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 109 毫秒
1.
Metallosurfactants are molecular compounds which combine the unique features of amphiphiles, like their capability of self-organization, with the peculiar properties of metal complexes like magnetism and a rich redox chemistry. Considering the high relevance of surfactants in industry and science, amphiphiles that change their properties on applying an external trigger are highly desirable. A special feature of the surfactant reported here, 1-(Z)-heptenyl-1′-dimethylammonium-methyl-(3-sulfopropyl)ferrocene (6), is that the redox-active ferrocene constituent is in a gemini-position. Oxidation to 6+ induces a drastic change of the surfactant''s properties accompanied by the emergence of paramagnetism. The effects of an external magnetic field on vesicles formed by 6+ and the associated dynamics were monitored in situ using a custom-made optical birefringence and dual dynamic light scattering setup. This allowed us to observe the optical anisotropy as well as the anisotropy of the diffusion coefficient and revealed the field-induced formation of oriented string-of-pearls-like aggregates and their delayed disappearance after the field is switched off.

The self-organization properties of a stimuli responsive amphiphile can be altered by subjecting the paramagnetic oxidized form to a magnetic field of 0.8 T and monitored in real time by coupling optical birefringence with dynamic light scattering.

Amphiphiles (or surfactants) combine hydrophilic (the so-called headgroups) and lipophilic entities (the so-called tails) as integral parts of their molecular structures. This particular construction principle provides them with the ability to display concentration-dependent self-organization in nonpolar and polar solvents.1 Amphiphiles with advanced functions that go far beyond the traditional ones as emulsifiers, stabilizing agents for interfaces, or detergents were meanwhile realized by skillful manipulation of any of its constituents.2–4 Recent examples are micellar LEDs,5,6 catalysts,7–9 or batteries.10 Such applications are important hallmarks on the way to even more sophisticated amphiphiles such as the ones found in nature, e.g. in the pockets of enzymes.11–18 An important milestone is the advent of (multi-) stimuli-responsive amphiphiles, whose encoded functionalities respond to (different) external triggers. Such systems are capable of adaptive self-assembly, which can be controlled using an external input such as the pH, temperature, ionic strength, or redox state.19–26Paramagnetic amphiphiles, recently reviewed by Eastoe and coworkers, constitute a fascinating family of stimuli-responsive surfactants.27 Particular attention has been paid to magnetic ionic liquids based on amphiphilic transition metal complexes, as their properties are often superior to those of conventional magnetic fluids (ferrofluids).28–31 Self-assembly results in high effective concentrations of the paramagnetic metal centers, and this in turn allows us to control their physico-chemical properties and the morphologies of their superstructures through an external magnetic field. Such a scheme has the added advantage that the external stimulus is non-invasive. In many current realizations of such systems, however, the magneto-active (transition) metal ion is only present as a constituent of the counterion of a cationic surfactant, but is not an integral constituent of the surfactant itself.21,30,31Some of us have previously reported redox-switchable as well as paramagnetic stimuli-responsive amphiphiles of relevance to the current work.32,33 We thought that ferrocene would be an ideal building block in order to combine both these kinds of stimuli within one single amphiphile.34–37 On oxidation, the diamagnetic, hydrophobic ferrocene nucleus is transformed into a paramagnetic S = 1/2 ferrocenium ion with a distinct hydrophilic character.38–41 Oxidation does hence not only generate a magnetic moment, but also transfers the ferrocene nucleus from the lipo- to the hydrophilic part of the amphiphile, thereby changing its entire structure. A 1,1′-disubstitution pattern of the ferrocene scaffold, which is synthetically well accessible,34,42–44 seemed particularly suited for such an endeavor.Studies on paramagnetic amphiphiles are often thwarted by the non-trivial analytics involved in their characterization. Detailed investigations often rely on small-angle neutron scattering (SANS), which is time-consuming and costly and suffers from poor availability.27,30,31,45–47 Moreover, SANS is only of limited value for following kinetically fast processes which would be desirable for the live monitoring of structural changes occurring in solution. Optical birefringence is a well-established method to monitor the dynamic response of materials to external fields.48–50 Although of high intrinsic value, optical birefringence measurements in magnetic fields were only rarely applied for the study of paramagnetic amphiphiles.29We here report the zwitterionic, ferrocene-based amphiphile FcNMe2SO3Heptene 6 (see Fig. 1, Fc = ferrocenyl) with a sultone headgroup. Compound 6 is unique in that its self-assembly properties can be controlled by three different external stimuli, namely the (i) addition of an electrolyte, (ii) addition of an oxidant/reductant, and (iii) exposure to an external magnetic field. We also demonstrate that optical birefringence in combination with dynamic light scattering (DLS) measurements in two orthogonal directions provides detailed insights into the functional response of aggregated magnetic nanoparticles formed by 6+ to an external magnetic field in real time. Specifically, we have observed the formation of string-of-pearls-like aggregates of 6+ in a magnetic field (0.8 T), the field-induced anisotropy of the diffusion of aggregated nanoparticles, and a hysteresis effect for their disappearance after the magnetic field is switched off. Thus, the anisotropy of larger aggregates persists for more than 5 min, while the structural alignment of smaller ones vanishes at a significantly faster rate.Open in a separate windowFig. 1Synthesis of FcNMe2SO3Heptene (6). (a) Synthesis of 6; (b) molecular structure of 6 crystallized from acetonitrile. C; dark grey, N; turquoise, Fe; orange, S; yellow, O; red, H atoms are omitted for clarity.  相似文献   

2.
Multicolor carbon dots (CDs) have been developed recently and demonstrate great potential in bio-imaging, sensing, and LEDs. However, the fluorescence mechanism of their tunable colors is still under debate, and efficient separation methods are still challenging. Herein, we synthesized multicolor polymeric CDs through solvothermal treatment of citric acid and urea in formamide. Automated reversed-phase column separation was used to achieve fractions with distinct colors, including blue, cyan, green, yellow, orange and red. This work explores the physicochemical properties and fluorescence origins of the red, green, and blue fractions in depth with combined experimental and computational methods. Three dominant fluorescence mechanism hypotheses were evaluated by comparing time-dependent density functional theory and molecular dynamics calculation results to measured characteristics. We find that blue fluorescence likely comes from embedded small molecules trapped in carbonaceous cages, while pyrene analogs are the most likely origin for emission at other wavelengths, especially in the red. Also important, upon interaction with live cells, different CD color fractions are trafficked to different sub-cellular locations. Super-resolution imaging shows that the blue CDs were found in a variety of organelles, such as mitochondria and lysosomes, while the red CDs were primarily localized in lysosomes. These findings significantly advance our understanding of the photoluminescence mechanism of multicolor CDs and help to guide future design and applications of these promising nanomaterials.

Understanding the origin and sensitivity of carbon dot emission will improve their utility in various applications.

