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 375 毫秒
1.
2.
Halogen-bonded (XB) complexes between halide anions and a cyclopropenylium-based anionic XB donor were characterized in solution for the first time. Spontaneous formation of such complexes confirms that halogen bonding is sufficiently strong to overcome electrostatic repulsion between two anions. The formation constants of such “anti-electrostatic” associations are comparable to those formed by halides with neutral halogenated electrophiles. However, while the latter usually show charge-transfer absorption bands, the UV-Vis spectra of the anion–anion complexes examined herein are determined by the electronic excitations within the XB donor. The identification of XB anion–anion complexes substantially extends the range of the feasible XB systems, and it provides vital information for the discussion of the nature of this interaction.

Spontaneous formation of “anti-electrostatic” complexes in solution demonstrates that halogen bonding can be sufficiently strong to overcome anion–anion repulsion when the latter is attenuated by the polar medium.

Halogen bonding (XB) is an attractive interaction between a Lewis base (LB) and a halogenated compound, exhibiting an electrophilic region on the halogen atom.1 It is most commonly related to electrostatic interaction between an electron-rich species (XB acceptor) and an area of positive electrostatic potential (σ-holes) on the surface of the halogen substituent in the electrophilic molecule (XB donor).2 Provided that mutual polarization of the interacting species is taken into account, the σ-hole model explains geometric features and the variation of stabilities of XB associations, especially in the series of relatively weak complexes.3 Based on the definition of halogen bonding and its electrostatic interpretation, this interaction is expected to involve either cationic or neutral XB donors. Electrostatic interaction of anionic halogenated species with electron-rich XB acceptors, however, seems to be repulsive, especially if the latter are also anionic. Yet, computational analyses predicted that halogen bonding between ions of like charges, called “anti-electrostatic” halogen bonding (AEXB),4 can possibly be formed5–12 and the first examples of AEXB complexes formed by different anions, i.e. halide anions and the anionic iodinated bis(dicyanomethylene)cyclopropanide derivatives 1 (see Scheme 1) or the anionic tetraiodo-p-benzoquinone radical, were characterized recently in the solid state.13,14 The identification of such complexes substantially extends the range of feasible XB systems, and it provides vital information for the discussion of the nature of this interaction. Computational results, however, significantly depend on the used methods and applied media (gas phase vs. polar environment and solvation models) and the solid state arrangements of the XB species might be affected by crystal forces and/or counterions. Unambiguous confirmation of the stability of the halogen-bonded anion–anion complexes and verification of their thermodynamic characteristics thus requires experimental characterization of the spontaneous formation of such associations in solution. Still, while the solution-phase complexes formed by hydrogen bonding between two anionic species were reported previously,15–17 there is currently no example of “anti-electrostatic” XB in solution.Open in a separate windowScheme 1Structures of the XB donor 1 and its hydrogen-substituted analogue 2.To examine halogen bonding between two anions in solution, we turn to the interaction between halides and 1,2-bis(dicyanomethylene)-3-iodo-cyclopropanide 1 (Scheme 1). Even though this compound features a cationic cyclopropenylium core, it is overall anionic, and calculations have demonstrated that its electrostatic potential is universally negative across its entire surface.13 The solution of 1 (with tris(dimethylamino)cyclopropenium (TDA) as counterion) in acetonitrile is characterized by an absorption band at 288 nm with ε = 2.3 × 104 M−1 cm−1 (Fig. 1). As LB, we first applied iodide anions taken as a salt with n-tetrabutylammonium counter-ion, Bu4NI. This salt does not show absorption bands above 290 nm, but its addition to a solution of 1 led to a rise of absorption in the 290–350 nm range (Fig. 1). Subtraction of the absorption of the individual components from that of their mixture produced a differential spectrum which shows a maximum at about 301 nm (insert in Fig. 1). At constant concentration of the XB donor (1) and constant ionic strength, the intensity of the absorption in the range of 280–300 nm (and hence differential absorbance, ΔAbs) rises with increasing iodide concentration (Fig. S1 in the ESI). This suggests that the interaction of iodide with 1 results in the formation of the [1, I]-complex which shows a higher absorptivity in this spectral range (eqn (1)):1 + X ⇌ [1, X]1Open in a separate windowFig. 1Spectra of acetonitrile solutions with constant concentration of 1 (0.60 mM) and various concentrations of Bu4NI (6.0, 13, 32, 49, 75, 115 and 250 mM, solid lines from the bottom to the top). The dashed lines show spectra of the individual solutions 1 (c = 0.60 mM, red line) and Bu4NI (c = 250 mM, blue line). The ionic strength was maintained using Bu4NPF6. Insert: Differential spectra of the solutions obtained by subtraction of the absorption of the individual components from the spectra of their corresponding mixtures.To clarify the mode of interaction between 1 and iodide in the complex, we also performed analogous measurements with the hydrogen-substituted compound 2 (see Scheme 1). The addition of iodide to a solution of 2 in acetonitrile did not increase the absorption in the 280–300 nm spectral range. Instead, some decrease of the absorption band intensity of 2 with the increase of concentration of I anions was observed (Fig. S2 in the ESI). Such changes are related to a blue shift of this band resulting from the hydrogen bonding between 2 and iodide (formation of hydrogen-bonded [2, I] complex is corroborated by the observation of the small shift of the NMR signal of the proton of 2 to the higher ppm values in the presence of I anions, see Fig. S3 in the ESI).§ Furthermore, since H-compound 2 should be at least as suitable as XB donor 1 to form anion–π complexes with the halide, this finding (as well as solid-state and computational data) rules out that any increase in absorption in this region observed with the I-compound 1 may be due to this alternative interaction.Likewise, the addition of NBu4I to a solution of TDA cations taken as a salt with Cl anions did not result in an increase in the relevant region. Hence, we could also rule out anion–π interactions with the TDA counter-ions as source of the observed changes, which is in line with previous reports on the electron-rich nature of TDA.18All these observations (supported by the computational analysis, vide infra) indicate that the [1, I] complex (eqn (1)) is formed via halogen bonding of I with iodine substituents in 1. The changes in the intensities of the differential absorption ΔAbs as a function of the iodide concentration (with constant concentration of XB donor (1) as well as constant ionic strength) are well-modelled by the 1 : 1 binding isotherm (Fig. S1 in the ESI). The fit of the absorption data produced a formation constant of K = 15 M−1 for the [1, I] complex (Table 1).|| The overlap with the absorption of the individual XB donor hindered the accurate evaluation of the position and intensity of the absorption band of the corresponding complex which is formed upon LB-addition to 1. As such, the values of Δλmax shown in Table 1 represent a wavelength of the largest difference in the absorptivity of the [1, I] complex and individual anion 1, and Δε reflects the difference of their absorptivity at this point (see the ESI for the details of calculations).Equilibrium constants and spectral characteristics of the complexes of 1 with halide anions X
Complexa K [M−1]Δλmaxc [nm]10−3Δεd [M−1 cm−1]
1·I15 ± 23029.0
1·Ib8 ± 23038.0
1·Br17 ± 23023.7
1·Cl40 ± 83023.0
Open in a separate windowaAll measurements performed in CH3CN at 22 °C, unless stated otherwise.bIn CH2Cl2.cWavelength of the maximum of the differential spectra.dDifferences in extinction coefficients of XB [1, I] complex and individual 1 at Δλmax.Since earlier computational studies demonstrated substantial dependence of formation of the AEXB complexes on polarity of the medium,6–12 interaction between 1 and I anions was also examined in dichloromethane. The spectral changes in this moderately-polar solvent were analogous to that in acetonitrile (Fig. S4 in the ESI). * The values for the formation constants of the [1, I] complex and Δε (obtained from the fitting of the ΔAbs vs. [I] dependence) in CH2Cl2 are lower than those in acetonitrile (Table 1). This finding is in line with the computational studies,6–12 predicting stronger binding in more polar solvents.The addition of bromide or chloride salts to an acetonitrile solution of 1 caused changes in the UV-Vis range which were generally similar to that observed upon addition of iodide. The variations of the magnitude of the differential absorption intensities with the increase in the bromide or chloride concentrations are less pronounced than that observed upon addition of iodide (in agreement with the results of the DFT computations of the UV-Vis spectra of the complexes, vide infra). Yet, they could also be fitted using 1 : 1 binding isotherms (see Fig. S5 and S6 in the ESI). The formation constants of the corresponding [1, Br] and [1, Cl] complexes resulted from the fitting of these dependencies are listed in Table 1. The values of K (which correspond to the free energy changes of complex formation in a range of −6 to −8 kJ mol−1) are comparable to those reported for complexes of neutral monodentate bromo- or iodosubstituted aliphatic or aromatic electrophiles with halides.19–22 Thus, despite the “anti-electrostatic” nature of XB complexes between two anions, the stabilities of such associations are similar to that observed with the most common neutral XB donors.In contrast to the similarity in thermodynamic characteristics, the UV-Vis spectral properties of the complexes of the anionic XB donor 1 with halides are substantially different from that reported for the analogous associations with the neutral XB donors. Specifically, a number of earlier studies revealed that intermolecular (XB or anion–π) complexes of halide anions are characterized by distinct absorption bands, which could be clearly segregated from the absorption of the interacting species.21–23 If the same neutral XB donor was used, the absorption bands of the corresponding complexes with chloride were blue shifted, and absorption bands of the complexes with iodide as LB were red shifted as compared to the bands of complexes with bromide. For example, XB complexes of CFBr3 with Cl, Br or I show absorption band maxima at 247 nm, 269 nm and 312 nm, respectively (individual CFBr3 is characterized by an absorption band at 233 nm).21 Within a framework of the Mulliken charge-transfer theory of molecular complexes,24 such an order is related to a rise in the energy of the corresponding HOMO (and electron-donor strength) from Cl to Br and to I anions. In the complexes with the same electron acceptor, this is accompanied by a decrease of the HOMO–LUMO gap, and thus, a red shift of the absorption band. The data in Table 1 shows, however, that the maxima of differential absorption spectra for these systems are observed at roughly the same wavelength. To clarify the reason for this observation, we carried out computational analysis of the associations between 1 and halide anions.The DFT optimization†† at M06-2X/def2-tzvpp level with acetonitrile as a medium (using PCM solvation model)25 produced thermodynamically stable XB complexes between 1 and I, Br or Cl anions (they were similar to the complexes which were obtained earlier via M06-2X/def2-tzvp computations with SMD solvation model13). The calculated structure of the [1, I] complex is shown in Fig. 2 and similar structures for the [1, Br] and [1, Cl] are shown in Fig. S7 in the ESI.Open in a separate windowFig. 2Optimized geometries of the [1, I] complex with (3, −1) bond critical points (yellow spheres) and the bond path (green line) from the QTAIM analysis. The blue–green disc indicates intermolecular attractive interactions resulting from the NCI treatments (s = 0.4 a.u. isosurfaces, color scale: −0.035 (blue) < ρ < 0.02 (red) a.u.).QTAIM analysis26 of these structures revealed the presence of the bond paths (shown as the green line) and (3, −1) bond critical points (BCPs) indicating bonding interaction between iodine substituent of 1 and halide anions. Characteristics of these BCPs (electron density of about 0.015 a.u., Laplacians of electron density of about 0.05 a.u. and energy density of about 0.0004 a.u., see Table S1 in the ESI) are typical for the moderately strong supramolecular halogen bonds.27 The Non-Covalent Interaction (NCI) Indexes treatment28 produced characteristic green–blue discs at the critical points'' positions, confirming bonding interaction in all these complexes.Binding energies, ΔE, for the [1, X] complexes are listed in Table 2. They are negative and their variations are consistent with the changes in experimental formation constants measured with three halide anions in Table 1. The ΔE value for [1, I] calculated in dichloromethane is also negative. Its magnitude is lower than that in acetonitrile, in agreement with the smaller formation constant of [1, I] in less polar dichloromethane.Calculated characteristics of the [1, X] complexesa
ComplexΔE, kJ mol−1 λ max,c nm10−4ε,c M−1 cm−1Δλmax,d M−1 cm−110−3Δε,d M−1 cm−1
1·I−14.22525.7025514
1·Ib−4.72536.07
1·Br−14.82525.022537.4
1·Cl−16.22514.782495.3
Open in a separate windowaIn CH3CN, if not noted otherwise.bIn CH2Cl2.cExtinction coefficient for the lowest-energy absorption band of the complex.dPosition and extinction coefficient of the differential absorption (see Fig. 3).The TD DFT calculations of the individual XB donor 1 and its complexes with halides (which were carried at the same level as the optimizations) produced strong absorption bands in the UV range (Fig. 3). The calculated spectrum of the individual anion 1 (λmax = 252 nm and ε = 4.27 × 104 M−1 cm−1) is characterized by somewhat higher energy and intensity of the absorption band than the experimental one, but the differences of about 0.6 eV in energy and about 0.3 in log ε are common for the TD DFT calculations.Open in a separate windowFig. 3Calculated spectra of 1 and its complexes (as indicated). The dashed lines show differential absorption obtained by subtraction of absorption of 1 from the absorption of the corresponding complex.The TD DFT calculations of the XB complexes with all three anions produced absorption bands at essentially the same wavelength as that of the individual XB donor 1, but their intensities were higher (in contrast, the hydrogen-bonded complex of 2 with iodide showed absorption band with slightly lower intensity than that of individual 2). The differential spectra obtained by subtraction of the spectra of individual anion 1 from the spectra of the complexes are shown in Fig. 3, and their characteristics are listed in Table 2. Similarly to the experimental data in Table 1, the calculated values of Δλmax are very close in complexes with different halides, and values of Δε are increasing in the order 1·Cl < 1·Br < 1·I.An analysis of the calculated spectra of the complexes revealed that the distinction in spectral characteristics of the XB complexes of anionic and neutral XB donors with halides are related to the differences in the molecular orbital energies of the interacting species. Specifically, the energy of the highest occupied molecular orbital (HOMO) of the anionic XB donor 1 is higher than the energies of the HOMOs of I, Br and Cl, and the energy of the lowest unoccupied molecular orbital (LUMO) of 1 is lower than those of the halides (Table S2 in the ESI). As such, the lowest-energy electron excitations (with the substantial oscillator strength) in the AEXB complexes involve molecular orbitals localized mostly on the XB donor (see Fig. S8 in the ESI). Accordingly, the energy of the absorption bands is essentially independent on the halide. Still, due to the molecular orbital interactions between the halides and 1, the small segments of the HOMOs of the complexes are localized on the halides, which affected the intensity of the transitions.‡‡ In contrast, in the XB complexes with the neutral halogenated electrophiles, the energies of the HOMOs and LUMOs of the halides are higher than the energies of the corresponding orbitals of the XB donors. As such, the HOMO of such complexes (as well as the other common molecular complexes) is localized mostly on the XB acceptors (electron donor), and the LUMO on the XB donor (electron acceptor). Accordingly, their lowest energy absorption bands represent in essence charge-transfer transition, and its energy vary with the energies of the HOMO of halides (the TD DFT calculations suggest that similar charge-transfer transitions in complexes of halides with 1 occur at higher energies, and they are overshadowed by the absorption of components).In summary, combined experimental (UV-Vis spectral) and computational studies of the interaction between halides and 1 demonstrated spontaneous formation of the anion–anion XB complexes in moderately-polar and polar solvents (which attenuate the electrostatic anion–anion repulsion and facilitate close approach of the interacting species§§). To the best of our knowledge, this constitutes the first experimental observation of AEXBs in solution. Stabilities of such “anti-electrostatic” associations are comparable to that formed by halide anions with the common neutral bromo- and iodo-substituted aliphatic or aromatic XB donors. These findings confirm that halogen bonding between our anionic XB donor 1 and halides is sufficiently strong to overcome electrostatic repulsion between two anions. It also supports earlier conclusions29 that besides electrostatics, molecular-orbital (weakly-covalent interaction) play an important role in the formation of XB complexes. Since the HOMO of 1 is higher in energy than those of the halides, the lowest-energy absorption bands in the anion–anion complexes is related mostly to the transition between the XB-donor localized MOs (in contrast to the charge transfer transition in the analogous complexes with neutral XB donors). Therefore, the energies of these transitions are similar in all complexes and the interaction with halides only slightly increase their intensities.  相似文献   

3.
Correction for ‘Graph neural network based coarse-grained mapping prediction’ by Zhiheng Li et al., Chem. Sci., 2020, 11, 9524–9531, DOI: 10.1039/D0SC02458A.

The authors regret that eqn (5) was missing the adjacency matrix term. The correct form of eqn (5) is shown below:5where σ is the bandwidth and is set to σ = 1 in the experiment. denotes the adjacency matrix (ij = 1 if atom i and atom j are bonded, otherwise ij = 0).The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

4.
Described here is a modular strategy for the rapid synthesis of β-functionalized electron-rich naphthalenes, a family of valuable molecules lacking general access previously. Our approach employs an intermolecular benzannulation of in situ generated isobenzopyrylium ions with various electron-rich alkynes, which were not well utilized for this type of reaction before. These reactions not only feature a broad scope, complete regioselectivity, and mild conditions, but also exhibit unusual product divergence depending on the substrate substitution pattern. This divergence allows further expansion of the product diversity. Control experiments provided preliminary insights into the reaction mechanism.

A substituent-controlled divergent benzannulation provides rapid access to various β-functionalized naphthalenes from electron-rich alkynes.

