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 15 毫秒
1.
Despite significant advances made on the synthesis of indole derivatives through photochemical strategies during the past several years, the requirement of equivalent amounts of oxidants, bases or other additional additives has limited their practical applications in the synthesis of natural products and pharmaceuticals as environment-friendly processes. Herein, we report LED visible-light-induced redox neutral desulfonylative C(sp2)–H functionalization for the synthesis of N-substituted indoles with a broad scope through γ-fragmentation under mild conditions in the absence of any additional additive. The reaction mechanism paradigm has been investigated on the basis of deuterium labeling experiments, kinetic analysis, Hammett plotting analysis and DFT calculations.

LED visible-light-induced redox neutral desulfonylative C(sp2)–H functionalization for the synthesis of N-substituted indoles in the absence of any additional additive has been established on the basis of KIE, Hammett plotting and DFT calculations.  相似文献   

2.
An economical, solvent-free, and metal-free method for peptide synthesis via C–N bond cleavage using lactams has been developed. The method not only eliminates the need for condensation agents and their auxiliaries, which are essential for conventional peptide synthesis, but also exhibits high atom economy. The reaction is versatile because it can tolerate side chains bearing a range of functional groups, affording up to >99% yields of the corresponding peptides without racemisation or polymerisation. Moreover, the developed strategy enables peptide segment coupling, providing access to a hexapeptide that occurs as a repeat sequence in spider silk proteins.

An economical, solvent-free, and metal-free method for peptide synthesis via C–N bond cleavage using lactams has been developed.  相似文献   

3.
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   

4.
The applications of peptides and peptidomimetics have been demonstrated in the fields of therapeutics, diagnostics, and chemical biology. Strategies for the direct late-stage modification of peptides and peptidomimetics are highly desirable in modern drug discovery. Transition-metal-catalyzed C–H functionalization is emerging as a powerful strategy for late-stage peptide modification that is able to construct functional groups or increase skeletal diversity. However, the installation of directing groups is necessary to control the site selectivity. In this work, we describe a transition metal-free strategy for late-stage peptide modification. In this strategy, a linear aliphatic side chain at the peptide N-terminus is cyclized to deliver a proline skeleton via site-selective δ-C(sp3)–H functionalization under visible light. Natural and unnatural amino acids are demonstrated as suitable substrates with the transformations proceeding with excellent regio- and stereo-selectivity.

We have developed a visible light-promoted selective editing of a peptide skeleton via C–N bond formation at N-terminal aliphatic side chains. A proline skeleton was constructed in peptides under such transition metal free conditions.  相似文献   

5.
With an ever-growing emphasis on sustainable synthesis, aerobic C–H activation (the use of oxygen in air to activate C–H bonds) represents a highly attractive conduit for the development of novel synthetic methodologies. Herein, we report the air mediated functionalisation of various saturated heterocycles and ethers via aerobically generated radical intermediates to form new C–C bonds using acetylenic and vinyl triflones as radical acceptors. This enables access to a variety of acetylenic and vinyl substituted saturated heterocycles that are rich in synthetic value. Mechanistic studies and control reactions support an aerobic radical-based C–H activation mechanism.

Herein we disclose a novel method for the aerobic C–H activation of ethereal-based heterocycles to generate various α-functionalised building blocks.  相似文献   

6.
Forging carbon–carbon (C–C) linkage in DNA-encoded combinatorial library synthesis represents a fundamental task for drug discovery, especially with broad substrate scope and exquisite functional group tolerance. Here we reported the palladium-catalyzed Suzuki–Miyaura, Heck and Hiyama type cross-coupling via DNA-conjugated aryl diazonium intermediates for DNA-encoded chemical library (DEL) synthesis. Starting from commodity arylamines, this synthetic route facilely delivers vast chemical diversity at a mild temperature and pH, thus circumventing damage to fragile functional groups. Given its orthogonality with traditional aryl halide-based cross-coupling, the aryl diazonium-centered strategy expands the compatible synthesis of complex C–C bond-connected scaffolds. In addition, DNA-tethered pharmaceutical compounds (e.g., HDAC inhibitor) are constructed without decomposition of susceptible bioactive warheads (e.g., hydroxamic acid), emphasizing the superiority of the aryl diazonium-based approach. Together with the convenient transformation into an aryl azide photo-crosslinker, aryl diazonium''s DNA-compatible diversification synergistically demonstrated its competence to create medicinally relevant combinatorial libraries and investigate protein–ligand interactions in pharmaceutical research.