Since the accidental discovery of luminescent carbon fragments in 2004,1 carbon dots (CDs) have attracted great research interest due to the diverse synthetic methods, tunable luminescence, and applicability in a broad range of fields, including bio-imaging,2–4 sensing,5,6 and light emitting diodes (LEDs).7,8 Typically, CDs are fluorescent carbon nanostructures of sizes less than 10 nm, composed of carbon, oxygen, and nitrogen.9–12 CDs can be produced through bottom-up methods, which involve small molecular precursors like citric acid, malic acid, urea, ethylenediamine, and so on.13–15 In a high temperature reaction, polymerization and dehydration occur among various functional groups, and the resulting products are usually a mixture of small molecule residues, oligomers, and long chain polymers.16 The unclear fluorescence mechanisms and poorly understood internal structure of CDs limit the ability to understand, tune, and fully exploit their fluorescence properties.Fortunately, in recent years, breakthrough syntheses of multicolor CDs have been achieved.17–19 Several different multicolor CDs have been synthesized with aromatic compounds such as phenylenediamine.4,20–22 However, it should be noted that precursors such as aniline and phenol may have toxic effects on human health and the environment,23,24 and thus should be avoided where possible. Syntheses of colorful CDs from non-aromatic compounds such as citric acid and urea often employ solvothermal methods. Utilizing different solvents such as formamide and dimethylformamide have been shown to play a significant role in tuning CD emission.25,26 In addition, chromatographic post-treatment of as-made CDs plays a critical role in obtaining different colored fractions, using techniques such as anion-exchange column chromatography,26 normal phase silica chromatography,27 and reversed phase silica chromatography.15 Compared with high performance liquid chromatography (HPLC), the aforementioned column chromatography techniques help to separate CDs on a larger scale. These separations are based on charge26 or polarity,21 and are efficient in isolating the desired fractions with distinct colors so that detailed structural characterization can be performed.To gain insight into the fluorescence mechanism of these multicolor CDs, researchers have considered three hypotheses: quantum size effects,28 the inclusion of molecular fluorophores,29 and surface state-induced emission.30 For example, Rogach and coworkers developed solid-state CDs with tunable fluorescence via the seeded growth method. They attributed the tunable emission to the size of π-conjugated domains.31 Yang and coworkers synthesized CDs by hydrothermal treatment of citric acid and ethylenediamine. They identified a small molecule fluorophore, IPCA (1,2,3,5-tetrahydro-5-oxo-imidazo[1,2-α]pyridiine-7-carboxylic acid) from CD column separation fractions, which contributed to the blue fluorescence.13 Xiong and coworkers synthesized CDs from urea and p-phenylenediamine that emitted a range of colors and separated them with silica column chromatography. They found the degree of carbon oxidation increased as the emission redshifted and thus, they endorsed the surface state hypothesis.21 In addition to the above mechanisms, computational methods such as density functional theory (DFT) have also been applied to analyze the fluorescence origins of CDs. The charge transfer between functional groups on the polymeric unit of CDs made from citric acid and ethylenediamine was found to facilitate blue emission.16The goal of present work is to understand the fluorescence origin of multicolor CDs. The model multicolor CDs were obtained by reacting citric acid and urea in formamide via a microwave-assisted hydrothermal treatment. An automated chromatographic apparatus was employed to separate as-made CD mixtures into distinct color fractions. The individual separation process took around 20 minutes, and the obtained CD fractions exhibit discrete illumination-induced emission throughout the visible region of the spectrum. Interestingly, the sizes of separated CD fractions are not statistically different from one another, suggesting that the quantum size effects are not the source of differential emission. Solvatochromism experiments showed that the blue and green fractions have similar fluorescence behavior as a function of solvent polarity, but the red fraction behaved differently. Using computational simulations, three models of the fluorescence origin were constructed and evaluated, showing that the formation of small blue fluorescent molecules is likely and pyrene analogs could be the origins for various emission colors. Moreover, two representative CD fractions, the blue- and red-emitting fractions, were chosen for subsequent cell imaging experiments. The localization pattern for the CD fractions differed: blue-emitting CDs were observed in a wide range of organelles, while red-emitting CDs were primarily enclosed in lysosomes. Understanding the origin and the sensitivity of CD emission will improve their utility in bioimaging applications.  相似文献   

3.
4.
Practically important metal electrodes are usually polycrystalline, comprising surface grains of many different crystallographic orientations, as well as grain boundaries. In this study, scanning electrochemical cell microscopy (SECCM) is applied in tandem with co-located electron backscattered diffraction (EBSD) to give a holistic view of the relationship between the surface structure and the electrochemical activity and corrosion susceptibility of polycrystalline Cu. An unusual aqueous nanodroplet/oil (dodecane)/metal three-phase configuration is employed, which opens up new prospects for fundamental studies of multiphase electrochemical systems, and mimics the environment of corrosion in certain industrial and automotive applications. In this configuration, the nanodroplet formed at the end of the SECCM probe (nanopipette) is surrounded by dodecane, which acts as a reservoir for oil-soluble species (e.g., O2) and can give rise to enhanced flux(es) across the immiscible liquid–liquid interface, as shown by finite element method (FEM) simulations. This unique three-phase configuration is used to fingerprint nanoscale corrosion in a nanodroplet cell, and to analyse the interrelationship between the Cu oxidation, Cu2+ deposition and oxygen reduction reaction (ORR) processes, together with nanoscale open circuit (corrosion) potential, in a grain-by-grain manner. Complex patterns of surface reactivity highlight the important role of grains of high-index orientation and microscopic surface defects (e.g., microscratches) in modulating the corrosion-properties of polycrystalline Cu. This work provides a roadmap for in-depth surface structure–function studies in (electro)materials science and highlights how small variations in surface structure (e.g., crystallographic orientation) can give rise to large differences in nanoscale reactivity.

Probing Cu corrosion in an aqueous nanodroplet/oil/metal three-phase environment revealed unique patterns of surface reactivity. The electrochemistry of high-index facets cannot be predicted simply from the low-index {001}, {011} and {111} responses.

Corrosion has long been studied, as a significant concern and a costly issue (ca. 3% of the GDP of industrialised countries) for the modern world.1,2 For metals, in particular, electrochemical techniques, allied to complementary analytical and microscopy methods, play a central role in unveiling corrosion and corrosion protection mechanisms.3–7 However, a limitation of many experimental approaches is that the electrochemical perturbation (and measurement) is applied globally at a macroscopic electrode immersed in a bulk solution,8 but most corrosion processes are initiated and perpetuated at (sub)microscopic surface sites (e.g., grain boundaries, inclusions, microscratches etc.).9–14 Mismatch between the scale of key corrosion phenomena and conventional electrochemical methods makes it difficult to unambiguously identify the key anodic/cathodic sites driving corrosion. This issue is compounded for the case of atmospheric corrosion,15 or corrosion in certain automotive/industrial environments (vide infra),16,17 which take place due to the action of small droplets on the surface in a confined system. Corrosion science needs electrochemical techniques that operate at the (sub)microscale, and allow activity and surface structure to be correlated commensurately at this scale.Among the limited library of electrochemical techniques that can routinely operate at the (sub)microscale,18,19 scanning electrochemical cell microscopy (SECCM) is attracting significant attention.20–22 SECCM maps electrochemistry locally and directly via a nanoscale electrochemical meniscus cell (formed at the end of a fluidic probe) that makes measurements over an array of points (typically thousands of discrete areas) on an electrode (or other) surface. For polycrystalline surfaces, SECCM measurements are powerfully combined with co-located electron backscattered diffraction (EBSD), to elucidate nanoscale structure–activity, as exemplified by studies of various electrochemical processes at a range of polycrystalline materials, including Pt,23–26 Au,27 Pd,28 low carbon steel,29–31 Zn32 and boron-doped diamond.33In addition to its high spatiotemporal resolution, the meniscus cell configuration of SECCM facilitates rapid reactant/product exchange with the surrounding environment, mimicking a gas diffusion electrode, with an enhanced flux of gases into the meniscus cell (i.e., at the three-phase boundary).24,27,34 When operated in air, SECCM emulates the configuration of atmospheric corrosion, with gas exchange (e.g., oxygen, O2) taking place across the liquid/gas interface of the meniscus in contact with a surface of interest. As recently reported, and expanded upon herein, SECCM can also be operated under oil immersion,32,35 which not only aids in confinement of the meniscus cell during prolonged measurements,35 but also opens up the possibility of studying the effect of oil-soluble species (e.g., corrosion inhibitors, organic contaminants, redox mediators etc.) on local reactions at the solid/liquid/liquid interface with high spatial-resolution. This configuration is regaining interest for fundamental studies,36,37 as well as being of great practical importance (e.g., phase-transfer reactions in industrial chemical processes, biology etc.).38A key attribute of SECCM is that a number of conventional dynamic electrochemistry techniques (e.g., potentiometry, amperometry and voltammetry) can be translated readily to the confines of the meniscus cell.20,22,39 Herein, the versatility of chronopotentiometry for local corrosion and electrochemical measurements is demonstrated. First, it is possible to make meniscus contact at zero applied current, corresponding to open circuit potential (OCP), which is measured. This corresponds to the corrosion (mixed) potential, where the rate of anodic dissolution of the metal (forming metal ions) and the rate of reduction of oxygen are balanced. Surface ion release under this condition is then analysed by subsequent “electrochemical titration” of a portion of the released metal ions, by applying a cathodic current and recording the resulting chronopotentiometric curve.40 This allows the evaluation of intrinsic corrosion susceptibility, in situ, with high spatial-resolution, for the entire range of crystallographic orientations of a polycrystalline metal (i.e., revealed through co-located EBSD analysis). Chronopotentiometry measurements with and without O2 present, and the use of an anodic pulse to induce the anodic dissolution (as well as the cathodic measurements mentioned) allow all of the key electrochemical processes underpinning localised corrosion to be studied. The patterns of surface reactivity establish the intimate link between corrosion susceptibility, electrochemical kinetics and surface structure at the nanoscale.  相似文献   

