首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 23 毫秒
1.
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   

2.
Intermolecular CDC amination of remote and proximal unactivated Csp3–H bonds through intrinsic substrate reactivity – expanding towards a traceless directing group     
Suresh Rajamanickam  Mayank Saraswat  Sugumar Venkataramani  Bhisma K. Patel 《Chemical science》2021,12(46):15318
An intermolecular radical based distal selectivity in appended alkyl chains has been developed. The selectivity is maximum when the distal carbon is γ to the appended group and decreases by moving from γδε positions. In –COO– linked alkyl chains, the same distal γ-selectivity is observed irrespective of its origin, either from the alkyl carboxy acid or alkyl alcohol. The appended groups include esters, N–H protected amines, phthaloyl, sulfone, sulfinimide, nitrile, phosphite, phosphate and borate esters. In borate esters, boron serves as a traceless directing group, which is hitherto unprecedented for any remote Csp3–H functionalization. The selectivity order follows the trend: 3° benzylic > 2° benzylic > 3° tertiary > α to keto > distal methylene (γ > δ > ε). Computations predicted the radical stability (thermodynamic factors) and the kinetic barriers as the factors responsible for such trends. Remarkably, this strategy eludes any designer catalysts, and the selectivity is due to the intrinsic substrate reactivity.

An intermolecular amination at the distal methylene carbon has been realized in an appended alkyl chain with electron withdrawing groups. Traceless remote Csp3–H functionalization has been accomplished using borate esters.  相似文献   

3.
Ni(0)-promoted activation of Csp2–H and Csp2–O bonds     
Sehye Min  Jonghoon Choi  Changho Yoo  Peter M. Graham  Yunho Lee 《Chemical science》2021,12(29):9983
A dinickel(0)–N2 complex, stabilized with a rigid acridane-based PNP pincer ligand, was studied for its ability to activate C(sp2)–H and C(sp2)–O bonds. Stabilized by a Ni–μ–N2–Na+ interaction, it activates C–H bonds of unfunctionalized arenes, affording nickel–aryl and nickel–hydride products. Concomitantly, two sodium cations get reduced to Na(0), which was identified and quantified by several methods. Our experimental results, including product analysis and kinetic measurements, strongly suggest that this C(sp2)–H activation does not follow the typical oxidative addition mechanism occurring at a low-valent single metal centre. Instead, via a bimolecular pathway, two powerfully reducing nickel ions cooperatively activate an arene C–H bond and concomitantly reduce two Lewis acidic alkali metals under ambient conditions. As a novel synthetic protocol, nickel(ii)–aryl species were directly synthesized from nickel(ii) precursors in benzene or toluene with excess Na under ambient conditions. Furthermore, when the dinickel(0)–N2 complex is accessed via reduction of the nickel(ii)–phenyl species, the resulting phenyl anion deprotonates a C–H bond of glyme or 15-crown-5 leading to C–O bond cleavage, which produces vinyl ether. The dinickel(0)–N2 species then cleaves the C(sp2)–O bond of vinyl ether to produce a nickel(ii)–vinyl complex. These results may provide a new strategy for the activation of C–H and C–O bonds mediated by a low valent nickel ion supported by a structurally rigidified ligand scaffold.

A structurally rigidified nickel(0) complex was found to be capable of cleaving both C(sp2)–H and C(sp2)–O bonds.  相似文献   

4.
Construction of congested Csp3–Csp3 bonds by a formal Ni-catalyzed alkylboration     
Amit Kumar Simlandy  Stephen R. Sardini  M. Kevin Brown 《Chemical science》2021,12(15):5517
Through the combination of a Ni-catalyzed alkene alkenylboration followed by hydrogenation, the synthesis of congested Csp3–Csp3-bonds can be achieved. Conditions have been identified that allow for the use of both alkenyl-bromides and -triflates. In addition, the hydrogenation creates another opportunity for stereocontrol, thus allowing access to multiple stereoisomers of the product. Finally, the method is demonstrated in the streamlined synthesis of a biologically relevant molecule.