Functionalized naphthalenes are important structural motifs widely present in bioactive natural products, pharmaceuticals and functional materials.1,2 They also serve as valuable intermediates in organic synthesis.3 Among them, naphthalenes bearing an electron-donating group (EDG) at the β-position (e.g., β-naphthols, β-naphthylamines, and β-naphthyl thioethers) are particularly versatile (Fig. 1).2,3 For example, BINOL, which is derived from β-naphthol, is an extensively utilized synthetic precursor toward a wide range of privileged chiral ligands and catalysts.3 While various approaches have been developed for the synthesis of naphthalenes, efficient and selective de novo strategies toward these β-functionalized ones still remain in high demand. Moreover, to the best of our knowledge, a general and modular approach for rapid access to all these electron-rich β-functionalized naphthalenes remains unknown.4Open in a separate windowFig. 1Useful β-naphthol, β-naphthylamine, and β-naphthyl thioether units.Isobenzopyrylium ions are readily accessible and versatile intermediates in organic synthesis (Scheme 1).5 They are known to participate in cycloaddition reactions with diverse carbon–carbon double bonds or triple bonds for the synthesis of naphthalenes.5,6 While extensive progress has been achieved in this topic, challenges still remain to be addressed. For example, the majority of these benzannulations with alkynes have to be executed at high temperature, except those intramolecular cases or with stoichiometric activators. Moreover, electron-rich alkynes have not been well explored as reaction partners for the benzannulations with isobenzopyrylium ions. Nevertheless, such reactions would provide expedient access to the valuable β-naphthol and β-napthylamine derivatives with diverse substitution patterns (Scheme 1). In this context, here we report our effort in achieving a general and modular strategy toward these electron-rich naphthalenes with high efficiency and regioselectivity under mild conditions. We also observed interesting selectivity divergence controlled by the substituent of the isobenzopyrylium substrates, giving rise to different types of naphthalene products (e.g., 2-hydroxy-1-naphthaldehydes, Scheme 1).Open in a separate windowScheme 1Benzannulation of isobenzopyryliums with electron-rich alkynes.We began our study with isochromene 1a as the isobenzopyrylium precursor. Siloxy alkyne 2a was first employed as the model electron-rich alkyne in view of the extraordinary versatility of this type of alkyne in various cyclization reactions, including benzannulation.7–9 Various Brønsted and Lewis acids were evaluated as catalysts for this reaction. Unfortunately, most of them did not show obvious catalytic activity toward the formation of a naphthalene product (Table 1). These reactions either had little conversion or resulted in a mixture of undesired products. Nevertheless, further screening led to the identification of AgOTf and BF3·OEt2 as capable catalysts (entries 5–6), with the latter being superior, leading to the formation of naphthalene 3a as the major product (75% yield, entry 6). Increasing the catalyst loading to 20 mol% further improved the product yield to 85% (with an isolated yield of 81%, entry 7). Further increasing the catalyst loading proved to be not beneficial (entry 8). Notably, the use of substrate 1a′ bearing an ethoxy leaving group also provided an equally good result. Other solvents, such as toluene and MeCN, were also evaluated, but all gave lower product yields. Finally, it is worth noting that the mild conditions used here are in sharp contrast to the typical high temperature required for the previously known catalytic intermolecular benzannulations involving isobenzopyryliums and alkynes.5dEvaluation of reaction conditions
EntryCatalystYielda (%)
1MeSO3H<5b
2HNTf2<5c
3Sc(OTf)3<5c
4TiCl4<5b
5AgOTf30c
6BF3·OEt275
7 BF 3 ·OEt 2 (20 mol%) 85 (81) d
8BF3·OEt2 (1.0 equiv.)83
9BF3·OEt2 (20 mol%), with 1a′(75)d
Open in a separate windowaReaction scale: 1a (0.05 mmol), 2a (0.06 mmol), catalyst (10 mol%), and DCM (0.5 mL). Yield is based on the analysis of the 1H NMR spectrum of the crude product using CH2Br2 as the internal standard.bUnreacted substrates account for the major remainder of the mass balance.cAn unidentifiable mixture accounts for the major remainder of the mass balance.dYield in parentheses is isolated yield.With the optimized conditions, the generality of this process was examined (Scheme 2). A range of isochromenes 1 and siloxy alkynes 2 participated in this benzannulation to form the desired silylated β-naphthols 3 with moderate to good efficiency. Notably, these products were all generated as a single isomer. This is particularly noteworthy when compared with the cases using phthalazines, which gave a mixture of regioisomers if the phthalazine was unsymmetrically substituted.8g The excellent regioselectivity observed in our reaction is likely attributed to the significant polarization of both cycloaddition partners. While alkyl-substituted siloxy alkynes reacted efficiently, unfortunately aryl-substituted ones led to low yield.Open in a separate windowScheme 2Reaction scope. Reaction scale: 1 (0.5 mmol), 2 (0.6 mmol), DCM (5 mL), 12 h, and isolated yield. a Run for 24 h.During the above scope study, we found that the reaction of 3-unsubstituted isochromene 1g and siloxy alkyne 2a did not form the desired product 3a. Instead, 2-hydroxy-1-naphthaldehyde 4a was formed as the major product (64% yield, eqn (1)). Careful analysis of this product structure indicated that the original leaving part in 1a (in the case of 3a) was not completely cleaved from the molecule. Instead, only the C–O bond cleavage took place, which led to a dangling aldehyde group. Another important observation is that the TIPS group was lost in the naphthalene product. It was proposed that the methoxide leaving group in 1a might help remove this silyl group and assist the above C–O bond cleavage. In contrast, in the formation of 3a, this methoxide group was a part of the whole leaving unit during rearomatization and thus unavailable for desilylation (vide infra).In fact, such 2-hydroxy-1-naphthaldehyde products, resembling the highly versatile salicylaldehyde family, are indeed a useful substructure of bioactive molecules, such as gossypol (Fig. 1).2b They can also serve as key intermediates toward functional materials, such as molecular sensors (e.g., AHN, Fig. 1).2e In view of these important applications, considerable efforts were next devoted to further improving the reaction efficiency (see the ESI for details). Finally, we found that the reaction yield of 4a could be improved to 85% with 2.0 equivalents of BF3·OEt2 and one equivalent of 2,4,6-collidine as the additive. We believe that the role of collidine is to help reversibly stabilize the isobenzopyrylium intermediate and prevent its decomposition during the reaction progress.9b1The above protocol for the synthesis of 2-hydroxy-1-naphthaldehydes is general for a range of 3-substituted isobenzopyryliums (Scheme 3). The reaction efficiency was not affected by electron-donating or electron-withdrawing groups. Various siloxy alkynes were also suitable reaction partners, including aryl- and alkyl-substituted ones. Again, these products were all formed as a single regioisomer. While the majority of these reactions were highly chemoselective toward the formation of aldehydes 4, it is worth noting that tBu- and TBS-substituted alkynes resulted in a mixture of 3 and 4, with the former being major. This is likely due to the substantial steric hindrance in close proximity to the silyl group, whose cleavage was obstructed and thus the driving force for the subsequent C–O bond cleavage was weakened, thereby altering the pathway to preferentially form 3 (vide infra). Finally, the structures of 4a and 4b were unambiguously confirmed by X-ray crystallography.Open in a separate windowScheme 3Scope for the synthesis of 2-hydroxy-1-naphthaldehydes 4. Reaction scale: 1 (0.5 mmol), 2 (0.6 mmol), DCM (5 mL), 12 h, and isolated yield. a Run for 16 h. b Run for 19 h.The divergent reaction patterns observed with siloxy alkynes prompted us to explore other electron-rich alkynes. Thioalkyne 5a was next employed for the reaction with 1a. Indeed, direct extension with BF3·OEt2 as the catalyst successfully resulted in regioselective formation of β-naphthyl thioethers 6a in about 70% yield (Table 2, entries 1 and 2). Further screening of other catalysts was performed to further improve the reaction efficiency (see the ESI for details). While Brønsted acids led to a significant drop in the yield, we were pleased to find that the Lewis acid AgNTf2 exhibited excellent catalytic activity, furnishing 6a in 89% yield (entry 6).Condition optimizationa
EntryCatalystYieldb (%)
1BF3·OEt270
2BF3·OEt2 (20 mol%)72 (68)b
3HNTf220
4MeSO3H7
5AgOTf80
6 AgNTf 2 89 (83) b
Open in a separate windowaReaction scale: 1a (0.05 mmol), 5a (0.06 mmol), catalyst (10 mol%), and DCM (0.5 mL). The yields are based on NMR analysis of the crude product using CH2Br2 as the internal standard.bIsolated yield.With the above conditions, we carried out a brief examination of the reaction scope (Scheme 4). Interestingly, a similar selectivity divergence was also observed with thioalkynes. Indeed, under identical conditions, the 3-butyl-substituted and 3-unsubstituted isobenzopyryliums all reacted efficiently, with the former leading to only 2-naphthyl thioethers 6, but the latter preferentially to 2-sulfenyl-1-naphthaldehydes 7. It is worth noting that only one regioisomer was observed in all these products.Open in a separate windowScheme 4Scope for the synthesis of 2-naphthyl thioethers. Reaction scale: 1 (0.3 mmol), 5 (0.36 mmol), 12 h, and isolated yield. a Run for 37 h. b Run for 19 h. The unreacted alkyne accounted for the major mass balance.Finally, we were curious about the reactivity of ynamides in such benzannulations.10 To our delight, in the presence of 20 mol% of BF3·OEt2, the reaction between isochromene 1a and different ynamides 8 successfully afforded the desired 2-naphthylamine products 9 with excellent regioselectivity and good efficiency (Scheme 5). Different electron-withdrawing groups on the ynamides were all compatible with this protocol. However, to our surprise, the use of 3-unsubstituted isochromene 1g did not lead to the expected naphthaldehyde 10 at all. Instead, the same product 9a was obtained as the only product, even if two equivalents of BF3·OEt2 were used together with collidine as the additive (conditions in Scheme 3). This result is in dramatic contrast to the cases of siloxy alkynes and thioalkynes.Open in a separate windowScheme 5Scope of ynamides. Reaction scale: 1 (0.5 mmol), 8 (0.6 mmol), DCM (5 mL), 12 h, and isolated yield. a Run for 30 h. b Yield in parentheses is obtained with BF3·OEt2 (2.0 equiv.), 2,4,6-collidine (1.0 equiv.), 0 °C, and DCM.Next, further studies were performed to help understand the unusual selectivity divergence observed with siloxy alkynes and thioalkynes. It is obvious that the substituent at the 3-position of the isochromene substrates has a crucial impact on the product distribution. To further probe the possible steric and electronic effects, we incorporated various other substituents, such as different aryl groups and the bulky tBu group (Table 3). We found that these reactions afforded products 3a and 4′ with variable ratios. With the tBu group, almost the same product distribution as the nBu group was observed, exclusively forming 3a (Table 3, entry 2). However, aryl substitution reversed this ratio, giving product 4′ as the major product (entries 3–5). More importantly, the preference toward 4′ is more pronounced with electron-deficient aryl groups. p-Fluorophenyl substitution led to almost only 4′. These results suggested that the product distribution appeared to be more related to the electronic effect than the steric effect.Influence of the 3-substituents in isochromenesa
Open in a separate windowaReaction scale: 1 (0.05 mmol), 2a (0.06 mmol), and 12 h. Yield is based on the analysis of the NMR yield of the crude mixture using CH2Br2 as an internal standard.bRun with 10 mol% of BF3·OEt2.Possible mechanisms are depicted in Scheme 6 using a siloxy alkyne as an example. The reaction begins with Lewis acid activation on the acetal motif to form the isobenzopyrylium intermediate I. The subsequent cycloaddition with the alkyne triple bond forms the key bicyclic oxonium intermediate II, which has a resonance form II′. This step might also be stepwise by forming one C–C bond first via ketenium III. In either case, the regioselectivity is precisely controlled by matching the polarity of the two partners, which explains the exclusive formation of only one regioisomer in all the cases. Depending on the R substituent, oxonium II can proceed via three possible pathways. In path a (R = alkyl), the methoxide attacks the oxonium carbon to form intermediate IV, which then undergoes retro-[4 + 2] cycloaddition with concomitant rearomatization to form naphthalene 3, together with the release of ester RCO2Me. However, if the oxonium carbon is unsubstituted (R = H), this intermediate is relatively unstable due to less stabilization of the positive charge (also viewed as a secondary cation in II′, versus tertiary carbocation if R ≠ H). Thus, the silyl enol ether motif tends to push the electron toward this unstable oxonium (a “push–pull” scenario) to cleave the bridging C–O bond. This step, likely assisted by the methoxide attack on the silyl unit, leads to 1,3-dicarbonyl intermediate V. The subsequent tautomerization/aromatization leads to the observed 2-hydroxy-1-naphthaldehyde 4.6b As an exception, if the siloxy alkyne is bulky (e.g., R′ = tBu or TBS), the reaction preferentially gives product 3 (Scheme 3, 4l–m). We believe that the increased steric repulsion by bulky R′ significantly disfavors the approach of methoxide to the neighboring silyl group in IIb. Instead, the methoxide prefers to add to the oxonium carbon, which allows path a to operate and forms 3 as the major product. Another exceptional case is the use of ynamides, which exclusively give products 3 even if R = H. Although actual rationalization would require more sophisticated mechanistic study, we currently reason that the enamide unit in the IIb analogue might not be electron-rich enough to exert sufficient driving force to cleave the C–O bond. This can also be viewed from the standpoint of the limited stabilization of the resulting iminium ion next to an electron-withdrawing group if this C–O bond is cleaved. Consequently, the methoxide prefers to attack the oxonium carbon to favor path a. Finally, with 3-aryl-substituted isochromenes (R = Ar), we believe that the oxonium intermediate IIc is well stabilized by aryl resonance. Thus, the effective delocalization of the positive charge makes this oxonium carbon less electrophilic for nucleophilic attack by methoxide. Instead, the methoxide serves as a base to deprotonate the bridgehead hydrogen, which triggers rearomatization and C–O cleavage to form product 4′. The more electron-deficient aryl group makes this bridgehead hydrogen more acidic, thus further favoring this pathway, which explains the trend in entries 3–5, Table 3.Open in a separate windowScheme 6Proposed mechanism.To further substantiate the above rationale, we carried out some control experiments. First of all, to confirm the fate of the leaving part in the isochromene substrates when β-naphthols 3 were formed, we used a large homobenzyl group in isochromene 1p (eqn (2)). After its reaction with 2a, we were able to isolate ester 11, which is consistent with path a of the proposed mechanism. Next, we also suspected that products 3 and 4 might interconvert to each other depending on the reaction conditions. To probe this possibility, a mixture of 3a and methyl formate was treated with BF3·OEt2 and 2,4,6-collidine, the standard conditions toward products 4 (eqn (3)). However, product 4a was not observed at all, indicating that 3 is not an intermediate toward 4. Similarly, subjecting 4a and MeOH (or MeONa) to BF3·OEt2 did not lead to the deformylation product β-naphthol 3a′, suggesting that 4 is unlikely an intermediate in the formation of 3 (eqn (4)). These observations are consistent with the proposed mechanism, in which the product distribution is kinetically controlled by the barrier in each case and these paths are likely irreversible, but not thermodynamically controlled by product stability.234In summary, we have developed a modular strategy for the rapid synthesis of valuable β-functionalized electron-rich naphthalenes, specifically, β-naphthol, β-naphthylamine, and β-naphthyl thioether derivatives. It also represents the first systematic study of the benzannulations of versatile isobenzopyryliums with general electron-rich alkynes. With suitable choice of the isochromene substrates and the Lewis acid catalysts, different types of electron-rich alkynes, such as siloxy alkynes, ynamides, and thioalkynes, participated in the intermolecular cycloaddition reactions under mild conditions with high efficiency and complete regioselectivity. Moreover, depending on the substitution pattern of the isochromene substrates, unusual divergence toward different naphthalene products was observed, thus allowing further diversification of the naphthalene products. Control experiments provided preliminary insights into the intriguing mechanism. Further detailed investigations toward a better understanding are ongoing.  相似文献   

5.
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   

6.
Here, we report the nitric oxide monooxygenation (NOM) reactions of a CoIII-nitrosyl complex (1, {Co-NO}8) in the presence of mono-oxygen reactive species, i.e., a base (OH, tetrabutylammonium hydroxide (TBAOH) or NaOH/15-crown-5), an oxide (O2− or Na2O/15-crown-5) and water (H2O). The reaction of 1 with OH produces a CoII-nitrito complex {3, (CoII-NO2)} and hydrogen gas (H2), via the formation of a putative N-bound Co-nitrous acid intermediate (2, {Co-NOOH}+). The homolytic cleavage of the O–H bond of proposed [Co-NOOH]+ releases H2via a presumed CoIII-H intermediate. In another reaction, 1 generates CoII-NO2 when reacted with O2−via an expected CoI-nitro (4) intermediate. However, complex 1 is found to be unreactive towards H2O. Mechanistic investigations using 15N-labeled-15NO and 2H-labeled-NaO2H (NaOD) evidently revealed that the N-atom in CoII-NO2 and the H-atom in H2 gas are derived from the nitrosyl ligand and OH moiety, respectively.

Base-induced hydrogen (H2) gas evolution in the nitric oxide monoxygenation reaction.