5.
Innovative and robust photosensitisation materials play a cardinal role in advancing the combined effort towards efficient solar energy harvesting. Here, we demonstrate the photocathode functionality of a Metal–Organic Framework (MOF) featuring cofacial pairs of photo- and electro-active 1,4,5,8-naphthalenediimide (NDI) ligands, which was successfully applied to markedly reduce the overpotential required for CO2 reduction to CO by a well-known rhenium molecular electrocatalyst. Reduction of [Cd(DPNDI)(TDC)]n (DPNDI = N,N′-di(4-pyridyl)-1,4,5,8-naphthalenediimide, H2TDC = thiophene-2,5-dicarboxylic acid) to its mixed-valence state induces through-space Intervalence Charge Transfer (IVCT) within cofacial DPNDI units. Irradiation of the mixed-valence MOF in the visible region generates a DPNDI photoexcited radical monoanion state, which is stabilised as a persistent species by the inherent IVCT interactions and has been rationalised using Density Functional Theory (DFT). This photoexcited radical monoanion state was able to undergo charge transfer (CT) reduction of the rhenium molecular electrocatalyst to effect CO generation at a lower overpotential than that required by the discrete electrocatalyst itself. The exploitation of cofacial MOFs opens new directions for the design philosophy behind light harvesting materials.

The photocathode functionality of a Metal–Organic Framework (MOF) featuring cofacial photo- and electro-active ligands provides a new approach to CO2 reduction via charge transfer with a rhenium electrocatalyst.

The development of photocathode materials for CO2 reduction and hydrogen evolution catalyses has traditionally focussed on photosensitising transition metal complexes or nanostructured solid state semiconductors.1,2 At the nascent frontier between robust solid state semiconductors and synthetically protean metal complexes are photo-/electro-active Metal–Organic Frameworks (MOFs) that consolidate the flexibility of homogeneous systems into the robust heterogeneous phase.3 Contrasting with reported MOF examples, natural photosynthesis remains one of the most efficient light harvesting systems.4 One common reaction centre adopted in photosynthesis features a redox-active cofacial dimer of chlorophyll pigment molecules.5 This cofacial moiety stabilises the photoexcited charge separated state through intra-dimer Intervalence Charge Transfer (IVCT) interactions, enabling the trapping and conversion of light to chemical energy. Recently, we characterised IVCT interactions upon reduction to the mixed-valence state in the MOF [Zn2(TDC)2(DPPTzTz)2]n (DPPTzTz = 2,5-bis(4-(4-pyridyl)phenyl)thiazolo[5,4-d]thiazole and H2TDC = thiophene-2,5-dicarboxylic acid) featuring cofacial dimers of the thiazolothiazole redox-active core, and probed its structure–activity dependence computationally and experimentally.6–9 Subsequently, we sought design a new MOF featuring cofacial pairs of the photo- and redox-active N,N′-di(4-pyridyl)-1,4,5,8-naphthalenediimide (DPNDI) ligand, as a conceptually neoteric photosensitiser for incorporation into systems relevant towards artificial photosynthesis.The naphthalene diimide (NDI) core was selected for its photoactive radical monoanion state.10 For a number of discrete systems, Wasielewski and coworkers have computationally and experimentally demonstrated the ability to photoexcite the easily accessible NDI radical monoanion using visible light, facilitating its transient photoelectrochemical reduction of Re based catalytic CO2 reduction sites.2,11–14 Recently, Goswami et al. synthesised a Zr NDI-based MOF, applying this as a radical state heterogeneous photosensitiser to decompose dichloromethane.15Here, we describe the synthesis of a new photo- and redox-active MOF [Cd(DPNDI)(TDC)]n, denoted csiMOF-6 (cofacial stacked IVCT), featuring cofacial dimers of the DPNDI ligand. Cofacial DPNDI MOFs have been reported previously by Takashima et al.16 and Sikdar et al.,17 where guest dependent charge transfer (CT) and neutral state photoexcitation behaviours were examined. Dinolfo et al. also incorporated DPNDI into a rhenium based cofacial complex, where its mixed-valence IVCT behaviour was probed using electrochemical and spectroelectrochemical (SEC) techniques.18 We envisaged that the cofacial NDI units in csiMOF-6 would stabilise its photoexcited radical monoanion state by IVCT interactions, akin to cofacial moieties in natural photosynthsesis processes. This strengthens the persistence of the NDI photoexcited radical monoanion state, thereby improving its efficacy at photoelectrochemical reduction of catalytically active sites. Effectiveness of the cofacial design principle behind csiMOF-6 photocathodes was verified using a combined experimental and computational approach. The successful photocathode performance of csiMOF-6 under broad band visible light irradiation encompassed its photoelectrochemical reduction of the [Re(bipy-tBu)(CO)3Cl] (bipy-tBu = 4,4′-di-tert-butyl-2,2′-bipyridine, developed by Smieja et al.19) CO2 reduction electrocatalyst, resulting in CO generation at reduced overpotential requirements.  相似文献   

6.
The biosynthetic gene cluster of the antifungal metabolite sporothriolide 1 was identified from three producing ascomycetes: Hypomontagnella monticulosa MUCL 54604, H. spongiphila CLL 205 and H. submonticulosa DAOMC 242471. A transformation protocol was established, and genes encoding a fatty acid synthase subunit and a citrate synthase were simultaneously knocked out which led to loss of sporothriolide and sporochartine production. In vitro reactions showed that the sporochartines are derived from non-enzymatic Diels–Alder cycloaddition of 1 and trienylfuranol A 7 during the fermentation and extraction process. Heterologous expression of the spo genes in Aspergillus oryzae then led to the production of intermediates and shunts and delineation of a new fungal biosynthetic pathway originating in fatty acid biosynthesis. Finally, a hydrolase was revealed by in vitro studies likely contributing towards self-resistance of the producer organism.

A new family of fungal biosynthetic pathways is elucidated based on the use of fatty acid and citrate-like intermediates.

Gamma-lactone and alkyl citrate compounds derived from oxaloacetate are widespread natural products in fungi and often possess potent biological activities. Examples include sporothriolide 1,1,2 piliformic acid 2,3 tyromycin 34 and the cyclic maleidrides including byssochlamic acid 45,6 among others (Fig. 1). In some cases, for example those of 4 and squalestatin S1 5,7 detailed molecular studies have revealed that a dedicated polyketide synthase (PKS) produces a carbon skeleton that is then condensed with oxaloacetate by a citrate synthase (CS) to give an early alkyl citrate intermediate that is further oxidatively processed. In other cases, such as 1 and the sporochartines 6, the biosynthetic pathways are not yet clear.Open in a separate windowFig. 1Structures of γ-lactone and alkyl citrate metabolites from fungi. Bold bonds show oxaloacetate-derived carbons where known.Sporochartines 6a–6d8,9 from the fungus Hypoxylon monticulosum CLL 205 (now referred to as Hypomontagnella spongiphila)10 possesses potent cytotoxicity (IC50: 7.2 to 21.5 μM) vs. human cancer cell lines and are proposed to be Diels Alder (DA) adducts of the furofurandione sporothriolide 1, itself a potent antifungal agent (EC50: 11.6 ± 0.8 μM),11 and trienylfuranol A 7,12 originally obtained from an endophytic fungus Hypoxylon submonticulosum DAOMC 242471 (now referred to as Hypomontagnella submonticulosa).13 Since the biosynthesis of sporothriolide 1 and related compounds is unknown, and biological DA reactions in fungi are currently of high interest,14 we decided to examine the biosynthesis of the sporochartines 6 in the Hypomontagnella spp. strains MUCL 54604 and CLL 205 (ref. 10 and 13) in detail.  相似文献   

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

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

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

8.
Signal Amplification by Reversible Exchange (SABRE) is a catalytic method for improving the detection of molecules by magnetic resonance spectroscopy. It achieves this by simultaneously binding the target substrate (sub) and para-hydrogen to a metal centre. To date, sterically large substrates are relatively inaccessible to SABRE due to their weak binding leading to catalyst destabilisation. We overcome this problem here through a simple co-ligand strategy that allows the hyperpolarisation of a range of weakly binding and sterically encumbered N-heterocycles. The resulting 1H NMR signal size is increased by up to 1400 times relative to their more usual Boltzmann controlled levels at 400 MHz. Hence, a significant reduction in scan time is achieved. The SABRE catalyst in these systems takes the form [IrX(H)2(NHC)(sulfoxide)(sub)] where X = Cl, Br or I. These complexes are shown to undergo very rapid ligand exchange and lower temperatures dramatically improve the efficiency of these SABRE catalysts.