Taking advantage of aryl diazonium intermediates, this work reported a DNA-compatible C–C bond formation strategy, achieving broad substrate scope, exquisite functional group tolerance, and orthogonality to aryl halide-based coupling reactions.  相似文献   

7.
8.
In this study, first direct access to aryl alkyl sulfides employing 2‐phenylpropanal as coupling partner is reported. Diaryl disulfides react with this aldehyde in the presence of morpholine and produce the corresponding sulfide products in high yields. In another part, disulfides are in situ generated in the reaction mixture from aryl halides/CuI/Cyanodithioformate and coupled with 2‐phenylpropanal to access aryl alkyl sulfides.  相似文献   

9.
A versatile silylation of heteroaryl C–H bonds is accomplished under the catalysis of a well-defined spirocyclic NHC Ir(iii) complex (SNIr), generating a variety of heteroarylsilanes. A significant advantage of this catalytic system is that multiple types of intermolecular C–H silylation can be achieved using one catalytic system at α, β, γ, or δ positions of heteroatoms with excellent regioselectivities. Mechanistic experiments and DFT calculations indicate that the polycyclic ligand of SNIr can form an isolable cyclometalated intermediate, which leaves a phenyl dentate free and provides a hemi-open space for activating substrates. In general, favorable silylations occur at γ or δ positions of chelating heteroatoms, forming 5- or 6-membered C–Ir–N cyclic intermediates. If such an activation mode is prohibited sterically, silylations would take place at the α or β positions. The mechanistic studies would be helpful for further explaining the reactivity of the SNIr system.

A versatile silylation of heteroaryl C–H bonds is accomplished under the catalysis of a well-defined spirocyclic NHC Ir(iii) complex (SNIr), generating a variety of heteroarylsilanes.  相似文献   

10.
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.
The use of acylnitroso compounds of the general formula RCONO as enophiles in the formation of carbon-nitrogen bonds is described. Both inter- and intramolecular ene reactions have been studied. For the intermolecular examples, nitrosocarbonylmethane, thermally liberated from its Diels-Alder adduct with 9,10-dimethylanthracene, is reacted with various olefins giving the corresponding N-alkylhydroxamic acids in moderate to high yields, providing an efficient method for allylic amidation. The regiochemistry of the intermolecular reaction is observed to be the result of kinetic control, and the direction of addition is consistent with attack by the olefin on electron-deficient nitrogen. Several examples of intramolecular ene cyclization are demonstrated, providing efficient entry into both spiro and fused bicyclic nitrogen containing systems which can be viewed as derived from annulation of 5- and 6- membered nitrogen containing rings onto 5- and 6-membered carbocycles, respectively. Various examples of this hetero-annulation scheme are described. Experimental details are also privided describing typical reaction procedures.  相似文献   

12.
α-Arylated carboxylic acids, esters and amides are widespread motifs in bioactive molecules and important building blocks in chemical synthesis. Thus, straightforward and rapid access to such structures is highly desirable. Here we report an organophotocatalytic multicomponent synthesis of α-arylated carboxylic acids, esters and amides from exhaustive defluorination of α-trifluoromethyl alkenes in the presence of alkyltrifluoroborates, water and nitrogen/oxygen nucleophiles. This operationally simple strategy features a unified access to functionally diverse α-arylated carboxylic acids, esters, and primary, secondary, and tertiary amides through backbone assembly from simple starting materials enabled by consecutive C–F bond functionalization at room temperature. Preliminary mechanistic investigations reveal that the reaction operates through a radical-triggered three-step cascade process, which involves distinct mechanisms for each defluorinative functionalization of the C–F bond.