As a radical species, nitric oxide (NO) has attracted great interest from the scientific community due to its major role in various physiological processes such as neurotransmission, vascular regulation, platelet disaggregation and immune responses to multiple infections.1 Nitric oxide synthase (NOS),2 and nitrite reductase (NiR)3 enzymes are involved in the biosynthesis of NO. NOSs produce NO by the oxidation of the guanidine nitrogen in l-arginine.4 However, in mammals and bacteria, NO2 is reduced to NO by NiRs in the presence of protons, i.e., NO2 + e + 2H+ → NO + H2O.5 Biological dysfunctions may cause overproduction of NO, and being radical it leads to the generation of reactive nitrogen species (RNS), i.e., peroxynitrite (PN, OONO)6 and nitrogen dioxide (˙NO2),7 upon reaction with reactive oxygen species (ROS) such as superoxide (O2˙),8 peroxide (H2O2),9 and dioxygen (O2).10 Hence, it is essential to maintain an optimal level of NO. In this regard, nitric oxide dioxygenases (NODs)11 are available in bio-systems to convert excess NO to biologically benign nitrate (NO3).12NO2 + FeII + H+ ↔ NO + FeIII + OH1[M–NO]n + 2OH → [M–NO2](n−2) + H2O2NOD enzymes generate NO3 from NO;11b,12−13 however, the formation of NO2 from NO is still under investigation. Clarkson and Bosolo reported NO2 formation in the reaction of CoIII-NO and O2.14 Nam and co-workers showed the generation of CoII-NO2 from CoIII-NO upon reaction with O2˙.15 Recently, Mondal and co-workers reported NO2 formation in the reaction of CoII-NO with O2.16 Apart from cobalt, the formation of CuII-NO2 was also observed in the reaction of CuI-NO and O2.17 For metal-dioxygen adducts, i.e., CrIII-O2˙ and MnIV-O22−, NOD reactions led to the generation of CrIII-NO2 (ref. 18) and MnV Created by potrace 1.16, written by Peter Selinger 2001-2019 O + NO2,19 respectively. However, the NOD reaction of FeIII-O2˙ and FeIII-O22− with NO and NO+, respectively, generated FeIII-NO3via FeIV Created by potrace 1.16, written by Peter Selinger 2001-2019 O and ˙NO2.20 Ford suggested that the reaction of ferric-heme nitrosyl with hydroxide leads to the formation of NO2 and H+.12 Lehnert and co-workers reported heme-based Fe-nitrosyl complexes21 showing different chemistries due to the FeII-NO+ type electronic structures. On the other hand, Bryan proposed that the one-electron reduction of NO2 to NO in ferrous heme protein is reversible (eqn (1)).22 Also, it is proposed that excess NO in biological systems is converted to NO2 and produces one equivalent of H+ upon reaction with ˙OH.23 Previously reported reactivity of M–NOs of Fe24 with OH suggested the formation of NO2 and one equivalent of H+, where H+ further reacts with one equivalent of OH and produces H2O (eqn (2)).25Here in this report, we explore the mechanistic aspects of nitric oxide monooxygenation (NOM) reactions of the CoIII-nitrosyl complex, [(12TMC)CoIII(NO)]2+/{Co(NO)}8 (1),15,26 bearing the 12TMC ligand (12TMC = 1,4,7,10-tetramethyl-1,4,7,10-tetraazacyclododecane) with mono-oxygen reactive species (O2−, OH and H2O) (Scheme 1). Complex 1 reacts with the base (OH, tetrabutylammonium hydroxide (TBAOH)/or NaOH in the presence of 15-crown-5 as the OH source) and generates the corresponding CoII-nitrito complex, [(12TMC)CoII(NO2)]+ (3), with the evolution of hydrogen gas (H2) via the formation of a plausible N-bound Co-nitrous acid intermediate ([Co-NOOH]+, 2) in CH3CN at 273 K (Scheme 1, reaction (I)). Also, when 1 reacts with the oxide (O2− or Na2O in the presence of 15-crown-5), it generates the CoII-nitrito complex (3) via a probable CoI-nitro, [(12TMC)CoI(NO2)] (4), intermediate (Scheme 1, reaction (II)); however, 1 does not react with water (Scheme 1, reaction (III)). Mechanistic investigations using 15N-labeled-15NO, D-labeled-NaOD and 18O-labelled-18OH demonstrated, unambiguously, that the N and O-atoms in the NO2 ligand of 3 resulted from NO and OH moieties; however, the H-atoms of H2 are derived from OH. To the extent of our knowledge, the present work reports the very first systematic study of CoIII-nitrosyl complex reactions with H2O, OH and O2−. This new finding presents an alternative route for NO2 generation in biosystems, and also illustrates a new pathway of H2 evolution, in addition to the reported literature.12,27Open in a separate windowScheme 1Nitric oxide monooxygenation (NOM) reactions of cobalt-nitrosyl complex (1) in the presence of a base (OH), sodium oxide (Na2O) and water (H2O).To further explore the chemistry of [(12TMC)CoIII(NO)]2+ (1),15,26 and the mechanistic insights of NOM reactions, we have reacted it with a base (OH), an oxide (O2−), and water (H2O). When complex 1 was reacted with TBAOH in CH3CN, the color of complex 1 changed to light pink from dark pink. In this reaction, the characteristic absorption band of 1 (370 nm) disappears within 2 minutes (Fig. 1a; ESI, Experimental section (ES) and Fig. S1a), producing a CoII-nitrito complex, [(12TMC)CoII(NO2)]+ (3), with H2 (Scheme 1, reaction (Ib)), in contrast to the previous reports on base induced NOM reactions (eqn (2)).12,25,28 The spectral titration data confirmed that the ratio-metric equivalent of OH to 1 was 1 : 1 (ESI, Fig. S1b). 3 was determined to be [(12TMC)CoII(NO2)](BF4) based on various spectroscopic and structural characterization experiments (vide infra).15,26bOpen in a separate windowFig. 1(a) UV-vis spectral changes of 1 (0.50 mM, black line) upon addition of OH (1 equiv.) in CH3CN under Ar at 273 K. Black line (1) changed to red line (3) upon addition of OH. Inset: IR spectra of 3-14NO2 (blue line) and 3-15NO2 (red line) in KBr. (b) ESI-MS spectra of 3. The peak at 333.2 is assigned to [(12TMC)CoII(NO2)]+ (calcd m/z 333.1). Inset: isotopic distribution pattern for 3-14NO2 (red line) and 3-15NO2 (blue line).The FT-IR spectrum of 3 showed a characteristic peak for nitrite stretching at 1271 cm−1 (CoII-14NO2) and shifted to 1245 cm−1 (CoII-15NO2) when 3 was prepared by reacting 15N-labeled NO (CoIII-15NO) with OH (Inset, Fig. 1a and Fig. S2). The shifting of NO2 stretching (Δ = 30 cm−1) indicates that the N-atom in the NO2 ligand is derived from CoIII-15NO. The ESI-MS spectrum of 3 showed a prominent peak at m/z 333.2, [(12TMC)CoII(14NO2)]+ (calcd m/z 333.2), which shifted to 334.2, [(12TMC)CoII(15NO2)]+ (calcd m/z 334.2), when the reaction was performed with CoIII-15NO (Inset, Fig. 1b; ESI, Fig. S3a); indicating clearly that NO2 in 3 was derived from the NO moiety of 1. In addition, we have reacted 1 with Na18OH (ES and ESI), in order to follow the source of the second O-atom in 3-NO2. The ESI-MS spectrum of the reaction mixture, obtained by reacting 1 with Na18OH, showed a prominent peak at m/z 335.2, [(12TMC)CoII(18ONO)]+ (calcd m/z 335.2), (SI, Fig. S3b) indicating clearly that NO2 in 3 was derived from 18OH. The 1H NMR spectrum of 3 did not show any signal for aliphatic protons of the 12TMC ligand, suggesting a bivalent cobalt center (Fig. S4).26b Furthermore, we have determined the magnetic moment of 3, using Evans'' method, and it was found to be 4.62 BM, suggesting a high spin Co(ii) metal center with three unpaired electrons (ESI and ES).29 The exact conformation of 3 was provided by single-crystal X-ray crystallographic analysis (Fig. 2b, ESI, ES, Fig. S5, and Tables T1 and T2) and similar to that of previously reported CoII-NO2/MII-NO2.15,26b Also, we have quantified the amount of nitrite (90 ± 5%), formed in the above reaction, using the Griess reagent (ESI, ES, and Fig. S6).Open in a separate windowFig. 2Displacement ellipsoid plot (20% probability) of 3 at 100 K. Disordered C-atoms of the TMC ring, anion and H-atoms have been removed for clarity.As is known from the literature, a metal-nitrous acid intermediate may form either by the reaction of a metal-nitrosyl with a base27 or by the metal-nitrite reaction with an acid (nitrite reduction chemistry);26b however, the products of both the reactions are different. Here, for the first time, we have explored the reaction of CoIII-nitrosyl (1) with a base. In this reaction, it is clear that the formation of CoII-nitrito would be accomplished by the release of H2 gas via the generation of a transient N-bound [Co-(NOOH)]+ intermediate (Scheme 2, reaction (II)). The formation of CoII-NO2 (3) from the [Co-(NOOH)]+ intermediate is likely to proceed by either (i) homolytic cleavage of the O–H bond and release of H2via the proposed CoIII-H transient species (CoIII-H = CoII + 1/2H2)30 (Scheme 2, reaction (III)), as reported in previous literature where the reduced cobalt, in a number of different ligand environments, is a good H+ reduction catalyst and generates H2 gas via a CoIII-H intermediate31 or (ii) heterolytic cleavage of the O–H bond and the formation of CoI-NO2 + H+.27 In the present study, we observed the formation of 3 and H2via the plausible homolytic cleavage of the NOO–H moiety of 2 as shown in Scheme 2, in contrast to the previous reports on base-induced reactions on metal-nitrosyls (eqn (3)).27 Taking together both possibilities, (i) is the most reasonable pathway for the NOM reaction of complex 1 in the presence of a base (as shown in Scheme 2, reaction (III)). And the reaction is believed to go through a CoIII-H intermediate as reported previously in CoI-induced H+ reduction in different ligand frameworks and based on literature precedence, we believe that complex 1 acts in a similar manner.31Open in a separate windowScheme 2NOM reaction of complex 1 in the presence of OH, showing the generation of CoII-nitrito (3) and H2via a Co(iii)-hydrido intermediate.In contrast to an O-bound CoII-ONOH intermediate, where N–O bond homolysis of the ON-OH moiety generates H2O2 (Scheme 2, reaction (IV)),26b the N-bound [Co-(NOOH)]+ intermediate decomposes to form NO2 and a Co(iii)-H transient species, arising from β-hydrogen transfer from the NOO–H moiety to the cobalt-center (Scheme 2, reaction (II)).30a,c,32 The Co(iii)-hydrido species may generate H2 gas either (a) by its transformation to the Co(ii)-nitrito complex (2) and H2 gas as observed in the case of CoIII-H intermediate chemistry30a,c,e−g as proposed in the chemistry of the CoI complex with H+ reduction31 and other metal-hydrido intermediates32 and also explained in O2 formation in PN chemistry17,33 or (b) by the reacting with another [Co-(NOOH)]+ intermediate (Scheme 2, reaction (III)).Furthermore, we have confirmed the H2 formation in the NOM reaction of 1 with OH by headspace gas mass spectrometry (Fig. 3a). Also, carrying out the reaction of 1 with NaOD leads to the formation of the [Co-(NOOD)]+ intermediate, which then transforms to a CoIII-D transient species. Further, as described above, the CoIII-D species releases D2 gas, detected by headspace gas mass spectrometry (Fig. 3b), which evidently established that H2 gas formed in the reaction of 1 with OH. In this regard, we have proposed that in the first step of this reaction, the nucleophilic addition of OH to {Co-NO}8 generates a transient N-bound [Co-(NOOH)]+ intermediate that is generated by an internal electron transfer to CoIII (Scheme 2, reaction (I)). By following the mechanism proposed in the case of CoIII-H,30a−c O2,15 and H2O2(ref. 26b) formation, we have proposed the sequences of the NOM reaction of 1, which leads to the generation of CoII-nitrito and H2 (Scheme 2, reaction (I)–(III) and Scheme 3). In the second step, O–H bond homolytic cleavage generates a CoIII-H transient species + NO2via a β-hydrogen elimination reaction of the [Co-(NOOH)]+ intermediate.32 The CoIII-H intermediate may undergo the following reactions to generate H2 gas and CoII-nitrito either (a) by the natural decomposition of the CoIII-H transient species to generate H2,30a,c,e−g or (b) by the H-atom abstraction from another [Co-(NOOH)]+ intermediate (Scheme 3). Also, to validate our assumption that the reaction goes through a plausible N-bound [Co-(NOOH)]+ intermediate followed by its transformation to the CoIII-H species (vide supra), we have performed the reaction of 1 with NaOH/NaOD (in 1 : 1 ratio). In this reaction, we have observed the formation of a mixture of H2, D2, and HD gases, which indicates clearly that the reaction goes through the formation of CoIII-H and CoIII-D transient species via the aforementioned mechanism (Fig. 3c). This is the only example where tracking of the H atoms has confirmed the H2 generation from an N-bound NOO–H moiety as proposed for H2 formation from CoIII-H.30Open in a separate windowFig. 3Mass spectra of formation of (a) H2 in the reaction of 1 (5.0 mM) with NaOH (5.0 mM), (b) D2 in the reaction of 1 (5.0 mM) with NaOD (5.0 mM), (c) D2, HD, and H2 in the reaction of 1 (5.0 mM) with NaOD/NaOH (1 : 1), and (d) H2 in the reaction of 1 (5.0 mM) with NaOH in the presence of 2,4 DTBP (50 mM).Open in a separate windowScheme 3NOM reaction of complex 1 in the presence of OH, showing the different steps of the reaction.While, we do not have direct spectral evidence to support the formation of the transient N-bound [Co-(NOOH)]+ intermediate and its decomposition to the CoIII-H transient species via β-hydrogen transfer from the NOOH moiety to the cobalt center, support for its formation comes from our finding that the reactive hydrogen species can be trapped by using 2,4-di-tert-butyl-phenol (2,4-DTBP).34 In this reaction, we observed the formation of 2,4-DTBP-dimer (2,4-DTBP-D, ∼67%) as a single product (ESI, ES, and Fig. S7). This result can readily be explained by the H-atom abstraction reaction of 2,4-DTBP either by [Co-(NOOH)]+ or CoIII-H, hence generating a phenoxyl-radical and 3 with H2 (Fig. 3d and Scheme 2, reaction (a)). Also, we have detected H2 gas formation in this reaction (ESI, ES, and Fig. 3d). In the next step, two phenoxyl radicals dimerized to give 2,4-DTBP-dimer (Scheme 2c, reaction (II)). Thus, the observation of 2,4-DTBP-dimer in good yield supports the proposed reaction mechanism (Scheme 2, reaction (a) and (b)). Further, the formation of 2,4 DTBP as a single product also rules out the formation of the hydroxyl radical as observed in the case of an O-bound nitrous acid intermediate.26bFurthermore, we have explored the NOM reactivity of 1 with Na2O/15-crown-5 (as the O2− source) and observed the formation of the CoII-nitrito complex (3) via a plausible CoI-nitro (4) intermediate (Scheme 1, reaction (IIa); also see the ESI and ES); however, 1 was found to be inert towards H2O (Scheme 1, reaction (III); also see the ESI, ES and Fig. S8). The product obtained in the reaction of 1 with O2− was characterized by various spectroscopic measurements.15,26b The UV-vis absorption band of 1 (λmax = 370 nm) disappears upon the addition of 1 equiv. of Na2O and a new band (λmax = 535 nm) forms, which corresponds to 3 (ESI, Fig. S9). The FT-IR spectrum of the isolated product of the above reaction shows a characteristic peak for CoII-bound nitrite at 1271 cm−1, which shifts to 1245 cm−1 when exchanged with 15N-labeled-NO (15N16O) (ESI, ES, and Fig. S10), clearly indicating the generation of nitrite from the NO ligand of complex 1.26b The ESI-MS spectrum recorded for the isolated product (vide supra) shows a prominent ion peak at m/z 333.1, and its mass and isotope distribution pattern matches with [(12-TMC)CoII(NO2)]+ (calc. m/z 333.1) (ESI, Fig. S11). Also, we quantified the amount of 3 (85 ± 5%) by quantifying the amount of nitrite (85 ± 5%) using the Griess reagent test (ESI, ES, and Fig. S6).In summary, we have demonstrated the reaction of CoIII-nitrosyl, [(12-TMC)CoIII(NO)]2+/{CoNO}8 (1), with mono-oxygen reactive species (O2−, OH and H2O) (Scheme 1). For the first time, we have established the clear formation of a CoII-nitrito complex, [(12TMC)CoII(NO2)]+ (3), and H2 in the reaction of 1 with one equivalent of OHvia a transient N-bound [Co-(NOOH)]+ (2) intermediate. This [Co-(NOOH)]+ intermediate undergoes the O–H bond homolytic cleavage and generates a CoIII-H transient species with NO2, via a β-hydrogen elimination reaction of the [Co-(NOOH)]+ intermediate, which upon decomposition produces H2 gas. This is in contrast to our previous report, where acid-induced nitrite reduction of 3 generated 1 and H2O2via an O-bound CoII-ONOH intermediate.26b Complex 1 was found to be inert towards H2O; however, we have observed the formation of 3 when reacted with O2−. It is important to note that H2 formation involves a distinctive pathway of O–H bond homolytic cleavage in the [Co-(NOOH)]+ intermediate, followed by the generation of the proposed CoIII-H transient species (CoII + 1/2H2)30 prior to H2 evolution as described in CoI chemistry with H+ in many different ligand frameworks.31 The present study is the first-ever report where the base induced NOM reaction of CoIII-nitrosyl (1) leads to CoII-nitrito (3) with H2 evolution via an N-bound [Co-(NOOH)]+ intermediate, in contrast to the chemistry of O-bound CoII-ONOH26b, hence adding an entirely new mechanistic insight of base induced H2 gas evolution and an additional pathway for NOM reactions.  相似文献   

7.
Radical electrons tend to localize on individual molecules, resulting in an insulating (Mott–Hubbard) bandgap in the solid state. Herein, we report the crystal structure and intrinsic electronic properties of the first single crystal of a π-radical metal, tetrathiafulvalene-extended dicarboxylate (TED). The electrical conductivity is up to 30 000 S cm−1 at 2 K and 2300 S cm−1 at room temperature. Temperature dependence of resistivity obeys a T3 power-law above T > 100 K, indicating a new type of metal. X-ray crystallographic analysis clarifies the planar TED molecule, with a symmetric intramolecular hydrogen bond, is stacked along longitudinal (the a-axis) and transverse (the b-axis) directions. The π-orbitals are distributed to avoid strong local interactions. First-principles electronic calculations reveal the origin of the metallization giving rise to a wide bandwidth exceeding 1 eV near the Fermi level. TED demonstrates the effect of two-dimensional stacking of π-orbitals on electron delocalization, where a high carrier mobility of 31.6 cm2 V−1 s−1 (113 K) is achieved.

The molecular arrangement that enables metallic conduction in a single-component pure organic crystal is revealed by single-crystal X-ray diffraction.

Organic molecular solids are typically insulating due to their paired electrons in spatially localized s- and p-orbitals. The concept of charge-transfer (CT) between donor and acceptor1 enabled the development of conducting molecular complexes (salts) including semiconducting perylene-bromine,2 metallic tetrathiafulvalene (TTF)-tetracyano-p-quinodimethane (TCNQ),3 and polyacethylene doped with halogen molecules.4 A different strategy was proposed in the 1970s based on organic radicals with an open-shell electronic structure.5 π-Radicals such as neutral-,6 fully-conjugated7 and zwitterionic (betainic)8 molecules, with an unpaired electron in their singly occupied molecular orbital (SOMO), offered potential candidates. However, all these π-radicals were insulators or semiconductors with a finite bandgap, which is due to the SOMO being localized on an individual molecule. In the Mott–Hubbard model,9 the case of the π-radical solids can be described by the on-site Coulomb repulsion U being larger than the electronic bandwidth W (U/W > 1), in contrast to U/W < 1 in molecular metals like CT metal systems (Fig. 1a).Open in a separate windowFig. 1(a) Schematic representation of the electronic band structure of Mott–Hubbard insulators (left) and possible molecular metals (right), respectively. Solids formed from typical π-radicals possess large on-site Coulomb repulsion U compared with the electronic bandwidth W, resulting a finite bandgap (left). This requires a new mechanism to expand W and overcome U to achieve a metallic state at ambient pressure in π-radical crystals. (b) Molecular structure of the zwitterionic radical, tetrathiafulvalene-extended dicarboxylate (TED) with a symmetric intramolecular hydrogen bond.A straightforward approach to realize high conductivity in π-radical systems is to enhance the intermolecular interaction by applying high pressure.10 Bisdithiazolyl radical crystal achieved W ∼ 1 eV near the Fermi level and the room temperature conductivity σRT = 2 S cm−1 under 11 GPa pressure.10c An alternative route is to decrease the interatomic spacing by incorporating a metal ion. Introduction of a semimetal Se and intermolecular hydrogen bonding in a donor-type radical succeeded to improve a conductivity to σRT = 19 S cm−1 but still required high pressure over 1 GPa for breaking its insulating character.10d An organometallic compound with a transition metal, [Ni(tmdt)2], by contrast, is known to form a three-dimensional (3D) Fermi surface with W = 0.48 eV and metallic conduction with σRT = 400 S cm−1 at ambient pressure.11 A breakthrough concept for expanding W at ambient pressure is desired for achieving metallization in pure organic π-radicals.Tetrathiafulvalene-extended dicarboxylate (TED) is an organic air-stable zwitterionic radical (Fig. 1b),12 which was designed based on carrier generation induced by a stably-introduced protonic defect (–H+) in hydrogen-bonding molecules without adopting CT between multiple molecules.13 A polycrystalline film of TED exhibited metallic conduction at ambient pressure, but the mechanism has not been clarified yet due to lack of single crystal information.12 Herein, we report the first crystal structure and intrinsic electronic properties of the recently grown single crystal TED. Structural analysis and quantum chemical simulations based on the single crystal reveal the origin of its metallic behavior.  相似文献   

8.
9.
α-Amylases are among the most widely used classes of enzymes in industry and considerable effort has gone into optimising their activities. Efforts to find better amylase mutants, such as through high-throughput screening, would be greatly aided by access to precise and robust active site titrating agents for quantitation of active mutants in crude cell lysates. While active site titration reagents designed for retaining β-glycosidases quantify these enzymes down to nanomolar levels, convenient titrants for α-glycosidases are not available. We designed such a reagent by incorporating a highly reactive fluorogenic leaving group onto unsaturated cyclitol ethers, which have been recently shown to act as slow substrates for retaining glycosidases that operate via a covalent ‘glycosyl’-enzyme intermediate. By appending this warhead onto the appropriate oligosaccharide, we developed efficient active site titration reagents for α-amylases that effect quantitation down to low nanomolar levels.