The scope of the hyperpolarisation method Signal Amplification by Reversible Exchange (SABRE) is dramatically expanded through the use of co-ligands to substrates that weakly interact with the active cataylst.

Hyperpolarised magnetic resonance is receiving increasing attention from both the analytical science and medical communities due to its ability to create signals that are many orders of magnitude higher than those normally detected under Boltzmann control.1–6 The time and cost benefits associated with this improvement have propelled this area of research forward over the past few decades. Two of the most prominent techniques used to create hyperpolarisation are dissolution Dynamic Nuclear Polarisation (d-DNP) and Para-Hydrogen Induced Polarisation (PHIP),7,8 which derive their non-Boltzmann spin energy level populations from interactions with unpaired electrons and para-hydrogen (p-H2, the singlet spin isomer of hydrogen), respectively. Both of these methods have been reviewed in detail.3–5,9,10Signal Amplification by Reversible Exchange (SABRE) is a PHIP method that does not involve the chemical incorporation of p-H2 into the target substrate.11,12 Instead, under SABRE, spin order transfer proceeds catalytically through the temporary formation of a scalar coupling network between p-H2 derived hydride ligands and the substrate''s nuclei whilst they are located in a transient metal complex. The most common catalysts are of the type [Ir(H)2(NHC)(sub)3]Cl (where NHC = N-heterocyclic carbene and sub = the substrate of interest, Fig. 1a),13,14 although other variants are known.15–17 For SABRE to be accomplished, the target substrate must be able to reversibly ligate to the metal centre and this limits the methods applicability; although several routes to overcome this have been reported.18–20 Recently, the use of bidentate ancillary ligands such as NHC-phenolates16 and phosphine-oxazoles21 has been shown to expand the applicability of SABRE for a variety of different ligands and solvents (Fig. 1b). For example, use of the PHOX ligand (PHOX = (2-diphenylphosphanyl)phenyl-4,5-dihydrooxazole) gives 1H NMR signal gains of up to 132-fold for 2-picoline; a substrate previously shown to be unpolarised under classic SABRE conditions.22Open in a separate windowFig. 1Development of the SABRE method for hyperpolarisation of a range of substrates.The use of co-ligands to stabilise the active SABRE catalyst has proven successful for substrates that weakly associate to the catalyst (Fig. 1c). Of particular note is the hyperpolarisation of sodium [1,2]-13C2-pyruvate23 and sodium 13C-acetate24 which could be used as in vivo metabolic probes. The importance of co-ligands in breaking the chemical symmetry of the SABRE catalyst is also well established and co-ligands such as acetonitrile,25 sulfoxides,23,26 1-methyl-1,2,3-triazole27 and substrate isotopologues28 have been employed.We report here on the use of co-ligands to allow the NMR hyperpolarisation of weakly binding N-heterocyclic derived substrates with functionality in the ortho-position that have proven to be routinely inaccessible to the SABRE technique (Fig. 1d). 1H signal gains of up to 1442 ± 84-fold were obtained for some of these substituted pyridines at 9.4 T and the expansion of this approach to 13C and 15N detection and other N-heterocyclic motifs is also exemplified.  相似文献   

9.
10.
Controlled protein functionalization holds great promise for a wide variety of applications. However, despite intensive research, the stoichiometry of the functionalization reaction remains difficult to control due to the inherent stochasticity of the conjugation process. Classical approaches that exploit peculiar structural features of specific protein substrates, or introduce reactive handles via mutagenesis, are by essence limited in scope or require substantial protein reengineering. We herein present equimolar native chemical tagging (ENACT), which precisely controls the stoichiometry of inherently random conjugation reactions by combining iterative low-conversion chemical modification, process automation, and bioorthogonal trans-tagging. We discuss the broad applicability of this conjugation process to a variety of protein substrates and payloads.

Controlled protein functionalization holds great promise for a wide variety of applications.

Applications of protein conjugates are limitless, including imaging, diagnostics, drug delivery, and sensing.1–4 In many of these applications, it is crucial that the conjugates are homogeneous.5 The site-selectivity of the conjugation process and the number of functional labels per biomolecule, known as the degree of conjugation (DoC), are crucial parameters that define the composition of the obtained products and are often the limiting factors to achieving adequate performance of the conjugates. For instance, immuno-PCR, an extremely sensitive detection technique, requires rigorous control of the average number of oligonucleotide labels per biomolecule (its DoC) in order to achieve high sensitivity.6 In optical imaging, the performance of many super-resolution microscopy techniques is directly defined by the DoC of fluorescent tags.7 For therapeutics, an even more striking example is provided by antibody–drug conjugates, which are prescribed for the treatment of an increasing range of cancer indications.8 A growing body of evidence from clinical trials indicates that bioconjugation parameters, DoC and DoC distribution, directly influence the therapeutic index of these targeted agents and hence must be tightly controlled.9Standard bioconjugation techniques, which rely on nucleophile–electrophile reactions, result in a broad distribution of different DoC species (Fig. 1a), which have different biophysical parameters, and consequently different functional properties.10Open in a separate windowFig. 1Schematic representation of the types of protein conjugates.To address this key issue and achieve better DoC selectivity, a number of site-specific conjugation approaches have been developed (Fig. 1b). These techniques rely on protein engineering for the introduction of specific motifs (e.g., free cysteines,11 selenocysteines,12 non-natural amino acids,13,14 peptide tags recognized by specific enzymes15,16) with distinct reactivity compared to the reactivity of the amino acids present in the native protein. These motifs are used to simultaneously control the DoC (via chemo-selective reactions) and the site of payload attachment. Both parameters are known to influence the biological and biophysical parameters of the conjugates,11 but so far there has been no way of evaluating their impact separately.The influence of DoC is more straightforward, with a lower DoC allowing the minimization of the influence of payload conjugation on the properties of the protein substrate. The lowest DoC that can be achieved for an individual conjugate is 1 (corresponding to one payload attached per biomolecule). It is noteworthy that DoC 1 is often difficult to achieve through site-specific conjugation techniques due to the symmetry of many protein substrates (e.g., antibodies). Site selection is a more intricate process, which usually relies on a systematic screening of conjugation sites for some specific criteria, such as stability or reactivity.17Herein, we introduce a method of accessing an entirely new class of protein conjugates with multiple conjugation sites but strictly homogenous DoCs (Fig. 1c). To achieve this, we combined (a) iterative low conversion chemical modification, (b) process automation, and (c) bioorthogonal trans-tagging in one workflow.The method has been exemplified for protein substrates, but it is applicable to virtually any native bio-macromolecule and payload. Importantly, this method allows for the first time the disentangling of the effects of homogeneous DoC and site-specificity on conjugate properties, which is especially intriguing in the light of recent publications revealing the complexity of the interplay between payload conjugation sites and DoC for in vivo efficacy of therapeutic bioconjugates.18 Finally, it is noteworthy that this method can be readily combined with an emerging class of site-selective bioconjugation reagents to produce site-specific DoC 1 conjugates, thus further expanding their potential for biotechnology applications.19  相似文献   

11.
Hybrid materials comprised of inorganic quantum dots functionalized with small-molecule organic chromophores have emerged as promising materials for reshaping light''s energy content. Quantum dots in these structures can serve as light harvesting antennas that absorb photons and pass their energy to molecules bound to their surface in the form of spin-triplet excitons. Energy passed in this manner can fuel upconversion schemes that use triplet fusion to convert infrared light into visible emission. Likewise, triplet excitons passed in the opposite direction, from molecules to quantum dots, can enable solar cells that use singlet fission to circumvent the Shockley–Queisser limit. Silicon QDs represent a key target for these hybrid materials due to silicon''s biocompatibility and preeminence within the solar energy market. However, while triplet transfer from silicon QDs to molecules has been observed, no reports to date have shown evidence of energy moving in the reverse direction. Here, we address this gap by creating silicon QDs functionalized with perylene chromophores that exhibit bidirectional triplet exciton transfer. Using transient absorption, we find triplet transfer from silicon to perylene takes place over 4.2 μs while energy transfer in the reverse direction occurs two orders of magnitude faster, on a 22 ns timescale. To demonstrate this system''s utility, we use it to create a photon upconversion system that generates blue emission at 475 nm using photons with wavelengths as long as 730 nm. Our work shows formation of covalent linkages between silicon and organic molecules can provide sufficient electronic coupling to allow efficient bidirectional triplet exchange, enabling new technologies for photon conversion.