Through the combination of a Ni-catalyzed alkene alkenylboration followed by hydrogenation, the synthesis of congested Csp3–Csp3-bonds can be achieved.  相似文献   

5.
Rapid formation of Csp3–Csp3 bonds through copper-catalyzed decarboxylative Csp3–H functionalization     
《中国化学快报》2023,34(1):107477
Transition-metal-catalyzed decarboxylative and CH functionalization strategy for the construction of Csp2-Csp2, Csp2-Csp, and Csp2-Csp3 bonds has been extensively studied. However, research surveys of this synthetic strategy for the Csp3-Csp3 bond forming reactions are surprisingly scarce. Herein, we present an efficient approach for the rapid formation of Csp3–Csp3 bond through copper-catalyzed decarboxylative Csp3H functionalization. The present method should provide a useful access to C3-substituted indole scaffolds with possible biological activities. Mechanistic experiments and DFT calculations supported a dual-Cu(II)-catalytic cycle involving rate-determining decarboxylation in an outer-sphere radical pathway and spin-crossover-promoted CC bond formation. This strategy offers a promising synthesis method for the construction of Csp3–Csp3 bond in the field of synthetic and pharmaceutical chemistry and extends the number of still limited copper-catalyzed decarboxylative Csp3–Csp3 bond forming reaction.  相似文献   

6.
Heteroarylation of unactivated C–H bonds suitable for late-stage functionalization     
Austin S. Miller  Erik J. Alexanian 《Chemical science》2022,13(40):11878
The late-stage introduction of diverse heterocycles onto complex small molecules enables efficient access to new medicinally relevant compounds. An attractive approach to such a transformation would utilize the ubiquitous aliphatic C–H bonds of a complex substrate. Herein, we report a system that enables direct C–H heteroarylation using a stable, commercially available O-alkenylhydroxamate with heterocyclic sulfone partners. The C–H heteroarylation proceeds efficiently with a range of aliphatic substrates and common heterocycles, and is a rare example of heteroarylation of strong C–H bonds. Importantly, the present approach is amenable to late-stage functionalization as the substrate is the limiting reagent in all cases.

The late-stage introduction of diverse heterocycles onto complex small molecules enables efficient access to new medicinally relevant compounds.  相似文献   

7.
Total syntheses of naturally occurring antiviral indolosesquiterpene alkaloids,xiamycins C–F via Csp3–H functionalization     
Mintu Munda  Rhituparna Nandi  Vipin R. Gavit  Sourav Kundu  Sovan Niyogi  Alakesh Bisai 《Chemical science》2022,13(39):11666
Concise total syntheses of naturally occurring antiviral indolosesquiterpene alkaloids, xiamycin C (2a), D (2b), E (2c) and F (2d), have been achieved via a late-stage oxidative δ-Csp3–H functionalization of an advanced pentacyclic enone intermediate 8. This strategy takes advantage of ipso-nitration of naturally occurring abietane diterpenoids to synthesize o-bromo nitroarene derivative 11. A Suzuki–Miyaura coupling of 11 with phenylboronic acid followed by Cadogan''s ring closure provided a modular approach to a carbazole ring required for a functionalized pentacyclic core of indolosesquiterpene alkaloids.

Enantioenriched enone 8 was synthesized via three key transformations: ipso-nitration of abietane diterpenoids to furnish o-bromo nitroarene 11, Suzuki coupling with phenylboronic acid, and Cadogan''s reductive ring closure to craft a carbazole ring.  相似文献   

8.
Organophotoredox/palladium dual catalytic decarboxylative Csp3–Csp3 coupling of carboxylic acids and π-electrophiles     
Kaitie C. Cartwright  Jon A. Tunge 《Chemical science》2020,11(31):8167
A dual catalytic decarboxylative allylation and benzylation method for the construction of new C(sp3)–C(sp3) bonds between readily available carboxylic acids and functionally diverse carbonate electrophiles has been developed. The new process is mild, operationally simple, and has greatly improved upon the efficiency and generality of previous methodology. In addition, new insights into the reaction mechanism have been realized and provide further understanding of the harnessed reactivity.

