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 93 毫秒
1.
Correction for ‘Highly selective acid-catalyzed olefin isomerization of limonene to terpinolene by kinetic suppression of overreactions in a confined space of porous metal–macrocycle frameworks’ by Wei He et al., Chem. Sci., 2022, 13, 8752–8758, https://doi.org/10.1039/d2sc01561g.

The authors regret that there were errors in Fig. 2, Fig. 5 and Fig. 6 in the original article and Fig. S18 of the ESI. The stereochemistry of the chemical structural formulas for (−)-α-pinene (6) and (−)-β-pinene (7) was incorrectly reversed. The correct versions of the figures are shown below, and in the updated version of the ESI.Open in a separate windowFig. 2Metal–macrocycle framework (MMF). (a) Self-assembly of asymmetrically twisted PdII-macrocycles into (b) a porous crystal MMF (sticks model) with five enantiomeric pairs of binding pockets (surface model). (c) Previously reported site-selective adsorption of (−)-α-pinene (6) (space-filling model) on the channel surface of the MMF.1 Blue, yellow, red, or black dashed circles indicate the ceiling-, side-, bottom-, or tubular-pockets of the MMF, respectively. MMF: Pd, yellow; Cl, green; N, blue; C, grey. 6: C, pink; H, white. Hydrogen atoms attached to the MMF were omitted for clarity. Green or blue surface represents exposed Cl or N–H groups of the MMF, respectively.Open in a separate windowFig. 5Investigation of the inhibitory effects of additives on the isomerization reaction of 2 using 2-NBSA@MMF at 25 °C for 102 h.Open in a separate windowFig. 6Crystallographic study of MMFs soaked in (a) a CHCl3 solution containing 1 (1.0 M), (b) a CHCl3 solution containing 2 (1.0 M), and (c) a CH3CN solution containing 7 (1.0 M). MMF: stick model or surface model; 1 and 7: space-filling model; water and CHCl3: stick model. Red dashed circles indicate the bottom pocket of the MMF. MMF: Pd, yellow; Cl, green; N, blue; C, grey. 1: C, yellow; H, white. 7: C, pink; H, white. Water and CHCl3: O, red; H, white; C, grey; Cl, green. Hydrogen atoms attached to the MMF were omitted for clarity. Green and blue surface represents exposed Cl and N–H groups of the MMF, respectively.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

2.
Correction for ‘Hydrogen-activation mechanism of [Fe] hydrogenase revealed by multi-scale modeling’ by Arndt Robert Finkelmann et al., Chem. Sci., 2014, 5, 4474–4482, DOI: 10.1039/C4SC01605J.

The authors regret that there were minor typographical errors in two figures. In Fig. 9 and and11,11, the internuclear distances were swapped. The Fe-bound hydrogen atoms are affected, where Hp is the hydrogen atom proximal to the oxypyridine ligand and Hd is the hydrogen atom distal to the oxypyridine ligand. In Fig. 9, left panel, the distance between Hp and the oxypyridine O atom was given as 1.82 Å and the distance between Hp and the Fe atom was given as 1.7 Å. However, it should read 1.82 Å between Hp and Fe and 1.70 Å between Hp and the oxypyridine O atom. In Fig. 11, top left panel, the distance between Hp and Fe was shown to be 1.70 Å and the distance between Hd and Fe was given as 1.73 Å. However, it should read 1.73 Å between Hp and Fe and 1.70 Å between Hd and Fe. The correct versions of these figures are given below. The results and conclusions are not affected by these typographical errors.Open in a separate windowFig. 9QM/MM-optimized reactant (left) and product (right) structures of the H2 cleavage reaction for the scenario with oxypyridine ligand. Distances are given in Å.Open in a separate windowFig. 11Top row: structures of the H2 adduct for the second scenario with neutral pyridinol; the pyridinol OH can be oriented away from Fe (top left) or towards Fe (top right). Bottom row: products of H2 cleavage, with the proton transferred to the thiolate; with the hydroxyl oriented away from Fe (bottom left) and towards Fe (bottom right). Distances are given in Å; relative energies with respect to the favoured adduct are indicated in red in kcal mol−1.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

3.
Correction for ‘In situ monitoring of functional activity of extracellular matrix stiffness-dependent multidrug resistance protein 1 using scanning electrochemical microscopy’ by Shuake Kuermanbayi et al., Chem. Sci., 2022, https://doi.org/10.1039/d2sc02708a.

The authors regret that an incorrect version of Fig. 5f was included in the original article. This error does not affect the conclusions of the original article as the correct Fig. 5f also proves that there is no significant difference in the mRNA levels of MRP1 in the MCF-7 cells on the PA gels with three stiffness. The correct version of Fig. 5 is presented below.Open in a separate windowFig. 1(a and b) Immunofluorescence images and (c and d) the normalized total MRP1 intensities of (a and c) MCF cells and (b and d) MDA-MB-231 cells on the PA gels with stiffness of 2.5, 17.1 and 26.2 kPa, respectively (scale bar: 40 μm). (e) Western blot analysis of the MRP1 expressions of the MCF-7 and MDA-MB-231 cells on the PA gels with stiffness of 2.5, 17.1 and 26.2 kPa, respectively. (f and g) The relative MRP1 mRNA expressions in (f) the MCF-7 cells and (g) the MDA-MB-231 cells on the PA gels with stiffness of 2.5, 17.1 and 26.2 kPa, respectively.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

4.
Correction for ‘HCOOH disproportionation to MeOH promoted by molybdenum PNP complexes’ by Elisabetta Alberico et al., Chem. Sci., 2021, 12, 13101–13119, DOI: 10.1039/D1SC04181A.

The authors regret that in Scheme 2 of the original article, complexes 7 and 8 were drawn incorrectly. The solid-state structure of both complexes, as established by X-ray analysis, had been previously reported (7 (ref. 1) and 8 (ref. 2)). In both complexes, the PNP ligand adopts a facial tridentate coordination to molybdenum and not a meridional one, as erroneously shown in Scheme 2 of the original article. The correct ligand arrangements in the metal coordination sphere for complexes 7 and 8 are reported below in Scheme 1.Open in a separate windowScheme 1Mo–PNP complexes tested in the dehydrogenation of HCOOH.Open in a separate windowScheme 2Proposed mechanisms for HCOOH dehydrogenation (red), disproportionation (blue) and decarbonylation (green) promoted by 5. Evidence for the formation of a Mo(iv) species is based on the detection by NMR of H2 and HD following addition of DCOOD to Mo(H)n species (see Fig. SI-31).Please note that complex 8 is also shown in Scheme 4 in the proposed mechanism for HCOOH decarbonylation (green part), and in Fig. 2. In both cases, the correct structure for complex 8 is reported below in Scheme 2 and Fig. 1.Open in a separate windowFig. 1 1H and 31P{1H} NMR spectra of a toluene-d8 solution of {Mo(CH3CN)(CO)2(HN[(CH2CH2P)(CH(CH3)2)2]2} 4 in the presence of 100 equivalents of HCOOH ([Mo] 10−2 M, [HCOOH] 1 M), before (a) and after heating at 90 °C for 1 hour (b). Spectra were recorded at room temperature. Signals related to complex 5 are marked by red dots.Open in a separate windowFig. 2Molecular structure of {Mo(CO)2(CH3CN)[CH3N(CH2CH2P(CH(CH3)2)2)2]} 9. Displacement ellipsoids correspond to 30% probability. Hydrogen atoms are omitted for clarity.Furthermore, a mistake was made in the caption of Fig. 6, showing the solid-state structure of complex 9: the latter has been incorrectly described as a Mo(i)-hydride species {Mo(H)(CO)2(CH3CN)[CH3N(CH2CH2P(CH(CH3)2)2)2]}. The correct formula, in agreement with the X-ray structure, is as follows and is shown above in Fig. 2: {Mo(CO)2(CH3CN)[CH3N(CH2CH2P(CH(CH3)2)2)2]}.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

5.
Simple enough to be understood and complex enough to be revealing, cascade cyclizations of diepoxides are introduced as new tools to characterize supramolecular catalysis. Decoded product fingerprints are provided for a consistent set of substrate stereoisomers, and shown to report on chemo-, diastereo- and enantioselectivity, mechanism and even autocatalysis. Application of the new tool to representative supramolecular systems reveals, for instance, that pnictogen-bonding catalysis is not only best in breaking the Baldwin rules but also converts substrate diastereomers into completely different products. Within supramolecular capsules, new cyclic hemiacetals from House–Meinwald rearrangements are identified, and autocatalysis on anion–π catalysts is found to be independent of substrate stereochemistry. Decoded product fingerprints further support that the involved epoxide-opening polyether cascade cyclizations are directional, racemization-free, and interconnected, at least partially. The discovery of unique characteristics for all catalysts tested would not have been possible without decoded cascade cyclization fingerprints, thus validating the existence and significance of privileged platforms to elucidate supramolecular catalysis. Once decoded, cascade cyclization fingerprints are easily and broadly applicable, ready for use in the community.

Hyperresponsive XL product space identifies polyether cascade fingerprinting as an attractive tool to elucidate supramolecular catalysis, including pnictogen-bonding, capsule and anion–π catalysts.

One general expectation from supramolecular catalysis1–10 is that new ways to interact will provide new ways to transform on the molecular level. This translates to access to new reactivity and products, at best contributing to new solutions for otherwise persistent challenges in science and society. While these high expectations are attracting attention to the development of supramolecular catalysts, their systematic characterization is much less advanced. Most classical and modern benchmark reactions1,9 are limited to one mechanism and cover little product space, also concerning chemo- and stereochemistry. To maximize the comparability of supramolecular catalysts, the ideal reaction would respond to as many parameters as possible at still manageable complexity. Epoxide opening polyether cascade cyclizations10–13 promise to meet these requirements for a privileged platform to evaluate supramolecular catalysts. Charismatic in chemistry and biology, they have attracted the attention of many giants in the field.11 They afford the largest polycyclic natural products, regularly featuring more than 10 rings made in one cascade. While product diversity of longer cascades is too complex and single cyclizations are too simple, minimalist cascades from diepoxide substrates such as 1 cover large structural space at tractable complexity (Fig. 1). In substrate 1, supramolecular catalysts can activate nucleophile, electrophile and leaving group, and stabilize cationic and anionic transition states and reactive intermediates (Fig. 1a). Cyclizations can follow either the 5-exo-tet selectivity predicted by the Baldwin rules (B) or anti-Baldwin (A) 6-endo-tet selectivity, leading to the four constitutional isomers 2–5 (Fig. 1a and and2).2). They can occur with normal or reverse directionally,12 forming ring 1 or ring 2 first, respectively (Fig. 1b). They can operate with pseudo SN2, SN1, or mixed mechanisms, and can integrate contributions from autocatalysis.10,13 The stereochemistry covers cistrans isomers at epoxide 1 and synanti isomers with regard to the two epoxides (Fig. 1c). This translates to the stereochemistry of products such as 2–5 at the ring junction and the exocyclic substituents of ring 2. Besides this expected diversity, the product space of the privileged substrate 1 further expands into structures that remain to be discovered, as demonstrated with two new products reported in this study.Open in a separate windowFig. 1(a) Epoxide-opening ether cascade cyclizations from diepoxide 1 as privileged platform to elaborate on supramolecular catalysis, with indication of possible contributions from electron-donating (red) and electron-accepting catalyst motifs (blue), exo-tet Baldwin (B) or endo-tet anti-Baldwin (A) chemoselectivity, (b) normal and reverse directionality, and (c) stereochemistry in selected substrates and products.Open in a separate windowFig. 2(a) Decoded product fingerprints for selected catalysts: Color-coded pie charts for products 2 (red), 3 (yellow), 4 (green), 5 (blue), 6 and 7 (teal) obtained from stereoisomers of cis and trans substrate isomers 1 with representative supramolecular catalysts 8–10 compared to general Brønsted acid (AcOH); results for cis,anti and trans,syn isomers of 1 are calculated (from data for the other diastereomer and the mixture of diastereomers in the respective series); estimated errors ± 5%. (b) Experimental results for cis,anti-1 cyclized with catalyst 9. (c) Selected X-ray structures from the BA series (p-bromobenzoyl derivatives). (d) Structure of catalysts, with indication of selected π-basic surfaces and hydrogen-bond donors on capsule 8 assembled from monomers 11, the cyclopean σ hole of pnictogen-bonding catalyst 9, and the π-acidic surface on anion–π catalyst 10.So far, substrate 1 has been used as a mixture of stereoisomers to characterize supramolecular catalysts.13 While results were intriguing, they could not be rationalized. Overlap of different trends obscured the key information and made product fingerprints dependent on the composition of the substrate mixtures. However, the observed hyperresponsiveness of the large product space suggested that decoded product fingerprints could provide a general tool to elucidate supramolecular catalysis.To assess the possibly privileged nature of diepoxide 1 as unifying substrate for supramolecular catalysis, we decided to synthesize and evaluate the necessary stereoisomers separately. The stereoisomers cis-1 and trans-1 were prepared by oxidation of the respective silyl protected cis- and trans-olefins with m-CPBA (meta-chloroperoxybenzoic acid), followed by deprotection (Fig. 2, Schemes S1 and S2). They were obtained as roughly equimolar mixtures of syn- and anti-diastereomers (cis-1: dr 54 : 46, trans-1: dr 50 : 50). Shi epoxidation14 in place of m-CPBA afforded enantioenriched cis,syn-1 (dr 89 : 11; dr 20 : 1 after purification) and trans,anti-1 (dr 82 : 18; dr 20 : 1 after purification) accordingly with unknown absolute configuration. These four substrates were sufficient to realize the complete analysis of the system because the product fingerprints for the remaining diastereomers cis,anti-1 and trans,syn-1 could be obtained from the difference of cis,syn-1 and trans,anti-1 and the respective mixture of diastereomers cis-1 and trans-1 (Fig. 2).To decode product fingerprints from different catalysts in their respective color-coded pie charts, all individual products were isolated and the diagnostic regions of their 1H NMR spectra and chiral GC traces were assembled for direct comparison (Fig. 3). In most GC traces, the two peaks were well resolved for each pair of enantiomers, confirming access to nearly all stereochemical information. The resulting unified fingerprint of the complete system then allowed to rapidly assign products obtained from different catalysts down to the level of enantiomers. The validity of most structures was confirmed by X-ray crystallography (Fig. 2c and S78–S83). If necessary, derivatives were prepared to facilitate the growth of single crystals.Open in a separate windowFig. 3Decoded product fingerprints: Diagnostic regions of 1H NMR spectra (a and b) and chiral GC (c and d) of purified cascade cyclization products from cis (a and c) and trans (b and d) substrate isomers 1 above representative examples of mixtures produced by AcOH and 9 combined (a and c), and by 8 (b and d).With the analytics in place, product fingerprints were recorded for representative supramolecular catalysts 8–10 in comparison to general Brønsted acid catalysis (Fig. 2). In the cis series, the product mixtures obtained from different catalysts contained all four constitutional isomers expected from Baldwin and anti-Baldwin cyclizations, that is cis-(BB)-2, cis-(BA)-3, cis-(AB)-4 and cis-(AA)-5 (Fig. 2a and and3a).3a). In contrast, trans-(AB)-4 was absent in the trans series, and two new products 6 and 7 were found instead (vide infra, Fig. 2a and and3b3b).In both the cis and the trans series, general Brønsted acid catalysis with AcOH was confirmed to follow the Baldwin rules almost exclusively, affording mostly (BB)-2 (Fig. 2). In the cis series, the supramolecular capsules 8 violated the Baldwin rules significantly (Fig. 2a). Capsules 8 self-assemble from resorcinarenes 11 and water (Fig. 2d).3,4 Their internal surface offers hydrogen-bond donors and π-basic aromatic planes for catalysis within their confined interior.3,4 Unique selectivities have been reported, also for bioinspired terpene cyclizations, for instance.4 From cascade cyclization with the mixture of cis-1 diastereomers in capsules 8, cis-(BA)-3 was obtained as the main product besides the still preferred cis-(BB)-2 (Fig. 2a). The pure cis,syn-1 showed a clearly different product distribution, characterized by an increased power to violate the Baldwin rule in cycle 2, affording cis-(BA)-3 as the main product. The calculated fingerprint for the products of cis,anti-1 gave the complementary dominance of the Baldwin conformant cis-(BB)-2 instead.Differences in selectivity for the syn- and anti-diastereomers in the cis series were most spectacular with the pnictogen-bonding catalyst 9 (Fig. 2a). Pnictogen-bonding catalysis has been introduced recently5–7 for consideration as the non-covalent counterpart of Lewis acid catalysis, analogous to hydrogen-bonding catalysis as non-covalent counterpart of Brønsted acid catalysis.7 Catalyst 9 is centered around an antimony V with one deep σ hole acting as pnictogen-bond donor to initiate catalysis.7 Catalyst 9 has been shown previously to efficiently break the Baldwin rules in polyether cyclizations.7,13 In the newly devised pie chart fingerprint, orthodox cis-(BB)-2 was indeed essentially absent (Fig. 2a). The mixture of diastereomers cis-1 afforded cis-(BA)-3 and cis-(AB)-4 as main products. In sharp contrast, diastereo-pure cis,syn-1 gave mostly cis,syn-(BA)-3. As a consequence, the calculated product fingerprint of cis,anti-1 showed the highly selective formation of cis,anti-(AB)-4.Selective access to cis,anti-(AB)-4 with pnictogen-bonding catalyst 9 was remarkable because none of the other stereoisomers of (AB)-4 were observed throughout the study (Fig. 3a and and4a).4a). Exclusive access to cis,anti-(AB)-4 from cis,anti-1 was understandable considering cascade cyclization with normal directionality (Fig. 1). Namely, the endo-tet cyclization of ring 1 will afford the reactive intermediate III (Fig. 4b). From this intermediate III, the exo-tet Baldwin conformant formation of ring 2 is possibly supported by an intramolecular hydrogen bond (Fig. 4b and c, arrows), which activates the nucleophile and places an epoxide in an equatorial position.Open in a separate windowFig. 4(a) The formation of only one out of four possible (AB) isomers 4 and (b) the origin of the selectivity and products found in the anti-Baldwin series with capsule 8 and pnictogen-bonding catalyst 9, with (c) selected X-ray structures.These favorable conditions to access cis,anti-(AB)-4 from cis,anti-1 contrasted sharply with the situation with all other diastereomers. In the cis series, access to the complementary cis,syn-(AB)-4 from cis,syn-1 is disfavored although the nucleophile in the reactive intermediate IV remains possibly activated by intramolecular hydrogen bonding. However, the axial orientations of epoxide in intermediate IV and a very bulky tertiary alcohol in product 4 make this reaction unlikely.With cis,syn-(AB)-4 from cis,syn-1 unfavorable, reactive intermediate IV obtained from an anti-Baldwin cyclization of ring 1 needs an alternative solution. An obvious choice is continuation with another endo-tet anti-Baldwin cyclization for ring 2 to result in cis,syn-(AA)-5 with a more flexible cis-fused oxepane ring. This cis,syn-(AA)-5 was indeed part of the product fingerprint of cis,syn-1 cyclized with pnictogen-bonding catalyst 9 (Fig. 2a). The markedly different amounts of cis,syn-(AA)-5 and cis,anti-(AB)-4 obtained from cis,syn-1 and cis,anti-1, respectively (Fig. 2a), would then suggest that normal cascade cyclizations are interconnected, possibly concerted (Fig. 1b).The differences of the selectivity of the cascade cyclization of cis,syn-1 and cis,anti-1 with pnictogen-bonding catalyst 9 (Fig. 2a) and the importance of the implications called for the experimental validation of the calculated results for cis,anti-1. Therefore, pure diastereomer cis,anti-1 was prepared and cyclized using catalyst 9. The experimental product fingerprint was very similar to the calculated one, confirming the unique cis,anti-(AB)-4 as the main product of the reaction (Fig. 2b). This results also validated the use of calculated data to decode complex product fingerprints completely.In the trans series, pnictogen-bonding catalyst 9 again broke the Baldwin rules most efficiently (Fig. 2a). For all diastereomers, trans-(AA)-5 was observed as the main product with more than 75% yield. This exceptional selectivity was understandable considering the reactive intermediates V and VI after the endo-tet cyclization of ring 1 (Fig. 4b). Contrary to intermediates III and IV in the cis series, the methyl substituent at the ring junction is in axial position also with regard to ring 2. 1,3-Diaxial interactions of the approaching electrophile with this methyl thus hinder the formation of this ring 2 by an exo-tet cyclization. Presumably for this reason, the trans-fused bis-oxane products trans,anti-(AB)-4 and trans,syn-(AB)-4 were not observed. With Baldwin cyclizations hindered, endo-tet anti-Baldwin cyclizations occurred instead to afford the respective trans,anti-(AA)-5 and trans,syn-(AA)-5 with very high selectivity (Fig. 2a).The supramolecular capsules 8 applied to the trans series yielded two new products 6 and 7 (Fig. 2). Product 6 was identified by 2D NMR spectroscopy to be a hemiacetal cyclized on an anti-Baldwin ring 1 (Fig. S72). It exists in equilibrium with the open ketone form 12, which results in dynamic epimerization at the “anomeric center” (Fig. 2 and and4).4). Derivatization of hemiacetal 6 with aromatic hydrazines gave the respective hydrazones (Fig. S76 and S77). Product 7 was identified as an acyclic allyl alcohol extending from an anti-Baldwin ring 1 (Fig. 2 and S73–S75). Both new products might originate from intermediate VII, which is generated from substrate 1 by endo-tet cyclization of ring 1 and the opening of epoxide 2 to afford the tertiary carbocation (Fig. 4b). From intermediate VII, the formation of allyl alcohol 7 only requires a proton abstraction from one of the two adjacent methyl groups. Ketone 12 originates from the same intermediate VIIvia House-Meinwald rearrangement,15 that is a 1,2-hydride shift. Similar processes might occur with trans-diepoxide 1 to give an alternative cationic intermediate VIII, which can proceed through reverse cyclization (Fig. 1b) to give products 6 and 7. Stabilization of carbocations via cation–π interactions is a distinct feature of this type of capsules.3,4The formation of these two new products in capsule 8 could be understood considering the inaccessibility of both AB products in the trans series, i.e., trans,syn-(AB)-4 and trans,anti-(AB)-4, with the explored catalysts (Fig. 4a). As already mentioned, the anti-Baldwin cyclization from trans,anti-1 and trans,syn-1 into intermediates V and VI with ring 1 is unproblematic, whereas continuation with exo-tet Baldwin cyclization of ring 2 is hindered by an axial methyl and, compared to the cis series, missing intramolecular activation of the nucleophile (Fig. 4b). With pnictogen-bonding catalyst 9, the solution was an alternative endo-tet anti-Baldwin cyclization into the trans-fused AA products 5, as discussed above (Fig. 2a and and4b).4b). In capsule 8, this endo-tet anti-Baldwin continuation of the cascade was not favorable. The reason for this distinctive selectivity within capsule 8 remains to be explored. In contrast to the other catalysts, the capsule may be able to stabilize cation VII better due to cation–π stabilization, making this pathway accessible.While the new oxanes 6 and 7 were obtained as main products from trans,anti-1 and trans,syn-1 with similar yields, the composition of the side products differed in the respective fingerprints (Fig. 2a). Cyclization of trans,syn-1 gave trans-(BB)-2 as the main side product, while trans,anti-1 gave trans,anti-(AA)-5 as the main side product. This difference was of interest because it could support that the cascade cyclizations might be interconnected, possibly concerted, at least in the present context.While capsules 8 excelled with access to new products in the trans series and pnictogen-bonding catalysts 9 with unique AB-BA selectivity on the level of diastereomers in the cis series, anion–π catalysts gave mostly Baldwin products like general Brønsted acid catalysis, independent of the stereochemistry of substrate 1 (Fig. 2). The largest deviation from Brønsted acid catalysis occurred with cis,anti-1, which gave a substantial percentage of cis-(BA)-3 and also a small amount of cis-(AB)-4 (Fig. 2a). The same trend, but less pronounced, was noted with the complementary trans,anti-1, which produced also small amounts of trans-(BA)-3 and trans-(AA)-5, formed instead of the inaccessible trans-(AB)-4 (see above, Fig. 2a).After investigation for anion transport, anion–π interactions have been introduced to catalysis in stabilizing anionic transition states on π-acidic surfaces.8,10 Over the past decade, catalysts from hexafluorobenzene to π-stacked foldamers, fullerenes, carbon nanotubes, artificial enzymes have been applied to many reactions, including enolate, enamine, imine, Diels–Alder chemistry.8 Polyether cyclizations have been introduced as a cascade transformation that should benefit best from the delocalized nature of anion–π interactions.10 On π-acidic surfaces, polyether cyclizations were autocatalytic,10 a unique emergent property that has not been observed in the many studies with systems without anion–π interactions.11With the privileged probe for supramolecular catalysis envisioned in this study, it was thus most interesting to assess the dependence of autocatalysis on the stereochemistry of the substrate. Significant dependence was conceivable considering the different products obtained from diastereomers of cis-1 with pnictogen-bonding catalyst 9 (Fig. 2a). Kinetics of all four test substrates converted with anion–π catalyst showed autocatalytic behavior (Fig. 5a and b). Moreover, autocatalysis was nearly independent of the stereochemistry of the substrate. This absence of diastereoselective autocatalysis was consistent with the computed model for transition-state stabilization by the product, and could explain why it is so difficult to achieve asymmetric autocatalysis on anion–π catalyst 10.13 Control experiments confirmed that general Brønsted acid catalysis does not show autocatalytic behavior, independent of the stereochemistry of substrate 1 (Fig. 5c and d).Open in a separate windowFig. 5Kinetics of the conversion of cis-1 (a and c, circles), cis,syn-1 (a and c, squares), trans-1 (b and d, circles) and trans,anti-1 (b and d, squares) with (a and b) anion–π catalyst 10 (10 mol%, rt) and (c and d) AcOH (500 mol%, 40 °C) in CD2Cl2, with hypothetical intermediate IX for autocatalysis on π-acidic surfaces.Taken together, the decoding of product fingerprints for cascade cyclizations that are simple enough to be tractable and complex enough to be interesting affords a privileged platform to characterize supramolecular catalysis. It is highly responsive to as many characteristics as possible, thus reporting on as many distinct advantages of the catalytic system as possible. The minimal substrate toolbox contains cis and trans di-epoxides as mixtures of synanti diastereomers, and at least one pure diastereomer. Most pairs of enantiomers are resolved in the chiral GC fingerprints. Applied to three model catalysts in comparison to general Brønsted acid catalysis, distinct fingerprints were found for all catalysts as well as for all different diastereomers of the substrate.In the cis series, most significant selectivity was observed with pnictogen-bonding catalysts, which give the unique AB product for anti and the more frequent BA product for the syn diastereomer of the diepoxide substrate with remarkably high selectivity. In the trans series, pnictogen-bonding catalysts broke the Baldwin rules most efficiently and independent of substrate stereochemistry, while within supramolecular capsules, completely new products were formed, including an interesting House–Meinwald rearrangement leading to cyclic hemiacetals. These distinct selectivities can be understood from the nature of the reactive intermediates. Together with particularly revealing details in the decoded product fingerprints, experimental support is obtained that the cascades are interconnected, possibly concerted. In clear contrast, anion–π catalysts gave mostly Baldwin products with fingerprints similar to general Brønsted acids. However, they showed unique autocatalytic behavior, a distinct emergent property that was independent of the stereochemistry of the substrate. All these distinctive characteristics found for representative supramolecular catalysts would be missed without the availability of decoded product fingerprints.These results thus validate the existence and significance of privileged substrate systems as general chemistry tools to characterize supramolecular catalysis. Once established, decoded polyether cascade fingerprints are very easy to use, ready to serve the community. For a new supramolecular catalyst to be characterized, the decoded fingerprints will reveal unique differences compared to controls. Importantly, because the system is hyperresponsive (Fig. 1a and and2a),2a), differences will be magnified. Due to the complexity required for hyperresponsiveness, the correlation of the fingerprint with the reactivity of a new catalyst will be mostly tentative and empirical at this point. For instance, AcOH-like fingerprints should reflect activation of epoxide opening to release the intramolecular leaving group, possibly supported by activation of the nucleophile as for autocatalysis on 10 (Fig. 5, IX). Fingerprints with more or even mostly A products should correlate with increasing SN1-like behavior. However, the generation of mostly B products with AcOH implies that the activation of epoxide opening needs to be supported by stabilization of the resulting carbocation with, e.g., cation–π interactions to afford A products. With pnictogen-bonding catalyst 9, this would be meaningful on the π-basic tetrachlorocatecholate plane next to the σ hole stabilizing the alcoholate (Fig. 2d). In fingerprints with the new HM-rearrangement products, so far unique for capsules 8, the existence of carbocation intermediates is experimentally confirmed and thus presumably most relevant, due to cation–π interactions, confinement effects, or both. From here, with the system trained with more and more fingerprints, the correlation of fingerprint with mechanism of a new catalyst should become increasingly informative. Sooner or later, this will enable high-level computational simulations at high confidence,7 which in turn will enhance the information on reactivity available from fingerprints of new catalysts. According to preliminary results on the difference between pnictogen-bonding and Lewis acid catalysis7 and the elucidation of more complex supramolecular systems,16 these future perspectives are most promising.  相似文献   