We demonstrate that silicon quantum dots can exchange spin triplet excitons with molecules covalently attached to their surface. Such hybrid materials can enable systems that upconvert incoherent far-red light into the visible spectral range.

Hybrid materials comprised of inorganic quantum dots (QDs) interfaced with small-molecule organic chromophores have emerged as a promising platform for materials that convert near-infrared radiation into the visible spectral range.1–3 In these structures, QDs act as light-harvesting antennas, absorbing long-wavelength photons and passing their energy to organic molecules bound to their surface in the form of spin-triplet excitons. These excitons can then be transferred into a surrounding medium, typically a solution or thin film, where pairs of them can fuse to form a bright spin-singlet state that can emit a short-wavelength photon.4–8 Due to the long lifetime of molecular triplet excitons, which can range from several microseconds to milliseconds, these materials can operate at low photon flux, enabling their integration into light-harvesting systems that operate under solar flux9,10 and limiting heat dissipation during their use in biological applications, such as phototherapy,11,12 live-cell imaging,13,14 and optogenetics.15 These hybrid materials can also be used to study interfacial energy transfer processes fundamental to the operation of solar cells that use triplet fusion''s inverse process, singlet fission, to enhance their performance.9,16–21 The simplest design for a cell of this type is one that interfaces a singlet fission material directly in line with a back-contacted semiconductor solar cell.22–24 In these structures, the singlet fission material acts as a light sensitizer that captures high-energy photons and uses their energy to generate pairs of triplet excitons that can be passed to the semiconductor to produce photocurrent. As molecules can be readily attached to QDs via a variety of chemical tethers, these materials allow detailed study of how the structure of the organic:inorganic interface impacts the ability of triplet excitons to move from one material to the other.For both triplet fusion-based light upconversion and singlet fission-based light harvesting, silicon represents a key material of interest. While several upconversion systems have been derived using QDs containing toxic elements, such as Cd5,7,25 or Pb,6,8,26,27 Si QDs are nontoxic, making them attractive for biological applications.28 Silicon also dominates the solar energy market, accounting for ∼90% of solar power production,29,30 making Si:organic interfaces that readily transmit triplet excitons a key design target for singlet fission-based solar cells.18,19,22 Previously, we have shown triplet exciton transfer from Si QDs to surface-bound anthracene molecules can power a photon upconversion system that operates with 7% efficiency.31 However, the inverse energy transfer process that is key for singlet fission devices, triplet exciton transfer from surface-bound molecules to Si, was not observed in our prior work.In this report, we address triplet exciton transfer from molecules to Si by demonstrating a hybrid Si QD:perylene system wherein photoexcitation of the Si QD establishes a spin-triplet exciton population that exists in a dynamic equilibrium between the QD and perylene molecules bound to its surface. While such exciton cycling has been reported for other QD:molecule systems,32–34 our work represents the first observation of this behavior in Si QD based systems. Using nanosecond transient absorption spectroscopy, we find triplet exciton transfer from Si to perylene takes place on a 4.2 μs timescale while energy transfer in the reverse direction occurs more than two orders of magnitude faster, on a 22 ns timescale. We attribute this difference in energy transfer rates to differences in the exciton density of states between perylene molecules and Si QDs. To demonstrate the utility of triplet excitons produced by this system for photon conversion applications, we have constructed a photon upconversion system by interfacing perylene-functionalized Si QDs with a complementary perylene-based triplet fusion annihilator. We find this system performs well, upconverting radiation with a wavelength as long as 730 nm into blue light centered near 475 nm. Under 532 nm illumination, the system upconverts light with an efficiency of 1.5% under incident light fluxes as low as 80 mW cm−2. This performance is comparable to that recently demonstrated using the same perylene annihilator coupled with a Pd-porphyrin light absorber.35 Our work demonstrates that the introduction of short, chemical linkers between molecules and Si can enable triplet exciton exchange between these materials for the design of new systems for both photon upconversion and light harvesting.  相似文献   

12.
13.
Ellman''s reagent has caused substantial confusion and concern as a probe for thiol-mediated uptake because it is the only established inhibitor available but works neither efficiently nor reliably. Here we use fluorescent cyclic oligochalcogenides that enter cells by thiol-mediated uptake to systematically screen for more potent inhibitors, including epidithiodiketopiperazines, benzopolysulfanes, disulfide-bridged γ-turned peptides, heteroaromatic sulfones and cyclic thiosulfonates, thiosulfinates and disulfides. With nanomolar activity, the best inhibitors identified are more than 5000 times better than Ellman''s reagent. Different activities found with different reporters reveal thiol-mediated uptake as a complex multitarget process. Preliminary results on the inhibition of the cellular uptake of pseudo-lentivectors expressing SARS-CoV-2 spike protein do not exclude potential of efficient inhibitors of thiol-mediated uptake for the development of new antivirals.

Thiol-reactive inhibitors for the cellular entry of cyclic oligochalcogenide (COC) transporters and SARS-CoV-2 spike pseudo-lentivirus are reported.