α-Amylases are among the most widely used classes of enzymes in industry and considerable effort has gone into optimising their activities.

Amylases are among the most common classes of enzymes employed in industrial settings, being used in detergents, bread, beer, biofuel, and many other sectors. Accordingly, α-amylases account for 25% of the world''s multi-billion dollar enzyme market.1,2 α-Amylases are endo-acting enzymes that cleave starch into malto-oligosaccharides, which are further degraded by exo-acting α-glucosidases, glucoamylases, β-amylases and α-glucan phosphorylases and lyases. They are found in CAZy GH families 13, 57, 119 and 126, with the vast majority in the large GH13 family.3 GH13 enzymes adopt a (β/α)8 fold with three highly conserved active site carboxylic acids.4–6 They employ a classical double-displacement mechanism7 in which one of the glutamic acids provides acid catalytic assistance to the leaving group departure while an aspartate attacks the anomeric centre, forming a covalent glycosyl enzyme intermediate. In a second step, water attacks the anomeric centre with base assistance from the glutamate residue (Fig. 1A and B).Open in a separate windowFig. 1Koshland mechanism of retaining β- and α-glycosidases (A & B). The same mechanism has been observed for the hydrolysis of “β”-valienols (C), and for “α”-valienols (D).Given their industrial importance, a huge amount of attention has been given to the discovery and improvement of α-amylases to attain optimal performance for particular applications. These approaches typically require high-throughput analysis of large numbers of gene products or mutants thereof.8–10 Identification of the best candidates then ideally requires high-throughput assay coupled with a method for determining the enzyme concentration in each sample. This can be a challenging task in the absence of purification, as would be the case for truly high-throughput approaches. The “gold standard” method to quantify active enzyme concentration is active site titration.11 Active site titrants react stoichiometrically with their target enzymes and release one equivalent of a quantifiable agent, which is typically either a chromophore or fluorophore. For enzymes that operate via a covalent intermediate, such as retaining glycosidases, the active site titrants are usually chromogenic or fluorogenic substrates that form this intermediate with a rate constant (kon) that is much greater than that for its hydrolysis (koff) – ideally with koff approaching zero.Our lab has previously developed active site titration reagents for several retaining β-glycosidases12,13 and neuraminidases.14,15 By replacing the substituent on the position adjacent to the anomeric centre of the sugar (the hydroxyl at C-2 for many monosaccharides) with a fluorine atom, both the formation and the hydrolysis of the glycosyl-enzyme intermediate are slowed, largely through inductive destabilisation of the transition state. Further incorporation of a reactive fluorogenic leaving group generates a reagent that, upon covalently inactivating the glycosidase, releases a stoichiometric and quantifiable amount of fluorophore. The fluorogenic response is then measured to determine the amount of active glycosidase that is present in solution.Unfortunately, this same strategy does not work for retaining α-glycosidases. In those cases, koff remains greater than kon, likely due to the inherently greater reactivity of the β-glycosyl-enzyme intermediate,16,17 and the compounds are simply substrates with low turnover numbers. By use of 2,2-dihalosugars with yet more reactive leaving groups, this problem could be solved in some cases, but their synthesis is challenging, and inactivation rates were low, or non-existent in some cases.18,19 Alternative approaches were called for.Recently, a new class of glycosidase substrates was reported in which the sugar moiety is replaced by an equivalently hydroxylated cyclohexene.20–23 Hydrolysis of these enol ethers likely occurs via an allylic cation of almost identical reactivity to that of the equivalent oxocarbenium ion. Glycosidases cleave these substrates via the classical Koshland mechanism7 (Fig. 1C and D), but considerably more slowly than their natural substrates. However, incorporation of a good leaving group will accelerate, relatively, the first step such that, in some cases, they act as mechanism-based inactivators making them candidates for development of an active site titrant for α-amylases.Since α-amylases are endo-acting enzymes that do not usually cleave monosaccharide glycosides, an ‘extended” oligosaccharide version containing a total of 2 or 3 sugar/pseudosugar moieties would be needed. Substrates longer than this would be prone to internal glycoside cleavage. Since 2-chloro-4-nitrophenyl maltotrioside (CNP-G3) functions as a substrate for most amylases, we focused on addition of a maltosyl unit to a valienol moiety containing a 6,8-difluorocoumarin (F2MU) leaving group at its “anomeric centre”. The low pKa of this coumarin, 4.7,14 results in a greater reactivity of the reagent and also ensures the coumarin will be deprotonated and thus fluorescent, upon release at neutral pH.Synthesis of partially protected alcohol 2 from gluconolactone 1via literature methods24 was followed by attachment of F2MU via a Mitsunobu reaction and subsequent removal of the protecting groups under acidic conditions, generating known pseudo-glycoside 3.23 To check this concept before we synthesized the longer version, we tested compound 3 as a titrant of a simple α-glucosidase and found that it did indeed titrate the enzyme (Fig. S5). Since elongation of this pseudosugar via classical organic synthesis would require substantial protecting group chemistry, we elected instead to employ an enzymatic coupling strategy using the GH13 cyclodextrin transglycosidase, CGTase. This enzyme can use glycosyl fluorides, such as α-maltosyl fluoride, to effect glycosyl transfer onto suitable acceptors. However, a significant competing reaction would involve self-condensation of glycosyl fluorides ultimately forming cyclodextrins. To avoid this problem, we employed a maltosyl fluoride donor (4), in which the 4′-hydroxyl had been capped with a methyl group.25,26 Incorporation of 4′-methoxy groups does not alter the reaction with α-amylases, as this site in the normal substrate is occupied by additional sugar residues. Thus CGTase-catalysed glycosylation between known glycosyl fluoride 4 and pseudo-glycoside 3, gave the pseudo-trisaccharide 5 in 64% isolated yield (Scheme 1).Open in a separate windowScheme 1Synthesis of titration reagent 5.With this reagent in hand, we proceeded to screen its ability to inactivate a small panel of α-amylases. As shown in Fig. 2, time-dependent inactivation was observed for all enzymes tested, with the most industrially relevant enzymes, Effusibacillus pohliae amylase (EPA) and Aspergillus oryzae amylase (AOA), being inactivated the fastest.Open in a separate windowFig. 2Time-dependent inactivation of a small panel of amylases, showing remaining % activity versus time. Red box with X: AOA (91 nM); blue square: EPA (66.7 nM); purple cross: PPA (500 nM); green triangle: HPA (125 nM). AOA = A. oryzae amylase; EPA = E. pohliae amylase; HPA = human pancreatic amylase; PPA = porcine pancreatic amylase.Kinetic parameters for inactivation were then determined by directly monitoring the release of F2MU by UV-Vis (Table 1). To determine kon and koff (Scheme 2), we monitored chromophore (F2MU) release by absorbance at 370 or 380 nm (dependent on the concentration of 5 in the measurements of each enzyme). After mixing 5 with each enzyme individually, a burst phase followed by a steady-state phase was observed. For each enzyme, this was then repeated with varying concentrations of 5. Initial rates of F2MU release versus concentration of 5 were fit to a Michaelis–Menten equation to provide kon. The rate constant of cyclitol release, koff, was determined by measuring rates of the steady-state region at a saturating concentration (5× Ki). We found that several amylases: Effusibacillus pohliae amylase (EPA), Aspergillus oryzae amylase (AOA), Rhizomucor pusillus amylase (RPA) and porcine pancreatic amylase (PPA), inactivated quickly (highest kon, lowest koff, and greatest kon/Ki), and are therefore ideal candidates for titration with compound 5. Human pancreatic amylase (HPA), on the other hand, while inactivating rapidly, binds the reagent relatively poorly.Open in a separate windowScheme 2Kinetic parameters for the hydrolysis of 5 by several amylases (at 25 °C for EPA, AOA, and RPA and 30 °C for human pancreatic amylase [HPA] and porcine pancreatic amylase [PPA])
Enzyme: K i (μM) k on (min−1) k on/Ki (min−1 nM−1) k off (min−1)
HPA30400.200.07n.d.
PPA2280.482.10.024
EPA45.01.7028.90.002
AOA73.00.344.60.001
RPA1600.251.60.004
Open in a separate windowConfirmation that the inactivation observed was a result of stoichiometric covalent derivatisation of the enzyme was obtained for three representative enzymes by monitoring the enzyme molecular weights before and after inactivation, by electrospray ionisation mass spectrometry. As shown in Table 2, Fig. 3 and in ESI Fig. S1, the mass of each enzyme after the reaction was increased by the expected 496 mass units relative to that of the unlabelled enzyme. Complete reaction is evident by the absence of any unlabelled enzyme peak after addition of 5 (Fig. 3 and S1). It should also be noted, that all glycoforms of AOA underwent complete reaction (Fig. 3).Open in a separate windowFig. 3MS-plot of AOA before (left side, blue) and after (right side, green) addition of 5 showing the expected increase of 496 mass units.Intact-MS results confirming the stoichiometric addition of pseudo-trisaccharide (+496 m/z) to each amylase
Enzyme:Enzyme (M + H+)Enzyme + 5 (M + H+)
HPA56 066.756 563.0
EPA75 190.675 686.0
AOA53 863.354 358.6
Open in a separate windowHaving demonstrated that these reagents function as effective time-dependent inactivators, we then evaluated their utility as active site titration agents. Initially, we tested the ability of compound 5 to titrate ∼100 nM EPA (Fig. 4A) and observed a classical burst of fluorescence followed by a steady-state turnover phase. The active enzyme concentration can be quantified from such plots by extrapolating the steady-state portion back to the y-intercept (t = 0) and fitting the burst to eqn (1), as described previously.11Burst = [E]0 × (kon/kon + koff)21Open in a separate windowFig. 4(A) Titration of EPA with 100 μM of 5. The release of F2MU was monitored fluorimetrically (λEx = 353 nm, λEm = 451 nm). The steady-state signal was extrapolated back to the y-axis to get the corresponding burst amplitude and fit to eqn (1) to give enzyme concentration. (B) Titration of a set of serial dilutions of EPA. Red small box with X: 100 nM; yellow triangle: 50 nM; green (small) square: 25 nM; blue-green cross: 12.5 nM; blue (large) square: 6.3 nM; pink star: 3.1 nM. Burst responses show a detectable response down to 3 nM. (C) Plot of dilution factor versus the calculated [Novamyl], based on the burst response in Fig. 4B with the colours corresponding to the concentrations indicated in Fig. 4B.The values we obtained in this way are shown in Table 3 and compared with the concentrations claimed by the manufacturers: agreement was excellent for EPA and AOA. To explore a case in which total protein concentration was likely to be different from active concentration, we titrated a commercial porcine pancreatic amylase (PPA) that was purchased from Sigma-Aldrich several years prior to this set of experiments. The concentration of the active enzyme was found to be almost 4-fold lower than the original stated value. This discrepancy is likely due to the degradation of the enzyme over the long storage period. Indeed, such discrepancies in active enzyme concentration and total protein concentration are exactly what the titration agent is designed to detect.Concentrations of several amylases determined by titration with reagent 5, compared with manufacturers'' listed concentrationa
Enzyme:[Manufacturers] (nM)[Titration] (nM)
AOA9193
EPA100116
PPA610164
Yeast α-glucosidaseb88007500
Open in a separate windowaDetermined by declared activity measurements.bSubstrate 3 was used for titration of this enzyme. The manufacturer''s concentration was determined by measuring the enzyme activity with PNP-Glc (Vmax) and using their reported units of activity vs. PNP-Glc.Finally, the sensitivity and linearity of the response of our titration reagent 5 were tested using a crude, commercial batch of EPA (5 mg mL−1; 66.7 μM). After initial dilution of this batch of EPA to 100 nM, a series of 2× dilutions gave a reliable and detectable fluorogenic response down to enzyme concentrations as low as 3 nM (Fig. 4B and C).  相似文献   

10.
We report here porphodilactol derivatives and their corresponding metal complexes. These systems show promise as “all-in-one” phototheranostics and are predicated on a design strategy that involves controlling the relationship between intersystem crossing (ISC) and photothermal conversion efficiency following photoexcitation. The requisite balance was achieved by tuning the aromaticity of these porphyrinoid derivatives and forming complexes with one of two lanthanide cations, namely Gd3+ and Lu3+. The net result led to a metalloporphodilactol system, Gd-trans-2, with seemingly optimal ISC efficiency, photothermal conversion efficiency and fluorescence properties, as well as good chemical stability. Encapsulation of Gd-trans-2 within mesoporous silica nanoparticles (MSN) allowed its evaluation for tumour diagnosis and therapy. It was found to be effective as an “all-in-one” phototheranostic that allowed for NIR fluorescence/photoacoustic dual-modal imaging while providing an excellent combined PTT/PDT therapeutic efficacy in vitro and in vivo in 4T1-tumour-bearing mice.

We report here porphodilactol derivatives and their corresponding metal complexes as “all-in-one” phototheranostics by controlling the relationship between intersystem crossing (ISC) and photothermal conversion efficiency following photoexcitation.  相似文献   

11.
12.
Gold nanoparticles (AuNPs) have been prepared and surface-functionalized with a mixture of 1-hexanethiol co-ligands and chiral discogen ligands separated from a disulfide function via a flexible spacer. Polarized optical microscopy together with differential scanning calorimetry showed that the organic corona of the nanocomposite forms a stable chiral discotic nematic phase with a wide thermal range. Synchrotron X-ray diffraction showed that gold NPs form a superlattice with p2 plane symmetry. Analysis indicated that the organic corona takes up the shape of a flexible macrodisk. Synchrotron radiation-based circular dichroism signals of thin films are significantly enhanced on the isotropic-LC transition, in line with the formation of a chiral nematic phase of the organic corona. At lower temperatures the appearance of CD signals at longer wavelengths is associated with the chiral organisation of the NPs and is indicative of the formation of a second helical structure. The decreased volume required and the chiral environment of the disc ligands drives the nanoparticles into columns that arrange helically, parallel to the shortest axis of the two dimensional lattice.

Gold nanoparticles (NPs) were surface functionalized with hexanethiol groups and chiral nematic discogen ligands. A superlattice, liquid crystal behaviour and helix formation of the discs and a second helical organisation of the NPs were detected.