Here we report an organophotocatalytic synthesis of α-arylated carboxylic acids, esters and amides from exhaustive defluorination of α-trifluoromethyl alkenes in the presence of alkyltrifluoroborates, water and nitrogen/oxygen nucleophiles.  相似文献   

13.
A palladium-catalyzed C–H activation of acetylated anilines (acetanilides, 1,1-dimethyl-3-phenylurea, 1-phenylpyrrolidin-2-one, and 1-(indolin-1-yl)ethan-1-one) with epoxides using O-coordinating directing groups was accomplished. This C–H alkylation reaction proceeds via formation of a previously unknown 6,4-palladacycle intermediate and provides rapid access to regioselectively functionalized β-hydroxy products. Notably, this catalytic system is applicable for the gram scale mono-functionalization of acetanilide in good yields. The palladium-catalyzed coupling reaction of the ortho-C(sp2) atom of O-coordinating directing groups with a C(sp3) carbon of chiral epoxides offers diverse substrate scope in good to excellent yields. In addition, further transformations of the synthesized compound led to biologically important heterocycles. Density functional theory reveals that the 6,4-palladacycle leveraged in this work is significantly more strained (>10 kcal mol−1) than the literature known 5,4 palladacycles.

The combined experimental and computational study on palladium-catalyzed regioselective C–H functionalization of O-coordinating directing groups with epoxides is described.  相似文献   

14.
The mechanical strength of individual polymer chains is believed to underlie a number of performance metrics in bulk materials, including adhesion and fracture toughness. Methods by which the intrinsic molecular strength of the constituents of a given polymeric material might be switched are therefore potentially useful both for applications in which triggered property changes are desirable, and as tests of molecular theories for bulk behaviors. Here we report that the sequential oxidation of sulfide containing polyesters (PE-S) to the corresponding sulfoxide (PE-SO) and then sulfone (PE-SO2) first weakens (sulfoxide), and then enhances (sulfone), the effective mechanical integrity of the polymer backbone; PE-S ∼ PE-SO2 > PE-SO. The relative mechanical strength as a function of oxidation state is revealed through the use of gem-dichlorocyclopropane nonscissile mechanophores as an internal standard, and the observed order agrees well with the reported bond dissociation energies of C–S bonds in each species and with the results of CoGEF modeling.

The mechanical strength of individual polymer chains is believed to underlie a number of performance metrics in bulk materials, including adhesion and fracture toughness.  相似文献   

15.
Introducing spin onto organic ligands that are coordinated to rare earth metal ions allows direct exchange with metal spin centres. This is particularly relevant for the deeply buried 4f-orbitals of the lanthanide ions that can give rise to unparalleled magnetic properties. For efficacy of exchange coupling, the donor atoms of the radical ligand require high-spin density. Such molecules are extremely rare owing to their reactive nature that renders isolation and purification difficult. Here, we demonstrate that a 2,2′-azopyridyl (abpy) radical (S = 1/2) bound to the rare earth metal yttrium can be realized. This molecule represents the first rare earth metal complex containing an abpy radical and is unambigously characterized by X-ray crystallography, NMR, UV-Vis-NIR, and IR spectroscopy. In addition, the most stable isotope 89Y with a natural abundance of 100% and a nuclear spin of ½ allows an in-depth analysis of the yttrium–radical complex via EPR and HYSCORE spectroscopy. Further insight into the electronic ground state of the radical azobispyridine-coordinated metal complex was realized through unrestricted DFT calculations, which suggests that the unpaired spin density of the SOMO is heavily localized on the azo and pyridyl nitrogen atoms. The experimental results are supported by NBO calculations and give a comprehensive picture of the spin density of the azopyridyl ancillary ligand. This unexplored azopyridyl radical anion in heavy element chemistry bears crucial implications for the design of molecule-based magnets particularly comprising anisotropic lanthanide ions.