A dual catalytic decarboxylative allylation and benzylation method for the construction of new C(sp3)–C(sp3) bonds between readily available carboxylic acids and functionally diverse carbonate electrophiles has been developed.  相似文献   

9.
Selective cleavage of unactivated arene ring C–C bonds by iridium: key roles of benzylic C–H activation and metal–metal cooperativity     
Yancong Tian  Martin Jakoobi  Roman Boulatov  Alexey G. Sergeev 《Chemical science》2021,12(10):3568
The cleavage of aromatic C–C bonds is central for conversion of fossil fuels into industrial chemicals and designing novel arene functionalisations through ring opening, expansion and contraction. However, the current progress is hampered by both the lack of experimental examples of selective oxidative addition of aromatic C–C bonds and limited understanding of the factors that favour insertion into the C–C rather than the C–H bonds. Here, we describe the comprehensive mechanism of the only reported chemo- and regioselective insertion of a transition metal into a range of substituted arene rings in simple iridium(i) complexes. The experimental and computational data reveal that this ring cleavage requires both reversible scission of a benzylic C–H bond and cooperativity of two Ir centres sandwiching the arene in the product-determining intermediate. The mechanism explains the chemoselectivity and scope of this unique C–C activation in industrially important methylarenes and provides a general insight into the role of metal–metal cooperativity in the cleavage of unsaturated C–C bonds.

The detailed mechanism of iridium-mediated C–C cleavage in unactivated arenes reveals the key factors enabling the process and helps predict the scope of the cleavage reaction.  相似文献   

10.
Intramolecular Csp3–H/C–C bond amination of alkyl azides for the selective synthesis of cyclic imines and tertiary amines     
Xiaojin Wen  Xinyao Li  Xiao Luo  Weijin Wang  Song Song  Ning Jiao 《Chemical science》2020,11(17):4482
The intramolecular Csp3–H and/or C–C bond amination is very important in modern organic synthesis due to its efficiency in the construction of diversified N-heterocycles. Herein, we report a novel intramolecular cyclization of alkyl azides for the synthesis of cyclic imines and tertiary amines through selective Csp3–H and/or C–C bond cleavage. Two C–N single bonds or a C Created by potrace 1.16, written by Peter Selinger 2001-2019 N double bond are efficiently constructed in these transformations. The carbocation mechanism differs from the reported metal nitrene intermediates and therefore enables metal-free and new transformation.

A novel intramolecular cyclization of alkyl azides for the synthesis of cyclic imines and tertiary amines has been developed. The aliphatic C–H or C–C bond was selectively cleaved with the efficient formation of two C–N single bonds or a C Created by potrace 1.16, written by Peter Selinger 2001-2019 N double bond.  相似文献   

11.
Palladium‐Catalyzed Homocoupling Reactions Between Two Csp3—Csp3 Centers.     
Aiwen Lei  Xumu Zhang 《ChemInform》2002,33(48):91-91
  相似文献   

12.
A general method for site-selective Csp3–S bond formation via cooperative catalysis     
Yuman Qin  Yujie Han  Yongzhen Tang  Junfa Wei  Mingyu Yang 《Chemical science》2020,11(5):1276
Herein, we report a copper-catalysed site-selective thiolation of Csp3–H bonds of aliphatic amines. The method features a broad substrate scope and good functional group compatibility. Primary, secondary, and tertiary C–H bonds can be converted into C–S bonds with a high efficiency. The late-stage modification of biologically active compounds by this method was also demonstrated. Furthermore, the one-pot preparation of pyrrolidine or piperidine compounds via a domino process was achieved.