The exploration of chiral nanomaterials is a rapidly evolving fascinating field with uses ranging from catalysis, chiral molecular sensing, stereoselective separations to super-resolution imaging.1–3 Bottom-up approach design strategies for nanoparticles (NPs) or nanorods (NRs) as composites in liquid crystal (LC) matrices have been developed, making use of progress in synthesis strategies and characterization techniques.4–6 For optical applications symmetry breaking by introducing chirality into the systems at the nanoscale is critical.7 So far, efforts to obtain chiroptical NP-LC systems have relied mainly on NPs doped in LC hosts: either chiral nematic templates disperse NPs into chiral assemblies8 or the NPs functionalized with chiral ligands are included in achiral nematic LCs;9 both strategies aim to transfer and amplify chiral effects.10,11 Beyond the scientific challenge of generating nanocomposites where the interplay of bottom-up structuring enables plasmonic interactions of the NPs, leading to novel, so called metamaterials properties, there are new uses ranging from novel display devices,12 to more advanced optical communications using spin information between LEDs13 and in quantum teleportation for optical computing.14Precise size control of NPs in LC matrices in combined systems has been found critical to avoid phase separation or precipitation of NPs.15 For intrinsic liquid crystalline NP-LC systems, targeting of the LC phase range to be close to ambient is of utmost importance, as delamination of ligands occurs at elevated temperatures, even for very stable thiol functions. Degradation tends to occur above 100 °C.16 Inclusion of chiral groups in the organic corona can lead to chiral LC phase behavior, but there is little evidence that the NPs experience a chiral distortion in their packing; experimental evidence indicates that the nanocomposites minimize energy by packing in non-chiral superlattices and chirality is confined to the corona.17Here we show that a combination of small NPs with a carefully designed chiral organic coating results in chiral assembly behaviour of both the NPs and the organic corona. Our approach is strictly modular. The surface of the gold NPs is covered and passivated by hexylthiol groups and chiral lipoic acid derivates bearing via a spacer the mesogenic groups. We note that lipoic acid, a naturally occurring chiral antioxidant explored elsewhere for the treatment of several diseases, is bound with a dithiolate bridge via two atoms to the Au surface (see Chart 1).18 This places the chiral moiety rigidly in position onto the NPs, making thus a chiral surface.19 A pentaalkynylbenzene (PA) derivative was selected as mesogenic group, as this structural motif is known to promote reliably discotic nematic phase (ND) behaviour.20 The much larger volume of the PA group, when compared to typical rod shaped mesogens, whilst also delivering LC behaviour close to ambient, is anticipated to impact more strongly on the spatial organization of neighbouring NPs when compared to calamitic functionalized NP systems.17 The formation of helically ordered superstructures is studied by a combination of dedicated thin film optical polarizing microscopy, X-ray diffraction and synchrotron radiation-based circular dichroism (SRCD). A structural model is proposed based on a chiral nematic phase made up of AuNPs coated with chiral shells forming flexible macrodisks which in turn assemble in a superlattice.21–23Open in a separate windowChart 1Sketch of the functionalized gold nanoparticle system AuDLC*.The design, synthesis and chemical characterization of the chiral discogen is described in the ESI. The precursor disk-like mesogen which forms exclusively a ND phase above 100 °C consists of six aromatic rings flanked by flexible pentyloxy chains and an undecyloxy chain bearing a terminal hydroxy group (Fig. S1). The attachment of the flexible (R)-(+)-1,2-dithiolane-3-pentanoic tail was predicted to promote the desired phase. The new target discogen was obtained in a high yield (93.1%) reaction. For the synthesis and purification of LC-functionalized AuNPs, denoted as AuDLC*, a synthetic pathway was developed involving first the preparing and purifying of the alkylthiol coated NPs to which LC groups are linked, using an exchange reaction (for details see the Scheme 1 in ESI).24,25The thermal behaviour of the chiral discogen was examined by differential scanning calorimetry (DSC) and polarized optical microscopy (POM). The target discogen shows clearly a chiral nematic phase with the characteristic bundles of oily streaks (Fig. S2). On heating (Fig. S3), phase formation was observed at 49.0 °C from the crystalline state, with the material turning into an isotropic liquid at 95.7 °C (ΔH = 0.20 J g−1 or 0.28 kJ mol−1). Formation of the mesophase from the isotropic state on cooling started from 94.5 °C with an enthalpy change of 0.12 J g−1 (0.17 kJ mol−1), followed by a slow crystallisation occurring at 5.1 °C. In brief, the chiral phase behaviour of the target ligand is exclusively chiral nematic, enantiotropic and with a wide thermal range of up to 89.4 °C.The 1H-NMR spectra of the investigated AuDLC* nanocomposite and the comparison with the spectra of the monomer ligand (Fig. S4) indicate that the chiral discogen with the disulfide group are chemically attached onto the surface of AuNPs with high efficiency. A typical feature for such NPs are the broadened 1H-NMR spectra caused by the reduced mobility of alkyl chains of the ligands when anchored to NPs and additionally the sharp peaks present in the free ligand have disappeared, typical for such systems.26 The absence of mobile free discogen using thin layer chromatography (silica, with dichloromethane as mobile phase) confirms this result. The amount of chiral discogen bounded onto the gold surface relative to the number of 1-hexanethiol groups was determined from the 1H-NMR spectra of AuDLC*. The ratio of hydrocarbons to discogens was found to be 2 : 3. Transmission Electron Microscopy (TEM) investigations confirm that this sequential synthesis affords nanoparticles of low polydispersity (Fig. 1a); the average particle size is ∼2.5 ± 0.5 nm. Thermo-gravimetric analysis (TGA) was performed (Fig. S6) and gives the weight fraction of gold in AuDLC* to be 47.8% (wt/wt). The mesogen density per Au NP was calculated (ESI Part 7) by combining the 1H-NMR results (Fig. S4) with the model developed by Gelbart et al.27 On the basis of these data, we obtain that there are on average 211 Au atoms per particle, and a total number of 53 ligands per particle, of which 32 are chiral discogens and 21 are hexylthiol groups (Table S1). This is in accord with a ligand density of up to ∼6.3 ligands per nm2.28 Additionally, the characteristic peaks in the UV-vis spectra of AuDLC* arise from the combination of gold (∼500 nm) and LC ligands (absorption maxima at 238, 261 and 351 nm with two shoulder peaks at 378 and 421 nm) (Fig. S7).Open in a separate windowFig. 1(a) TEM image of 2.5 nm AuDLC* nanocomposite (inset: the size distribution of AuDLC*). POM micrographs of AuDLC* (b) 86.5 °C (90° crossed polarizer) (×100 μm). (c) 37.8 °C (90° crossed polarizer) (×100 μm). (d) DSC of AuDLC* at heating and cooling rate of 10.0 °C min−1 (e) SAXS diffractograms of AuDLC* on heating from 30 °C to 140 °C. (f) SRCD spectra of sheared AuDLC*. Recorded in 5 °C steps on cooling from 100 °C to 30 °C. The enlarged plots in dotted regions A–C are shown in Fig. S12b–d. The CD intensities at a selected wavelengths in each region (A, B and C) are plotted as a function of temperature in Fig. S13 and these plots clearly show the difference in the formation process between the two helical structures.The mesophase behaviour of the AuDLC* nanocomposite was first characterized by POM observation. As shown in Fig. 1b and c, upon cooling from the isotropic liquid, the colour in the POM texture changes from light green to yellow and we associate this with the helical pitch in AuDLC* increasing with decreasing temperature. In some regions of the POM slides, a chiral nematic phase with the typical Grandjean texture (Fig. S8a and b) is found. The LC textures in AuDLC* do recover after gently pressing the microscope coverslip indicating that the LC-decorated NPs form a stable phase and birefringent textures remain stable at ambient (Fig. S9). DSC investigations (Fig. 1d) confirm that the AuDLC* nanocomposite shows a thermodynamically stable (enantiotropic) mesophase. In DSC experiments on cooling a small peak associated with the isotropic to phase transition (92.2 °C) can be detected, and at a low temperature (3.8 °C), the glass transition occurs. The transition peak for the transition is wider than in the pure chiral discogen, typical for such systems.4 This is associated with the reduced mobility of the LC groups attached to the NPs. The thermal transitions determined by DSC for the pure chiral discogen and AuDLC* are collected in Table 1, showing the low values typical for transitions.20 In contrast to the free chiral ligand, the NPs coated with the chiral ligand and hexylthiol groups show a small decrease in transition temperatures. This could be attributed to both the plastifying effects of the hexylthiol co-ligands and packing constraints of the LC groups due to the attachment to the NPs. This view is supported by the DSC data for the transition at 0.043 J g−1 (0.12 kJ mol−1, considering the evaluated 32 chiral discogen ligands per particle).Transition temperatures (°C) of free chiral discogen and AuDLC*, as determined by DSC (second cooling at rate of 10.0 °C min−1)a
CompoundTransition temperature/°C (enthalpy)
MonomerIso94.5 (0.12 J g−1 or 0.17 kJ mol−1) 5.1Cr
AuDLC*Iso92.2 (0.043 J g−1 or 0.12 kJ mol−1) 3.8 T g
Open in a separate windowaIso = isotropic, Cr = crystalline, Tg = glass transition, = chiral discotic nematic phase.In synchrotron XRD studies of the AuDLC* nanocomposite, three peaks are obtained in the small angle regions (Fig. 1e) with values of q: 2.0160, 3.3289 and 4.9015 nm−1; equivalent to values of 3.12 nm, 1.89 nm and 1.28 nm (Table S2). The peaks could be indexed as (10), (01) and (11̄) reflections of a columnar lattice with a p2 plane group (Fig. 2a). The unit cell parameters are a = 4.17 nm, b = 2.52 nm and γ ∼ 48.4°. When comparing the estimated average volume, the area of the lattice is 10.51 nm2. On cooling from the isotropic to room temperature, a slight increase of the lattice parameters by 2–3% (∼0.1 nm) is noticeable and γ decreases from 48.4° to 46.2° at ambient (Fig. S10). The increase of a and b at lower temperatures is concomitant with shrinking of the NP corona in the third dimension, considering the volume reduction of organics with decreasing temperature. The superlattice is stable beyond the stability of the LC phase, indicating that overall the shape of the organic corona is maintained over the chemical stability region of the materials. Using the lattice area of 10.51 nm2 and the calculated average volume of an AuDLC* particle of 64.30 nm3, the value of 6.12 nm in the third dimension, likely the main axis of the orientationally ordered mesogens is needed to account for the total volume. This value is somewhat shorter than the estimated maximum possible extension of ∼8.70 nm of the overall system. This indicates that the overall shape of AuDLC* could be best described as a flat ellipsoid, directed by the packing requirements of the rigid discogens. AuDLC* can thus be viewed as a macrodisk (Fig. 2b); this is distinctly different to the shape of calamitic functionalized NPs which are often viewed as sphero-cylinders. In other words, for the investigated system the shape of the corona replicates to some extend the shape of the mesogens.Open in a separate windowFig. 2Schematic representation of (a) p2 lattice with parameters a = 4.17 nm, b = 2.52 nm and the angle γ ∼48.4°; spatial direction c not defined by the lattice. Red arrow: helix axis; purple arrow: plasmonic interactions. (b) AuDLC* macrodisk. Yellow: gold core; red: PA unit; green: chiral sheath. (c) phase formed by functionalized NPs. (d) Chiral assembly behaviour of gold NPs along the axis of smallest lattice parameter b.The formation of the macroscopic phase of the mesogens of the column forming AuNPs is due to the long flexible alkyl spacer, decoupling mesogens from the NP cores. As a result, even at high levels of surface coverage; no steric crowding occurs; orientational mobility is maintained. This is different for shorter spacer lengths of discogen ligands as evidenced in the results for a non-chiral system where no superstructure was detected.20 As the lattice parameters change very little with temperature this supports the view that the helix is perpendicular to the base of the lattice; characterized by only orientational ordering of the system; similar to other chiral materials.17To characterize the chirality further, series of CD spectra as a function of temperature for the nanocomposite AuDLC* and the free chiral discogen in thin films were recorded over a range of temperatures. For AuDLC* (Fig. S11a and b), the largest increase in ellipticity at 460 nm on cooling is between 95 °C and 90 °C. This corresponds to the transition, and for the free ligand (Fig. S11c) this occurs at 95 °C. To systematically exclude the contribution of the birefringence to the signal of the LC phase, samples were sheared in the preparation of a quartz cell before CD experiments. The LC domains were found to be small and numerous after shearing, thus the contributions from linear dichroism and birefringence are expected to cancel out. This hypothesis was confirmed by rotating the sheared sample in an interval of 90° (Fig. S11d) and additionally turning the sample over; only small changes in value but not in the sign of the CD signal occurred.17The CD bands in the UV region for AuDLC* are linked to the helical arrangement of ligands attached to AuNPs. On further cooling to 70 °C (Fig. 1f and S13), an increase of the positive CD signal in the wavelength region of 355–437 nm occurred as well as a negative signal in the plasmonic region of the gold core (Fig. S11b) (around 520 nm at 70 °C) appears. This plasmonic signal fluctuated in value on cooling further. The consistently negative value of this CD signal is in line with a tangential dipole of the attached organic corona.29,30 On further cooling, between 60 °C and 55 °C (see Fig. S12b and S13), an enhancement of the negative signal at 325 nm was observed. We associate these processes with an increased twisting of the mesogenic structures in the chiral nematic. We see fundamentally two processes, the first is associated with phase formation from the isotropic liquid, associated with the increased chiral distortions of the organic corona. The second process, commencing at ∼70 °C, is linked to chiral interactions of the NPs. Here, the packing distance measured by the lattice parameter b (2.52 nm), similar to the NP diameter (2.5 ± 0.5 nm) is the primary candidate. In the p2 lattice, the position of the gold cores along the axis c perpendicular to the lattice plane is not defined (see Fig. 2a).Due to size variation and NPs being out of position (translated along c), some disordering exists and this leads to formation of a p2 lattice by time and space averaging. On cooling the system contracts along the third dimension c and the proximity of the gold cores enveloped by a chiral surface increases. As the CD signal is small in value, when compared to theoretical models, this suggests short range ordering is developed.31To reconcile XRD, POM, DSC and CD data we propose a model which is schematically shown in Fig. 2. To minimize energy and to enhance attractive van der Waals interactions of the mesogens, the particles form an overall elongated ellipsoid shape with the mesogens at the periphery and the gold cores at the centre (Fig. 2b); determined by the volume content of the NPs, their size and size distribution a columnar superstructure of the NPs is formed, similar to other systems.32 In this model, the chiral centers at the gold surface form a chiral envelope (green corona). On entering the phase a chiral nematic helix orthogonal to the lattice plane is formed (Fig. 2c). With lowering the temperature, the CD signal in the plasmonic region of the NPs; indicative an additional helical structure along the axis of the smallest lattice parameter, as schematically shown in Fig. 2d is detected. Small translations of the system parallel to the main axis of the mesogens and packing requirements decreasing the overall volume with lower temperature are responsible for this. This is in accord with results for other chiral coated surfaces in assemblies.33,34 The CD signals in the UV region are associated with the changing chiral environment of the mesogens. The signals in the plasmonic region are through the interactions of the gold NPs.  相似文献   

13.
In situ single-crystal diffraction and spectroscopic techniques have been used to study a previously unreported Cu-framework bis[1-(4-pyridyl)butane-1,3-dione]copper(ii) (CuPyr-I). CuPyr-I was found to exhibit high-pressure and low-temperature phase transitions, piezochromism, negative linear compressibility, and a pressure induced Jahn–Teller switch, where the switching pressure was hydrostatic media dependent.

In situ high-pressure single-crystal diffraction and spectroscopic techniques have been used to study a previously unreported Cu-framework bis[1-(4-pyridyl)butane-1,3-dione]copper(ii) (CuPyr-I).