Unambiguous characterization of the first 2,2′-azobispyridine radical-containing rare earth metal complex through X-ray crystallography, DFT computations, EPR and HYSCORE spectroscopy.  相似文献   

16.
Correction for ‘Cu-catalyzed C–C bond formation of vinylidene cyclopropanes with carbon nucleophiles’ by Jichao Chen et al., Chem. Sci., 2019, 10, 10601–10606.

We regret that in the original article the structure of compound 1 in Tables 1–3 was incorrect. The correct structure is given below.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

17.
C–O bond cleavage is often a key process in defunctionalization of organic compounds as well as in degradation of natural polymers. However, it seldom occurs regioselectively for different types of C–O bonds under metal-free mild conditions. Here we report a facile chemo-selective cleavage of the α-C–O bonds in α-carboxy ketones by commercially available pinacolborane under the catalysis of diazaphosphinane based on a mechanism switch strategy. This new reaction features high efficiency, low cost and good group-tolerance, and is also amenable to catalytic deprotection of desyl-protected carboxylic acids and amino acids. Mechanistic studies indicated an electron-transfer-initiated radical process, underlining two crucial steps: (1) the initiator azodiisobutyronitrile switches originally hydridic reduction to kinetically more accessible electron reduction; and (2) the catalytic phosphorus species upconverts weakly reducing pinacolborane into strongly reducing diazaphosphinane.

Diazaphosphinyl radical-catalyzed chemo-selective deoxygenation of α-carboxy ketones with pinacolborane was achieved through the mechanism switch from direct to stepwise hydride transfer of diazaphosphinane.