6.
7.
8.
Heterogeneous Fenton-like processes are very promising methods of treating organic pollutants through the generation of reactive oxygen containing radicals. Herein, we report novel 0D–1D hybrid nanoarchitectonics (necklace-like structures) consisting of FeCo@N–C yolk–shell nanoreactors as advanced catalysts for Fenton-like reactions. Each FeCo@N–C unit possesses a yolk–shell structure like a nanoreactor, which can accelerate the diffusion of reactive oxygen species and guard the active sites of FeCo. Furthermore, all the nanoreactors are threaded along carbon fibers, providing a highway for electron transport. FeCo@N–C nano-necklaces thereby exhibit excellent performance for pollutant removal via activation of peroxymonosulfate, achieving 100% bisphenol A (k = 0.8308 min−1) degradation in 10 min with good cycling stability. The experiments and density-functional theory calculations reveal that FeCo dual sites are beneficial for activation of O–O, which is crucial for enhancing Fenton-like processes.

Novel 0D–1D hybrid nanoarchitectonics consisting of FeCo@N–C yolk–shell nanoreactors are developed for Fenton-like reaction. With the multilevel advantages of this design, FeCo@N–C nano-necklaces exhibit excellent performance for BPA removal.

Advanced oxidation processes (AOPs) are one of the most promising strategies to eliminate organic contaminants, sustainably generating reactive oxygen species (ROS) to ideally destroy all non-biodegradable, recalcitrant, toxic, or membrane-permeable organic impurities.1–4 Among these AOPs, sulfate radical (SO4˙)-based Fenton-like processes have gained increasing attention as a water treatment strategy because of the strong oxidation potential of SO4˙ (3.1 V vs. normal hydrogen electrode) at wider pH ranges. SO4˙ is mainly produced by physical or chemical methods for activation of persulfate salts, such as peroxymonosulfate (PMS) and persulfate.5–9 Over the past two decades, heterogeneous catalysis has emerged as the most effective approach to water treatment, with much effort dedicated to developing better catalysts, including transition metal-based and carbonaceous materials.10,11 Unfortunately, most metal-based catalysts suffer from leaching of toxic metal ions, which can thwart their practical application,12,13 and although carbonaceous catalysts produce no secondary pollution, their cycle performance is always depressed.14 There is therefore an urgent need to find robust catalysts with adequate activity and stability for Fenton-like processes.To achieve superior performance, an ideal Fenton-like catalyst should contain oxidants with favorably reactive centers for cleavage of peroxyl bonds (O–O), have structure optimized for target pollutant attraction, and have chainmail to protect the vulnerable active sites for long periods.15–17 Recent studies have demonstrated Co–N–C active sites prefer to activate the O–O of PMS.18 Furthermore, introducing Fe-doping into the Co–N–C system not only suppresses Co2+ leaching, but also modulates the pyrrolic-N content, which is the adsorption site for capture of bisphenol A (BPA).19 We previously discovered that Co@C yolk–shell nanoreactors could enhance the catalytic activity because of the confinement effect in the nano-spaces between the core and shell, while the carbon shell acted like a chainmail protecting the Co active sites, keeping them highly reactive after five cycles.20,21Combining different kinds of materials to generate novel hybrid material interfaces can enable the creation of new kinds of chemical and physical functionalities that do not currently exist. However, one cannot simply mix these materials in an uncontrolled manner, because the ensemble of interfaces created by random mixing tends to favour thermodynamically stable interfaces that are functionally less active. Therefore, to prepare new materials with high functionality, it is necessary to carefully control the hybridization of components in interfacial regions with nanometric or atomic precision. By further hybridization of different components e.g., zero to one dimension (0D–1D) hybrid structures, we can prepare the structure to increase not only the specific surface area but also the interfacial region between different materials.In this work, we report novel 0D–1D hybrid nanoarchitectonics (necklace-like structures) consisting of FeCo@N–C yolk–shell nanoreactors as a PMS activator for Fenton-like processes. This catalyst has multilevel advantages: (i) each FeCo@N–C unit is a well-formed yolk–shell nanoreactor, which can guarantee sufficient contact of reactants and active sites, as well as defend them for good durability; (ii) all single nanoreactors are threaded along the carbon fibers, providing a highway for electron transport; and (iii) all the carbon fibers constructed into a thin film with macroscopic structure, which overcomes the complex recyclability of powder catalysts. Benefiting from favorable composition and unique structure, the FeCo@N–C catalyst delivers excellent performance for BPA removal via activation of PMS accompanied with good stability.The synthesis processes of necklace-like nanoarchitecture containing FeCo@N–C yolk–shell nanoreactors are illustrated in Fig. 1a. First, uniform Fe–Co Prussian blue analogue (Fe–Co PBA) nanocubes with an average size of 800–900 nm (Fig. 1b) are encapsulated in polyacrylonitrile (PAN) nanofibers by electrospinning. The obtained necklace-like FeCo PBA–PAN fibers (Fig. 1c) are then pyrolyzed at 800 °C in N2 atmosphere to produce FeCo@N–C nano-necklaces. The scanning electron microscopy (SEM) image (Fig. 1d) of the FeCo@N–C shows this necklace-like morphology with its large aspect ratio, with the FeCo@N–C particles strung along the PAN-derived carbon fibers. A broken particle (Fig. 1e) shows that the FeCo@N–C has a yolk–shell architecture, which is also identified by transmission electron microscopy (TEM). Fig. 1f and g show the well-defined space between the inner yolk and outer shell, which is attributed to the volume shrinkage of the original Fe–Co PBAs. During pyrolysis, Fe–Co PBA is reduced to FeCo (inner yolk) and PAN is carbonized (outer carbon shell), resulting in the unique necklace-like nanoarchitecture.22–24 The high-resolution TEM in Fig. 1h shows a lattice fringe of 0.20 nm, which matches well with the (110) plane of FeCo alloy.25 The scanning transmission electron microscopy (STEM) image (Fig. 1i) and corresponding elemental map (Fig. 1j) indicate that FeCo nanocrystals are well distributed in the inner core with some small FeCo nanocrystals located on external carbon shells. Furthermore, the control samples of Fe@N–C and Co@N–C nano-necklaces, prepared by only replacing the Fe–Co PBA nanocubes with Fe–Fe PB and Co–Co PBA (Fig. S1), also demonstrate the versatility of this synthetic strategy. The formation of hierarchical porous structure, beneficial to the PMS transportation on the surface of catalysts, could be determined by N2 adsorption–desorption isotherms and corresponding pore volume analysis (Fig. S2 and Table S1).Open in a separate windowFig. 1(a) Preparation of FeCo@N–C necklace-like nanoarchitecture. SEM images of (b) Fe–Co PBA cubic particles and (c) the electrospun FeCo PBA–PAN fibers. (d and e) SEM, (f and g) TEM, and (h) high-resolution TEM images of FeCo@N–C nano-necklaces. (i) STEM and (j) the corresponding elemental mappings of C, N, Fe, and Co.The X-ray diffraction patterns of the as-prepared products are depicted in Fig. S3, with one prominent diffraction peak centered at 44.8° corresponding to the (110) lattice plane of FeCo alloy. All the products also have a characteristic signal at 26°, implying that graphite carbon is formed during pyrolysis. Raman spectroscopy further analyzed the crystal structures and defects of the FeCo@N–C nano-necklaces (Fig. S4), where peaks found at 1349 cm−1 and 1585 cm−1 index the disordered (D band) and graphitic carbon (G band), respectively.26 X-ray photoelectron spectroscopy investigated the composition and valence band spectra of FeCo@N–C nano-necklaces. The survey spectrum (Fig. S5a) reveals the presence of Fe (1.4%), Co (1.2%), C (86.4%), N (4.5%), and O (6.5%) in the composite. The high-resolution N 1s spectrum (Fig. S5b) exhibits broad peaks at 398.1, 401.1, and 407.4 eV, corresponding to the pyridinic-N, graphitic-N, and σ* excitation of C–N, respectively.27 The high-resolution Fe 2p spectrum (Fig. S5c) shows a broad peak at 707.4 eV, attributed to Fe0. Similarly, the 777.5 eV peak observed in the Co 2p spectrum (Fig. S5d) corresponds to Co0, implying that FeCo dual sites have formed.28 The oxidation state of these sites was investigated by 57Fe Mössbauer spectroscopy, which found a sextet in the Mössbauer spectrum of the FeCo@N–C nano-necklaces attributed to FeCo dual sites (Fig. 2a and Table S2).29 The coordination environment of the FeCo dual sites was also verified by X-ray absorption fine structure (XAFS) spectroscopy. Fig. 2b shows that the X-ray absorption near-edge structure (XANES) spectra of the Fe K-edge, which demonstrates a similar near-edge structure to that of Fe foil, illustrating that the main valence state of Fe in FeCo@N–C nano-necklaces is Fe0. Furthermore, the extended-XAFS (EXAFS) spectra (Fig. 2c) displays a peak at 1.7 Å, which is ascribed to the Fe–N bond, and a remarkable peak at approximately 2.25 Å corresponding to the metal–metal band.10,30 The Co K-edge and EXAFS spectra (Fig. S6) also confirm the presence of Co–N and the metal–metal band. These results provide a potential structure of the FeCo dual sites in the FeCo@N–C nano-necklaces, as illustrated in Fig. 2d.Open in a separate windowFig. 2(a) 57Fe Mössbauer spectra of FeCo@N–C nano-necklaces at 298 K. (b) Fe K-edge XANES spectra of FeCo@N–C nano-necklaces and Fe foil. (c) Corresponding Fourier transformed k3-weighted of the EXAFS spectra for Fe K-edge. (d) Possible structure of the FeCo dual sites.This dual-metal center and necklace-like structure may be beneficial to enhance catalytic performance. Fig. 3a shows the Fenton-like performance for BPA degradation compared to Fe@N–C nano-necklaces, Co@N–C nano-necklaces, and FeCo@N–C particles (Fe–Co PBA directly carbonized without electrospinning). Here, the FeCo@N–C nano-necklaces display a higher catalytic performance, with BPA completely removed in 7 min. To clearly compare their catalytic behavior, the kinetics of BPA degradation was fitted by the first-order reaction. As shown in Fig. 3b, FeCo@N–C nano-necklaces exhibit the highest apparent rate constant (k = 0.83 min−1), which is approximately 6.4, 2.6, and 1.2 times that of FeCo@N–C particles, Fe@N–C nano-necklaces, and Co@N–C nano-necklaces, respectively. The significantly enhanced performance of FeCo@N–C nano-necklaces suggests that the FeCo dual sites and necklace-like nanoarchitecture are crucial. Furthermore, the concentration of BPA and PMS in the solution is higher than that in yolk–shell nanoreactor, resulting a concentration gradient which helps to accelerate the diffusion rates of reactants (Fig. 3c).31,32 For these nano-necklaces, the carbon shell acts like a chainmail protecting the FeCo active sites from attack by molecules and ions, and all the nanoreactors are threaded along the carbon fibers, providing a highway for electron transport, which is important for SO4˙ generation (SO4˙ production as eqn, HSO5 + e → SO4˙ + OH). Electrochemical impedance spectroscopy further confirms the good conductivity of the FeCo@N–C nano-necklaces (Fig. 3d). In addition, the concentration of metal-ion leaching and cycling performance (Fig. 3e and f) reveal the high reusability of FeCo@N–C nano-necklaces, with 95% BPA removal in 20 min after five cycles, which is also proved by the SEM and TEM characterization (Fig. S7). The effect of other reaction parameters on the BPA degradation, such as pH, reaction temperature, PMS or catalysts dosage, and common anions, were investigated in detail (Fig. S8–S11). All the results demonstrate that FeCo@N–C nano-necklaces deliver a better performance for PMS catalysis. In addition, the turnover frequency (TOF) value of FeCo@N–C nano-necklaces is 5.5 min−1 for BPA degradation, which is higher than many previously reported catalysts (detailed catalytic performance comparison as shown in Table S3).Open in a separate windowFig. 3(a) BPA degradation efficiency in different reaction systems and (b) the corresponding reaction rate constants. (c) Schematic illustration of PMS activation in FeCo@N–C nano-necklaces. (d) Nyquist plots of the catalysts. (e) The metal leaching in different reaction systems. (f) Cycling performance of FeCo@N–C nano-necklaces for BPA removal. Reaction conditions: [catalyst] = 0.15 g L−1, [BPA] = 20 mg L−1, [PMS] = 0.5 g L−1, T = 298 K, and initial pH = 7.0.To examine the enhanced catalytic activity, radical quenching experiments were conducted. As shown in Fig. 4a, when NaN3 is added to the reaction solution as a scavenger for 1O2, there is no significant reduction of BPA decomposition, implying that non-radicals are not the dominant reactive species. By comparison, when tert-butanol (TBA) (radical scavenger for ˙OH) is added, there is a slight (2.8%) decrease in BPA removal. However, if methanol (radical scavenger for SO4˙ and ˙OH) is added, the efficiency of BPA degradation declines by up to 59.2%, indicating that the major radicals generated from the PMS activation are SO4˙;33 the presence of these radicals is also verified by electron paramagnetic resonance (EPR) (Fig. 4b). Furthermore, the significant inhibition ratio can be observed when KI (quencher for the surface) is added, demonstrating that BPA degradation is mainly attributed to reactions with SO4˙, which is produced by a surface catalytic process.34Open in a separate windowFig. 4(a) Effects of the radical scavengers on BPA degradation. (b) EPR spectra of SO4˙ and ˙OH. (c) The energy profiles of PMS on FeCo@N–C nano-necklaces surface. (d) Optimized configurations of PMS adsorbed on FeCo@N–C nano-necklaces.Density-functional theory was applied to calculate the surface energy of PMS activation at FeCo dual sites (Fig. 4c, d and S12). The dissociation barrier of PMS into SO4˙ and OH is −2.25 eV, which is much lower than that on an Fe or Co single site, suggesting that cleavage of O–O bonds of PMS occurs more easily on FeCo dual sites. This is because FeCo dual sites provide two anchoring sites for the dissociated O atoms, leading to more efficient activation of O–O. The FeCo@N–C nano-necklaces can reduce the energy barrier of O–O bond breaking, which results in high activity for PMS activation and thus high productivity of SO4˙.  相似文献   

9.
We report computationally-guided protein engineering of monomeric streptavidin Rh(iii) artificial metalloenzyme to enhance catalysis of the enantioselective coupling of acrylamide hydroxamate esters and styrenes. Increased TON correlates with calculated distances between the Rh(iii) metal and surrounding residues, underscoring an artificial metalloenzyme''s propensity for additional control in metal-catalyzed transformations by through-space interactions.

We report computationally-guided protein engineering of monomeric streptavidin Rh(iii) artificial metalloenzyme to enhance catalysis of the enantioselective coupling of acrylamide hydroxamate esters and styrenes.