High-pressure crystallographic experiments over the last 25 years or so have proved to be a unique tool in probing the mechanical properties of the organic solid state,1 metal-complexes, and 2D/3D coordination compounds.2 In particular, high-pressure techniques have been used to study an array of mechanical and chemical properties of crystals, such as changes in electrical and thermal conductivity,3 pressure-induced melting,4 solubility,5 amorphisation,6 post-synthetic modification,7 and chemical reactions such as polymerisation,8 cycloaddition9 and nanoparticle formation.10 Previous high-pressure experiments on porous metal-organic framework (MOF) materials have shown that on loading a diamond anvil cell (DAC) with a single-crystal or polycrystalline powder, the hydrostatic medium that surrounds the sample (to ensure hydrostatic conditions) can be forced inside the pores on increasing pressure, causing the pore and sample volume to increase with applied pressure.11 This technique has also been used to determine the position of CH4 and CO2 molecules inside the small pores of a Sc-based MOF at room temperature using a laboratory X-ray diffractometer, and has proved useful in experimentally determining the maximum size of guest molecules that can penetrate into a pore.12On direct compression of more dense frameworks, negative linear compressibility (NLC) has also been observed, which results in an expansion of one or more of the unit cell dimensions with an overall contraction in volume. Such changes in the compressibility behaviour of metal-containing framework materials is usually as a result of common structural motifs which rotate or bend in order to accommodate increases in length along particular crystallographic directions.13 Changes in coordination environment can also be induced at pressure, as metal–ligand bonds are more susceptible to compression than covalent bonds.2a In previous high pressure studies on metal complexes or coordination compounds, in which the metal ion has an asymmetric octahedral environment caused by Jahn–Teller (JT) distortions for example (such as those observed in Cu2+ and Mn3+ complexes), the application of pressure can result in compression of the JT axis, and can even be switched to lie along another bonding direction within the octahedron.14 Such distortions often result in piezochromism, often observed within a single crystal.15Here, we present a high-pressure crystallographic study on a novel and unreported Cu-framework bis[1-(4-pyridyl)butane-1,3-dione]copper(ii) (hereafter referred to as CuPyr-I). On application of pressure, CuPyr-I is highly unusual in that it demonstrates several of these phenomena within the same framework, including a single-crystal to single-crystal phase transition, a switching of the JT axis that depends on the hydrostatic medium used to compress the crystal, piezochromism and NLC behaviour. To date, we are unaware of any other material which exhibits all of these phenomena, with the first ever reported hydrostatic media ‘tuneable’ JT-switching.Under ambient temperature and pressure CuPyr-I crystallises in the rhombohedral space group R3̄ (a/b = 26.5936(31) Å, c = 7.7475(9) Å). Each Cu-centre is coordinated to four 1-(4-pyridyl)butane-1,3-dione linkers, two of these ligands are bound through the dione O-atoms, with the final two bonding through the N-atom of the pyridine ring to form a 3D polymer. The crystal structure of CuPyr-I is composed of an interpenetration of these 3D polymers to form one-dimensional porous channels (∼2 Å in diameter) that run along the c-axis direction (Fig. 1).Open in a separate windowFig. 1Ball and stick model showing the coordination environment around the Cu2+ ion in CuPyr-I, and 3D-pore structure as viewed along the c-axis direction. The yellow sphere represents the available pore-space. Colour scheme is red: oxygen, blue: nitrogen, black: carbon, white: hydrogen and cyan: copper. The Cu2+ octahedron is illustrated in green.On increasing pressure from 0.07 GPa to 1.56 GPa using Fluorinert FC-70 (a mixture of large perfluorinated hydrocarbons) as a hydrostatic medium, compression of the framework occurs, resulting in a 9.89% reduction in volume, while the a/b-axes and c-axis are reduced by 4.46% and 1.25% respectively (Fig. 2 (blue triangles) and Table S1).Open in a separate windowFig. 2 a/b and c-axes as a function of pressure in a hydrostatic medium of FC-70 (blue triangles) and MeOH (red/black circles). The vertical line indicates the transition from CuPyr-I (red circles) to CuPyr-II (black circles) above 2.15 GPa. Errors in cell-lengths are smaller than the symbols plotted.On increasing pressure to 1.84 GPa, the framework became amorphous, though this is unsurprising as the hydrostatic limit for FC-70 is ∼2 GPa, and compression of frameworks in non-hydrostatic conditions usually results in amorphisation.16 On increasing pressure to 1.56 GPa, the three-symmetry independent Cu–O/N bond lengths to the ligand were monitored (Fig. 3 and Table S5). Under ambient pressure conditions, the two Cu–N1 pyridine bonds are longer than the four Cu–O1/O2 dione bonds, typical for an elongated JT distorted Cu2+ complex. However, on increasing pressure the direction of the JT axis gradually changed from Cu–N1 to the Cu–O1 bond (the dione oxygen in the 3-position), becoming equidistant at ∼0.57 GPa. By 1.56 GPa, the lengths of the Cu–N1 and Cu–O1 bonds had steadily reduced and increased by 12.3% and 8.9%, respectively. Throughout this the Cu–O2 bond remained essentially unchanged.Open in a separate windowFig. 3Cu–O1 (orange), Cu–N1 (blue) and Cu–O2 (green) bond lengths on increasing pressure in both FC-70 (triangles and dashed lines) and MeOH (circles and solid lines).Pressure induced JT switching has been observed in other systems, including a Mn12 single-molecule magnet cluster that re-orientates the JT axis on one of the Mn centres at 2.5 GPa.14a A similar transition was also observed in [CuF2(H2O)2(pyz)] (pyz = pyrazine) and Rb2CuCl4(H2O)2,15 where the JT axis was reoriented from the Cu–N bond to the perpendicular Cu–O bond, though this occurs during a crystallographic phase transition at 1.8 GPa.18 Here, in CuPyr-I, no phase transition takes place, and unusually the JT switching appears to occur progressively on increasing pressure with no phase transition.14bUsing methanol (MeOH) as the hydrostatic medium, CuPyr-I was compressed in two separate experiments, from 0.52 GPa to 5.28 GPa using synchrotron radiation, and from 0.34 GPa to 2.95 GPa using a laboratory X-ray diffractometer. On increasing pressure to 2.15 GPa, the a/b and c-axes compressed by 6.22% and 0.39% respectively (Fig. 2, Tables S2 and S3). On increasing pressure from 2.15 GPa to 2.78 GPa, CuPyr-I underwent a single-crystal to single-crystal isosymmetric phase transition to a previously unobserved phase (hereafter referred to as CuPyr-II).The transition to CuPyr-II resulted in a doubling of the a/b-axes, whilst the c-axis remained essentially unchanged. On increasing the pressure further, the a/b-axes continued to be compressed, whilst the c-axis increased in length, exhibiting negative linear compressibility (NLC) until the sample became amorphous at 5.28 GPa. The diffraction data were of poor quality after the phase transition, and only the connectivity of the CuPyr-II phase could be determined at 3.34 GPa. Above 3.34 GPa, only unit cell dimensions could be extracted. The occurrence of positive linear compressibility (PLC) followed by NLC is unusual in a framework material, and we could find only a few examples in the literature where this occurs.19During the NLC, the c-axis expanded by 1.46%, to give a compressibility of KNLC = −5.3 (0.8) TPa−1p = 2.23–4.90 GPa). KNLC is calculated using the relationship K = −1/l(∂l/∂p)T, where l is the length of the axis and (∂l/∂p)T is the length change in pressure at constant temperature.20 The value of KNLC here is rather small compared to the massive NLC behaviour observed in the low pressure phase of Ag3[Co(CN)6]9 (KNLC = −76(9) TPa−1, Δp = 0–0.19 GPa) or the flexible MOF MIL-53(Al) (KNLC = −28 TPa−1, Δp = 0–3 GPa) for example,17b and is much more comparable to the dense Zn formate MOF [NH4][Zn(HCOO)3] (−1.8(8) TPa−1p = 0–0.94 GPa)).21 Because of the quality of the data, the exact nature, or reason for the NLC in CuPyr-II is unknown, although we aim to investigate this in the future.On increasing pressure using MeOH, the JT axis was again supressed on compression, with the Cu–N1 bond reducing in length by 0.288 Å (12%) between 0.34 and 2.15 GPa, while the Cu–O1 bond length increased by 0.216 Å (11%). The pressure at which Cu–N1 and Cu–O1 became equidistant was 1.28 GPa, measuring 2.140(5) Å and 2.131(6) Å respectively (Fig. 3 and Table S6). Across the entire pressure range, little to no compression or expansion was observed in the Cu–O2 bond in the 1-position of the dione in CuPyr-I, the same trend observed when compressed in FC-70. The JT switching pressure in MeOH however was 0.71 GPa higher than observed by direct compression in FC-70 (0.57 GPa). This, to our knowledge, is the first time that pressure induced JT switching has been observed to be hydrostatic media dependent.Changes to the Cu–N and Cu–O bond lengths were supported by high-pressure Raman spectroscopy of CuPyr-I, using MeOH as the hydrostatic medium (Fig. S10). Gradual growth of a shoulder on a band at ∼700 cm−1 during compression is tentatively assigned to the Cu2+ coordination environment shifting from elongated to compressed JT distorted geometry. The shouldered peak becomes split above 2 GPa, after which the isosymmetric phase transition occurs.The gradual JT switch is thought to be principally responsible for reversible piezochromism in single crystals of CuPyr-I, which change in colour from green to dark red under applied pressure (Fig. 4b, S1 and S2). UV-visible spectroscopy confirms a bariometric blue-shift in the absorption peak at ∼700 nm assigned to d–d electronic transitions, and a red-shift of the tentatively assigned ligand-to-metal charge-transfer (LMCT) edge around 450 nm during the elongated to compressed switch (Fig. 4a and S8), accounting for this colour change. The red-shift is observed during compression in both Fluorinert® FC-70 and MeOH hydrostatic media, with a slightly suppressed shift measured in the latter due to filling of the framework pores (Table S2). Geometric switching at the metal centre leads to electronic stabilisation of the Cu2+ ion, as electrons transfer from higher energy dx2y2 (Cu–O) orbitals to the lower energy dz2 (Cu–N) state (Fig. S7), evidenced by the blue-shift of the d–d intraconfigurational band as the dz2 (Cu–N) is progressively mixing with dx2y2 (Cu–O) increasing its energy with respect to the lower energy dxy and dxz,yz levels becoming the highest energy level at the nearly compressed rhombic geometry. On the other hand, the redshift in the hesitantly assigned O2− to Cu2+ LMCT band below 450 nm is ascribed to increase of the Cu–O bond distance and a likely bandwidth broadening with pressure both yielding a pressure redshift of the absorption band gap edge (Fig. 4a).Open in a separate windowFig. 4(a) UV-visible spectroscopy of CuPyr-I during compression in Fluorinert® FC-70 showing a gradual BLUE-shift in the d–d intraconfigurational band (∼700 nm) and a gradual red-shift of the absorption band assigned to LMCT (∼450 nm) with increasing pressure. (b) Gradual pressure-induced Jahn–Teller switch of the Cu2+ octahedral coordination environment in CuPyr-1 from tetragonal elongated (left, green) to rhombic compressed (right, red), causing piezochromism. Atom colouring follows previous figures.Compression of the coordination bonds was not the only distortion to take place in CuPyr-I, with the Cu-octahedra also twisting with respect to the 1-(4-pyridyl)butane-1,3-dione linkers on increasing pressure. Twisting of the Cu-octahedra in CuPyr-I with respect to the dione section of the linker could be quantified by measuring both the ∠N1Cu1O2C4 and the ∠N1Cu1O1C2 torsion angles from the X-ray data, which in MeOH gradually decrease and increase by 12.2° and 7.3°, respectively, to 2.15 GPa (Table S8). In FC-70, ∠N1Cu1O2C4 and ∠N1Cu1O1C2 decrease and increase by 5.4° and 2.8°, respectively, to 1.56 GPa. On increasing pressure to 1.57 GPa in a hydrostatic medium of MeOH, a difference of ∼5° for both angles was observed compared to FC-70 at 1.56 GPa. Twisting about the octahedra allows compression of the channels to take place in a ‘screw’ like fashion and has been observed in other porous materials with channel structures.22 The overall effect is to reduce the pore volume, and decrease the size of the channels (Tables S2 and S3). Using MeOH as a hydrostatic medium therefore appears to reduce this effect by decreasing the compressibility of the framework.It was not possible to determine the pressure dependence in other longer-chain alcohols, including ethanol (EtOH) and isopropanol (IPA), due to cracking of the crystal upon loading into the diamond anvil cell (Fig. S1). We believe this is a result of these longer chain alcohols acting as reducing agents, as indicated by the loss in colour of the crystals.To ascertain the origin of the hydrostatic media-induced change in the JT switching pressure and unit cell compressibility, the pore size and content were monitored as a function of pressure. A dried crystal of CuPyr-I was collected at ambient pressure and temperature in order to compare to the high-pressure data and is included in the ESI. The pore volume and electron density were estimated and modelled respectively using the SQUEEZE algorithm within PLATON (Tables S1–S3).23 CuPyr-I under ambient pressure conditions has three symmetry equivalent channels per unit cell with a total volume of ∼1152 Å3 containing diethyl ether (2.5 wt%) trapped in the pores during the synthesis of the framework, confirmed by TGA analysis (Fig. S6).On surrounding the crystal with FC-70, direct compression of the framework occurred. The pore content remained almost constant during compression up to 0.88 GPa, inferring no change in the pore contents. On increasing pressure further to 1.56 GPa, an increase in the calculated electron density was observed (23%), though the data here were of depreciating quality and less reliable. During compression of CuPyr-I in MeOH to 0.52 GPa, the pore volume and electron density in the channels increased by 4.5% and 54%, respectively, reflecting ingress of MeOH into the pores. The electron density in the channels continued to increase to a maximum of 0.466e A−3 by 0.96 GPa, although the pore volume began to decrease at this pressure. The uptake of MeOH into the pores therefore results in the marked decrease in compressibility, as noted above.Previous high-pressure experiments on porous MOFs have resulted in similar behaviour on application of pressure, with the uptake of the media significantly decreasing the compressibility of the framework.24 However, using different hydrostatic media to control the JT switch in any material is, to the best of our knowledge, previously unreported. On increasing pressure above 0.96 GPa, the electron density in the pores decreases, and coincides with a steady reduction in volume of the unit cell. Both an initial increase and then subsequent decrease in uptake of hydrostatic media is common in high-pressure studies of MOFs, and has been seen several times, for example in HKUST-1 (ref. 24c) and MOF-5.24a The ingress of MeOH into the pores on initially increasing pressure to 0.52 GPa is also reflected in a twisting of the octahedra, in-particular the ∠N1Cu1O2C4 angle decreases by 5.8° in MeOH, whereas on compression in FC-70, little to no change is observed in the ∠N1Cu1O2C4 angle to 1.56 GPa. These angles represent a twisting of the dione backbone, which we speculate must interact with the MeOH molecules which penetrate into the framework.Upon compression in n-pentane, the lightest alkane that is a liquid at ambient temperature, we see different behaviour to that in MeOH or FC-70. Poor data quality permitted only the extraction of unit cell parameters but from this it can be seen that CuPyr-I has undergone the transition to CuPyr-II by 0.77 GPa. This is a significantly lower pressure than is required to induce the phase transition in MeOH (ca. 2.15 GPa). We speculate that this difference in pressure is caused by the n-pentane entering the channels at a lower pressure than MeOH due to the hydrophobic nature of the channels. This can be overcome by MeOH but not until substantially higher pressures, as seen in other MOFs that contain hydrophobic pores.25On undergoing the transition to CuPyr-II at 2.78 GPa the unit cell volume quadruples, resulting in three symmetry independent channels (12 per unit cell), with the % pore volume continuing to decrease (Table S4). Additionally, the reflections become much broader, significantly depreciating the data quality. Nevertheless, changes in metal–ligand bond lengths and general packing features can be extracted. In particular, the transition to CuPyr-II results in two independent Cu-centres, with six independent Cu–N/O bond distances per Cu. Each exhibits a continuation of the trend seen in CuPyr-I, with the Cu–O bonds (equivalent to the Cu–O1 bond in CuPyr-I) remaining longer than the JT suppressed Cu–N bonds. However, the transition to CuPyr-II results in both an increase and decrease in three of the four Cu–N and Cu–O bonds respectively, compared to CuPyr-I at 2.15 GPa (Table S6). The net result is a framework which contains a Cu-centre where the coordination bonds are more equidistant, while the JT axis becomes much more prominent in the other Cu-centre, with the Cu–O dione bond continuing to increase in length. The data for CuPyr-II depreciates rapidly after the phase transition, and more work would be required to study the effect of the anisotropic compression of the JT axis in CuPyr-II on increasing pressure further.It is difficult to determine the mechanism behind the NLC behaviour observed upon compression of CuPyr-II because the phase transition results in a significant reduction in data quality. Further work will be carried out computationally in order to elucidate the structural mechanism that gives rise to the PLC followed by NLC. However, we propose this effect is inherent to this framework and the ingress of MeOH molecules into the channels allows the retention of crystallinity to allow this behaviour to be observed crystallographically.In order to determine whether the JT switch could be induced by decreasing temperature and remove any effect the ingress of hydrostatic media has into the pores on the JT switch, variable temperature X-ray diffraction measurements were undertaken on a powder and single-crystal sample. On cooling below 175 K and 150 K in a powder and single-crystal sample respectively, a phase transition was observed, however, this was to a completely different triclinic phase, hereafter referred to as CuPyr-III. The transition here appears to occur when the disordered diethyl ether becomes ordered in the pores, confirmed by determination of the structure by single-crystal X-ray diffraction, where the diethylether could be modelled inside the pore-channel (see ESI Sections 7 & 8 for details).In conclusion, we have presented a compression study on the newly synthesised Cu-based porous framework bis[1-(4-pyridyl)butane-1,3-dione]copper(ii), referred to as CuPyr, compressed in FC-70 to 1.56 GPa and MeOH to 4.90 GPa. In both FC-70 and MeOH hydrostatic media, the JT axis, which extends along the Cu–N pyridyl bond, steadily compresses and then switches to lie along one of the Cu–O dione bonds. Compression in MeOH results in ingress of the medium into the framework pores, which increases the JT switching pressure to 1.47 GPa, compared with 0.64 GPa during compression in Fluorinert® FC-70. Interaction of stored MeOH with the host framework prompts twisting of the ligand backbone, which is not observed in the absence of adsorbed guest. Suppression of the JT axis is accompanied by a piezochromic colour change in the single crystals from green to dark red, as confirmed by crystallographic and spectroscopic measurements. Increasing the applied pressure to at least 2.15 GPa causes the framework to undergo an isosymmetric phase transition to a previously unobserved phase, characterised by a doubling of the a/b axes. Between 2.15 GPa and 4.90 GPa, NLC behaviour is observed.This is to the best of our knowledge the first time a phase transition, NLC, piezochromic and pressure induced JT switching behaviour have been observed within the same material. We have also reported for the first time a pressure induced JT axis switch which is hydrostatic media dependent. In further analysis of this system, we intend to study the magnetic properties under ambient and high pressure.  相似文献   

14.
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   

15.
Correction for ‘Rational design of a “dual lock-and-key” supramolecular photosensitizer based on aromatic nucleophilic substitution for specific and enhanced photodynamic therapy’ by Kun-Xu Teng et al., Chem. Sci., 2020, 11, 9703–9711, DOI: 10.1039/D0SC01122C.

The authors regret an error in Fig. 1f. The correct image is shown below.Open in a separate windowFig. 1(f) Fluorescence decay curves of BSG and BIBSG, and the detection wavelength is 600 nm in DMSO.Additionally, there was a minor error in Fig. 4a. The correct image is shown below.Open in a separate windowFig. 4(a) Evaluation of 1O2 generation in HepG2 cells with DCFH-DA and SOSG. The scale bar represents 20 μm.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

16.
Brønsted acid catalyzed formal [4 + 4]-, [4 + 3]-, and [4 + 2]-cycloadditions of donor–acceptor cyclobutenes, cyclopropenes, and siloxyalkynes with benzopyrylium ions are reported. [4 + 2]-cyclization/deMayo-type ring-extension cascade processes produce highly functionalized benzocyclooctatrienes, benzocycloheptatrienes, and 2-naphthols in good to excellent yields and selectivities. Moreover, the optical purity of reactant donor–acceptor cyclobutenes is fully retained during the cascade. The 1,3-dicarbonyl product framework of the reaction products provides opportunities for salen-type ligand syntheses and the construction of fused pyrazoles and isoxazoles that reveal a novel rotamer-diastereoisomerism.

Brønsted acid catalysis realizes formal [4 + 4]-, [4 + 3]-, and [4 + 2]-cycloadditions of donor–acceptor cyclobutenes, cyclopropanes, and siloxyalkynes with benzopyrylium ions.