The importance of reductive deoxygenation can be gauged by the wide use of Barton–McCombie deoxygenation in organic syntheses.1 Such C–O bond cleavage is also a crucial step in the degradation of natural polymers (e.g., sugars and lignins) to recycle sustainable resources.2 Consequently, a great variety of methodologies were explored for activation of these strong C–O bonds.3 Among them, deoxygenation of α-acyloxy ketones3b,c,4 (represented by benzoin derivatives, stemming from simple aldehydes via benzoin condensation5) has attracted considerable attention, because it may provide a facile way for accessing commonly useful building blocks (aryl ketones).6 As known, benzoin derivatives bear two types of C–O bonds—the carbonyl π-C Created by potrace 1.16, written by Peter Selinger 2001-2019 O bond and the benzyl σ-C–O bond (Scheme 1). While reduction of the carbonyl π-C Created by potrace 1.16, written by Peter Selinger 2001-2019 O bonds has been well established through transition metal-7 or Lewis acid-mediated8 hydride transfers, chemo-selective cleavage of the benzyl σ-C–O bonds is challenging and has seldom been achieved.9 The later process is occasionally seen, however, in some radical or electron reductions, but toxic tin hydrides1c or aggressive metal reagents (like Raney nickel,10 zinc dust,11etc.) are inevitably employed.Open in a separate windowScheme 1Possible reactive sites for benzoin reduction.The recent successful development of super electron donors (SEDs),4,12 which are defined as ground-state organic electron-donors capable of reducing aryl halides to aryl radicals or aryl anions,13 and photocatalytic systems3b,c may provide alternative protocols for reductive cleavage of the σ-C–O bonds in O-acetylated benzoin. However, these electron transfer-initiated reductions also suffer from some drawbacks, such as, excessive use of SEDs and their tedious synthetic procedures, expensive photoredox catalysts and ligands, and group-tolerance issues. In fact, there have been few reports to date on metal-free systems for efficiently catalytic deoxygenation with commercially available inexpensive reductants.3h Given the ubiquity of C–O bonds in nature, it is still an unmet need for development of efficient and economical methods for their degradation. N-Heterocyclic phosphines (NHPs)8c,14 have recently found plentiful applications in hydridic reductions8b,15 owing to their outstanding hydricity.16 However, this seems to blind one to search for their other promising reaction patterns, like radical and electron transfer reactivities. Up to now, the catalytic potential of NHPs in radical or electron reductions has never been explored. Given the logical understanding that a deliberately manipulated mechanism variation usually leads to diverse reactivity and selectivity, we anticipate that an intended mechanism switch for NHP-based reactions from the conventional hydride transfer to an alternative electron transfer might provide a chance for originally inaccessible chemo-selectivity in the reduction of the substrates bearing multiple reactive sites. As known from previous studies, NHPs could transfer a hydride ion to carbonyl C Created by potrace 1.16, written by Peter Selinger 2001-2019 O bonds to deliver the corresponding alcohol counterparts (Scheme 2a).8b This is indeed what we have seen. When NHPs are mixed with O-acetylated benzoins, an exclusive hydridic reduction of the carbonyl π-C Created by potrace 1.16, written by Peter Selinger 2001-2019 O bonds is observed, leaving the benzyl σ-C–O bonds intact. How could we make the propensity of NHP reduction to switch from the original hydridic path to a radical one? Inspired by our recent findings that NHPs are also capable of serving as good hydrogen-atom donors (by P–H bond homolysis) and their corresponding phosphinyl radicals are excellent electron donors17 (Scheme 2b, bottom), we envisioned that if phosphinyl radicals can be in situ generated, their super electron-donicity may promote the initial electron transfer to benzoin, and trigger the subsequent benzyl σ-C–O bond scission. If this is realizable, chemo-selective deoxygenation of benzoin derivatives with NHPs may be achieved via such a mechanism switch.Open in a separate windowScheme 2Chemical transformations of N-heterocyclic phosphines NHPs in reduction reactions.It is noted that the phosphorus species NHP-OR′ is produced in either the hydride or electron reduction of benzoin derivatives (Scheme 2c). Based on the previous knowledge that NHP-OR′ can be recycled back to NHP through a σ-bond metathesis between its exocyclic P–O bond and the B–H bond of pinacolborane (HBpin),8b we envisioned that the present deoxygenation may operate in a catalytic fashion with readily available HBpin as the terminal reductant to avoid the use of stoichiometric NHP. To verify this plot, we chose dimethyl 4,4′-(1-acetoxy-2-oxoethane-1,2-diyl)dibenzoate X as the testing substrate, and 1,3-di-tert-butyl-1,3,2-diazaphosphinane 1a as the catalyst based on its compatible reducing capacity (Scheme 3, for structure of 1a, cf.Table 1).17a It is observed that, under the previously established catalytic conditions for carbonyl reduction (20 mol% of 1a and 1.2 equiv. HBpin),8b the product X1 of hydridic reduction was obtained in 86% yield and 1 : 0.69 of diastereomer ratio in 12 h. On the other hand, when 10% azodiisobutyronitrile (AIBN) was added as a radical initiator, the σ-C–O bonds were, indeed, selectively cleaved to give the anticipated product X2 in 87% yield. This distinct chemo-selectivity did echo our proposed mechanism switch from the direct hydridic pathway to an electron reduction. In the following, we report this catalytic transformation in a more inclusive fashion. To our best knowledge, this is the first example of catalytic electron reduction mediated by NHPs.Optimization of reaction conditions for C–O bond cleavage