Artificial metalloenzymes (ArMs) can be made by anchoring a non-natural (metal) cofactor into a protein scaffold, with the goal of imbuing new-to-nature reactivity.1 One of the most common ArM platforms is the biotin-tetrameric(strept)avidin (biotin-tSav) system pioneered by Whitesides and Ward.2,3 These ArMs utilize high-affinity (up to KD ∼10−14 M) interactions between tSav and biotin–metal conjugates. tSav-based ArMs have appeared in an increasing number of transition-metal catalyzed transformations.4–6 In collaboration with the Ward group, we have previously described a tetrameric streptavidin (tSav) system containing a biotinylated Rh(iii) cofactor for the asymmetric synthesis of dihydroisoquinolones using benzhydroxamate esters and acrylate partners.7 Monomeric streptavidin (mSav), a streptavidin/rhizavidin hybrid designed to resist tetramerization, retains its high affinity for biotin (KD ∼10−9 M).8,9 We recently described the use of mSav as a new ArM,10 whose simpler topology encourages protein engineering via a site-directed mutagenesis approach.Traditional manipulation of a metal''s reactivity has been accomplished by modification of the electronic and steric properties of the bound ligands (Fig. 1a).11,12 For example, we have documented and parsed the impact of Cp electronics and sterics on a number of Rh(iii) catalyzed transformations, by structural changes to the ligand in the primary coordination sphere of Rh.13 On the other hand, ArMs have traditionally been used as modifiers of a metal''s steric environment largely focusing on inducing asymmetry in the bond-forming events. Less broadly appreciated is the fact that any mutations in residues proximal to the active site may also impact the metal''s electronic properties via changes to the secondary coordination sphere (Fig. 1b), with the prospect of delivering more active catalysts for a given transformation.Open in a separate windowFig. 1Methods to modify the (a) primary and (b) secondary coordination sphere of a Rh(iii) catalyst.Previously, we described a mSav·Rh(iii) catalyst and demonstrated its use in the direct enantioselective coupling of acrylamide hydroxamate esters and styrenes.10 The reaction allows rapid access to piperidines – the most common N-heterocycle found in FDA-approved pharmaceuticals.14 One of the most interesting aspects of this reaction was our observation of a 7-fold increase in turnover number (TON) by embedding the cofactor into mSav''s active site.15 It has been a long-standing goal of ArMs to not only enable new-to-nature reactivity, but also for them to achieve the stellar kinetics of a native metalloenzyme. As these systems lack the evolutionary privilege of a natural metalloenzyme, extensive mutation of the protein scaffold may be required to find the optimal environment of the metal cofactor.Predicting the effects of specific mutations can prove very challenging, as any alterations to the protein conformation and charge distribution can impact reactivity regardless of the mutation''s distance from the active site.16–19 In order to design a better mutant, we embarked on a collaborative experimental and computational study to define the role of the protein scaffold and how single point mutations affect reactivity. We identified two key residues that play a pivotal role in mSav·Rh(iii) ArM''s secondary coordination sphere, and have used this insight to design a more active mutant.For the purposes of this study, we focused on the mSav·Rh(iii) ArM-catalyzed coupling of methacrylamide with 4-methoxystyrene as our model reaction (Fig. 2a). Using a small model of the catalyst, the lowest energy pathway of this reaction''s proposed mechanism was generated (Fig. S9). The calculations were performed in Turbomole20–32 with the M06 density functional.33 Geometries were optimized with the def2-SVP basis set, and final electronic energies were calculated with the def2-TZVP basis set.34 The conductor-like screening model (COSMO)35 was used as implicit solvent with a dielectric of 80 to simulate water. These calculations predicted similar barriers for the N–H activation, the C–H activation, and the migratory insertion (differences less than 3 kcal mol−1). Isotope-exchange experiments revealed that the C–H activation step is reversible, implicating the migratory insertion step as turnover-limiting.10Open in a separate windowFig. 2(a) Model transformation. (b) Snapshot of the transition state for alkene insertion illustrating key nearby residues Y112 (red), E124 (blue), and S119 (purple). (c) Computed barrier to alkene insertion in the presence and absence of phenol and acetate (shown in blue).The Cp* moiety of the Cp*biotinRhX2 cofactor is non-covalently localized in the active site likely due to a π–π stacking interaction with Y112 (Fig. 2b). This assignment is supported by the observation that mutant Y112A leads to lower yield and enantioselectivity.10 We hypothesized that we could further manipulate both the sterics and electronics of the Cp* moiety by either directly mutating Y112 or indirectly by mutating other residues that affect the Y112-Cp* interaction.To generate a model of mSav''s protein scaffold and active-site we used QM/DMD36 – a hybrid quantum mechanics/molecular mechanics method that simulates proteins piecewise. Discrete molecular dynamics (DMD) equilibrates the entire system except for the metal and part of the substrate.36 After a trajectory of ∼0.5 ns, quantum mechanics (QM) is used to optimize the metal region plus sidechains and residues immediately surrounding it. This process is repeated, providing efficient sampling of the entire protein scaffold while treating the metal environment quantum-mechanically. For this study, the migratory insertion transition state was modeled in WT by freezing the coordinates of the rhodium atom and the two carbon atoms forming a bond. For each system, five replicate simulations were run for ∼20 ns each.Residues E124 and S119 both hydrogen bond to Y112 and are in close proximity to the RhCp* catalytic site (Fig. 2b).37 To estimate the electronic effects of these three residues on the reaction, an acetate ion, methanol molecule, and 4-methylphenol (p-cresol) molecule were added to a small catalyst model without constraints but initially positioned to mimic the sidechains of these residues (Fig. 2c). The migratory insertion energy barrier decreases by 2 kcal mol−1 with incorporation of the three residues. However, this energy barrier decreases by an additional 3 kcal mol−1 upon the deletion of the methanol molecule representing S119. Not only does this imply that these amino acid sidechains may be the primary reason for the increased activity of the protein-installed catalyst, but also suggest that a longer Y112–S119 distance is favorable, so long as no water can insert in this region and replace S119 in its H-bond with Y112. We hypothesize that the carboxylate group of E124 acts as a hydrogen bond acceptor, donating electron density to the Y112 phenol ring, which in turn donates electron density to the catalyst via π–π charge transfer. This could enhance the electron donation of the metal and decrease the energy barrier to the migratory insertion step. On the other hand, S119 acts as a hydrogen bond donor which would remove electron density from Y112 and subsequently the Rh(iii) moiety.Unfortunately, mutation of Y112 (Y112F and Y112W) results in negligible protein yields. We thus identified three flanking residues (T111, E113, H87) that may be expected to have a significant impact on Y112''s position, and one distal (T32) residue, chosen as distal mutations sometimes have significant impact (Fig. 3). Through this subset of mutants, we attempted to increase TON and establish a correlation between the Y112–Rh distance and Y112–S119 distance of the mutants and their reactivities.Open in a separate windowFig. 3Structure of mSav from two different views highlighting some of the mutated residues including their TON and enantioselectivity.We used QM/DMD to simulate a representative set of these mutants spanning a wide range of TONs measured in the experiment. The Y112–Rh and Y112–S119 distances were measured every ∼0.5 ps for every simulation. The results can be represented by a 3-dimensional plot with Y112–Rh distance on the X axis, Y112–S119 distance on the Y axis, and probability density on the Z axis (Fig. 4). We find the best correlation between TON and probability density in the conformational region where the Y112–Rh distance is the shortest and the Y112–S119 interaction is not energetically relevant.38Open in a separate windowFig. 4Three-dimensional probability distributions from select mutants by simultaneous sampling of Rh–Y112 and S119–Y112 distances. Probabilities for the outlined regions are also shown.To clarify this correlation, we calculated the probability of having a Y112–S119 distance between 3.5–6 Å and a Y112–Rh distance less than 5.65 Å. This Y112–S119 distance corresponds to negligible hydrogen bonding.39 Additionally, we constrained the small model catalyst shown in Fig. 5b (ref. 40) and calculated the corresponding energy barriers at different Y112–Rh distances (Fig. 5a). Since rate increases exponentially as the barrier decreases,35 differences in probabilities in the region where the Y112–Rh distance is between 5.4–5.65 Å have the greatest impact on the relative TONs of our model methacrylamide styrene coupling. We conclude that mutants with increasing probability in this region provide increasing TON.Open in a separate windowFig. 5a) Theoretical dependence of migratory insertion barrier on Rh-phenol distance. (b) Small-model catalyst.Theoretically, a shorter Y112–Rh distance relative to WT would result in increased reactivity. Residue G49 is located under the Rh(iii) moiety (Fig. 6). We hypothesize that by mutating the glycine into an alanine, steric congestion would force the biotinylated Rh(iii) cofactor to shift upwards closer to the electron donating phenol side chain of residue Y112. Analyzing the critical portions of the Y112–Rh and Y112–S119 distances in tandem reveals that G49A has the highest probability density in this region (Fig. 4). Indeed, experimentally, this mutant gives 97 TON and 91% ee (Fig. 6). The combination of a short Y112–Rh distance and long Y112–S119 distance leads to an increase in reactivity. This is an approximate 3-fold improvement in the TON relative to WT. The G49A mutant serves as an experimental proof of concept that a computational analysis of an ArMs secondary coordination sphere can lead to the design of a more efficient ArM.Open in a separate windowFig. 6Snapshot of the transition state for alkene insertion highlighting the position of G49 (purple) relative to Rh. Y112 is shown in red and E124 is shown in blue.In summary, we have identified three key residues that contribute to accelerating the rate of a Rh(iii)-catalyzed reaction by electronic communication to the metal via the secondary coordination sphere. E124 hydrogen bonds to Y112 transferring electron density via π–π charge transfer, an effect that is attenuated by hydrogen bonding from S119. Optimal interaction of these residues can be described computationally by finding mutants that have multiple conformations bearing short Y112–Rh distances coupled with negligible bonding between Y112 and S119. This hypothesis was experimentally verified by a mutant that enforces a closer Y112–Rh distance leading to improved TON. This result demonstrates the use of a hypothesis-based site-directed mutagenesis of the secondary sphere residues, to optimize the metal''s electronic environment within the protein scaffold and enhance an ArM''s activity.  相似文献   

10.
The organized assembly of nanoparticles into complex macroarchitectures opens up a promising pathway to create functional materials. Here, we demonstrate a scalable strategy to fabricate macroarchitectures with high compressibility and elasticity from hollow particle-based carbon nanofibers. This strategy causes zeolitic imidazolate framework (ZIF-8)-polyacrylonitrile nanofibers to assemble into centimetre-sized aerogels (ZIF-8/NFAs) with expected shapes and tunable functions on a large scale. On further carbonization of ZIF-8/NFAs, ZIF-8 nanoparticles are transformed into a hollow structure to form the carbon nanofiber aerogels (CNFAs). The resulting CNFAs integrate the properties of zero-dimensional hollow structures, one-dimensional nanofibers, and three-dimensional carbon aerogels, and exhibit a low density of 7.32 mg cm−3, high mechanical strength (rapid recovery from 80% strain), outstanding adsorption capacity, and excellent photo-thermal conversion potential. These results provide a platform for the future development of macroarchitectured assemblies from nanometres to centimetres and facilitate the design of multifunctional materials.

A scalable strategy is established to generate macroarchitectures based on MOF-related nanofibers. The modular assembly of macroarchitectures with luffa-like structures exhibits high mechanical strength and low densities.

The assembly of simple nanoparticles (such as silica, polystyrene and metal–organic frameworks) into macroarchitectures has a unique attraction for engineering materials due to their variable sizes, shapes, and chemical and physical properties.1–3 As a novel nanomaterial, the formed macroarchitecture with three-dimensional (3D) porous interconnected network structures has broad application prospects in various fields, including environment treatment, chemical sensing, energy storage, catalysis, and advanced electronic devices.4–7 Moreover, the functions of macroarchitectures are mainly determined by the fundamental building blocks. On account of large surface area, high porosity and more exposure to active sites, the complex macroarchitectures, which are assembled by building blocks with hollow structures, possess greater advantages.8,9In the past few decades, in order to seek high-performance hollow building blocks for macroarchitectures, much effort has been put into it. In particular, the emergence of one-dimensional (1D) carbon hollow nanostructures, including hollow porous carbon nanofibers (HPCNs)10 and carbon nanotubes (CNTs)11 promotes the rapid development of this field. The HPCN- and CNT-based macroarchitectures realize the transformation from 1D nanomaterials to three-dimensional (3D) macroscopic materials with excellent properties (e.g., electrochemical energy storage and antimicrobial air filtration).12,13 Furthermore, these macroarchitectures can not only retain the characteristics of the 1D material, but also generate many new kinds of features (e.g., high specific surface area, high mechanical strength, and low density) that the components do not possess. Unfortunately, though the HPCN- and CNT-based macroarchitectures exhibit improved conductivity and stability properties, the synthesis of their building blocks is usually expensive and complex.14 Besides, the assembly of building blocks into 3D macroarchitectures usually exhibits relatively poor mechanical properties and requires some adhesives or templates, which have to be eliminated by extra strategies.15 These complex synthetic procedures and less favourable structural stability largely hinder the scale-up production of carbon aerogels and their practical applications.To address these issues, we explore a novel and scalable method to synthesize functional macroarchitectures with robust mechanical properties fabricated from MOF-derived carbon nanofibers through manipulating nano-sized particles (MOFs) and micron-sized fibers. First, zeolitic imidazolate framework (ZIF-8) nanoparticles, which are spectacular for their large nitrogen content and surface area,16–18 are embedded into polyacrylonitrile (PAN)/polyvinylpyrrolidone (PVP) nanofibers to form ZIF-8-PAN/PVP composite nanofibers. Subsequently, the ZIF-8-PAN/PVP nanofibers are assembled into centimetre-sized nanofiber aerogels (ZIF-8/NFAs) by a freeze-drying technique. After preoxidation and carbonization of ZIF-8/NFAs, the carbon nanofiber aerogels with hollow and porous interlayer structures are fabricated (named C-ZIF-8-CNFAs). The interlayer structure of C-ZIF-8-CNFAs is very similar to natural luffa consisting of a network of elastic frameworks. As with the interconnected nanofibers in the interlayer structure, cellulose skeletons are in the interior of luffa interconnected in a highly uniform manner to maximize strength, a porous structure. Therefore, we refer to our porous structure as luffa-like. The prepared C-ZIF-8-CNFAs exhibit a low density of 7.32 mg cm−3, high specific surface area (288 m2 g−1), large hierarchical pore volume (0.22 cm−3 g−1), high mechanical strength (rapid recovery from 80% strain), outstanding adsorption capacity, and excellent photo-thermal conversion potential. Fig. 1a depicts the assembly strategy of macroarchitectures (C-ZIF-8-CNFAs) schematically. The preparation process begins with the fabrication of ZIF-8/nanofibers (Fig. S1b) using the method that ZIF-8 nanoparticles (Fig. S1a) are incorporated into nanofibers by electrospinning (specific preparation methods in the ESI). Upon homogenization in a mixed solution of ultrapure water and tert-butanol, the ZIF-8/nanofibers become wrapped around each other and dispersed uniformly. Subsequently, the homogenized nanofiber dispersion is frozen in a mold followed by freeze-drying into uncrosslinked ZIF-8/NFAs. To build further robust bonding among nanofibers, the obtained uncrosslinked ZIF-8/NFAs are preoxidized at 250 °C to form crosslinked ZIF-8/NFAs with a welding structure under the action of polyvinylpyrrolidone (PVP), providing elastic resilience to the resultant NFAs. Ultimately, the resulting preoxidized ZIF-8/NFAs are carbonized at 900 °C to form hollow C-ZIF-8-CNFAs under a N2 atmosphere. As a result of the carbonization, the organic ligands of ZIF-8 and molecules of PAN and PVP are decomposed and transformed into N-doped carbon materials. In addition, the Zn2+ in ZIF-8 nanoparticles is reduced to metallic Zn and then evaporated at high temperature.19–21 Nanoscale 0D MOFs are assembled into micron-scale 1D fibers, which are then assembled into centimeter-scale 3D carbon aerogels (Fig. 1b). This novel approach enables the super assembly on a multi-dimensional scale, which realizes the macroscopic application of nanoparticles and the functionalization of CAs. In order to verify that the performance of C-ZIF-8-CNFAs is improved after the introduction of ZIF-8, pure CNFAs (Fig. S2) without ZIF-8 (the synthetic process is shown in the ESI) are also prepared.Open in a separate windowFig. 1Preparation steps for C-ZIF-8-CNFAs. (a) More fabrication details for C-ZIF-8-CNFAs. (b) Schematic illustration of the fabrication of CNFAs.As illustrated in Fig. 2a, when the ZIF-8/nanofiber dispersion solution is frozen, the uniformly dispersed ZIF-8/nanofibers are extruded by the growth of ice crystals and assembled among ice crystals. After the sample is frozen completely, the nanofibers become lapped and locked into a 3D nanofibrous network. Subsequently, the ZIF-8/NFAs with a luffa-like structure are obtained after the sublimation of ice crystals through the freeze-drying process.22 Moreover, ZIF-8/NFAs can be made into diverse desired shapes such as cylinders, cubes, moon-like shapes, star-like shapes, heart-like shapes and intricate shapes of the letters (Fig. 2b). Fig. 2c shows the obvious reduction of the intensity of ZIF-8/NFA XRD patterns compared to the original ZIF-8, but the site hardly changes, which confirms that the introduced ZIF-8 nanoparticles are not destroyed during the electrospinning and the preparation process of aerogels. After the preoxidation and carbonization steps, the typical C-ZIF-8-CNFAs with an ultra-low density of 7.32 mg cm−3 can freely stand on the tip of a red maple leaf (Fig. 2b). The scanning electron microscopy (SEM) images in Fig. 2d–f show that C-ZIF-8-CNFAs have the hierarchical porous luffa-like structure with three kinds of pores (the picture of an actual luffa shown in the inset of Fig. 2d). The porous structure exhibits obvious rectangular pores of ∼25 μm, and the wall of these pores is made of interconnecting nanofibers (Fig. 2e). Meanwhile, the secondary pores of ∼1.5 μm are formed by the welded nanofibers that are interconnected with each other. The nanoscale pores of ∼200 nm also exist in these nanofibers, which come from the carbonization of ZIF-8 nanoparticles (Fig. 2f). As observed from the magnified SEM image (Fig. S3), the welded structure resulting from preoxdiation is still preserved through carbonization. Remarkably, ZIF-8 nanoparticles encapsulated by PAN and PVP are transformed into a hollow structure after carbonization. As can be seen from the TEM image (Fig. 2g), the ZIF-8-derived hollow structure is evenly dispersed in PAN/PVP-derived carbon nanofibers. Because Zn2+ ions coordinate with –C Created by potrace 1.16, written by Peter Selinger 2001-2019 N groups existing on the surface of PAN/PVP nanofibers, the ZIF-8 particles become tightly encapsulated by the PAN/PVP layers. During carbonization, the PAN/PVP layers make ZIF-8 shrink from inside to outside, thereby leading to the generation of a hollow structure. Moreover, the confined carbonization process within the PAN/PVP matrix prevents the irreversible fusion and aggregation of carbonized ZIF-8 nanoparticles.19Open in a separate windowFig. 2(a) Schematical illustration of the formation principles for the hierarchical cellular structure. (b) Photographs of ZIF-8/NFAs with diverse shapes and the lightweight C-ZIF-8-CNFAs standing on the tip of a red maple leaf. (c) Wide-angle XRD patterns. (d–g) SEM (d–f) and TEM (g) images showing the microstructure of C-ZIF-8-CNFAs at various magnifications. (h and i) N2 adsorption–desorption isotherm and pore-size distribution curve of CNFAs and C-ZIF-8-CNFAs.After ZIF-8 nanoparticles were introduced into CNFAs, the properties have been improved significantly. To comprehend the variation of the porous characteristics in C-ZIF-8-CNFAs, nitrogen (N2) adsorption–desorption measurements were carried out. As observed from Fig. 2h and i, C-ZIF-8-CNFAs have a larger specific surface area of 288.3 m2 g−1 and pore volume of 0.22 cm−3 g−1, while the specific surface area and pore volume of CNFAs are only 12.1 m2 g−1 and 0.01 cm−3 g−1, respectively. Meanwhile, C-ZIF-8-CNFAs also have a hierarchical porous structure with micropores, mesopores, and macropores (Fig. 2i). Because of the existence of the hierarchical porous structure in C-ZIF-8-CNFAs, they have a lower density relative to CNFAs (20.73 mg cm−3) (Table S1). The chemical compositions and graphitic structure of C-ZIF-8-CNFAs are investigated by X-ray powder diffraction (XRD) and X-ray photoelectron spectroscopy (XPS). The XRD pattern (Fig. S4a) of C-ZIF-8-CNFAs only exhibits two broad peaks at about 25° and 44°, corresponding to the (002) and (101) diffraction facets of the graphitic structure, respectively.23 The correlative XPS spectrum shows that C-ZIF-8-CNFAs consist of C, N and O (Fig. S4b). The high-resolution N 1s spectra (Fig. S4c) can be deconvoluted into four peaks: pyridinic N (398.6 eV), pyrrolic N (399.4 eV), graphitic N (400.9 eV), and oxidized N (403.7 eV), respectively. The percentage of nitrogen and its types are listed in Table S2. There is no Zn content thus indicating that it had evaporated during the carbonization process at high temperature (at 900 °C for 3 hours).24 Furthermore the N content of C-ZIF-8-CNFAs increases with the introduction of ZIF-8 compared to CNFAs clearly (Table S3).In stark contrast to the hard and brittle characteristics of conventional carbon aerogels, C-ZIF-8-CNFAs show robust mechanical properties, sustaining large compressive strain without fracture (insets in Fig. 3a and movie S1). The compressive stress–strain (σε) curves (Fig. 3a) show the compressive process of C-ZIF-8-CNFAs and two typical deformation regimes could be recognized: a Hookean or linear elastic regime of ε < 50% with a stable tangent modulus, and a densification regime of ε > 50% with σ and dσ/dε increasing sharply. When the maximum compressive strain increases from 30% to 80%, the maximum compressive stress steeply increases from 1.2 to 25.9 kPa, indicating that C-ZIF-8-CNFAs can bear over 25 000 times their own weight without cracking. Moreover, the cyclic compression test of the C-ZIF-8-CNFAs is performed to validate their durable cycling performance by applying 50 loading–unloading fatigue cycles at a large ε of 50% (Fig. 3b). C-ZIF-8-CNFAs hardly undergo plastic deformation at all after the 50th cycle, which is a huge improvement over traditional CAs with a crisp character. As shown in Fig. 3c, C-ZIF-8-CNFAs retain nearly 100% of the initial value of the Young''s modulus, maximum stress and energy loss coefficient, indicating that their strength or stiffness has no significant decline highlighting their exceptional structural robustness. Two reasons could account for the excellent mechanical elasticity of C-ZIF-8-CNFAs. On the one hand, the compressive strain is absorbed by the bending of the connected ZIF-8/nanofibers between layers; on the other hand, the strain that continues to grow (beyond 50%) is absorbed by the densification of minor pores, which is formed by the welding of PVP (Fig. 3d).25 The two reasons can also be confirmed by the sharp increase of the compressive stress after ε > 50% (Fig. 3a).Open in a separate windowFig. 3(a) Compressive stress–strain curve of C-ZIF-8-CNFAs at different strains. The insets are photographs of C-ZIF-8-CNFAs under a compressing and releasing cycle (ε = 80%). (b) A 50-cycle compressive fatigue test with ε of 50%. (c) The Young''s modulus, energy loss coefficient, and maximum stress versus compressive cycles. (d) Sketch of the changes in the hierarchical porous structure with compressive deformation.MOF-based derived CNFAs with hollow structures, ultra-low density, extraordinary mechanical elasticity, and large surface area are conducive to applications in various fields, including environment governance, energy absorption, and energy storage.26–28 As a proof of concept, we evaluated the organic solvent absorption and photo-thermal conversion performance of C-ZIF-8-CNFAs. Excellent hydrophobicity is one of the important factors to ensure the absorption efficiency of organic solvents. As demonstrated in Fig. 4a, C-ZIF-8-CNFAs exhibit high hydrophobicity with a water contact angle of 142° and the water contact angle has no significant change after 120 s. The water droplet and the absorbed oil droplet are on the surface of C-ZIF-8-CNFAs, indicating the effective oil/water selectivity of C-ZIF-8-CNFAs (Fig. S5). As shown in Fig. 4b, C-ZIF-8-CNFAs demonstrate extraordinary absorption capacities for common oils and various organic solvents, 90–200 times their own weight, principally depending on the density and viscosity of the solvents. For example, the adsorption capacity for carbon tetrachloride (ρ = 1.595 g cm−3) is much higher than the adsorption capacity for n-hexane (ρ = 0.66 g cm−3). We also compared the maximum adsorption capacity of C-ZIF-8-CNFAs and CNFAs for several common organic solvents and oils. The adsorption performance of C-ZIF-8-CNFAs is also demonstrated to be better than that of CNFAs (Fig. S6). According to the cross-sectional diagram of the adsorption of organic solvents (Fig. 4c), CNFAs only absorb organic solvents through the capillary phenomenon of the channel, which is formed by nanofibers, while the absorbed organic solvent can also enter into the hollow structure formed by carbonization of ZIF-8 inside the nanofiber of C-ZIF-8-CNFAs. Therefore, the performance of C-ZIF-8-CNFAs is significantly improved after the addition of ZIF-8 nanoparticles. Moreover, the adsorption capacity of C-ZIF-8-CNFAs for organic solvents is greater than that of previously reported aerogels (Fig. S7). To verify the excellent adsorption performance of C-ZIF-8-CNFAs in practical application, the oil/water separation and cyclic adsorption experiments were carried out. As illustrated in Fig. S8, C-ZIF-8-CNFAs can quickly absorb heavy organic solvents such as carbon tetrachloride (dyed with oil red) sunk at the bottom of water, thus indicating their potential application for selectively removing oils from water. Recyclability and reusability are also crucial to evaluate the practical application possibility of adsorption materials. In view of the outstanding elasticity and structural robustness of C-ZIF-8-CNFAs, we chose ethanol as the absorption solvent for recycling tests. Through simple heating, ethanol absorbed by C-ZIF-8-CNFAs can be readily removed. As demonstrated in Fig. S9a and b, even after 10 cycles, C-ZIF-8-CNFAs still retained over 90% adsorption capacity for ethanol and their mass was reduced by less than 10%.Open in a separate windowFig. 4(a) Dynamic behaviors of a water droplet on the surface of C-ZIF-8-CNFAs. (b) Adsorption efficiency of C-ZIF-8-CNFAs towards commonly used organic solvents and oils. (c) Schematic illustration of the organic solvent adsorption process inside the CNFAs and C-ZIF-8-CNFAs. (d) Mass changes of evaporated water versus time under 1 sun illumination. (e) The temperature of C-ZIF-8-CNFAs and seawater under 1 sun illumination as a function of irradiation time. (f) The infrared images (IR) show the temperature distribution of C-ZIF-8-CNFAs and seawater under 1 sun illumination with an irradiation time of 0, 30, and 60 min.Given the abundant porous structure and blackbody characteristic, C-ZIF-8-CNFAs are also promising materials for interfacial solar steam generation (ISSG). To investigate the ISSG performance, the evaporation mass change of seawater and C-ZIF-8-CNFAs is measured under one sun illumination. As shown in Fig. 4d, the seawater in C-ZIF-8-CNFAs achieves the maximum evaporation rate of 3.74 kg m−2 h−1, which is 5.12 times the evaporation rate of bulk seawater (0.73 kg m−2 h−1) and 2.34 times faster than the evaporation rate of general 2D ISSG (∼1.6 kg m−2 h−1). An infrared camera is used to trace the surface temperatures of C-ZIF-8-CNFAs and bulk seawater under one-sun illumination to evaluate the photothermal behavior of ISSG (Fig. 4e). The surface temperature of C-ZIF-8-CNFAs presents a quick increase in 5 min and eventually reaches a stable state (∼48 °C) after 10 min, while the bulk seawater temperature stays unchanged (∼25 °C). The consecutive infrared images in Fig. 4f show the equilibrium temperature distribution and heat localization effect of C-ZIF-8-CNFAs and bulk seawater in 60 min under one-sun illumination.  相似文献   