Thiol-mediated uptake1–10 has been developed to explain surprisingly efficient cellular uptake of substrates attached to thiol-reactive groups, most notably disulfides. The key step of this mechanism is the dynamic covalent thiol-disulfide exchange between disulfides of the substrates and exofacial thiols on cell surfaces (Fig. 1). The covalently bound substrate then enters the cell either by fusion, endocytosis, or direct translocation across the plasma membrane into the cytosol. Thiol-disulfide exchange has been confirmed to play an essential role in the cellular entry of some viruses1,11–14 and toxins.2 Indeed, diphtheria toxin and HIV were among the first to be recognized to enter cells via thiol-mediated uptake.1,2 The involvement of cell-surface thiols in cellular uptake is most often probed by inhibition with Ellman''s reagent (DTNB). However, this test is not always reliable, in part due to the comparably poor reactivity of DTNB, and the comparably high reactivity of the disulfide obtained as a product. Thus, the importance of thiol-mediated uptake for viral entry and beyond remains, at least in part, unclear.Open in a separate windowFig. 1In thiol-mediated uptake, dynamic covalent exchange with thiols on the cell surface precedes entry through different mechanisms. Inhibition of thiol-mediated uptake by removal of exofacial thiols and disulfides could thus afford new antivirals.We became interested in thiol-mediated uptake3–5 while studying the cytosolic delivery of substrates such as drugs, probes and also larger objects like proteins or quantum dots with cell-penetrating poly(disulfide)s.6 Our recent focus shifted to cyclic oligochalcogenides (COCs) to increase speed and selectivity of dynamic covalent thiol-oligochalcogenide exchange, and, most importantly, to assure reversibility, i.e., mobility during uptake, with a covalently tethered, intramolecular leaving group.7 With increasingly unorthodox COC chemistry, from strained disulfides7,8 and diselenides9 to adaptive dynamic covalent networks produced by polysulfanes,10 uptake activities steadily increased. Their high activities suggested that the same, or complementary, COCs could also function as powerful inhibitors of thiol-mediated uptake that ultimately might perhaps lead to antivirals. In the following, this hypothesis is developed further.Fluorescently labeled COCs 18 and 210 were selected as reporters for the screening of thiol-mediated uptake inhibitors because of their high activity, their destination in the cytosol, and their different characteristics (Fig. 2). The COC in 1 is an epidithiodiketopiperazine (ETP). With a CSSC dihedral angle ∼0°, ETPs drive ring tension to the extreme.15,16 Ring-opening thiol-disulfide exchange is ultrafast, and the released thiols are acidic enough to continue exchanging in neutral water, including ring closure.8 This unique exchange chemistry coincides with efficient cellular uptake and poor retention on thiol affinity columns.8Open in a separate windowFig. 2Structure of reporters 1 and 2 and inhibitor candidates 3–30 with their concentrations needed to inhibit by ∼15% (MIC) the uptake of 1 (1 h pre-incubation with inhibitors, 30 min incubation with reporter, filled symbols) and 2 (4 h pre-incubation, empty symbols). Red squares: ETPs; orange circles: BPSs; blue upward triangles: heteroaromatic sulfones; purple diamonds: thiosulfonates; magenta downward triangles: di- and polysulfides; brown hexagons: thiosulfinates. Symbols with upward arrows: MIC not reached at the highest concentration tested. Symbols with downward arrows indicate the lowest concentration tested already exceeds the MIC. (a) Similarly active upon co-incubation of reporters and inhibitor; (b–d) similarly (b), less (c), or more (d) active upon co-incubation in the presence of serum (mostly 6 h); (e) pre-incubation for 15 min; (f) isomerizes into cis22; (g) V-shaped DRC (see Fig. 3f); (h) pre-incubation for 30 min, co-incubation with 2; (i) mixture of regioisomers.The COC in 2 is a benzopolysulfane (BPS). Like ETPs, BPSs occur in natural products and have inspired total synthesis.17 Unlike ETPs, BPSs are not strained but evolve into adaptive networks of extreme sulfur species for cells to select from. Uptake efficiencies and retention on thiol affinity columns exceed other COCs clearly.10,18With COCs 1 and 2 as cell-penetrating reporters, a fully automated, fluorescent microscopy image-based high-content high-throughput (HCHT)19 inhibitor screening assay was developed. HeLa cells in multiwell plates are incubated with a reporter at constant and inhibitors at varying concentrations and incubation times. Hindered reporter uptake then causes decrease of fluorescence inside of cells (Fig. 3a). Automated data analysis19 was established to extract average fluorescence intensity per cell and, at the same time, cell viability from propidium iodide negative nuclei count (Fig. 3 and S3–S6). Standard assay conditions consisted of pre-incubation of HeLa cells with inhibitors for different periods of time, followed by the removal of inhibitors and the addition of reporters, thus excluding possible interactions between the two in the extracellular environment. In alternative co-incubation conditions, inhibitors were not removed before the addition of reporters to allow for eventual interactions between the two.Open in a separate windowFig. 3(a) Fluorescence image of HCHT plates (4 images per well) with HeLa cells pre-incubated with 6 (30 min) followed by co-incubation with 1 (left) and 2 (right, 10 μM each) for constant 30 min. (b–f) HCHT data showing relative fluorescence intensity (filled symbols) and cell viability (empty symbols) of HeLa cells after (b) pre-incubation with 4 for 1 h, followed by washing and incubation with 1 (top), or pre-incubation with 4 for 30 min, followed by co-incubation with 4 and 2 (bottom). (c) As in (b) with 18. (d) As in (b) after incubation for 4 h with 16 followed by incubation with 2. (e) As in (b) after pre-incubation with 11 (circles), 14 (crosses), or 21 (diamonds) for 15 min, followed by washing and incubation with 1. (f) As in (b) after pre-incubation with 20 (30 min), followed by washing and incubation with 1.Among the very high number of thiol-reactive probes, compounds 3–30 were selected based on promise, experience, availability and accessibility. Main focus was on COCs offering increasingly extreme sulfur chemistry because dynamic covalent thiol-oligochalcogenide exchange with different intramolecular leaving groups promises access to different exchange cascades for the intramolecular and, perhaps, also intermolecular crosslinking of the target proteins. More hydrophilic, often anionic COCs were preferred to prevent diffusion into cells and thus minimize toxicity. The expectation was that from such a sketchy outline of an immense chemical space, leads could be identified for future, more systematic exploration. Reporters 1 and 2 and candidates 3–30 were prepared by substantial multistep synthesis (Schemes S1–S11 and Fig. S47–S93, commercially available: 20, 25, 30). Inhibitors were numbered in the order of efficiency against reporter 1, evaluated by their minimum inhibitory concentrations (MICs), i.e., concentrations that cause a ∼15% reduction of reporter uptake in cells (Fig. 2 and Tables S1–S37). We chose to use MICs because half-maximal inhibitions could not always be reached due to the onset of toxicity, formally anticooperative, or even V-shaped dose–response curves (DRCs, e.g., Fig. 3b–f, all DRCs can be found in the ESI, Fig. S7–S43). MICs are usually below the half-maximal cell growth inhibition concentration (GI50, Tables S1–S37).Among the most potent inhibitors of ETP reporter 1 were ETPs 4 and 5 (Fig. 2, ,3b).3b). This intriguing self-inhibition was even surpassed by the expanded cyclic tetrasulfide ETP43 (MIC < 0.1 μM), which was of interest because they are much poorer transporters.10 Further formal ring expansion leads to cyclic pentasulfides BPS56 as equally outstanding inhibitors (MIC ≈ 0.3 μM). This trend toward the adaptive networks, reminiscent of elemental sulfur chemistry, did not extend toward inorganic polysulfides 13 (MIC ≈ 20 μM). ETPs 4 and 5 were sensitive to modification of the carboxylate, with the cationic 12 being the worst (MIC ≈ 30 μM) and the neutral glucose hemiacetal 7 the most promising (MIC ≈ 0.5 μM).Although this study focuses on increasingly extreme dynamic covalent COC chemistry, the inclusion of one example for covalent C–S bond formation was of interest for comparison. The classical iodoacetamides7 and maleimides4 were more toxic than active (not shown). However, nucleophilic aromatic substitution of heteroaromatic sulfones,20 just developed for the efficient bioorthogonal conversion of thiols into sulfides, was more promising. Weaker than dynamic covalent COCs, this irreversible inhibition was best with benzoxazole 11 (MIC ≈ 15 μM) and decreased in accordance with reactivity toward free thiols to oxadiazole 14 and benzothiazole 21 (MIC ≈ 300 μM, Fig. 3e).At constant pH, Ellman''s reagent 20 was confirmed to be erratic also in this assay. The DRC showed minor inhibition up to around 2 mM, which disappeared again at higher concentrations (Fig. 3f). Other cyclic disulfides were inactive as well (28–30). Also disappointing were oxidized disulfides, that is thiosulfinates, including allicin 25, the main odorant component of garlic,21,22 oxidized cystine 26 and oxidized lipoic acid 27. Thiosulfinates were of interest because they should selectively target the vicinal thiols of reduced disulfides bridges, producing two disulfides.23 The most active trans dithioerythrol (DTE) thiosulfinate 17 isomerized with time into the less active, hydrogen-bonded cis isomer 22 (Fig. S46).Reporter 2 was more difficult to inhibit than 1, as expected from high activity with extreme retention on thiol affinity columns.10,18 For instance, BPS 6 was very efficient against ETP 1 but much less active against BPS 2 (Fig. 3a), although longer pre-incubation could lower the MIC down to 4 μM (Fig. 2, S41). The complementary ETP 4 “self-inhibited” ETP 1 but was also unable to inhibit BPS 2 as efficiently (Fig. 3b). Among the best inhibitors of BPS 2 upon co-incubation were disulfide bridged γ-turn24 peptides 18 and 19 (MIC ≈ 5 μM), both less active against 1 (MIC ≈ 300 μM, Fig. 3c). Disulfide-bridged γ-turn CXC peptides consist of an 11-membered ring with significant Prelog strain. They were introduced by Wu and coworkers as transporters for efficient cytosolic delivery.5 The cyclic thiosulfonates 15 and 16 showed promising activities against both 1 and 2, and were tolerant toward the presence of serum (Fig. 2d, S33 and S42). Contrary to thiosulfinate 27, the oxidation of lipoic acid to pure thiosulfonates was not successful so far. However, weakly detectable activity of the lipoyl-glutamate conjugate oxidized to the thiosulfinate (MIC ≈ 350 μM, not shown) compared to the inactive thiosulfinate 27 implied that lipoic acid oxidized to the thiosulfonate would also be less active than the glutamate conjugate 15.The oxidized DTE 1625–28 was particularly intriguing because it was more potent against 2 and could achieve nearly complete inhibition (MIC ∼ 20 μM, Fig. 3d). Highly selective for thiols, the cyclic thiosulfonate 16 was stable for weeks at room temperature, without precaution, in all solvents tested. The disulfides and sulfinates obtained from exchange with thiols were stable as well, and the latter can further react with disulfides27 for intramolecular or eventually intermolecular crosslinking of the target proteins.The overall mismatched inhibition profiles found for reporters 1 and 2 supported that thiol-mediated uptake proceeds through a series of at least partially uncoupled parallel multitarget systems instead of a specific single protein or membrane target. From proteomics studies with cysteine-reactive irreversible probes, it is known that different probes generally target different proteins.29b Proteomics analysis29a for asparagusic acid derived transporters supports the involvement of many targets beyond the commonly considered protein disulfide isomerases and the confirmed transferrin receptor.12–14,26–30 The unusual, formally anti-cooperative (Hill coefficients < 1) DRCs further supported thiol-mediated uptake as complex multitarget systems.Despite the complexity of these systems, results did not much depend on assay conditions. Compared to the standard protocol of pre-incubation with inhibitors followed by inhibitor removal and incubation with reporters 1 or 2 for detection, the co-incubation protocol, in which pre-incubation with inhibitors is followed by co-incubation with reporters 1 or 2 without inhibitor removal, gave reasonably similar results (Fig. 2). Inhibition characteristics naturally depended on pre-incubation time, with weaker activities at shorter and longer times, reflecting incomplete exchange and cellular response or other ways of inhibitor destruction, respectively. The presence of serum also did not affect the activities much (Fig. 2b–d).Preliminary studies on antiviral activity were performed with pseudo-lentivectors31 that express the D614G mutant11 of the SARS-CoV-2 spike protein and code for a luciferase reporter gene, which is expressed by the infected cells.12 A549 human lung alveolar basal epithelium cell line constitutively overexpressing ACE2 and TMPRSS2 was selected to facilitate the entry of the SARS-CoV-2 spike pseudo-lentivirus. The most significant activities were found for DTE thiosulfonate 16 with an IC50 around 50 μM, while toxicity was detected only at 500 μM (Fig. S44). The onset of inhibition could be observed for tetrasulfide ETP 3 at 50 μM, but it coincided with the appearance of cytotoxicity. Protease inhibition is less likely to be the mode of action, as similar activity was found with wild type A549 cells transduced with a standard lentivirus expressing vesicular-stomatitis virus G surface protein VSVG (Fig. S45).13 Short incubation times of cells and inhibitors before the addition of viruses disfavored contributions from changes in gene expression. More detailed studies are ongoing.The lessons learned from this study are that, firstly, thiol-mediated uptake can be inhibited efficiently by thiol-reactive reagents, confirming that thiol-mediated uptake exists and transporters like ETP 1 and BPS 2 do not simply diffuse into cells; the best inhibitors are more than 5000 times better than Ellman''s reagent. Secondly, inhibitor efficiencies vary with the transporters, supporting that thiol-mediated uptake operates as a complex multitarget system. The best inhibitors are COCs that operate with fast dynamic covalent exchange, suggesting that the reversibility provided by COCs is important. The inhibition of thiol-mediated uptake might contribute to activities of thiol-reactive antivirals such as 16, ETPs or ebselen, although they have been shown to bind to zinc fingers or inhibit proteases.16,25,32–34 Finally, the inhibitors reported here could also be of interest for delivery applications and might be worth investigation with regard to antiviral activity. We currently plan to focus more systematically on the most promising leads within COCs, particularly cyclic thiosulfonates, and to expand the screening campaign toward new attractive motifs.33–35  相似文献   