EntryCatalystConditionaYieldb
1 1a Standard condition92%
2 1a 10 mol% 1a62%
3 1b Standard condition<10%c
4 1c Standard condition<5%c
5 1d Standard condition<5%c
6 1e Standard condition46%
7 1a NH3BH3 as reductant<5%c
8 1a No AIBN<5%c
9 1a No heat<5%c
10Standard condition<5%c
Open in a separate windowaConditions for C–O bond activation: 2 (0.4 mmol), AIBN (0.04 mmol), 1a (0.08 mmol), HBpin (0.48 mmol) in toluene (1.0 mL).bIsolated yields.cNMR yields using 1,3,5-trimethoxybenzene as the internal standard.Open in a separate windowScheme 3Chemo-selectively reductive cleavage of C–O bonds in O-acetylated benzoin X by diazaphosphinane 1a.To verify the necessity of each component in the above catalytic system, a series of comparative experiments were conducted. We commenced the condition optimization with simple O-acetylated benzoin 2a as the standard substrate – an attractive precursor for accessing α-aryl ketone which is a common pharmacophore and also present in numerous biologically active natural products.18 As shown in Table 1, treatment of 2a with 20 mol% catalyst 1a, 10 mol% initiator AIBN and 1.2 equiv. HBpin in toluene solution harvested the product 1,2-diphenylethanone 3a in 92% isolated yield (entry 1). Decreasing the catalyst loading led to an inferior result (62%, entry 2). Replacement of 1a with structurally similar 1b gave a much lower yield (<10%, entry 3), which is primarily because the weak reducing capacity of 1b-derived phosphinyl radical (Eox = −1.94 V vs. Fc in MeCN)17a prevents its electron transfer to 2a. The same reason can be applied to account for the poor results of 1c and 1d catalysts (<5%, entry 4 and 5). When a stronger hydride donor 1e was employed, a moderate yield (46%, entry 6) was obtained along with 40% byproduct of direct hydride transfer. This may be because enhancing the reducing ability of 1e can simultaneously accelerate its hydride transfer to carbonyl groups, which competes with the electron transfer between its derived phosphinyl radical and benzoin. Commercially available borane ammonia (NH3·BH3) was also examined, furnishing no desired product (<5%, entry 7). In addition, the absence of AIBN, heating or catalyst 1a cannot render efficient C–O bond cleavage (entry 8–10). Therefore, 20 mol% 1a, 10 mol% AIBN and 1.2 equiv. HBpin in toluene solution were eventually used as the standard conditions.Next, we explored the substrate scope starting with different benzoin derivatives 2 (Scheme 4). Besides the acetate, the reaction presented here also worked very well for other leaving groups, such as pivalate 2b, benzoate 2c and 4-cyanophenolate 2d, affording the product 1,2-diphenylethanone 3a in good to excellent yields (72–99%). Then, a series of benzoin derivatives with diverse substituents (2e–i) were synthesized to examine the functional group tolerance. As seen, the substrates with electron-withdrawing F (2e) and Cl (2f) groups gave almost quantitative yields (99%). Noteworthily, in contrast to the previously reported Ru-based photocatalytic deoxygenation,3b the reaction presented here could tolerate the ester group well and gave 3g in 87% yield. As for electron-donating substituents, such as methyl (2h) and methoxy groups (2i), the reaction yields were slightly reduced (72% and 75%), which may be ascribed to their lower reduction potentials. Replacement of the phenyl group with naphthyl (2j) afforded the product 3j in a good yield (80%). Furthermore, some cross-benzoin analogues were also investigated. The unsymmetrical counterpart 2k gave 3k in a moderate yield (62%). Similarly, heteroaromatic substrates (2l and 2m) generated corresponding products in 65% and 55% yields, respectively. Additionally, we examined the acyloin derivative 2n which was previously reported to give a base promoted aldol-type cyclization byproduct in the SED system.4 Notably, our conditions are mild enough for selective cleavage of its C–O bond in a moderate yield (52%), although 1a was necessarily employed as a stoichiometric reductant. However, the analogs 2o and 2p gave poor yields, possibly due to the less stability of their corresponding radical intermediates.Open in a separate windowScheme 4Substrate scope for C–O bond activation. Conditions unless otherwise specified: 2 (0.4 mmol), AIBN (0.04 mmol), 1a (0.08 mmol), HBpin (0.48 mmol) in toluene (1.0 mL). Isolated yields were given. [a] 0.4 mmol of 1a was used. [b] NMR yields using 1,3,5-trimethoxybenzene as the internal standard.Desyl is a classical protection group in organic chemistry and biology.3c,19 We wondered whether the same reaction could serve as a practical strategy to realize catalytic deprotection of various desyl-protected carboxylic acids under metal-free conditions. To assess its feasibility, we tested some carboxylic acids, including aromatic, aliphatic, and amino acids. The results revealed a good tolerance for the present method. As shown in Scheme 5, the substrate 4a gave benzoic acid 5a in a quantitative yield (99%) under the standard conditions. And, the reaction was compatible well with the susceptive acetal moiety and furnished 5b in a good yield (88%). This result indicated the high selectivity of our system to the targeted C–O bond. Substrate 4c with electron-donating groups was also found feasible, and afforded the deprotected product 5c in a slightly lower yield (74%). Interestingly, for the isophthalic acid system whose two carboxylic groups were both protected by desyl groups, the deprotection was proved to be highly reactive, and afforded the fully-deprotected product 5d in an excellent yield (92%). 