11.
A molecular keypad lock that displays photodynamic activity when exposed to glutathione (GSH), esterase and light in the given order, is fabricated and its efficacy in drug resistant MCF7 cancer cells is investigated. The first two inputs are common drug resistant tumor markers. GSH reacts with the agent and shifts the absorption wavelength. Esterase separates the quencher from the structure, further activating the agent. After these sequential exposures, the molecular keypad lock is exposed to light and produces cytotoxic singlet oxygen. Among many possible combinations, only one ‘key’ can activate the agent, and initiate a photodynamic response. Paclitaxel resistant MCF7 cells are selectively killed. This work presents the first ever biological application of small molecular keypad locks.

Information processing therapeutics with an implemented keypad lock logic gate selects input order for activation in drug resistant cancer cells.

The complex nature of diseases such as cancer necessitates smarter drugs that can discriminate each disease state or regulate drug efficacy spatially and/or temporally. With this intention, activatable drugs, drugs with on demand release properties are developed with promising selectivity.1–4 Information processing therapeutics which are based on molecular logic gate operations are another approach to solve this problem.5–7 Molecular logic gates are small compounds using Boolean logic operations to process inputs (i.e. the analyte concentration), and give an output as a result (fluorescence, and therapeutic activity etc.).8 Selective drug activation, release, multiple-analyte sensing and theranostic applications of these devices have been explored by us and others.5,9–19Among the operations that can be carried out using small molecules, keypad locks provide an alternative application in information security.20 This logic operation can give a specific output when the inputs are given in the correct form and correct sequence. For the device, each input is considered as an AND logic operation where the history of the process is also considered. A pioneering example was reported by Margulies and Shanzer in 2007 where energy transfer is modulated by chelation of Fe3+ in a pH dependent manner.21 Later, various other devices were introduced with advanced properties such as more than 2 input responsiveness and error detection capability.22–24 All-photonic logic gates to address chemical waste production is extensively studied by Gust, Andréasson and Pischel.25,26 Beside small molecule keypad locks, enzymes, antibodies, and DNA hybrids are used to achieve the same goal.27–30 Although their potential use in molecular cryptology is highlighted, so far, there is no solid biological application of small molecule keypad locks.In the research presented here, a molecular keypad lock is developed which displays a photodynamic therapeutic output when a molecule is exposed to analytes in the correct order and type (PS3, Fig. 1). Two inputs of the system are chosen to be the common markers of drug resistant tumours: glutathione (GSH) and esterase enzyme (E). Cancer cells develop resistance to traditional chemotherapy in time by changing the protein expression or metabolite content of the cell. This adaptation of cancer cells is an obstacle for their treatment and needs to be addressed. Glutathione is a tripeptide used in reductive biochemical synthesis and it is known to be present in elevated levels in rapidly dividing cells such as cancer cells.31 A high GSH level is reported to contribute to drug resistance, since GSH adducts of the drugs are exported out of the cell much more rapidly.32,33 Likewise, esterase enzyme activity is known to be associated with drug detoxification as this enzyme contributes to the chemical conversion of the drug.34,35 Glutathione and esterase enzyme are chosen to be the first two inputs of the molecular keypad lock, the first two digits of the password. In the research, light is used as the final input. Although trivial, light is essential for photodynamic activity and spatiotemporal control of irradiation, further improving selectivity of the therapy.Open in a separate windowFig. 1Chemical structures of model photosensitizers (PS1 and PS2) and a molecular keypad lock (PS3). Ester bonds (red) are prone to hydrolysis by the esterase enzyme. Distyryl sites of the photosensitizers (blue) can react with thiol nucleophile provided that it is bound to an electron deficient group (i.e. pyri-dinium).Keypad lock PS3 is a photodynamic therapy (PDT) agent. PDT is a non-invasive method used for the treatment of surface cancers and certain other diseases ranging from atherosclerosis to macular degeneration.36–39 In this therapy, a photosensitizer is excited with light, and produces cytotoxic singlet oxygen (1O2) thereby triggering apoptosis or necrosis of the cell, initiating an immune response and blocking microvasculature.40 In the research, a boradiazaindecene (BODIPY) photosensitizer is used to benefit from versatile chemistry and spectroscopic properties.41–45Near-IR absorbing PS3 shown in Fig. 1 is the molecular keypad lock and it is synthesized in 13 steps (Scheme S1). PS3 and model compound PS2 have heavy atoms on the structure to favour intersystem crossing required for transition to the triplet state and hence 1O2 generation occurs.43 Ester bonds on the structure of PS3 are prone to cleavage by esterase enzyme. Distryryl bonds on the PS3 (blue) tend to reduce or form an adduct with thiol nucleophiles when it is activated by the pyridinium electron withdrawing group.46 This property lies at the heart of sequential operation of esterase and GSH. When GSH reacts with electron poor double bonds, the extended conjugated structure is broken and PS3-a is generated (Fig. 2). This structure has absorption below 550 nm, like brominated core BODIPY molecules (compound 8, Scheme S1), and therefore can be excited with a green light. A quencher (green) is attached to ensure that photodynamic activity is OFF until esterase cleaves the ester bond. This is because of the energy transfer from the photosensitizer to this module, until esterase separates the photosensitizer. Since PS3 lacks absorption around the 500–550 nm region, it is inactive until GSH reacts with the compound. However, the GSH reacted photosensitizer does show absorption in this region; so, in order to avoid full activation just by GSH, a quencher module is attached. Spectral overlap between the BODIPY core (see the structure of compound 8 in the ESI, similar to that of PS3-a in terms of conjugation) and quencher (Q) can be seen from UV-Vis absorption and fluorescence spectra (Fig. 3 and S1). By this way, the photosensitizer is chemically modulated by GSH to ensure excitation, and then esterase enzyme inhibits energy transfer by removing the quencher. Lastly a green light is used to excite the photosensitizer leading to generation of photodynamic action. Since light is necessary for the final excitation of the molecule, it should always be the last input. If the order of esterase and GSH changes, as shown in Fig. 2, activation is not expected to take place since cleavage of the ester bonds generates 4-hydroxybenzyl derivative on PS3, which spontaneously faces 1,4-elimination to generate pyridine (Fig. S2).47 Pyridine on its own is not sufficiently electron withdrawing to favour nucleophilic attack of double bonds by GSH and to activate it as demonstrated below. Therefore, the photosensitizer preserves extended conjugation and essentially lacks absorption at the wavelength of excitation.Open in a separate windowFig. 2Sequential operation of GSH and esterase. GSH can only react with BODIPY distyryl units when the structure has electron withdrawing pyridinium, either reducing it or forming an adduct. Esterase enzyme cleaves ester bonds, liberating the photosensitizer from the quencher module (green). Initial esterase activity converts the pyridinium unit to pyridine, thereby decreasing the reactivity of double bonds with GSH.Open in a separate windowFig. 3Normalized UV-Vis absorption and fluorescence spectra of PS1–3 in 2% water in THF (a and b). Samples are excited at 600 nm. Spectral changes of PS3 (10 μM) alone (black) or PS3 upon exposure to 0.5 mM GSH (c) and 10U esterase (d) for 90 min and 60 min at 37 °C, in 2% water in THF, respectively. A new peak at 544 nm appears upon incubation with GSH which is attributed to reduced PS3 and/orthe GSH-adduct. Esterase treatment increases the relative intensity of the shoulder peak around 600 nm.In order to understand the response of the PS3 to GSH, a molecule is incubated with 0.5 mM of GSH at 37 °C for 90 min. A new peak at 544 nm appears in UV-Vis absorption spectra consistent with the hypothesis (Fig. 3c, S1 and S9). The formation of the GSH adduct (PS3-a) is demonstrated by Liquid Chromatography Mass spectrometry analysis (Fig. S3). When control module PS1 is exposed to the same conditions, this new peak is not detected indicating that the pyridine bearing structure is neither activated enough for the nucleophilic substitution by GSH nor did it display PDT activity (Fig. S4 and S5). On the other hand, GSH treated pyridinium bearing PS2 immediately displayed a colour change indicative of broken conjugation (Fig. S6). When PS3 is incubated with esterase for 1 h, a small hypsochromic shift in the absorption peak is detected as a shoulder to the parent peak which is attributed to the conversion of pyridinium to pyridine (PS3-c, Fig. 3d). The control PS3 sample which is incubated under the same conditions but lacks esterase does not show an enhancement of this peak (Fig. 3d, black). High Resolution Mass Spectrometry analysis of the esterase treated PS2 samples confirm the hydrolysis of the ester and subsequent formation of the pyridine compound (Fig. S7). Esterase treated samples display an increase in the emission intensity when excited at 620 nm (Fig. S8). This is attributed to the initial quenching of the quencher module by the pyridinium photosensitizer. Analysis of the absorption and emission spectra suggest that the quencher module of PS3 can induce energy transfer to the pyridinium photosensitizer (Fig. 3). Once separated by esterase, fluorescence of the quencher module increases. In the case of GSH treated sample, a small enhancement in emission upon excitation at 500 nm is observed (Fig. S9). Note that the GSH adduct (or PS3 with reduced double bonds) has higher absorption at this wavelength, which would be the reason for the increase in emission intensity. In the spectral analysis organic solvents with a low water content are used to monitor the formation of water-insoluble, neutral, pyridine-bearing intermediate species.In the project, the molecular keypad lock is aimed to unlock in the presence of drug resistant tumour markers and get activated. Activation cannot take place when the input order differs. To demonstrate this, photodynamic action in the presence of all three inputs in a different order is investigated. 1O2 production can be followed by using trap molecule, 1,3-diphenylisobenzofuran (DPBF).48 This molecule reacts with 1O2 and loses its absorption at 418 nm. The effect of different input combinations on the PDT action are given in Fig. 4. In the first 15 min, all samples are kept in the dark. Under such conditions no 1O2 generation is detected, which indicates lack of dark activity. DPBF is exposed to light from a LED source (peak 505 nm) under the same experimental conditions and no decrease in the absorption is detected. This control experiment eliminates the photodegradation of DPBF in the absence of a photosensitizer. Upon irradiation before the activation of the photosensitizer by GSH and esterase, no 1O2 generation is observed as expected. The results show that 1O2 generation, and the subsequent decrease in DPBF absorption, are significantly more in the input order of glutathione, esterase enzyme and light, consistent with the proposed mode of activation.Open in a separate windowFig. 4 1O2 generation ability of PS3 (0.1 μM) when three inputs are given in a different order. All samples contain 50 μM of 1O2 trap molecule DPBF. In the first 15 minutes samples are kept in the dark. GSH is added in 0.5 mM concentration and incubated for 90 min at 37 °C. Samples are incubated with 10U esterase for 1 h at 37 °C. An LED light is irradiated from a 30 cm distance for 45 min.To analyse the effect of PDT action in the cell, a drug resistant cell line is generated. MCF7 cells are exposed to an increased dose of traditional cancer therapeutic agent paclitaxel as described in the literature.49 When the spindle-shaped morphology is obtained following maximum drug dose application, cells are reported to have drug resistance. At this stage, PS3 is applied to both normal and drug resistant cells. When cell viabilities at various concentrations are analysed, it has been found that the light toxicity of PS3 is significantly enhanced in drug resistant cells (Fig. 5). The IC50 values of irradiated samples are calculated to be 124.8 μM for MCF7 cells. This value is reduced to 52.5 μM in paclitaxel resistant MCF7 (Pac-MCF7) indicating improved cytotoxicity in these cells. Efficient induction of apoptosis is also proved by Annexin V and PI staining (Fig. 6). Under dark conditions, cells do not have significant loss of viability. Upon irradiation, resistant cells are more prone to apoptosis by the photosensitizer. Relative singlet oxygen generation abilities and results of cell culture experiments altogether confirm selective activation in drug resistant cells.Open in a separate windowFig. 5Change in the cell viability of normal and paclitaxel resistant MCF7 cells (Pac-MCF7) in the presence of PS3 at various concentrations. For each group, cell viability is analysed both after incubation in the dark or after irradiation with a 505 nm LED light source from a distance of 10 cm. Average values of three independent experiments are used.Open in a separate windowFig. 6Apoptosis induction by PS3 (25 μM) in normal and paclitaxel resistant MCF7 cancer cells under dark conditions and upon irradiation with a 505 nm LED light from 10 cm distance. Scale bars: 50 μm.  相似文献   

12.
Correction for ‘Suppressing carboxylate nucleophilicity with inorganic salts enables selective electrocarboxylation without sacrificial anodes’ by Nathan Corbin et al., Chem. Sci., 2021, DOI: 10.1039/D1SC02413B.

We regret that there was a minor error in the structure of the benzyl chloride in Scheme 2, Fig. 2 and the ESI. The structure of the benzyl chloride should be 4-methyl benzyl chloride but was instead given as 3-methyl benzyl. The correct figure and scheme are shown below, and the ESI has been updated.Open in a separate windowFig. 2(A) Comparison of acid yields for non-sacrificial-anode and sacrificial-anode carboxylation of various substrates. (B) Ratio of carboxylic acid to nucleophilic side products (ester + carbonate + alcohol) for various systems and substrates. Effect of adding MgBr2 to the sacrificial-anode system on the (C) acid yield and (D) ratio of acid to SN2 side products for benzyl bromide. Acid yields are tabulated in Table S6.† ND: acid not detected (acid-to-SN2 ratio <0.1).Open in a separate windowScheme 2Substrate scope for the sacrificial-anode-free electrochemical carboxylation of organic halides. aStandard reaction conditions: 100 mM electrolyte, 100 mM substrate, 100 mM MgBr2, silver cathode, platinum anode, 20 sccm CO2, 2.2 mL DMF, −20 mA cm−2 for 3.5 h. TBA-Br was used for chlorinated substrates because bromide oxidizes more readily than chloride, and only a small amount of chloride was replaced by bromide (<1% for the alkyl chloride, ∼4% for the benzylic chloride). Yields are referenced to the initial amount of substrate and were calculated from 1H NMR spectroscopy using either 1,3,5-trimethoxybenzene or ethylene carbonate as internal standards. b−15 mA cm−2 instead of −20 mA cm−2. c150 mM MgBr2 instead of 100 mM MgBr2.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

13.
Multisubstituted pyrroles are important fragments that appear in many bioactive small molecule scaffolds. Efficient synthesis of multisubstituted pyrroles with different substituents from easily accessible starting materials is challenging. Herein, we describe a metal-free method for the preparation of pentasubstituted pyrroles and hexasubstituted pyrrolines with different substituents and a free amino group by a base-promoted cascade addition–cyclization of propargylamides or allenamides with trimethylsilyl cyanide. This method would complement previous methods and support expansion of the toolbox for the synthesis of valuable, but previously inaccessible, highly substituted pyrroles and pyrrolines. Mechanistic studies to elucidate the reaction pathway have been conducted.

This method is a toolbox for the synthesis of valuable, but previously inaccessible, highly substituted pyrroles and pyrrolines.

