-transfer reagents.  相似文献   

19.
Visible light driven deuteration of formyl C–H and hydridic C(sp3)–H bonds in feedstock chemicals and pharmaceutical molecules     
Yulong Kuang  Hui Cao  Haidi Tang  Junhong Chew  Wei Chen  Xiangcheng Shi  Jie Wu 《Chemical science》2020,11(33):8912
Deuterium labelled compounds are of significant importance in chemical mechanism investigations, mass spectrometric studies, diagnoses of drug metabolisms, and pharmaceutical discovery. Herein, we report an efficient hydrogen deuterium exchange reaction using deuterium oxide (D2O) as the deuterium source, enabled by merging a tetra-n-butylammonium decatungstate (TBADT) hydrogen atom transfer photocatalyst and a thiol catalyst under light irradiation at 390 nm. This deuteration protocol is effective with formyl C–H bonds and a wide range of hydridic C(sp3)–H bonds (e.g. α-oxy, α-thioxy, α-amino, benzylic, and unactivated tertiary C(sp3)–H bonds). It has been successfully applied to the high incorporation of deuterium in 38 feedstock chemicals, 15 pharmaceutical compounds, and 6 drug precursors. Sequential deuteration between formyl C–H bonds of aldehydes and other activated hydridic C(sp3)–H bonds can be achieved in a selective manner.

A selective hydrogen deuterium exchange reaction with formyl C–H bonds and a wide range of hydridic C(sp3)–H bonds has been achieved by merging tetra-n-butylammonium decatungstate photocatalyst and a thiol catalyst under 390 nm light irradiation.  相似文献   

20.
Substrate-controlled C–H or C–C alkynylation of cyclopropanes: generation of aryl radical cations by direct light activation of hypervalent iodine reagents     
Tin V. T. Nguyen  Matthew D. Wodrich  Jerome Waser 《Chemical science》2022,13(43):12831
We report the first oxidative C–H alkynylation of arylcyclopropanes. Irradiation of ethynylbenziodoxolone (EBX) reagents with visible light at 440 nm promoted the reaction. By the choice of the aryl group on the cyclopropane, it was possible to completely switch the outcome of the reaction from the alkynylation of the C–H bond to the oxyalkynylation of the C–C bond, which proceeded without the need for a catalyst, in contrast to previous works. The oxyalkynylation could also be extended to aminocyclopropanes as well as styrenes. Computations indicated that the C–H activation became a favoured nearly barrierless process in the presence of two ortho methyl groups on the benzene ring.

C–C or C–H alkynylation of aryl cyclopropanes was realized by direct activation of EBX reagents with visible light. With ortho methyl groups on the benzene, C–H functionalization became favoured.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司    京ICP备09084417号-23

京公网安备 11010802026262号