Medium-sized rings are the core skeletons of many natural products and bioactive molecules,1 and a growing number of strategies have been developed for their synthesis.2 Because of their enthalpic and entropic advantages, ring expansion is a highly efficient methodology for these constructions.3–6 For example, Sun and co-workers have developed acid promoted ring extensions of oxetenium and azetidinium species formed from siloxyalkynes with cyclic acetals and hemiaminals (Scheme 1a).4 Takasu and co-workers have reported an elegant ring expansion with a palladium(ii) catalyzed 4π-electrocyclic ring-opening/Heck arylation cascade with fused cyclobutenes (Scheme 1b).5 Each transformation is initiated by the formation of fused bicyclic units followed by ring expansion or rearrangement to give medium-sized rings.Open in a separate windowScheme 1Cycloaddition/ring expansion background and this work.Strategies for the formation of fused bicyclic compounds rely on cycloaddition of dienes or dipoles with unsaturated cyclic compounds7 and, if the reactant cyclic compound is strained and chiral, the resulting bicyclic compound is activated toward ring opening that results in retention of chirality. We have recently reported access to donor–acceptor cycloalkenes by [3 + n] cycloaddition that have the prerequisites of unsaturated cyclic compounds suitable for cycloaddition.8 Donor–acceptor (D–A) cyclopropenes9 and cyclobutenes10 have sufficient strain in the resulting bicyclic compounds to undergo ring opening. We envision that the selection of a diene or dipolar reactant and suitable reaction conditions could realize cycloaddition and subsequent ring expansion. Benzopyrylium species,11,12 which are generated by metal or acid catalysis, have attracted our attention. We anticipated that their high reactivity would overcome the conventional unfavorable kinetic and/or thermodynamic factors that typically impede medium-sized ring formation. From a mechanistic perspective, benzopyrylium species are often formed by transition metal catalyzed reactions with 2-alkynylbenzaldehydes.11 Recently, the reaction between 1H-isochromene acetal and Brønsted acid catalyst forms the 2-benzopyrylium salts that could react with functional alkenes to give same cycloaddition products.12 Consequently, we believed that the [4 + 2]-cyclization between benzopyrylium species and donor–acceptor cyclobutenes, cyclopropenes, or siloxyalkynes would give bridged oxetenium intermediates, which contain high strain energy that should provide the driving force for ring expansion. The “push and pull” electronic effect of donor–acceptor functional groups facilitates deMayo-type ring-opening of the cyclobutane or cyclopropane skeletons (Scheme 1c).13Here we report bis(trifluoromethanesulfonyl)imide (HNTf2) catalyzed formal [4 + 4]-, [4 + 3]- and [4 + 2]-cycloaddition reactions of D–A cyclobutenes, cyclopropenes, and siloxyalkynes with benzopyrylium salts. Polysubstituted benzo-cyclooctatrienes, benzocycloheptatrienes and 2-naphthols, are produced in good to excellent yields and selectivities. Complete retention of configuration occurs using chiral cyclobutenes, and opportunities for further functionalization are built into these constructions.Initially, we conducted transition metal catalyzed reactions with 2-alkynylbenzaldehyde intending to produce the corresponding benzopyrylium ion and explore the possibility of cycloaddition/ring opening with donor–acceptor cyclobutene 2a. Use of Ph3PAuCl/AgSbF6, Pd(OAc)2 and Cu(OTf)2, which were efficient catalysts in previous transformations,11 gave only a trace amount of cycloaddition product (Fig. 1). Spectral analysis showed that mostly starting material remained (Fig. 1-i and ii). Increasing the reaction temperature led only to decomposition of 2-alkynylbenzaldehyde (Fig. 1-iv) or donor–acceptor cyclobutene 2a (Fig. 1-iv).Open in a separate windowFig. 1Reaction of 2-alkynylbenzaldehyde with donor–acceptor cyclobutene 2a.From these disappointments we turned our attention to 1H-isochromene acetal 1a as the benzopyrylium ion precursor. Various Lewis acid catalysts were employed with limited success, but we were pleased to observe the formation of the desired formal [4 + 4]-cycloaddition benzocyclooctatriene product 3aa, albeit in low yields (Table 1, entries 1–6). Selection of the Brønsted super acid HNTf2 (ref. 14) proved to be the most promising, producing 3aa in 40% yield (Table 1, entry 7). Increasing the amount of D–A cyclobutene 2a by 30% gave a much higher yield of 3aa (Table 1, entry 8 vs. 7). Optimization of the stoichiometric reaction between 1a and 2a by increasing the reaction temperature from rt to 35 °C led to the formation of the desired product in 76% isolated yield (Table 1, entry 9 vs. 8). However, a further increase in the reaction temperature (Table 1, entry 10) or reducing the catalyst loading to 5 mol% did not improve the yield of 3aa (Table 1, entry 11).Optimization of reaction conditionsa
EntryCat (10 mol%)Temp. (°C)Yieldb
1Sc(OTf)3rt23
2Yb(OTf)3rtTrace
3In(OTf)3rt16
4TiCl4rtTrace
5BF3·OEt2rt20
6TMSOTfrt17
7HNTf2rt40
8cHNTf2rt67
9cHNTf23578(76)d
10c,eHNTf26062
11c,fHNTf23550
Open in a separate windowaReactions were performed by adding the catalyst (10 mol%) to 1a (0.1 mmol) and 2a (0.1 mmol) in CH2Cl2 (2 mL) at the corresponding temperature for 24 h.bYields were determined by 1H NMR spectroscopic analysis with CH2Br2 as the internal standard.c1.3 equiv. 2a was used.dIsolated yield.eReaction performed at 60 °C for 12 h.f5 mol% catalyst loading.With optimized conditions using 1a in hand, we examined the scope of the formal [4 + 4]-cycloaddition reactions of D–A cyclobutenes 2 with a diverse set of acetal compounds 1. As shown in Scheme 2, a wide range of acetal substrates (1a–1i) with different substituents at different positions all reacted smoothly with D–A cyclobutene 2a to form the corresponding benzocyclooctatriene products 3 in good to excellent yields. Structural variations in the acetals produced only modest changes in product yields which ranged from 55 to 87%. Similarly, both electron-withdrawing and electron-donating substituents at the 4-position of the cyclobutene phenyl ring produced the corresponding products (3db–3dd) in good yields, and 2-naphthyl (2e) and 2-thienyl (2f) substituted cyclobutenes were suitable substrates (85% and 51% product yields, respectively). trans-1,2,3,4-Tetrasubstituted (R2 = CH3) 2-sil-oxycyclobutenecarboxylate 2g also underwent [4 + 4]-cycloaddition with 1d in good yield and fully retained its diastereoselectivity. The structure of 3fa was confirmed by X-ray diffraction (Scheme 2).15Open in a separate windowScheme 2Scope of the [4 + 4]-cycloaddition reaction of D–A cyclobutenes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (10 mol%) to 1 (0.1 mmol) and 2 (0.13 mmol) in CH2Cl2 (2 mL) at 35 °C for 24 h. Isolated yields are reported.Chiral donor–acceptor cyclobutenes with high enantiomeric excess and diastereoselectivity are conveniently obtained by catalytic [3 + 1]-cycloaddition of enoldiazoacetates with acyl ylides of sulfur.10 To determine if optical purity is retained, chiral D–A cyclobutene 2a (80% ee) was reacted with 1d under the optimized conditions, and the corresponding benzocyclo-octatriene product 3da was obtained in good yield with complete retention of configuration (Scheme 3, eqn (1)). With trans-disubstituted 2g and 2h that have higher optical purity, however, 3dg and 3dh were obtained in moderate yields with full retention of diastereo- and enantioselectivities, but addition products 8dg and 8dh were formed competitively (Scheme 3, eqn (2)). These compounds resulted from initial addition then desilylation, indicating that the [4 + 2]-cyclization is a stepwise reaction. Attempts to suppress the competing pathway by changing solvents or using the isopropyl acetal substrate to form a bulky 2-propanol nucleophile (1j) failed (for details, see ESI). However, p-methoxy (R1) substituted 1g that would stabilize the incipient benzopyrylium ion gave higher selectivity (4 : 1 vs. 2 : 1), and the desired [4 + 4] cycloaddition product 3gg was isolated in 50% yield with 97% ee and >19 : 1 dr (Scheme 3, eqn (3)).Open in a separate windowScheme 3Stereochemical features of the [4 + 4]-cycloaddition of D–A cyclobutenes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (10 mol%) to 1 (0.1 mmol) and 2 (0.13 mmol) in CH2Cl2 (2 mL) at 35 °C for 24 h. Isolated yields are reported.To further expand the generality of this strategy, we investigated its use with donor–acceptor cyclopropenes 4. The desired formal [4 + 3]-cycloaddition products 5 were obtained in good to excellent yields (Scheme 4). Optimized conditions used 20 mol% HNTf2 catalyst with 4 Å molecular sieves at room temperature for 2 h. The scope of this [4 + 3]-cycloaddition reaction with cyclopropenes 4 showed that acetals bearing both electron-donating and electron-withdrawing substituents on the aromatic ring were tolerated. However, as with [4 + 4]-cycloaddition reactions, the [4 + 3] reactions of 1 with R1 = alkyl or H did not produce any of the desired products. Furthermore, 3-substituted cyclopropenes 4b–4d participated in this reaction, and their products (5ab–5ad) were obtained in 63%–81% yields.Open in a separate windowScheme 4Scope of the [4 + 3]-cycloaddition reaction of D–A cyclopropenes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (20 mol%) to 1 (0.2 mmol), 4 (0.24 mmol) and 4 Å (50 mg) in CH2Cl2 (2 mL) at rt for 2 h. Isolated yields are reported.Siloxyalkynes 6, as electron-rich alkynes, have been widely used in diverse cyclization reaction.4,16 We expected that 6 could also participate in [4 + 2]-cyclization/ring-expansion cascade processes, giving substituted 2-naphthol products. Interestingly, substituent controlled diverse products were obtained in good to excellent reactivity and selectivity (Scheme 5). Aryl (R1) substituted acetals (1a–1d, 1f, 1h, 1i) reacted with siloxyalkynes 6a–6d, giving 2-naphthols (7aa–7ad and 7ba–7ia) in 35%–96% yields. With electron-donating group (EDG) substituents (7ba and 7ca) on the aromatic ring, higher reactivity was observed relative to those with electron-withdrawing groups (7da–7ia). In addition, the acetal with R1 = H (1m) reacted with siloxyalkynes 6 to form the 2-naphthol-1-carboxaldehyde derivative in good yield, and the structure of 7ma was confirmed by X-ray diffraction (Scheme 5b).15 Intriguingly, the n-butyl (R1) substituted acetal 1n reacted with siloxyalkynes via a [4 + 2]-cyclization with loss of methyl pentanoate (BuCO2CH3), affording siloxy naphthalenes 7na–7nd that are important precursors to the widely used axially chiral 2,2′-binols.17 The substrate scope of siloxyalkynes 6 for their formal [4 + 2]-cycloaddition reaction with n-butyl substituted acetal 1n was also explored (Scheme 5c). In all cases methyl pentanoate was eliminated to form 2,3-disubstituted naphthalene products (7na–7nd). Alkyl substituted siloxyalkynes (6a–6c) showed higher reactivity compared with phenyl substituted siloxyalkynes 6d. It should be mentioned that, recently, a similar transformation using BF3·OEt2 as catalyst or in excess (2 equiv.) with 2,4,6-collidine (1 equiv.) was reported,18 and HNTf2 was stated to be much less effective. To clarify this discrepancy, we carefully repeated these transformations (7ma and 7na) and found that all starting materials are completely consumed in less than 10 min to deliver [4 + 2]-cycloaddition products in good yields. Prolonging the reaction time to 12 hours, which was the reaction time used by the authors, results in their decomposition to a complex mixture of materials.Open in a separate windowScheme 5Scope of the [4 + 2]-cycloaddition reaction of siloxyalkynes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (10–30 mol%) to 1 (0.2 mmol) and 6 (0.3–0.4 mmol) in CH2Cl2 (2 mL) at rt for 10 min. Isolated yields are reported. b40 mol% HNTf2 catalyst was used.To illustrate the utility of this process, a large scale catalytic [4 + 4] cycloaddition was performed, and adduct 3da was obtained in 87% yield. Further transformations were conducted for the synthesis of pyrazole and isoxazole structures based on its 1,3-dicarbonyl skeleton (Fig. 2a). Compound 3da reacted with hydrazine and hydroxylamine in refluxed ethanol, affording pyrazole 14da and isoxazole 15da in 89% yield or 66% yield, respectively. The structure of 15da was confirmed by X-ray diffraction.15 Interestingly, two NMR distinguishable interconvertible diastereoisomers were detected for each of these eight-membered cyclic products (2.5 : 1 and 5 : 1 dr for 14da and 15da, respectively, in CDCl3). These diastereoisomers are rotamers (for details, see ESI) that exist at equilibrium with each other in solution but form one crystalline product (X-ray structure of 15da). In addition, the cycloaddition product 7ma reacted with chiral 1,2-cyclohexanediamine and 1,2-diphenylethylenediamine to give salen-type ligands L1 and L2 in 53% yield and 74% yield, respectively, which provides new opportunities for ligand screening (Fig. 2b).19Open in a separate windowFig. 2Large scale reaction, further transformations, and ligands synthesis.In the mechanistic possibility considered for these HNTf2 catalyzed cycloaddition reactions (Fig. 3), protonation of acetal 1 with HNTf2 gives the corresponding highly reactive benzopyrylium intermediate int-I, which reacts with donor–acceptor cyclobutenes 2, cyclopropenes 4, or siloxyalkynes 6 affording addition intermediates int-II that undergo ring closure to int-III. Ring expansion then occurs to deliver 3, 5 and 7 in good to excellent yields with fully retained stereoselectivities. Furthermore, the formed TIPSNTf2 ([4 + 4]- and [4 + 3]-cycloaddition) or HNTf2 {[4 + 2] cycloaddition} are active acid catalysts for the conversion of 1 to benzopyrylium intermediate int-I that continues the catalytic cycle. With sterically larger D–A cyclobutenes or when a less ring-strained benzocyclopentene 12 is employed (for details, see ESI), the competing direct desilylation of int-II occurs, delivering addition byproducts 8 or 13. Compounds 7na–7nd arise from the analog to int-III from which ketene formation or methanol displacement effects 1,4-elimination.Open in a separate windowFig. 3Proposed mechanism.  相似文献   

17.
18.
β-Difluoroalkylborons, featuring functionally important CF2 moiety and synthetically valuable boron group, have great synthetic potential while remaining synthetically challenging. Herein we report a hypervalent iodine-mediated oxidative gem-difluorination strategy to realize the construction of gem-difluorinated alkylborons via an unusual 1,2-hydrogen migration event, in which the (N-methyliminodiacetyl) boronate (BMIDA) motif is responsible for the high regio- and chemoselectivity. The protocol provides facile access to a broad range of β-difluoroalkylborons under rather mild conditions. The value of these products was demonstrated by further transformations of the boryl group into other valuable functional groups, providing a wide range of difluorine-containing molecules.

A hypervalent iodine-mediated gem-difluorination allows the facile synthesis of β-difluoroalkylborons. An unusual 1,2-hydrogen migration, triggered by boron substitution, is involved.

Organofluorine compounds have been widely applied in medicinal chemistry and materials science.1ad In particular, the gem-difluoro moiety featuring unique steric and electronic properties can act as a chemically inert isostere of a variety of polar functional groups.2ac Therefore, the construction of gem-difluoro-containing compounds has received considerable attention in recent years. Efficient methods including deoxyfluorination of carbonyl compounds,3a,b photoredox difluorination,4 radical difluorination,5 and cross-coupling reactions with suitable CF2 carriers6af are well developed. Alternatively, iodoarene-mediated oxidative difluorination reactions provide valuable access to these motifs by using simple alkenes as starting materials.7ai Previously, these reactions were generally associated with a 1,2-aryl or 1,2-alkyl migration (Scheme 1a).7af Recent developments also allowed the use of heteroatoms as migrating groups, thereby furnishing gem-difluoro compounds equipped with easily transformable functional groups (Scheme 1b). In this regard, Bi and coworkers reported an elegant 1,2-azide migrative gem-difluorination of α-vinyl azides, enabling the synthesis of a broad range of novel β-difluorinated alkyl azides.7g Jacobsen developed an iodoarene-catalyzed synthesis of gem-difluorinated aliphatic bromides featuring 1,2-bromo migration with high enantioselectivity.7h Almost at the same time, research work from our group demonstrated that not only bromo, but also chloro and iodo could serve as viable migrating groups.7iOpen in a separate windowScheme 1Hypervalent iodine-mediated β-difluoroalkylboron synthesis.We have been devoted to developing new methodologies for the assembly of boron-containing building blocks by using easily accessible and stable MIDA (N-methyliminodiacetyl) boronates8ac as starting materials.9ae Recently, we realized a hypervalent iodine-mediated oxidative difluorination of aryl-substituted alkenyl MIDA boronates.9d Depending on the substitution patterns, the reaction could lead to the synthesis of either α- or β-difluoroalkylborons via 1,2-aryl migration (Scheme 1c). Recently, with alkyl-substituted branched alkenyl MIDA boronates, Szabó and Himo observed an interesting bora-Wagner–Meerwein rearrangement, furnishing β-difluorinated alkylboronates with broader product diversity (Scheme 1d).10 While extending the scope of our previous work,9d we found that the use of linear alkyl-substituted alkenyl MIDA boronates also delivers β-difluoroalkylboron products. Intriguingly, instead of an alkyl- or boryl-migration, an unusual 1,2-hydrogen shift takes place. It should be noted that internal inactivated alkenes typically deliver the 1,2-difluorinated products, with no rearrangement taking place.11ad Herein, we disclose our detailed study of our second generation of β-difluoroalkylborons synthesis (Scheme 1e). The starting linear 1,2-disubstituted alkyl-substituted alkenyl MIDA boronates, unlike the branched ones,10 could be readily prepared via a two-step sequence consisting of hydroborylation of the terminal alkyne and a subsequent ligand exchange with N-methyliminodiacetic acid. This intriguing 1,2-H shift was found to be closely related to the boron substitution, probably driven thermodynamically by the formation of the β-carbon cation stabilized by a σ(C–B) bond via hyperconjugation.12adTo start, we employed benzyl-substituted alkenyl MIDA boronate 1a as a model substrate (9d the use of F sources such as CsF, AgF and Et3N·HF in association with PhI(OAc)2 (PIDA) as the oxidant and DCM as the solvent led to no reaction (entries 1 to 3). The use of Py·HF (20 equiv) successfully provided β-difluorinated alkylboronate 2a, derived from an unusual 1,2-hydrogen migration, in 39% yield (entry 4). By simply increasing the loading of Py·HF to 40 equivalents, a higher conversion and thus an improved yield of 61% was obtained (entry 5). No further improvement was observed by using a large excess of Py·HF (100 equiv) (entry 6). Other hypervalent iodine oxidants such as PhIO or PIFA were also effective but resulted in reduced yields (entries 7 and 8). A brief survey of other solvents revealed that the original DCM was the optimal one (entries 9 and 10).Optimization of reaction conditions
EntryF (equiv)OxidantSolventYield (%)
1CsF (2.0)PIDADCM0
2AgF (2.0)PIDADCM0
3Et3N·HF (40.0)PIDADCM0
4Py·HF (20.0)PIDADCM39
5 Py·HF (40.0) PIDA DCM 61
6Py·HF (100.0)PIDADCM55
7Py·HF (40.0)PIFADCM52
8Py·HF (40.0)PhIODCM26
9Py·HF (40.0)PIDADCE49
10Py·HF (40.0)PIDAToluene46
Open in a separate windowWith the optimized reaction conditions in hand, we set out to investigate the scope and limitation of this gem-difluorination reaction. The reaction of a series of E-type 1,2-disubstituted alkenyl MIDA boronates were first examined. As shown in Scheme 2, the reaction of substrates with primary alkyl (1b, 1e–g), secondary alkyl (1c, 1d), or benzyl (1h–k) groups proceeded efficiently to give the corresponding gem-difluorinated alkylboronates in moderate to good yields. Halides (1i–k, 1m) and cyano (1l) were well tolerated in this reaction. Of note, cyclic alkene 1n is also a viable substrate, affording an interesting gem-difluorinated cyclohexane product (2n).Open in a separate windowScheme 2Scope of 1,2-H migratory gem-difluorinations. a 4 h. b PIFA was used.To define the scope further, the substrates with Z configuration were also employed under the standard reaction conditions (eqn (1) and (2)). The same type of products were isolated with comparable efficiency, suggesting that the reaction outcome is independent of the substrate configuration and substrates with Z configuration also have a profound aptitude of 1,2-hydrogen migration. Nevertheless, the reaction of t-butyl substituted alkenyl MIDA boronate (1p) delivered a normal 1,2-difluorinated alkylboron product (eqn (3)). The 1,2-hydrogen migration was completely suppressed probably due to unfavorable steric perturbation. With an additional alkyl substituent introduced, a 1,2-alkyl migrated product was formed as expected (eqn (4)).1The gem-difluorination protocol was amenable to gram-scale synthesis of 2a (Scheme 3, 8 mmol scale of 1a, 1.24 g, 50%). To assess the synthetic utility of the resulting β-difluorinated alkylborons, transformations of the C–B bond were carried out (Scheme 3). Ligand exchange of 2a furnished the corresponding pinacol boronic ester 4 without difficulty, which could be ligated with electron-rich aromatics to obtain 5 and 6 in moderate yields. On the other hand, 2a could be oxidized with high efficiency to alcohol 7 using H2O2/NaOH. The hydroxyl group of 7 could then be converted to bromide 8 or triflate 9. Both serve as useful electrophiles that can undergo intermolecular SN2 substitution with diverse nitrogen- (10, 13), oxygen- (14), phosphorus- (11) and sulfur-centered (12) nucleophiles.Open in a separate windowScheme 3Product derivatizations. PMB = p-methoxyphenyl.To gain insight into the reaction mechanism, preliminary mechanistic studies were conducted. The reaction employing deuterated alkenyl MIDA boronate [D]-1a efficiently afforded difluorinated product [D]-2a in 72% isolated yield, clearly demonstrating that 1,2-H migration occurred (Scheme 4a). However, when the MIDA boronate moiety was replaced with a methyl group (15), no difluorinated product (derived from 1,2-migration) was detected at all, suggesting an indispensable role of boron for promoting the 1,2-migration event (Scheme 4b). Also, with a Bpin congener of 1a, the reaction led to large decomposition of the starting material, with no desired product being formed (Scheme 4b).Open in a separate windowScheme 4Mechanistic studies and proposals.Based on the literature precedent and these experiments, a possible reaction mechanism is proposed in Scheme 4c. With linear alkenyl MIDA boronates, the initial coordination of the double bond to an iodium ion triggered a regioselective fluoroiodination to deliver intermediate B. The regioselectivity could arise from an electron-donating inductive effect from boron due to its low electronegativity, consistent with previous observations.13a,b Thereafter, a 1,2-hydrogen shift, rather than the typical direct fluoride substitution of the C–I bond, provides carbon cation C. The formation of a hyperconjugatively stabilized cation is believed to be the driving force for this event.12ad The trapping of this cation finally forms the product.In conclusion, we demonstrated herein our second generation of β-difluoroalkylboron synthesis via oxidative difluorination of easily accessible linear 1,2-disubstituted alkenyl MIDA boronates. An unexpected 1,2-hydrogen migration was observed, which was found to be triggered by a MIDA boron substitution. Mild reaction conditions, moderate to good yields and excellent regioselectivity were achieved. The applications of these products allowed the facile preparation of a wide range of gem-difluorinated molecules by further transformations of the boryl group.  相似文献   

19.
The direct catalytic asymmetric hydrogenation of (Z)-α-dehydroamino boronate esters was realized. Using this approach, a class of therapeutically relevant alkyl-substituted α-amidoboronic esters was easily synthesized in high yields with generally excellent enantioselectivities (up to 99% yield and 99% ee). The utility of the products has been demonstrated by transformation to their corresponding boronic acid derivatives by a Pd-catalyzed borylation reaction and an efficient synthesis of a potential intermediate of bortezomib. The clean, atom-economic and environment friendly nature of this catalytic asymmetric hydrogenation process would make this approach a new alternative for the production of alkyl-substituted α-amidoboronic esters of great potential in the area of organic synthesis and medicinal chemistry.

The direct catalytic asymmetric hydrogenation of (Z)-α-dehydroamino boronate esters was realized.