Pyrroles are molecules of great interest in a variety of compounds including pharmaceuticals, natural products and other materials. Pyrrole fragments for example are key motifs in bioactive natural molecules, forming the subunit of heme, chlorophyll and bile pigments, and are also found in many clinical drugs, including those in Fig. 1a.1 Although many classical methods of pyrrole synthesis, including the Paal–Knorr condensation,2 the Knorr reaction,3 the Hantzsch reaction,4 transition metal-catalyzed reactions,5 and multicomponent coupling reactions,6 have been developed over many years, the efficient synthesis of multisubstituted pyrroles is still challenging. In condensation syntheses of pyrroles, the major problems lie in the extended syntheses of complex precursors and limited substitution patterns that are allowed. Multicomponent reactions are superior when building pyrrole core structures with more substituents. Among these, the [2+2+1] cycloaddition reaction of alkynes and primary amines is attractive because of the readily available alkyne and amine substrates and the ability to construct fully substituted pyrroles.7 However, with the exception of some rare examples,8 most [2+2+1] cycloaddition reactions afford pyrroles with two or more identical substituents. The synthesis of multisubstituted pyrroles with all different substituents from simple starting materials therefore remains a major challenge.Open in a separate windowFig. 1Previous reports and this work on propargylamides transformation.Easily accessible propargylamides are classical, privileged building blocks broadly utilized for the synthesis of a large variety of heterocyclic molecules such as pyrroles, pyridines, thiazoles, oxazoles and other relevant organic frameworks.9 For example, Looper10et al. reported the synthesis of 2-aminoimidazoles from propargyl cyanamides and Eycken11 reported a method starting from propargyl guanidines which undergo a 5-exo-dig heterocyclization as shown in Fig. 1b. Subsequently, Wan12et al. revealed the cyclization of N-alkenyl propargyl sulfonamides into pyrroles via sulfonyl migration. Inspired by these transformations and multi-substituted pyrrole synthesis, we report herein a direct synthesis of pentasubstituted pyrroles and hexasubstituted pyrrolines with all different substituents from propargyl sulfonylamides and allenamides.Previously, Zhu,13 Ji14 and Qiu13b,15 reported efficient syntheses of 2-aminopyrroles from isocyanides. Ye16 and Huang17 independently developed gold-catalyzed syntheses of 2-amino-pentasubstituted pyrroles with ynamides. Despite the many advantages of these methods, they all afford protected amines, rather than free amines. The deprotection of these amines may cause problems in further transformations of the products. Our method delivers pyrroles with an unprotected free amino group and are often complementary to the previously well-developed classical methods.Initially, the cyclization reaction of N-(1,3-diphenylprop-2-yn-1-yl)-N-ethylbenzenesulfonamide (1a) with trimethylsilyl cyanide (TMSCN) was carried out with Ni(PPh3)2Cl2 as a catalyst, a base (Cs2CO3) and DMF as a solvent. Different metal catalysts, such as Ni(PPh3)2Cl2, Pd(OAc)2, Cu(OAc)2, and Co(OAc)2 provided the desired product with similar yields (
EntryCat.BaseSolventYield
1Ni(PPh3)2Cl2Cs2CO3DMF67%
2Pd(OAc)2Cs2CO3DMF65%
3Cu(OAc)2Cs2CO3DMF65%
4Co(OAc)2Cs2CO3DMF63%
5Cs2CO3DMF65%
6KFDMFTrace
7K3PO4DMFTrace
8K2CO3DMF48%
9KOHDMF52%
10KOtBuDMF46%
11Et3NDMFTrace
12Cs2CO3CH3CN18%
13Cs2CO3DME23%
14Cs2CO3TolueneTrace
15Cs2CO3DCETrace
16Cs2CO3DioxaneTrace
Open in a separate windowaReaction conditions: 1a (0.1 mmol, 1 equiv.), TMSCN (0.3 mmol, 3 equiv.), cat. (0 or 10 mol%), base (0.3 mmol, 3 equiv.) and solvent (1 mL), at 80 °C for 10 h; isolated yield.With the optimal reaction conditions in hand, we investigated the scope of this reaction. As shown in Fig. 2, the transformation tolerates a broad variety of substituted propargylamides (1). The R1 group could be an aryl group containing either electron-donating groups or electron-withdrawing groups, and the corresponding products (2b–2h) were obtained in yields of 62–80%. The substituent R1 could also be an alkyl group such as 1-hexyl in which case the reaction provided the corresponding pyrrole (2i) in 53% yield. Exploration of the R2 substituent was also conducted. Electron-rich and electron-deficient substituents in the aromatic ring of R2 gave the desired products (2j–2o) with yields of 70–81%. The product bearing a furyl group (2p) can be produced in 61% yield. However, when R2 group is an aliphatic group, the reaction failed to provide the desired product. Substituent groups R3, such as benzyl (2q) or 3,4-dimethoxyphenylethyl (2r) were also compatible in the reaction, providing the corresponding products in moderate yields. Significantly, this method has the potential to produce core structures (for example 2s) similar to that in Atorvastatin. Interestingly, when alkynyl substituted isoquinolines (1t–1v) were used as the substrates, the reactions smoothly afforded fused pyrrolo[2,1-α]isoquinoline derivatives (2t–2v), members of a class of compounds that are found widely in marine alkaloids and exhibit anticancer and antiviral activity.18Open in a separate windowFig. 2Substrate scope of propargylamides. Reaction conditions: 1 (0.20 mmol, 1 equiv.), TMSCN (0.60 mmol, 3 equiv.), Cs2CO3 (0.60 mmol, 3 equiv.) and DMF (2 mL), at 80 °C for 10 h; isolated yield.Allenes are key intermediates in the synthesis of many complex molecules.19 As a subtype of allenes, allenamines are also useful as reaction intermediates.20 Although the transformation of allenamides to multisubstituted pyrroles has not been previously recorded, this reaction probably goes through the allenamide intermediates which can be derived from propargyl sulfonamides under basic conditions. To verify this hypothesis, the trisubstituted allenamide (3) was synthesized and subjected to the standard reaction conditions. A pyrrole (2a) was isolated in 82% yield from this reaction (Fig. 3). This result confirmed our assumption and raised a new question: is it possible to build hexasubstituted pyrrolines from tetrasubstituted allenamides? A range of tetrasubstituted allenamides21 was tested under the standard reaction conditions, and the hexasubstituted pyrrolines were obtained as is shown in Fig. 4. The R1 group could be an aryl substituent or an alkyl chain, and the corresponding products (5a–5e) were obtained with good yields. Various aryl groups with either electron-donating groups or electron-withdrawing groups in the aromatic ring of R2 provided the desired products (5f–5k) in 62–83% yields. In addition, the difluoromethyl group can also be replaced by a phenyl group, and the reaction provided the corresponding product 5l in 82% yield. It is worth noting that these pyrroline products are not easily accessible from other methods.Open in a separate windowFig. 3Synthesis of substituted pyrroles from allenes.Open in a separate windowFig. 4Substrate scope of tetrasubstituted allenamides. Reaction conditions: 4 (0.10 mmol, 1 equiv.), TMSCN (0.30 mmol, 3 equiv.), K2CO3 (0.30 mmol, 3 equiv.) and DMF (1 mL), at 80 °C for 10 h, isolated yield.Some synthetic applications of this method are shown in Fig. 5. The amide is a naturally occurring and ubiquitous functional group. When using benzoyl chloride to protect the free amino group of the fully-substituted pyrrole (2a), a bis-dibenzoyl amide (6) was obtained in the presence of a base, triethylamine while the monobenzoyl protected amide (7) was obtained in the presence of pyridine as the base (Fig. 5a). This method also provides a straightforward approach to pyrrole fused lactam structures (Fig. 5b). For examples, a five-membered lactam and a six-membered lactam were generated separately in a one pot reaction, directly from, (8 and 10), respectively. Taking advantage of this method, an analogue of the drug Atorvastatin was synthesized in 5 steps (Fig. 5c), demonstrating the synthetic value of the reaction.Open in a separate windowFig. 5Synthetic applications.Mechanistic experiments were performed (Fig. 6) to explore the mechanism of the reaction. When 3 equivalents of TEMPO were added, the reaction was not inhibited and the desired product (2a) was formed in 62% yield (Fig. 6a). This result suggested that the reaction might not involve a radical process. To probe the reaction further, a kinetic study was conducted (Fig. 6b). According to this study, the propargylamide (1a) was completely converted to an allenamide (3a) in 10 min under the standard conditions. The multi-substituted pyrrole (2a) was then gradually produced from the intermediate allenamide and no other reaction intermediates were observed or identified. On the other hand, DFT calculations of substrates 3b and 4a were carried out at the B3LYP-D3(SMD)/Def2-TZVP//B3LYP-D3/Def2-SVP level of theory to identify the natural bond orbital (NBO) charges on the carbons of the allene moieties. NBO charges on the internal carbon in both 3b and 4a are 0.11 and 0.18, respectively (Fig. 6c) indicating that the nucleophilic addition of cyanide anion onto the internal carbon should be reasonable as opposed to its addition onto the terminal carbon. Pathways of the cyano addition to 3b were also calculated (Fig. 6d). The transition state of cyano addition on the internal carbon (TS1), is indeed much lower than addition on the terminal carbon (TS2). The intermediate of internal carbon addition int1, is more stable than int2, implying that the internal carbon addition pathway is not only kinetically but also thermodynamically favoured.Open in a separate windowFig. 6Mechanistic studies and proposed mechanism.Based on the results of these mechanistic studies, a plausible reaction mechanism for the synthesis of pentasubstituted pyrroles and hexasubstituted pyrrolines is proposed and is shown in Fig. 6e. First, under basic conditions, the propargylamide isomerizes to an intermediate allenamide (A), which can be attacked nucleophilically by the cyanide anion to afford an intermediate imine (B) with release of the sulfonyl group. Then, the second cyanide anion attacks the imine to form an intermediate (C), which can undergo cyclization and protonation to afford the fully substituted pyrrole (2). Similarly, the hexasubstituted pyrroline product (5) can be obtained from double nucleophilic attack of the intermediate (A) by the cyanide ion.  相似文献   

14.
Correction: Plasticizer and catalyst co-functionalized PEDOT:PSS enables stretchable electrochemical sensing of living cells     
Jing Yan  Yu Qin  Wen-Ting Fan  Wen-Tao Wu  Song-Wei Lv  Li-Ping Yan  Yan-Ling Liu  Wei-Hua Huang 《Chemical science》2021,12(47):15771
Correction for ‘Plasticizer and catalyst co-functionalized PEDOT:PSS enables stretchable electrochemical sensing of living cells’ by Jing Yan et al., Chem. Sci., 2021, 12, 14432–14440, DOI: 10.1039/d1sc04138j.

The authors regret that there was an error in the equation of the calibration curve of PPL/PDMS in Fig. 3. The correct version is shown below.Open in a separate windowFig. 3(A) CVs of different electrodes obtained in 10 mM K3[Fe(CN)6]. Inset: the enlarged view for CV of the PP electrode. (B) Schematic illustration of the electrocatalysis mechanism. (C) CVs of different electrodes with and without 1 mM H2O2. (D) Amperometric responses of PPL/PDMS (black lines) and PPLC/PDMS (red lines) electrodes to H2O2 at a potential of +0.55 V (vs. Ag/AgCl) to increasing H2O2 concentrations. Inset: the enlargements of amperometric responses framed in blue. (E) Calibration curves of PPL/PDMS and PPLC/PDMS electrodes to increasing H2O2 concentrations (data presented as mean ± standard error, n = 3). (F) Calculated LOD and sensitivity of PPL/PDMS and PPLC/PDMS electrodes to H2O2.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

15.
Construction of vicinal 4°/3°-carbons via reductive Cope rearrangement     
Kristin M. Sobie  Matthew Albritton  Yinuo Yang  Mariana M. Alves  Adrian Roitberg  Alexander J. Grenning 《Chemical science》2022,13(7):1951
Herein reported is a strategy for constructing vicinal 4°/3° carbons via reductive Cope rearrangement. Substrates have been designed which exhibit Cope rearrangement kinetic barriers of ∼23 kcal mol−1 with isoenergetic favorability (ΔG ∼ 0). These fluxional/shape-shifting molecules can be driven forward by chemoselective reduction to useful polyfunctionalized building blocks.

Herein reported is a strategy for constructing vicinal 4°/3° carbons via reductive Cope rearrangement.

Constructing sterically congested vicinal quaternary–tertiary carbons (4°/3° carbons) via Cope rearrangement is currently quite limited with only a handful of papers on the subject published over the past 40 years. This stands in stark contrast to the plethora of other methods for establishing sterically congested vicinal carbons.1–5 Central to the challenge are kinetic and thermodynamic issues associated with the transformation. In the simplest sense, Cope rearrangements proceed in the direction that results in highest alkene substitution (Fig. 1).6,7 To forge 4°/3° motifs by Cope rearrangement, additional driving forces must be introduced to reverse the [3,3] directionality and compensate for the energetic penalty associated with the steric and torsional strain of the targeted vicinal 4°/3° motif. With limited reports in all cases, oxy-Cope substrates (Scheme 1, eqn (1)),8–14 divinylcyclopropanes (Scheme 1, eqn (2)),15–20 and vinylidenecyclopropane-based 1,5-dienes21 (Scheme 1, eqn (3)) have demonstrated favourability for constructing vicinal 4°/3° carbons. Malachowski et al. put forth a series of studies on the construction of quaternary centers via Cope rearrangement driven forward by a conjugation event (Scheme 1, eqn (4)).22–25 In their work, a single example related to the construction of vicinal 4°/3° centers was disclosed, though kinetic (180 °C) and thermodynamic (equilibrium mixtures) challenges are also observed.23 And of particular relevance to this work, Wigfield et al. demonstrated that 3,3-dicyano-1,5-dienes with the potential to generate vicinal 4°/3° carbons instead react via an ionic mechanism yielding the less congested products (Scheme 1, eqn (5)).26Open in a separate windowFig. 1Cope equilibrium of 1,1,6-trisubstituted 1,5-dienes.Open in a separate windowScheme 1(A) Cope rearrangements for constructing vicinal 4°/3°-centers (B) this report.Our group has been examining strategies to decrease kinetic barriers and increase the thermodynamic favourability of 3,3-dicyano-1,5-diene-based Cope substrates.27–31 Beyond the simplest, unsubstituted variants, this class of 1,5-diene is not particularly reactive in both a kinetic and thermodynamic sense (e.g.Scheme 1, eqn (5)).26,32 Reactivity issues aside, these substrates are attractive building blocks for two main reasons: (1) they have straightforward accessibility from alkylidenemalononitriles and allylic electrophiles by deconjugative allylic alkylation.33 (2) The 1,5-diene termini are substantially different (malononitrile vs. simple alkene) thus allowing for orthogonal functional group interconversion facilitating target and analogue synthesis.34 Herein we report that a combination of 1,5-diene structural engineering28,31 and reductive conditions (the reductive Cope rearrangement29,30) can result in the synthesis of building blocks containing vicinal gem-dimethyl 4°/3° carbons along with orthogonal malononitrile and styrene functional groups for interconversion (Scheme 1B). On this line, malononitrile can be directly converted to amides34 yielding functionally dense β-gem-dimethylamides, important pharmaceutical scaffolds.35This project began during the Covid-19 pandemic lockdown (ca. March–May 2020). As such, we were not permitted to use our laboratory out of an abundance of caution. We took this opportunity to first computationally investigate a Cope rearrangement that could result in vicinal 4°/3° carbons (Scheme 2). Then, when permitted to safely return to the lab, we would experimentally validate our findings (vide infra). From our previous work, it is known that by adding either a 4-aromatic group28 or a 4-methyl group31 to a 3,3-dicyano-1,5-diene, low barrier (rt – 80 °C) diastereoselective Cope rearrangements can occur. Notably, the 4-substituent was found to destabilize the starting material (weaken the C3–C4 bond, conformationally bias the substrate for [3,3]), and stabilize the product side of the equilibrium via resonance (phenyl group) or hyperconjugation (methyl group). In this study, we modelled substrates 1, 3, and 5 that have variable 4-substitution and would result in vicinal gem-dimethyl- and phenyl-containing 4°/3° carbons upon Cope rearrangement to 2, 4, or 6, respectively. We chose to target this motif due to likely synthetic accessibility from simple starting materials but also because of the important and profound impact that gem-dimethyl groups impart on pharmaceuticals.35 Substrate 1 lacking 4-substitution had an extremely unfavourable kinetic and thermodynamic profile (ΔG = 31.6; ΔG = +5.3 kcal mol−1). When a 4-methyl group was added, the kinetic barrier (ΔG) dropped appreciably to 28.2 kcal mol; however, the thermodynamics were still quite endergonic (ΔG = +4.4 kcal mol−1). Most excitingly, it was uncovered that the 4-phenyl group dramatically impacted the kinetics and thermodynamics: the [3,3] has a barrier of 22.9 kcal mol−1G) and is ∼isoenergetic (ΔG = +0.17 kcal mol−1). Thus, the reaction appears to be fluxional/shape-shifting at room temperature.36–40 For this substrate, we also modelled the dissociative pathway (Scheme 2D). It was found that bond breakage to two allylic radical intermediates is a higher energy process than the concerted transition state (Scheme 2Cvs.Scheme 2D). Specifically, the dissociative pathway was found to be kinetically less favourable (ΔG ∼ 27.6 kcal mol; ΔG = 26.2 kcal mol−1) than the concerted process (ΔG = 22.9 kcal mol−1). While the dissociative pathway is less favourable than the concerted transformation, we surmised that the two-step process becomes accessible at elevated temperature (vide infra). Finally, the ionic pathway was calculated to be significantly higher for this substrate (see the ESI).Open in a separate windowScheme 2Computational analysis of 3,3-dicyano-1,5-diene that in theory could result in vicinal 4°/3° carbons. (A) 4-Unsubstituted 3,3-dicyano-1,5-diene. (B) 4-Methyl 3,3-dicyano-1,5-diene. (C) 4-Phenyl 3,3-dicyano-1,5-diene. (D) The dissociative mechanism for substrate 5 is higher than the closed transition state. (E) visualization of the kinetic- and thermodynamic differences of transformations (A–D).The class of substrate uncovered from our computational investigation could be accessed from γ,γ-dimethyl-alkylidenemalononitrile (7a) and 1,3-diarylallyl electrophiles (such as 8a) by Pd-catalyzed deconjugative allylic alkylation (Scheme 3A).33 As such, model 1,5-diene 5a was prepared to verify the computational results. It was found that upon synthesis of 5a, an inseparable 21 : 79 mixture of 1,5-diene 5a and the 1,5-diene 6a was observed. The predicted ratio of 5a to 6a was 57 : 43 (Scheme 2C). These two results are within the error of the calculations (predicted; slightly endergonic, observed; slightly exergonic). To determine whether the transformation was progressing through the predicted concerted pathway (Scheme 2C) over the dissociative pathway (Scheme 2D), substrate 5b was prepared by an analogous deconjugative allylic alkylation reaction. Similarly, two Cope equilibrium isomers 5b and 6b are observed at room temperature in a 12 : 88 ratio. Upon heating at 100 °C for 3 h, the 1,5-dienes “scramble” (e.g. iso-6b is observed; 0.2 : 1.0 : 1.5 ratio of 5b : 6b : iso-6b) indicating that the dissociative pathway is only accessible at elevated temperature. This is all in good agreement with the calculated kinetics and thermodynamics of this system (Scheme 2).Open in a separate windowScheme 3(A) Observation of fluxional [3,3] and confirmation of calculated predictions. (B) Optimization of a reductive Cope rearrangement protocol for constructing vicinal 4°/3° centers. (C) The Pd-catalyzed deconjugative allylic alkylation must be regioselective.With respect to the synthetic methodology, we aimed to increase the overall efficiency and applicability of the sequence (Scheme 3B). Specifically, we wanted to avoid [3,3] equilibrium mixtures and sensitive/unstable substates and intermediates. It was found that the direct coupling of 7a with diphenylallyl alcohol 9a could take place in the presence of DMAP, Ac2O, and Pd(PPh3)4. When the coupling was complete, methanol and NaBH4 were added to drive the Cope equilibrium forward, yielding the reduced Cope rearrangement product 10a in 76% isolated yield. In terms of practicality and efficiency, this method utilizes diphenylallyl alcohols, which are more stable and synthetically accessible than their respective acetates, and the [3,3] equilibrium mixture can be directly converted dynamically to a single reduced product.With an efficient protocol in hand for constructing malononitrile–styrene-tethered building blocks featuring central vicinal 4°/3° carbons, we next examined the scope of the transformation (Scheme 4). We chose diarylallyl alcohols with the propensity to react regioselectively via an electronic bias (Scheme 3C).41,42 The combination of p-nitrophenyl and phenyl (10b) or p-methoxyphenyl (10c) yielded regioselective outcomes with the electron-deficient arene at the allylic position. This is consistent with the expected regiochemical outcome where the nucleophile reacts preferentially at the α-position and the electrophile reacts at the allylic position bearing the donor-arene (Scheme 3C).41,42 Then, reductive Cope rearrangement occurs to position the electron-deficient arene adjacent to the gem-dimethyl quaternary center. This is an exciting outcome as many pharmaceutically relevant (hetero)arenes are electron deficient. Thus, fluorinated arenes were installed at the allylic position of products 10d–10k. While the phenyl group resulted in poor regioselectivity (1 : 1–3 : 1), the p-methoxyphenyl group enhanced the regiomeric ratios in all cases (3 : 1–15 : 1). The degree of selectivity is correlated with the number and position of fluorine atoms. N-Heterocycles could be incorporated with excellent regioselectivity, generally speaking (10l–10q). For example, 3-chloro-4-pyridyl (10l/10m) groups were installed at the allylic position with >20 : 1 rr. 4-Chloro-3-pyridyl was poorly regioselective (10n), but the combination of 4-trifluomethyl-3-pyridyl/p-methoxyphenyl (10o) gave good regioselectivity of 11 : 1. 2-Pyridyl/p-methoxyphenyl (10q) was also a regioselective combination. We also examined a few other heterocycles including quinoline (10s) and thiazole (10t and 10u) with excellent and modest regioselectivity observed, respectively. As a general trend, when the arenes on the allylic electrophile become less polarized, poor regioselectivity is observed in the Pd-catalyzed allylic alkylation. For example, the combination of p-chlorophenyl and p-methoxyphenyl (10v) or phenyl (10w) yields regioisomeric mixtures of products. This can be circumvented by utilizing symmetric electrophiles (to 10x).Open in a separate windowScheme 4Scope of the 4°/3°-center-generating reductive Cope rearrangement.The phenyl or the p-methoxyphenyl group is necessary to achieve the 4°/3° carbon-generating Cope rearrangement: it functions as an “activator” by lowering the kinetic barrier and increasing thermodynamic favourability. These activating groups can be removed through alkene C Created by potrace 1.16, written by Peter Selinger 2001-2019 C cleavage reactions (e.g. metathesis (Scheme 5) and ozonolysis (Scheme 6B)). In this regard, highly substituted cycloheptenes 11 were prepared by allylation and metathesis (Scheme 4).28,43 The yields were modest to excellent over this two-step sequence. In many cases, where 10 exists as a mixture of regioisomers, the major allylation/RCM products 11 could be chromatographically separated from their minor constituents. As shown in Scheme 6A, the malononitrile can be transformed via oxidative amidation34 to products 12 containing a dense array of pharmaceutically relevant functionalities (amides, gem-dimethyl, fluoroaromatics, and heteroaromatics). Following this transformation, ozonolysis terminated with a NaBH4 quench installs an alcohol moiety on small molecule 13a.Open in a separate windowScheme 5Removal of the “activating group” by ring-closing metathesis.Open in a separate windowScheme 6(A) oxidative amidation of malononitrile. (B) Removal of “activating group” by ozonolysis.These first computational and experimental studies utilizing 3,3-dicyano-1,5-dienes as substrates for constructing vicinal 4°/3° centers sets the stage for much further examination and application. For example, while we focused our efforts on gem-dimethyl-based quaternary carbons, it is likely that other functionality can be installed at this position. For example, while unoptimized, it appears the protocol is reasonably effective at incorporating a piperidine moiety in addition to heteroarenes from the allylic electrophile (7b + 9f → 14a; Scheme 7A). Similar functional group interconversion chemistry as described in Schemes 5 and and66 can thus yield functionally dense building blocks 15 and 16 in good yields.Open in a separate windowScheme 7(A) The construction of 4/3° centres on piperidines. (B) Promoting endergonic [3,3] rearrangements is possible, assuming the [3,3] kinetic barrier is sufficiently low.While the 4,6-diaryl-3,3-dicyano-1,5-dienes offered the most attractive energetic profile (low kinetic barrier, isoenergetic [3,3] equillibrium; Scheme 2C), the 4-methyl analogue is also intriguing to consider as a viable substrate class for reductive Cope rearrangement (Scheme 2B). The challenge here is that the kinetics and thermodynamics are quite unfavourable (not observable by NMR), but potentially not prohibitively so. It is extremely exciting to find that Cope equilibria that are significantly endergonic in the desired, forward direction (e.g.3a to 4a) can be promoted by a related reductive protocol (Scheme 7B). While unoptimized, we were able to isolate product 17 in xx% yield by heating at 90 °C in the presence of Hantzsch ester in DMF.  相似文献   

16.
Synthesis and enantioseparation of chiral Au13 nanoclusters protected by bis-N-heterocyclic carbene ligands     
Hong Yi  Kimberly M. Osten  Tetyana I. Levchenko  Alex J. Veinot  Yoshitaka Aramaki  Takashi Ooi  Masakazu Nambo  Cathleen M. Crudden 《Chemical science》2021,12(31):10436
  相似文献   

17.
Radical generation enabled by photoinduced N–O bond fragmentation     
Edward J. McClain  Alan K. Wortman  Corey R. J. Stephenson 《Chemical science》2022,13(41):12158
Recent advances in synthetic chemistry have seen a resurgence in the development of methods for visible light-mediated radical generation. Herein, we report the development of a photoactive ester based on a quinoline N-oxide core structure, that provides a strong oxidant in its excited state. The heteroaromatic N-oxide provides access to primary, secondary, and tertiary radical intermediates, and its application toward the development of a photochemical Minisci alkylation is reported.