A copper-catalyzed site-selective thiolation of Csp3–H bonds of aliphatic amines was developed. The method features a broad substrate scope and good functional group tolerance.  相似文献   

13.
Site-selective functionalization of remote aliphatic C–H bonds via C–H metallation     
Qi Zhang  Bing-Feng Shi 《Chemical science》2021,12(3):841
Directing group assistance provided a paradigm for controlling site-selectivity in transition metal-catalyzed C–H functionalization reactions. However, the kinetically and thermodynamically favored formation of 5-membered metallacycles has greatly hampered the selective activation of remote C(sp3)–H bonds via larger-membered metallacycles. Recent development to achieve remote C(sp3)–H functionalization via the C–H metallation process largely relies on employing specific substrates without accessible proximal C–H bonds. Encouragingly, recent advances in this field have enabled the selective functionalization of remote aliphatic C–H bonds in the presence of equally accessible proximal ones by taking advantage of the switch of the regiodetermining step, ring strain of metallacycles, multiple non-covalent interactions, and favourable reductive elimination from larger-membered metallacycles. In this review, we summarize these advancements according to the strategies used, hoping to facilitate further efforts to achieve site- and even enantioselective functionalization of remote C(sp3)–H bonds.

Recent advances in site-selective functionalization of remote aliphatic C–H bonds in organometallic pathways are summarized.  相似文献   

14.
The role of hypervalent iodine(iii) reagents in promoting alkoxylation of unactivated C(sp3)–H bonds catalyzed by palladium(ii) complexes     
Payam Abdolalian  Samaneh K. Tizhoush  Kaveh Farshadfar  Alireza Ariafard 《Chemical science》2021,12(20):7185
Although Pd(OAc)2-catalysed alkoxylation of the C(sp3)–H bonds mediated by hypervalent iodine(iii) reagents (ArIX2) has been developed by several prominent researchers, there is no clear mechanism yet for such crucial transformations. In this study, we shed light on this important issue with the aid of the density functional theory (DFT) calculations for alkoxylation of butyramide derivatives. We found that the previously proposed mechanism in the literature is not consistent with the experimental observations and thus cannot be operating. The calculations allowed us to discover an unprecedented mechanism composed of four main steps as follows: (i) activation of the C(sp3)–H bond, (ii) oxidative addition, (iii) reductive elimination and (iv) regeneration of the active catalyst. After completion of step (i) via the CMD mechanism, the oxidative addition commences with an X ligand transfer from the iodine(iii) reagent (ArIX2) to Pd(ii) to form a square pyramidal complex in which an iodonium occupies the apical position. Interestingly, a simple isomerization of the resultant five-coordinate complex triggers the Pd(ii) oxidation. Accordingly, the movement of the ligand trans to the Pd–C(sp3) bond to the apical position promotes the electron transfer from Pd(ii) to iodine(iii), resulting in the reduction of iodine(iii) concomitant with the ejection of the second X ligand as a free anion. The ensuing Pd(iv) complex then undergoes the C–O reductive elimination by nucleophilic attack of the solvent (alcohol) on the sp3 carbon via an outer-sphere SN2 mechanism assisted by the X anion. Noteworthy, starting from the five coordinate complex, the oxidative addition and reductive elimination processes occur with a very low activation barrier (ΔG 0–6 kcal mol−1). The strong coordination of the alkoxylated product to the Pd(ii) centre causes the regeneration of the active catalyst, i.e. step (iv), to be considerably endergonic, leading to subsequent catalytic cycles to proceed with a much higher activation barrier than the first cycle. We also found that although, in most cases, the alkoxylation reactions proceed via a Pd(ii)–Pd(iv)–Pd(ii) catalytic cycle, the other alternative in which the oxidation state of the Pd(ii) centre remains unchanged during the catalysis could be operative, depending on the nature of the organic substrate.