Since FDA approval of bortezomib1 for the treatment of multiple myeloma, chiral α-aminoboronic acids have been recognized as key pharmacophores for the design of proteasome inhibitors.2 The incorporation of chiral α-aminoboronic acid motifs at the C-terminal position of a peptide3 to develop potential clinical drug candidates has drawn increasing interest4 (Fig. 1). Meanwhile, chiral α-amidoboronic acids and their derivatives are useful synthetic building blocks for the stereospecific construction of chiral amine compounds.5 The biological and synthetic value of α-amidoboronates has led to considerable efforts for the development of efficient synthetic methods. However, up to now, limited transition-metal-catalyzed asymmetric approaches have been reported. The widely used strategies to synthesize these compounds are stepwise Matteson homologation/N-nucleophilic replacement,6 borylation of imines,7 and alkene functionalization.8 Recently, two other elegant approaches, Ni-catalyzed decarboxylative borylation of α-amino acid derivatives9 and enantiospecific borylation of lithiated α-N-Boc species,10 were reported by the Baran and Negishi groups, respectively. To the best of our knowledge, the majority of the methods relied on either stoichiometric amounts of chiral auxiliaries6,7a,b or substrate-control strategies9 and most of these methods enable the construction of aryl-substituted α-aminoboronates. Enantioselective methods to access unfunctionalized alkyl-substituted α-aminoboronic esters are still rarely developed and so far only two examples have been realized by the Miura8a and Scheidt7f groups, respectively. Considering that most therapeutically relevant α-amidoboronic acid fragments contain an alkyl subunit and the fact that the options for the synthesis of alkyl-substituted α-amidoboronic esters in an enantioselective manner are still rare, the development of other distinct approaches would be highly desirable. Herein, we report a new alternative to access these compounds by catalytic asymmetric hydrogenation of (Z)-α-dehydroamidoboronate esters. With this approach, the desired chiral alkyl-substituted α-amidoboronic esters could be obtained in high yields and generally excellent enantioselectivities (up to 99% yield and 99% ee) with simple purification.Open in a separate windowFig. 1Selected inhibitors containing chiral alkyl-substituted α-amidoboronic acids.Catalytic asymmetric hydrogenation of olefins is an atom-economic, environmentally friendly and clean process for the synthesis of valuable pharmaceuticals, agricultural compounds and feedstock chemicals.11 Recently, hydrogenation of vinylboronic compounds has emerged for the preparation of chiral boronic compounds in a regiodefined manner.12,13 However, surprisingly α-dehydroamido boronate esters and their derivatives, as elegant precursors to access alkyl-substituted α-amidoboronic compounds, have never been used as substrates in asymmetric hydrogenation and remain a challenging project. To our knowledge, only one efficient hydrogenation approach to (1-halo-1-alkenyl) boronic esters was reported for indirect synthesis of alkyl-substituted α-aminoboronic esters but it was accompanied by inevitable de-halogenated by-products14 (Scheme 1). Given the catalytic efficiency and atom economy of the hydrogenation method, the development of a new direct hydrogenation approach to construct these important chiral alkyl-substituted α-amidoboronic esters would be very appealing.Open in a separate windowScheme 1Approaches towards the synthesis of chiral alkyl-substituted α-aminoboronic esters.The inspiration for our approach to the hydrogenation of α-dehydroamido boronates came from the molecular structures of relevant biologically active inhibitors containing alkyl-substituted α-amidoboronic acid fragments. Due to the limited stability of free α-aminoboronic acids, an electron-withdrawing carboxylic N-substituent is often required.15 Thus, we envisaged that N-carboxyl protected α-dehydroamido boronate esters could serve as a potential precursor for the synthesis of alkyl-substituted α-amidoboronates through Rh-catalyzed asymmetric hydrogenation of the C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bond16 (Fig. 1), a strategically distinct approach to the construction of unfunctionalized alkyl-substituted α-amidoboronic esters. However, challenges still remain, including: (1) how to synthesize α-dehydroamido boronates; (2) the facile transmetalation process of the starting materials leading to deboronated by-products in the hydrogenation process;17 (3) the unknown stability of α-amidoboronic compounds in the presence of a transition-metal catalyst and hydrogen molecules. As part of our continuous efforts to develop efficient hydrogenation approaches to construct valuable motifs,18 here we present the results of the investigation to address the aforementioned challenges.The desired aryl-substituted (Z)-α-dehydroamido boronates could be obtained by Cu-catalyzed regioselective hydroborylation of ynamide according to a previous report.19 However, different α/β-regioselectivity was observed for the preparation of alkyl-substituted (Z)-α-dehydroamido boronate esters and a new synthetic route was developed (Scheme 2, see the ESI for details). Of note, (Z)-α-dehydroamido boronate esters should be purified with deactivated silica gel,7c or else protodeborylation would occur readily with flash chromatography.Open in a separate windowScheme 2Synthetic route to (Z)-α-dehydroamino boronates.In order to check the feasibility of our hypothesis, three substrates were prepared with Rh(NBD)2BF4 and examined and our group prepared (Rc,Sp)-DuanPhos under 50 atm hydrogen pressure (Table 1). Gratifyingly, substrate 1b reacted smoothly to provide the desired product 2b in high yield and enantioselectivity (>99% conv., 98% ee, entry 2) whilst the reaction with substrate 1a yielded a mixture of deborylation products and 1c did not work at all (entries 1 and 3). Of note, we did not observe deborylation products with 1b under the current reaction conditions and we did not select (Z)-α-dehydroamido boronic acid 1a as the model substrate because of its poor solubility in most solvents. Then, a variety of chiral diphosphine ligands were investigated along with Rh(NBD)2BF4 and the results are shown in Table 1. In most cases, the reaction proceeded smoothly to furnish the desired products and the best results were obtained when (Rc,Sp)-DuanPhos was used as the ligand (entries 2 and 4–12). Poor results were obtained with axially bidentate phosphine ligands (entries 5, 6 and 9). (R,R)-QuinoxP* and (R,R)-Ph-BPE also gave good conversion with a slightly decreased ee whilst (R,R)-iPr-DuPhos exhibited poor results (entries 4, 7 and 10). Subsequent solvent screening revealed that the desired products could be obtained in most of the solvents and 1,2-DCE was the best solvent. (Entry 13, see the ESI).Condition optimization for catalytic asymmetric hydrogenation of 1a
EntrySubLigandConv.b (%)eec (%)
1d 1a (Rc,Sp)-DuanPhos89n.d.
2e 1b (Rc,Sp)-DuanPhos>9998
3 1c (Rc,Sp)-DuanPhosn.r.n.d.
4 1b (R,R)-QuinoxP*>9997
5 1b (S)-SegPhos>9917
6 1b (S)-BINAP>9910
7 1b (R,R)-iPr-DuPhos>993
8 1b (R,S)-Cy-JosiPhos>9914
9 1b (R)-BIPHEP>99−30
10 1b (R,R)-Ph-BPE>99−86
11 1b (S,S)-f-Binaphane>9961
12 1b (2S,4S)-BDPP>9959
13e,f,g 1b (Rc,Sp)-DuanPhos>99(99)99
Open in a separate windowaUnless otherwise mentioned, the reactions were performed with 1 (0.1 mmol), Rh(NBD)2BF4 (10 mol%), and a ligand (11 mol%) in 1.0 mL THF at 50 °C for 15 h.bDetermined by crude 1H NMR.cDetermined with chiral HPLC.dThe reaction was performed in iPrOH.eRh(NBD)2BF4 (1.0 mol%) and ligand (1.05 mol%) were used.fIsolated yield in parentheses.g1,2-DCE was used as the solvent. Pin = 2,3-dimethyl-2,3-butanediol; dan = 1,8-diaminonaphthalene.With the optimized reaction conditions in hand, a series of (Z)-α-dehydroamido boronate esters were tested and the results are summarized in Table 2. All the substrates reacted smoothly to give the corresponding alkyl-substituted α-amidoboronates in high yields with good to excellent enantioselectivities (2b, 2d–2r, and 2u, 99% yield, 57–99% ee). Alkyl-substituted (Z)-α-dehydroamido boronate esters were well tolerated in the current reaction, providing the corresponding α-amidoboronates in high yields and excellent enantioselectivities (2d–2i, 99% yield, 96–99% ee). Aryl-substituted (Z)-α-dehydroamido boronate esters with electron-donating (2j–l, 2n and 2p–r) and withdrawing (2m and 2o) substituents could also give the desired products in excellent yield with excellent enantioselectivities (90–99% ee). The ortho-methyl-substituted substrate 1r reacted smoothly to give the desired product with excellent enantioselectivity, but the 2,6-dimethyl-substituted substrate 1z could not react at all. Functional groups such as ether, halo and benzyl were well tolerated in the current reaction (2k, 2l, 2m and 2o–q). Replacement of the N-substituents with acyclic carbamate was also tolerated but with a decreased ee (2u and 2z). Substrates containing a chiral oxazolidin-2-one unit bearing bulky Ph-substituents around the nitrogen and oxygen were also competent, yielding the desired products with good to excellent diastereoselectivities (2s, 2t, and 2v–y). Of note, the substrate 1s bearing an N-Ms substituent and the cyclic substrate 1t did not work in the current reaction. The absolute configuration of generated α-amidoboronates was assigned as (S) by X-ray crystallographic analysis of 2i (Scheme 3).20Substrate scope.a,b,c
Open in a separate windowaUnless otherwise mentioned, the reactions were performed with 1 (0.1 mmol), Rh(NBD)2BF4 (1.0 mol%), and a ligand (1.05 mol%) in 1.0 mL 1,2-DCE at 50 °C under 50 atm H2 for 15 h.bIsolated yield.cDetermined with chiral HPLC.dDetermined by crude 1H NMR.Open in a separate windowScheme 3Scale-up synthesis and synthetic utility.To demonstrate the utility of the products, a scale-up reaction (0.62 g) was successfully performed with 0.1 mol% catalytic loading, giving 2b in 99% yield and 98% ee, and 2b could be easily transformed to a more stable α-amidoborate 3b with KHF2,6d,21 followed by hydrolysis with TMSCl to yield α-amido boronic acid 4b in 46% yield,22 which could also be obtained from 2b by treating it with BCl3 in 84% yield, without loss of the optical purity.8b2m could easily be transformed to 4m in 68% yield by a Pd-catalyzed borylation reaction. Meanwhile, after hydrogenation of 1x to 2x′ and transformation of 2x′ to its trifluoroborate derivative 3x′, removal of the benzyl group of 3x′ with Pd/C under hydrogenation conditions23 yielded the primary α-aminoborate 4x in 62% yield in three steps, which could serve as a potential precursor15 to synthesize bortezomib.  相似文献   

20.
Palladium-catalyzed regioselective di- or mono-arylation of o-carboranes was achieved using weakly coordinating amides at room temperature. Therefore, a series of B(3,4)-diarylated and B(3)-monoarylated o-carboranes anchored with valuable functional groups were accessed for the first time. This strategy provided an efficient approach for the selective activation of B(3,4)–H bonds for regioselective functionalizations of o-carboranes.

B–H: site-selective B(3,4)–H arylations were accomplished at room temperature by versatile palladium catalysis enabled by weakly coordinating amides.

o-Carboranes, icosahedral carboranes – three-dimensional arene analogues – represent an important class of carbon–boron molecular clusters.1 The regioselective functionalization of o-carboranes has attracted growing interest due to its potential applications in supramolecular design,2 medicine,3 optoelectronics,4 nanomaterials,5 boron neutron capture therapy agents6 and organometallic/coordination chemistry.7 In recent years, transition metal-catalyzed cage B–H activation for the regioselective boron functionalization of o-carboranes has emerged as a powerful tool for molecular syntheses. However, the 10 B–H bonds of o-carboranes are not equal, and the unique structural motif renders their selective functionalization difficult, since the charge differences are very small and the electrophilic reactivity in unfunctionalized o-carboranes reduces in the following order: B(9,12) > B(8,10) > B(4,5,7,11) > B(3,6).8 Therefore, efficient and selective boron substitution of o-carboranes continues to be a major challenge.Recently, transition metal-catalyzed carboxylic acid or formyl-directed B(4,5)–H functionalization of o-carboranes has drawn increasing interest, since it provides an efficient approach for direct regioselective boron–carbon and boron–heteroatom bond formations (Scheme 1a),9 with major contributions by the groups of Xie,10 and Yan,11 among others.12 Likewise, pyridyl-directed B(3,6)–H acyloxylations (Scheme 1b),13 and amide-assisted B(4,7,8)–H arylations14 (Scheme 1c) have been enabled by rhodium or palladium catalysis, respectively.15,16 Despite indisputable progress, efficient approaches for complementary site-selective functionalizations of o-carboranes are hence in high demand.17 Hence, metal-catalyzed position-selective B(3,4)–H functionalizations of o-carboranes have thus far not been reported.Open in a separate windowScheme 1Chelation-assisted transition metal-catalyzed cage B–H activation of o-carboranes.Arylated compounds represent key structural motifs in inter alia functional materials, biologically active compounds, and natural products.18 In recent years, transition metal-catalyzed chelation-assisted arylations have received significant attention as environmentally benign and economically superior alternatives to traditional cross-coupling reactions.19 Within our program on sustainable C–H activation,20 we have now devised a protocol for unprecedented cage B–H arylations of o-carboranes with weak amide assistance, on which we report herein. Notable features of our findings include (a) transition metal-catalyzed room temperature B–H functionalization, (b) high levels of positional control, delivering B(3,4)-diarylated and B(3)-monoarylated o-carboranes, and (c) mechanistic insights from DFT computation providing strong support for selective B–H arylation (Scheme 1d).We initiated our studies by probing various reaction conditions for the envisioned palladium-catalyzed B–H arylation of o-carborane amide 1a with 1-iodo-4-methylbenzene (2a) at room temperature (Tables 1 and S1). We were delighted to observe that the unexpected B(3,4)-di-arylated product 3aa was obtained in 59% yield in the presence of 10 mol% Pd(OAc)2 and 2 equiv. of AgTFA, when HFIP was employed as the solvent, which proved to be the optimal choice (entries 1–5).21 Control experiments confirmed the essential role of the palladium catalyst and silver additive (entries 6–7). Further optimization revealed that AgOAc, Ag2O, K2HPO4, and Na2CO3 failed to show any beneficial effect (entries 8–11). Increasing the reaction temperature fell short in improving the performance (entries 12 and 13). The replacement of the amide group in substrate 1a with a carboxylic acid, aldehyde, ketone, or ester group failed to afford the desired arylation product (see the ESI). We were pleased to find that the use of 1.0 equiv. of trifluoroacetic acid (TFA) as an additive improved the yield to 71% (entry 14). To our delight, replacing the silver additive with Ag2CO3 resulted in the formation of B(3)–H mono-arylation product 4aa as the major product (entries 15–16).Optimization of reaction conditionsa
EntryAdditiveSolventYield of 3aa/%Yield of 4aa/%
1AgTFAPhMe00
2AgTFADCE00
3AgTFA1,4-Dioxane00
4AgTFATFE213
5AgTFAHFIP594
6AgTFAHFIP00b
7HFIP00
8AgOAcHFIP5<3
9Ag2OHFIP<3<3
10K2HPO4HFIP00
11Na2CO3HFIP00
12AgTFAHFIP534c
13AgTFAHFIP423d
14 AgTFA HFIP 71 <3 e
15Ag2CO3HFIP934f
16 Ag 2 CO 3 HFIP 5 55 f , g
Open in a separate windowaReaction conditions: 1a (0.20 mmol), 2 (0.48 mmol), Pd(OAc)2 (10 mol%), additive (0.48 mmol), solvent (0.50 mL), 25 °C, 16 h, and isolated yield.bWithout Pd(OAc)2.cAt 40 °C.dAt 60 °C.eTFA (0.2 mmol) was added.f 1a (0.20 mmol), 2a (0.24 mmol), Pd(OAc)2 (5.0 mol%), and Ag2CO3 (0.24 mmol).g 2a was added in three portions every 4 h. DCE = dichloroethane, TFE = 2,2,2-trifluoroethanol, HFIP = hexafluoroisopropanol, and TFA = trifluoroacetic acid.With the optimized reaction conditions in hand, we probed the scope of the B–H di-arylation of o-carboranes 1a with different aryl iodides 2 (Scheme 2). The versatility of the room temperature B(3,4)–H di-arylation was reflected by tolerating valuable functional groups, including bromo, chloro, and enolizable ketone substituents. The connectivity of the products 3aa and 3ab was unambiguously verified by X-ray single crystal diffraction analysis.22Open in a separate windowScheme 2Cage B(3,4)–H di-arylation of o-carboranes.Next, we explored the effect exerted by the N-substituent at the amide moiety (Scheme 3). Tertiary amides 1b–1f proved to be suitable substrates with optimal results being accomplished with substrate 1a. The effect of varying the cage carbon substituents R1 on the reaction''s outcome was also probed, and both aryl and alkyl substituents gave the B–H arylation products and the molecular structures of the products 3dd, 3ea and 3fa were fully established by single-crystal X-ray diffraction.Open in a separate windowScheme 3Effect of substituents on B–H diarylation. aAt 50 °C.The robustness of the palladium-catalyzed B–H functionalization was subsequently investigated for the challenging catalytic B–H monoarylation of o-carboranes (Scheme 4). The B(3)–H monoarylation, as confirmed by single-crystal X-ray diffraction analysis of products 4aa and 4ai, proceeded smoothly with valuable functional groups, featuring aldehyde and nitro substituents, which should prove invaluable for further late-stage manipulation.Open in a separate windowScheme 4Cage B(3)–H mono-arylation of o-carboranes.To elucidate the palladium catalysts'' working mode, a series of experiments was performed. The reactions in the presence of TEMPO or 1,4-cyclohexadiene produced the desired product 3aa, which indicates that the present B–H arylation is less likely to operate via radical intermediates (Scheme 5a). The palladium catalysis carried out in the dark performed efficiently (Scheme 5b). Compound 4aa could be converted to di-arylation product 3aa with high efficiency, indicating that 4aa is an intermediate for the formation of the diarylated cage 3aa (Scheme 5c).Open in a separate windowScheme 5Control experiments.To further understand the catalyst mode of action, we studied the site-selectivity of the o-carborane B–H activation for the first B–H activation at the B3 versus B4 position and for the second B–H activation at the B4 versus B6 position using density functional theory (DFT) at the PBE0-D3(BJ)/def2-TZVP+SMD(HFIP)//TPSS-D3(BJ)/def2-SVP level of theory (Fig. 1). Our computational studies show that the B3 position is 5.8 kcal mol−1 more favorable than the B4 position for the first B–H activation, while the B4 position is 3.4 kcal mol−1 more favorable than the B6 position for the second B–H activation. It is noteworthy that here the interaction between AgTFA and a cationic palladium(ii) complex was the key to success, being in good agreement with our experimental results (for more details, see the ESI).Open in a separate windowFig. 1Computed relative Gibbs free energies in kcal mol−1 and the optimized geometries of the transition states involved in the B–H activation at the PBE0-D3(BJ)/def2-TZVP+SMD(HFIP)//TPSS-D3(BJ)/def2-SVP level of theory. (a) First B–H activation transition states at the B3 and B4 positions. (b) Second B–H activation transition states at the B4 and B6 positions. Irrelevant hydrogen atoms in the transition states are omitted for clarity and the bond lengths are given in Å.A plausible reaction mechanism is proposed which commences with an organometallic B(3)–H activation of 1a with weak assistance of the amide group and assistance by AgTFA to form the cationic intermediate I (Scheme 6). Oxidative addition with the aryl iodide 2 affords the proposed cationic palladium(iv) intermediate II, followed by reductive elimination to give the B(3)-mono-arylation product 4aa. Subsequent B(4)-arylation occurs assisted by the weakly coordinating amide to generate the B(3,4)-di-arylation product 3aa. Due to the innate higher reactivity of the B(4)–H bond in intermediate 4aa – which is inherently higher than that of the B(6)–H bond – the B(3,6)-di-arylation product is not formed.Open in a separate windowScheme 6Proposed reaction mechanism.In summary, room temperature palladium-catalyzed direct arylations at cage B(3,4) positions in o-carboranes have been achieved with the aid of weakly coordinating, synthetically useful amides. Thus, palladium-catalyzed B–H activations enable the assembly of a wealth of arylated o-carboranes. This method features high site-selectivity, high tolerance for functional groups, and mild reaction conditions, thereby offering a platform for the design and synthesis of boron-substituted o-carboranes. Our findings offer a facile strategy for selective activations of B(3,4)–H bonds, which will be instrumental for future design of optoelectronics, nanomaterials, and boron neutron capture therapy agents.  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