1-Naphthoic acid 5e could be obtained in a good yield of 85% after deprotection. Furthermore, the deprotection of aliphatic acids 4f furnished 5f in an almost quantitative yield (99%). However, similar 5g with an additional conjugated double bond was obtained in a diminished yield (71%). More importantly, our protocol is also applicable in amino acid systems. As seen, 5h was obtained in 90% yield with conformational retention, and the deprotection of 4i was not affected by other commonly-used protecting group Boc, giving the product 5i in 91% yield.Open in a separate windowScheme 5Substrate scope for catalytic deprotection with desyl as the protecting group. Conditions: 4 (0.4 mmol), AIBN (0.04 mmol), 1a (0.08 mmol), HBpin (0.48 mmol) in toluene (1.0 mL). Isolated yields were given. [a] 0.96 mmol of HBpin was used.Furthermore, we investigated the reaction mechanism by taking substrate 2a as the template compound. As previously established in SED systems, benzoin derivates were deemed to be reduced via a successive double-electron transfer mechanism, affording enolates as the intermediates which eventually captured a proton from the solvent.4 Different from this double-electron transfer pathway, our system would operate in a single-electron reduction mechanism, however. This was deduced from the fact that one equivalent reductant 1a could afford almost quantitative product 3a (eqn (1)). Consequently, the present process clearly displays a superiority in atom economy over the previous SED systems. With respect to the catalyst regeneration, we conducted the reaction of the intermediate 1a-OAc with HBpin in toluene-d8 at room temperature (eqn (2)). Through monitoring the 1H NMR and 31P NMR spectra of the reaction mixture, it is found that as the intermediate 1a-OAc gradually disappeared in about one hour (see ESI for details), 1a P–H bond was formed synchronously. This confirmed the effective regeneration of 1a from HBpin. Moreover, when 20 mol% 1a-OAc was used as the catalyst, the reduction could also work quite well to furnish the desired product in 63% yield (eqn (3)). Therefore, 1a-OAc can be regarded as an intermediate in the catalytic cycle to regenerate 1a. In addition, to exclude the possibility of a radical chain process, that is, a direct oxygen abstraction from benzoin by the phosphinyl radical, DFT calculations were conducted (see ESI for details). The results showed that 1a-[P]˙ and 1b-[P]˙ have a comparable ability in abstracting the oxygen atom (with an energy difference of 0.78 kcal mol−1, eqn (4)). This failed to explain the disparate yields for 1a and 1b systems (90% vs. <10%). Besides, the difference in the oxidation potentials of 1a-[P]˙ (Eox = −2.39 V) and of 1b-[P]˙ (Eox = −1.94 V)17a is consistent well with the observed diverse reduction results. All these preferentially support an electron-transfer initiated reduction.1234Based on the above control experiments and the computation, we outlined the catalytic cycle for reductive cleavage of C–O bonds in Scheme 6. The reaction is turned on by the isobutyronitrile radical, which abstracts a hydrogen-atom from diazaphosphinane 1a to produce the actual reductant phosphinyl radical. This potent electron donor (Eox = −2.39 V) then transfers an electron to 2a (Ered = ∼−2.3 V),3c,4 furnishing the ketyl radical anion 6 and the corresponding phosphonium cation. The σ-C–O bond of the intermediate 6 is readily cleaved to afford the ketyl 7 and acetate. The ketyl 7 would not be further reduced into the corresponding enolate, but instead abstracts a hydrogen-atom from 1a and simultaneously triggers the next catalytic cycle. Meanwhile, a combination of the stable phosphonium cation with acetate produces 1a-OAc, which could regenerate the catalyst 1a from the terminal reductant HBpin. Accordingly, the success of the present deoxygenation primarily attributes to two crucial factors: the mechanism switch from the originally hydridic reduction to a kinetically more accessible electron reduction by the initiator azodiisobutyronitrile, and the “upconversion” of weakly reducing HBpin into strongly reducing diazaphosphinane by catalytic phosphorus species.20 Moreover, in our systems, HBpin serves as both the electron and hydrogen-atom sources, namely the apparent hydride donor. This is different from what was known for the previous SED systems, in which the reductants only provide the electron, and hence, extraneous hydrogen sources are necessarily employed.Open in a separate windowScheme 6Proposed mechanism of C–O bond cleavage.  相似文献   