Recent advances in synthetic chemistry have seen a resurgence in the development of methods for visible light-mediated radical generation.

Photoinduced radical generation has become a focal point of contemporary chemical research. Historically, the photochemical formation of radicals has been achieved via the direct irradiation of organic chromophores with high energy UV light, with notable examples including Norrish – type 1 reactions and the photochemical decomposition of azo compounds, peroxides, N-(acyloxy)-pyridones, and xanthates.1 While a powerful tool, the propensity of organic functional groups to absorb UV light leads to undesired excitation events that ultimately result in uncontrolled, deleterious reactivity.2 The use of visible light irradiation to drive reactivity offers a solution to the problems presented by UV light photochemistry, as typical organic functionalities do not absorb in the visible region of the electromagnetic spectrum. In this regard, photoredox catalysis has emerged as a powerful tool for the generation of radical intermediates, as the visible light absorbing catalysts can access high energy excited states that engage organic substrates in redox events.3 While incredibly versatile, the manipulation of substrate oxidation states can limit the scope of reactivity, as functionalities that are predisposed to oxidation (or reduction) may undergo undesired side reactions. To circumvent this limitation, redox auxiliaries are employed to alter the redox properties of the substrate and facilitate reactivity under mild reducing (or oxidizing) conditions.4Radical formation via photoinduced dissociation of an auxiliary, or complex, represents a complementary strategy that, in principle, should be tolerant of redox sensitive functionalities. However, despite the ubiquity of this strategy in UV light mediated reaction manifolds, the development of reagents that undergo efficient photodissociation upon visible light irradiation is limited. In 2017, Melchiorre and coworkers reported the dissociation of 4-alkyl-1,4-dihydropyridines (alkyl-DHPs) upon irradiation with 405 nm light (Fig. 1A).5 Study of the photoactive alkyl-DHPs revealed that they are reductants in their excited state (D+˙/D* = −2.0 V vs. SCE) and, as such, both the dissociative and redox properties of the excited state could be exploited for ipso-substitution of aryl nitriles and nickel catalyzed C(sp2)–C(sp3) cross-coupling reactions.Open in a separate windowFig. 1Overview of photochemical radical generation in synthetic chemistry.In 2020, Ohmiya and colleagues reported that boracene, when reacted with organolithium or Grignard reagents, forms a photoactive alkyl borate salt that liberates an equivalent of an alkyl radical upon irradiation with 440 nm light (Fig. 1B).6 Characterization of the boracene-based alkyl borate revealed a strongly reducing excited state (B+˙/B* = −2.2 V vs. SCE), and this auxiliary was demonstrated to be effective for nickel catalyzed C(sp2)–C(sp3) cross-coupling reactions with tertiary alkyl radicals.Pioneering work by Barton and coworkers in the 1980''s demonstrated that the thermal or photochemical decomposition of N-(acyloxy)-2-thiopyridones (commonly referred to as Barton esters) could efficiently generate carbon centered radicals, ultimately delivering the corresponding alkane or thioether products.7 However, state-of-the-art methods employing pyridine N-oxide or its derivatives as radical precursors have demonstrated limited intermolecular reactivity, as the generated radical intermediates are competitively trapped by the pyridine-based auxiliary.8,9 To overcome these limitations, we sought to design a pyridine N-oxide based auxiliary that does not undergo undesired alkylation reactions. Inspired by Barton''s work and informed by our previous studies,8,9 we envisioned that the fast fragmentation of N–O bonds in pyridine N-oxide derivatives could be leveraged in the design of a photocleavable activator that would deliver carbon centered radicals from readily available carboxylic acids as precursors. Importantly, the development of a photoactive ester derived from simple pyridine N-oxide core structure delivers a species that is a strong oxidant in its excited state, complementing the reductive excited states of the previously developed photoactive auxiliaries. Moreover, by using a core structure that is derived from an abundant heteroaromatic building block, the designed photoactive esters will contain highly tunable core structures, allowing for control over the photoexcitation and fragmentation events. Herein, we report the development of a photoactive ester derived from a quinoline N-oxide core structure and its application to achieve an efficient intermolecular Minisci alkylation (Fig. 1C).At the outset of our study, we established several requirements to be satisfied by potential photoactive esters (Fig. 2B): (1) the pyridine N-oxide core structure needed to be preserved, (2) the heteroaromatic core would need to have blocking substituents at the sites of alkylation to slow deleterious functionalization of the ester (i.e. 2-, 4-, and 6-substituted pyridine N-oxide derivatives), (3) the heteroaromatic N-oxide would need to maintain sufficient nucleophilicity to form the activated N-acyloxy pyridinium, (4) the heteroaromatic backbone would need to deliver a sterically accessible N-oxide functionality. Additionally, it was recognized that pyridine N-oxide derivatives bearing alkyl substituents with benzylic C–H bonds were not suitable photoactive esters, as a deleterious Boekelheide reaction occurred upon acylation of the N-oxide functionality. Ultimately, 2,4,6-triphenyl pyridine N-oxide (TPPNO), methyl 2-phenylquinoline-4-carboxylate N-oxide (PQCNO), and methyl acridine-9-carboxylate N-oxide (ACNO) were identified as potential photoactive ester precursors that met all aforementioned requirements.Open in a separate windowFig. 2(A) Comparison of pyridinium esters. (B) Overview of photoactive ester design principles. (C) Photophysical characterization of Ac-TPPNO, Ac-PQCNO, and Ac-ACNO.Initial investigations focused on the photophysical characterization of the three heteroaromatic N-oxides (Fig. 2C). Uv-vis spectroscopy in acetonitrile revealed TPPNO to have two maximum absorbances at 319 nm and 365 nm, with the latter absorbance tailing off beyond 400 nm. PQCNO and ACNO also displayed a strong absorbance feature at 365 nm, however, for ACNO two additional lower energy absorbance features at 418 nm and 444 nm were observed. Addition of acetyl chloride to TPPNO and PQCNO resulted in an increase of intensity for the absorbance features, with no apparent shift in the absorbance maxima. Acylation of the ACNO auxiliary with acetyl chloride caused the maximum at 274 nm to diverge into two distinct sharp absorbance features with maxima at 264 nm and 275 nm. Additionally, the absorbance feature at 364 nm both increased in intensity and became more structured upon acylation, revealing an apparent shoulder at 352 nm. Measurement of the fluorescence spectra for the heteroaromatic N-oxides (irradiated at 361 nm) revealed that Ac-TPPNO, Ac-PQCNO, and Ac-ACNO have emission maxima at 435 nm, 468 nm and 519 nm, respectively (Fig. 3A). In the absence of an acyl equivalent, no emission was observed for the heteroaromatic N-oxides. This finding is consistent with previous reports that pyridine N-oxides and quinoline N-oxides possess non-emissive excited states under basic conditions due to a propensity to undergo a fast rearrangement on the singlet surface.10Open in a separate windowFig. 3(A) Absorbance and emission spectra for Ac-TPPNO, Ac-PQCNO, and Ac-ACNO. Emission recorded upon irradiation at 361 nm. (B) Fluorescence spectra of Ac-PQCNO monitored over 25 successive scans (excitation 335 nm). (C) Decomposition of Ac-PQCNO upon irradiation by a 427 nm Kessil lamp, monitored by Uv-vis. (D) Proposed mechanism for photochemical decomposition of Ac-PQCNO.Investigation of the photoinduced N–O bond cleavage of Ac-PQCNO revealed that upon successive fluorescence scans (excitation at 335 nm), Ac-PQCNO is observed to decompose steadily with the concomitant appearance of the fluorescence signal corresponding to the generation of deoxygenated quinoline (Fig. 3B). Interrogation of the decomposition by Uv-vis spectroscopy revealed that after 60 s of irradiation of Ac-PQCNO with a 427 nm Kessil lamp, a decrease in the absorbance feature at 365 nm was observed. Upon extending irradiation time out to 30 min, significant degradation of the Ac-PQCNO was observed.Together, the photophysical characterization shows the photoactive esters possess conserved absorption features at 365 nm. Consistent with previous studies on quinoline N-oxide photochemistry, the absorption feature at 365 nm is thought to be π,π* in character and responsible for the observed deoxygenation reactivity.11 The deoxygenation of the aromatic N-oxides is hypothesized to arise from crossing over of the π,π* excited state to a dissociative π,σ* state. Computational evidence in support of the hypothesized mechanism can be found in Fukuda and Ehara''s study on the deoxygenation of structurally related N-hydroxypyridine-2(1H)-thione.12 Experimental support for this proposed mechanism can be found in Hata and Tanaka''s study of the gas phase photolysis of pyridine N-oxide13 as well as Hata''s subsequent translation of this work to the solution phase photodeoxygenation of heteroaromatic N-oxides in the presence of a strong Lewis acid, BF3·Et2O.14Interrogation of the thermal reactivity of Ac-PQCNO revealed there was no deoxygenation of the heteroaromatic N-oxide after heating to 90 °C for 3 hours. This demonstrates that the remarkable reactivity of Ac-PQCNO is due to dissociation of the N–O bond from a low energy photoexcited state.Electrochemical analysis of the heteroaromatic N-oxide photoactive esters using cyclic voltammetry in acetonitrile revealed the Ac-TPPNO, Ac-PQCNO, and Ac-ACNO to have low energy reduction waves, with measured E1/2 of −0.79 V vs. SCE, − 0.45 V vs. SCE, and −0.47 V vs. SCE. Applying information gathered from absorbance and emission spectroscopy, as well as the electrochemical data, the standard potential for oxidation of Ac-TPPNO, Ac-PQCNO and Ac-ACNO in the excited state was estimated to be +2.30 V [E0(T+*/T˙)], +2.57 V [E0(P+*/P˙)], and +2.32 V [E0(A+*/A˙)] vs. SCE according to the Rehm–Weller approximation (Fig. 2C).15Due to the mild photolytic conditions for decarboxylative radical generation from N-oxides and characteristics as excited state oxidants, we sought to assess the reactivity of the photocleavable esters towards the development of an intermolecular Minisci alkylation (Fig. 4). Initial studies focused on the addition of the tert-butyl radical to 4-chloroquinoline in acetonitrile. Assessment of reactivity revealed TPPNO and PQCNO to perform identically under unoptimized conditions delivering 27% of the desired 2-tert-butyl-4-chloroquinoline product, while ACNO produced less than 5% of the desired product. Due to the low cost of the parent quinoline and its high yielding N-oxidation,16 we elected to focus our investigations on the application of PQCNO as a photoactive ester. Degassing the reaction resulted in a substantial increase in yield, providing 90% isolated yield of the desired product. Further investigation of solvents and additives did not result in increased reactivity.Open in a separate windowFig. 4Scope of intermolecular Minisci alkylation. Isolated yields unless otherwise noted. Standard conditions: substrate (0.2 mmol, 1 equiv.), PQCNO (2 equiv.), acyl chloride (2.2 equiv.), CaCl2 (1 equiv.), MeCN (0.5 mL), N2 atmosphere, 427 nm Kessil lamp, 30 min. aReaction run for 1 hour. bReaction run for 2 hours. cReaction run for 1 hour with a 1 : 1 mixture of MeCN to DCM as solvent. NMR yield with methyl tert-butyl ether as internal standard, in brackets.Assessment of the scope for the intermolecular Minisci alkylation revealed the tert-butyl radical addition was the most efficient of the simple alkyl fragments (90%, 4a), followed by isopropyl addition (50%, 4b), ethyl addition (48%, 4c), and finally methyl addition (43%, 4d) (Fig. 4). Simple carbocyclic radical fragments such as cyclohexyl (4e), cyclobutyl (4f), and cyclopropyl (4g) provided the corresponding alkylation products in 88%, 57%, and 34% yield, respectively. Interestingly, Minisci alkylation products from more complex carbocyclic radical fragments were also accessible under-developed reaction conditions as 1-phenylcyclopropyl (4h), adamantyl (4i), 4-(methoxycarbonyl)-[2.2.2]-bicycloctyl (4j), and 3-(methoxycarbonyl)-[1.1.1]-bicyclopentyl (4k) fragments gave the corresponding radical addition products in 37–80% yield. Application of tetrahydropyran-4-carbonyl chloride (4l), N-Boc azetidine 3-carbonyl chloride (4m), N-Boc piperidine 4-carbonyl chloride (4n), and N-Boc piperidine 3-carbonyl chloride (4o) provided the desired coupling products in 41–61% yield, respectively.Assessment of the heterocyclic coupling partners revealed that substituted pyridine and quinoline derivatives performed well in the Minisci alkylation, however, over-alkylation of the heteroarene was often observed (4p–4t). Notably, 4-cyanopyridine (4p) and lepidine (4s) performed well under the developed intermolecular reaction conditions; these substrates were either low yielding or inaccessible under our previously reported fragment coupling conditions.8 Modestly complex heteroaromatic scaffolds such as quinoxaline (4u), and 3-chloro-6-phenylpyridazine (4v) also performed well under the tert-butylation reaciton conditions. Biologically active scaffolds such as the 4-chloroquinazoline core of erlotinib (4w) and the imidazopyrazine core structure of gandotinib (4x) each provided a single regioisomer of the tert-butyl addition product in high yield. Finally, nicotine (4y) was observed to undergo tert-butyl addition in 37% yield with retention of the configuration at the benzylic stereocenter.During the exploration of the Minisci alkylation reaction, it was realized that the deoxygenated quinoline could be recovered in high yields (87% recovery). Resubjecting the recovered quinoline to oxidation conditions, PQCNO could be re-generated in 71% yield. Regenerated PQCNO was observed to show no decrease in reactivity when recycled through three consecutive reactions (see ESI).To probe the mechanism of the photoinduced Minisci alkylation reaction, we monitored the reaction by employing an in situ LED NMR device equipped with a 430 nm LED source. Investigation of the N–O bond fragmentation revealed that the rate of deoxygenation was not dependent on the rate of decarboxylation for the acyloxy group. Benzoyloxy and pivaloyloxy substituted N-oxides provided similar rates of deoxygenation (kobs = −5.8 × 10−5 M s−1 (BzO–) vs. kobs = −6.3 × 10−5 M s−1 (PivO–); see ESI), a result that is interesting given the difference in rates of decarboxylation form the respective carboxylic acids.17,18 Alteration of the aryl substituent in the 2-position of the PQCNO auxiliary was observed to impact the rate of deoxygenation, as ortho substituted 2,6-dimethylphenyl substituent provided an increased rate of deoxygenation (kobs = −7.9 × 10−5 M s−1), whereas electron rich para-methoxyphenyl substituent led to a substantial decrease in the observed rate for deoxygenation (kobs = −2.2 × 10−5 M s−1).Monitoring the deoxygenation of Piv-PQCNO under reaction conditions revealed an increase in the observed rate of deoxygenation in the presence of 4-chloroquinoline (Fig. 5A). We hypothesize that the change in rate for deoxygenation is indicative of a propagative reaction mechanism in which electron transfer from the radical addition product (III) to acylated PQCNO (I) leads to formation of the final C–H alkylated product as well as generating a second equivalent of R˙ through a reductive decarboxylation of acylated PQCNO (I). Determination of the quantum yield (Φ) for the decomposition of Piv-PQCNO supported the proposed propagative chain mechanism, in the presence of 4-chloroquinoling Φ = 0.983 whereas in the absence of a substrate Φ = 0.218 for Piv-PQCNO decomposition. The increase in observed quantum yield is indicative of chain mechanism promoting Piv-PQCNO decomposition in the presence of a substrate. Further support was found when subjecting 2-phenylpropionyl chloride to the reaction conditions in the absence of a heteroaromatic substrate, 1-chloro-1-phenylethane was observed (see ESI), demonstrating the ability of acyl PQCNO to oxidize stabilized radical intermediates.Open in a separate windowFig. 5(A) Reaction profile in the presence (blue) and absence (yellow) of substrate, monitored using 430 nm LED NMR apparatus. (B) Proposed mechanism for photomediated Minisci reaction.19On the basis of these findings, we propose the following mechanism (Fig. 5B). Initiation of the Minisci alkylation occurs by the photoinduced decomposition of acyl-PQCNO (I) to generate an equivalent of a reactive radical intermediate (R˙). Addition of R˙ to an equivalent of the protonated heteroaromatic substrate provides intermediate (II). Deprotonation of (II) generates radical intermediate (III) that is in turn oxidized by a second equivalent of acyl-PQCNO (I), providing the desired C–H alkylated product while generating a second equivalent of R˙ through the decomposition of reduced acyl-PQCNO (IV), thereby propagating the reaction.Finally, we sought to explore alternate radical transformations that can be promoted by PQCNO-based esters (Scheme 1). Lactonization of 2-phenylbenzoyl chloride was carried out, providing moderate yield for the 3,4-benzocoumarin product.20 By employing trifluoroacetic anhydride (TFAA) as an acyl equivalent in the presence of tert-butyl anisole, radical trifluoromethylation of the electron-rich arene was achieved in moderate yields.8 Further assessments of reactivity are currently ongoing within our laboratory.Open in a separate windowScheme 1Alternate radical transformations.In conclusion, we have developed a photoactive ester based upon the quinoline N-oxide core which delivers a strong oxidant in its excited state. The designed photocleavable ester enabled the development of a photochemical Minisci alkylation, providing a reaction platform that leveraged both the photochemical dissociation and the oxidizing characteristics of the photoactive esters. The photochemical reactivity of the PQCNO ester was also demonstrated to effect radical lactonization and trifluoromethylation reactions.  相似文献   

18.
One step synthesis of unsymmetrical 1,3-disubstituted BCP ketones via nickel/photoredox-catalyzed [1.1.1]propellane multicomponent dicarbofunctionalization     
Weichen Huang  Sebastian Keess  Gary A. Molander 《Chemical science》2022,13(40):11936
Bicyclo[1.1.1]pentanes (BCPs), utilized as sp3-rich bioisosteres for tert-butyl- and aryl groups as well as internal alkynes, have gained considerable momentum in drug development programs. Although many elegant methods have been developed to access BCP amines and BCP aryls efficiently, the methods used to construct BCP ketones directly are relatively underdeveloped. In particular, the preparation of unsymmetrical 1,3-disubstituted-BCP ketones remains challenging and still requires multiple chemical steps. Herein, a single-step, multi-component approach to versatile disubstituted BCP ketones via nickel/photoredox catalysis is reported. Importantly, installing a boron group at the carbon position adjacent to the BCP structure bypasses the limitation to tertiary BF3K coupling partners, thus expanding the scope of this paradigm. Further transformation of disubstituted-BCP ketones into a variety of other BCP derivatives demonstrates the synthetic value of this developed method.

Bicyclo[1.1.1]pentanes (BCPs), utilized as sp3-rich bioisosteres for tert-butyl- and aryl groups as well as internal alkynes, have gained considerable momentum in drug development programs.