14.
Improved understanding of the effect of protein glycosylation is expected to provide the foundation for the design of protein glycoengineering strategies. In this study, we examine the impact of O-glycosylation on the binding selectivity of a model Family 1 carbohydrate-binding module (CBM), which has been shown to be one of the primary sub-domains responsible for non-productive lignin binding in multi-modular cellulases. Specifically, we examine the relationship between glycan structure and the binding specificity of the CBM to cellulose and lignin substrates. We find that the glycosylation pattern of the CBM exhibits a strong influence on the binding affinity and the selectivity between both cellulose and lignin. In addition, the large set of binding data collected allows us to examine the relationship between binding affinity and the correlation in motion between pairs of glycosylation sites. Our results suggest that glycoforms displaying highly correlated motion in their glycosylation sites tend to bind cellulose with high affinity and lignin with low affinity. Taken together, this work helps lay the groundwork for future exploitation of glycoengineering as a tool to improve the performance of industrial enzymes.

Improved understanding of the effect of protein glycosylation is expected to provide the foundation for the design of protein glycoengineering strategies.

The cell walls of terrestrial plants primarily comprise the polysaccharides cellulose, hemicellulose, and pectin, as well as the heterogeneous aromatic polymer, lignin. In nature, carbohydrates derived from plant polysaccharides provide a massive carbon and energy source for biomass-degrading fungi, bacteria, and archaea, which together are the primary organisms that recycle plant matter and are a critical component of the global carbon cycle. Across the various environments in which these microbes break down lignocellulose, a few known enzymatic and chemical systems have evolved to deconstruct polysaccharides to soluble sugars.1–6 These natural systems are, in several cases, being evaluated for industrial use to produce sugars for further conversion into renewable biofuels and chemicals.From an industrial perspective, overcoming biomass recalcitrance to cost-effectively produce soluble intermediates, including sugars for further upgrading remains the main challenge in biomass conversion. Lignin, the evolution of which in planta provided a significant advantage for terrestrial plants to mitigate microbial attack, is now widely recognized as a primary cause of biomass recalcitrance.7 Chemical and/or biological processing scenarios of lignocellulose have been evaluated8 and several approaches have been scaled to industrial biorefineries to date. Many biomass conversion technologies overcome recalcitrance by partially or wholly removing lignin from biomass using thermochemical pretreatment or fractionation. This approach enables easier polysaccharide access for carbohydrate-active enzymes and/or microbes. There are however, several biomass deconstruction approaches that employ enzymes or microbes with whole, unpretreated biomass.9,10 In most realistic biomass conversion scenarios wherein enzymes or microbes are used to depolymerize polysaccharides, native or residual lignin remains.11,12 It is important to note that lignin can bind and sequester carbohydrate-active enzymes, which in turn can affect conversion performance.13Therefore, efforts aimed at improving cellulose binding selectivity relative to lignin have emerged as major thrusts in cellulase studies.14–25 Multiple reports in the past a few years have made exciting new contributions to our collective understanding of how fungal glycoside hydrolases, which are among the most well-characterized cellulolytic enzymes given their importance to cellulosic biofuels production, bind to lignin from various pretreatments.15,17 Taken together, these studies have demonstrated that the Family 1 carbohydrate-binding modules (CBMs) often found in fungal cellulases are the most relevant sub-domains for non-productive binding to lignin,15,17,20,26 likely due to the hydrophobic face of these CBMs that is known to be also responsible for cellulose binding (Fig. 1).27Open in a separate windowFig. 1Model of glycosylated CBM binding the surface of a cellulose crystal. Glycans are shown in green with oxygen atoms in red, tyrosines known to be critical to binding shown in purple, and disulfide bonds Cys8–Cys25 and Cys19–Cys35 in yellow.Furthermore, several studies have been published recently using protein engineering of Family 1 CBMs to improve CBM binding selectivity to cellulose with respect to lignin. Of particular note, Strobel et al. screened a large library of point mutations in both the Family 1 CBM and the linker connecting the catalytic domain (CD) and CBM.21,22 These studies demonstrated that several mutations in the CBM and one in the linker led to improved cellulose binding selectivity compared to lignin. The emerging picture is that the CBM-cellulose interaction, which occurs mainly as a result of stacking between the flat, hydrophobic CBM face (which is decorated with aromatic residues) and the hydrophobic crystal face of cellulose I, is also likely the main driving force in the CBM-lignin interaction given the strong potential for aromatic–aromatic and hydrophobic interactions.Alongside amino acid changes, modification of O-glycosylation has recently emerged as a potential tool in engineering fungal CBMs, which Harrison et al. demonstrated to be O-glycosylated.28–31 In particular, we have revealed that the O-mannosylation of a Family 1 CBM of Trichoderma reesei cellobiohydrolase I (TrCel7A) can lead to significant enhancements in the binding affinity towards bacterial microcrystalline cellulose (BMCC).30,32,33 This observation, together with the fact that glycans have the potential to form both hydrophilic and hydrophobic interactions with other molecules, led us to hypothesize that glycosylation may have a unique role in the binding selectivity of Family 1 CBMs to cellulose relative to lignin and as such, glycoengineering may be exploited to improve the industrial performance of these enzymes. To test this hypothesis, in the present study, we systematically probed the effects of glycosylation on CBM binding affinity for a variety of lignocellulose-derived cellulose and lignin substrates and investigated routes to computationally predict the binding properties of different glycosylated CBMs.  相似文献   

15.
While the development of chiral molecules displaying circularly polarized luminescence (CPL) has received considerable attention, the corresponding CPL intensity, glum, hardly exceeds 10−2 at the molecular level owing to the difficulty in optimizing the key parameters governing such a luminescence process. To address this challenge, we report here the synthesis and chiroptical properties of a new family of π-helical push–pull systems based on carbo[6]helicene, where the latter acts as either a chiral electron acceptor or a donor unit. This comprehensive experimental and theoretical investigation shows that the magnitude and relative orientation of the electric (μe) and magnetic (μm) dipole transition moments can be tuned efficiently with regard to the molecular chiroptical properties, which results in high glum values, i.e. up to 3–4 × 10−2. Our investigations revealed that the optimized mutual orientation of the electric and magnetic dipoles in the excited state is a crucial parameter to achieve intense helicene-mediated exciton coupling, which is a major contributor to the obtained strong CPL. Finally, top-emission CP-OLEDs were fabricated through vapor deposition, which afforded a promising gEl of around 8 × 10−3. These results bring about further molecular design guidelines to reach high CPL intensity and offer new insights into the development of innovative CP-OLED architectures.