18.
The practical synthesis of 4,4,4-trifluorocrotonaldehyde (1) and its application to enantioselective 1,4-additions are described. The organocatalytic 1,4-addition of 1 with several nucleophiles such as heteroaromatics, alkylthiols and aldoximes afforded the corresponding products, each bearing a trifluoromethylated stereogenic center with high optical purity. A resulting product was converted into an MAO-A inhibitor, befloxatone.  相似文献   

19.
The one-electron oxidation of metal thiolates results in an increased oxidation state of the metal ion or the formation of a sulfur-based, thiyl radical in limiting extremes. For complexes with highly covalent M-S bonds, the unpaired electron may be delocalized over the metal and the sulfur, yielding a metal-stabilized thiyl radical. Oxidation of the metal thiolate precursors [Ru(DPPBT)(3)](-), [Ru-1](-), and Re(DPPBT)(3), Re-1 (DPPBT = diphenylphosphinobenzenethiolate), generates metal-stabilized thiyl radicals that react with alkenes to yield dithioether-metal products. Alkene addition to [Ru-1](+) and [Re-1](+) is symmetry-allowed due to the meridional arrangement of the DPPBT chelates. Combined bulk electrolysis and cyclic voltammetry experiments reveal the addition of alkenes to [Ru-1](+) as an irreversible process with experimentally determined rate constants ranging from 4.6(5) × 10(7) M(-1) s(-1) for electron-rich alkenes to 2.7(2) × 10(4) M(-1) s(-1) for electron-poor alkenes. Rate constants for cyclic alkenes range from 4(2) × 10(7) to 2.9(3) × 10(3) M(-1) s(-1). Chemical oxidation of [Ru-1](-) by ferrocenium hexafluorophosphate (FcPF(6)) in the presence of m-methylstyrene or p-methylstyrene yields the dithioether complexes [Ru-1·m-methylstyrene](+) and [Ru-1·p-methylstyrene](+), respectively. Each complex was crystallized and the structure determined by single-crystal X-ray diffraction. (31)P NMR of the samples reveals a major and minor product, each displaying a second-order spectrum. The oxidized intermediate [Re-1](+) binds alkenes reversibly with equilibrium binding constants that vary with the complex charge from 1.9 × 10(-11) M(-1) for n = 0 to 4.0 M(-1) for n = +1 to 2.5 × 10(9) M(-1) for n = +2. The three binding regimes are separated by 240 mV. Crystalline samples of [Re-1·C(2)H(4)](2+) are obtained upon chemical oxidation of Re-1 with silver hexafluorophosphate (AgPF(6)) in the presence of ethylene. Strategies for the addition of alkenes to other metal-stabilized thiyl radicals are suggested.  相似文献   

20.
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