Three-dimensional (3D) molecular scaffolds have received considerable attention in drug molecular design to improve physicochemical properties of drug candidates.1 Among the promising 3D scaffolds in this area are the bicyclo[1.1.1]pentanes (BCPs), which serve as bioisosteres of aromatic rings as well as tert-butyl- and alkyne groups in medicinal chemistry.2 In Stepan''s pioneering work,2a the replacement of the fluorinated aryl ring of a gamma secretase inhibitor with a BCP moiety resulted in improved permeability and kinetic solubility. Since this landmark work, the number of patents published with BCP-containing drugs has skyrocketed. Despite considerable interest from the medicinal chemistry community, the incorporation of BCPs into specific structural classes found in bioactive molecules remains an unsolved challenge.BCP ketones could be considered as bioisosteres of aryl ketones, which widely exist in FDA-approved drugs (Fig. 1A).3 They can also be used as vehicles for the synthesis of other important BCP derivatives, including BCP amides and BCP esters through efficient transformations. Nevertheless, the methods that are used to construct BCP ketones efficiently are relatively underdeveloped, especially compared with well-developed approaches to access amino BCPs and aryl BCPs (Fig. 1B).4 Specifically, the Wiberg,5a Walsh,5b and Pan5c groups have reported methods for acylation of [1.1.1]propellane with aldehydes to form monosubstituted-BCP ketones. In contrast, the preparation of unsymmetrically 1,3-disubstituted-BCP ketones remains challenging and still requires multiple chemical steps. For example, Wills and coworkers reported a method for the synthesis of BCP ketones by reacting [1.1.1]propellane and Grignard reagents, followed by addition to an aldehyde and oxidation with MnO2 (Fig. 2A).6a This method requires the use of metal reagents and multiple synthesis steps, which are incompatible with the construction of complex targets containing sensitive functional groups. The Knochel group developed a similar two-step strategy to construct 1,3-disubstituted BCP ketones by opening the [1.1.1]propellane with allylzinc halides, followed by addition to acyl chlorides (Fig. 2A).6b However, this method is only suitable for some special organozinc reagents, which limits the diversity of the BCP ketones. Chemists at SpiroChem also reported a two-step method for construction of 1,3-disubstituted BCP ketones through a process involving radical addition to [1.1.1]propellane, followed by engagement with different arylmetal reagents (Fig. 2A).6c In this case, the other substituent on the BCP ring is limited to an ester functional group. Furthermore, there are some individual examples showing that disubstituted BCP ketones can be obtained from the corresponding BCP redox active ester. Specifically, the Ohmiya group developed the N-heterocyclic carbene-catalyzed acylation of BCP redox active ester, but the yield was only 20% (Fig. 2A).6d The Yuan group also conducted the cross-coupling of BCP redox active esters with pyridyl esters to access BCP ketones (Fig. 2A).6e Considering the five-step synthesis of BCP ketones from [1.1.1]propellane, these methods cannot meet the requirements of rapid synthesis of a library of products in the medicinal chemistry setting. Clearly, the drawbacks of stepwise synthetic approaches to 1,3-disubstituted BCP ketones hamper the broad application of bicyclo[1.1.1]pentanes. Thus, more efficient methods for the preparation of disubstituted BCP ketones are urgently needed.Open in a separate windowFig. 1(A) Examples of bioactive diaryl ketones. (B) Representative BCP derivatives.Open in a separate windowFig. 2(A) Previous strategies to access unsymmetrically 1,3 disubstituted BCP ketones. (B) Research reported herein. HE = Hantzsch ester; RAE = Redox active ester [N-(acyloxy)phthalimide]; NHC = N-heterocyclic carbene; CzIPN = 1,2,3,5-tetrakis(carbazol-9-yl)-4,6-dicyanobenzene.Multicomponent reactions (MCRs) that allow one-step access to complex and diverse disubstituted BCP products are synthetically advantageous to current stepwise approaches to BCP derivatives. However, achieving such a transformation is still challenging because of competing two-component coupling or propellane oligomerization. Uchiyama,7a MacMillan,7b and our group7c,d have successfully developed multi-component approaches to versatile BCP derivatives based on the differentiated reactivity of BCP radicals and substrate alkyl radicals. In our previous report,7d we successfully took advantage of the slow capture of tertiary radicals by Ni species as a key mechanistic aspect to achieve a one-step, multicomponent reaction for the synthesis of BCP-aryl derivatives. Meanwhile, our group has successfully developed an efficient photoredox/Ni dual catalysis paradigm for transition metal-catalyzed cross-couplings of alkylboron- or alkylsilicon reagents with various electrophiles, including aryl halides, acyl chlorides, alkenyl halides, and isocyanates based on a single-electron transfer (SET) transmetalation pathway.8 Inspired by these results, we questioned whether acyl chlorides or other electrophiles could also serve as partners in the three-component radical coupling of [1.1.1]propellane to access a diverse array of BCP derivatives of high importance in the pharmaceutical industry. Herein we report a one-step, three-component radical coupling of [1.1.1]propellane to afford diversely functionalized bicycles using various electrophiles.To determine the chemoselectivity of the proposed MCR pathway, the reactivity of tertiary alkyl and BCP radicals in the nickel/photoredox-catalyzed cross-couplings with acyl chlorides was first examined (Fig. 3). The results indicated that BCP bridgehead radicals engage the nickel catalyst to enter the cross-coupling catalytic cycle, generating the product BCP ketone, while acyclic tertiary radicals do not take part in this catalytic cycle. Encouraged by this promising reactivity pattern, we explored the possibility of achieving a multi-component reaction forging two C–C bonds in a single operation using [1.1.1]propellane.Open in a separate windowFig. 3Control experiments.Initial investigations utilized t-BuBF3K, [1.1.1]propellane, and benzoyl chloride as a model reaction to optimize the reaction conditions ( EntryDeviation from standard conditionsNMR yield (%) 1 None 63 2No base3230.01 M5240.025 M555427 mm2562 mol% [Ir] cat. 10 mol% [Ni]4972 mol% [Ir] cat. 20 mol% [Ni]588No [Ni] catalyst09No [Ir] catalyst010No light011 t-BuCOOCs, instead of 10Open in a separate windowaOptimization of reaction conditions: 1 (0.15 mmol), 2 (0.3 mmol), 3 (0.10 mmol) under purple Kessil irradiation (λ max = 390 nm) for 16 h at rt; NMR yield was calculated using 1,3,5-trimethoxybenzene as an internal standard (IS) from the crude mixture.With suitable conditions in hand, the generality of this metallaphotoredox protocol with respect to a broad range of aliphatic- and aromatic acyl chlorides was investigated. As summarized in Fig. 4, both electron-rich and electron-poor aromatic acyl chlorides were coupled under the developed reaction conditions with 28−64% yields (4–15). For example, aromatic acyl chlorides containing common functional groups such as ether (5), fluoro (6), chloro (7), trifluoromethoxy (8), cyano (10) and trifluoromethylthio (12) proved to be suitable. Heteroaromatic acyl chlorides (14, 15) also react smoothly to afford the desired product in acceptable yield. Furthermore, the success of the reaction with ethyl succinyl chloride (19), which was not compatible utilizing previous methods employing metal reagents, further demonstrates the functional group compatibility of this protocol. Notably, alkyl bromide or -chloride handles (20, 21) have been incorporated, thus enabling further modification by substitution. Finally, other electrophiles including isocyanates and alkenyl halides (24, 25) have been embedded within the substrates, although the efficiency is not ideal in these cases.Open in a separate windowFig. 4Scope of aliphatic and aromatic acyl chloride substrates. Reaction conditions: acyl chlorides (0.30 mmol, 1.0 equiv.), [1.1.1]propellane (0.90 mmol, 3.0 equiv.), alkyltrifluoroborates (0.45 mmol, 1.5 equiv.), [Ir(dFCF3ppy)2dtbbpy]PF6 (2 mol%, 0.006 mmol), Ni(dtbbpy)Br2 (20 mol%, 0.06 mmol), Cs2CO3 (1.5 equiv., 0.45 mmol), DME (0.05 M), irradiating with purple Kessil irradiation (λ max = 390 nm) for 16 h at rt.To explore the generality of this transformation further, a variety of structurally diverse tertiary- and secondary alkyltrifluoroborates were investigated using the developed conditions. Thanks to the development of powerful synthetic methods, tertiary boronate esters9 are quite readily available from diverse feedstocks including carboxylic acids, alkenes, alkyl halides, and ketones.10 As demonstrated in Fig. 5, ester-, nitrile-, ketone-, alkene-, and even hydroxyl-containing trifluoroborates were incorporated into the established protocol (26–32). These sensitive functional groups would be difficult to integrate within previously reported synthetic methods, especially those using metal reagents as the coupling partner. Additionally, alkyltrifluoroborates possessing various ring sizes reacted smoothly to afford the coupled products (33–39). Interestingly, the secondary radical derived from a benzyltrifluoromethyl-substituted alkyltrifluoroborate was engaged in this MCR process (40), with no evidence for formation of the two-component product. Aliphatic acyl chlorides were also tested and found to be compatible with the reaction conditions, affording the corresponding products 41–43. Finally, we applied this method to the late-stage modification of drug-like molecules. Several alkyltrifluoroborate-containing natural products and drug scaffolds were incorporated under the standard conditions to afford the desired products in moderate to good yields (44–48), demonstrating the applicability of the developed method in complex molecular settings.Open in a separate windowFig. 5Scope of aliphatic- and aromatic acyl chloride substrates. Reaction conditions: acyl chlorides (0.30 mmol, 1.0 equiv.), [1.1.1]propellane (0.90 mmol, 3.0 equiv.), alkyltrifluoroborates (0.45 mmol, 1.5 equiv.), [Ir(dFCF3ppy)2dtbbpy]PF6 (2 mol%, 0.006 mmol), Ni(dtbbpy)Br2 (20 mol%, 0.06 mmol), Cs2CO3 (1.5 equiv., 0.45 mmol), DME (0.05 M), irradiating with purple Kessil irradiation (λ max = 390 nm) for 16 h at rt.Although the established one-step, three-component radical coupling enabled by nickel/photoredox dual catalysis provides an efficient method for rapid construction of disubstituted BCP ketones, this protocol was only applicable to tertiary radicals or a specific secondary radical that limits its generality. To resolve this issue, we examined the feasibility of incorporating a substituent on the carbon adjacent to the BF3K group that would serve as a versatile surrogate group. As an example, if a Bpin group was installed into this position, it could be proto-deborylated or even further manipulated in downstream transformations, greatly expanding the scope of the overall process. The Masarwa group reported a method for the desymmetrization of gem-diborylalkanes,11 allowing ready access to the requisite trifluoroborates. Gratifyingly, when the desymmetrized 1,1-dibora substrate was subjected to the developed reaction conditions with an aromatic- and aliphatic acyl chloride, the desired products 49 and 50 were formed in good yield (Fig. 6A).Open in a separate windowFig. 6(A) the Synthesis of β-Bpin-substituted BCP ketones. (B) Further transformations.Ketones have long been used as important intermediates to provide access to other functional groups. We have utilized the efficient access to BCP ketones provided by the method developed herein to demonstrate their conversion into a variety of diverse BCP substructures, including carboxylates and amides through classical functional group interconversions (Fig. 6). Considering the lack of efficient methods for synthesis of such building blocks, the current protocol takes on added significance. Reduction of ketones with NaBH4 produces the corresponding secondary alcohol 51 with a good yield. By using a Baeyer–Villiger oxidation, the corresponding BCP carboxylate 52 was formed. Alternatively, the ketone was further transformed into BCP amide 24via a Beckmann rearrangement. Finally, a BCP ketone was used to generate the corresponding alkene in 75% yield through a Wittig olefination.To gain insights into the reaction mechanism, we conducted a series of control experiments. First, competition experiments demonstrated that a tertiary radical participates in the three-component reaction exclusively, while a secondary radical was only involved in the classical cross-coupling reaction (Fig. 7A). TEMPO trapping experiments showed that the reaction was completely suppressed in the presence of this reagent, and only TEMPO adducts 56 derived from the radical precursors were observed (Fig. 7B). The reaction of the alkyltrifluoroborate generated from verbenone under the standard conditions afforded ring-opened product 58 (Fig. 7C).12 Therefore, the radical nature of the MCR process was confirmed. Based on these results and previous reports,10 a plausible reaction mechanism for this dual nickel/photoredox catalyzed three-component cross-coupling is depicted in Fig. 7D. Initially, under light irradiation, the photocatalyst is excited to provide *Ir(iii). The alkyltrifluoroborates reductively quench the excited photocatalyst *Ir(iii) to generate tertiary alkyl radical V. Because the metal–carbon bond between the nickel center and tertiary alkyl group is quite fragile, the acyclic tertiary radical favorably dissociates from the Ni(iii) center to form free alkyl tertiary radicals.13 Tertiary alkyl radical V undergoes irreversible radical addition to [1.1.1]propellane, leading to BCP radical VI, which is then trapped by Ni(0), forming an alkyl Ni(i) species VIII. Subsequently, VIII undergoes rapid oxidative addition with acyl chlorides. Alternatively, as shown in blue, BCP radical VI can also be captured by Ni(ii) oxidative addition complex IX. Both pathways lead to Ni(iii) complex X, which subsequently undergoes rapid and productive C–C bond formation to yield the BCP ketone products.Open in a separate windowFig. 7Mechanistic study. (A) Secondary versus tertiary radical competition (the ratio was determined by GC-MS analysis). (B) Radical-trapping experiment. (C) Radical ring-opening reaction. (D) Proposed mechanism.In conclusion, the multi-component radical cross-coupling reaction involving [1.1.1]propellane reported herein enables rapid access to a diverse array of disubstituted BCP ketones and offers an expedient alternative to traditional routes for the synthesis of BCP ketones via pre-functionalization of [1.1.1]propellane. The method exhibits several advantages over previously reported routes, including excellent chemoselectivity, mild reaction conditions, and good functional group tolerance. Importantly, the usefulness of this method is further boosted by installing boronate esters (Bpin) at the carbon adjacent to the BCP substructure, which in principle could be protodeborylated or even further manipulated in downstream transformations. Overall, the reaction described herein enables access to unprecedented BCP structures of interest to the organic chemistry synthetic community, especially in the drug discovery sector.  相似文献   

19.
Radical 1,2,3-tricarbofunctionalization of α-vinyl-β-ketoesters enabled by a carbon shift from an all-carbon quaternary center     
Qi Zhang  Mong-Feng Chiou  Changqing Ye  Xiaobin Yuan  Yajun Li  Hongli Bao 《Chemical science》2022,13(23):6836
Herein, we report an intermolecular, radical 1,2,3-tricarbofunctionalization of α-vinyl-β-ketoesters to achieve the goal of building molecular complexity via the one-pot multifunctionalization of alkenes. This reaction allows the expansion of the carbon ring by a carbon shift from an all-carbon quaternary center, and enables further C–C bond formation on the tertiary carbon intermediate with the aim of reconstructing a new all-carbon quaternary center. The good functional group compatibility ensures diverse synthetic transformations of this method. Experimental and theoretical studies reveal that the excellent diastereoselectivity should be attributed to the hydrogen bonding between the substrates and solvent.

Herein, we report an intermolecular, radical 1,2,3-tricarbofunctionalization of α-vinyl-β-ketoesters to achieve the goal of building molecular complexity via the one-pot multifunctionalization of alkenes.

A leading motive for the impressive achievements in the area of assembling molecular complexity is the transformation of simple feedstock chemicals into complex molecular skeletons with superior bioactive properties. In this respect, the direct functionalization of alkenes has been demonstrated as one of the most effective and simple strategies to meet this criterion at a high level. While the difunctionalization of alkenes in a one-pot process is the major theme of considerable interest in this field,1 the multifunctionalization of alkenes,2 for example, a 1,2,3-trifunctionalization of alkenes, has the power to simultaneously incorporate multifunctional groups. Therefore, this multifunctionalization reaction model can be regarded as an efficient and novel strategy to afford molecules with high structural diversity and complexity. However, such methods are elusive.During the last decades, radical alkene functionalizations have been revealed to be a powerful tool for building complex molecular frameworks by employing a radical initiator, a transition metal catalyst, or a photocatalyst.1fi However, only several successful methods for the radical multifunctionalization of alkenes have been achieved. For example, the Studer group reported an elegant 1,2-boryl shift-enabled radical 1,2,3-trifunctionalization of allylboronic esters using AIBN as the radical initiator (Fig. 1a).3 Shi et al. disclosed an excellent photocatalytic perfluoroalkylation of a vinyl-substituted all-carbon quaternary center through 1,2-aryl migration (Fig. 1b).4 Herein, we report a new one-pot protocol to realize an intermolecular, radical 1,2,3-tricarbofunctionalization of α-vinyl-β-ketoesters through a cascade process of deconstruction–reconstruction of the all-carbon quaternary center (Fig. 1c).5Open in a separate windowFig. 1Radical 1,2,3-trifunctionalization of alkenes. (a) Studer''s work; (b) Shi''s work; (c) This work.The direct incorporation of a fluorine atom or fluorinated moieties into organic compounds has been extensively investigated and proved to be an significant synthetic strategy in the field of discovering new pharmaceuticals.6 Recently, we are interested in the radical functionalization of alkenes with fluoroalkyl groups,7 and we envisioned that, different from the typical Dowd–Beckwith8 ring expansion reaction,9 the addition of a fluoroalkyl radical to the C Created by potrace 1.16, written by Peter Selinger 2001-2019 C double bond would generate an adduct radical species I, which will transform into the radical intermediate II upon ring expansion (Fig. 1c). Finally, the cascade C–C coupling affords the product with a reconstructed all-carbon quaternary center. However, there are several challenging issues that need to be addressed: (1) the carbon shift from an all-carbon quaternary center to afford a tertiary carbon center which is bulkier than the tertiary carbon center formed in a typical Dowd–Beckwith ring expansion reaction; (2) the reconstruction of all-carbon quaternary center from tertiary carbon radical II will meet the associated conformational restriction and steric congestion; (3) side reactions, such as 1,2-radical addition to the alkenyl group, homolytic couplings of the carbon radical intermediates I and II, and direct H-atom abstraction;10 (4) how to control the diastereomeric ratio of the products. To meet these challenges, we developed a novel method for the 1,2,3-trifunctionalization of alkenes using alkynyl triflones as both the CF3 (ref. 6) and alkynyl sources, providing the ring-expanded cyclic β-ketoesters with excellent diastereoselectivity and functional group diversity. In addition, good functional group compatibility of this method was observed, which ensures the diverse synthetic transformations. Moreover, hydrogen bonding between the substrates and 2,2,2-trifluoroethanol solvent was revealed to be the key factor for the excellent diastereoselectivity obtained in this reaction, and this result was confirmed by both experimental and theoretical studies.This study began by surveying radical initiators for 1,2,3-tricarbofunctionalizing α-vinyl-β-ketoester 1a with alkynyl triflone 2a11 (12 (13 dramatically increased the diastereoselectivity and (±)-3a could be obtained in an identical yield with an even higher dr value (dr > 20 : 1) (14 Without the addition of a radical initiator, a reaction did not happen ( EntrySolventYieldb (%)1EA60 (dr = 13 : 1)c2EA55 (dr = 11 : 1)d3EA63 (dr = 12 : 1)4MTBE45 (dr = 10 : 1)5DCE63 (dr = 15 : 1)6TolueneTrace7DMFTrace8MeOHTrace9TFE63 (dr > 20 : 1)10eTFE60 (dr > 20 : 1)11fTFE56 (dr > 20 : 1)12gTFE70 (dr > 20 : 1)13hTFE76 (65)i (dr > 20 : 1)14jTFE71 (dr > 20 : 1)15TFETraceOpen in a separate windowaReaction conditions: alkene 1a (0.2 mmol, 1 equiv.), 2a (0.6 mmol, 3.0 equiv.), and AIBN (0.3 equiv.) in 3 mL of solvent at 85 °C for 18 h in a sealed tube under a nitrogen atmosphere.bCrude yield and crude diastereomeric ratio were determined by 19F NMR.cLPO was used as the initiator.dBPO was used as the initiator.eThe reaction was performed at 100 °C.fThe reaction was performed at 120 °C.gAIBN (60 mol%) was used.h2a (3.0 equiv.) and AIBN (60 mol%) were added as two equal portions with an interval of 9 h.iIsolated yield in parentheses.j2a (3.0 equiv.) and AIBN (60 mol%) were added as three equal portions with an interval of 6 h.Under optimal conditions, a diverse array of α-vinyl-β-ketoesters serve as substrates in this metal-free deconstruction–construction of all-carbon quaternary centers for the synthesis of carbon-ring expanded cyclic β-ketoesters (Fig. 2). In most of the cases, excellent diastereoselectivities (dr > 20 : 1) were observed by crude 19F NMR analysis. Substrates with the substituents at the 5- or 6-position of the α-vinyl-β-ketoesters generally produced the corresponding product (±)-3 in higher yields than those with the substituents at the 4-position. Apart from the carbonyl group and the ester group, functional groups such as chloride ((±)-3b and (±)-3f), fluoride ((±)-3c), a methoxyl group ((±)-3d and (±)-3h), a methyl group ((±)-3e and (±)-3g) and a phenyl group ((±)-3i) can be tolerated under the reaction conditions. Notably, the phenyl ring of the core structure with two substituents reacted smoothly to afford the corresponding products ((±)-3j and (±)-3k). When substrate 1l that lacks the fused benzene ring was used for this carbon-ring expansion reaction, a dramatical loss of diastereoselectivity was detected, presumably because of the feasible interconversion of the boat and chair conformations of the intermediate. Substrates with an ethyl ester or a benzyl ester group, as opposed to a methyl ester group, delivered the corresponding products ((±)-3m and (±)-3n) with moderate yields and excellent diastereoselectivity. When the CH2 unit of the six membered-ring was replaced by a CMe2 group, only a trace amount of the desired product (±)-3o was detected. A reaction with the purpose of realizing an extension from the six-membered ring was also carried out and (±)-3p was obtained, although with a low yield and low diastereoselectivity. Notably, the diastereochemistries of products (±)-3e and (±)-3h have been confirmed by X-ray crystallography.Open in a separate windowFig. 2Substrate scope of α-vinyl-β-ketoesters. aThe reaction was performed with 1p and 2b.The scope with respect to the alkynyl triflones was also investigated and the results are summarized in Fig. 3. Generally, substituents on the phenyl ring of the arylethynyl moiety have little impact on the yields of the corresponding products. The functional groups at the para-, meta-, or ortho-position of the phenyl ring produced the desired products ((±)-4a–(±)-4k) with excellent diastereoselectivities. Furthermore, the method is compatible with alkynyl triflones that have a thienyl group or a perfluorobutyl group and the reactions afforded the product ((±)-4l or (±)-4m) with an excellent dr value, respectively. However, when the arylethynyl moiety was replaced by an alkylethynyl or a silylethynyl part, the reaction failed to produce the targeted tricarbofunctionalization product ((±)-4n or (±)-4o).15 Moreover, when triflic azide or (Z)-TolCH Created by potrace 1.16, written by Peter Selinger 2001-2019 CHSO2CF3 was used in place of the alkynyl triflone, the desired product was not obtained and most of the starting material was recovered. Notably, the diastereochemistry of product (±)-4a has been confirmed by X-ray crystallography.Open in a separate windowFig. 3Substrate scope of alkynyl triflones.This 1,2,3-trifunctionalization reaction not only allows the deconstruction and reconstruction of all-carbon quaternary centers, but features good functional group tolerance and excellent diastereoselectivity. Regarding the diverse reactivities of these functional groups, many valuable synthetic transformations have been successfully achieved (Fig. 4). For example, the C–C triple bond of (±)-4a can be completely reduced to a CH2CH2 unit ((±)-5) in the presence of hydrogen and a Pd/C catalyst,16 while the selective reduction of (±)-4a gives rise to a Z-alkene (±)-6 when quinoline is added as an additive for the Lindlar reduction.17 The diastereochemistry of (±)-6 has been confirmed by X-ray crystallography. The selective reducing methods afford formal approaches for radical 1,3-trifluoromethylalkylation and 1,3-trifluoromethylalkenylation of α-vinyl-β-ketoesters, respectively, to produce the corresponding products which are otherwise difficult to obtain. In addition, the C–C triple bond can be oxidized under oxidative conditions with RuCl3/NaIO4, and (±)-4a can be smoothly transformed into the trifluoromethylated triketone (±)-7 in 65% yield.18 With a large excess amount of reducing agent LiAlH4, the carbonyl group and the ester group, together with the C–C triple bond, can be unexpectedly reduced simultaneously, affording the alkenyl diol (±)-8 in excellent regioselectivity. The hydrolysis process under basic conditions provided a reliable method for access to a free carboxylic acid (±)-9. Interestingly, when the reaction was performed under milder conditions compared to those for the synthesis of (±)-8, (±)-4a was successfully converted into an alkynyl diol (±)-10, which can be cyclized into a spiro compound (±)-11 (ref. 19) and an endocyclic compound (±)-12,20 respectively. Notably, in the majority of these cases, the excellent diastereoselectivity was reserved. These synthetic applications can demonstrate the significant value of this method.Open in a separate windowFig. 4Synthetic transformations.In order to gain some mechanistic insights into this radical cascade reaction, subsequent efforts have been made (Fig. 5). First, the detection of trifluoromethylated toluene (with toluene as the solvent, Fig. 5a, see ESI for details). Second, we were curious about the excellent diastereoselectivity associated with the use of TFE as the solvent. As can be seen in Fig. 5b, 1H NMR titration of 1a with increasing amounts of TFE showed a chemical shift of the resonance signal corresponding to protons. The 2D NOESY spectrum indicates the existence of an interaction between 1a and TFE (Fig. 5c). Moreover, Job plot studies by both 1H NMR and 19F NMR imply a 1 : 1.5 stoichiometry of the complex adduct resulting from 1a and TFE (Fig. 5d). These mechanistic studies strongly suggest that the excellent diastereoselectivity of this reaction might be attributed to the hydrogen bonding between TFE and the α-vinyl-β-ketoester.Open in a separate windowFig. 5Mechanism studies. (a) Radical probe; (b) 1H NMR titration; (c) 2D NOESY; (d) Job plot studies.On the other hand, density functional theory (DFT) calculations have also been performed at the B3LYP-D3(SMD)/Def2-TZVP//B3LYP-D3(SMD)/Def2-SVP level of theory in the TFE solvent model to further investigate the reaction pathways (Fig. 6). On the basis of the experimental results, herein, the radical pathway was considered. Initially, the CF3 radical addition onto 1a was calculated, and a transition state, TS1, was located with a free energy barrier of 10.9 kcal mol−1 to deliver the radical intermediate int1 with an exergonicity of 20.5 kcal mol−1. Then, a bicyclic transition state, TS2,21 with a barrier of 11.0 kcal mol−1 through a concerted 1,2-shift route was found to be the lower barrier TS for int2 formation than that of the addition to 2b for the byproduct (see Fig. S5 in ESI), which is consistent with the experimental results of the mainly hexacyclic products. Moreover, the intrinsic reaction coordinate (IRC) calculations and the root mean square (RMS) gradient of the potential energy surface from TS2 suggested that no transition state for the formation of the previously proposed strained alkoxyl radical was found. Next, the radical intermediate int2 attacking 2b was calculated. To understand the diastereoselectivity of this step, the transition states of the addition of 2b onto the Re and Si faces of C3 in int2 were located with barriers of 12.5 and 17.4 kcal mol−1 (TS3 and TS3′), respectively. It is noteworthy that the torsion angle of C1–C2–C3–C4 in TS3′ is −62.3°, larger than that of −40.9° in int2 and −49.0° in TS3, indicating that the distortion factor in TS3′ is large due to the steric effect from the trifluoroethyl group in int2 and, therefore, increases the barrier. The transition states of 2b addition were also optimized in solvents DCE and EA, and the free energy barrier differences between TS3 and TS3′ [ΔG = G(TS3′) − G(TS3)] are 3.6 and 3.0 kcal mol−1, respectively, in agreement with the experimental observations. Finally, dissociation of a SO2 molecule with a CF3 radical from int3 to deliver the product was conducted, and a transition state TS4 with a much lower barrier of only 7.1 kcal mol−1 was located, which led to the major product (±)-4a with a relative free enthalpy of −51.6 kcal mol−1.Open in a separate windowFig. 6Gibbs free energy profile for the synthesis of 4a in the TFE solvent model.  相似文献   

20.
HydroFlipper membrane tension probes: imaging membrane hydration and mechanical compression simultaneously in living cells     
Jos García-Calvo  Javier Lpez-Andarias  Jimmy Maillard  Vincent Mercier  Chlo Roffay  Aurlien Roux  Alexandre Fürstenberg  Naomi Sakai  Stefan Matile 《Chemical science》2022,13(7):2086
HydroFlippers are introduced as the first fluorescent membrane tension probes that report simultaneously on membrane compression and hydration. The probe design is centered around a sensing cycle that couples the mechanical planarization of twisted push–pull fluorophores with the dynamic covalent hydration of their exocyclic acceptor. In FLIM images of living cells, tension-induced deplanarization is reported as a decrease in fluorescence lifetime of the dehydrated mechanophore. Membrane hydration is reported as the ratio of the photon counts associated to the hydrated and dehydrated mechanophores in reconvoluted lifetime frequency histograms. Trends for tension-induced decompression and hydration of cellular membranes of interest (MOIs) covering plasma membrane, lysosomes, mitochondria, ER, and Golgi are found not to be the same. Tension-induced changes in mechanical compression are rather independent of the nature of the MOI, while the responsiveness to changes in hydration are highly dependent on the intrinsic order of the MOI. These results confirm the mechanical planarization of push–pull probes in the ground state as most robust mechanism to routinely image membrane tension in living cells, while the availability of simultaneous information on membrane hydration will open new perspectives in mechanobiology.