This work uses DFT calculations to explore Pd(ii)-catalysed iodine(iii)-mediated alkoxylation of unactivated C(sp3)–H bonds and reveals how important the isomerization is in triggering the oxidative addition of ArIX2 to Pd(ii).  相似文献   

15.
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   

16.
Palladium-catalyzed oxygenation of unactivated sp3 C-H bonds     
Desai LV  Hull KL  Sanford MS 《Journal of the American Chemical Society》2004,126(31):9542-9543
This communication describes a new palladium-catalyzed method for the oxygenation of unactivated sp3 C-H bonds. A wide variety of alkane substrates containing readily available oxime and/or pyridine directing groups are oxidized with extremely high levels of chemo-, regio-, and in some cases diastereoselectivity. The substrate scope of these reactions is discussed, and the high selectivities are rationalized on the basis of the requirements of putative palladium alkyl intermediates.  相似文献   

17.
Biomimetic catalytic aerobic oxidation of C–sp(3)–H bonds under mild conditions using galactose oxidase model compound CuIIL     
Xiao-Hui Liu  Hai-Yang Yu  Jia-Ying Huang  Ji-Hu Su  Can Xue  Xian-Tai Zhou  Yao-Rong He  Qian He  De-Jing Xu  Chao Xiong  Hong-Bing Ji 《Chemical science》2022,13(33):9560
Developing highly efficient catalytic protocols for C–sp(3)–H bond aerobic oxidation under mild conditions is a long-desired goal of chemists. Inspired by nature, a biomimetic approach for the aerobic oxidation of C–sp(3)–H by galactose oxidase model compound CuIIL and NHPI (N-hydroxyphthalimide) was developed. The CuIIL–NHPI system exhibited excellent performance in the oxidation of C–sp(3)–H bonds to ketones, especially for light alkanes. The biomimetic catalytic protocol had a broad substrate scope. Mechanistic studies revealed that the CuI-radical intermediate species generated from the intramolecular redox process of CuIILH2 was critical for O2 activation. Kinetic experiments showed that the activation of NHPI was the rate-determining step. Furthermore, activation of NHPI in the CuIIL–NHPI system was demonstrated by time-resolved EPR results. The persistent PINO (phthalimide-N-oxyl) radical mechanism for the aerobic oxidation of C–sp(3)–H bond was demonstrated.

A biomimetic catalytic approach for the aerobic oxidation of C–sp(3)–H bonds using galactose oxidase model compound was developed. EPR showed that the CuI-radical intermediate species was critical for O2 activation.  相似文献   

18.
Iridium(III)‐Catalyzed Regioselective Intermolecular Unactivated Secondary Csp3−H Bond Amidation          下载免费PDF全文
Dr. Xinsheng Xiao  Dr. Cheng Hou  Zhenhui Zhang  Prof. Zhuofeng Ke  Jianyong Lan  Prof. Huanfeng Jiang  Prof. Wei Zeng 《Angewandte Chemie (International ed. in English)》2016,55(39):11897-11901
For the first time, a highly regioselective intermolecular sulfonylamidation unactivated secondary Csp3?H bond has been achieved using IrIII catalysts. The introduced N,N’‐bichelating ligand plays a crucial role in enabling iridium–nitrene insertion into a secondary Csp3?H bond via an outer‐sphere pathway. Mechanistic studies and density functional theory (DFT) calculations demonstrated that a two‐electron concerted nitrene insertion was involved in this Csp3?H amidation process. This method tolerates a broad range of linear and branched‐chain N‐alkylamides, and provides efficient access to diverse γ‐sulfonamido‐substituted aliphatic amines.  相似文献   

19.
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   

20.
Palladium-catalyzed fluorination of carbon-hydrogen bonds     
Hull KL  Anani WQ  Sanford MS 《Journal of the American Chemical Society》2006,128(22):7134-7135
This communication describes the development of a new Pd-catalyzed method for the fluorination of carbon-hydrogen bonds. A key step of these transformations involves palladium-mediated carbon-fluorine coupling-a much sought after, but previously unprecedented, transformation. These reactions were successfully achieved under oxidative conditions using electrophilic N-fluoropyridinium reagents. Microwave irradiation in the presence of catalytic palladium acetate served as optimal conditions for the fluorination of C-H bonds in a variety of substituted 2-arylpyridine and 8-methylquinoline derivatives.  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