A CPL intensity of up to 3 × 10−2 is achieved in π-extended 6-helicene derivatives, owing to an intense helicene-mediated exciton coupling. Corresponding top-emission CP-OLEDs afforded a promising gEl of around 8 × 10−3.

The design of chiral emitters displaying intense circularly polarized luminescence (CPL) has attracted significant interest, thanks to the potential of CP light in a diverse range of applications going from chiroptoelectronics (organic light-emitting diodes (OLEDs), optical information processing, etc.) to bio-imaging and chiral sensing.1 Recently, designing OLEDs with CP electroluminescence (CP-OLEDs) has emerged as an interesting approach to improve high-resolution display performance. Namely, using unpolarised OLEDs, up to 50% of the emitted light can be lost due to the use of antiglare polarized filters.2 In CP-OLEDs, the electro-generated light can pass these filters with less attenuation owing to its circular polarization and thus lead to an increase of the image brightness with lower power consumption.3 To develop CP-OLED devices, the main approach relies on the doping of the device''s emitting layer by a CPL emitter, which should ensure simultaneously high exciton conversion and a high degree of circular polarization. The harvesting of both singlet and triplet excitons has been successfully addressed using either chiral phosphorescent materials or thermally activated delayed fluorescence (CP-TADF) emitters with device efficiencies of up to 32%.4 However, the intensity of circularly polarized electroluminescence (CPEL), evaluated by the corresponding dissymmetry factor gEl, remains inefficient and typically falls within the range of 10−3 with limited examples reaching gEl > 10−2 based on polymeric materials and lanthanide complexes.5 For CP-OLEDs using a molecular chiral emissive dopant, gEl, defined as the ratio between the intensity difference of left- and right-CPEL, and the total generated electroluminescence, 2(ElL − ElR)/(ElL + ElR), can be generally related to the luminescence dissymmetry factor glum measured in diluted solution.2 Accordingly, it is of crucial importance to design luminescent molecules with high glum values,3,28a–d,29 in order to reach strong CP electro-luminescence when going to practical devices. However, structural and electronic factors that govern the CPL of chiral compounds are still poorly understood even if a few studies have recently tried to rationalize and establish molecular guidelines to obtain high glum values.6Our team has contributed to the research in this area by developing extended π-helical molecular architectures resulting from the association of carbo[6]helicene and achiral dyes,7 which afforded enhanced chiroptical properties, with notably a glum up to 10−2, owing to an uncommon chiral exciton coupling process mediated by the chiral helicenic unit.8 In addition, we also described an unusual solvent effect on the intensity of CPL of π-helical push–pull helicene–naphthalimide derivatives,7b which showed a decrease of glum from 10−2 to 10−3 upon increasing the polarity of solvent.7b This solvatochromism effect was shown to be related to a symmetry breaking of the chiral excited state before emission,9 which modifies the relative intensity of the magnetic (μm) and electric (μe) dipole transition moments, and the angle, θ, between them (Fig. 1), ultimately impacting glum. The latter is well approximated as 4|m|cos θ/(|μ|) for an electric dipole-allowed transition.10Open in a separate windowFig. 1Chemical structures of “push–pull” 2,15-diethynylhexahelicene-based emitters with their polarized luminescence characteristics including their calculated electric and magnetic transition dipole moments and the angle between them corresponding to the S1 → S0 transition.While these results highlight interesting aspects regarding the key parameters influencing the CPL of organic emitters, this type of “helical push–pull design” remains limited to only one example, which render the systematic rationalization of these findings difficult. Accordingly, we decided to develop a complete family of new chiral push–pull compounds to explore the structural and electronic impact of the grafted substituents on the helical π-conjugated system. In addition, we went a step further and incorporated the designed chiral emitter into proof-of-concept CP-OLEDs using a top-emission architecture,11 which remains scarcely explored for CP-light generation despite its considerable potential for micro-display applications. To the best of our knowledge, only one example of such type of electroluminescent device has been reported, using a CP-TADF emitter, affording a modest gEl of 10−3.11aHerein, we report the synthesis and chiroptical properties of a new family of π-helical push–pull systems based on chiral carbo[6]helicene, functionalized by either electron donor or acceptor units. Interestingly, the chiral π-conjugated system of the helicene may act as either an electron acceptor or a donor, depending on the nature of the attached substituents, thereby impacting the chiroptical properties, notably the resulting CPL. By optimizing the chiral exciton coupling process through the modulation of the magnitude and relative orientation of the electric (μ) and magnetic (m) dipoles, the chiroptical properties of classical carbo[6]helicene-based emitters can be dramatically enhanced and reach high glum values at the molecular level, i.e. up to 3–4 × 10−2. Experimental and theoretical investigations revealed that the mutual orientation of the electric and magnetic dipoles in the excited-state is a crucial parameter and is optimal when the substituents attached to the helicene core possess a rather weak electron withdrawing or donating ability. Finally, proof of concept top-emission CP-OLEDs were fabricated through vapor deposition of π-helical push–pull derivatives and afforded a gEl of around 8 × 10−3, which represents a significant improvement for the polarization of electroluminescence emitted using this device architecture.  相似文献   

16.
Site-selective fluorination of aliphatic C–H bonds remains synthetically challenging. While directed C–H fluorination represents the most promising approach, the limited work conducted to date has enabled just a few functional groups as the arbiters of direction. Leveraging insights gained from both computations and experimentation, we enabled the use of the ubiquitous amine functional group as a handle for the directed C–H fluorination of Csp3–H bonds. By converting primary amines to adamantoyl-based fluoroamides, site-selective C–H fluorination proceeds under the influence of a simple iron catalyst in 20 minutes. Computational studies revealed a unique reaction coordinate for the catalytic process and offer an explanation for the high site selectivity.

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

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

17.
In situ single-crystal diffraction and spectroscopic techniques have been used to study a previously unreported Cu-framework bis[1-(4-pyridyl)butane-1,3-dione]copper(ii) (CuPyr-I). CuPyr-I was found to exhibit high-pressure and low-temperature phase transitions, piezochromism, negative linear compressibility, and a pressure induced Jahn–Teller switch, where the switching pressure was hydrostatic media dependent.

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

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

18.
19.
20.
Here, we report the nitric oxide monooxygenation (NOM) reactions of a CoIII-nitrosyl complex (1, {Co-NO}8) in the presence of mono-oxygen reactive species, i.e., a base (OH, tetrabutylammonium hydroxide (TBAOH) or NaOH/15-crown-5), an oxide (O2− or Na2O/15-crown-5) and water (H2O). The reaction of 1 with OH produces a CoII-nitrito complex {3, (CoII-NO2)} and hydrogen gas (H2), via the formation of a putative N-bound Co-nitrous acid intermediate (2, {Co-NOOH}+). The homolytic cleavage of the O–H bond of proposed [Co-NOOH]+ releases H2via a presumed CoIII-H intermediate. In another reaction, 1 generates CoII-NO2 when reacted with O2−via an expected CoI-nitro (4) intermediate. However, complex 1 is found to be unreactive towards H2O. Mechanistic investigations using 15N-labeled-15NO and 2H-labeled-NaO2H (NaOD) evidently revealed that the N-atom in CoII-NO2 and the H-atom in H2 gas are derived from the nitrosyl ligand and OH moiety, respectively.

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

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

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