HydroFlippers respond to membrane compression and hydration in the same fluorescence lifetime imaging microscopy histogram: the responses do not correlate.

The detection and study of membrane mechanics in living cells is a topic of current concern.1–14 To enable this research, appropriate chemistry tools, that is small-molecule fluorescent probes that allow imaging of membrane tension, are needed.15 With the direct imaging of physical forces being intrinsically impossible, design strategies toward such probes have to focus on the suprastructural changes caused by changes in membrane tension.15 These suprastructural changes are divers, often interconnected, and vary with the composition of the membrane.15–25 Beyond the fundamental lipid compression and decompression, they include changes in membrane curvature, from rippling, buckling and budding to tubules extending from the membrane and excess lipid being ejected. Of similar importance are changes in membrane organization, particularly tension-induced phase separation and mixing, i.e. assembly and disassembly of microdomains. Consequences of these suprastructural changes include microdomain strengthening and softening and changes in membrane hydration and viscosity.16–25The currently most developed fluorescent flipper probes have been introduced26,27 to image membrane tension by responding to a combination of mechanical compression and microdomain assembly in equilibrium in the ground state.15 Extensive studies, including computational simulations,28 have shown that flipper probes align non-invasively along the lipid tails of one leaflet and report changes in membrane order and tension as changes in fluorescent lifetimes and shifts of excitation maxima.15 Among other candidates, solvatochromic probes respond off-equilibrium in the excited state to changes in membrane hydration and have very recently been considered for the imaging of membrane tension in living cells.29–36 So far not considered to image tension, ESIPT probes also report off equilibrium in the excited state on membrane hydration, but for different reasons.37,38 Mechanosensitive molecular rotors respond off equilibrium in the excited state to changes in microviscosity.17,30,32,39–53 The same principle holds for the planarization of bent, papillon or flapping fluorophores.54–57 The response of all possible probes to tension can further include less desired changes in positioning and partitioning between different domains, not to speak of more catastrophic probe aggregation, precipitation, disturbance of the surrounding membrane structure, and so on. Although the imaging of membrane tension is conceivable in principle with most of above approaches, the complex combination of parameters that has to be in place can thus far only be identified empirically, followed by much optimization.15The force-induced suprastructural changes are accompanied by the alteration in several unrelated physical properties of membranes. It is, for instance, well documented that membrane hydration increases with membrane disorder, from solid-ordered (So) to liquid-disordered (Ld) phases.29,58 Increasing cholesterol content decreases membrane hydration in solid- and liquid-ordered membranes.59 However, studies in model membranes also indicate that membrane hydration and membrane fluidity do not necessarily correlate.59 The dissection of the individual parameters contributing to the response of fluorescent membrane tension probes would be important for probe design and understanding of their responses, but it remains a daunting challenge. In this study, we introduce fluorescent flipper probes that simultaneously report on mechanical membrane compression and membrane hydration at equilibrium in the ground state. Changes of both in response to changes in membrane tension and membrane composition are determined in various organelles in living cells.The dual hydration and membrane tension probes are referred to as HydroFlippers to highlight the newly added responsiveness to membrane hydration. The mechanosensing of lipid compression in bilayer membranes by flipper probes has been explored extensively.15 Fluorescent flippers27 like 1 are designed as bioinspired60 planarizable push–pull probes26 (Fig. 1). They are constructed from two dithienothiophene fluorophores that are twisted out of co-planarity by repulsion of methyls and σ holes on sulfurs61,62 next to the twistable bond. The push–pull system is constructed first from formal sulfide and sulfone redox bridges in the two twisted dithienothiophenes. These endocyclic donors and acceptors are supported by exocyclic ones, here a trifluoroketone acceptor and a triazole donor.63 To assure stability, these endo- and exocyclic donors are turned off in the twisted ground state because of chalcogen bonding and repulsion, respectively.62Open in a separate windowFig. 1The dual sensing cycle of HydroFlippers 1–5, made to target the indicated MOIs in living cells and responding to membrane compression by planarization and to membrane hydration by dynamic covalent ketone hydration. With indication of excitation maxima (ref. 63) and fluorescence lifetimes (this study).Mechanical planarization of the flipper probe establishes conjugation along the push–pull systems, electrons flow from endocyclic donors to acceptors, which turns on the exocyclic donors and acceptors to finalize the push–pull system.62 This elaborate, chalcogen-bonding cascade switch has been described elsewhere in detail, including high-level computational simulations.62 The planar high-energy conformer 1dp excels with red shifted excitation and increased quantum yield and lifetime compared to the twisted conformer 1dt because the less twisted Franck-Condon state favors emission through planar intramolecular charge transfer (PICT) over non-radiative decay through twisted ICT, or conical intersections.15Flipper probe 1 was considered for dual responsiveness to membrane tension and hydration because of the trifluoroketone acceptor.63 Dynamic covalent hydration of 1dt yields hydrate 1ht.64–76 Blue-shifted excitation and short lifetime of 1ht are not expected to improve much upon planarization because the hydrate is a poor acceptor and thus, the push–pull system in 1hp is weak. The dynamic covalent chemistry of the trifluoroketone acceptor has been characterized in detail in solution and in lipid bilayer membranes.63To explore dual responsiveness to membrane tension in any membrane of interest (MOI) in living cells, HydroFlippers 2–5 were synthesized. While HydroFlipper 1 targets the plasma membrane (PM), HydroFlippers 2–4 were equipped with empirical targeting motifs.77 HydroFlipper 5 terminates with a chloroalkane to react with the self-labeling HaloTag protein, which can be expressed in essentially any MOI.78 Their substantial multistep synthesis was realized by adapting reported procedures (Schemes S1–S4).The MOIs labeling selectivity of HydroFlippers was determined in HeLa Kyoto (HK) cells by confocal laser scanning microscopy. Co-localization experiments of flippers 1–4 with the corresponding trackers gave Pearson correlation coefficients (PCCs) >0.80 for the targeting of mitochondria, lysosomes and the endoplasmic reticulum (ER, Fig. S4–S6). HydroFlipper 5 was first tested with stable HGM cells, which express both HaloTag and GFP on mitochondria (referred to as 5M).78,79 The well-established chloroalkane penetration assay demonstrated the efficient labeling of HaloTag protein by 5 as previously reported HaloFlippers (Fig. S3).78 By transient transfection, HydroFlippers 5 were also directed to lysosomes (5L), Golgi apparatus (GA, 5G)80 and peroxisomes (5P) with HaloTag and GFP expressed on their surface.78 PCCs >0.80 for co-localization of flipper and GFP emission confirmed that MOI labeling with genetically engineered cells was as efficient as with empirical trackers (Fig. S7–S11).Dual imaging of membrane compression and hydration was envisioned by analysis of fluorescence lifetime imaging microscopy (FLIM) images using a triexponential model (Fig. 2).81 FLIM images of ER HydroFlipper 4 in iso-osmotic HK cells were selected to illustrate the concept (Fig. 3a). Contrary to classical flipper probes, the fluorescence decay curve of the total FLIM image (Fig. 2a, grey) showed a poor fit to a biexponential model (Fig. 2a, cyan, b). Consistent with their expected dual sensing mode, a triexponential fit was excellent (Fig. 2a, dark blue, c). Lifetimes τ1i = 4.3 ns () were obtained besides background. This three-component model was then applied to every pixel of FLIM images (Fig. 3c). The resulting reconvoluted FLIM histogram revealed three clearly separated populations for τ1 (red), τ2 (green), and background (τ3, blue, Fig. 2d). Maxima of these three clear peaks were at the lifetimes estimated by triexponential fit of the global decay curve, thus demonstrating the validity of the methodology at necessarily small photon counts. Irreproducible fitting would give randomly scattered data without separated peaks.Open in a separate windowFig. 2(a) Fluorescence decay curve (grey, corresponding to the total image, not to a single pixel) with biexponential (cyan) and triexponential fit (dark blue). (b, c) Residual plots for bi- (b) and triexponential fit (c). (d) Histogram with the intensities associated with the τ1 (red), τ2 (green), and τ3 (blue, background) components obtained by triexponential fit of the fluorescence decay curve of each pixel of the FLIM image, fit to Gaussian function (black solid curves).Open in a separate windowFig. 3FLIM images of HK cells labelled with ER flipper 4 before (a, c) and after (b, d) hyper-osmotic shock, showing average lifetimes τav (a, b) and τ1 (c, d) from triexponential reconvolution; scale bars = 10 μm. (e) Distribution of the photon counts associated with the τ1 component of 4 in HK cells after triexponential reconvolution of FLIM images before (c, τ1i) and after (d, τ1h) hyper-osmotic shock, showing decreasing lifetimes for τ1 (4d). (f) The dehydration factor dhi defined as total integrated photon counts for τ1τ1) divided by Στ2 (i.e., dhi = area Στ1i/area Στ2i) for 4 in strongly hydrated ER (dhi < 2, turquoise) and 1 in weakly hydrated plasma membrane (dhi > 6, purple) of HK Kyoto cells under iso-osmotic conditions.Dual response of HydroFlippers to changes in membrane tensiona
ProbebdhicdhhdΔdhe (%) τ 1i f (ns) τ 1h g (ns)Δτ1h (%)
11 (PM)6.36.5-34.84.48
21 (-C)i6.18j4.83k
32 (Lyso)2.92.844.44.010
43 (Mito)2.31.9174.44.08
54 (ER)1.81.5174.33.715
64 (–C)i1.139l4.110m
75G (GA)n2.52.384.23.810
85E (ER)o1.71.2293.83.75
91 (Lo)p115.2
101 (Ld)q1.23.4
Open in a separate windowaFrom triexponential fit of FLIM images in HK cells (errors, see ESI).bFlipper (target MOI).cdhi = area Στ1i/area Στ2i in FLIM histogram under iso-osmotic (i) conditions (e.g.Fig. 3f).ddhh = area Στ1h/area Στ2h in FLIM histogram under hyper-osmotic (h) conditions.eFlipper hydration change in response to membrane tension: Δdh = (1 – dhh/dhi) × 100%.fFluorescence lifetime value of the slowest component from the fitted fluorescence decay under iso-osmotic (i) conditions (e.g.Fig. 2d).gSame as f, under hyper-osmotic (h) conditions.hFlipper planarization in response to membrane tension: Δτ1 = (1 – τ1h/τ1i) × 100%.iMeasured after cholesterol (C) removal from cells with MβCD.jCompared to dhi of 1 (6.6) in untreated cells measured on the same day.kCompared to τih of 1 (5.0) in untreated cells measured on the same day.lAs j using 4 and compared to dhi = 1.8.mAs k using 4 compared to τih = 4.5.nMeasured in transiently transfected HK cells with ST-HaloTag-HA expressed inside GA.80oMeasured in transiently transfected HK cells with HaloTag-Sec61B expressed inside ER.78pMeasured in SM/C GUVs.qMeasured in DOPC GUVs.Extensive lifetime data for monofunctional flipper probes supported that the intensities associated to τ1i (i for iso-osmotic, see below) originate from at least partially planarized flippers 4d in the ER (Fig. 2d, red, 3c, 1). The population of the τ2i component in the reconvoluted FLIM histogram was attributed to the presence of hydrated 4h in the ER (Fig. 2d, green, 1). This assignment was consistent with lifetime differences in solution between τ = 2.7 ns for the dehydrated and τ = 0.7 ns for the hydrated form of a hydrophobic flipper analog in dioxane-water mixtures (Fig. S2), and model studies in GUVs (see below).63The ratio between the τ1i (red) and τ2i (green) populations in the reconvoluted FLIM histogram was used to extract a quantitative measure for hydration of the MOI (Fig. 2d, ,3f).3f). A dehydration factor dh was defined by dividing the total integrated counts for τ1τ1) by Στ2. For 4 in iso-osmotic ER, dhi = 1.8 ± 0.1 was obtained (Fig. 3f, 63 Thus, these results implied that the dehydration factor dh obtained from reconvoluted triexponential FLIM images reports quantitatively on membrane hydration, that is the local water concentration around HydroFlippers in their MOI.In uniform model membranes composed of only one lipid, flipper probes like 1 respond to increasing membrane tension with decreasing lifetimes.15,18 This response can be explained by flipper deplanarization upon lipid decompression. In the mixed membranes composed of different lipids, flipper probes reliably respond to increasing membrane tension with increasing lifetimes, and lifetime changes can be calibrated quantitatively to the applied physical force.18,77 This indicates that in these biologically relevant membranes, the response is dominated by factors other than lipid decompression. Tension-induced microdomain formation is confirmed to account for, or at least contribute to, increasing lifetimes with increasing tension, or membrane decompression.15,18 Not only microdomain disassembly but also changes in membrane curvature from rippling, budding and microdomain softening to tube formation and lipid ejection combine to afford decreasing lifetimes with membrane compression, or decreasing tension.17,18Membrane tension was applied to the ER by extracellular hyper-osmotic stress. This causes membrane tension to decrease, i.e., membrane compression to increase.18,77 Consistent with tension-induced deplanarization from 4p to 4t (Fig. 1), lifetimes of 4 visibly decreased in response to decreasing membrane tension (Fig. 3b). The reconvoluted FLIM histogram clearly shows that compression caused the decrease of τ1 of 4 in the ER from τ1i = 4.3 ns to τ1h = 3.7 ns, whereas τ2i = 1.5 ns was less mechanosensitive (τ2h = 1.4 ns, Fig. 3e, 4a–c). These different mechanosensitivities were meaningful considering that in three-component histograms, τ1 originates from dehydrated HydroFlipper 4d that loses a strong push–pull dipole and thus shortens lifetime upon tension-induced deplanarization from 4dp to 4dt (Fig. 1). In contrast, hydrated HydroFlipper 4h accounting for τ2 lacks a strong dipole and thus features short lifetimes with poor sensitivity for tension-induced deplanarization from 4hp and 4ht. This result was consistent with the central importance of turn-on push–pull systems for flipper probes to function as mechanosensitive planarizable push–pull probes.81Open in a separate windowFig. 4(a) Reconvoluted FLIM histograms for 1–5 obtained by fitting each pixel of the FLIM image to a three-exponential model under iso-osmotic (top) and hyper-osmotic (bottom) conditions in HK cells; *dhi analysis in Fig. 3f; **Δτ1 analysis in Fig. 3e. (b–e) Trend plots for membrane compression (τ1) and hydration (dh) for 1–5 in HK cells without (b, e) and in response to hyper-osmotic membrane tension (c–e). (b) τ1i (iso-osmotic compression) vs. dhi (iso-osmotic hydration). (c) τ1iτ1hvs. τ2iτ2h (compression response in ns). (d) Δτ1 (compression response, %) vs. Δdh (hydration response, %), (e) Δτ1 and Δdh upon compression (σ) and cholesterol depletion (C). #Discontinuous, see 17,18The uniform response of HydroFlipper planarization and hydration thus provided corroborative support that membrane deformation and reorganization dominate the fluorescence imaging of membrane tension under the condition that the probe partitions equally between different phases.63 However, the dual response HydroFlipper dissects the consequences of these tension-induced suprastructural changes. HydroFlipper planarization 4t/4p detected by τ1 reports on lipid compression in the local environment in the MOI. HydroFlipper hydration 4d/4h detected by the dehydration factor dh reports on local membrane hydration. Pertinent reports from model membranes in the literature indicate that the two do not have to be the same.59To elaborate on these implications, FLIM images were recorded for all HydroFlippers 1–5 in their respective MOIs before and after the application of hyper-osmotic stress and then analyzed using the three-component model (Fig. 4a, Fig. 4a) and estimated by global triexponential fit (Fig. 3f, ,4a).4a). However, these changes do not affect dhi, which compares areas rather than maxima in the histograms.Trends for membrane hydration and compression reported by dhi and τ1i, respectively, should reflect the overall composition and thus nature of the different membranes. For PM 1, Lyso 2, GA 5G and ER 5E, coinciding trends were found for hydration (dhi, blue) and compression (τ1i, red, Fig. 4b). Hydration and deplanarization increased in parallel, consistent with increasingly disordered membranes. With Mito 3 and ER 4, increasing hydration (blue) was not reflected in increasing deplanarization (red, Fig. 4b).For the comprehensive analysis of the changes caused by hyper-osmotic stress, the differences in lifetimes for τ1 and τ2 were clarified first. Whereas τ1iτ1h values (red) around 0.3 ns were large and significant in all MOIs, τ2iτ2h values (pink) were negligible (Fig. 4c). The mechano-insensitive τ2, corresponding to hydrate 4h, were thus not further considered as a valid measure of membrane compression.To facilitate direct comparability, membrane compression Δτ1 and membrane dehydration Δdh in response to hyper-osmotic stress were converted in percentage of decrease (positive) or increase (negative) from the value under iso-osmotic conditions (Fig. 4d, Fig. 4d, red). In clear contrast, dehydration Δdh varied from 3% increase to 29% decrease (Fig. 4d, blue). The most extreme deviations concerned ER probes with maximal Δτ1 responsiveness for tracker 4 and minimal Δτ1 responsiveness for Halo flipper 5E. For dehydration Δdh, both probes showed high responsiveness. These extremes could reflect the diverse membrane properties of the ER, with τ = 4.1, 3.5 and 3.4 ns reported previously for different flipper mechanophores in tubular, sheet, and nuclear membranes of COS7 cells, respectively.15,77 Although less resolvable in HK cells, this heterogeneity of ER membranes is also visible in the FLIM images with 4 (Fig. 3). Tracker 4 and Halo flipper 5E both react covalently with membrane proteins and report on the respective surrounding ER membrane, which differs significantly according to the two HydroFlipper probes. The extreme values for Halo flipper 5E suggested that other factors like fractions of mispositioned flipper in more hydrophilic environment could also contribute to the global outcome (Fig. 4b, Fig. 4d, blue) increased with membranes disorder characterized by shorter τ1i and low dhi (Fig. 4b), while Δτ1 remained more constant until the possible onset of decreases at very high hydration (5E, Fig. 4d, red). Both observations - independence of mechanical flipper planarization and dependence of dynamic covalent hydrate formation on the water concentration in the surrounding membrane - were chemically meaningful.The validity of these conclusions was tested by removing cholesterol with methyl-β-cyclodextrin (MβCD). As expected for the increased hydration level and decreased order of cholesterol depleted membranes, Δdh and Δτ1 of 1 and 4 increased by MβCD treatment compared to those obtained on the same day without the treatment (Fig. 4e, C). Stronger response of ER HydroFlipper 4 to the cholesterol removal can be attributed to the poorer cholesterol content in ER membranes than in PM.82 Consistent with the overall trend, Δdh was more significantly affected by changes of the MOI by MβCD treatment than by tension change (Fig. 4e, blue, C vs. σ), while Δτ1 responded better to membrane tension than MOI change (Fig. 4e, red, C vs. σ).Taken together, these results reveal HydroFlippers as first dual mode fluorescent membrane tension probe, reporting on membrane hydration and membrane compression at the same time. Mechanical compression is reported as shift in τ, while tension-induced hydration is reported as change in relative photon counts for hydrated and dehydrated probes in the reconvoluted FLIM histograms. The response of flipper deplanarization to membrane tension is robust and less dependent on the nature of the MOI, including plasma membrane, ER, mitochondria, lysosomes and Golgi. In contrast, the responsiveness of flipper hydration to membrane tension depends strongly on the nature of the MOI, generally increasing with increasing intrinsic disorder, that is hydration, already under iso-osmotic conditions. These results validate the flipper probes as most reliable to routinely image membrane tension in cells, while the simultaneous information provided on membrane dehydration provides attractive possibilities for biological applications.  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